You are on page 1of 15

SPE-172952-MS

Comprehensive Spectral and Thermal Characterization of Oil Shales


S. Guven and S. Akin, Middle East Technical University; B. Hascakir, Texas A&M University

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Middle East Unconventional Resources Conference and Exhibition held in Muscat, Oman, 26 –28 January 2015.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents
of the paper have not been reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect
any position of the Society of Petroleum Engineers, its officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written
consent of the Society of Petroleum Engineers is prohibited. Permission to reproduce in print is restricted to an abstract of not more than 300 words; illustrations may
not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
The heterogeneous nature of oil shale resources associated to the depositional environments, lithology,
and organic content make the reserve estimation complex and unpredictable. However, comprehensive
laboratory studies on organic rich shale samples collected from different regions can increase the
understanding about the organic content of oil shales, interaction of shale with organic matter and injected
fluid used during enhanced oil recovery method. This study investigates the characterization of eight
different Turkish and American oil shale samples with several spectral methods and a thermal analysis.
The main purpose of this study is to characterize the oil shale samples to increase the understanding about
the organic content and interaction of shale with organic matter.
In this study, we used Thermal Gravimetric Analysis/Differential Scanning Calorimetry (TGA/DSC)
analysis to estimate organic content of each oil shale sample in air and nitrogen environments. X-Ray
Diffraction (XRD) was used to define minerals in oil shale. Fourier Transform Infrared Spectroscopy
(FTIR) was used to detect the mineral and kerogen in oil shale before and after the TGA/DSC analysis.
Scanning Electron Microscope (SEM) was used to characterize the depositional environment of each oil
shale samples. TGA/DSC results verified that oil shale samples have up to 50% of organic matter. XRD
and FTIR results helped to identify the organic and inorganic compounds. Effects of minerals and ions
were recognized by comparing TGA/DSC curves and FTIR spectra. It was recognized that the more
carbonate ion in the oil shale the more increase in weight loss occurred. Diatoms identified from SEM
results showed that depositional environments of the some oil shale samples are marine environments.
This study provides insight for the reserve estimation of the eight different oil shale samples with
comprehensive spectral and thermal characterization.

Introduction
The unconventional resources are so huge that it is several times larger than the conventional oil resources
(World Energy Council, 2010). Moreover, the abundance of the components of oil shale is immense when
it is compared to conventional fossil fuels (Yen, 1976). The hydrocarbons in oil shale can be used as an
alternative to conventional petroleum and natural gas.
As source rock for conventional oil, the distributions of the oil shale samples are generally close to the
oil fields in the world and seen in 100 major deposits. The United States, China, Russia, Brazil,
2 SPE-172952-MS

Figure 1—Locations of the oil shale samples

Democratic Republic of the Congo, Italy, Morocco, Jordan, Australia, Canada and Estonia are top known
countries that have the major oil shale deposits (World Energy Council, 2010).
United States is the leading country with more than 1,000 billion barrels technically recoverable oil
shale resources (International Energy Agency, 2010). Eocene Green River deposit is the major resource
and 83% of the United States’ oil shale resources are supplied from the Green River. Turkey also has a
remarkable amount of oil shale capacity such that it is the second fossil fuel potential in Turkey (Güleç
& Önen, 1992). The total oil shale reserves in Turkey are predicted to be more than 3-5 billion tonnes;
2.2 billion tonnes of which are thought to be technically recoverable (Biglarbigi, Mohan, & Killen, 2009).
Oil shale is generally defined as a sedimentary rock containing organic matter. Oil shale is the source
rock for most of the conventional oils; therefore, does not originally contain mature oil, but it can be
upgraded into hydrocarbons by releasing the organic matter inside with sufficient heat. Oil shales consist
of both organic and inorganic components: kerogen, bitumen and inorganic matrix (quartz; feldspars;
clays, mainly illite and chlorite; carbonates, calcite and dolomite; pyrite; and other minerals). The
insoluble part, kerogen comprises the major part of the organic matter; while; the soluble part, bitumen
constitutes insignificant amount. On the basis of dry weight, oil shales can contain 10-60% organic matter,
20-70% carbonate minerals and 15-60% sandy-clay minerals (Razvigorova et al., 2008).
Two vital factors for assessment of potential of oil shale formation as an oil resource are: how much
organic matter that formation bears (related to the total quantity of organic matter in the sediment) and
how much organic matter can be converted to petroleum (related to the quality or type of the organic
matter). Profitable oil extraction from oil shale can only be accomplished when the organic content of the
oil shale resource is approximately 2.5 wt% or more (Pamir, 1997). Moreover, type of the kerogen and
inorganic content of oil shale are important to be able to generate oil. Therefore, it is essential to analyze
oil shale samples qualitatively and quantitatively with different techniques. For this purpose several
thermal and spectral analyses are used in the literature. Fischer assay, elemental analysis, and Thermal
Gravimetric Analysis/Differential Scanning Calorimetry (TGA/DSC) are common methods used as
thermal analysis; while, X-Ray Diffraction (XRD), Nuclear Magnetic Resonance (NMR), and Infrared
Spectroscopy (IR) are used as spectral methods. Among the thermal methods, Fischer assay is the most
common method. However, Bhargava, Awaja, & Subasinghe (2005) stated that this technique is complex
and expensive. They used Fischer assay, TGA, XRD and IR techniques in their study to characterize
Australian oil shale samples and they found that using combination of XRD, TGA, and IR not only gives
same results as Fischer assay offers, but provides them in more economic and easy way. Moreover, they
SPE-172952-MS 3

Figure 2—SEM images of (a) Minerals (b) Organic content

Figure 3—FTIR spectra of original (a) Turkish and (b) American oil shale samples

stated that combination of TGA and IR gives consistent results to detect organic matter; while, combi-
nation of IR and XRD may be used to detect inorganic matter. On the other hand, Yan, Jiang, Han, & Liu
(2013) used the combination of Fourier Transform Infrared Spectroscopy (FTIR) and TGA/DSC to see the
effects of mineral matrix in oil shale pyrolysis and combustion of Huadian oil shale. They concluded that
inorganic minerals have a resultant effect which increases the reactivity of the organic matter. Moreover,
Ballice, Yüksel, Saglam, Schulz, & Hanoglu (1995) used FTIR and TGA/DSC to detect type of the
kerogen and kerogen - inorganic decomposition periods of the pyrolysis of Goynuk, Beypazari and
Timahdit oil shales. They classified the oil shale samples by using aliphatic, carboxyl/carbonyl and
aromatic compounds in their study; and found that if oil shale has high amount of organic matter,
secondary pyrolysis effect (coking) cannot be seen easily during the TG analysis. Thakur & Nuttall (1987)
studied on Moroccan oil shale under inert atmosphere by using TG analysis. They provided fundamentals
4 SPE-172952-MS

Figure 4 —Weight loss curves of Turkish and American oil shale samples

of the thermal decomposition stages for oil shales in this study. They stated that thermal decomposition
of Moroccan oil shale has two stages: First stage has decompositions of kerogen and bitumen (as an
intermediate product); while, the second stage has decomposition of inorganic matter. Kok, Senguler,
Hufnagel, & Sonel (2001) studied thermal behavior of Seyitomer oil shale under air injection by using
TGA/DSC. They recognized that at lower temperatures under air injection, organic decomposition
behavior of shales is similar to the pyrolysis (under nitrogen injection). However, they stated that at higher
temperatures this similarity cannot be seen due to oxidative characteristics of the organic content. Apart
from thermal and spectral analysis, Scanning Electron Microscope (SEM) is a different method used to
obtain physical information of samples about morphology. SEM shows surface structures and organic -
mineral existance of the oil shale samples in microscopic level.
In this work, we characterize eight oil shale samples from Turkey and USA by using FTIR, XRD,
TGA/DSC, and SEM analyses. The objectives of this work are to characterize the eight oil shale samples
and comprehensively investigate the rock-kerogen interaction under oxidative and inert environments at
elevated temperatures by using four different techniques. Moreover, we aim to compare Turkish and
American oil shales to understand the real potential of Turkish oil shales. In this paper, first, materials and
SPE-172952-MS 5

Figure 5—FTIR spectra of the original samples (on top) and pyrolysis postmortems (on bottom) under nitrogen

methods will be mentioned, and then, results of the analysis will be discussed. Finally, this study will be
concluded by making some arguments.
Materials and Methods
Samples
Turkish oil shale samples were collected from Hatildag, Himmetoglu, Seyitomer, and Ulukisla. The
American oil shale samples were obtained from Barnett, Eagle Ford, Green River, and Marcellus
formations (See Fig. 1). All four Turkish oil shale samples are from Tertiary Period and located near Bolu
6 SPE-172952-MS

Figure 6 —FTIR spectra of the original samples (on top) and combustion postmortems (on bottom) under air

(Hatildag and Himmetoglu), Kutahya (Seyitomer) and Nigde (Ulukisla). Similar to Turkish oil shales,
Green River (Colorado, Wyoming, and Utah) also deposited during Early Eocene Epoch of Tertiary
Period. Barnett and Eagle Ford oil shales deposited in Texas during Late Mississippian and Cretaceous
Periods, respectively. Marcellus is in eastern North America from Middle Devonian Period. For all oil
shale samples, characterization studies have been conducted on crushed samples that have 3-5 mm particle
sizes.
SPE-172952-MS 7

Material Analyses
Characterization studies were achieved with Fourier Transform Infrared Spectroscopy (FTIR) to
determine the molecular structure, X-Ray Diffraction (XRD) to define minerology, and Scanning Electron
Microscope (SEM) to visualize the surface stratigraphy. Thermogravimetric Analyses (TGA/DSC) were
conducted under nitrogen and air atmosphere to determine the reactivity of rock samples at varying
temperature ranges. After TGA/DSC, FTIR analyses were performed again to understand the change in
mineralogical and organic content of the samples. TGA/DSC analyses were performed with Netzsch STA
449 F3 Jupiter instrument at heating rate of 10 °C/min with aluminum oxide (Al2O3) crucibles. Samples
were grinded to size of ⬍60 mesh and 30-40 mg of each was used in the experiments. Two sets of
experiments were conducted with TGA/DSC on original oil shale samples. In the first set, samples were
subjected to heat (from 24 °C to 900 °C) under 50 ml/min air (purge) and 20 ml/min nitrogen protective
gas flow to check the organic and inorganic decomposition periods. In the second set, original samples
were heated under 50 ml/min nitrogen (purge) and 20 ml/min nitrogen protective gas flow (24 and 900
°C). The samples at the end of the first and second test were kept for further analysis. For simplicity, the
samples from the first set will be named combustion samples and from the second set, pyrolysis samples.
FTIR analyses were performed using Agilent Cary 630 FT-IR Spectrometer. The range of the spectra
is 4000-650 cm-1 at a spectral resolution of 8 cm-1 for 32 background and sample scans with Happ-Genzel
apodization. Analyses have been repeated three times for each sample. Measurements were conducted on
original oil shale samples and on the combustion and pyrolysis samples, to observe the mineralogical
changes with absorbance finger prints. Before each FTIR measurement, all samples were brought to 60
mesh size.
High magnification JEOL/JSM-7500F Field Emission Scanning Electron Microscope was used to
check the surface morphology of original samples. Samples were uncoated and placed on brass sample
mounts. Original sample size was maintained for SEM analyses.
XRD analyses were performed with Bruker D8 Advanced XRD instrument. The samples were
powdered and placed in the sample holder of a two circle goniometer. The X-ray source was a 2.2 kW
Cu X-ray tube and maintained at an operating current of 40 kV and 40 mA. The two-circle 250 mm
diameter goniometer was computer controlled with independent stepper motors and optical encoders for
the ␪ and 2␪ circles with the smallest angular step size of 0.0001°2␪.
Results and Discussion
The aim of this work is to analyze the reactivity and extraction potential of oil shale samples under
oxidative (air) and inert (nitrogen) atmosphere; and to extend our knowledge by analyzing different types
of oil shales samples with varity of mineralogical changes and which have different kerogen type and
amount.
The minerals of the oil shale samples obtained from XRD analysis show that quartz and dolomite/
calcite are the leading minerals (See Table 1). Other minerals commonly seen in samples are gypsum,
aragonite, muscovite, albite and pyrite minerals. Big hump seen in XRD spectrum of Himmetoglu oil
shale indicates low degree of crystallinity; showing that the sample is mostly amorphous (See Appendix
A – Fig. 7 (b)). Amorphous content in the oil shale represent Type I kerogen. Moreover, Largeau et al.
(1990) stated that kerogens in oil shale appear to be composed of homogeneous and amorphous organic
matter; therefore, we can say that Himmetoglu oil shale has significant amount of kerogen when compared
to others.
SEM images confirm the presence of quartz, pyrite and albite. The existence of those minerals is
confirmed with simultaneous Energy Dispersive X-Ray Spectroscopy (EDS) analysis. Figure 1 (a) shows
quartz from Green River, and pyrite and albite from Himmetoglu oil shale samples, respectively.
Algeas and some diatoms are indicative of marine environments. During the SEM analysis, fossils of
algeas and diatoms were recognized in Himmetoglu and Seyitomer samples (See Fig. 2 (b)). The
8 SPE-172952-MS

Table 1—Minerals from XRD analysis


Oil Shale Minerals

Ulukisla Gypsum, aragonite, quartz, dolomite, muscovite


Himmetoglu Sphalerite, pyrite, albite, quartz
Hatildag Pyrite, dolomite, albite, analcime, muscovite, gypsum
Seyitomer Gypsum, aragonite, pyrite, quartz, dolomite, albite, muscovite, clinochlore, lithiophorite
Barnett Fluorapatite, quartz, clinochlore, muscovite, albite
Marcellus Muscovite, quartz, pyrite, calcite
Eagle Ford Quartz, gypsum, calcite, kaolinite
Green River Quarz, albite, siderite, dolomite, pyrite, dawsonite, analcime

simultaneous EDS analysis with SEM confirms the organic content of these fossils with high oxygen
counts and silica content.
FTIR uses the interaction of infrared radiation with the vibrating dipole moments of molecules. FTIR
spectra represent the molecular absorption of the different chemical bonds and certain bonds of the
original oil shale samples (Fig. 3). It is seen that almost all of the samples have similar kerogen bands from
FTIR spectra as stated in other studies (Alstadt, Katti, & Katti, 2012; Levy & Stuart, 1984; Rouxhet,
Robin, & Nicaise, 1980). Between 3570 and 3200 cm-1 a hydroxy group, H-bonded OH stretch can be
seen as a broad band in some of the samples (Coates & Ed, 2000). The FTIR spectra show that almost
all the original oil shale samples have asymmetric and symmetric CH2 bridge bands of kerogen at around
2920 and 2850 cm-1, respectively. Moreover, from near 1700 cm-1 carbonyl C⫽O stretch bands and from
1600 cm-1 C⫽C aromatic structures can be noticed. The band around 1440 cm-1 shows C-H bending
vibrations of CH2 and CH3, and the peak at around 1375 cm-1 represents the C-H bending of CH3. Right
hand side of the FTIR spectra shows the mineral bands of shale samples. The distinct broad peak at around
1436 cm-1 represents the CO32- ion from dolomite and calcite. Other two sharp peaks at 876 and 727 cm-1
represent carbonate ion from dolomite and calcite (Alstadt et al., 2012; Downs, 2006). Other broad peak
around 1050 cm-1 with small peaks around 770 and 700 cm-1 represents Si-O stretch of quartz and silica.
From Figure 3, it can be seen that Turkish oil shale samples have more significant OH stretch than
American oil shale samples. Based on the intensity of the peaks around 2920, 2850, 1440, and 1375 cm-1,
we can say that the kerogen of Himmetoglu oil shale has high aliphatic content. Hatildag, Seyitomer, and
Green River oil shales also have relatively high aliphatic content with high asymmetric and symmetric
stretch of methyl groups. The kerogen of Barnett and Ulukisla oil shale samples are more aromatic than
others due to the relatively high C⫽C aromatic structures around 1600 cm-1.
The reaction kinetics of oil shale samples are studied under air and nitrogen injection wih TGA/DSC.
Figure 4 summaries these results. TG curves indicate weight loss precentages and DSC curves represent
heat flow during the analysis. Differential thermogravimetric (DTG) curves of combustion analysis are
provided in Appendix B. For all samples in Figure 4, y axis on the left hand side represents weight loss
percentages and y axis on the right hand side gives heat flow behavior of oil shale samples. Therefore, the
solid curves in Figure 4 represent TG curves and dashed curves represent heat flow.
Energy generation of Himmetoglu oil shale is increased under air injection, which might be due to its
high organic content and the type of kerogen (Type I). Fuel formation reactions are observed in all but
Ulukisla oil shale. This could be due to low organic content of Ulukisla oil shale. Complete thermal
decoposition of kerogen is accomplished in the ascending order for: Barnett, Eagle Ford, Seyitomer,
Hatildag, Marcellus, Green River, and Himmetoglu samples. As the thermal decomposition temperature
increases, the energy generation increases.
In the first set of experiments, pyrolysis section takes place within nitrogen environment. Figure 4 (b)
and (d) show the total weight loss and heat flow curves of pyrolysis analysis from 24 °C to 900 °C. There
is no significant weight loss observed during pre-heating period (24-200 °C), indicating very low moisture
SPE-172952-MS 9

Table 2—Weight loss percentages of oil shales

content in the samples except for Ulukisla oil shale with 9% moisture content. As explained by Alstadt
et al. (2012), the relatively high hump between 3600 and 3000 cm-1 of the FTIR spectrum of original
Ulukisla oil shale shows the H-OH hydrogen bonds from water also supports the presence of moisture
content. Table 2 shows the moisture contents of the samples separately.
Two stages during pyrolysis process are identified as organic and inorganic decomposition periods. The
first stage occurs roundly between 200 °C and 600 °C and represents organic decomposition; where
degradation of kerogen and bitumen molecules occurs (Al-Harahsheh et al., 2011). During organic
decomposition, kerogen decomposition occurs generally until 350 °C. Then, the produced gas, bitumen,
and carbon residue of kerogen devolatilization are decomposed into oil, coke (char), and gas between 350
and 600 °C. The organic decomposition stage is completed with these two stages.
However, several studies show that, sometimes according to the type of the oil shale the organic
decomposition stage can be completed with one-stage (Al-Harahsheh et al., 2009; Fang-Fang, Ze,
Wei-Gang, & Wen-Li, 2010; Qing, Baizhong, Aijuan, Jingru, & Shaohua, 2007). Although the real
mechanism of the reaction is a little bit complicated, Thakur & Nuttall (1987) explained the two-stage of
the organic decomposition periods as follows:

The second stage refers to inorganic decomposition period and takes place generally between 600 °C
and 800 °C. In the second stage, weight loss is seen due to the decomposition of the inorganic compounds
of oil shale like carbonates. At high temperatures the produced carbondioxide during carbonate decom-
position reacts with residual char and forms carbonmonoxide. This process results in additional weight
loss (Qing et al., 2007).
Total weight losses during pyrolysis are shown in Table 2. As seen from Table 2, the total weight loss
of Himmetoglu oil shale is high (67%); while, Marcellus has the lowest (8.2%) weight loss. Weight loss
due to the organic decomposition is also highest in Himmetoglu oil shale; which indicates Himmetoglu
oil shale has the highest organic content. As Bhargava et al. (2005) stated, this situation can be confirmed
by FTIR spectra (See Fig. 3). The intensity of the peaks showing asymmetric and symmetric stretching
methylene bridge, carbonyl C⫽O stretch and C⫽C aromatic structure bands of kerogen verifies that
Himmetoglu oil shale has the highest organic content. Other oil shales (Seyitomer, Green River, and
Hatildag) that have methylene bridge bands also have relatively high organic decomposition. Similarly,
the oil shale samples that have significantly high carbonate minerals (from dolomite and calcite) show
10 SPE-172952-MS

high inorganic decomposition. Eagle Ford oil shale has the highest weight loss for the inorganic
decomposition with 25.2%. Hatildag, Green River, Ulukisla and Seyitomer are the other oil shales that
have high inorganic decomposition due to dolomite and calcite decomposition (See Table 2). As expected,
lack of carbonate ion results in the low inorganic decomposition period for Himmetoglu, Barnett, and
Marcellus oil shales. The effect of the other common mineral, pyrite, cannot be recognized during the
pyrolysis stage as Ballice (2005) reported in his study on Turkish oil shales.
FTIR spectra given below show that almost all of the asymmetric and symmetric CH2 bridge bands of
kerogen at around 2920 and 2850 cm-1 and carbonyl C⫽O stretch at around 1696 cm-1 disappear after
pyrolysis. Carbonate minerals of Seyitomer, Ulukisla, Eagle Ford, and Green River oil shales are
decomposed entirely; while, small amount of carbonate mineral of Hatildag oil shale sample remains after
the analysis. FTIR spectra of the postmortems of the Barnett and Marcellus oil shale samples are very
similar to the originals; verifying the low weight loss during pyrolysis.
In the second set, combustion analyses take place under air injection. Figure 4 (a) and (c) show the
weight loss and heat flow curves of the combustion analysis from 24 °C to 900 °C. In this set, weight loss
periods and postmortem characteristics differ from pyrolysis section by additional oxidation processes of
organic matters. The area under the DSC curve gives the enthalpy change during the thermal analysis. It
can be clearly seen that the enthalpy changes of the combustion analysis are higher than pyrolysis;
meaning that, in combustion of shales the difference between the energies is used to break the chemical
bonds and formation of new chemical bonds is higher.
Derivative of the TG curve (DTG) helps to identify the weight loss periods of the combustion analysis
(See Appendix B). Basically, we can say that there are two groups of peaks representing oxidation of the
kerogen and the inorganic constituents. First group of peaks between 200 and 600 °C are seen as a result
of combustion of the aliphatic components of kerogen and the char - product of the aliphatic component’s
combustion. Second group of peaks (after 600 °C) represents the combustion of the inorganic components.
The weight loss until 200 °C represents moisture content of the samples. Unlike other samples, Ulukisla
oil shale has two sharp derivative peaks before 200 °C. First peak below 100 °C represents water content
of the sample and the second peak between 100 and 200 °C represents the combustion of the light organic
contents.
In their study on Huadian oil shales, Yan, Jiang, Han, & Liu (2013) suggested that the first derivative
peak between 200 and 600 °C represents the oxidation of the aliphatic content of the kerogen. Later, the
residuals of this oxidation process are dehdrogenated and polymerized into the char. The second peak
represents the oxidation of that char. The chemical reactions of the combustion analysis are so complex
that each sample shows different mechanisms. Therefore, we can not recognize every derivative peak of
oxidation stages, or we can see additional peaks. For example, Seyitomer oil shale sample shows an
additional derivative peak between 200 and 600 °C. Lithiophorite mineral of Seyitomer oil shale may
cause this additional DTG peak. On the other hand, any chemical reaction between the inorganic matrix
and the kerogen of the Seyitomer oil shale sample may cause this additional weight loss period.
The oxidation of the inorganic matrix can be recognized above 600 °C. The oil shale samples that have
high carbonate ion show high weight loss above 600 °C (Levy & Stuart, 1984). Hatildag, Ulukisla,
Seyitomer, Eagle Ford, and Green River oil shale samples have distinctive peaks between 600 and 800 °C,
indicating that the weight loss in this period is due to the calcite and dolomite decomposition (See Fig.
3). Moreover, Marcellus oil shale sample has small DTG peak which is due to the presence of a low
amount of carbonate ion (Also see Fig. 6). Other oil shale samples, Himmetoglu and Barnett, do not have
typical DTG peak indicating that inorganic decomposition is taking place. Note that, these samples do not
have any carbonate ion from calcite or dolomite.
Table 2 shows the weight loss percentages of the oil shale samples during combustion analyses. Similar
to the pyrolysis section, Himmetoglu oil shale sample has the highest weight loss percentage. It is seen
SPE-172952-MS 11

that Marcellus and Barnett oil shale samples have the lowest weight loss percentages during combustion:
12.6% and 18.6%, respectively.
Figure 6 represents the FTIR spectra of the samples before and after the combustion analysis. Each of
the FTIR spectra of the combustion postmortems shows similar peaks; the only remainings are Si-O
stretch of quartz and silica. This indicates almost all organic content and calcite ions are decomposed
during the combustion. Unlike other samples, the FTIR spectra of the original and postmortem of the
Barnett oil shale are almost same. During the combustion analysis only small amount of the organic matter
is decomposed and inorganic decomposition does not occur. This means that the weight loss (18.6%) is
only due to oxidation of the organic matter. Similarly, FTIR spectra of the Marcellus oil shale sample
before and after the combustion analyses are almost same such that only small amount of the carbonate
is decomposed during the combustion. However, this time weight loss (12.6%) is mostly due to the
inorganic decomposition; since, Marcellus oil shale sample has very little amount of organic matter (See
Table 2).

Conclusions
In this paper, four Turkish and American oil shale samples are characterized using XRD, SEM, FTIR and
TGA/DSC analyses. XRD analyses give information about the mineral contents of the samples. SEM is
used to detect minerals and some organic content in the samples in microscopic size. FTIR analyses help
to specify the bonds of the kerogen and inorganic matrix; additionally, with the combination of
TGA/DSC, the changes after pyrolysis and combustion analyses are identified. Finally, the analyses of
TGA/DSC show the pyrolysis - combustion behaviors and organic - inorganic amount of each sample. The
combination of these four techniques is found to be useful to characterize the oil shale samples and
following conclusions are drawn:
1. It is found that the combination of TGA/DSC and FTIR is useful to detect organic content and the
combination of XRD and FTIR to detect inorganic compounds.
2. It is seen that amount of organic decomposition is propoportional to the aliphatic content of the
kerogens. Himmetoglu oil shale shows highest organic decomposition during pyrolysis and
combustion due to high aliphatic content recognized from FTIR spectra.
3. When inorganic decompositions are analyzed, it is observed that the reasons for weight loss are
highly related with the carbonate ion from calcite or dolomite. Eagle Ford oil shale has highest
amount of calcite ion shows the highest inorganic decomposition.
4. SEM verifies the inorganic content of oil shales and gives clue about the depositional environ-
ments of some samples. Fresh water diatoms and algeas from Seyitomer and Himmetoglu oil
shales indicate marine depositional environment for these oil shale samples.

Acknowledgement
We acknowledge the financial support and the opportunity provided by the Texas A&M University to
conduct expreiments in the Ramey Thermal Recovery Laboratory. We also acknowledge the member of
Heavy Oil, Oil shales, Oil sands, & Carbonate Analysis and Recovery Methods (HOCAM) Research
Team at Texas A&M University, Petroleum Engineering Department, for their help.

References
Al-Harahsheh, M., Al-Ayed, O., Robinson, J., Kingman, S., Al-Harahsheh, A., Tarawneh, K., . . .
Barranco, R. (2011). Effect of demineralization and heating rate on the pyrolysis kinetics of
Jordanian oil shales. Fuel Processing Technology, 92(9), 1805–1811. doi: 10.1016/j.fu-
proc.2011.04.037
Al-Harahsheh, M., Kingman, S., Saeid, A., Robinson, J., Dimitrakis, G., & Alnawafleh, H. (2009).
12 SPE-172952-MS

Dielectric properties of Jordanian oil shales. Fuel Processing Technology, 90(10), 1259 –1264.
doi: 10.1016/j.fuproc.2009.06.012
Alstadt, K. N., Katti, D. R., & Katti, K. S. (2012). An in situ FTIR step-scan photoacoustic
investigation of kerogen and minerals in oil shale. Spectrochimica Acta. Part A, Molecular and
Biomolecular Spectroscopy, 89, 105–13. doi: 10.1016/j.saa.2011.10.078
Ballice, L. (2005). Effect of demineralization on yield and composition of the volatile products
evolved from temperature-programmed pyrolysis of Beypazari (Turkey) Oil Shale. Fuel Process-
ing Technology, 86(6), 673–690. doi: 10.1016/j.fuproc.2004.07.003
Ballice, L., Yüksel, M., Saglam, M., Schulz, H., & Hanoglu, C. (1995). Application of infrared
spectroscopy to the classification of kerogen types and the thermogravimetrically derived pyrol-
ysis kinetics of oil shales. Fuel, 74(11), 1618 –1623. doi: 10.1016/0016-2361(95)00093-K
Bhargava, S., Awaja, F., & Subasinghe, N. (2005). Characterisation of some Australian oil shale using
thermal, X-ray and IR techniques. Fuel, 84(6), 707–715. doi: 10.1016/j.fuel.2004.11.013
Biglarbigi, K., Mohan, H., & Killen, J. (2009). Oil Shale 1: US, World Possess Rich Resource Base
- Oil & Gas Journal.
Coates, J., & Ed, R. A. M. (2000). Interpretation of Infrared Spectra, A Practical Approach Interpre-
tation of Infrared Spectra, A Practical Approach, 10815–10837.
Downs, R. T. (2006). An Integrated Study of the Chemistry, Crystallography, Raman and Infrared
Spectroscopy of Minerals. In The RRUFF Project. Kobe, Japan: 19th General Meeting of the
International Mineralogical Association.
Fang-Fang, X., Ze, W., Wei-Gang, L., & Wen-Li, S. (2010). Study on Thermal Conversion of Huadian
Oil Shale Under N2 and Co2 Atmospheres. Oil Shale, 27(4), 309. doi: 10.3176/oil.2010.4.04
Güleç, K., & Önen, A. (1992). Turkish Oil Shales: Reserves, Characterization and Utilisation Studies.
In Eastern Oil Shale Symposium (pp. 12–24).
International Energy Agency. (2010). World Energy Outlook (p. 738). Paris.
Kok, M. V., Senguler, I., Hufnagel, H., & Sonel, N. (2001). Thermal and geochemical investigation
of Seyitomer oil shale. Thermochimica Acta, 371(1-2), 111–119. doi: 10.1016/S0040-
6031(01)00415-4
Largeau, C., Derenne, S., Casadevall, E., Berkaloff, C., Corolleur, M., Lugardon, B., . . . Connan, J.
(1990). Occurrence and origin of “ultralaminar” structures in “amorphous” kerogens of various
source rocks and oil shales. Organic Geochemistry, 16(4-6), 889 –895. doi: 10.1016/0146-
6380(90)90125-J
Levy, J. H., & Stuart, W. I. (1984). Thermal Properties of Australian Oil Shales: Characterization by
Thermal Analysis and Infrared Spectrophotometry. In Symposium on Characterization and Chem-
istry of Oil Shales Presented before the Divisions of Fuel Chemistry and Petroleum Chemistry,
Inc. (pp. 61–70).
Pamir, M. R. (1997). Thermal Characterization of Turkish Oil Shales. Middle East Technical
University.
Qing, W., Baizhong, S., Aijuan, H., Jingru, B., & Shaohua, L. (2007). Pyrolysis characteristics of
Huadian oil shales. Oil Shale, 24(2), 147–157.
Razvigorova, M., Budinova, T., Tsyntsarski, B., Petrova, B., Ekinci, E., & Atakul, H. (2008). The
composition of acids in bitumen and in products from saponification of kerogen: Investigation of
their role as connecting kerogen and mineral matrix. International Journal of Coal Geology, 76(3),
243–249. doi: 10.1016/j.coal.2008.07.011
Rouxhet, P. G., Robin, P. L., & Nicaise, G. (1980). Characterization of kerogens and of their evolution
by infrared spectroscopy. In Kerogen: Insoluble Organic Matter from Sedimentary Rocks (pp.
165–190).
SPE-172952-MS 13

Thakur, D. S., & Nuttall, H. E. (1987). Kinetics of pyrolysis of Moroccan oil shale by thermogravi-
metry. Ind. Eng. Chem. Res., 26(1932), 1351–1356.
World Energy Council. (2010). 2010 Survey of Energy Resources (pp. 93–102). Retrieved from
http://www.worldenergy.org/documents/ser_2010_report_1.pdf
Yan, J., Jiang, X., Han, X., & Liu, J. (2013). A TG – FTIR investigation to the catalytic effect of
mineral matrix in oil shale on the pyrolysis and combustion of kerogen. Fuel, 104, 307–317. doi:
10.1016/j.fuel.2012.10.024
Yen, T. F. (1976). Science and technology of oil shale. Advances in Astronomy and Space Physics
(Vol. -1). Michigan, USA: Ann Arbor Science Publishers Inc. Retrieved from . . .ѠY⬙⬎http://
adsabs.harvard.edu/abs/1976aasp.book. . .ѠY
14 SPE-172952-MS

Appendix A

Figure 7—Bulk XRD results of oil shale samples


SPE-172952-MS 15

Appendix B

Figure 8 —TG and DTG Curves of the Combustion Analysis

You might also like