Evolution of Nanoporosity in Organic-Rich Shales During Thermal

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Fuel 129 (2014) 173–181

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Evolution of nanoporosity in organic-rich shales during thermal


maturation
Ji Chen a,b, Xianming Xiao a,⇑
a
State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, Guangzhou 510640, China
b
University of Chinese Academy of Sciences, Beijing 100049, China

h i g h l i g h t s

 Laboratory-matured shale samples were produced by an anhydrous pyrolysis experiment.


 Low-pressure gas adsorption was used to characterize the pore structure of laboratory-matured shales.
 There are substantial differences in evolution of nanoporosity between organic-rich and organic-poor samples.
 Evolution of organic matter-hosted nanopores of gas shales can be roughly divided into three stages.

a r t i c l e i n f o a b s t r a c t

Article history: Artificial shale samples with equivalent vitrinite reflectance values (VRo) ranging from 0.69% to 4.19%
Received 7 January 2014 were obtained from an anhydrous pyrolysis experiment. Microporous and mesoporous characteristics
Received in revised form 18 March 2014 of these samples were investigated by low-pressure nitrogen and carbon dioxide adsorption techniques.
Accepted 25 March 2014
The result shows that the nanoporosity (microporosity plus mesoporosity) increases with thermal
Available online 13 April 2014
maturity after the oil window stage, and this increase is attributed to the formation of porosity within
organic matter and/or mineral–organic matter groundmass, rather than in the pure clay minerals. By
Keywords:
combining the gas generation and porosity evolution of these shales, a general model for formation
Gas shale
Anhydrous pyrolysis
and development of the nanoporosity is proposed.
Thermal maturation Ó 2014 Elsevier Ltd. All rights reserved.
Nanoporosity evolution

1. Introduction pressure (MICP), and scanning electron microscopy (SEM) [5,13–


17]. The geological controls of pore structure in gas shales include
A gas shale reservoir is characterized by abundant pores having the total organic carbon (TOC) content, thermal maturity and
sizes in a range of several to several hundreds of nanometers [1–8]. mineralogy, which have preliminarily been discussed in previous
Pores are subdivided into micropores (pore width <2 nm), works [3,4,10,18,19].
mesopores (pore width between 2–50 nm) and macropores (pore Effect of thermal maturation on porosity has recently attracted
width >50 nm) according to the classification of International more attention because of the growing recognition that porosity in
Union of Pure and Applied Chemistry (IUPAC) [9]. It has been found organic matter, to a large extent, is a function of thermal maturity
that a considerable part of gas occurs in an adsorbed state within [4,6,18,19]. The recent efforts toward understanding porosity
micro- and mesopores of gas shales [3,4,10–12]. Elucidating the evolution with maturation have been matched with documenting
complex pore networks in gas shales has become a strategic the pore networks in gas shales. For example, Curtis et al. [20] used
subject because many studies have shown that shale pore struc- focused ion beam milling combined with scanning electron
ture is one of the most important factors controlling gas capacity microscopy (FIB–SEM) techniques to investigate the organic poros-
[1–4,10–12]. Several measurement techniques have successfully ity within Woodford Shale samples with vitrinite reflectance
been developed to characterize the complex pore systems such values (VRo) ranging from 0.51% to 6.36%, finding that secondary
as small-angle and ultra small angle neutron scattering (SANS/ organic porosity is absent in samples with VRo values less than
USANS), low-pressure gas adsorption, mercury injection capillary 0.90%, but does occur in samples with VRo values greater than
1.23% except for a 2.0% VRo sample. The exception suggested that
maturity alone is not a reliable predictor of porosity in organic
⇑ Corresponding author. Tel.: +86 20 85290176; fax: +86 20 85290706. matter and other factors such as organic matter composition
E-mail address: xmxiao@gig.ac.cn (X. Xiao).

http://dx.doi.org/10.1016/j.fuel.2014.03.058
0016-2361/Ó 2014 Elsevier Ltd. All rights reserved.
174 J. Chen, X. Xiao / Fuel 129 (2014) 173–181

complicate this predication. While SEM observations of organic- 2.3. Organic geochemistry and petrology analysis
rich mudstones from the Upper Jurassic Kimmeridge Clay Forma-
tion indicate that there are no differences of organic pore shape The total organic carbon (TOC) of the three initial shales was
and size between immature and mature samples, and the minimal measured by a Leco CS230 Carbon/Sulfur analyzer after the sam-
preservation of organic pores during hydrocarbon generation is ples were treated by hydrochloric acid to remove the carbonates.
attributed to the highly ductile nature of clays and organic materi- The Rock–Eval pyrolysis was conducted using a Rock–Eval 6 Turbo
als [21]. Mastalerz et al. [22] examined the evolution of porosity analyzer to assess their hydrocarbon generation potential. Stan-
with maturation in a suit of five New Albany Shale samples span- dard IFP55000 was applied to calibrate the instrument before
ning a maturity range from immature (VRo = 0.35%) to postmature and after sample analyses.
(VRo = 1.41%), observing that total porosity and total pore volume A 3Y-Leica DMR XP microscopy equipped with a microphotom-
exhibit a significant decline from immature sample to late mature eter was used to identify the macerals of the initial shales and mea-
sample with a reduction of more than 80%, and then regain their sure the vitrinite reflectance (VRo) values of the coal samples
higher values in the post mature sample. After addressing the obtained from the different pyrolysis temperatures. According to
influence of TOC content and mineralogy on porosity, they further the maturities achieved from different pyrolysis temperatures, a
suggested that maturity appears to exert the dominant control standard with a reflectance closest to the measured value was
upon porosity development in shales. However, their data was lim- selected from three available standards: yttrium aluminum garnet
ited to shales with VRo value less than 1.41%. YAG-08-57 (Ro = 0.904%), NR1149 (Ro = 1.24%), and cubic zirconia
Therefore, response of shale porosity, especially organic (Ro = 3.11%). An oil objective of 50  /0.85 and a measuring fiber
porosity to hydrocarbon generation and cracking caused by ther- with a diameter of 0.6 mm were employed. VRo% for each sample
mal maturation remains puzzling in gas shale systems (especially was determined by averaging 50 measurements on vitrinite
highly-matured shales), additional systematic studies need to be particles.
carried out. In the present study, three suites of artificial shale
samples across a maturation gradient were produced by laboratory 2.4. Mineral composition analysis
anhydrous pyrolysis of three low maturity shale samples with dif-
ferent kerogen types or TOC contents at a wide range of tempera- X-ray diffraction (XRD) analysis of the shale powders (including
tures. The evolution of their micro- and mesoporous characteristics the three initial shales and some of their heated samples) was car-
with thermal maturation was evaluated by low pressure nitrogen ried out on a Bruker D8 Advance X-ray diffractometers at 40 kV
and carbon dioxide adsorption measurements. The relationship and 30 mA with a Cu Ka radiation (k = 1.5406 for Cu Ka1). Step-
between gas generation and nanopore formation is also discussed. wise scanning measurements were performed at a rate of 4°/min
in the range of 3–85° 2h. The relative mineral percentages were
estimated semi-quantitatively using the area under the curve for
2. Samples and experiments
the major peaks of each mineral with correction for Lorentz Polar-
ization [4].
2.1. Samples
2.5. Gas adsorption and data analysis
Three suites of shale samples covering a range of maturities
were produced by anhydrous pyrolysis of three low maturity
Nitrogen adsorption at 77 K is universally used to examine the
shales/mudstones (approximately 0.60% VRo). The starting samples
micro-, meso-, and macroporosity of solid materials if a wide range
used to be pyrolyzed originated from three different deposits across
of relative pressures (P/P0) were applied [26]. Carbon dioxide
China, including two organic-rich shales (samples LCG and DL) and
adsorption at 273.15 K is often the preferred choice to investigate
one organic-lean mudstone (sample EP). Sample LCG was collected
the microporosity of carbonaceous materials, since the ambient
from an Upper Permian outcrop (Yaomoshan, Urumqi) of the south-
temperature promote the gas diffusion in the microporous system
ern Junggar Basin, sample DL from an Upper Permian outcrop
compared to the low temperature (77 K) used in nitrogen adsorp-
(Changjianggou, Guangyuan) in the northern Sichuan Basin, and
tion [26]. However, a limitation of CO2 adsorption at 273.15 K is
Sample EP from an Oligocene deposit of well WC19-1M-1 in the
that CO2 molecular cannot fill larger micropores and also may have
western Pearl River Mouth Basin. Their locations were present in
additional interactions with organic matter [26,27]. With more
the Refs. [23–25]. A coal sample with a similar initial maturity
attention focused on the shale pore systems, nitrogen and carbon
(VRo = 0.56%) was also pyrolyzed under the same conditions as
dioxide adsorption techniques have widely been introduced to
the three shales to give a measure of the thermal maturity of the
study the pore characteristics of shales [5,13–17]. The experimen-
shale samples at different pyrolysis temperatures.
tal methods and parameter calculations may be briefly described
as follows.
2.2. Anhydrous pyrolysis Low-pressure gas adsorption measurements were conducted on
a Micromeritics ASAP 2020M surface area and porosimetry ana-
The anhydrous pyrolysis experiment was carried out under lyzer. Nitrogen adsorption and carbon dioxide adsorption were
inert conditions in a programmed muffle furnace. The four raw used to obtain information about meso- and micropores respec-
samples were first crushed to <250 lm, dried for 24 h at 60 °C in tively. The term ‘‘nanopore’’ or ‘‘nanoporosity’’ used in this article
a vacuum oven, and then each was separated into 10 fractions, referred in particular to the micropores and mesopores. Each sam-
transferred to a series of quartz tubes (with an inner diameter of ple was degassed at 110 °C for 16 h in a vacuum chamber prior to
30 mm) and sealed under vacuum condition. The four raw samples the analysis to remove its residual volatile material. Nitrogen iso-
were simultaneously heated in the furnace from room temperature therms were collected within a relative pressure (P/P0) range of
to a preset temperature at a heating rate of 200 °C/h, and then kept 0.01–0.995 at liquid nitrogen (77 K). The equivalent surface area
for 24 h. The target temperature was set within a range of 300– was calculated within a relative pressure range of 0.05–0.20 by
750 °C at 50 °C intervals. Every heating run was performed on a the multi-point BET (Brunauer–Emmett–Teller) equation [28]:
new batch of powdered samples. The pyrolysis products at each
1 1 C1 P
temperature point were recovered and were subject to an ongoing ¼ þ  ð1Þ
VðP0 =P  1Þ V m C V m C P0
multidisciplinary analytical program.
J. Chen, X. Xiao / Fuel 129 (2014) 173–181 175

where P is the gas pressure (Pa), P0 is the saturation pressure (Pa), V examined after 500 °C, which can be interpreted to the formation
is the volume of sorbed gas at equilibrium pressure (cm3/g, STP), Vm of amorphous metakaolinite caused by the dehydroxylation of kao-
is the volume of monolayer (cm3/g, STP), C is the BET constant. linite [32]. The dolomite in sample LCG disappears at 600 °C
The mesopore volume was calculated by the BJH (Barrett– because of its thermal decomposition [33]. The contents of quartz,
Joyner–Halenda) theory using the adsorption branch of nitrogen feldspar and illite–smectite show an irregular change with the
isotherms [29]. This classical pore size model is based on the Kelvin temperature. It is believed that this change is mainly related to
equation assuming a cylindrical pore and corrected for multilayer the heterogeneity of their sub-samples as well as the above decom-
adsorption. position or dehydroxylation which affects their percentage.
Carbon dioxide adsorption isotherms were collected within a
relative pressure range of 105–3.0  102 at ice/water bath
3.2. Pyrolysis temperature and vitrinite reflectance
(273.15 K) [30]. Micropore surface area was calculated by the
well-known Dubinin–Radshkevich (D–R) equation [31], which is
The vitrinite reflectance values of the coal samples at different
based on the theory of volume filling of micropores (TVFM) and fol-
pyrolysis temperatures are presented in Table 3. They show an
lows the fundamental relation as below:
approximately linear correlation. The sub-samples of samples
 2  2 LCG and DL range from 150 to 200 mg at each temperature point,
RT P0
ln V ¼ ln V 0  ln ð2Þ which make difficulty to prepare their polished blocks to measure
E P
the vitrinite reflectance. It is generally accepted that the thermal
where V is the volume of sorbed gas at equilibrium pressure (cm3/g, maturity of sedimentary organic matter or the rank of coal depends
STP), V0 is the micropore capacity (cm3/g, STP), R is the universal gas mainly on its temperature history [34–36]. The famous tempera-
constant (8.314 J mol1 K1), T is the analysis temperature (K), E is ture and time index model [34,35] and the widely used EASY%Ro
the characteristic energy (kJ mol1). model [37] quantified the relationships between maturity (VRo),
The micropore volume was determined by the Dubinin–Astak- temperature and time. According to these models, different
hov (D–A) equation based on the CO2 isotherms [31]. They intro- sedimentary organic matter (shale or coal) will achieve the same
duced the structural factor n (Astakhov index) to expand the D–R maturity if they experience the same thermal history (temperature
equation on heterogeneous adsorbents. This equation can be and duration). The shale samples have a similar initial maturity
defined as: with the coal sample and the pyrolysis experiment for each sample
 n  n and temperature point was carried out in the same batch, thus
RT P0
ln V ¼ ln V 0  ln ð3Þ the VRo values of the coal sub-samples can be used to characterize
E P the maturities of the shale sub-samples obtained from the
same temperatures.
3. Results and discussion
3.3. Adsorption isotherms
3.1. Organic geochemistry and mineralogy
The nitrogen isotherms for the three suits of samples are shown
The organic geochemistry data of the three initial shales are in Figs. 2–4, which are corresponding to type IV isotherms as
shown in Table 1. Although the three shales have a similar matu- defined by IUPAC [9]. Such type of isotherm is characterized by
rity with VRo range from 0.56% to 0.62%, their kerogen types are its hysteresis loop, which is associated with capillary condensation
different and the TOC contents vary widely. Sample LCG has a taking place in mesopore structures [9]. The shape of hysteresis
TOC content of as high as 29.1% with a HI (hydrogen index) of loop is close to type H3 [9]. Although the effect of various factors
839 mg/g TOC (Table 1). The shale is rich in lamalginite and poor on hysteresis loop is not fully understood, the type H3 loop is usu-
in vitrinite and inertinite (Fig. 1). The TOC content and HI of sample ally thought to be associated with aggregates of plate-like particles
DL are 8.63% and 278 mg/g TOC (Table 1), respectively, and it is giving rise to slit-shaped pores based on the interpretation of
characterized by liptodetrinites and telalginites with a few vitrinite IUPAC [9]. Nevertheless, this interpretation is subject to error in
and inertinite (Fig. 1). These data suggest that samples LCG and DL the case of natural materials such as gas shales [16,17]. Many
have kerogens classified as type I and type IIb respectively. Sample SEM observations have repeatedly shown that pores in organic
EP is poor in organic matter with a TOC content of 0.18% (Table 1). matter of shales are not slit-like but bubble- or foam-like
Table 2 presents the mineralogical data of the three initial [5,6,20,21]. SANS/USANS data from Clarkson et al. also suggested
shales and their sub-samples at some selected pyrolysis tempera- that the assumption of slit-shaped pores inferred from hysteresis
tures. The initial samples LCG and DL have similar amounts of brit- loop shape was incompatible with SANS scattering results [14].
tle minerals and clays, but the former is dominated by quartz plus For the nitrogen amounts adsorbed at P/P0 around 0.995, sample
feldspar (total 66%) and mixed illite–smectite (20.1%), the latter by LCG shows a decrease from 6.50 cm3/g at 0.69% VRo to 1.67 cm3/
Quartz (60.8%) and mixed illite–smectite (18.9%). The initial sam- g at VRo of 0.89%, and then turns to increase to 22.50 cm3/g at
ple EP has a clay content of 56.8%, including 37.9% kaolinite and 4.19% VRo (Fig. 2). Sample DL shows an increase from 12.67 cm3/
18.9% mixed illite–smectite. g at 0.69% VRo to 36.75 cm3/g at 4.19% VRo (Fig. 3), whereas sam-
During the pyrolysis, the clay mineral types of samples LCG and ple EP shows a relatively stable value of about 22 cm3/g before
DL are not changed, while the kaolinite of sample EP was not 1.95% VRo and then a sharp decline with further increasing

Table 1
Geochemical data of the three initial shales.

Sample Age VRo (%) TOC (wt.%) S1 (mg/g) S2 (mg/g) Tmax (°C) HI (mg/g TOC)
LCG Upper permian 0.61 29.1 2.92 230.4 435 839
DL Upper permian 0.62 8.63 1.24 25.4 437 278
EP Oligocene 0.56a 0.18 / / / /
a
Data from organic-rich mudstone at similar depth.
176 J. Chen, X. Xiao / Fuel 129 (2014) 173–181

Fig. 1. Photomicrographs (oil immersion) of samples LCG and DL before anhydrous pyrolysis taken under the fluorescent light.

Table 2
Mineralogical compositions (in wt.%) of three initial shales and their sub-samples at selected pyrolysis temperatures.

Sample Quartz (wt.%) Feldspar (wt.%) Calcite (wt.%) Dolomite (wt.%) Illite–smectite (wt.%) Kaolinite (wt.%) Pyrite (wt.%)
LCG-initial 37.8 28.2 0 13.9 20.1 0 0
LCG-300 °C 35.5 23.8 0 11.0 29.7 0 0
LCG-400 °C 39.6 22.9 0 13.7 23.8 0 0
LCG-500 °C 40.8 27.9 0 12.3 19.0 0 0
LCG-600 °C 43.8 29.4 0 0 26.8 0 0
LCG-700 °C 44.4 30.8 0 0 24.8 0 0
DL-initial 60.8 0 13.5 0 18.9 0 6.8
DL-300 °C 61.2 0 16.1 0 22.7 0 0
DL-400 °C 65.1 0 17.3 0 17.6 0 0
DL-500 °C 66.5 0 14.8 0 18.7 0 0
DL-600 °C 63.8 0 16.1 0 20.1 0 0
DL-700 °C 67.7 0 14.0 0 18.3 0 0
EP-initial 28.8 14.4 0 0 18.9 37.9 0
EP-300 °C 25.1 17.0 0 0 30.8 27.1 0
EP-400 °C 27.8 17.6 0 0 41.4 13.2 0
EP-500 °C 30.1 13.0 0 0 45.4 11.5 0
EP-600 °C 31.1 21.7 0 0 47.2 0 0
EP-700 °C 30.0 23.6 0 0 46.4 0 0

Table 3 3.4. Response of pore structure parameters to thermal maturity


Vitrinite reflectance values of coal samples obtained from different pyrolysis
temperatures.
Specific surface area and pore volume for the three suites of
Temperature (°C) Vitrinite reflectance VRo (%) samples are shown in Fig. 6. For sample LCG (Fig. 6A1 and A2),
300 0.69 the specific surface area and volume of its micro- and mesopores
350 0.89 pass through a minimum value around the oil window
400 1.25 (VRo = 0.89%). After the oil window, its micropore surface area
450 1.51
and volume as well as its mesopore surface area increase to a max-
500 1.95
550 2.60 imum at 3.48% VRo, and then decrease with further increase of
600 3.02 maturity. But its mesopore volume increases up to the maximum
650 3.48 maturity achieved in this study. Compared with sample LCG, sam-
700 3.89 ple DL has similar variations of the four parameters with increasing
750 4.19
maturity except for the lack of minimum value around the oil win-
dow, in which range there is a slight increase (Fig. 6B1 and B2).
Valenza et al. [18] reported the geochemical controls on micro-
structure of North American shales having maturities from approx-
maturity from 19.41 cm3/g at 2.60% VRo to 5.69 cm3/g at 4.19% VRo imately 0.5% to 2.7% VRo/VRb (vitrinite or bitumen reflectance),
(Fig. 4). also observing a marked increase in surface area and pore volume
Fig. 5 gives an overview on the carbon dioxide isotherms, which with increasing maturity. In the case of sample LCG, its evolution of
are close to type I isotherm given by microporous solids [9]. The porosity indicates that porosity can undergo a reduction to some
limiting uptake of these isotherms with increasing relative pres- extent in the oil window stage, which is different from a monoton-
sure is governed by the accessible micropore volume rather than ically growing trend as indicated by previous studies [4,18,19]. A
the internal surface area [9]. The carbon dioxide adsorption capac- similar reduction of pore volume or porosity around the matura-
ity of samples LCG and DL at P/P0 around 0.03 shows an increase tion stage is also observed in naturally buried samples such as
with increasing maturity after 0.89% VRo, with a maximum of gas shales and coals, interpreted to be the result of bitumen swell-
6.84 cm3/g and 2.60 cm3/g respectively at VRo of 3.48%, and then ing and compaction [22,38,39]. Effect of pore clogging by bitumen
a decrease with further increasing maturity. However, the carbon has been assessed by Valenza et al. via comparing surface areas and
dioxide adsorption capacity of sample EP gradually decreases from micropore volumes before and after solvent extraction [18]. They
0.71 cm3/g (VRo = 0.69%) to 0.22 cm3/g (VRo = 4.19%), which is observed a largely consistent rise in surface area and micropore
consistent with its nitrogen adsorption characteristics. volume after the solvent extraction, especially in the oil window.
J. Chen, X. Xiao / Fuel 129 (2014) 173–181 177

Fig. 2. Nitrogen isotherms for sample LCG with different maturities.

Fig. 3. Nitrogen isotherms for sample DL with different maturities.

Bernard et al. [40] documented the chemical evolution of Missis- generate a relatively small amount of bitumen, thus the influence
sippian Barnett Shale samples from Fort Worth Basin ranging from on the nanopores would be small.
immature to postmature, suggesting that kerogen exhibits a trend A further examination of the pore volume of samples LCG and
of aromatization, demethylation and deoxygenation during matu- DL reveals that micropore volume exhibits a noticeable decrease
ration. In contrast, bitumen generated in the oil window stage with VRo > 3.48%, while the mesopore volume and total volume
appears aromatic-poor, aliphatic-rich and oxygen-rich, which is (micropore volume plus mesopore volume) still increase with the
likely to fill in intra- and intergranular spaces, thereby reducing maturity (Fig. 6A2 and B2). This phenomenon implicates the trans-
the porosity [40]. Thus, the difference of the nanopore evolution formation of micropore to mesopore and expansion of the meso-
between sample LCG and sample DL in the oil window stage can porous pore size with further increasing maturity. In general,
be referred to their oil generation potentials. Sample LCG contains there is an inverse relationship between pore size and specific sur-
oil-prone kerogen which could generate a large amount of face area for a given pore volume [4,5], thus this transformation
asphaltene-rich bitumen, leading to a pore occlusion in the oil win- and expansion also result in a reduction of micro- and mesopore
dow stage. Sample DL contains gas-prone kerogen, which could surface areas (Fig. 6A1 and B1).
178 J. Chen, X. Xiao / Fuel 129 (2014) 173–181

Fig. 4. Nitrogen isotherms for sample EP with different maturities.

micro- and mesopores present a monotonically decreasing trend


with increasing maturity (Fig. 6C1 and C2). As compared to the ini-
tial sample, the micro- and mesopore volumes reduce by 86.7% and
83.8% respectively, and its micro- and mesopore surface areas
reduce by 50.1% and 71.3% respectively, when it is matured to
the VRo value of 4.19 %. The result is in agreement with the heating
treatment of clay minerals [41,42]. Noyan et al. [42] observed an
overall trend of decline in both specific surface areas and micro-
mesopore volumes with increasing heating temperature from100
to 900 °C for Hancili bentonite from Turkey. In consideration of
the high content of clay minerals (56.8%), but a very small amount
of organic matter (TOC = 0.18%) in sample EP, it is believed that the
evolution of its nanopore structure is primarily controlled by
changes in the clay minerals. Diagenesis of clay minerals mainly
includes the dehydration and microstructural change (such as
reduction of interplanar spacing) [43,44], resulting in an obvious
decrease of the porosity [41,42], as well as the transformation of
mineral type (typically smectite to illite through a mixed-layer
illite–smectite) [45–48]. This transformation of clay mineral types
exists in normal geological environments [45,46] and also happens
under strictly controlled laboratory conditions [47,48]. However, it
does not occur under the experimental conditions of this study
(Table 2). The trends exhibited by sample EP can be ascribed to
the dehydration and framework collapse of clay minerals during
the process of thermal maturation [41–44].
Types and contents of quartz can also affect the pore volumes in
shales [10,17,49,50]. Especially for quartz of biogenic origin, it is
produced by precipitation during diagenesis with silica derived
from dissolution of opal if the geofluids are active, which may also
change the pore volume and structure in shales [10,51–53].
Although types of quartz were not examined for the studied shales,
any quartz should be stable under the anhydrous pyrolysis condi-
Fig. 5. Carbon dioxide isotherms for three suits of shale samples with different tion of the present study. From Table 2, the quartz contents vary
maturities. irregularly within a narrow range with increasing pyrolysis tem-
perature, indicating that quartz has not a significant effect on the
nanoporosity evolution of the studied shale samples.
In order to evaluate the effect of thermal maturation on nano- These observations demonstrate two evolutionary routes of
pore structure of clay minerals, sample EP containing rare organic nanopore structure in these shales under the present study condi-
matter (TOC = 0.18%) was pyrolyzed under the same conditions as tions: increase of nanoporosity in the organic matter (regardless of
the two organic-rich shales. The surface areas and volumes of its the possible reduction taking place in the oil window stage) and
J. Chen, X. Xiao / Fuel 129 (2014) 173–181 179

Fig. 6. Surface areas and pore volumes of micro- and mesopores, as functions of thermal maturity as indicated by vitrinite reflectance (VRo). A1, A2: sample LCG; B1, B2:
sample DL; C1, C2: sample EP.

reduction of nanoporosity in the clay minerals. This opposite evo- and residual kerogen are nanoporous due to the expulsion process
lutionary tendency will largely affect how the nanopore structure of hydrocarbon gas during oil cracking, which make a contribution
of the shales as a whole develops with thermal maturity. Thus, it to the total porosity of shales [6,19,40,50]. Therefore, naoporosity
can be concluded that an increase in shale porosity with thermal of oil-prone shales, in this stage, will typically shows a decrease,
maturation is mostly associated with changes in the porosity of and then an increase.
organic matter, which is consistent with previous studies for
natural shale samples [1–4,18,22,54], and evolution of the OM- (2) Developing stage of OM-hosted nanopores (2.0% < VRo < 3.5%)
hosted nanopores will become an important factor to influence
gas storage capacity of shales with very high maturity. In this stage, methane is generated further from kerogen and
pyrobitumen, and the secondary cracking of heavy hydrocarbon
3.5. Evolutionary characteristics of nanoporosity in shales gases (C2–5) [56,57], with more matured residual kerogen and
pyrobitumen. These solid organic matters will gradually evolve
Gas generation of organic matter in shales mainly includes the towards the development of a graphitic-like structure with a large
three pathways [55]: (1) kerogen cracking to asphaltene and gas; number of nanopores, leading to a rapid increase of the pore
(2) asphaltene cracking to oil and gas; (3) oil cracking to gas and volume and specific surface area of micro- and mesopores.
pyrobitumen. Gas generation is always accompanied by the forma-
tion and transformation of porosity. Based on the results of the (3) Conversion and destruction stage of OM-hosted nanopores
present study, the evolution of OM-hosted nanoporosity of gas (VRo > 3.5%)
shales can be roughly divided into three stages.
A large amount of nitrogen will be generated through further
(1) Formation stage of OM-hosted nanopores (0.6% < VRo < 2.0%) decomposition of solid organic matter in shales at very high matu-
rity with further increasing temperature and pressure, and trans-
In this stage, oil is generated from kerogen and the generated oil formation and destruction of nanopores will occur. Under the
is cracked to gas, with the formation of organic porosity in shales. experimental conditions of the present study, the nanopores are
Primary and secondary migration of the generated oils will occur in enlarged after VRo > 3.5%, with a decrease of microporosity and
the oil window stage, with residual oil rich in asphaltene and resin, an increase of mesoporosity, which can be attributed to the trans-
filling in intra- and intergranular spaces, resulting in a decrease of formation of micropores to mesopores and/or mesopores to
shale porosity. At higher maturation, pyrobitumen is produced by macropores. However, organic porosity may also be affected by
oil cracking to gaseous hydrocarbons [56,57]. This pyrobitumen post-compaction under geological conditions. Using field-emission
180 J. Chen, X. Xiao / Fuel 129 (2014) 173–181

scanning electron microscopy (FE–SEM) techniques on a suit of application of an integrated formation evaluation. AAPG Bull
2008;92(1):87–125.
Marcellus Shale samples from northern Pennsylvania ranging from
[3] Chalmers GRL, Bustin RM. The organic matter distribution and methane
VRo 1.0 to 2.1%, Milliken et al. [54] found that the preservation of capacity of the Lower Cretaceous strata of Northeastern British Columbia,
organic porosity is related to the content of organic matter, and a Canada. Int J Coal Geol 2007;70:223–39.
reduction of organic nanoporosity occurs by compaction when [4] Chalmers GRL, Bustin RM. Lower Cretaceous gas shales in Northeastern British
Columbia, Part I: geological controls on methane sorption capacity. Bull Can
the organic matter content exceeds a certain value, for instance Pet Geol 2008;56(1):1–21.
TOC > 5.5% for Marcellus Shale. [5] Chalmers GRL, Bustin RM, Power IM. Characterization of gas shale pore
It should be noted that the proposed pattern is a general model, systems by porosimetry, pycnometry, surface area, and field emission
scanning electron microscopy/transmission electron microscopy image
and it should be celebrated by substantial natural shale samples. analyses: examples from the Barnett, Woodford, Haynesville, Marcellus, and
Many factors such as types, quantities and occurrence of organic Doig units. AAPG Bull 2012;96(6):1099–119.
matter in shales, variations in mineral components and composi- [6] Loucks RG, Reed RM, Ruppel SC, Jarvie DM. Morphology, genesis, and
distribution of nanometer-scale pores in siliceous mudstones of the
tions, as well as features of geothermal fluids may, on further Mississippian Barnett Shale. J Sediment Res 2009;79:848–61.
investigation, result in significant differences in nanoporosity [7] Loucks RG, Reed RM, Ruppel SC, Hammes U. Spectrum of pore types and
evolution with increasing maturity. networks in mudrocks and a descriptive classification for matrix-related
mudrock pores. AAPG Bull 2012;96:1071–98.
[8] Nelson PH. Pore-throat sizes in sandstones, tight sandstones, and shales. AAPG
Bull 2009;93:329–40.
4. Conclusions [9] Sing KSW, Everett DH, Haul RAW, Moscou L, Pierotti RA, Rouquerol J, et al.
Reporting physisorption data for gas/solid systems with special reference to
the determination of surface area and porosity. Pure Appl Chem
Three suites of artificially matured shale samples spanning a 1985;57:603–19.
range of 0.69–4.19% VRo were obtained from anhydrous pyrolysis [10] Ross DJK, Bustin RM. The importance of shale composition and pore structure
experiments, and their microporosity and mesoporosity were upon gas storage potential of shale gas reservoirs. Mar Pet Geol
2009;26:916–27.
investigated by low pressure nitrogen and carbon dioxide adsorp- [11] Rexer TFT, Benham MJ, Aplin AC, Thomas KM. Methane adsorption on shale
tion. The following conclusions were obtained. under simulated geological temperature and pressure conditions. Energy Fuels
2013;27:3099–109.
[12] Mosher K, He J, Liu Y, Rupp E, Wilcox J. Molecular simulation of methane
(1) The two organic-rich shales have a substantially similar adsorption in micro- and mesoporous carbons with applications to coal and
evolutionary tendency in their micropore and mesopore gas shale systems. Int J Coal Geol 2013;109:36–44.
characteristics. Their micropore volumes, micro- and meso- [13] Mastalerz M, He L, Melnichenko YB, Rupp JA. Porosity of coal and shale:
Insights from gas adsorption and SANS/USANS techniques. Energy Fuels
pore specific surface areas show an increase within the
2012;26:5109–20.
maturity range from the oil window to about 3.50% VRo, [14] Clarkson CR, Freeman M, He L, Agamalian M, Melnichenko YB, Mastalerz M,
and then turns to a slight decrease with further increasing et al. Characterization of tight gas reservoir pore structure using USANS/SANS
and gas adsorption analysis. Fuel 2012;95:371–85.
maturity, but their mesopore volume shows an increase
[15] Clarkson CR, Solano N, Bustin RM, Bustin AMM, Chalmers GRL, He L, et al.
after 3.50% VRo up to the maturity (VRo = 4.19%) achieved Pore structure characterization of North American shale gas reservoirs
by the pyrolysis experiment. using USANS/SANS, gas adsorption, and mercury intrusion. Fuel 2013;103:
(2) The micro- and mesopore specific surface areas and pore 606–16.
[16] Schmitt M, Fernandes CP, da Cunha Neto JAB, Wolf FG, dos Santos VSS.
volumes of the organic-poor shale show a simple decrease Characterization of pore systems in seal rocks using nitrogen gas adsorption
with maturity. From 0.69% to 4.19% of VRo, the reduction combined with mercury injection capillary pressure techniques. Mar Pet Geol
is 86.7% and 83.8% respectively for the micro- and mesopore 2013;39:138–49.
[17] Tian H, Pan L, Xiao X, Wilkins RWT, Meng Z, Huang B. A preliminary study on
volumes, and 50.1% and 71.3% respectively for its micro- and the pore characterization of Lower Silurian black shales in the Chuandong
mesopore surface areas. Thrust Fold Belt, southwestern China using low pressure N2 adsorption and
(3) The increase of the nanoporosity in organic-rich shales dur- FE–SEM methods. Mar Pet Geol 2013;48:8–19.
[18] Valenza JJ, Drenzek N, Marques F, Pagels M, Mastalerz M. Geochemical
ing thermal maturation is mainly attributed to the formation controls on shale microstructure. Geology 2013;41:611–4.
of a large number of nanopores within organic matter. The [19] Modica CJ, Lapierre SG. Estimation of kerogen porosity in source rocks as a
formation and evolution of the OM-hosted nanopores in function of thermal transformation: example from the Mowry Shale in the
Powder River Basin of Wyoming. AAPG Bull 2012;96:87–108.
shales can be roughly subdivided into three stages: forma-
[20] Curtis ME, Cardott BJ, Sondergeld CH, Rai CS. Development of organic porosity
tion stage (0.60% < VRo < 2.0%), development stage in the Woodford Shale with increasing thermal maturity. Int J Coal Geol
(2.0% < VRo < 3.5%), and conversion or destruction stage 2012;103:26–31.
[21] Fishman NS, Hackley PC, Lowers HA, Hill RJ, Egenhoff SO, Eberl DD, et al. The
(VRo > 3.5%).
nature of porosity in organic-rich mudstones of the Upper Jurassic
Kimmeridge Clay Formation, North Sea, offshore United Kingdom. Int J Coal
Acknowledgements Geol 2012;103:32–50.
[22] Mastalerz M, Schimmelmann A, Drobniak A, Chen Y. Porosity of Devonian and
Mississippian New Albany Shale across a maturation gradient: insights from
The authors are indebted to three anonymous reviewers for organic petrology, gas adsorption, and mercury intrusion. AAPG Bull
their insightful comments and suggestions that have significantly 2013;97:1621–43.
[23] Carroll AR. Upper Permian lacustrine organic facies evolution, Southern
improved the manuscript, and to Dr. Wilkins, R.W.T. for improving Junggar Basin, NW China. Org Geochem 1998;28(11):649–67.
the English. This study was jointly supported by National Key Basic [24] Zhou Q, Xiao XM, Tian H, Wilkins RWT. Oil charge history of bitumens of
Research Program of China (973 Program, No. 2012CB214705), differing maturities in exhumed Palaeozoic reservoir rocks at Tianjingshan,
NW Sichuan Basin, Southern China. J Pet Geol 2013;36(4):363–82.
Chinese Academy of Sciences (XDB10040300) and a Special Found-
[25] Cheng P, Tian H, Huang BJ, Wilkins RWT, Xiao XM. Tracing early-charged oils
ing from the State Major Research Programme of China (No. and exploration directions for the Wenchang A sag, western Pearl River Mouth
2011ZX05008-002-40). Basin, offshore South China Sea. Org Geochem 2013;61:15–26.
[26] Groen JC, Peffer LAA, Perez-Ramirez J. Pore size determination in modified
micro- and mesoporous materials. Pitfalls and limitations in gas adsorption
data analysis. Microporous Mesoporous Mater 2003;60:1–17.
References [27] Feng B, Bhatia SK. Variation of the pore structure of coal chars during
gasification. Carbon 2003;41:507–23.
[1] Ross DJK, Bustin RM. Shale gas potential of the Lower Jurassic Gordondale [28] Brunauer S, Emmett PH, Teller EJ. Adsorption of gases in multimolecular
member, Northeastern British Columbia, Canada. Bull Can Pet Geol layers. J Am Chem Soc 1938;60:309–19.
2007;55:51–75. [29] Barrett EP, Joyner LG, Halenda PP. The determination of pore volume and area
[2] Ross DJK, Bustin RM. Characterizing the shale gas resource potential of distributions in porous substances. I. computations from nitrogen isotherms. J
Devonian–Mississippian strata in the Western Canada sedimentary basin: Am Chem Soc 1951;73(1):373–80.
J. Chen, X. Xiao / Fuel 129 (2014) 173–181 181

[30] Ghosal R, Smith DM. Micropore characterization using the Dubinin–Astakhov [44] Bala P, Samantaray BK, Srivastava SK. Dehydration transformation in Ca-
equation to analyze high pressure CO2 (273 K) adsorption data. J Porous Mater montmorillonite. Bull Mater Sci 2000;23(1):61–7.
1996;3:247–55. [45] Lynch FL. Frio shale mineralogy and the stoichiometry of the smectite-to-illite
[31] Dubinin MM, Stoeckli HF. Homogeneous and heterogeneous micropore reaction: the most important reaction in clastic sedimentary diagenesis. Clays
structures in carbonaceous adsorbents. J Colloid Interface Sci Clay Miner 1997;45:618–31.
1980;75(1):34–42. [46] Van de Kamp PC. Smectite-illite-muscovite transformations, quartz
[32] Gasparini E, Tarantino SC, Ghigna P, Riccardi MP, Cedillo-González EI, Siligardi dissolution, and silica release in shales. Clays Clay Miner 2008;56:66–81.
C, et al. Thermal dehydroxylation of kaolinite under isothermal conditions. [47] Eberl DD, Velde B, McCormick T. Synthesis of illite–smectite from smectite at
Appl Clay Sci 2013;80–81:417–25. earth surface temperature and high pH. Clay Miner 1993;28:49–60.
[33] Rodriguez-Navarro C, Kudlacz K, Ruiz-Agudo E. The mechanism of thermal [48] Metwally YM, Chesnokov EM. Clay mineral transformation as a major source
decomposition of dolomite: new insights from 2D-XRD and TEM analyses. Am for authigenic quartz in thermo-mature gas shale. Appl Clay Sci
Miner 2012;97:38–51. 2012;55:138–50.
[34] Lopatin NV. Temperature and geologic time as factors in coalification. Akad [49] Chalmers GRL, Ross DK, Bustin RM. Geological controls on matrix permeability
Nauk SSSR Izv Ser Geol 1971;3:95–106. of Devonian Gas Shales in the Horn River and Liard basins, Northeastern
[35] Waples DW. Time and temperature in petroleum formation: application of British Columbia, Canada. Int J Coal Geol 2012;103:120–31.
Lopatin’s method to petroleum exploration. AAPG Bull 1980;64(6):916–26. [50] Chalmers GRL, Bustin RM. Geological evaluation of Halfway–Doig–Montney
[36] Goff JC. Hydrocarbon generation and migration from Jurassic source rock in hybrid gas shale-tight gas reservoir, northeastern British Columbia. Mar Pet
East Shetland basin and Viking Graben of the northern North Sea. J Geol Soc Geol 2012;38:53–72.
(London, UK) 1983;140:445–74. [51] DeMaster DJ. The diagenesis of biogenic silica: chemical transformations
[37] Sweeney JJ, Burnham AK. Evaluation of a simple model of vitrinite reflectance occurring in the water column, seabed, and crust. Treatise Geochem
based on chemical kinetics. AAPG Bull 1990;74(10):1559–70. 2003;7:87–98.
[38] Mathia E, Rexer T, Bowen L, Aplin A. Evolution of porosity and pore systems in [52] Volpi V, Camerlenghi A, Hillenbrand CD, Rebesco M, Ivaldi R. Effects of
organic-rich Posidonia and Wealden Shales. In: 75th EAGE conference & biogenic silica on sediment compaction and slope stability on the Pacific
exhibition incorporating SPE EUROPEC 2013, London, UK; June 10–13, 2013. margin of the Antarctic Peninsula. Basin Res 2003;15:339–63.
[39] Prinz D, Littke R. Development of the micro- and ultramicroporous structure of [53] Blood R, Lash G, Bridges L. Biogenic silica in the Devonian Shale Succession of
coals with rank as deduced from the accessibility to water. Fuel the Appalachian Basin, USA. In: AAPG search and discovery article 50864
2005;84:1645–52. presented at AAPG 2013 annual convention and exhibition, Pittsburgh,
[40] Bernard S, Wirth R, Schreiber A, Schulz HM, Horsfield B. Formation of Pennsylvania; May 19–22, 2013.
nanoporous pyrobitumen residues during maturation of the Barnett Shale [54] Milliken KL, Rudnicki M, Awwiller DN, Zhang T. Organic matter-hosted pore
(Fort Worth Basin). Int J Coal Geol 2012;103:3–11. system, Marcellus Formation (Devonian), Pennsylvania. AAPG Bull
[41] Neaman A, Pelletier, Villieras F. The effects of exchanged cation, compression, 2013;97:177–200.
heating and hydration on textural properties of bulk bentonite and its [55] Jarvie DM, Hill RJ, Ruble TE, Pollastro RM. Unconventional shale-gas systems:
corresponding purified montmorillonite. Appl Clay Sci 2003;22:153–68. the Mississippian Barnett Shale of north-central Texas as one model for
[42] Noyan H, Onal M, Sarikaya Y. The effect of heating on the surface area, thermogenic shale-gas assessment. AAPG Bull 2007;91:475–99.
porosity and surface acidity of a bentonite. Clays Clay Miner 2006;54(3): [56] Hill RJ, Tang Y, Kaplan IR. Insights into oil cracking based on laboratory
375–81. experiments. Org Geochem 2003;34:1651–72.
[43] Bray Helen, Redfern SAT, Clark SM. The kinetics of dehydration in Ca- [57] Tian H, Xiao X, Wilkins RW, Tang Y. New insights into the volume and pressure
montmorillonite: an in situ X-ray diffraction study. Mineral Mag 1998;62(5): changes during the thermal cracking of oil to gas in reservoirs: implications for
647–56. the in situ accumulation of gas cracked from oils. AAPG Bull 2008;92:181–200.

You might also like