You are on page 1of 14

Computers & Geosciences 156 (2021) 104895

Contents lists available at ScienceDirect

Computers and Geosciences


journal homepage: www.elsevier.com/locate/cageo

Characterization of pore and grain size distributions in porous geological


samples – An image processing workflow
Hossein Safari a, Bruce J. Balcom b, Armin Afrough c, *
a
Reservoir Studies Division, Pars Petro Zagros Engineering & Services Company, 42 38th St, JahanAra Ave, Tehran, Iran
b
MRI Research Centre, University of New Brunswick, 8 Bailey Dr, Fredericton, E3B 9K6 NB, Canada
c
The Danish Hydrocarbon Research and Technology Centre, Technical University of Denmark, Kongens Lyngby 2800, Denmark

A R T I C L E I N F O A B S T R A C T

Keywords: An image processing workflow is presented for the characterization of pore and grain size distributions in porous
X-ray microcomputed tomography (μCT) geological samples from X-ray microcomputed tomography (μCT) and scanning electron microscopy (SEM)
Scanning electron microscopy (SEM) images. The pore and grain size distributions of five sandstone samples including Berea, Buff Berea, Nugget,
Pore size distribution
Castlegate, and Bentheimer, and one carbonate sample, Indiana limestone, are extracted using the proposed
Grain size distribution
Image processing
workflow. Two-dimensional size distributions acquired from SEM images were found to be biased toward smaller
Rocks sizes misrepresenting the actual 3D distributions. Stereological techniques unfolded the measured 2D size dis­
Porous materials tributions from SEM images to 3D distributions comparable with μCT results. While larger pores and grains can
easily be detected from μCT and SEM images, the quantification of small-scale heterogeneities is severely
influenced by their limits of resolution. We show that microstructural details resolved by SEM can significantly
impact the pore and grain size distributions in sandstone and carbonate rock samples. For example, SEM-resolved
microporosities in Indiana limestone result in bimodal distributions of pore and grain sizes, whereas μCT ob­
servations exhibit unimodal distributions. The acquired images and processed results are openly available and
may be used by researchers investigating image processing, magnetic resonance relaxation or fluid flow simu­
lations in natural rocks. The proposed methodology can be implemented to process μCT and SEM images of
natural rocks as well as other types of porous materials.

1. Introduction on macroscopic properties has been studied in a broad range of physical


processes, for example in the multiphase flow in porous media (Yang
The characterization of pore and particle/crystal/grain size distri­ et al., 2016), permeability estimation in hydrate-bearing sediments
butions in porous materials has a broad spectrum of applications in (Minagawa et al., 2009), phase transition in nanoconfined fluids (Luo
geosciences (Berger et al., 2011; Ghasemi et al., 2018; Roostaei et al., et al., 2019), propagation of wormholes in reactive fluid transport
2020; Srisutthiyakorn and Mavko, 2019), sedimentology (Gaafar et al., (Dubetz et al., 2016), compressive strength of cementitious materials
2014; Palavecino et al., 2018), petroleum engineering (Burdine, 1953; (Jin et al., 2021), and particle diffusion in filamentous biopolymer
Luo et al., 2019; Song et al., 2019, 2019b; Tian et al., 2019), chemical networks (Mickel et al., 2008).
engineering (Bardestani et al., 2019; Kenvin et al., 2015; Occelli et al., The description of particle, crystal, and grain size distributions is of
2003), and biomedical sciences (Bagherzadeh et al., 2013; Bartoš et al., significant practical importance in various fields aiming to understand
2018; Udenni Gunathilake et al., 2017). In the context of rock physics, and characterize components of microstructures (Berger et al., 2011;
the transport and mechanical properties of rocks are strongly influenced Higgins, 2000; Roostaei et al., 2020). The particle size distribution is
by the size and arrangement of their pores and grains, respectively. Most used amongst others in soil sciences for soil classification, the estimation
naturally occurring geological materials are composed of heterogeneous of soil hydraulic properties, and evaluation of depositional history of
and complex pore and grain structures. The pore size distribution is used transported soils (Roostaei et al., 2020). Crystal and grain size distri­
as a key measure in many engineering calculations to quantify the butions are used to quantify the texture of solid rocks, characterize
complex geometry of the pore space. The impact of pore size distribution sediments based on their depositional environments, and control sand

* Corresponding author.
E-mail address: armafr@dtu.dk (A. Afrough).

https://doi.org/10.1016/j.cageo.2021.104895
Received 23 March 2021; Received in revised form 28 June 2021; Accepted 20 July 2021
Available online 26 July 2021
0098-3004/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
H. Safari et al. Computers and Geosciences 156 (2021) 104895

production in wellbores drilled in unconsolidated and consolidated In this work, we propose an image processing workflow for
sandstone rocks (Ghasemi et al., 2018; Higgins, 2000; Srisutthiyakorn extracting pore and grain size distributions of natural rocks from X-ray
and Mavko, 2019). microcomputed tomography and scanning electron microscopy images.
Various experimental methods have been developed for analyzing We employ stereological techniques to convert 2D distributions ob­
the pore and grain size distributions in porous rocks over the years. tained from SEM images to 3D distributions comparable with μCT re­
Practical methods for determining the grain size distribution include sults. While most commercial image processing software do not offer
sieve analysis, laser diffraction, point-count method, and more recently stereological algorithms, this study emphasizes the role of stereology in
the dynamic image analysis method (Roostaei et al., 2020; Srisutthiya­ obtaining representative pore and grain size distributions from 2D SEM
korn and Mavko, 2019). Although these methods are widely used for images. Pore and grain size distributions are presented in terms of
determining the particle size and shape in unconsolidated materials, volumetric probabilities as opposed to frequency- and count-based
they cannot be applied to consolidated porous media (Ghasemi et al., methods. Volume fractions are of significant importance in engineer­
2018). Experimental methods for measuring the pore size distribution ing and theoretical calculations concerning aquifer capacity, distribu­
include mercury injection porosimetry (MIP) for the pore throat size tion of fluid saturations in oil reservoirs, and geo-mechanical properties
distribution, nitrogen adsorption isotherms for the pore body size dis­ of rocks. The pore and grain size distribution of six common
tribution of micropores, and nuclear magnetic resonance (NMR) for the commercially-accessible rock samples, including five sandstones and
pore body size distribution (Afrough et al., 2019; Arns, 2004; Kruschwitz one carbonate, are investigated. The rock samples span a variety of rock
et al., 2020; Shuaibing Song et al., 2019). These methods normally types with different and similar pore and grain size distributions and are
require the injection of a fluid phase into porous samples and mostly used by researchers worldwide in their verifying of measurement
describe the spatial distribution of connected pore space. As a result, methods, such as in magnetic resonance relaxation (Afrough et al., 2019;
disconnected pore structures which might become reconnected in pro­ Arns, 2004). The algorithms presented in this study could directly be
cesses such as reactive fluid transport still remain undetected (Al-Khu­ applied or easily extended to other porous materials. The acquired im­
laifi, 2018). ages and obtained results can also be used by those processing μCT and
The emergence of X-ray microcomputed tomography (μCT) as a non- SEM images, analyzing magnetic resonance relaxation, or those per­
invasive and non-destructive method along with recent improvements in forming fluid flow simulation in rocks.
the computational power has made it possible to visualize the internal
structure of porous materials. The μCT scanning techniques have been 2. Material and methods
broadly used to characterize the geometrical structure of pore, grain,
and mineral phases in rock samples (Arns, 2004; Elkhoury et al., 2019; 2.1. Description of rock samples
Ghasemi et al., 2018; Lai et al., 2015; Schmitt et al., 2015; Srisutthiya­
korn and Mavko, 2019; Zhang et al., 2018). The resolution of 3D scans is Five different sandstone samples including Berea (BA), Nugget (NU),
directly proportional to the sample size, beam quality, and specifications Buff Berea (BB), Castlegate (CG), and Bentheimer (BT), and one car­
of the detector (Blunt et al., 2013). The true resolution of μCT systems is bonate sample, Indiana Limestone (IL), were acquired from Kocurek
limited to approximately 0.3 μm in projection imaging, in comparison Industries (Caldwell, TX, USA). The selected rock samples present a wide
with a sub-50 nm imaging resolution in lens-based X-ray CT systems range of pore and grain size distributions.
(Withers, 2007). High-resolution scans can reveal submicron textures Berea and Buff Berea samples are Upper Devonian grain-supported
and heterogeneities in porous materials at the cost of significant sample sandstones from the Kipton formation. Porosity of the Berea sample
size reduction (Bai et al., 2013; Blunt et al., 2013). Scanning electron ranges from 18 to 21 percent, while Buff Berea has slightly higher po­
microscopy (SEM) is another method widely employed in describing the rosities ranging from 20 to 22 percent. Castlegate is a Late Cretaceous
submicron structure of porous materials. SEM offers high-resolution sandstone sample from the Price River formation of Mesaverde group
images with limited field of views (FOVs) that could be used in the which has a porosity ranging from 26 to 29 percent. Nugget is a Late
characterization of small-scale heterogeneities in natural rocks. SEM, in Triassic heterogeneous sandstone sample from the Utah formation
the backscattered electron detection mode, can observe compositional which compared to other samples has a lower porosity ranging from 10
variations with a lateral spatial resolution of approximately 10–100 nm to 15 percent. Bentheimer is a well-sorted Valaginian sandstone from the
depending on the specimen composition and the beam energy selected Germany formation and has a porosity ranging between 21 and 26
(Goldstein et al., 2018). Two-dimensional SEM images have been mostly percent. The sandstone samples are composed dominantly of quartz,
considered as a qualitative tool and their quantitative analysis to obtain with other present minerals being feldspar, clays, and sometimes calcite
pore and grain size distributions have been limited (Song et al., 2019; S. (Afrough et al., 2017; Dalton et al., 2020).
Song et al., 2019b; Zhang et al., 2015). Indiana limestone is a Mississippian fossiliferous limestone from the
Two-dimensional size distributions obtained from SEM images do Salem limestone formation in the Bedford-Bloomington area which is
not represent the actual 3D distributions because of the sectioning bias mainly composed of calcite (more than 97 %) with minor amounts of
(Liu et al., 2020; Reppel and Weinberg, 2019; Song et al., 2019b). Sri­ other minerals, such as alumina, iron oxide, silica, and magnesite pre­
sutthiyakorn and Mavko (2019) presented a general approach for con­ sent (Ji et al., 2012). Its porosity with a bimodal pore size distribution is
verting 2D grain size distributions acquired from thin section images to characterized by intergranular micro-connected macropores and ranges
3D distributions by providing a forward solution to the Wicksell’s between 12 and 19 percent.
corpuscle problem. They obtained transformation matrices relating 2D
(from CT slices) to 3D grain size distributions using μCT images of 2.2. X-ray microcomputed tomography
natural rocks. We show in this study how their method could fail when
inherent microstructures in porous rocks are not properly resolved in Standard 1.5 inch core plugs were cut into 2–3 mm thick slabs. Small
μCT images. A variety of methods have been presented in the literature samples were chipped from these slabs and installed on stubs by glue.
for the characterization of pore and grain size distribution in porous These samples had varying dimensions approximately close to 2 × 2 × 4
geological samples. However, proposed methodologies are limited from mm3, but were not perfectly shaped. High-resolution μCT images were
one image processing environment to another and are not checked acquired using a Skyscan 1072 system at the Microscopy and Micro­
against several other methods and several samples. Therefore, a general analysis Facility of the University of New Brunswick in Fredericton,
image processing workflow that could be applied to both 2D and 3D Canada. Some of the samples were slightly larger than the field of view
images of rock samples for quantitative analysis of pore and grain size and were imaged in the region of interest imaging mode. Imaging was
distributions is still lacking. performed with a tube voltage of 91 kV and a source current of 110 μA

2
H. Safari et al. Computers and Geosciences 156 (2021) 104895

with a 1-mm aluminum filter. Source-to-object distance (SOD) was set to


29.73 mm. Samples were rotated one step at a time (0.9◦ ) through 360◦
and 16-bit images were recorded at each rotation step with a 1024 ×
1024 detector. The 3D reconstruction of X-ray radiographs was per­
formed with the Skyscan volumetric reconstruction program “NRecon”
(Version: 1.6.4.8) with a voxel size of 2.7 μm/pixel.

2.3. Scanning electron microscopy

For SEM observations, resin-impregnated samples were cut, grinded,


and polished to prepare thin sections. Imaging was performed by a JEOL
6400 scanning electron microscope operating at a voltage of 15 kV. All
SEM micrographs were acquired in the backscattered electron mode.
Prepared thin sections were scanned at different FOVs and between
twelve to nineteen 4096 × 4096 16-bit images were acquired for each
sample. Table 1 summarizes the typical resolution of the SEM images
captured for each sample. The resolution scales are fixed for all the
images acquired from each thin section sample except for Berea sand­
stone which has a few images captured at different magnifications.

3. Image processing workflow

The Avizo image processing software (Thermo Fisher Scientific, V.


2019.1) was employed in analyzing the 3D reconstructed μCT images of
the rock samples. For SEM images, a similar procedure was followed by
developing a computer program in the Python programming language.

3.1. X-ray microcomputed tomography

The acquired 16-bit images were first converted into 8-bit images
and their beam hardening and region of interest imaging artifacts were
corrected. The maximum possible rectangular sub-volumes were
extracted from each image for the quantification of the samples’ pore
and grain size distributions (see Appendix). Fig. 1 shows the general
workflow employed in this study for analyzing the μCT images. To
improve the signal-to-noise ratio, a 2D edge-preserving non-local means
filter was applied on every slice in the xy-plane. Two-dimensional,
instead of 3D, non-local means filtering was employed because of the
poor computational performance of the 3D filter in Avizo, even in the Fig. 1. The general workflow employed in the extraction of pore and grain size
GPU-processing mode. Table 2 summarizes the non-local means filter distributions from μCT images. An edge-preserving non-local means filter is
applied for denoising purposes and the filtered images are segmented into two
settings along with other image parameters including image resolution,
binary images representing pores and grains. A Euclidean distance transform is
and dimensions of the extracted sub-volumes. The pore size distribution
applied on the binary images which is then masked by a medial axis transform.
was not very sensitive to the variation of the search window size in the An H-maxima filter is applied on the masked distance map to obtain markers for
non-local means filtering. watershed segmentation. The labeled images from the watershed algorithm are
In the next step, we applied image segmentation techniques to used to obtain volumetric pore and grain size distributions. Each side of a cube
segment each gray-scale image into two binary images representing the is 675 μm.
pores and grains of the rock samples. This is the most crucial step as it
affects all the subsequent quantitative analysis. Although a variety of transform was applied to the segmented images of pores and grains. A
segmentation algorithms have been proposed in the literature, global medial axis transform was applied to binary images to extract connected
thresholding technique is still the most commonly applied approach in voxel skeletons which was then used to mask the calculated distance
porous media research (Iassonov et al., 2009). Automatic thresholding maps. An H-Maxima filter was applied on the new distance map to find
algorithms such as Otsu’s method (Otsu, 1979) tended to over-segment the peaks in every region. The calculated peaks were passed as seeds to a
the gray-scale images in almost all the samples. This could be attributed marker-based watershed segmentation algorithm to provide labeled
to their sensitivity to partial volume effects arising from the unresolved images for pore and grain volumes. Segmented regions were analyzed
sub-resolution features and limited contrast-to-noise ratio in μCT images and equivalent diameters, defined as the diameter of a sphere of the
(Sheppard et al., 2014). An interactive thresholding procedure with same volume as that of a segmented region, were calculated. Therefore,
immediate visual feedback was found to provide more accurate and the term diameter in the context of pore and grain diameters is referring
consistent segmentation results. However, manual thresholding still to the equivalent diameter throughout this paper. These spheres with
remains labor intensive and might introduce some operator bias. varying equivalent diameters are represented as nodes in the generated
Having performed image segmentation, a Euclidean distance pore and grain network models. Volumetric probability functions for
pore and grain size distributions were calculated using the regional
Table 1 volumes and their respective equivalent diameters.
Typical resolution scales of the captured SEM images for each sample.
Samples NU CG BT BB BA IL

Resolution (μm) 0.406 0.406 0.338 0.406 0.239 0.406

3
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Table 2
Parameters of the μCT images including their resolution, the dimensions of extracted sub-volumes, and the non-local means filter settings for each sample.
Parameters NU CG BT BB BA IL

Resolution (μm) 2.7 2.7 2.7 2.7 2.7 2.7


Size (voxels) 528 × 281 × 433 × 603 × 378 × 406 × 584 × 584 × 664 × 400 × 628 × 320 ×
700 794 520 600 700 519
Non-Local Means Filter Search Window (pixels) 10 10 10 5 5 10
(2D) Local Neighbourhood 3 3 3 3 3 3
(pixels)
Similarity Value 0.6 0.6 0.6 0.6 0.6 0.6

3.2. Scanning electron microscopy

A computer program was developed in Python for quantitative


analysis and processing of all the 2D SEM images captured at different
FOVs from the prepared thin section samples. This program integrates
pore and grain size distributions from several fields of view and unfolds
the resulting 2D size distribution to a 3D size distribution for each
sample.
The “ncempy” module was first imported to load the metadata from
DM4 files into Python. Image data and magnification scales were then
extracted from all the captured views. Next, a nonlocal-means filter from
the skimage library was employed with a patch size and patch distance
of 3 to denoise all the SEM images. Backscattered SEM images provided
sufficient contrast to observe the mineral phases present in the rock
samples. Quartz, feldspar, and clays were the dominant minerals present
in sandstone samples, while for Indiana limestone, calcite was the only
mineral phase. A multiphase automatic segmentation was performed on
SEM images employing a fast multi-Otsu thresholding algorithm
developed by Liao et al. (2001). The method provides universal
threshold values by maximizing the inter-class variances. Unlike μCT
images, automatic segmentation algorithms provide fairly accurate re­
sults in SEM images due to limited partial volume effects and good
contrast-to-noise ratio. Sandstone samples were segmented into four
different phases consisting of pores and three main minerals namely
quartz, feldspar, and clays. Indiana limestone was segmented into two
phases representing pores and solids (dominated by calcite). Fig. 2
shows the general workflow employed for the extraction of pore and
grain size distributions from SEM images.
Similar to the μCT workflow, an H-Maxima filter was applied to the
masked distance maps of pores and grains to find markers for watershed
segmentation. However, resolved microstructural details in SEM images
resulted in homogeneous distance maps with spurious local maxima,
with ridges and plateaus, leading to an over-segmented watershed. To
solve this problem, the methodology proposed by Gostick (2017) was
used to dilate the distance maps by trimming the saddle points and
nearby peaks with a cubic structuring element. The new markers were
passed into the watershed algorithm to generate labeled images for
pores and grains. Although dilation of distance maps resulted in lesser
bisection of larger regions, several pores and grains appear to have been Fig. 2. The general workflow used for the extraction of pore and grain size
labeled incorrectly as shown in Fig. 2 (e, f). Despite the improved image distributions from SEM images, (a) backscattered SEM images of resin-
processing procedures employed in this study, automatic segmentation impregnated sandstones. The pore space is black, quartz is medium gray,
still remains challenging and a few segmentation errors are inevitable feldspar is white, and clays are dark gray, (b) multiphase segmentation of
minerals and the pore space, (c), (d) a distance transform performed on the pore
(Sheppard et al., 2014). The “regionprops” method from the “measure”
and grain space, respectively (e), (f) the watershed segmentation of pores and
module of the “skimage” library was then used to analyze the regions
grains, respectively, using markers obtained from dilated distance maps. The
labeled by the watershed algorithm. The areas of regions were obtained
scale bar is equivalent to 500 μm.
by summing the pixels in each region and equivalent diameters of circles
with the same area as the regions were calculated afterwards. Using
sectioning bias. Stereological solutions to this problem are categorized
regional areas and their respective equivalent diameters, area-weighted
into two classes of (1) direct methods, for which no shape and size
probability functions were obtained for the pore and grain size
distributions are assumed, and (2) indirect methods, in which a few
distributions.
assumptions are necessary (Higgins, 2015). Although direct methods
provide more precise size distributions, they are rarely employed mainly
3.3. Stereological unfolding of 2D distributions to 3D
due to their requirements of many closely obtained serial sections from
the 3D object. In most cases, only a few slices are available or could be
The pore and grain size distributions obtained from 2D SEM images
obtained from the 3D structure, hence leaving the indirect methods as
do not represent the actual 3D distributions, mainly because of the

4
H. Safari et al. Computers and Geosciences 156 (2021) 104895

the only viable solution (Higgins, 2015). r


Stereological techniques have been widely used to obtain volume ϕ(r) = √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ (1)
R2 − r 2
fractions and 3D particle size distributions from measurements per­
formed on 2D cross-sections (Berger et al., 2011; Jutzeler et al., 2012; Employing the workflow described in section 3.2, 2D pore and grain
Reppel et al., 2019). A classical problem in stereology which addresses size distributions are obtained from the SEM images taken at different
the unfolding of size distributions from 2D to 3D is the well-known FOVs. Then, for each bin in the 2D size distribution histogram, the
Wicksell’s corpuscle problem in which Wicksell investigated the size number of particles per unit volume Nvi (or 3D size distribution) is
distribution of spherical corpuscles in the human spleen (Reppel and directly calculated from the number of sections per unit area NAi (or 2D
Weinberg, 2019; Srisutthiyakorn and Mavko, 2019; Wicksell, 1925). In size distribution) via the following formulations:
this study, we employed the methodology proposed by Sahagian and (
i−∑
1
)
Proussevitch (1998) for poly-dispersed spherical particles to unfold the 1
Nvi = ’ α1 NAi − αj+1 NA(i− j) (2)
population of 2D diameters obtained from processing SEM images into Hi j=1

3D size distributions. Their technique is applicable to any particle size or ( )


size distribution regardless of the lognormality of distribution. Fig. 3 1 ∑
i− 2
αi = α1 Pi − αj+1 Pi− (3)
provides a simplified representation of the stereological unfolding P1 j=1
j

process.
In Fig. 3, the occurrence frequency ϕ, for spherical particles with where i = 1, .., n with n being the number of bins in the histogram, Hi is

radius R intersecting with circles of radius r is: the mean projected height which is equal to 1 for spherical particles, α1
is the conversion coefficient of the first class, and P represents the cross-
section size probabilities. The probability of a particular class in the
cross section with radius r (r1 < r < r2 ) intersecting with a sphere of
radius R is given by:
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
1
P(r|R) = ( (R2 − r21 ) − (R2 − r22 )) (4)
R

where r1 and r2 are the lower and upper limits of the classes in 2D his­
togram, respectively. Mathematical details of the above formulations
could be found in the work of Sahagian and Proussevitch (1998).
The “GrainSize Tools” Python scripts developed by Lopez-Sanchez
(2018) for grain size analysis were modified to account for the volu­
metric probabilities and employed in this study to calculate 3D distri­
butions of pore and grain sizes from 2D distributions.

4. Results and discussion

Fig. 4 shows the 3D gray-scale reconstruction of carbonate and


sandstone samples used in this study for the characterization of pore and
grain size distributions along with their segmented pore and grain
spaces. Berea, Buff Berea, and Castlegate sandstone samples have similar
pore and grain structures. Nugget has smaller pore sizes compared to
other samples, while pore and grain sizes in Bentheimer are visibly
larger. The backscattered SEM images and their multiphase segmenta­
tion results shown in Fig. 5 provide a 2D impression of the inherent
structures of pores, grains, and minerals specific to each rock sample.
The relatively low contrast of μCT images and the close gray levels of
clay minerals and pores made it difficult to appropriately segment clays
in sandstone samples imaged by our instrument. Other researchers have
successfully segmented clay minerals from the surrounding pore space
by acquiring high-contrast μCT images (Cnudde et al., 2011; Lai et al.,
2015). However, in most cases the microporosities in clay minerals are
not resolved. In contrast to μCT images, SEM images provided sufficient
contrast to observe various minerals as well as small-scale heterogene­
ities in rock samples.
SEM observations show that sandstone samples are predominantly
Fig. 3. Stereological unfolding of 2D size distributions from SEM images to 3D
composed of moderately well-sorted quartz grains and small amounts of
distributions, (a) typical backscattered SEM images of resin-impregnated
feldspar and clays. Clay minerals are distributed either as thin layers on
sandstone thin sections, (b) binary segmentation of the rock matrix and the
pore space. The pore space is filled with inscribed circles of various radii, (c) a other minerals’ surfaces or as agglomerates filling the pore space. Ben­
representation of the Wicksell’s corpuscle problem indicating the likelihood of a theimer featured weathered feldspar grains that were consistently
certain circular region of radius r intersecting with a sphere of radius R, (d) observed in μCT images and also by light microscopy.
stereological conversion of 2D size distributions (NA) obtained from 2D cross- Indiana limestone shows a monomineralic texture predominantly
sections to 3D distributions (Nv) comparable with those of μCT images. While composed of calcite with allochems such as fossils and ooids constituting
2D size distributions always show large negative skewness toward smaller sizes, a major fraction of the solid structure. SEM images reveal a high con­
the occurrence probability of most of the small pores almost disappears during centration of microporosity regions at the boundaries of allochems
stereological conversion to 3D size distributions. The scale bar is equivalent to which account for the connectivity of larger macropores throughout the
500 μm. The inscribed circles are for demonstration only; an equivalent
structure. These microporosity details are not resolved in μCT images
diameter definition is employed in the rest of the manuscript.

5
H. Safari et al. Computers and Geosciences 156 (2021) 104895

sandstones appear to have similar mineral compositions with quartz


mineral constituting a major part of their grain structures. These mineral
compositions are consistent with other XRD observations in the litera­
ture (Afrough et al., 2017; Churcher et al., 1991; Peksa et al., 2015;
Shehata and Nasr-El-Din, 2015).

4.1. Pore size distribution

Fig. 6 compares the pore size distribution of rock samples derived


from SEM and μCT images in terms of their volumetric probabilities. The
red histograms represent the 3D pore size distributions obtained from
the stereological unfolding process. Pore size distributions obtained
from processing μCT images for all sandstone samples are approximately
lognormal, while 3D unfolded SEM distributions indicate skewness to­
ward smaller pore diameters. This skewness is attributed to the SEM-
resolved microporosity in clay minerals. For Indiana limestone, the
distribution of pore diameters from μCT shows negative skewness to­
ward smaller pores, whereas SEM observations present a clear bimodal
pore size distribution which is attributed to the well-resolved micropo­
rosity regions in SEM images. The methodology proposed by Sri­
sutthiyakorn and Mavko (2019) overlooks the effect of these
microstructural heterogeneities in converting size distributions from 2D
to 3D.
Table 5 summarizes the statistical descriptions of pore diameters
obtained from μCT, 2D SEM, and the stereologically unfolded 3D SEM
distributions for all rock samples. The skewness in this table describes
the lack of symmetry in lognormal distributions whereas kurtosis mea­
sures the weight of distribution’s tail relative to its center. The original
pore sizes obtained from 2D SEM analysis prior to 3D unfolding present
very large values of kurtosis and skewness for all the rock samples
indicating that their size distributions are heavily tailed toward smaller
pore diameters. This is due to the sectioning bias originating from the
fact that the majority of pores and grains are not necessarily sectioned
through their maximum possible dimensions. As a result, 2D size dis­
tributions of the cross-sections are always biased toward smaller di­
ameters and do not represent the actual 3D distributions.
For Berea sandstone, the mean pore diameter from μCT and 3D SEM
analysis are 22 and 26 μm, respectively. For Indiana limestone, the
average pore size from μCT is 36 μm, while the mean pore diameters
Fig. 4. The 2D and 3D gray-scale reconstruction of sandstone and carbonate
samples and their segmented pore and grain space. Berea, Buff Berea, and
from processed SEM images are 9.5 and 50 μm for small and large pores,
Castlegate have similar pore and grain structures. Nugget shows smaller pore respectively. These results are in close agreement with the absolute pore
sizes, whereas the pore and grain sizes of Bentheimer are clearly larger. It is sizes measured by Afrough et al. (2019) using non-ground eigenvalues of
difficult to distinguish clay minerals in sandstone samples due to their gray magnetic resonance relaxations.
level that is similar to that of the pore space. The intermediate gray levels in The average pore sizes and pore size distributions obtained from μCT
Indiana limestone represent unresolved microporosity regions in the recon­ and SEM images for Berea and Buff Berea are relatively close as expected
structed μCT image. Each side of the cubes (slices) has a length and size of 250 from their similar pore structures. Mean pore diameters of Castlegate
voxels (pixels) and 675 μm, respectively. and Nugget sandstones are almost identical with Castlegate exhibiting a
wider range of pore sizes. Among sandstone samples, Bentheimer has
and are represented mostly with intermediate gray levels as a result of the highest mean pore diameter consistent with its relatively large and
the partial volume effect. well connected pore space. The small pore size measured from SEM
Table 3 compares the experimental porosity of the samples with images of Bentheimer is ascribed to the presence of fine weathered
those calculated from processing the μCT and SEM images. The experi­ feldspar particles and possible smearing of clay minerals that might have
mental porosity was determined by calculating (1) the pore volume from had been introduced to the sample surface in the sample preparation
the saturated weight and dry weight employing the ‘Liquid Saturation stage. This is the likely reason behind deviation between the μCT and
Method’ and (2) the bulk volume employing the ‘Calipering’ technique – SEM mean pore size for Bentheimer.
both by largely following the procedures described in the API Recom­
mended Practices for Core Analysis. While volumetric estimates from 4.2. Grain size distribution
μCT are largely consistent with experimental data, SEM observations
appear to underestimate the void fractions in sandstones and provide Fig. 7 shows the grain size distribution of minerals present in rock
slightly higher values for Indiana limestone. The porosity obtained for samples from 3D unfolded SEM observations. The predominant minerals
the Castlegate sandstone from SEM images is underestimated by more are quartz, feldspar, and clay minerals in sandstones and calcite in
than 20 % compared to its experimental value. The SEM porosity of Indiana limestone.
Bentheimer might had been altered by the presence of fine weathered Table 6 summarizes the statistical description of mineral grain sizes
feldspar particles from sample preparation. Table 4 provides area frac­ in sandstone samples. Clay exhibits a wide range of grain sizes with the
tions of the mineral phases present in sandstone samples derived from majority of grains occurring in the range of 1–40 μm. The appearance of
backscattered SEM images. Comparing the bulk minerology, all the large clay grains is attributed to the agglomeration of clays within the

6
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Fig. 5. Typical SEM images of sandstone and carbonate samples and their multiphase segmentation results. Sandstone samples are mainly composed of quartz grains
and small amounts of feldspar and clays. Clay minerals are observed either as thin layers or as agglomerates filling the pore space. Indiana limestone features a
monomineralic texture predominantly composed of calcite mineral. Well-resolved SEM images reveal microporosity regions in sandstones and Indiana limestone. The
scale bars are equivalent to 500 μm except for Berea sandstone which has a scale bar equivalent to 200 μm.

Table 3 Table 4
Comparison of the experimental porosities obtained from gravimetry with those Area fractions of the minerals present in sandstone samples derived from SEM
calculated from μCT and SEM images. observations.
Porosity (%) NU CG BT BB BA IL Minerals NU CG BT BB BA

μCT 15 27 22 22 21 14 Clays 0.057 0.074 0.052 0.060 0.074


SEM 13 19 12 18 19 16 Feldspar 0.051 0.040 0.050 0.054 0.065
Exp. 15 26 21 22 21 14 Quartz 0.76 0.69 0.78 0.71 0.67

7
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Fig. 6. The volumetric probability of pore diameters from μCT (blue) and unfolded SEM (red) for Bentheimer (BT), Berea (BA), and Indiana limestone (IL) (Top), and
for Castlegate (CG), Buff Berea (BB), and Nugget (NU) (Bottom). SEM-resolved microporosity regions in clay minerals has shifted pore size distributions of sandstone
samples toward smaller sizes. The bimodal pore size distribution obtained from SEM images of Indiana limestone is also attributed to highly resolved microporosity
regions at the grain boundaries. The pore size distributions from μCT and SEM images are consistent. (For interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of this article.)

Table 5
The statistical description of pore diameters from μCT, 2D SEM, and the unfolded 3D SEM distributions for all rock samples. All pore diameters are in μm.
Statistical Parameters NU IL CG

μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D)

mean 42 8.3 42 36 4.8 39 40 8.0 39


std 25 8.8 29 27 4.3 42 24 8.3 29
min 3.3 0.6 10 3.3 0.5 2.9 3.3 0.5 8.3
25 % 23 3.2 18 15 2.4 8.4 22 3.2 16
50 % 40 5.4 33 30 3.5 22 38 5.3 30
75 % 59 9.9 60 49 5.6 56 56 9.4 56
max 151 113 108 170 166 158 193 111 106
kurtosis 0.3 17 0.1 2.8 36 1.3 0.1 25.1 0.1
skewness 1.5 3.7 0.9 1.6 3.8 1.4 1.5 4.4 0.9

Statistical Parameters BT BB BA
μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D)

mean 53 5.3 28 36 7.9 44 22 4.6 26


std 34 4.7 23 27 8.0 36 22 4.8 21
min 3.3 0.4 4.9 3.3 0.5 8.7 3.3 0.3 4.7
25 % 27 2.6 9.6 13 3.2 17 4.8 1.9 9.4
50 % 48 4.0 19 31 5.3 32 11 3.2 19
75 % 73 6.4 41 52 9.3 63 35 5.5 37
max 212 91 87 252 155 147 163 84 80
kurtosis 0.5 64 0.3 0.5 24 1.3 0.3 35 0.4
skewness 1.4 6.9 1.1 1.5 4.4 1.3 1.4 5.2 1.1

pore space. These agglomerates even resulted in a bimodal grain size size distribution is observed for feldspar. The calcite mineral in Indiana
distribution of clays in Berea sandstone (see Fig. 7). Quartz and feldspar limestone also exhibits a bimodal grain size distribution consistent with
minerals show similar distributions with grain sizes spanning from 1 to its highly resolved microstructural details in SEM images.
250 μm. These minerals demonstrate a unimodal grain size distribution To compare the grain size distribution histograms with those ob­
in all sandstone samples, except for Castlegate where a bimodal grain tained from μCT, the mineral phases were merged into a single solid

8
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Fig. 7. The volumetric probability of mineral grain diameters from SEM images for Bentheimer (BT), Berea (BA), and Indiana limestone(IL) (Top), and for Castlegate
(CG), Buff Berea (BB), and Nugget (NU) (Bottom). Clay minerals in sandstones exhibit a wide range of grain sizes ranging from 1 to 40 μm. The appearance of large
grain sizes are due to agglomeration of clays within the pore space resulting in a bimodal grain size distribution in Berea sandstone. The distribution of grain sizes in
quartz and feldspar minerals is unimodal in almost all sandstone samples except for Castlegate sandstone where a bimodal grain size distribution is observed for
Feldspar. Calcite in Indiana limestone also shows a bimodal grain size distribution due to its well-resolved microstructural details from SEM images.

Table 6
Statistical description of the quartz, feldspar, and clay grain diameters from the unfolded 3D SEM distributions for sandstone samples. All grain diameters are in μm.
Statistical Parameters NU CG BT

Quartz Feldspar Clay Quartz Feldspar Clay Quartz Feldspar Clay

mean 126 67 14 55 96 12 47 104 9.1


std 65 55 8.9 61 59 7.4 50 60 5.9
min 48 11 4.0 3.9 13 3.6 1.1 32 1.0
25 % 73 24 6.9 9.0 50 6.0 4.4 56 4.5
50 % 111 52 12 21 82 10 35 90 7.5
75 % 168 93 20 84 134 17 77 143 12
max 256 231 34 217 218 30 173 227 25
kurtosis 0.7 2.3 0.5 0.8 0.6 0.3 0.0 0.6 0.5
skewness 0.7 1.4 0.8 1.3 0.6 0.8 0.9 0.7 1.0

Statistical Parameters BB BA
Quartz Feldspar Clay Quartz Feldspar Clay

mean 96 61 12 44 38 9.6
std 47 51 7.8 35 35 9.5
min 39 11 3.3 1.0 1.1 1.0
25 % 57 22 6.1 17 10 3.3
50 % 85 43 10 34 24 6.4
75 % 127 86 18 65 55 13
max 188 190 30 125 127 46
kurtosis 0.7 0.6 0.5 0.2 0.3 5.6
skewness 0.6 1.2 0.8 0.9 1.1 2.1

phase. Fig. 8 illustrates the volumetric probabilities of grain size dis­ agreement between volume-weighted grain sizes, the merging of min­
tribution in sandstone and carbonate rock samples. Although the 3D eral grain components resulted in large arithmetic mean grain diameters
volumetric grain size distributions obtained from SEM images are from SEM images, compared to those of μCT results. Similar to pore size
largely consistent with μCT results, the probabilities of small-scale distributions, 2D grain size distributions from SEM images represent
grains like those of clay minerals have vanished due to merging. high values of kurtosis and skewness compared to 3D distributions.
Table 7 provides the statistical description of the grain diameters Among sandstones, Bentheimer has the largest average grain diam­
derived from μCT and SEM images for all samples. Despite the eter (82 μm from μCT and 208 μm from SEM), while Berea represents the

9
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Fig. 8. Volumetric probability of grain diameters from μCT (blue) and unfolded SEM (red) for Bentheimer (BT), Berea (BA), and Indiana limestone (IL) (Top), and for
Castlegate (CG), Buff Berea (BB), and Nugget (NU) (Bottom). The merging process resulted in the disappearance of small-scale grains like clays in SEM images and
shifted the mean grain diameters toward larger values. Despite discrepancies in arithmetic means, volume-weighted averages of grain diameters from SEM images are
largely consistent with μCT results. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)

Table 7
Statistical description of the grain diameters from μCT, 2D SEM, and the unfolded 3D SEM distributions for all rock samples. All grain diameters are in μm.
Statistical Parameters NU IL CG

μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D)

mean 58 16 161 71 10 73 64 13 153


std 46 28 71 55 8.2 66 41 20 67
min 3.3 0.5 74 4.2 0.5 9.6 3.3 0.5 71
25 % 21 3.9 104 31 4.9 22 32 4.2 100
50 % 43 6.7 147 56 8.2 48 56 7.2 140
75 % 86 12 207 96 13 107 90 12 197
max 261 310 292 392 251 238 263 294 278
kurtosis 0.1 16 0.1 2.0 95 0.3 0.5 18.5 0.1
skewness 0.6 3.3 0.6 1.3 6.4 1.1 0.7 3.5 0.6

Statistical Parameters BT BB BA
μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D) μCT SEM (2D) SEM (3D)

mean 82 9.3 208 63 13 150 51 7.5 95


std 64 17 97 41 21 55 35 12 40
min 3.3 0.5 81 3.3 0.6 81 3.3 0.4 44
25 % 27 3.2 132 31 3.9 107 23 2.4 62
50 % 66 5.4 189 54 6.7 142 42 4.1 87
75 % 125 9.1 271 88 12 187 73 7.1 122
max 350 414 390 299 261 247 246 180 171
kurtosis 0.8 40 0.1 1.1 20 0.1 1.3 31 0.1
skewness 0.8 4.7 0.6 1.0 3.6 0.5 1.3 4.3 0.5

smallest (51 μm from μCT and 95 μm from SEM). Castlegate, Buff Berea, 4.3. Representativeness
and Nugget represent relatively similar mean grain diameters. For
Indiana limestone, the average grain diameter from μCT is 71, whereas All the rock samples used in this study showed homogeneous char­
the mean grain sizes from SEM observations are 12 and 88 for small and acteristics. The agreements between the pore and grain size distributions
large grains, respectively. The grain size distribution of Indiana lime­ obtained from μCT and SEM images confirm this homogeneity and
stone obtained from μCT exhibits a left-tailed unimodal distribution indicate that the selected volumes for μCT imaging are fairly represen­
skewed toward smaller grain diameters. tative. Yet, discrepancies exist between the amount of microstructural
details, such as small pores and grains, resolved in μCT and SEM images

10
H. Safari et al. Computers and Geosciences 156 (2021) 104895

which is attributed to their limits of resolution. High resolution μCT minerals within the pore space resulted in a bimodal grain size distri­
imaging requires even smaller sample sizes which might result in their bution in Berea sandstone. The clay microporosities were also observed
loss of representativeness as compared to other characterization meth­ to influence the pore size distributions in sandstone samples. In Indiana
odologies. For the rock samples considered in this study, the pore vol­ limestone, the SEM-resolved microporosity regions led to bimodal pore
ume fractions from μCT images were quite consistent with experimental and grain size distributions as opposed to unimodal distributions ob­
porosities obtained by gravimetry, whereas void fractions derived from tained from μCT images.
SEM observations presented noticeable discrepancies. The mean pore We showed that our workflow can be successfully employed to
sizes obtained from μCT and SEM images, however, were found to be in process and extract quantitative information from μCT and SEM images
close agreement with each other as well as with the absolute pore sizes of natural rocks. We addressed the impact of sectioning bias in SEM
measured using magnetic resonance relaxation. Unlike μCT and SEM images and highlighted the role of stereology in obtaining representa­
imaging techniques, experimental methods such as NMR and MIP, only tive 3D size distributions. The rock materials implemented in this study
provide information about the connected pore space. span a variety of rock types with different and similar pore and grain size
distributions. These rock samples are widely used by researchers
5. Conclusions worldwide for analyzing magnetic resonance relaxation or performing
pore-scale fluid flow simulations.
We have shown the application of a proposed image processing
workflow for the extraction of pore and grain size distributions from X-
ray microcomputed tomography and scanning electron microscopy im­ Declaration of competing interest
ages in five different sandstones and one carbonate sample. While it was
difficult to distinguish between the mineral phases in rock samples from The authors declare that they have no known competing financial
μCT images, particularly clay minerals in sandstones, SEM images pro­ interests or personal relationships that could have appeared to influence
vided a sufficient contrast to observe different minerals from 2D cross- the work reported in this paper.
sections. The predominant minerals in sandstone rocks were quartz,
feldspar, and clays, whereas Indiana limestone exhibited a mono­ Acknowledgements
mineralic texture of calcite mineral. Two-dimensional pore and grain
size distributions from SEM images were found to be misrepresenting Authors thank Steven R. Cogswell from Microscopy and Microanal­
the actual 3D distributions due to the sectioning bias. Stereological ysis for microscopy, Stephen Delahunty from the Earth Sciences
methods converted the 2D size distributions obtained from SEM images Department for sample preparation, Tatiana Zaraiskaya from UNB Li­
to 3D distributions and the results were compared with μCT observa­ braries for data management, and Florea Marica from UNB MRI
tions. The distribution histograms derived from processing μCT and SEM Research Center for assistance, all from the University of New Bruns­
images presented noticeable differences with regard to their limits of wick. Armin Afrough acknowledges the Danish Hydrocarbon Research
resolution. We observed that microstructural details resolved in SEM and Technology Centre for funding. Bruce J. Balcom acknowledges
images can significantly affect the distribution of pore and grain sizes in NSERC of Canada for a Discovery Grant and the Canada Chairs Program
both sandstone and carbonate samples. The agglomeration of clay for a Chair in Material Science MRI.

Appendix

Figure A.1 illustrates an example of the artifact correction, sub-volume extraction, and nonlocal means filtering for X-ray CT images.

11
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Fig. A.1. Illustration of the image processing steps performed on μCT images, (a) original image (b) artifact correction. There are two different artifacts in the
original image including (1) the region of interest imaging artifact and (2) the beam hardening artifact. We applied the beam hardening module in Avizo software to
correct for the intensity variations caused by both artifacts, (c) maximum possible rectangular sub-volume extraction from the image, (d) cropping of the selected
rectangular sub-volume, and (e) denoising the extracted sub-volume via applying a nonlocal means filter.

Author contributions

All authors contributed to this paper.

Data and code availability

The data and code/procedures of this manuscript are available from a data repository.

Data and computer code availability

Data, Python codes, and Avizo procedures associated with this manuscript are available in the UNB institutional repository Dataverse at DOI htt
ps://doi.org/10.25545/A19OPN. The Python code is available as a Jupyter Notebook SEM_Analyzer.ipynb, version 1.0, and the Avizo 2019.1 pro­
cedure, including all parameters for all samples, is available in Avizo_Worflow.csv. The data and codes are available as CC0 – Public Domain
Dedication.

12
H. Safari et al. Computers and Geosciences 156 (2021) 104895

References and mercury intrusion. Langmuir 31, 1242–1247. https://doi.org/10.1021/


la504575s.
Kruschwitz, S., Halisch, M., Dlugosch, R., Prinz, C., 2020. Toward a better understanding
Afrough, A., Zamiri, M.S., Romero-Zerón, L., Balcom, B.J., 2017. Magnetic-resonance
of low-frequency electrical relaxation — an enhanced pore space characterization.
imaging of fines migration in Berea sandstone. SPE J. 22, 1385–1392. https://doi.
Geophysics 85, MR257–MR270. https://doi.org/10.1190/geo2019-0074.1.
org/10.2118/186089-PA.
Lai, P., Moulton, K., Krevor, S., 2015. Pore-scale heterogeneity in the mineral
Afrough, A., Vashaee, S., Romero Zerón, L., Balcom, B., 2019. Absolute measurement of
distribution and reactive surface area of porous rocks. Chem. Geol. 411, 260–273.
pore size based on nonground eigenstates in magnetic-resonance relaxation. Phys.
https://doi.org/10.1016/j.chemgeo.2015.07.010.
Rev. Appl. 11 https://doi.org/10.1103/PhysRevApplied.11.041002.
Liao, P.-S., Chen, T.-S., Chung, P.-C., 2001. A fast algorithm for multilevel thresholding.
Al-Khulaifi, Y.A., 2018. Pore-scale Imaging and Analysis of Carbonate Dissolution during
J. Inf. Sci. Eng. 17, 713–727. https://doi.org/10.1688/JISE.2001.17.5.1.
Reservoir- Condition CO2-Acidified Brine Flow: Influence of Chemical and Physical
Liu, Jiangfeng, Song, S., Cao, X., Meng, Q., Pu, H., Wang, Y., Liu, Jianfeng, 2020.
Heterogeneity. Imperial College London. https://doi.org/10.25560/68513.
Determination of full-scale pore size distribution of Gaomiaozi bentonite and its
Arns, C.H., 2004. A comparison of pore size distributions derived by NMR and X-ray-CT
permeability prediction. J. Rock Mech. Geotech. Eng. 12, 403–413. https://doi.org/
techniques. Phys. A Stat. Mech. its Appl. 339, 159–165. https://doi.org/10.1016/j.
10.1016/j.jrmge.2019.12.005.
physa.2004.03.033.
Lopez-Sanchez, M.A., 2018. GrainSizeTools: a Python script for grain size analysis and
Bagherzadeh, R., Najar, S.S., Latifi, M., Tehran, M.A., Kong, L., 2013. A theoretical
paleopiezometry based on grain size. J. Open Source Softw. 3, 863. https://doi.org/
analysis and prediction of pore size and pore size distribution in electrospun
10.21105/joss.00863.
multilayer nanofibrous materials. J. Biomed. Mater. Res. Part A 101A 2107–2117.
Luo, S., Lutkenhaus, J.L., Nasrabadi, H., 2019. Experimental study of pore size
https://doi.org/10.1002/jbm.a.34487.
distribution effect on phase transitions of hydrocarbons in nanoporous media. Fluid
Bai, B., Zhu, R., Wu, S., Yang, W., Gelb, J., Gu, A., Zhang, X., Su, L., 2013. Multi-scale
Phase Equil. 487, 8–15. https://doi.org/10.1016/j.fluid.2018.11.026.
method of Nano(Micro)-CT study on microscopic pore structure of tight sandstone of
Mickel, W., Münster, S., Jawerth, L.M., Vader, D.A., Weitz, D.A., Sheppard, A.P.,
Yanchang Formation, Ordos Basin. Petrol. Explor. Dev. 40, 354–358. https://doi.
Mecke, K., Fabry, B., Schröder-Turk, G.E., 2008. Robust pore size analysis of
org/10.1016/S1876-3804(13)60042-7.
filamentous networks from three-dimensional confocal microscopy. Biophys. J. 95,
Bardestani, R., Patience, G.S., Kaliaguine, S., 2019. Experimental methods in chemical
6072–6080. https://doi.org/10.1529/biophysj.108.135939.
engineering: specific surface area and pore size distribution measurements—BET,
Minagawa, H., Sakamoto, Y., Komai, T., Narita, H., Mizutani, K., Ohga, K., Takahara, N.,
BJH, and DFT. Can. J. Chem. Eng. 97, 2781–2791. https://doi.org/10.1002/
Yamaguchi, T., 2009. Relation between pore-size distribution and permeability of
cjce.23632.
sediment. In: Proceedings of the International Offshore and Polar Engineering
Bartoš, M., Suchý, T., Foltán, R., 2018. Note on the use of different approaches to
Conference, pp. 25–32.
determine the pore sizes of tissue engineering scaffolds: what do we measure?
Occelli, M.L., Olivier, J.P., Petre, A., Auroux, A., 2003. Determination of pore size
Biomed. Eng. Online 17, 110. https://doi.org/10.1186/s12938-018-0543-z.
distribution, surface area, and acidity in fluid cracking catalysts (FCCs) from
Berger, A., Herwegh, M., Schwarz, J., Putlitz, B., 2011. Quantitative analysis of crystal/
nonlocal density functional theoretical models of adsorption and from
grain sizes and their distributions in 2D and 3D. J. Struct. Geol. 33, 1751–1763.
microcalorimetry methods. J. Phys. Chem. B 107, 4128–4136. https://doi.org/
https://doi.org/10.1016/j.jsg.2011.07.002.
10.1021/jp022242m.
Blunt, M.J., Bijeljic, B., Dong, H., Gharbi, O., Iglauer, S., Mostaghimi, P., Paluszny, A.,
Otsu, N., 1979. A threshold selection method from gray-level histograms. IEEE Trans.
Pentland, C., 2013. Pore-scale imaging and modelling. Adv. Water Resour. 51,
Syst. Man. Cybern. 9, 62–66. https://doi.org/10.1109/TSMC.1979.4310076.
197–216. https://doi.org/10.1016/j.advwatres.2012.03.003.
Palavecino, M., Torres-Verdín, C., Strobel, J., 2018. Grain-size distribution, grain
Burdine, N.T., 1953. Relative permeability calculations from pore size distribution data.
arrangement, and fluid transport properties: an integrated rock classification method
J. Petrol. Technol. 5, 71–78. https://doi.org/10.2118/225-G.
for tight-gas sandstones. In: SPWLA 59th Annual Logging Symposium 2018.
Churcher, P.L., French, P.R., Shaw, J.C., Schramm, L.L., 1991. Rock properties of Berea
Peksa, A.E., Wolf, K.-H.A.A., Zitha, P.L.J., 2015. Bentheimer sandstone revisited for
sandstone, baker dolomite, and Indiana limestone. In: SPE International Symposium
experimental purposes. Mar. Petrol. Geol. 67, 701–719. https://doi.org/10.1016/j.
on Oilfield Chemistry. Society of Petroleum Engineers. https://doi.org/10.2118/
marpetgeo.2015.06.001.
21044-MS.
Reppel, T., Weinberg, K., 2019. Stereological transformation of pore size distributions
Cnudde, V., Boone, M., Dewanckele, J., Dierick, M., Hoorebeke, L. Van, Jacobs, P., 2011.
with application to soft polymer and FDM-printed specimens 1–15. https://doi.
3D characterization of sandstone by means of X-ray computed tomography.
org/10.1002/zamm.201800287.
Geosphere 54–61. https://doi.org/10.1130/GES00563.1.
Reppel, T., Korzeniowski, T.F., Weinberg, K., 2019. Stereological transformation of pore
Dalton, L.E., Tapriyal, D., Crandall, D., Goodman, A., Shi, F., Haeri, F., 2020. Contact
size distributions with application to soft polymer and FDM-printed specimens.
angle measurements using sessile drop and micro-CT data from six sandstones.
ZAMM - J. Appl. Math. Mech./Z. Angew. Math. Mech. 99, 1–15. https://doi.org/
Transport Porous Media 133, 71–83. https://doi.org/10.1007/s11242-020-01415-y.
10.1002/zamm.201800287.
Dubetz, D., Cheng, H., Zhu, D., Hill, A.D., 2016. Characterization of rock pore-size
Roostaei, M., Soroush, M., Hosseini, S.A., Velayati, A., Alkouh, A., Mahmoudi, M.,
distribution and its effects on wormhole propagation. In: Day 1 Mon, September 26,
Ghalambor, A., Fattahpour, V., 2020. Comparison of various particle size
2016. SPE. https://doi.org/10.2118/181725-MS.
distribution measurement methods: role of particle shape descriptors. In: Day 2 Thu,
Elkhoury, J.E., Shankar, R., Ramakrishnan, T.S., 2019. Resolution and limitations of X-
February 20, 2020. SPE, pp. 19–21. https://doi.org/10.2118/199335-MS.
ray micro-CT with applications to sandstones and limestones. Transport Porous
Sahagian, D.L., Proussevitch, A.A., 1998. 3D particle size distributions from 2D
Media 129, 413–425. https://doi.org/10.1007/s11242-019-01275-1.
observations: stereology for natural applications. J. Volcanol. Geoth. Res. 84,
Gaafar, G.R., Altunbay, M., Bal, A., Anuar, N.B., 2014. Ascendancy of continuous profiles
173–196. https://doi.org/10.1016/S0377-0273(98)00043-2.
of grain-size distribution for depositional environment studies. In: All Days. IPTC.
Schmitt, M., Halisch, M., Muller, C., Fernandes, C.P., 2015. Classification and
https://doi.org/10.2523/IPTC-17754-MS.
quantification of pore shapes in sandstone reservoir rocks with 3-D X-ray micro-
Ghasemi, K., Mahmoudi, M., Roostaei, M., Fattahpour, V., Soroush, M., Nouri, A., 2018.
computed tomography. Solid Earth Discuss 7, 3441–3479. https://doi.org/10.5194/
Determination of particle shape and size distribution from micro X-ray CT scans for
sed-7-3441-2015.
petrophysical evaluation and sand control design. In: Day 1 Mon, June 25, 2018.
Shehata, A.M., Nasr-El-Din, H.A., 2015. Zeta potential measurements: impact of salinity
SPE. https://doi.org/10.2118/191193-MS.
on sandstone minerals. In: Day 1 Mon, April 13, 2015. SPE, pp. 789–805. https://doi.
Goldstein, J.I., Newbury, D.E., Michael, J.R., Ritchie, N.W.M., Scott, J.H.J., Joy, D.C.,
org/10.2118/173763-MS.
2018. Scanning Electron Microscopy and X-Ray Microanalysis. Springer New York,
Sheppard, A., Latham, S., Middleton, J., Kingston, A., Myers, G., Varslot, T., Fogden, A.,
New York, NY. https://doi.org/10.1007/978-1-4939-6676-9.
Sawkins, T., Cruikshank, R., Saadatfar, M., Francois, N., Arns, C., Senden, T., 2014.
Gostick, J.T., 2017. Versatile and efficient pore network extraction method using marker-
Techniques in helical scanning, dynamic imaging and image segmentation for
based watershed segmentation. Phys. Rev. E 96, 023307. https://doi.org/10.1103/
improved quantitative analysis with X-ray micro-CT. Nucl. Instrum. Methods Phys.
PhysRevE.96.023307.
Res. Sect. B Beam Interact. Mater. Atoms 324, 49–56. https://doi.org/10.1016/j.
Higgins, M.D., 2000. Measurement of crystal size distributions. Am. Mineral. 85,
nimb.2013.08.072.
1105–1116. https://doi.org/10.2138/am-2000-8-901.
Song, Shuaibing, Ding, Q., Wei, J., 2019. Improved algorithm for estimating pore size
Higgins, M.D., 2015. Measurement of crystal size distributions. https://doi.org/10.2138/
distribution from pore space images of porous media. Phys. Rev. E 100, 053314.
am-2000-8-901.
https://doi.org/10.1103/PhysRevE.100.053314.
Iassonov, P., Gebrenegus, T., Tuller, M., 2009. Segmentation of X-ray computed
Song, Shuai-bing, Liu, J., Yang, D., Ni, H., Huang, B., Zhang, K., Mao, X.-B., 2019. Pore
tomography images of porous materials: a crucial step for characterization and
structure characterization and permeability prediction of coal samples based on SEM
quantitative analysis of pore structures. Water Resour. Res. 45 https://doi.org/
images. J. Nat. Gas Sci. Eng. 67, 160–171. https://doi.org/10.1016/j.
10.1029/2009WR008087.
jngse.2019.05.003.
Ji, Y., Baud, P., Vajdova, V., Wong, T. -f., 2012. Characterization of pore geometry of
Srisutthiyakorn, N., Mavko, G., 2019. Computation of grain size distribution in 2-D and
Indiana limestone in relation to mechanical compaction. Oil Gas Sci. Technol. – Rev.
3-D binary images. Comput. Geosci. 126, 21–30. https://doi.org/10.1016/j.
d’IFP Energies Nouv. 67, 753–775. https://doi.org/10.2516/ogst/2012051.
cageo.2019.01.019.
Jin, S., Zhou, J., Zhao, X., Sun, L., 2021. Quantitative relationship between pore size
Tian, W., Lu, S., Huang, W., Wang, S., Gao, Y., Wang, W., Li, J., Xu, J., Zhan, Z., 2019.
distribution and compressive strength of cementitious materials. Construct. Build.
Study on the full-range pore size distribution and the movable oil distribution in
Mater. 273, 121727. https://doi.org/10.1016/j.conbuildmat.2020.121727.
glutenite. Energy Fuels 33, 7028–7042. https://doi.org/10.1021/acs.
Jutzeler, M., Proussevitch, A.A., Allen, S.R., 2012. Grain-size distribution of
energyfuels.9b00999.
volcaniclastic rocks 1: a new technique based on functional stereology. J. Volcanol.
Udenni Gunathilake, T., Ching, Y., Ching, K., Chuah, C., Abdullah, L., 2017. Biomedical
Geoth. Res. 239–240, 1–11. https://doi.org/10.1016/j.jvolgeores.2012.05.013.
and microbiological applications of bio-based porous materials: a review. Polymers
Kenvin, J., Jagiello, J., Mitchell, S., Pérez-Ramírez, J., 2015. Unified method for the total
9, 160. https://doi.org/10.3390/polym9050160.
pore volume and pore size distribution of hierarchical zeolites from argon adsorption

13
H. Safari et al. Computers and Geosciences 156 (2021) 104895

Wicksell, S.D., 1925. The corpuscle problem. A mathematical study of a biometric Zhang, Y., Jin, S., Wang, Y., Wang, Y., 2015. Characterization of the pore size
problem. Biometrika 17, 84–99. https://doi.org/10.1093/biomet/17.1-2.84. distribution with SEM images processing for the tight rock. In: 2015 IEEE
Withers, P.J., 2007. X-ray nanotomography. Mater. Today 10, 26–34. https://doi.org/ International Conference on Information and Automation, pp. 653–656. https://doi.
10.1016/S1369-7021(07)70305-X. org/10.1109/ICInfA.2015.7279367.
Yang, Y., Liu, P., Zhang, W., Liu, Z., Sun, H., Zhang, L., Zhao, J., Song, W., Liu, L., An, S., Zhang, Z., Kruschwitz, S., Weller, A., Halisch, M., 2018. Enhanced pore space analysis by
Yao, J., 2016. Effect of the pore size distribution on the displacement efficiency of use of μ-CT, MIP, NMR, and SIP. Solid Earth 9, 1225–1238. https://doi.org/
multiphase flow in porous media. Open Phys. 14, 610–616. https://doi.org/ 10.5194/se-9-1225-2018.
10.1515/phys-2016-0069.

14

You might also like