Sunteți pe pagina 1din 31

Marine and Petroleum Geology 114 (2020) 104141

Contents lists available at ScienceDirect

Marine and Petroleum Geology


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

Review article

Pore-scale modeling of carbonates T


Ayaz Mehmani , Rahul Verma, Maša Prodanović

Hildebrand Department of Petroleum and Geosystems Engineering, The University of Texas at Austin, Texas, USA

ARTICLE INFO ABSTRACT

Keywords: Carbonate rocks display complex pore textures with heterogeneities that extend over several length scales.
Pore-scale modeling Identifying and capturing their significant pore space properties in a computationally tractable model is
Carbonates therefore nontrivial. Pore-scale models bridge predictions of transport properties from sub-pore to core scale
Petrophysics while attempting to render such methods cost effective. This chapter reviews several methods for fabricating
Digital rock physics
pore-scale models of carbonate rocks and discusses approaches to simulating their fluid flow and solute transport
Pore-structure reconstruction
properties. The advantages and limitations of each method in reducing the reservoir development uncertainties
are explained. We argue that pore-scale models can play an important role in predicting the petrophysical
properties of carbonate rocks. Even more importantly, they provide sensitivity analyses of pore-scale features on
macroscopic transport observations and offer explanations for anomalous flow behaviors.

1. Introduction priori are due to the fact that traditional relationships such as Archie's
equation relating resistivity to saturation, Brooks-Corey for relative
The length scales involved in modeling subsurface formations span permeability and Carman-Kozeny for absolute permeability (Peters,
from sub-micrometers/micrometers (size of a pore) to meters/kilo- 2012) were originally derived for rocks with well sorted grains and
meters (e.g. formation thickness or distance between wells). Although it narrowly distributed pores such as unconsolidated sandstones. Such
is desirable to conduct transport simulations solely on the field scale, relationships are erroneous for carbonates because of ranges in pore
typically sparse formation evaluation datasets (such as log and core size distribution that span orders of magnitude and high heterogeneity
data) may not be sufficient to statistically extrapolate necessary trans- in pore connectivity. It becomes therefore imperative to delve into the
port properties to the entire reservoir domain. In general, a posteriori pore scale to gain a fundamental understanding of flow through porous
prediction of the fluid behavior in subsurface porous media is proble- media for finding closure relationships.
matic given the wide range of values in core-scale macroscopic mea- Pore-scale modeling is rooted in first principles. If the pore structure
surements and their spatial heterogeneity in a given reservoir. Although is correctly represented and the initial conditions, such as initial contact
constraining the data based on lithology or rock typing schemes im- angle values, and details of a physical phenomenon, such as evolution
proves predictions, obtaining such classes is challenging when it comes of wettability during two-phase flow, are understood, then any emer-
to heterogeneous formations such as carbonates. gent macroscopic fluid behavior, such as relative permeability, can be
In addition to rock typing, determining the transport properties captured accurately. Pore-scale models however require to be “up-
based on direct formation evaluation can become complicated in car- scaled” to reservoir grid domains which can become a nontrivial pro-
bonates as commonly used formulas needed to interpret log/core blem. In addition, uncertainties of the relevant input information for a
measurements may result in high errors. For instance, researchers have pore-scale model, such as in situ contact angle values, inevitably create
pointed out that utilizing the Carman-Kozeny model for predicting tuning parameters that will require adjustments to match macroscopic
absolute permeability from porosity measurements in carbonates re- experimental measurements.
quires fitting the model to core data in order to incorporate a quanti- Pore-scale modeling allows direct identification and quantification
fication of the “rock fabric” as an independent parameter. For accurate of parameters related to pore structure and fluid properties for linking
predictions, the model needs to be continuously adjusted as data be- the rock transport properties with geological information and estab-
come available (Jennings and Lucia, 2001) – which are not necessarily lishing reliable rock classes. Once corroborated, pore-scale models can
available. offer data points away from wells where empirical relationships are
The uncertainties in predicting carbonate transport properties a unavailable. Conducting experimental measurements is difficult in tight


Corresponding author.
E-mail address: mehmani.ayaz@utexas.edu (A. Mehmani).

https://doi.org/10.1016/j.marpetgeo.2019.104141
Received 27 July 2019; Received in revised form 7 November 2019; Accepted 14 November 2019
Available online 20 November 2019
0264-8172/ © 2019 Published by Elsevier Ltd.
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 1. Schematic of the application of pore-scale mod-


eling. The green line depicts a well log (such as porosity)
identifying the flow unit (a lithologically preferable zone
for production). At a particular flow unit, core plugs are
extracted and images of them acquired to quantify the
pore structure. The information is fed to a pore-scale
model in order to determine the relationship between
several macroscopic properties. These relationships are
used subsequently in other flow units. Pc: capillary
pressure; Sw: water saturation; ϕ: porosity; K: perme-
ability; XCT: X-ray computed tomography; SEM: scanning
electron microscopy. (For interpretation of the references
to color in this figure legend, the reader is referred to the
Web version of this article.)

media. Heavily cemented but microporous limestones certainly fall into geometry preparation in various pore scale modeling methods, as well
this category for which obtaining experimental measurements become as highlight the transport modeling methods themselves. That is a hard-
scarce due to time and financial limitations. A cost-effective pore-scale balancing act, and we apologize to any researchers in advance for not
model can optimize the number of experimental measurements needed referencing their papers.
for linking petrophysical rock types to depositional classifications Since pore-scale models hinge upon capturing key characteristics of
(Chandra et al., 2015; Skalinski and Kenter, 2015; van der Land et al., the pore geometry, we specifically discuss pore geometry generation
2013). methods that follow the geological history of the rock, utilize three-
A robust pore-scale model provides general understanding of re- dimensional image acquisition methods, or extrapolate two-dimen-
lationships between various petrophysical properties and is therefore sional information to three dimensions via statistical algorithms.
valuable for long term reservoir development. An a priori knowledge Integration of pores from multiple length scales can be a key element in
obtained through pore-scale modeling of fluid phase distributions can correctly capturing macroscopic behavior of carbonates and therefore
assist in establishing the performance of water flooding, surfactant must be addressed in each generation method. We subsequently discuss
flooding, and the efficacy of wettability altering chemicals during im- pore-network modeling and direct numerical simulation as two com-
proved oil recovery (Lerdahl et al., 2000). In addition such a model can plementary approaches for upscaling sub-pore scale transport dynamics
clarify the potential of environmental operations by predicting the fate to a representative elementary volume (REV) or a volume approaching
of pollutants during hazardous waste storage, determining NAPL re- it. The REV is a minimum volume for which continuum assumption of a
covery, and capacity of formations for carbon storage. particular rock property is valid and was first introduced by Bear
Fig. 1 shows an example of utilizing pore-scale modeling in for- (1972) and Hill (1963). As is discussed later, pore-network models are
mation evaluation. During well log acquisition, core plugs from a flow computationally more effective than direct numerical simulation
unit can be extracted to construct a pore-scale model. The model can methods; but both are viable for making a priori predictions of a
subsequently be used to establish the relationships between several transport property if they capture the correct pore geometry, fluid/fluid
macroscopic petrophysical properties. Such relationships can then be (for example interfacial tension) and fluid/rock conditions (for example
deployed to other flow units for which laboratory measurements are wettability). The ultimate choice of the method to be used depends on
expensive and not feasible. the available information for constraining the input parameters as well
This paper reviews methods and challenges in constructing pore- as computational budget associated with the physical process to be
scale models representative of rocks and elaborates on the various ap- modeled: the methods should be viewed as complementary rather than
proaches used in simulating their fluid/solute behavior. Fig. 2 illus- competing. In addition, direct numerical simulations can determine
trates the topics covered in this paper and their contributions to the input variables for pore-network models and therefore assist in further
creation of a pore-scale model. We start by covering three major pore- upscaling macroscopic calculations (Raeini et al., 2014a).
space generation methods. Subsequently a discussion on pore-network
modeling and direct numerical simulation of fluid and solute through 2. Methods
the reconstructed pore space is provided. Algorithms for integrating
pores from multiple length scales is discussed next. We note that direct 2.1. Pore geometry creation
numerical simulations can be used to benchmark reduced-ordered pore-
network models as well and a synthesis of these two methods is given in Capturing pore geometry is essential to fabricating a reliable pore-
Section 3 (Discussion). We focus on challenges in developing pore-scale scale model. Three approaches that pursue this purpose gather their
models for carbonate rocks as much as possible. Each section is written information from images (image-based method), replicate the geologic
such that the reader can obtain the most relevant concepts and litera- processes of the rock (process-based method) or are based on statistical
ture independently from other sections. When a relevant review article description of the porous medium (stochastic).
is referenced, we provide a succinct summary of already reviewed
concepts (to the extent that the reader can gather the most pertinent 2.1.1. Imaging as a source of information for rock models
information) and discuss supplemental topics not synthesized in the Imaging techniques have the potential for providing detailed de-
literature. We have attempted to review a wide range of methods for scriptions of the pore space that can be utilized for numerical

2
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 2. Topics covered in this paper and their contributions


to the creation of a pore-scale model. Since pore-network
modeling is a reduced-ordered method, direct numerical
simulations can not only be used for predicting the rock
sample's transport properties, but also inform pore-network
modeling algorithms. In carbonates, similar to other un-
conventional systems, experimental data is not always
available for benchmarking purposes. Since direct numer-
ical simulations sacrifice pore space geometry and fluid flow
behaviors to a lesser extent than pore-network models, they
can serve a valuable purpose in reducing the latter's un-
certainty.

simulation schemes. X-ray computed tomography (X-ray CT) is a well- network representation of the pore space (to be discussed later in the
established and rapid approach to acquiring three-dimensional visua- chapter). In this paper, we refer to a pore space generation that is
lizations of the pore structure. The method distinguishes phases based completely based on a three-dimensional image as an “image-based
on density and atomic number variations that attenuate the X-rays pore-scale model”. As we discuss other pore space generation methods,
differently. X-ray CT is non-destructive, meaning that unlike other overlaps between the three methods exist and our classification serves
imaging techniques such as petrographic light microscopy and scanning as a point of reference rather than a rigid framework.
electron microscopy, the rock sample is not cut or polished. The re- Light microscopy and scanning electron microscopy are two-di-
solution of X-ray CT ranges from 100 nm to 1 mm (Bultreys et al., 2016) mensional image acquisition techniques that provide micrometer and
and has an inverse relationship with the domain size. nanometer scale information respectively (see Fig. 4a–b for an example
Once a three-dimensional visualization is obtained and processed of each). For light microscopy, the pore space of a rock sample is filled
such that its quality is acceptable, the image is “segmented” into phases with blue epoxy and a 30-μm-thick thin-section is cut. Polarized light is
of interest, normally the pore space and the solid phase. Any pore-scale emitted through the sample to determine the grain mineralogy, visible
numerical analysis of the image is thus conducted based on the seg- porosity and paragenesis (the chronology of geological events occurring
mented image. Image processing, although conceptually simple, can be during rock formation). For scanning electron microscopy, electrons are
challenging in geomaterials due to the presence of artefacts such as bombarded on a polished carbon coated rock surface. The surface
beam hardening (Van de Casteel et al., 2002) or when the sample is not scattered electrons (called secondary electrons) and those that are
held in a stable position properly. Given the many sources of un- backscattered due to elasticity, which carry information from a higher
certainty in processing a micro-CT image, completely automatic pro- depth, are analyzed to infer the surface topography of the solid phase.
cessing workflows are not yet established, leaving many image filtering Using similar bombardment, mineral atoms can be excited which re-
and segmentation attempts dependent on the operator's judgement lease X-rays when they return to their stable state. This is referred to as
(Iassonov et al., 2009). The current subjective nature of image analysis Energy Dispersive Spectroscopy (EDS) which gives information of the
highlights the importance of core data measurements to validate a mineral composition of the rock. The disadvantage of utilizing light and
processed image prior to using it to predict unavailable transport scanning electron microscopy is the fact that they are two-dimensional.
properties. Such validations may become especially important in car- Numerical methods for extrapolating such information to three-di-
bonates given the higher sensitivity of near-resolution pores of be- mensions are non-trivial and will be discussed further in Section 2.1.3
coming dismissed during image filtering and segmentation. Carbonates as part of stochastic methods. Confocal laser scanning microscopy
present an additional challenge for capturing their pore-space in a (CLSM), with a prior application in biology, has been leveraged to
single micro-CT image given their multiscale pore structure. Pore scales quantify the pore texture of rocks (Fredrich, 1999; Fredrich et al.,
can range from sub-resolution nanometer sized pores to interparticle 1995). In this method, a laser beam is focused on the domain of interest
pores and up to larger scale vugs and microfractures. This range in scale and its reflection is detected and analyzed. This method produces sub-
can make the capture of carbonate pore spaces in a single image un- micron lateral resolution with lower vertical resolution depending on
feasible (Victor, 2017). To circumvent this limitation, a top-to-bottom material type and magnification (with the higher magnification having
approach can be adopted where following a coarse-scale three-dimen- larger vertical resolutions). Given the limitations in vertical resolution,
sional CT image, high resolution micro-CT images from targeted regions two-dimensional CLSM images can benefit from extrapolation techni-
are acquired (Arns et al., 2005). A combination of other imaging ques discussed in Section 2.1.3 as well. Recent work by Jobe et al.
methods to gather the higher resolutions information can be im- (2018) has shown CLSM's application in acquiring two-dimensional
plemented as well. The transport properties determined from the high- images of micropores in carbonate samples (Fig. 4c) and utilizing them
resolution pore structures are subsequently scaled up to establish the to capture their pore topology with stochastic extrapolation methods
transport property of the larger domain (we provide a discussion on this (Wu et al., 2006).
topic in Section 2.3). For a complete review on image processing refer Focused ion beam scanning electron microscopy (FIBSEM) is a
to Heilbronner and Barrett (2013), Serra (1982), and Wildenschild and technique with nanometer scale resolution for capturing the pore tex-
Sheppard (2013). Fig. 3 shows an example of (a) a micro-CT image, (b) ture of sub-resolution micro-CT pores in three dimensions (Bera et al.,
pore-solid surface visualization after segmentation and (c) a pore- 2012; Wirth, 2009). Fig. 5 shows an example FIBSEM image stack of a

3
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 3. (a) Example of micro-CT images of a Ketton limestone sample at a resolution of 2.65 µ m (b) pore-solid surface extracted after segmentation and (c) a pore-
network representation of the pore space (Blunt et al., 2013).

chalk sample. In FIBSEM the sample is sliced with ion beams and an diagnostic purposes and as a supplemental tool along with other image
SEM image is taken for each stack. Although valuable information can acquisition methods for carbonates (Sok et al., 2010). When micro-CT
be gained with this imaging modality, namely the 3D interconnectivity imaging becomes insufficient to reveal the sub-resolution pore space,
of pores, the financial and time cost of acquiring a FIBSEM image stack FIBSEM can provide information on the presence of critical pore con-
renders them intractable for routine use. Given the destructive nature of nectivity within sub-resolution pores as well as at the interface of sub-
FIBSEM acquisition during beam slicing, uncertainties associated with resolution and resolved pores. Furthermore, given the complex carbo-
pore space connectivity can be exasperated as well. Finally, in addition nate diagenesis that result in amorphous pore shapes, FIBSEM can
to challenges in image registration, the rendered domain is very small provide a clearer picture of pore morphology. FIBSEM uncertainties
such that a quantitative assessment of flow properties become ques- associated with ion beam slicing, electron charging and resolution
tionable (Kelly et al., 2016). FIBSEM therefore is an ideal technique for limitations (Dewers et al., 2012) can be mitigated by calibration with

Fig. 4. Examples of (a) a plane polarized light microscopy image of ooids in a carbonate rock sample – the horizontal axis is 2.0 mm (Scholle and Ulmer-Scholle,
2003), (b) an Indiana limestone secondary electron image (Freire-Gormaly et al., 2015) and (c) confocal laser microscopy images of intraparticle microporosity of
Arab D carbonate rock samples (Jobe et al., 2018).

4
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 5. (a) A two-dimensional FIB image of a chalk sample and (b) its corresponding three-dimensional reconstruction. From Tomutsa et al. (2007).

macroscopic measurements such as nitrogen-sorption hysteresis 2.1.2. Process-based rock models


(Mehmani and Prodanović, 2014a; Xu and Prodanovic, 2018). How- Process-based methods are based on replicating the chemical and
ever, much work is needed to establish robust workflows for such ca- mechanical rock formation processes on grain scale. Following this
libration purposes. approach, the pore-scale model can link formation properties with any
Data storage and management are challenges in image-based char- burial stage process throughout the diagenesis. This pore space gen-
acterization of porous media. For instance, X-ray microtomography eration method is important not only for understanding the first prin-
images typically have 20001 voxels (the storage necessary is 8 GB or ciples of the rock transport properties, but also as a practical tool in
16 GB depending on whether 1-byte or 2-byte integers are used to store extrapolating information from one portion of the reservoir to the next
attenuation coefficients for each voxel) and can triple in size if stitching (that is extrapolating transport information from one burial stage to
along axial direction or if helical scanning is conducted. Furthermore, another). Process-based implementations on image-based methods (a
recent multi-beam scanning electron microscopes (e.g. Zeiss MultiSEM) hybrid process-image-based method) is considerably difficult since
use up to 91 parallel beams and can collect areal images with ap- capturing the mechanical and chemical processes in images throughout
proximately 5-nm resolution at a scale of several square millimeters. the formation stages is not trivial. Most often it is essentially the “end
While the resulting tiled images are getting close to covering re- result” that is imaged and characterized. The starting point of process-
presentative areas (Keller et al., 2014; Kemen et al., 2015) for hetero- based methods is typically an unconsolidated grain pack. At each burial
geneous media, with a size of more than 200 GB their display and stage, pertinent diagenetic processes are included. Such processes can
analysis is challenging. Due to the resultant “big data”, a repository for be grain rearrangement and ductile deformation due to compaction,
retaining, retrieval, sharing, and organization of acquired images and cementation of various minerals as well as grain dissolution. Process-
their analyses can be valuable for enhancing research for modeling/ based approach, though powerful in its motivation, compromises
prediction of porous material properties. complexity by simplifying the grains to ideal shapes such as spheres or
One of the new research data management tools, Digital Rocks ellipsoids. This approach is in particular challenging in carbonates
Portal (Prodanovic et al., 2015a) allows managing, preserving, visua- given their complex diagenesis rendering it difficult to formulate each
lization and basic image analysis of available images of porous mate- diagenetic step to be quantitatively.
rials and experiments performed on them. It can accommodate storage In a pioneering work on process-based methods, Bryant et al.
of any accompanying measurements (porosity, capillary pressure, per- (1993b) modeled compaction and cementation on a disordered mono-
meability, electrical, NMR and elastic properties, etc.) required for both disperse grain pack. In this technique, the grain deposit is replicated
validations on modeling approaches and the upscaling and building of based on experimental measurements of ball-bearings packed in a
larger (hydro)geological models. As of the end of 2018, Digital Rocks rubber bladder (Finney, 1970, 2016). Compaction is modeled by
Portal currently published 58 samples, and is expected to have a re- overlapping the grains and for cementation, the radii are increased
presentative population of samples for every lithology. Carbonate ex- without moving the grain centers. The researchers subsequently used a
amples range from petrophysical characterization of pore/vug/fracture pore-network model to simulate permeability and drainage capillary
space, to wettability (contact angle) measurements and precipitation pressure and relative permeability curves (Bryant and Blunt, 1992;
experiments (Alhammadi et al., 2018; Baek and Tae-Hyuk Kwon, 2018; Bryant et al., 1993a). Bakke and Øren (1997) and Øren et al. (1998)
Bultreys et al., 2017a, Bultreys et al., 2017b; 2018; Heidari and developed low-energy and high-energy sedimentation modeling
Posenato Garcia, 2016; Khan et al., 2018; Khan and Pope, 2018; schemes for the initial grain pack and subsequently added cementation
Muljadi, 2015; Prodanovic et al., 2016; Scanziani et al., 2018; Singh and compaction to represent the formation of the rock. Thin-section
and Blunt, 2018; Victor and Prodanovic, 2017). While other general analysis of a Bentheimer sandstone sample was used to inform the
data storage options exist (e.g. Mendeley Data for any data,2 and En- model in a case study. In addition, Torskaya et al. (2014) have included
ergy Data Exchange3 for energy/subsurface focus), they lack tools for pore-filling and throat-filling cementations into a disordered packing
viewing 3D datasets within browser and specialized analysis or ability and studied its effects on absolute permeability and drainage fluid
to produce digital object identifiers for referencing in publications. properties. Hosa and Wood (2017) utilized a single-scale process-based
method to investigate the impacts of early calcite cementation on
porosity-permeability relationships in carbonate rocks. They differ-
entiated grains into monocrystalline (capable of syntaxial cementation)
1 and polycrystalline (with a characteristic isopachous cementation) and
http://www.palabos.org.
2
https://data.mendeley.com/.
simulated single-phase incompressible flow directly on the pore space
3
https://edx.netl.doe.gov/. with a lattice Boltzmann algorithm. Although their method captured

5
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

the log-log relationship between permeability and porosity observed in


experiments, they overestimated permeability potentially due to the
very high permeability value of the initial synthetic grain pack.
Establishing the initial grain pack is not trivial. One of the recent,
popular grain packing algorithms is cooperative sphere rearrangement
(Thane, 2006). In this method, a set of initial points in a three-dimen-
sional domain is randomly selected. Spheres of a predetermined size
distribution are grown with their centers at randomly selected points.
Each sphere grows at a rate proportional to its ultimate size. When
overlaps occur, the sphere centers are moved apart. If the removal of
overlap is not possible, a different sphere size is selected. Another ap-
proach to establish the starting point grain pack is based on a discrete
element method (DEM) in which the translation, rotation as well as
interaction of grains are numerically modeled (Cundall and Strack,
1979; Holtzman et al., 2008; Jin et al., 2003).
Torskaya et al. (2014) studied the impact of grain shapes on per-
meability, formation factor and drainage capillary pressure. They con-
cluded that permeability can have a difference of up to 1.6 times for a
sandstone sample if grain shapes are ignored. Formation factor and
drainage capillary pressure values remain within a 20% difference in
their study. Garcia et al. (2009) developed a method to study the impact Fig. 7. Two-dimensional schematic (top) of explicitly including (a) grain-filling
of irregular grain shape on single phase flow properties. Each irregular and (b) pore-filling microporosity. Red lines indicate throats that connect two
grain was a cluster of several spherical ones that made the im- different scales (that have been referred to as micro and macro network). Three-
dimensional visualizations of the pore-network models are at the bottom
plementation of DEM in simulating its sedimentation feasible. DEM has
(Mehmani and Prodanović, 2014b). (For interpretation of the references to
also been coupled with drainage capillary movements (Prodanović
color in this figure legend, the reader is referred to the Web version of this
et al., 2009) to investigate the impact of invasion forces to grain dis- article.)
placement. In other work, Mirabolghasemi et al. (2015) coupled DEM
with computational fluid dynamic modeling for single phase flow to
predict the amount of particle entrapment in porous media. Modeling methods as discussed in Section 2.1.3. A secondary smaller scale pore
ductility has been introduced by Mousavi and Bryant (2012) and Thane network was then connected with the pre-dissolution pore network
(2006) for a grain pack. In this method, a ductile grain is assumed to model following a stochastic method developed by Jiang et al. (2013b).
have a hard center. The ductile shell is penetrated by an adjacent rigid Mehmani and Prodanović (2014b) introduced novel deterministic
grain and the ductile materials are conserved by increasing the radius of methods to explicitly integrate intraparticle microporosity with inter-
the ductile grain. For carbonate rocks, several classifications for pore- particle macroporosity following geological causes. In their method to
scale process-based modeling has been introduced based on thin-section include microporosity, a network model is scaled down and planted to
observations (Mousavi et al., 2012). Fig. 6 shows three of such diage- targeted grains and interparticle pores based on information gathered
netic processes namely, cementation, fracturing and grain dissolution from light or scanning electron microscopy. The micro-network is then
that originated from sedimentary packings. Because certain types of stitched to its adjacent (interparticle) macropores. Fig. 7 shows a
diagenesis, such as dissolution and micro-fracture generation, impose schematic of the method. Mehmani et al. (2015) developed algorithms
integration of pores from multiple scales, the development of such to explicitly include vugs and microcracks adjacent to interparticle
models is not straightforward. pores. Vugs are included as larger pores centered on a randomly chosen
Recent advances in process-based methods have focused on mod- interparticle pore and subsequently connected to the vug's surrounding
eling the integration of pores over multiple length scales for both car- pores (Fig. 8a). A microcrack is included by two parallel planes with a
bonates and tight porous media. van der Land et al. (2013) used con- particular orientation that is imposed on a single-scale network model.
ceptual thin-section models of several carbonate rock types and All pores falling within the planes are removed and microcrack pores
predicted the impacts of cementation and dissolution on the perme- are added to the empty space instead (Fig. 8b). Mehmani and
ability, drainage capillary pressure and relative permeability curves. Prodanović (2014b) present several simulations of drainage and im-
Pore-network models were extracted after translating the two-dimen- bibition for networks that include microporosity. Fig. 9 presents the
sional conceptualizations into three dimensions using stereographical primary quasi-static capillary pressure drainage simulations for vuggy

Fig. 6. Examples of modeling approaches for (a) cementation (b) inclusion of fractures and (c) grain dissolution (Mousavi et al., 2012).

6
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

significant volume of a well-connected network of vugs. As was shown


by Pak et al. (2016), such percolation thresholds can be used as cali-
bration points to fine-tune multiscale pore networks by matching
mercury injection data. Systematic comparisons of two-phase flow be-
haviors of multiscale pore-network models with experimental data is
however currently pending.

2.1.3. Stochastic methods for geometry creation


Stochastic methods rely on several critical source information to
generate the pore space. This method is in contrast to image-based
methods which capture the entire pore space from a three-dimensional
image, and process-based methods, which attempt to generate the pore
space by following the diagenesis. It is important to note that the three
pore-space generation techniques can have overlaps. For instance, two-
dimensional images with process-based adjustments (van der Land
et al., 2013) can be a source for stochastic methods. Given the afore-
mentioned uncertainties in capturing the carbonate pore texture via
images and difficulties in replicating the geological processes for a
process-based approach, statistical descriptions have the potential to
constrain pore-scale models if accurate quantitative descriptions of the
pore space are provided. Stochastic approaches can in addition expand
the limited domain size of an image by extrapolating the mathematical
descriptions to larger domains. Furthermore, three-dimensional images
of a porous medium may not be available, may not provide important
information on mineralogy where microscopy is crucial or may cover
too small of a volume to infer representative information on porosity
(or any other property). In such cases, extrapolating two-dimensional
information to three-dimensions can be possible via stochastic methods.
One stochastic method is multiple-point statistics which utilizes two-
dimensional images as “training images” to construct the pore-space
pattern in three-dimensions (Caers, 2001; Okabe and Blunt, 2004;
Tahmasebi and Sahimi, 2013). In the context of predicting the transport
properties of porous media, it has shown to perform better than two
Fig. 8. (a) 2D schematic on top shows how a vug is incorporated into the pore other stochastic reconstruction methods: object-based methods, which
throat network: the grains as well as original pores are removed from the lo- relies on the geometrical descriptions of the “objects” comprising the
cation where a disk-shaped (sphere-shaped) vug is placed. 3D model on bottom porous medium such as grains, and pixel-based methods, that typically
shows the original network, as well as the model after placement of a number of aim to minimize the lower order statistical properties of a sample.
vugs. (b) 2D schematic and a 3 d example model of including a fracture (plane) Conceptually, stochastic creation can apply to both pore space
into an interparticle pore-network model (Mehmani et al., 2015). images and pore networks. For pore networks, however, Sok et al.
(2002) showed that a sole reliance on a finite number of statistical
carbonates with different porosities (ϕ) and a vug to pore scale differ- descriptors extracted from images, and attempting to replicate them by
ence of 5, and carbonates with a single microcrack with varying angles modifying a regular network without including the topological prop-
θ of the crack axis with respect to the flow direction. The initial grain erties of the pore space, can result in high errors in flow predictions.
pack of each network has a uniform radius of unit 1 and the side length Multiple point statistics for imaged pore space reconstruction, on the
of the cube is 20 units. The considerably lower percolation threshold of other hand, honors the pore space patterns and spatial relationships of
the vuggy pore-network model at a porosity of 0.62 is due to the the pore space by matching a multipoint statistical property such as the
cross-correlation function of a “training image”. Figs. 10 and 11 give

Fig. 9. Primary drainage capillary pressure curves of (a) a vuggy carbonate rock (with a scale difference of 5) and (b) a carbonate rock with a single microcrack (with
a scale difference of 5) at three axis angles with respect to the pressure gradient direction. Drainage is simulated on pore-network models described in Fig. 8.

7
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 10. Example multipoint statistics method for extrapolating two-dimensional images of a carbonate sample to three-dimensions and predicting the permeability-
porosity relationship (Karimpouli and Tahmasebi, 2016).

examples of utilizing multiple point statistics in reconstructing the pore images in a computationally effective way. The method is capable of
space of carbonate rock samples. The method's accuracy, however, re- capturing pore space heterogeneity in all three dimensions and can
lies on the representativeness of the two-dimensional training image. A store the trained hyper-parameters and reuse them for multiple reali-
two-dimensional training image, as was discussed in Section 2.1.1, zations. The researchers applied GAN to recreate Keystone limestone
might not capture the multiscale pores of a carbonate sample that exist and the statistical descriptions of pore morphology and single-phase
in close proximity due to the selected field-of-view (which has an in- permeability between the original and reconstruction were in close
verse relationship with voxel resolution). Further, image acquisition agreement. Experimental verification of such methods is currently
artefacts, as a result of milling in SEM for example, propagate to the pending.
three-dimensional reconstruction. Finally, the reconstruction extra-
polates the cross-section information (pore space shape, etc.) into the
orthogonal direction, whether that is correct or not. 2.2. Simulation methods
Although predictions of single phase permeability using multiple-
point statistics reconstructions are encouraging when compared to ex- 2.2.1. Pore-network generation schemes
perimental data (Okabe and Blunt, 2004), further work is required to Modeling the transport properties of a porous medium depends on
investigate the reliability of multiple point statistics, or other stochastic its pore structure. It is therefore imperative in a predictive model to
methods (van der Land et al., 2013; Wu et al., 2006), for multi-phase capture the pore space of a rock as accurately as possible. Once the
flow behaviors. This necessity for additional work is because multi- pore-space is fully resolved, so-called “direct numerical” methods can
phase transport properties in general are more sensitive to pore struc- be implemented to conduct multiphase flow simulations (see Section
tural properties. Recent work on stochastic methods have involved 2.2.4). A major challenge in direct numerical methods is the compu-
machine learning techniques. Mosser et al. (2017) for example utilized tational cost that comes with the detailed description of the pore space.
a generative adversarial neural network (GAN) method to reconstruct Pore-network modeling was first introduced by Fatt, 1956, Fatt, 1956,
three-dimensional pore space representations using training micro-CT Fatt (1956a, 1956b, 1956c) to introduce interconnectivity and three-
dimensional spatial organization; which was in contrast to a previous

Fig. 11. (a) A binarized image of a back scattered electron image of a carbonate rock sample (512 × 512 with 0.345 µ m/pixel) and (b) a three-dimensional
reconstruction of it using a multi-point statistical reconstruction method (Okabe and Blunt, 2005). Predicted permeabilities for carbonates have shown to over-
estimate by a factor of 3 when compared with experimental data but were considered acceptable given the model and experimental domain size differences (Okabe
and Blunt, 2004).

8
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

representation of pore space as bundle of tubes. Although pore-network throat channel and those are often taken as straight lines between pore
modeling is conceptually represented as three-dimensional diagrams centroids (though in reality, those paths are not straight). Some level of
composed of balls (for pores) and sticks (for throat), details for their arbitrariness thus exists in establishing image-based pore-throat net-
generation vary. Fluid interaction rules in pore-network modeling are works. In an extension of medial-axis pore partitioning algorithms,
based on analytical derivations of meniscus instability based on ex- Jiang et al. (2017) have extracted pore-network models from fractured
perimental fluid dynamic observations which were pioneered by rocks and demonstrated good agreement with analytical equations and
Lenormand and Zarcone (1984). Although in its early conception pore- direct numerical simulations for single phase flow and experimental
network modeling was considered a computationally cost-effective values for drainage residual saturations. For unconsolidated/granular
method for fluid flow predictions in porous media, with the advance- packs (in general, more porous materials) watershed and/or Delaunay
ments in computational power, modern pore-network models have tessellation perform better.
become essentially upscaling schemes that incorporate sub-pore scale In maximal balls algorithm (Silin et al., 2003), Euclidian distance
physics derived from either direct numerical simulations or experi- labels are interpreted as a radius of a maximal ball that can be inscribed
ments (Aghaei and Piri, 2015; Zolfaghari and Piri, 2017a, Zolfaghari inside of the pore space at that point. Subsequently, the balls are pro-
and Piri, 2017b). Pore-network models and direct numerical simulation cessed to identify intersections. For two balls that intersect, the smaller
should therefore not be considered as rivals. one is identified as a “slave” of the larger (“master”). At the end of the
In a pore-network model, throats are defined by Bear (1991) as fluid algorithm, master balls represent pore centers. If two master balls have
pathway cross-sections with minimal hydraulic radius (the ratio of area two intersecting slave balls, that location (and radius) identifies a
to perimeter of the cross-section). Pores are then spaces enclosed by throat. If throat location and radius are ample information for char-
grains and throats. Although the concept of a pore-network re- acterization and/or modeling, these algorithms work well. The actual
presentation is simple, dividing any given pore space into pores sepa- throat barrier such as the ones exemplified in Fig. 15 (or those by
rated by throats in a unique way is non-trivial. Fig. 12 demonstrates watershed algorithm in Appendix) are more difficult to obtain.
how an image-based pore-network construction algorithm can add Watershed partitioning locates throat surfaces and is relatively easy
uncertainties in numbers of pores. However, as mentioned in the cap- to implement. This makes it one of the most popular algorithms: we
tion of Fig. 12, regardless of the local choice made in identifying a local describe it next and provide an example in Appendix. In the watershed
throat (and/or whether it separates two pores), the flow simulation pore partitioning algorithm, the binarized image is processed to pro-
should ultimately make that local choice irrelevant. duce the Euclidean distance field (i.e. each voxel of the phase of interest
We next describe several generations of pore partitioning algorithms is labeled by the distance to the closest voxel of the other phase). The
for image-based porous media. We start with medial-axis algorithms. Euclidean distance is thought of as the topographic height function and
Maximal balls and watershed algorithms are described next. The latter inverted so that local maxima become local minima. The algorithm
two use intuitive ideas to partition pores and are less computationally partitions the phase of interest into “watershed basins”: water dropped
demanding. Thus, they are currently most commonly used in the digital inside a region will end up at the same low point. This type of algorithm
rock petrophysics community. is widely implemented in software (e.g. ImageJ, MATLAB, Avizo) and is
Early pore-throat network extraction methods utilize medial axis as used in a number of partitioning applications, e.g. for separating in-
a search tool (Liang et al., 2000; Lindquist and Venkatarangan, 1999). dividual grains in a sand or sandstone image (Thompson et al., 2006)
After image filtering and segmentation, the medial axis (Lee et al., and pore-throat networks (Sheppard et al., 2006). Since watersheds are
1994), that is, the pore space skeleton that runs approximately in its very intuitive and widely implemented, we provide an example in the
middle, is extracted. Throats are identified on a medial axis path that Appendix. All watershed algorithms suffer from noise, however, in the
has a continuous sequence of grains/particles and minimal area/peri- sense that every local minimum (maximum in the original image) will
meter (Lindquist et al., 2005; Shin et al., 2005) along that path is re- produce a corresponding region. Thus, one has to filter the number of
cognized as the throat perimeter. Pores are subsequently determined starting local minima in some way in order to produce meaningful
following a labeling algorithm after throat barriers are identified in partitioning or find a criterion for region merging afterwards (see 3D
digital space (Prodanović et al., 2006) – see Fig. 13 for a schematic from example in the Appendix for example). If one was to compare to medial-
(Mehmani et al., 2011), and Fig. 14 and Fig. 15 for the example 3D axis based search algorithm result in Fig. 14, the local minima corre-
throat and pore. The algorithms in this group work well for con- sponding to the medial axis branching points would have to be merged
solidated materials where continuous path of grain voxels bounding a or else the algorithm would result in labeling two pores instead of one.
throat barrier (see Fig. 14a) can be identified in a desired location. Such discrepancies in pore indexing should not affect the average flow
Despite the throat channel visualization shown in Fig. 14, its limits are properties.
somewhat arbitrary. The algorithm finds each throat surface and barrier For granular materials or weakly consolidated sandstones with re-
(barrier refers to a collection of voxels intersected by a throat surface) latively narrow distribution of grain sizes, a simple pore-network gen-
but there are no clear criteria based on an image that can identify the eration method based on the early work of Finney (1970) can be used.
beginning and end of a corresponding throat channel. This uncertainty This method, which is very suitable for process-based modeling, assigns
is the case for most algorithms. Either pore-throat-network visualiza- the pores and throats to the interiors and faces of tetrahedra identified
tion or pore-network models, however, require definition of a straight in tessellation of grain centers.

Fig. 12. In each of these conceptual cases


(figure reprinted from (Prodanović et al.,
2006)), pore partitioning algorithms will
likely place a throat labeled “a-b” given the
local minimum of hydraulic radii. From the
point of drainage, however, case (a) acts
like a single pore, and the fact that it was
identified as two pores does not make a
difference in averaged flow properties. Case
(b) can be considered having two pores or
be only one pore and case (c) is a two-pore
system.

9
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 13. Image-based pore-throat network extraction. After segmenting the raw image into grain and pore spaces, the medial axis of the pore space is extracted using
a thinning algorithm (Lee and Chu, 1994). The medial axis is then used to identify throat barriers and partition pores.

Fig. 16 shows a schematic of extracting a pore-throat network from


a grain pack. Some merging of the pores (identified by corresponding
overlap of inscribed spheres) is typically necessary (Al-Raoush et al.,
2003). The above pore-network generation algorithms are general and
any adjustment to the pore-network that is required to establish a
model more suitable for a carbonate sample (such as integration of sub-
resolution micropores or inclusion of rugosity for film flow) should be
implemented on the extracted pore-network.

2.2.2. Pore-network modeling of single-phase flow and solute transport


Mechanisms contributing to the transport of a passive tracer are
advection and molecular diffusion. Their relative dominance is quan-
Fig. 15. An example pore from a sandstone image with throats barriers, medial
tified by the dimensionless Péclet number (vL/D), where v is fluid ve- axis paths in the search are identified in red. Note that there are two branching
locity, L is a characteristic length scale and D is the bulk fluid diffu- points of the medial axis identified within this pore: the branching points of
sivity. When majority of solute particles have had the opportunity to medial axis and pore centers are not in one-to-one correspondence. (For in-
experience the entire range of interstitial fluid velocity, in other words, terpretation of the references to color in this figure legend, the reader is referred
solutes have traveled beyond the correlation length of the porous to the Web version of this article.)
medium, the traversing plume takes the form of a Gaussian distribution.
In this scenario, diffusion follows Fick's law in that mass diffusion rate models, species balance is imposed at each pore. It is also often assumed
becomes linearly proportional to the concentration gradient. In addi- that pore concentrations are uniform, which can potentially lead to
tion, diffusion length scales with the square root of time and the errors in high Péclet numbers. This is referred to as the Mixed Cell
average solute velocity equals to Darcy velocity. This solute transport Method (MCM) (Acharya et al., 2007; Bryntesson, 2002; Kim et al.,
behavior is referred to as “Fickian”. Due to the complex and hetero- 2011; Li et al., 2006; Mehmani et al., 2012; Nogues et al., 2013). Y.
geneous pore structure of carbonates, solute transport oftentimes takes Mehmani et al., 2014, improved MCM by accounting for imperfect
a “non-Fickian” form and its prediction has been the focus of pore-scale mixing within pores using a streamline splitting based method. Sig-
modeling efforts (Bijeljic et al., 2004, 2011). nificant improvements in macroscopic transverse dispersion were re-
The direct modeling of mass transport requires the solution of the ported in 2D ordered micromodels. The improvements however were
Navier-Stokes equation coupled with a set of species’ mass balance minor for 3D disordered sphere packs. Mehmani and Balhoff, (2015a)
equations. The numerical framework can either be grid-based addressed another shortcoming of MCM by incorporating shear-induced
(Eulerian) or based on tracking Brownian particles within the pore dispersion caused by non-uniform velocity profiles within pore throats.
space (Lagrangian). Given the computationally intensive nature of di- The result was an improvement in macroscopic longitudinal dispersion
rect methods, pore-network models are often used for modeling single- in disordered media at high Péclet numbers. Lagrangian methods,
phase and solute transport, although direct numerical simulation of which are more accurate but computationally expensive for continuous
single-phase flow has recently become more tractable. Prior to mod- injection problems, require the tracking of particles via discrete (Bijeljic
eling transport, the flow field is computed. In Eulerian pore-network

Fig. 14. (a) Medial axis path voxels extracted from a 3D Berea sandstone image in red with a throat perimeter voxels identified in brown. Grain and pore voxels are
not shown. Voxels in blue are throat barrier voxels used in subsequent pore labeling. (b) Corresponding triangulated throat surface is shown in white. (c) Dark red
shows the pore-grain surface corresponding to the channel where the throat (surface shown in white) is identified. (For interpretation of the references to color in this
figure legend, the reader is referred to the Web version of this article.)

10
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 16. Schematic of Delaunay tessellation


for identifying a pore-throat network in
granular media. (a) Delaunay tessellation
input is a packing of grains (spheres in this
work). (b) Tetrahedral cell representing a
pore defined by four neighboring spheres.
The pore space within each tetrahedron re-
presents pore, and the tetrahedral sides are
the tightest cross-sections (throats) on the
path between two neighboring pore centers.
(c) The final result is a network of pores
(shown as spheres, colored/shaded by size)
and throats (identified as cylindrical con-
nections). Figure modified after Mehmani
and Prodanović, 2014b.

et al., 2004; Bruderer and Bernabé, 2001; Jha et al., 2011) or con- quasi-static or dynamic methods to model fluid flow, this is an im-
tinuous temporal transition probabilities (Rhodes et al., 2009; Sahimi portant feature. As one goes down the rows of the table, towards
et al., 1986; Sorbie and Clifford, 1991). However, they have shown to more modern pore networks, the physics becomes increasingly so-
reliably predict the non-Fickian behavior of carbonate rocks when im- phisticated, incorporating both two- and three-phase flow, wett-
plemented directly on pore-scale images (DNS method) (Bijeljic et al., ability and trapping.
2013). Mehmani and Tchelepi (2017) showed that a pore-network
framework for solute transport at moderate to high Péclet numbers is The table reveals a number of interesting trends in pore network
accurate only if (1) network extraction is informed by the interstitial modeling. Many early works were focused on building 2D networks
flow rate (2) a Lagrangian framework is used for particle tracking and with the pores arranged on a regular square lattice (Dodd and Kiel,
(3) an accurate mapping of particles between pores is implemented. 1959; Fatt, 1956a). These regular grids continued in 3D (Chatzis and
The accuracy of using pore-networks for carbonates has not been es- Dullien, 1977). These networks modeled pore size distribution by ad-
tablished yet. For a detailed review of pore-network modeling of solute justing the pore sizes at each lattice node and could also change con-
transport in porous media we refer the reader to Mehmani and Balhoff, nectivity by adjusting the number of bonds from one lattice site to the
(2015b). next. In most cases, these networks were not directly representative of
any physical sample (except micromodels (Chen and Wilkinson, 1985)).
2.2.3. Quasi-static versus dynamic pore-network modeling of multiple In addition, a number of researchers only studied drainage capillary
phases pressure curves, using invasion-percolation algorithms (Chatzis and
In a pore-network model, pores hold the volume and contribute to Dullien, 1977; Dodd and Kiel, 1959; Fatt, 1956a). With only drainage,
the critical curvature for percolation during imbibition whereas throats no hysteresis was present in their models, but they could model wett-
represent conductivity and critical curvature during drainage. Drainage ability effects by assigning different contact angles in individual pores.
and imbibition at low capillary numbers are modeled with an invasion- Payatakes and co-workers (Payatakes et al., 1973, 1980), for-
percolation algorithm. In this method, a given capillary pressure is mulated a population balance model for the evolution of oil ganglia
imposed on the network domain at the inlet and the invading phase during water flooding. They studied granular media, where the sim-
occupies the pores based on the critical curvature of pores/throats and plistic pore networks they used were more representative of the real
certain trapping criteria (Bryant and Blunt, 1992; Mason and Mellor, sample. Shortly afterwards, in a series of papers by different researchers
1995). The hydraulic conductivity (also referred to as effective per- (Dias and Wilkinson, 1986; Wilkinson, 1986; Wilkinson and Willemsen,
meability) for each phase is then calculated by writing systems of 1983), advances in percolation theory led to a much better under-
equations based on imposing mass conservation at each pore. standing of displacement patterns formed during both drainage and
Table 1 shows a number of representative publications on pore imbibition, including viscous effects. These papers had far-reaching
network modeling over the past several decades. The rows of the table impacts on how to upscale pore-scale model relationships. Bryant et al.
are organized chronologically, starting with Fatt's seminal work (Fatt, (1993a), Bryant et al. (1993b) introduced sophisticated methods of
1956a, 1956b, 1956c) which first introduced 2D pore networks as a modeling compaction and cementation in granular media. They derived
way to study porous media flow properties. The columns have been unstructured grids from the resulting pore networks, and computed
organized as follows: relative permeabilities from them, with very satisfactory matches to
experimental data in strongly water-wet sandpacks and sandstones.
• Contributing author(s) and year of publication. This work was later extended by Bakke and Øren (1997) and Øren et al.
• A short summary of the work, including key innovations and/or (1998), who pioneered a process-based method of reconstructing the
findings. pore space in consolidated media like sandstones, incorporating in-
• Type of grids used, including nature of pore types: This also includes formation from high resolution thin sections of rocks. They also mod-
size of the grid used to represent REV. If applicable, this also in- eled effects of heterogeneous wettability. This work was later extended
cludes the resolution and size of the physical sample image from to three phases by other researchers (Piri and Blunt, 2004, 2005;
which the pore network was derived, or which the pore network was Valvatne and Blunt, 2004). Recent papers have focused on modeling
targeting. If no physical image was used, then the sides of the grid trapping and the structure of residual phases by improving the pore-
are dimensionless. If multiple similar samples with different grid filling models used for modeling imbibition processes - areas where
sizes were used, the largest grid size is reported. The size is very earlier pore networks have struggled (Ruspini et al., 2017; Ryazanov
interesting for future use in designing a pore-scale study, as it shows et al., 2010, 2014). Multi-scale pore networks are also a fairly recent
increasing computational power does not necessarily mean using development (Bultreys et al., 2015; Jiang et al., 2013b; Mehmani et al.,
larger sample sizes. 2011; Mehmani and Prodanović, 2014b; Prodanović et al., 2015b) but
• The type of physics incorporated: Since pore networks can use either are vital for modeling systems with significant microporosity like

11
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Table 1
Pore-network modeling literature: key representative journal publications over the years.
Author(s), year Key innovation/summary Grid type(s)/size Physics incorporated

Fatt, (1956) Introduced pore-network modeling. Examined, qualitatively, 2D Regular lattice of square and hexagonal shapes Quasi-static, 2 phase
effects of different properties of a 2D network on Pc-Sw, kr-
Sw curves
Dodd, Kiel (1959) Generated Pc-Sw curves for drainage, incorporating trapping and 2D, square, hexagonal networks, targeting Quasi-static, 2 phase,
mixed-wet conditions sandstone rocks including wettability
Chatzis, Dullien (1977) Examined both 2D and 3D networks for percolation properties, 2D and 3D regular lattices, targeting sandstone Quasi-static, 2 phase
and drainage Pc-Sw curves, against experimental data from rocks.
sandstones
Payatakes et al. (1980) Developed stochastic model for oil ganglion dynamics 2D regular lattice, interconnected unit cells, each Quasi-static, 2 phase
(mobilization, breakup and stranding) in granular media cell a constricted tube. Targeted granular media -
glass beads, sand
Lin, Slattery (1982) Predicted hysteresis in kr-Sw, Pc-Sw curves, though the method 3D regular lattice, targeting packings of glass Quasi-static, 2 phase
did not work well for consolidated media beads, sand
Mohanty (1982) In I, tracer flows for a wide range of fractional flows were done. 3D cubic network, 12 × 12 × 20, 12 × 12 × 40, Quasi-static, 2 phase
In II, a network model was developed to describe the physics. 16 × 16 × 20, targeting sandstones
Predicted consolidated media well for the first time.
Koplik, Lasseter (1985) Simulated waterflooding at different Ca, incorporating dynamic 2D network, 10 × 10 Dynamic, 2 phase
viscous effects, and capillary forces, through an approximate
numerical procedure in the pore network.
Chen, Wilkinson (1985) Studied viscous fingering during dynamic flow, ignoring surface 2D Square network, 100 × 100 Dynamic, 2 phase
tension term, in micromodel experiments and square lattice
networks. Confirmed DLA-like patterns
Lenormand et al. (1988) Modeled effects of Ca and viscosity ratio on flow patterns for 2D square lattice, 100 × 100 Dynamic, 2 phase
drainage in a 2D network. Percolation theory, DLA and anti-DLA
were used for explaining results
Jerauld, Salter (1990) Two-phase kr-Sw and Pc-Sw curves, including hysteresis. 3D network of spheres connected by tubes, both Quasi-static, 2 phase
Compared predictions with experiments, but not quantitatively cubic and Voronoi, size 203, targeting sandstones
(used assumed properties for the network) and bead packs
Blunt, King (1991) Demonstrate effects of both viscous and capillary forces on a 2D Delaunay triangulation of points in circular region - Quasi-static, 2 phase
and 3D networks 80,000 for 2D, Voronoi networks with 50,000
points in 3D
Blunt et al., 1992 Predicted kr-Sw curves with pore networks, examine results 3D cubic, cylindrical bonds/spherical bodies, Dynamic, 2 phase
against different types of percolation 32 × 32 × 64
Bryant, Blunt (1992) Used realistic geometries to derive pore networks. Predicted 3D unstructured grid, from a mono-size sphere Quasi-static, 2 phase
absolute permeability without adjustable parameters. pack (Finney packing of 3367 spheres)
Bryant et al., 1993, Bryant Physically representative network models of transport in porous 3D unstructured grid, from Delaunay tessellation of Single phase
et al. (1993a, 1993b) media central 2000 spheres from Finney pack
McDougall, Sorbie (1995) Study Pc-Sw and kr-Sw curves, recovery efficiency, for 3D cubic, 203 Quasi-static, 2 phase, with
waterflooding in systems of varying wettability. Also relate wettability
macroscopic wettability indices with pore-scale physics
Lowry, Miller (1995) Examine various geometrical factors affecting non-wetting phase 3D cubic, cylindrical bonds/spherical bodies, 8000 Quasi-static, 2 phase
residual after imbibition in a strongly water-wet system pore bodies
Dixit et al. (1996) Use pore-network models to explain higher oil recovery for 3D cubic, equal-length cylindrical pores Quasi-static, 2 phase, with
intermediate-wet systems, propose “regimes” for wettability wettability
classification
Bakke, Oren (1997) Proposed technique for generating 3D pore networks based on 3D network, with angular pores, derived from Quasi-static, 2 phase
geological processes for sandstone creation, and information Bentheimer sandstones, 1703, 10 μm resolution
from thin sections, applying it to predict macroscale flow
properties for water-wet systems
Fenwick, Blunt (1998) Modeled three-phase flow using pore networks, incorporating 3D cubic lattice, cubic pores, 30 × 15 × 15 Quasi-static, 3-phase
micromodel observations and theory of three-phase flow
Oren et al. (1998) Extended Bryant and Blunt (1992) to unequal size spheres. Used 3D unstructured, angular pores, derived from Quasi-static, 2 phase, with
reconstruction algorithm from thin sections to generate pore different sandstones, 3003, 10 μm resolution wettability
space, and then derived networks. Also modeled wettability
effects
Mogensen, Stenby (1998) Dynamic network model to study residual oil saturation, with 3D cubic lattice, cubic pores, 153 Dynamic, 2 phase,
effects of Ca, M, contact angle, aspect ratio including wettability
3
Dixit et al. (1998b) Hysteresis in kr-Sw curves, between primary drainage, secondary 3D cubic lattice, cylindrical bonds, 15 Quasi-static, 2 phase,
imbibition and secondary drainage for strongly water-wet and water-wet systems
weakly water-wet systems kr-Sw curves
Dixit et al. (1998a) Hysteresis in kr-Sw curves in mixed-wet systems 3D cubic lattice, cylindrical bonds, 153 Quasi-static, 2 phase,
including wettability
Dixit et al. (2000) Derive relationships between Amott-Harvey and USBM 3D cubic lattice, cylindrical bonds, 253 Quasi-static, 2 phase,
wettability indices including wettability
Lerdahl et al. (2000) Developed a stochastic reconstruction of Berea, and predicted 3D, unstructured lattice, angular pores, Quasi-static, 3 phase
Oak's [Oak, 1990] data. This pore network was then used in Reconstructed 3003, 10 μm resolution, targeting
several subsequent publications sandstones
Oren and Bakke (2002) Develop a method to reconstruct sandstones, using 3D, unstructured lattice, with angular pores, Single-phase
petrographical information from thin sections related to reconstructed 3003, resolution 7.5 μm
sedimentation, compaction and diagenesis
Oren and Bakke (2003) Extend earlier work to present a unified workflow for computing 3D, unstructured lattice, with angular pores, Quasi-static, 2 phase,
both single and multi-phase transport properties from reconstructed 1283, resolution 10 μm, targeting including wettability
reconstructed sandstones sandstones
(continued on next page)

12
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Table 1 (continued)

Author(s), year Key innovation/summary Grid type(s)/size Physics incorporated

Singh, Mohanty (2003) Developed a rule-based dynamic pore-network model, studied 3D cubic lattice, square pores/throats, 30 × 8 × 8 Dynamic, 2 phase
effect of viscous and capillary forces. Generated qualitative
capillary desaturation curves.
Valvatne andBlunt (2004) Predicted kr-Sw and Pc-Sw curves on mixed-wet sand packs and 3D unstructured lattice, angular pores, from Bakke Quasi-static, 2 phase,
Berea and Oren (1997) including wettability
Piri and Blunt (2004) Developed thermodynamic relationships for three-phase Individual angular pores Quasi-static, 3 phase,
displacements in a single, constant cross-section angular pore including wettability
Piri and Blunt (2005) Expanded on earlier work in single pore to develop three-phase 3D unstructured lattice, angular pores, 3003, Quasi-static, 3 phase,
flow in pore networks, using thermodynamics to determine resolution 10 μm, from Berea sandstone including wettability
stability
Løvoll et al. (2005) Drainage of more viscous wetting fluid by less viscous non- 2D, square network, 80 × 160 nodes Dynamic, 2 phase
wetting fluid- experiment, numerics and percolation theory
3
Nguyen et al. (2006) Developed a dynamic pore-network model for imbibition 3D unstructured, 300 , from Lerdahl et al. (2000), Dynamic, 2 phase
incorporating film flow, film swelling and snap-off. Quantified Valvatne andBlunt (2004)
effects of flow rate in inhibiting snap-off
Spiteri et al. (2008) Use pore-network simulations to study new trapping and relative 3D unstructured from [Valvatne andBlunt 2004, Quasi-static, 2 phase
permeability hysteresis models Lerdahl et al., 2000]
Joekar-Niasar et al. (2007) Study two types of pore-network models: only tubes, and spheres 3D cubic lattice, 403 Quasi-static, 2 phase
and tubes - with and without trapping, respectively. Also relate
interfacial area to capillary pressure, rel perm
Raeesi, Piri (2009) Used mixed-wet pore-network models to study Pc-Sw-area 3D unstructured grid from [Piri and Blunt 2005], Quasi-static, 2 phase,
relationships Berea (3003, resolution 10 μm) and Saudi reservoir including wettability
sandstone (2503, resolution 10 μm)
Idowu and Blunt (2010) Study dynamic effects during waterflooding, incorporating 3D unstructured grid, 3003 from [Valvatne Dynamic, 2 phase, with
effects of capillary number, viscosity ratio and contact angles andBlunt 2004] wettability
Gharbi and Blunt (2012) Study effects of wettability in several different carbonate 3D unstructured, from carbonate CT images of size Quasi-static, 2 phase, with
samples 3203-3503, with angular pores, resolution 7.7–9 μm wettability
Mehmani et al. (2013) Presented a new multi-scale model for flows in shales, 3D unstructured, multiscale Single-phase, with slip-
combining nano-meter size pores with micro-meter size pores, flow and Knudsen
and studied effect of pore connectivity and fraction of diffusion
microporosity on permeability
Ryazanov et al. (2014) Figure out structure of residual oil after water flooding as a 3D unstructured lattice, Berea 3003, 10 μm Quasi-static, 2 phase, with
function of wettability resolution, from [Oren and Bakke, 2003] wettability
Bultreys et al. (2015) Developed a dual pore-network model to model microporosity 3D unstructured, multiscale, spherical pores, with Quasi-static, 2 phase
cylindrical links, Estallaides limestone, 10003,
resolution 3.1 μm
Ruspini et al. (2017) Improved pore-network models so they predict trapping curves 3D unstructured, from [Øren et al., 1998], 4503- Quasi-static, 2 phase, with
accurately, based on a new pore-body filling approach 12003, resolution 1.5–10 μm wettability

Fig. 17. Primary drainage relative permeability


curves for (a) a vuggy carbonate rock (with a scale
difference of 5) and (b) a carbonate rock with a
single microcrack (with a scale difference of 5 (at
three axis angles with respect to the direction of the
pressure gradient). Drainage is simulated on pore-
network models described in Fig. 8, and the corre-
sponding capillary pressure curve is shown in Fig. 9.

carbonates. As examples of these multiscale networks, Fig. 17 shows the (Andrew et al., 2015; Berg et al., 2013; Singh et al., 2017) have ob-
drainage relative permeability curves for carbonates with vugs and served rapid movements and phase rearrangements at low capillary
microcracks as was described in Section 2.1.2. Multiscale networks are numbers, which challenge the traditional assumption of quasi-static
a significant advancement over earlier works which focused on more flow. In particular, Pak et al. (2015) and Reynolds et al. (2017) have
homogeneous porous media like sandstones and sandpacks. Carbonates shown some completely new pore-scale displacement mechanisms in
have a bimodal pore size distribution, as well as a much more com- carbonates that cannot be modeled with traditional pore-network ap-
plicated wettability state, and are hence much harder to model. proaches.
The shortcoming of quasi-static modeling is its assumption that the Two computational schemes that are used in dynamic pore-network
movements of phases do not affect each other. This is not the case in modeling are the single-pressure (Aker et al., 1998; van der Marck
high capillary numbers when the phase distributions become functions et al., 1997) and two-pressure (Thompson, 2002) schemes. In the
of viscosity ratios and flow rates as well as the interfacial tension. We single-pressure scheme, regardless of the simultaneous occupancy of a
note that recent experiments using fast synchrotron x-ray imaging pore by both phases, a single pressure is assigned to each pore as an

13
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

unknown. The assumption therefore is that the local capillary pressure operator is given by the Bhatnagar Gross Krook (BGK) relationship,
within a pore is negligible. In addition, corner flow cannot be im- which performs very well for a wide variety of situations. Without
plemented using the single-pressure scheme and therefore it is only going into mathematical details, the BGK operator is based on a single
applicable for lowly angular and smooth granular media. In the two- relaxation time parameter, which results in viscosity-dependent per-
phase pressure scheme, the local capillary pressure within a pore is not meability. A more complex operator is the Multiple Relaxation Time
ignored and therefore each phase's pressure becomes an unknown in the (MRT), which is more computationally expensive, but performs better
systems of equations. The problem that arises using this scheme in low at computing permeabilities in porous media images (Pan et al., 2006).
capillary numbers is a numerical instability called “capillary pinning” For multiphase flow, many different lattice Boltzmann models exist -
(Koplik and Lasseter, 1985) at which the saturation of a pore is non- the color gradient model (Gunstensen et al., 1991), the Shan-Chen
convergent. Joekar-Niasar et al. (2010) proposed a semi-implicit sa- model (Shan and Chen, 1993), the free energy model (Swift et al., 1996)
turation update at each time step by which the numerical instabilities and others. It is not clear if any one model is always superior to the
improved. For a further discussion on dynamic pore-network modeling other (Liu et al., 2016) and it is up to the user to decide which one to
the reader is referred to Joekar Niasar (2010) and the references use for their application. However, as reported by Liu et al. (2016) some
therein. From discretization standpoint, these pore-network models models, such as the color gradient model and the stabilized diffuse-
resemble reservoir simulators (for instance, they update saturation), but interface model, are better at handling high viscosity ratios at low ca-
operate on a smaller length scales and are thus incorporating capillary pillary numbers which is relevant in porous media flows, though this
forces directly (as opposed to through averaged functional relation- comes at a higher computational cost and more complex equations.
ships). Similarly, the larger the porous domain is the longer the computational
Unlike voxel-based discretization in direct numerical simulation time becomes. The issue of domain size is pertinent in carbonates as
methods (Section 2.2.4), discretization done via pore-network models is simulating the injection of one pore-volume when large volumes of
not scalable. In other words, voxel-based discretization can offer an vugs are present can become computationally difficult. It is currently
optimum space interval at which the solution will be satisfyingly ac- not clear when these costs would be justified for a given case.
curate and yet computationally feasible. In pore-network modeling, Hazlett (1995) was one of the earliest works in direct modeling of
besides choosing the domain size that does not fall below the re- flow in porous media images. He simulated multiphase flow using a
presentative elementary volume (REV), the concept of an optimum sphere-algorithm, and calculated relative permeabilities using the
discretization does not exist. Hybrid methods when coupling from dif- multiphase color-gradient lattice Boltzmann method. Subsequently,
ferent length scales are an exception and this is discussed in Section 2.3. Martys and Chen (1996) performed multiphase flow simulations using
Open source pore-network modeling code is available - please see the Shan-Chen lattice Boltzmann method. These two early pioneering
(Gostick et al., 2016). papers had limited predictive power, due to the limited resolution and
computational resources at the time. Kats and Egberts, (1999) per-
2.2.4. Principles of direct numerical simulation formed the first three-phase porous media simulations, but only in 2D.
Modeling the multiphase fluid behavior directly in the pore space As computational power improved, other researchers focused efforts on
domain is an attempt to achieve accuracy that pore-network models studying hysteresis in capillary pressure curves, relating it to interfacial
may sacrifice for computational speed (Raeini et al., 2014a). In addi- area (McClure et al., 2004; Pan et al., 2004; Porter et al., 2009; Schaap
tion, direct numerical methods provide sub-pore scale fluid dynamic et al., 2007). Liu et al. (2016) performed a detailed benchmark study on
insights that can be utilized for improving pore-network models, which what viscosity ratios can be simulated with each model in static and
could potentially become a very useful tool in upscaling simulation dynamic cases. Open source parallelized codes for lattice Boltzmann
results – particularly for multi-scale systems like carbonates. Direct method are available, e.g. Palabos3 and Taxila-LBM (Coon et al., 2014).
numerical methods are implemented once the pore space has been ei- The other popular class of direct simulation methods are based on
ther segmented from an image (Iassonov et al., 2009) or digitally re- discretization of Navier-Stokes equations. These techniques have ex-
constructed through a process-based or stochastic-based method. isted longer than the lattice Boltzmann method, and have very well-
The most popular methods for direct modeling are lattice Boltzmann developed mathematical foundations. Difficulties with these methods
methods and finite volume methods based on discretization of Navier- arise when more than one phase becomes involved. This is due to
Stokes equations (and addition of interfacial tension when modeling challenges in tracking the two-phase fluid interface and its contact line
immiscible flow). Lattice Boltzmann methods evolved from lattice gas with the solid phase. The volume-of-fluid (VOF) method (Hirt and
automata (LGA) techniques. They are pseudo-molecular methods, Nichols, 1981) is a continuum-based method that solves the Navier-
which track the distribution functions of an ensemble of fluid particles. Stokes equation for each phase using saturation-averaged values for
They work at a scale in between molecular dynamics, and the con- density and viscosity. The interfacial tension is treated as a body force
tinuum scale. Macroscopic variables such as velocity, pressure and in this method (Brackbill et al., 1992). The drawback of VOF is the
density, which are part of traditional computational fluid dynamics emergence of spurious currents when capillary forces become high
(CFD) methods, are recovered by moment integration of the particle (Dupont and Legendre, 2010). The problem of spurious currents at low
distribution functions. The particle distributions at each lattice point capillary numbers is particularly relevant for multiphase flows in
evolve by interacting only with its immediate neighbors. Since each porous media. Algorithms on correctly tracking the immiscible inter-
lattice point interacts only with its neighbors, the problem is highly face has been the focus of eliminating spurious current (Popinet and
parallelizable, and well suited for modern multi-thread/core pro- Zaleski, 1999; Renardy and Renardy, 2002; Abu-Al-Saud et al., 2018),
cessors. In addition, boundary conditions can be easily handled using for instance Raeini et al. (2014a) and Shams et al. (2018) showed that
the bounce-back condition. This enables representing complex geome- the non-physical instabilities can be improved by using a sharp surface
tries easily using an underlying orthogonal grid, without using complex force model. For an open-source platform that incorporates single-
meshing algorithms. It has hence become a very popular choice in the phase and multiphase immiscible flow using finite-volume methods
porous media community which deals with very complex 3D pore refer to www.openfoam.org.
spaces, which can have features on many different scales. Level set methods are often proposed as an alternative to VOF for
LB methods follow a two-step procedure. First, there is an advection tracking moving interfaces (Sussman, 2005; Sussman et al., 1994,
step, where the values of the distribution function are propagated to 1999). These methods have their own problems, such as mass loss
their neighboring lattice points along defined velocity vectors. Second, (Enright et al., 2002) and high computational costs. Several authors
there is a collision step, where the particles at each lattice point are have proposed methods to address the mass loss issues in level set
redistributed according to a collision operator. The simplest collision methods, such as the particle-level set method (Enright et al., 2002),

14
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

and non-graded adaptive grids (Min and Gibou, 2007). Other re- as well as highly non-uniform pore size distributions result in even
searchers have combined the level set and VOF techniques to give hy- more challenging systems which have not yet been validated compre-
brid methods which can take advantage of the features of both hensively. Frank et al. (2018a), Frank et al. (2018) have recently ap-
methods, for example the Coupled Level Set VOF method (CLSVOF) plied a finite element formulation with the discontinuous Galerkin
(Sussman and Puckett, 2000). In a benchmark study of these methods, method which shows promise in resolving some numerical issues which
Gerlach et al. (2006) concluded that the CLSVOF method was the best surround simulation of moving interfaces.
one of these techniques. Thömmes et al. (2009) attempted to couple
level set methods with lattice Boltzmann to improve characteristics of 2.3. Combining multi-scale pore-geometries in pore-scale models
the flow behavior for large viscosity ratios. However, they only applied
the method to very simple cases, and the method has not been widely Given the continuous and wide distribution of pore sizes over
used since. multiple length scales in carbonate rocks, a proper integration of mul-
An alternative method for predicting the multiphase fluid distribu- tiscale pores is essential for predicting their transport properties. Biswal
tion at a given capillary pressure is the quasi-static level-set method et al. (2007) proposed a stochastic continuum reconstruction scheme
(Prodanović and Bryant, 2006). In this method, the immiscible interface for microporous carbonates, based on implementing a spatial correla-
is defined by a level set function that progresses with a velocity that is tion for various grains that is established from training images. The
directly linked to the balance between pressure and surface tension. model was used to predict absolute permeability, formation factor and
Open source code is available (Prodanovic, 2010). Although the level- elastic moduli for carbonate samples. In their method, each voxel is
set method has a high accuracy in describing the fluid morphology, it segmented into pores, matrices or undecided. Undecided voxels are
cannot be used for modeling viscosity-dominated regimes. Later on, alternately assigned as pores or matrices. The method was implemented
Jettestuen et al. (2013) and Verma et al. (2018) added the ability to for a microporous oolithic dolostone. The predicted transport properties
incorporate arbitrary contact angles to this simulation method. Fig. 18 however were highly dependent on the chosen resolution which high-
shows an example of each of the discussed direct numerical methods lights the potential advantage of reduced-order computational methods
(Harting et al., 2005; Prodanovic and Bryant, 2009; Raeini et al., such as pore-network modeling if they are supplied with correct a priori
2014a). transport rules. Bauer et al. (2012) presented a dual pore-network
Table 2 summarizes publications over the past two decades for di- model for carbonates where a three-dimensional micro-CT image is
rect modeling. Like the pore-network modeling table, it is arranged segmented into solid, macropores and microporous regions. The mi-
chronologically. It includes papers based on lattice Boltzmann and croporous regions however are not resolved but represented as con-
Navier-Stokes discretization. It also incorporates other less popular tinuous regions. The explicit integration of pores from two scales into a
techniques like the level set method and smoothed particle hydro- single-entity pore-network model has been developed via an image-
dynamics (SPH). The columns are organized similarly to Table 1. based method in which a micropore network is stochastically created
However, the grids are always related to some physical sample, unlike and indiscriminately (Jiang et al., 2013b), or with a spatial correlation
pore-network models for which the concept of grid does not apply. (Jiang et al., 2013a), superimposed to a macropore network. In order to
Over the years, as more computational power and better imaging capture micropore generating diagenesis, Mehmani and Prodanović,
techniques have become available, the resolution used has grown (2014b) have proposed a process-based approach to integrating mi-
better, increasing the REV size as well as match with the experiments. cropores with macropores which was discussed previously in Section
For instance, Ramstad et al. (2012) generated relative permeability 2.1.2. The method can additionally be implemented to three-dimen-
curves for a reservoir sandstone but matching with experimentally sional images where grains are segmented and subsequently filled with
measured values (on a larger core) improved predictions dramatically micropores (referred to as “grain-based” approach). For images where,
by using end-point saturations from experimental data. This is the same due to considerable diagenesis, grain segmentation is not feasible, a
procedure used by pore-network modelers, though one has more “tile-based” approach has been developed. In this method, the domain
parameters to adjust with in pore networks. Other methods like the is tiled and micropores are implanted on the appropriate areas
quasi-static level set method (Prodanović and Bryant, 2006) and (Prodanović et al., 2015b) (see Fig. 19 for schematics of the methods).
smoothed particle hydrodynamics (Tartakovsky et al., 2007; As one may intuit, the details captured by explicitly incorporating mi-
Tartakovsky and Meakin, 2006) have been limited by computational cropores into the pore-network model results in high numbers of un-
resources, but the VOF technique has been applied to study relative knowns and can significantly reduce the computational feasibility of
permeability (Raeini et al., 2014b) and trapping (Raeini et al., 2015). solving the systems of equations. In an effort to amend the computa-
Trapping is in general typically modeled better using VOF or a direct tional limitations of coupling micropores with macropores, Bultreys
method, though a carefully matched pore network model recently et al. (2015) have represented the upscaled properties of microporous
(Ruspini et al., 2017) showed promising results using a new pore-body regions with microthroats. Determining the properties of the micro-
filling model in the same pore network. For an additional general dis- porous clusters and microthroats a priori is challenging given the ne-
cussion on pore-scale modeling refer to Golparvar et al. (2018). In the cessity for high resolution images for pore structure characterization.
context of carbonates, their two main challenges are their multiscale Tuning the micropore properties by matching predicted transport
nature (discussed in Section 2.3) and their generally heterogeneous-wet properties with macroscopic experimental measurement is not always
condition. Current research efforts are focused on incorporating such possible either given the scarcity of experimental data points in highly
heterogeneous-wet conditions to an accurate model. Zhao et al. (2016) microporous and tight carbonates. Verifying the fidelity of multiscale
have shown the importance of film flow in determining macro-scale pore structure reconstruction is nontrivial and will likely require ad-
behavior under uniform wettability conditions. In a follow-up work, vanced real-time image acquisitions tools such as fast-synchrotron to-
Zhao et al. (2019) have compared 14 different pore-scale modeling mography to clarify the transport communication between the micro-
methodologies for capturing the behavior seen in the micromodels. We pores and macropores.
note here that the micromodel experiments of Zhao et al. (2016) are not Direct numerical simulations have recently been implemented in
very representative of usual flow conditions in porous media - in par- imaged multiscale pore systems to predict their single-phase and two-
ticular the capillary number is extremely high, which means entirely phase properties. Soulaine et al. (2016) employed the Darcy-Brinkman
different macro-scale behavior from typical subsurface flow conditions. equation in a computationally parallel setting to demonstrate the high
However, this study still shows that current simulation methods are not impact of even small percentage of sub-resolution pores on absolute
quite up to the task of simulating all the complexities in a porous permeability in a sandstone sample. The Darcy-Brinkman formulation
medium. Moreover, in carbonates, film flow, heterogeneous wettability, renders the solution of the velocity field at the interface between sub-

15
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. 18. Examples of (a) utilizing volume-of-fluid method to simulate drainage (throat length is 37.5 µ m) (Raeini et al., 2014a) (b) level-set method for a naturally
fractured carbonate rock, (b-1) is a tomogram of size 700 × 1 × 700, (b-2) is a media surface of size 200 × 230 x 190 and (b-3) is the non-wetting fluid distribution
at a particular curvature during drainage (red is the fluid/fluid and gray is the fluid/solid interface) all with a voxel size of 3.1 µ m (Prodanovic and Bryant, 2009) and
(c) drainage simulation for a Bentheimer sample (512 × 512 x 521 with a voxel size of 4.9 µ m) using lattice Boltzmann method (Harting et al., 2005). (For
interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

resolution and resolved pores tractable and provides a direct adjust- method for carbonates that have pores at inseparable scales is the
ment of the equations based on the microporous region's permeability. mortar domain decomposition method (Y. Mehmani, 2014). In this
A similar formulation for two-phase immiscible flow with nanoscale method, a pore-scale subdomain can be coupled with a continuum
effects corrections including slip flow and Knudsen diffusion was ex- subdomain by defining an interface function space (the mortar space).
tended by Soulaine et al. (2019) to predict the transport properties of For each subdomain an appropriate transport algorithm can be solved
shale formations. Such nanoscale corrections can become relevant for iteratively until the flux at the interface coincides with the two adjacent
carbonates as well (Yoon and Dewers, 2013). However, the practicality subdomains. Much work however on adopting mortar domain decom-
of the Darcy-Brinkman formulation imposes an abrupt change from the position for predicting immiscible flow in carbonates is needed. Lattice
Darcy velocity in the sub-resolution porous systems to the interstitial Boltzmann methods have also been extended to predict fluid transport
velocity in the resolved pores. Beavers and Joseph (1967) for instance in multiscale settings. This approach is generally referred to as Gray
demonstrated a slip velocity region in the boundary between the pore Lattice Boltzmann (GLB) as the sub-resolved porous areas would have
space and the sub-resolution regions. Additional research is required to geometric properties in-between resolvable pores and the solid phase.
determine the importance of this additional interface condition to flow The “gray” areas are considered to have a grain density in proportion to
properties, including in case of immiscible flow, and their sensitivity to their porosities and X-ray attenuation intensities. In earlier models (Gao
any associated parameters in carbonates. Another promising hybrid and Sharma, 1994), the impacts of sub-resolution pores on fluid

16
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Table 2
Direct pore-scale modeling literature: key representative journal publications over the years.
Author(s), year Key innovation/summary Sample type(s)/size Physics incorporated

3
Hazlett (1995) First work to directly try simulations on actual rock images. Berea sandstone, 256 , resolution 10 μm Inscribed spheres for displacement,
Predicts fluid distributions based on inscribed spheres in including wettability, lattice
image. Lattice Boltzmann was used to get permeabilities from Boltzmann for permeability
equilibrium distributions
Martys, Chen (1996) First work to model two-phase flow directly using lattice- Fontainebleau sandstone, 7.5 μm, resolution, 2-phase lattice Boltzmann, Shan-Chen
Boltzmann 643 samples model
Hazlett et al. (1998) CMT image of sandstone used for drainage in strongly water- 1283 reservoir sandstone, 10 μm resolution 2 phase lattice Boltzmann color
wet and mixed-wet cases gradient model, including wettability
Kats and Egberts, First lattice Boltzmann work to model 3 phase flow in porous 210 × 350 2D lattice with square elements 3 phase lattice Boltzmann color
(1999) media gradient model
Manwart et al. (2002) Computed permeability for Fontainebleau samples, generating 3003, 4 Fontainebleau sandstone samples, Single phase lattice Boltzmann BGK
matches with experiment. Shows first order nature of bounce- resolution 7.4 μm, one real, and three model, finite difference solver for
back BCs generated Stokes equation
Pan et al. (2004) LB simulations on ideal sphere packing, to give Pc-Sw curves, 1283, 150 dimensionless spheres 2 phase lattice Boltzmann Shan-Chen
including hysteresis, and subsequent comparison with model
experiments
Kang et al. (2004) Simulations of fingering during immiscible fluid 2D channel, 400 × 66 2 phase lattice Boltzmann Shan-Chen
displacements in a channel, studying effects of capillary model
number, viscosity ratio and wettability
Huang et al. (2005) Studied multiphase flow in unsaturated fractures, including 2D fractures, 60 × 400, resolution 0.05 mm Volume of fluid (VOF) method Navier-
effects of viscosity, contact angle, and gravity Stokes solver
3
Harting et al. (2005) Performed lattice Boltzmann simulations in porous media Benthemer sandstone, 512 , 4.9 μm resolution 2 phase lattice Boltzmann Shan-Chen
samples as part of a larger review on simulations for other model
cases
Kang et al. (2006) Simulation of reactive flow at the pore scale, considering 2D channel, 80 × 80 Lattice Boltzmann equations coupled
effects of advection, diffusion, mineralogy, and reaction with chemical reactions
between species
Pan et al. (2006) Investigation of BGK and MRT lattice Boltzmann models to 323 BCC sphere array; and 643 random sphere Singe phase BGK and MRT lattice
compute permeability, as well as effect of using different packing, with 23 spheres Boltzmann models
boundary conditions.
Prodanović et al. Introduced a quasi-static level set method, studying Mono-sized sphere packs, 383 2 phase quasi-static level set, with zero
(2006) multiphase displacements in idealized sphere packs, with zero contact angle
contact angle
Tartakovsky, Meakin Simulated immiscible and miscible flows in idealized 2D 2D 32 × 16 geometries Smoothed particle hydrodynamics, 2
(2006) geometries phase, both immiscible and miscible
Schaap et al. (2007) Study of water-air and water-soltrol displacement experiments Glass bead packs, 405 × 405 × 100, 2-phase lattice Boltzmann Shan-Chen
in glass bead packings, relating interfacial areas to capillary resolution 17 μm model
pressure characteristics
Hatiboglu, Babadagli Studied spontaneous imbibition using micromodel 2D micromodel, 4002, 0.125 mm resolution 2 phase lattice Boltzmann Shan-Chen
(2008) experiments and LB model
Sukop et al. (2008) Compared LB vs experimental fluid distributions, slice by slice 2563, 20 μm resolution, quartz grain packing. 2 phase lattice Boltzmann Shan-Chen
(1283 really) model
Thömmes et al. Combined LB with level set methods. Simulated flows with 2D Couette flow, bubble, Hele-Shaw cell, 1282 2 phase, lattice Boltzmann BGK model
(2009) large viscosity ratios and density differences, for simple 2D coupled with level sets
and 3D cases.
Porter et al. (2009) Matched interfacial area measurement by CMT in glass beads Glass bead pack, 205 × 205 × 150, resolution 2 phase lattice Boltzmann Shan-Chen
with LB simulations, and related it to hysteresis in Pc- 34 μm model
Sw curves
Boek, Venturoli Performed simulations for a 2D sample of Berea and 3D 1283 sample of Bentheimer sandstone, 2 phase lattice Boltzmann Shan-Chen
(2010) Bentheimer sample resolution 4.9 μm model
Ramstad et al. (2010) Simulations in CMT and reconstructed sandstone images. Bentheimer sandstone, 2563, resolution 2 phase, modified lattice Boltzmann
Reproduce Pc-Sw drainage curves well, but underpredict rel 6.67 μm (CMT), 5 μm (reconstructed) color-gradient model Latva-Kokko,
perms Rothman [2005]
Ramstad et al. (2012) Relative permeability from Bentheimer and Berea sandstone Bentheimer, 2563, resolution 6.67 μm, Berea 2 phase, modified lattice Boltzmann
using LB 2563, resolution 5.345 μm color-gradient model Latva-Kokko and
Rothman (2005)
Jettestuen et al. Proposed extension of model by Prodanovic and 3D sphere packs, 66 × 66 × 62, Castlegate 2 phase quasi-static level set, with non-
(2013) Bryant Prodanović, Bryant [2006] to incorporate wettability sandstone, 60 × 60 × 70, resolution 5.6 μm zero contact angle
effects
Li et al. (2013) Extended the Rothman-Keller multiphase LB model with MRT, Berea, resolution 2.89 μm, 3003 for 2-phase 2 and 3 phase, Rothman-Keller MRT LB
and simulated drainage-imbibition experiments Berea simulation, 643 for 3-phase model
samples. Also tried preliminary 3 phase simulations
Ferrari, Lunati (2013) Simulated immiscible displacements in 2D, and studied the 2D 1760 × 2280, resolution 50 μm, 10028 Navier-Stokes VOF solver
relation between macroscopic capillary pressure and pore- particles
scale properties
Raeini et al. [2014] Used their own modified VOF formulation Raeini et al. Berea sandstone (400 × 256 × 256), Navier-Stokes VOF solver
[2012] to relate pore scale forces with Darcy-scale properties resolution 4.58 μm; Sandpack
including rel perms, proposing an upscaling algorithm (282 × 174 × 174), resolution 8.2 μm
Liu et al. (2014) Used a new color-fluid model to simulate flows in 2D 2D micromodel, 2816 × 2689, resolution 2 phase lattice Boltzmann color-fluid
geometries of heterogeneous permeabilities 1.5 μm model
Landry et al. (2014) Used LB to get rel perms for strongly water-wet (glass beads) Glass bead pack, 1003, resolution 26 μm 2 phase lattice Boltzmann Shan-Chen
and weakly oil-wet (polyethylene beads) cases, comparing model
results with bead pack CMT experiments
(continued on next page)

17
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Table 2 (continued)

Author(s), year Key innovation/summary Sample type(s)/size Physics incorporated

Raeini et al. (2015) Studied trapping trends in water-wet consolidated and Berea sandstone (330 × 210 × 210), Navier-Stokes VOF solver
unconsolidated media, matching against experimental data resolution 5.549 μm; Sandpack
(230 × 142 × 142), resolution 10 μm
Liu et al. (2016) Extensive review of popular LB multiphase models, and points 2D samples, cylinder grids (512 × 1280), 2 phase Shan-Chen, color gradient, free
out limitations of each, in the context of porous media flows reproduced from other works energy, mean-field theory, stabilized
diffuse-interface
McClure et al. (2016) Revisit hysteresis in Pc-Sw curves, conclude that inclusion of Randomly packed 1964 spheres, 9003 2-phase lattice Boltzmann modified
interfacial area and Euler characteristic makes relationship color-gradient model [McClure et al.,
non-hysteretic 2014]

propagation are incorporated in the collision step as described in Sec- (Carman, 1937; Kozeny, 1927) and the Brooks–Corey parameterization
tion 2.2.4. Subsequent works have focused on increasing the accuracy of relative permeability (Brooks and Corey, 1964), were developed for
and computational performance of GLB by modifying the collision step rocks whose pores are mostly intergranular, single length scale and well
(Walsh et al., 2009). Recently, multiscale algorithms for accelerating connected. Since carbonates exhibit pore space heterogeneity such as
fluid dynamics simulations on digitized porous media have been pro- high variations in pore size, shape, topology and wettability, traditional
posed (Mehmani and Tchelepi 2018, 2019; Guo et al., 2019). The Pore- transport relationships are rarely satisfactory. Reliable pore-scale
Level Multiscale Method (PLMM) has been shown to produce highly models have the potential to provide the petrophysical properties of all
accurate approximations to single-phase incompressible, immiscible rocks at the REV scale (Fig. 1), and for carbonates such workflows are
two-phase, and compressible gas flow in single (e.g., sandstone) and especially valuable given their complex flow behavior and scarcity of
dual porosity (e.g., shale) media. No simplification of the underlying data during formation evaluation.
geometry or governing equations are made. Broadly speaking, PLMM An accurate pore-scale model depends on the correct capturing of
consists of four steps: (1) decomposition of the image into physical the pore space. While pore scale modeling (specifically, pore-network
pores, (2) computation of local basis functions on each pore, (3) solu- modeling) has been in existence since the first use of bundle-of-tubes
tion of a global interface problem, and (4) reconstruction of the final models by Purcell (1949) and the first interconnected pore-scale models
solution. Step (1) utilizes watershed segmentation, similar to PNM, and by Fatt (1956a, 1956b, 1956c), in this paper we have focused on re-
the subdomains created correspond to physical pores in the image. viewing methods that have a degree of physical representativeness in
However, the decomposition cost is much lower than typical network terms of pore space geometry and interconnectedness. We reviewed
extraction, which involves multiple additional steps. Step (2) can be three philosophies for reconstructing the pore-space. The process-based
thought of as a parametrization of upscaled flow parameters for each method hinges on an appropriate description of the original granular
pore, which is loosely analogous to computing hydraulic conductivities packing followed by modeling diagenesis of the sample. The algorithm's
for throats in certain PNM variants (Prodanović et al., 2015b). Step (3) basis on sphere packing with a desired grain size distribution and tes-
involves solving a coarse system, which is comparable in size to typical sellating of the sphere centers results in computational expedience.
(non-)linear systems encountered in PNM. Step (4) is optional and is Multiple realizations of a diagenetically altered granular medium can
only executed if a sub-pore scale solution is desired. If only macroscopic therefore be generated to cover a range of pore-scale parameters. Ad-
parameters are of interest (e.g., permeability), step (4) may be omitted. vances in imaging technology and a continuous increase in computa-
The errors in PLMM can be further estimated a priori and controlled tional power have catalyzed the explosion in pore-scale modeling since
(not possible in PNM) through a rapidly convergent iterative strategy in the late 1990s. Technologies such as computed tomography, X-Ray
(Mehmani and Tchelpi, 2018). The DNS solver used in step (2) can be microtomography and (focused ion beam) scanning electron micro-
chosen from a number of available approaches (FVM, FEM, SPH, LBM). scopy, provide a window into geologic materials that can be analyzed
All computations in PLMM are massively scalable in parallel environ- with a digital rock physics framework for characterization and creation
ments due to independence of calculations. of pore-scale models. The image-based and stochastic-based methods
depend on acquiring a quality image using resolutions that can capture
relevant pore space features. Once the pore space is rendered, either a
3. Discussion
pore-network method or a direct numerical simulation technique is
chosen to achieve the desired macroscopic property.
Conventional functional relationships, such as Archie's law for re-
The choice of the pore-scale model depends on several criteria
sistivity (Archie, 1942), the Carman–Kozeny permeability estimate

Fig. 19. Schematic of algorithms for (a) “grain-


based” and (b) “tile-based” image-based two-scale
pore-network creation (Prodanović et al., 2015b).
Impermeable grains and microporous regions are
colored black and gray respectively. Throats linking
pores from the same scale are colored red whereas
throats connecting pores from two scales are blue.
(For interpretation of the references to color in this
figure legend, the reader is referred to the Web ver-
sion of this article.)

18
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

including microstructure image availability, importance of sensitivity helpful source of information for parameters such as minerology and
analysis, type of the transport process one wishes to study and com- pore size. In addition, they can be expanded into 3D with stochastic-
plexity/availability of the software. If one wants to analyze a specific based methods. The spatial distance of the extrapolation however can
sample and it can be appropriately imaged so that pores are well-re- be limited due to anisotropy. Even though image-based direct simula-
solved, then direct numerical simulation is the clear choice. It could be tion methods are limited by domain size and resolution, they can be
that pores are of complex shape so that their representation with a pore- useful in providing input parameters for pore-network models that can
throat network is not well-defined. Further, if single phase permeability subsequently be expanded to larger domains. Hybrid DNS and PNM
is desired, off-the-shelf parallelized direct simulators exist that make its workflows however require further research to establish their benefits.
calculation relatively straightforward. Nonetheless, direct-numerical Wettability is an important parameter in determining the capillary
methods can become unfeasible in “large” porous domain. A domain pressure and relative permeability behaviors of porous media that we
size for interparticle pores that is manageable with direct numerical did not cover in detail in this paper but find that it is an essential re-
methods can be considered “small” when vugs with sizes orders of search area for obtaining reliable pore-scale models for carbonates. In
magnitude larger are present. In this case even though all pores are pore-scale modeling, it is typical to assume a uniform and unchanging
technically resolvable in the three-dimensional image, a pore-network value throughout the pore space. Recent visualizations of in situ contact
model, despite uncertainties with defining the pores and throats during angle subsequent to core-flooding (AlRatrout et al., 2017; Andrew et al.,
its extraction, is preferable. If an appropriate image is unavailable or 2014) have shown that carbonate rocks demonstrate a wide contact
sensitivity analyses for geometry in addition to transport parameters is angle distribution. This is partially due to the varying mineral makeup
needed, then process-based models are a preferred option. In creating a of the rock. In addition, given the commonly considerable chemical
process-based model for carbonate samples, one faces the challenge of diagenesis of carbonates, surface textural heterogeneity such as surface
quantitatively describing the paragenesis given the uncertainty of par- roughness, can create a non-uniform mixed-wet system. An important
ticular diagenetic processes such as dissolution. Further, if geometry property of carbonates that is difficult to capture in models, is its po-
details such as surface roughness are important to the transport process, tential to change its wettability during two-phase flow. This dynamic
then direct numerical simulation is suggested as well. On the other wettability behavior was proposed (Kovscek et al., 1993) by linking the
hand, due to geometry simplification and use of analytical formulas, breakup of monolayer wetting films during primary drainage. The au-
pore-network models offer “infinite resolution”, i.e. there is no nu- thors showed that once the imposed capillary pressure exceeds the
merical discretization involved in pore and throat cross-sectional geo- summation of pore entry capillary pressure threshold and the wetting
metries. Therefore, if film flow for instance is of interest, pore-network surface's disjoining pressure, oil is able to contact the grain surface
models are a good choice. One must bear in mind however, that the previously protected by wetting films. Recent efforts to incorporate the
original complex geometry details of the pore space (such as surface dynamic alteration of wettability during two-phase has been conducted
roughness) have been discarded so that the term “infinite resolution” (Kallel et al., 2016) but much work remains. Questions that remain are:
should be taken with caution. It might be that the ability to study (a) Can wettability alteration rules based on pore morphology for pore-
phenomena on one's laptop (as opposed to having access to a super- network modeling be developed? (b) What insights can direct numer-
computer) is important or that many images (and in large domain sizes) ical simulation provide in regard to oil/water emulsion lens stability
are needed for investigating relevant rock/fluid parameters. Analytical and contact line pinning? (c) Can core-flood experiments shed light into
geometry descriptions allow for less computational effort and more wettability alteration in real-time or should functionalized micro-
complex physical processes and boundary conditions (e.g. three-phase fluidics experiments be leveraged? Microfluidics or lab-on-a-chip ex-
flow, mixed wettability, combination of diffusion and convection in periments are two-dimensional replicas of porous media. For a review
nano-scale flows) are explored with relative ease. Accordingly, to date, of lab-on-a-chip devices see Anbari et al. (2018), Karadimitriou and
pore-network studies are more developed than direct simulation Hassanizadeh (2012), and Mehmani et al. (2019).
methods. Finally, the complexity of a particular pore-network genera- Upscaling is an essential component in utilizing pore-scale modeling
tion method is proportional to its number of adjustable parameters, and prediction of flow transport properties that we did not cover. Porous
since such parameters (for example detailed mineralogy of pore/grain media with distinct hierarchies in scale, such as unconsolidated sand-
surfaces) may not be available, uncertainties increase. Multiple ad- stones, allow upscaling to be conducted by establishing average values
justable parameters, of course, create situations where there is no un- for transport properties. Volume average and homogenization techni-
ique optimal solution when matching a core-scale experiment. Having ques (Hornung, 1997; Whitaker, 1999), fall into this category. In highly
multiple plausible network model representations then creates un- heterogeneous media, such as carbonates, where distinction between
certainty in modeling a different experiment. The predictive capability scale hierarchies is not possible, nonlocal hybrid multiscale methods
of pore-network models has been discussed in depth by Sorbie and are utilized. In hybrid multiscale methods, the parameter of interest is
Skauge (2012). They argued that pore-network modeling cannot be calculated in microscale directly and extrapolated to larger scales based
predictive of the multiphase flow properties of mixed-wet systems. This on macroscale information. Scheibe et al., (2015) provides a review of
is due to the increased number of adjustment parameters that is needed various hybrid methods and a flowchart referred to as MAP (multiscale
in order to capture the pore-scale displacements accurately. In such analysis platform) that guides the researcher to a suitable upscaling
cases, one might prefer to rely on empirical models and tune them with approach following a series of questions. Given the high heterogeneity
core measurements that are typically fewer than the number of ad- of carbonates, assessment of MAP for various carbonate rock types can
justable parameters for a predictive pore-network model. Pore-network shed light on the most useful upscaling scheme.
models are still valuable for mixed-wet systems in that they allow
transport properties to be extrapolated for varying flow conditions after Acknowledgement
calibrating them with core measurements. Furthermore, they provide
description of interstitial fluid flow behavior which can be insightful for We acknowledge support from National Science Foundation
any type of environmental remediation or enhanced oil recovery. CAREER grant 1255622. We are grateful to Prof. Martin Blunt, Prof.
Capturing the heterogeneity of carbonate rocks is difficult using a Sebastian Geiger, and Dr. Yashar Mehmani for constructive comments
single imaging modality. High resolution 2D SEM maps remain a throughout the manuscript.

19
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Symbols

Pc Capillary pressure
Sw Water saturation
Φ Porosity
K Permeability
Kr Relative permeability
V Velocity
L Characteristic length scale
D Bulk fluid diffusivity
Ca Capillary number
M Mobility ratio
Acronyms

NAPL Nonaqueous phase liquid


XCT X-ray computed tomography
SEM Scanning electron microscopy
μCT X-ray microtomography
VOF Volume-of-fluid method
CLSVOF Coupled level set method
SPH Smoothed particle hydrodynamics
GLB Gray lattice Boltzmann
PLMM Pore-level multiscale method
PNM Pore-network model
MAP Multiscale analysis platform
DNS Direct numerical simulation
FVM Finite volume method
FEM Finite element method
BC Boundary condition
BGK Bhannagar Gross Krook
MRT Multiple relaxation time
EDS Energy dispersive spectroscopy
CLSM Confocal laser scanning microscopy
CFD Computational fluid dynamics
LBM Lattice Boltzmann method
DLA Diffusion-limited aggregation
BCC Body-centered cubic

Appendix. Example pore-throat network construction

We provide two simple examples in ImageJ/Fiji software (http://fiji.sc/) that outline acquiring basic grain and pore space analysis from seg-
mented porous material images (with one additional analysis done in MATLAB/Octave). ImageJ is a widely used and free image analysis software,
and Fiji is one of its variants with a number of useful plugins. Since contribution to ImageJ is public, the number of plugins are growing and there are
often multiple implementations of the same algorithm. MATLAB programming environment (https://www.mathworks.com/products/matlab.html)
is used for one small part of the example, and Octave (https://www.gnu.org/software/octave/) is its free alternative. The objective of this appendix
is a demonstration on the fundamentals on how a pore network is obtained from images. Algorithms were chosen based on their conceptual
simplicity and availability. Some advanced algorithms remain part of the individual research groups’ software solutions and are not readily available.

Sphere packing example

We start with a digitized two-dimensional cross-section from a synthetically generated monodisperse sphere pack (Finney, 1970). The center of
each sphere and a discretization of the pack can be found in Digital Rocks Portal (Finney, 2016). Slice 6 from the discretized image stack is selected
for this example. The data are binarized into values of 0 and 1 that represent the pore and solid space respectively. The pixel values are converted to
255 (white) for pore space and 0 (black) for solid space in order to comply with the ImageJ input requirements. The file (shown in Figure A1a) is also
available in supplementary information. Commands used to produce images are provided in the next section.
Duplicate the image and run Distance Map (see Figure A1b). Distance Map for each pore space pixel (labeled 255 in the original image) calculates
the distance to the closest solid pixel and displays it in the proportional shade of gray. Local maxima in distances are the starting point for the
watershed algorithm (the starting points can be observed by executing Ultimate Points). The watershed algorithm starts growing regions from the
local maxima. When the regions meet, the borders are established such as those shown in Figure A2a. The regions can be considered individual pores.
Subsequent to separating them with borders (which can be considered throats), we can label them using Find Connected Components (see result in
Figure A2b).

20
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. A1. Slice of the (a) original sphere pack and (b) the distance map corresponding to the pore space .

Fig. A2. Watershed algorithm result for (a) pore space (the borders that are identified in the process) and (b) subsequent labeling of individual connected com-
ponents in the pore space (individual pores).

Find Connected Components gives the area of each region (expressed in number of pixels in each region). The results can be saved as a.csv file (and
opened in a spreadsheet software for analysis as needed). The histogram of the areas can be produced with ImageJ as well (see Figure A3).

Fig. A3. Histogram of pore areas using Analyze- > Distribution applied to the Results from Find Connected Components and the snapshots of the entered parameters.

21
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

We can use similar tools to identify the throat information. However, in order to identify throats, we need to manipulate the images first. This is
done by subtracting segmented and watershed-separated images (we remind the reader that images are arrays of numbers). The watershed image
labels individual pores with pixel value of 255 and solid disks as well as throats with pixel value of 0. The original image labels pores and throats
with pixel value of 255 and solid disks with 0. By subtracting the two we obtain an image with label 255 where throats are, and 0 everywhere else
(see Figure A4a). We can run Find Connected Components on the subtracted image (Figure A4b), and obtain labeling and statistics on throats.
The procedures above have identified the objects that are required to create a pore-throat network. Capturing the pore-throat connectivity,
however, requires post-processing and we provide an example script for this purpose in MATLAB/Octave. The pore coordination number statistics is
shown in Figure A5.

Fig. A4. (a) Throats identified by subtracting the original and watershed images (a). (b)Throats labeled (colored) by Find Connected Components .

The aforementioned tools can be used to acquire grain statistics. The original image can be inverted so that the solid phase is white (labeled with
pixel value of 255), and the watershed procedure can be repeated. Most of the grains in this image are well separated, but the few that overlap can be
segmented with watersheds and finding the connected components. The result is shown in Figure A6.

Fig. A5. Pore coordination number histogram (result from MATLAB/Octave script shown below).

22
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. A6. Labeled individual grains.

List of relevant commands in ImageJ.

1. Plugins- > Macro- > Record (records trace of all commands in the ImageJ session)
2. File- > Open (for 2D images)
3. File- > Import (for 3D images)
4. Image- > Adjust- > Brightness/Contrast (when data is in 0,1 format, contrast needs to be adjusted in order to see the different image phases).
5. Edit- > Invert (2D, 3D) to invert phases (black to white, and white to black).
6. File- > Binary- > Watershed (2D only)
7. File- > Binary- > Distance Map (2D only)
8. File- > Binary- > Ultimate Points (2D only) will show the extreme points of the distance map from which the watershed algorithm has started.
9. Process- > Image Calculator

Trace of commands in ImageJ

The following are commands obtained by Plugins- > Macro- > Record in ImageJ. The commands can be copied into an ImageJ Macro and
executed via Plugins- > New- > Macro (the file with the commands SpherePack_Slice6_analysis_macro.ijm is provided in supplementary material).

//Reading in input image (Figure A1a)


open(“SpherePack_Slice6_Inverted.tif");

//Creating distance map of the pore space (Figure A1b)


run(“Duplicate … ", " ");
run(“Distance Map");
saveAs(“PNG”, “SpherePack_Slice6_DistanceMap.png");

//Figures A2,A3: watershed and identifying pores


selectWindow(“SpherePack_Slice6_Inverted.tif");
run(“Watershed");
saveAs(“PNG”, “SpherePack_Slice6_WatershedResult.png");

selectWindow(“SpherePack_Slice6_WatershedResult.png");
run(“Find Connected Regions”, “allow_diagonal display_one_image display_results regions_for_values_over = 100 minimum_number_of_points = 1 stop_after = −1″);
saveAs(“PNG”, “SpherePack_Slice6_LabeledPores.png");
saveAs(“Results”, “Results.csv");
run(“Distribution … ", “parameter = [Points In Region] or = 20 and = 0–2375″);

//save raw data for analysis in MATLAB


selectWindow(“SpherePack_Slice6_LabeledPores.png");
saveAs(“Raw Data”, “SpherePack_Slice6_LabeledPores_nx500ny500_16bit.raw");

23
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

//Throat analysis in Figure A4:

open(“SpherePack_Slice6_Inverted.tif”);//open again, just in case


imageCalculator(“Subtract create","SpherePack_Slice6_Inverted.tif","SpherePack_Slice6_WatershedResult.png");
selectWindow(“Result of SpherePack_Slice6_Inverted.tif");
saveAs(“PNG”, “SpherePack_Slice6_Throats.png");

run(“Find Connected Regions”, “allow_diagonal display_one_image display_results regions_for_values_over = 100 minimum_number_of_points = 1 stop_after = −1″);
saveAs(“PNG”, “SpherePack_Slice6_LabeledThroats.png");

//save raw data for analysis in MATLAB


selectWindow(“SpherePack_Slice6_LabeledThroats.png");
saveAs(“Raw Data”, “SpherePack_Slice6_LabeledThroats_nx500ny500_16bit.raw");

//Grain analysis in Figure A5:

open(“SpherePack_Slice6.tif");
run(“Watershed");
saveAs(“PNG”, “SpherePack_Slice6_GrainsWatershed.png");
run(“Find Connected Regions”, “allow_diagonal display_one_image display_results regions_for_values_over = 100 minimum_number_of_points = 1 stop_after = −1″);
saveAs(“PNG”, “SpherePack_Slice6_GrainsLabeled.png");

MATLAB/Octave commands to establish pore-throat network

%% This script reads in images with labeled pores and throats


%% For each throat, we search the neighborhood for neighboring
%% pores and record this information. At the same time
%% we record how many throats does each pore have in its
%% neighborhood (pore coordination number).

pore_labels = imread(‘SpherePack_Slice6_LabeledPores.png');
throat_labels = imread(‘SpherePack_Slice6_LabeledThroats.png');

% number of throats
num_throats = max(throat_labels());
num_pores = max(pore_labels());

% initialize array that will store pores that neighbor


% each throat
throat_neighbors = zeros(num_throats,2);
pore_coord_number = zeros(num_pores,1);

for t = 1:num_throats

% this is logical array corresponding to throat t


individual_throat = (throat_labels = = t);

% dilate this individual throat to get its immediate


% neighborhood, and then and then investigate which
% pore labels are found in its neighborhood
se = strel(‘sphere',2);
dilated_throat = imdilate(individual_throat, se);

% find pore labels in the neighborhood of a throat above


A = pore_labels(dilated_throat);

% 0 in pore label array is throat or solid - setdiff avoids it


% only unique labels are left
A = unique(setdiff(A,0));

if numel(A) < 2
fprintf(‘\nThroat %d not enough pores',t);
elseif numel(A) > 2
fprintf(‘\nThroat %d too many pores',t);
end

% record information
if numel(A) > 1
throat_neighbors(t,1) = A(1);
pore_coord_number(A(1)) = pore_coord_number(A(1))+1;
end

24
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

if numel(A) > 1
throat_neighbors(t,2) = A(2);
pore_coord_number(A(2)) = pore_coord_number(A(2))+2;
end
end

% Plot histogram of pore coordination numbers


max_coord_number = max(pore_coord_number);
hist(pore_coord_number,0:max_coord_number);
colormap gray
set(gca,'fontsize',16);
xlabel(‘pore coordination number');
ylabel(‘frequency');
axis tight
saveas(gcf,'SpherePack_Slice6_PoreCoordNumber','png');

Ketton carbonate example

We downloaded Ketton carbonate image sample (available online4). The downloaded image is segmented; the tomographic original is not
available. We work with slice 800 from that image (it is available for download in supplementary material).
Grain and pore labeling results of a watershed operation is demonstrated in Figure A7. The microporosity within some of the grains can create
seemingly multiple “broken” components. Similarly, small pieces of grain as well as grain surface roughness results in exceeding number of pores.
This is a common problem with such algorithms and typically can be resolved by reducing the number of the initial seeds of the watershed operation.
Tools for such analysis are available in an ImageJ plugin (Ollion et al., 2013).
Furthermore, we observe that because the Ketton carbonate pore space does not percolate in a single 2D cross-section, thus its pore-network
representation can be of little value. In contrast, a grain network model is connected.

ImageJ commands producing Figure A7

//Grain analysis
open(“Ketton_Substack_Slice800.tif");
run(“Watershed");
saveAs(“PNG”, “Ketton_Substack_Slice800_GrainWatershed.png");
run(“Find Connected Regions”, “allow_diagonal display_one_image display_results regions_for_values_over = 100 minimum_number_of_points = 1 stop_after = −1″);
saveAs(“PNG”, “Ketton_Substack_Slice800_GrainLabeled.png");
selectWindow(“Ketton_Substack_Slice800_GrainWatershed.png");

//Pore analysis
open(“Ketton_Substack_Slice800.tif");
run(“Invert");
run(“Watershed");
run(“Find
Connected Regions”, “allow_diagonal display_one_image display_results regions_for_values_over = 100 minimum_number_of_points = 1 stop_after = −1″);
saveAs(“PNG”, “Ketton_Substack_Slice800_PoresLabeled.png");

4
http://www.imperial.ac.uk/earth-science/research/research-groups/perm/
research/pore-scale-modelling/micro-ct-images-and-networks/.

25
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. A7. (a) Ketton carbonate slice , (b) labeled grains and (c) labeled pores .

3D analysis

The reader will notice that some of the commands are labeled ‘2D only’ in the list provided above. Three-dimensional image operations are
memory and computational time intensive and have been developed separately. Three-dimensional implementations are available (with no cost) in
‘3D ImageJ Suite’ plugin (Ollion et al., 2013), with documentation available online.5 We show an example watershed performed to identify different
pores in Figure A8. ImageJ instructions for creating it (including how to reduce number of starting points for watershed function, and thus noise) are
available in Supplementary material m).

5
http://imagejdocu.tudor.lu/doku.php?id=plugin:stacks:3d_ij_suite:start.

26
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Fig. A8. A slice through 3D watershed result on pore space. Throat barriers dividing pores are shown in blue, and the rest of the colors show different pores.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.marpetgeo.2019.104141.

References Beavers, G.S., Joseph, D.D., 1967 Oct. Boundary conditions at a naturally permeable wall.
J. Fluid Mech. 30 (1), 197–207. https://doi.org/10.1017/S0022112067001375.
Bear, J., 1972. Dynamics of Fluids in Porous Media. Dover, New York.
Abu-Al-Saud, M.O., Popinet, S., Tchelepi, H.A., 2018. A conservative and well-balanced Bear, J., 1991. Modelling transport phenomena in porous media. In: Convective Heat and
surface tension model. J. Comput. Phys. 371, 896–913. Mass Transfer in Porous Media. Springer, Dordrecht, pp. 7–69. https://doi.org/10.
Acharya, R.C., Van der Zee, S.E.A.T.M., Leijnse, A., 2007. Approaches for modeling 1007/978-94-011-3220-6_2.
longitudinal dispersion in pore-networks. Adv. Water Resour. 30 (2), 261–272. Bera, B., Gunda, N.S.K., Mitra, S.K., Vick, D., 2012. Characterization of nanometer-scale
https://doi.org/10.1016/j.advwatres.2005.11.015. porosity in reservoir carbonate rock by focused ion beam–scanning electron micro-
Aghaei, A., Piri, M., 2015. Direct pore-to-core up-scaling of displacement processes: dy- scopy. Microsc. Microanal. 18 (01), 171–178. https://doi.org/10.1017/
namic pore network modeling and experimentation. J. Hydrol. 522, 488–509. S1431927611012505.
https://doi.org/10.1016/j.jhydrol.2015.01.004. Berg, S., Ott, H., Klapp, S.A., Schwing, A., Neiteler, R., Brussee, N., Makurat, A., Leu, L.,
Aker, E., Måløy, K.J., Hansen, A., Batrouni, G.G., 1998. A two-dimensional network si- Enzmann, F., Schwarz, J.-O., Kersten, M., Irvine, S., Stampanoni, M., 2013. Real-time
mulator for two-phase flow in porous media. Transp. Porous Media 32 (2), 163–186. 3D imaging of Haines jumps in porous media flow. Proc. Natl. Acad. Sci. 110 (10),
https://doi.org/10.1023/A:1006510106194. 3755–3759. https://doi.org/10.1073/pnas.1221373110.
Alhammadi, A., Bijeljic, B., Blunt, M., 2018. X-ray Micro-tomography Datasets of Mixed- Bijeljic, B., Mostaghimi, P., Blunt, M.J., 2011. Signature of non-fickian solute transport in
Wet Carbonate Reservoir Rocks for in Situ Effective Contact Angle Measurements. complex heterogeneous porous media. Phys. Rev. Lett. 107 (20). https://doi.org/10.
https://doi.org/10.17612/P7VQ2G. 1103/PhysRevLett.107.204502.
Al-Raoush, R., Thompson, K., Willson, C.S., 2003. Comparison of network generation Bijeljic, B., Muggeridge, A.H., Blunt, M.J., 2004. Pore-scale modeling of longitudinal
techniques for unconsolidated porous media. Soil Sci. Soc. Am. J. 67 (6), 1687–1700. dispersion. Water Resour. Res. 40 (11), W11501. https://doi.org/10.1029/
https://doi.org/10.2136/sssaj2003.1687. 2004WR003567.
AlRatrout, A., Raeini, A.Q., Bijeljic, B., Blunt, M.J., 2017. Automatic measurement of Bijeljic, B., Raeini, A., Mostaghimi, P., Blunt, M.J., 2013. Predictions of non-Fickian solute
contact angle in pore-space images. Adv. Water Resour. 109, 158–169. https://doi. transport in different classes of porous media using direct simulation on pore-scale
org/10.1016/j.advwatres.2017.07.018. images. Phys. Rev. 87 (1). https://doi.org/10.1103/PhysRevE.87.013011.
Anbari, A., Chien, H.-T., Datta, S.S., Deng, W., Weitz, D.A., Fan, J., 2018. Microfluidic Biswal, B., Øren, P.-E., Held, R.J., Bakke, S., Hilfer, R., 2007. Stochastic multiscale model
model porous media: fabrication and applications. Small 14 (18), 1703575. https:// for carbonate rocks. Phys. Rev. 75 (6), 061303. https://doi.org/10.1103/PhysRevE.
doi.org/10.1002/smll.201703575. 75.061303.
Andrew, M.G., Bijeljic, B., Blunt, M.J., 2014. Reservoir-condition pore-scale imaging - Blunt, M.J., Bijeljic, B., Dong, H., Gharbi, O., Iglauer, S., Mostaghimi, P., Paluszny, A.,
contact angle, wettability, dynamics and trapping. In: Presented at the 76th EAGE Pentland, C., 2013. Pore-scale imaging and modelling. Adv. Water Resour. 51,
Conference and Exhibition 2014, Amsterdam, Netherlands, https://doi.org/10.3997/ 197–216. https://doi.org/10.1016/j.advwatres.2012.03.003.
2214-4609.20140848. Blunt, M., King, M.J., Scher, H., 1992. Simulation and theory of two-phase flow in porous
Andrew, M., Menke, H., Blunt, M.J., Bijeljic, B., 2015. The imaging of dynamic multi- media. Phys. Rev. A 46 (12), 7680–7699. https://doi.org/10.1103/PhysRevA.46.
phase fluid flow using synchrotron-based X-ray microtomography at reservoir con- 7680.
ditions. Transp. Porous Media 110 (1), 1–24. https://doi.org/10.1007/s11242-015- Blunt, M., King, P., 1991. Relative permeabilities from two- and three-dimensional pore-
0553-2. scale network modelling. Transp. Porous Media 6 (4), 407–433. https://doi.org/10.
Archie, G.E., 1942. The electrical resistivity log as an aid in determining some reservoir 1007/BF00136349.
characteristics. Transactions of the AIME 146 (1), 54–62. https://doi.org/10.2118/ Boek, E.S., Venturoli, M., 2010. Lattice-Boltzmann studies of fluid flow in porous media
942054-G. with realistic rock geometries. Comput. Math. Appl. 59 (7), 2305–2314. https://doi.
Arns, C.H., Bauget, F., Limaye, A., Sakellariou, A., Senden, T., Sheppard, A., Sok, R.M., org/10.1016/j.camwa.2009.08.063.
Pinczewski, V., Bakke, S., Berge, L.I., Oren, P.E., 2005 Dec. Pore scale characteriza- Brackbill, J.U., Kothe, D.B., Zemach, C., 1992. A continuum method for modeling surface
tion of carbonates using X-ray microtomography. SPE J 10 (4), 475–484. tension. J. Comput. Phys. 100 (2), 335–354. https://doi.org/10.1016/0021-9991(92)
Baek, S.-H., Kwon, Tae-Hyuk, 2018. X-ray CMT Images of Abiotic Carbonate Precipitation 90240-Y.
in Porous Media from a Supersaturated Solution [Data Set]. Digital Rocks Portal. The Brooks, R., Corey, T., 1964. Hydraulic Properties of Porous Media. Hydrology Papers.
University of Texas at Austinhttps://doi.org/10.17612/p70963. Colorado State University.
Bakke, S., Øren, P.-E., 1997. 3-D pore-scale modelling of sandstones and flow simulations Bruderer, C., Bernabé, Y., 2001. Network modeling of dispersion: transition from Taylor
in the pore networks. SPE J. 2 (02), 136–149. https://doi.org/10.2118/35479-PA. Dispersion in homogeneous networks to mechanical dispersion in very heterogeneous
Bauer, D., Youssef, S., Fleury, M., Bekri, S., Rosenberg, E., Vizika, O., 2012. Improving the ones. Water Resour. Res. 37 (4), 897–908. https://doi.org/10.1029/2000WR900362.
estimations of petrophysical transport behavior of carbonate rocks using a dual pore Bryant, S., Blunt, M., 1992. Prediction of relative permeability in simple porous media.
network approach combined with computed microtomography. Transp. Porous Phys. Rev. A 46 (4), 2004–2011. https://doi.org/10.1103/PhysRevA.46.2004.
Media 94 (2), 505–524. https://doi.org/10.1007/s11242-012-9941-z. Bryant, S.L., King, P.R., Mellor, D.W., 1993. Network model evaluation of permeability

27
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

and spatial correlation in a real random sphere packing. Transp. Porous Media 11 (1), Fredrich, J.T., 1999. 3D imaging of porous media using laser scanning confocal micro-
53–70. https://doi.org/10.1007/BF00614635. scopy with application to microscale transport processes. Phys. Chem. Earth A Solid
Bryant, S.L., Mellor, D.W., Cade, C.A., 1993. Physically representative network models of Earth Geod. 24 (7), 551–561. https://doi.org/10.1016/S1464-1895(99)00079-4.
transport in porous media. AIChE J. 39 (3), 387–396. https://doi.org/10.1002/aic. Fredrich, J.T., Menendez, B., Wong, T.-F., 1995. Imaging the pore structure of geoma-
690390303. terials. Science 268 (5208), 276–279. https://doi.org/10.1126/science.268.5208.
Bryntesson, L.M., 2002. Pore network modelling of the behaviour of a solute in chro- 276.
matography media: transient and steady-state diffusion properties. J. Chromatogr. A Freire-Gormaly, M., Ellis, J.S., Bazylak, A., MacLean, H.L., 2015. Comparing thresholding
945 (1–2), 103–115. https://doi.org/10.1016/S0021-9673(01)01485-6. techniques for quantifying the dual porosity of Indiana Limestone and Pink Dolomite.
Bultreys, T., Boone, M.A., De Kock, T., De Schutter, G., Vontobel, P., Van Hoorebeke, L., Microporous Mesoporous Mater. 207, 84–89. https://doi.org/10.1016/j.micromeso.
Cnudde, V., 2017. Massangis Jaune Carbonate [Data Set]. Digital Rocks Portal. The 2015.01.002.
University of Texas at Austin. Gao, Y., Sharma, M.M., 1994. A LGA model for fluid flow in heterogeneous porous media.
Bultreys, T., De Boever, W., Cnudde, V., 2016. Imaging and image-based fluid transport Transp. Porous Media 17 (1), 1–7.
modeling at the pore scale in geological materials: a practical introduction to the Garcia, X., Akanji, L.T., Blunt, M.J., Matthai, S.K., Latham, J.P., 2009a. Numerical study
current state-of-the-art. Earth Sci. Rev. 155, 93–128. https://doi.org/10.1016/j. of the effects of particle shape and polydispersity on permeability. Phys. Rev. 80 (2),
earscirev.2016.02.001. 021304. https://doi.org/10.1103/PhysRevE.80.021304.
Bultreys, T., Hoorebeke, L.V., Cnudde, V., 2018. Estaillades Carbonate #2 [Data Set]. Gerlach, D., Tomar, G., Biswas, G., Durst, F., 2006. Comparison of volume-of-fluid
Digital Rocks Portal. https://doi.org/10.17612/P7C09J. methods for surface tension-dominant two-phase flows. Int. J. Heat Mass Transf. 49
Bultreys, T., Stappen, J.V., Kock, T.D., Boever, W.D., Boone, M.A., Hoorebeke, L.V., (3–4), 740–754. https://doi.org/10.1016/j.ijheatmasstransfer.2005.07.045.
Cnudde, V., 2017. Savonnières Carbonate [Data Set]. Digital Rocks Portal. The Gharbi, O., Blunt, M.J., 2012. The impact of wettability and connectivity on relative
University of Texas at Austin. permeability in carbonates: a pore network modeling analysis. Water Resour. Res. 48
Bultreys, T., Van Hoorebeke, L., Cnudde, V., 2015. Multi-scale, micro-computed tomo- (12), W12513. https://doi.org/10.1029/2012WR011877.
graphy-based pore network models to simulate drainage in heterogeneous rocks. Adv. Golparvar, A., Zhou, Y., Wu, K., Ma, J., Yu, Z., 2018 Dec 25. A comprehensive review of
Water Resour. 78, 36–49. https://doi.org/10.1016/j.advwatres.2015.02.003. pore scale modeling methodologies for multiphase flow in porous media. Adv. Geo-
Caers, J., 2001. Geostatistical reservoir modelling using statistical pattern recognition. J. Energy Res. 2 (4), 418–440.
Pet. Sci. Eng. 29 (3–4), 177–188. https://doi.org/10.1016/S0920-4105(01)00088-2. Gostick, J., Aghighi, M., Hinebaugh, J., Tranter, T., Hoeh, M.A., Day, H., Spellacy, B.,
Carman, P.C., 1937. Fluid Flow through Granular Beds, vol. 15. Transactions of Sharqawy, M.H., Bazylak, A., Burns, A., Lehnert, W., Putz, A., 2016. OpenPNM: a
Institution of Chemical Engineering (London), pp. 150–166. pore network modeling package. Comput. Sci. Eng. 18 (4), 60–74. https://doi.org/
Chandra, V., Barnett, A., Corbett, P., Geiger, S., Wright, P., Steele, R., Milroy, P., 2015. 10.1109/MCSE.2016.49.
Effective integration of reservoir rock-typing and simulation using near-wellbore Guo, B., Mehmani, Y., Tchelepi, H.A., 2019. Multiscale formulation of pore-scale com-
upscaling. Mar. Pet. Geol. 67, 307–326. https://doi.org/10.1016/j.marpetgeo.2015. pressible Darcy-Stokes flow. J. Comput. Phys. 397, 108849. https://doi.org/10.1016/
05.005. j.jcp.2019.07.047.
Chatzis, I., Dullien, F.a.L., 1977. Modelling pore structure by 2-D and 3-D networks with Gunstensen, A.K., Rothman, D.H., Zaleski, S., Zanetti, G., 1991. Lattice Boltzmann model
ApplicationTo sandstones. J. Can. Pet. Technol. 16 (01). https://doi.org/10.2118/77- of immiscible fluids. Phys. Rev. A 43 (8), 4320–4327. https://doi.org/10.1103/
01-09. PhysRevA.43.4320.
Chen, J.-D., Wilkinson, D., 1985. Pore-scale viscous fingering in porous media. Phys. Rev. Harting, J., Chin, J., Venturoli, M., Coveney, P.V., 2005. Large-scale lattice Boltzmann
Lett. 55 (18), 1892–1895. https://doi.org/10.1103/PhysRevLett.55.1892. simulations of complex fluids: advances through the advent of computational Grids.
Coon, E.T., Porter, M.L., Kang, Q., 2014. Taxila LBM: a parallel, modular lattice Phil. Trans. R. Soc. Lond. A: Math. Phys. and Eng. Sci. 363 (1833), 1895–1915.
Boltzmann framework for simulating pore-scale flow in porous media. Comput. https://doi.org/10.1098/rsta.2005.1618.
Geosci. 18 (1), 17–27. https://doi.org/10.1007/s10596-013-9379-6. Hatiboglu, C.U., Babadagli, T., 2008. Pore-scale studies of spontaneous imbibition into
Cundall, P.A., Strack, O.D.L., 1979. A discrete numerical model for granular assemblies. oil-saturated porous media. Phys. Rev. 77 (6), 066311. https://doi.org/10.1103/
Geotechnique 29 (1), 47–65. https://doi.org/10.1680/geot.1979.29.1.47. PhysRevE.77.066311.
Dewers, T.A., Heath, J., Ewy, R., Duranti, L., 2012. Three-dimensional pore networks and Hazlett, R.D., 1995. Simulation of capillary-dominated displacements in microtomo-
transport properties of a shale gas formation determined from focused ion beam serial graphic images of reservoir rocks. Transp. Porous Media 20 (1–2), 21–35. https://doi.
imaging. Int. J. Oil Gas Coal Technol. 5 (2–3), 229–248. org/10.1007/BF00616924.
Dias, M.M., Wilkinson, D., 1986. Percolation with trapping. J. Phys. A Math. Gen. 19 (15), Hazlett, R.D., Chen, S.Y., Soll, W.E., 1998. Wettability and rate effects on immiscible
3131–3146. https://doi.org/10.1088/0305-4470/19/15/034. displacement: lattice Boltzmann simulation in microtomographic images of reservoir
Dixit, A.B., Buckley, J.S., McDougall, S.R., Sorbie, K.S., 2000. Empirical measures of rocks. J. Pet. Sci. Eng. 20 (3), 167–175. https://doi.org/10.1016/S0920-4105(98)
wettability in porous media and the relationship between them derived from pore- 00017-5.
scale modelling. Transp. Porous Media 40 (1), 27–54. https://doi.org/10.1023/ Heidari, Z., Posenato Garcia, A., 2016. Austin chalk [data set]. Digital Rocks Portal.
A:1006687829177. https://doi.org/10.17612/P73011.
Dixit, A.B., McDougall, S.R., Sorbie, K.S., 1998. A pore-level investigation of relative Heilbronner, R., Barrett, S., 2013. Image Analysis in Earth Sciences: Microstructures and
permeability hysteresis in water-wet systems. SPE J. 3 (2), 115–123. https://doi.org/ Textures of Earth Materials. Springer Science & Business Media.
10.2118/37233-PA. Hill, R., 1963. Elastic properties of reinforced solids: some theoretical principles. J. Mech.
Dixit, A.B., McDougall, S.R., Sorbie, K.S., 1998. Analysis of Relative Permeability Phys. Solids 11 (5), 357–372. https://doi.org/10.1016/0022-5096(63)90036-X.
Hysteresis Trends in Mixed-Wet Porous Media Using. Society of Petroleum Engineers. Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the dynamics of free
Dixit, A.B., McDougall, S.R., Sorbie, K.S., Buckley, J.S., 1996. Pore Scale Modelling of boundaries. J. Comput. Phys. 39 (1), 201–225. https://doi.org/10.1016/0021-
Wettability Effects and Their Influence on Oil Recovery. Society of Petroleum 9991(81)90145-5.
Engineershttps://doi.org/10.2118/35451-MS. Holtzman, R., Silin, D.B., Patzek, T.W., 2008. Micromechanics of hydrate dissociation in
Dodd, C.G., Kiel, O.G., 1959. Evaluation of Monte Carlo methods in studying fluid–fluid marine sediments by grain-scale simulations. In: Presented at the SPE Western
displacements and wettability in porous rocks. J. Phys. Chem. 63 (10), 1646–1652. Regional and Pacific Section AAPG Joint Meeting. Society of Petroleum Engineers.
https://doi.org/10.1021/j150580a019. https://doi.org/10.2118/114223-MS.
Dupont, J.-B., Legendre, D., 2010. Numerical simulation of static and sliding drop with Hornung, U. (Ed.), 1997. Homogenization and Porous Media. Springer, New York.
contact angle hysteresis. J. Comput. Phys. 229 (7), 2453–2478. https://doi.org/10. Hosa, A., Wood, R., 2017. Quantifying the impact of early calcite cementation on the
1016/j.jcp.2009.07.034. reservoir quality of carbonate rocks: a 3D process-based model. Adv. Water Resour.
Enright, D., Fedkiw, R., Ferziger, J., Mitchell, I., 2002. A hybrid particle level set method 104, 89–104. https://doi.org/10.1016/j.advwatres.2017.02.019.
for improved interface capturing. J. Comput. Phys. 183 (1), 83–116. https://doi.org/ Huang, H., Meakin, P., Liu, M., 2005. Computer simulation of two-phase immiscible fluid
10.1006/jcph.2002.7166. motion in unsaturated complex fractures using a volume of fluid method. Water
Fatt, I., 1956a. The network model of porous media I. Capillary pressure characteristics. Resour. Res. 41 (12), W12413. https://doi.org/10.1029/2005WR004204.
Pet. Transa., AIME 207, 144–159. Iassonov, P., Gebrenegus, T., Tuller, M., 2009. Segmentation of X-ray computed tomo-
Fatt, I., 1956b. The network model of porous media II. Dynamic properties of a single size graphy images of porous materials: a crucial step for characterization and quantita-
tube network. Pet. Transa., AIME 207, 160–163. tive analysis of pore structures. Water Resour. Res. 45 (9). https://doi.org/10.1029/
Fatt, I., 1956c. The network model of porous media III. Dynamic propeties of networks 2009WR008087.
with tube radius distribution 207, 164–177. Idowu, N.A., Blunt, M.J., 2010. Pore-scale modelling of rate effects in waterflooding.
Fenwick, D.H., Blunt, M.J., 1998. Three-dimensional modeling of three phase imbibition Transp. Porous Media 83 (1), 151–169. https://doi.org/10.1007/s11242-009-
and drainage. Adv. Water Resour. 21 (2), 121–143. https://doi.org/10.1016/S0309- 9468-0.
1708(96)00037-1. Jennings, J.W., Lucia, F.J., 2001. Predicting permeability from well logs in carbonates
Ferrari, A., Lunati, I., 2013. Direct numerical simulations of interface dynamics to link with a link to geology for interwell permeability mapping. In: SPE Annual Technical
capillary pressure and total surface energy. Adv. Water Resour. 57, 19–31. https:// Conference and Exhibition. Society of Petroleum Engineers, New Orleans, Louisiana.
doi.org/10.1016/j.advwatres.2013.03.005. https://doi.org/10.2118/71336-MS.
Finney, J.L., 1970. Random packings and the structure of simple liquids. I. The geometry Jerauld, G.R., Salter, S.J., 1990. The effect of pore-structure on hysteresis in relative
of random close packing. Proc. R. Soc. Lond.: Math. Phys. and Eng. Sci. 319 (1539), permeability and capillary pressure: pore-level modeling. Transp. Porous Media 5 (2),
479–493. https://doi.org/10.1098/rspa.1970.0189. 103–151. https://doi.org/10.1007/BF00144600.
Finney, J.L., 2016. Finney Packing of Spheres [Data Set]. Digital Rocks Portal. https:// Jettestuen, E., Helland, J.O., Prodanović, M., 2013. A level set method for simulating
doi.org/10.17612/P78G69. capillary-controlled displacements at the pore scale with nonzero contact angles.
Frank, F., Liu, C., Alpak, F.O., Riviere, B., 2018. A finite volume/discontinuous Galerkin Water Resour. Res. 49 (8), 4645–4661. https://doi.org/10.1002/wrcr.20334.
method for the advective Cahn–Hilliard equation with degenerate mobility on porous Jha, R.K., Bryant, S., Lake, L.W., 2011. Effect of diffusion on dispersion. SPE J. 16 (01),
domains stemming from micro-CT imaging. Comput. Geosci. 22 (2), 543–563. 65–77. https://doi.org/10.2118/115961-PA.
Frank, F., Liu, C., Scanziani, A., Alpak, F.O., Riviere, B., 2018. An energy-based equili- Jiang, Z., van Dijke, M., Geiger, S., Kronbauer, D., Mantovani, I., Fernandes, C., 2013.
brium contact angle boundary condition on jagged surfaces for phase-field methods. Impact of the Spatial Correlation of Microporosity on Fluid Flow in Carbonate Rocks.
J. Colloid Interface Sci. 523, 282–291. https://doi.org/10.1016/j.jcis.2018.02.075. Society of Petroleum Engineers Paper number 166001.

28
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

Jiang, Z., van Dijke, M.I.J., Geiger, S., Ma, J., Couples, G.D., Li, X., 2017. Pore network Liang, Z., Ioannidis, M.A., Chatzis, I., 2000. Geometric and topological analysis of three-
extraction for fractured porous media. Adv. Water Resour. 107, 280–289. https://doi. dimensional porous media: pore space partitioning based on morphological skeleto-
org/10.1016/j.advwatres.2017.06.025. nization. J. Colloid Interface Sci. 221 (1), 13–24. https://doi.org/10.1006/jcis.1999.
Jiang, Z., van Dijke, M.I.J., Sorbie, K.S., Couples, G.D., 2013. Representation of multiscale 6559.
heterogeneity via multiscale pore networks. Water Resour. Res. 49 (9), 5437–5449. Lin, C.-Y., Slattery, J.C., 1982. Three-dimensional, randomized, network model for two-
https://doi.org/10.1002/wrcr.20304. phase flow through porous media. AIChE J. 28 (2), 311–324. https://doi.org/10.
Jin, G., Patzek, T.W., Silin, D.B., 2003. Physics-based reconstruction of sedimentary rocks. 1002/aic.690280221.
In: Presented at the SPE Western Regional/AAPG Pacific Section Joint Meeting. Lindquist, W.B., Lee, S.M., Oh, W., Venkatarangan, A.B., Shin, H., Prodanović, M., 2005.
Society of Petroleum Engineers. https://doi.org/10.2118/83587-MS. 3DMA-Rock: A software package for automated analysis of rock pore structure in 3-D
Jobe, T.D., Geiger, S., Jiang, Z., Zhang, S., Agar, S., 2018. Micropore network modelling computed microtomography images. Department of Applied Mathematics and
from 2D confocal imagery: impact on reservoir quality and hydrocarbon recovery. Statistics, Stony Brook University, Stony Brook, NY. Retrieved from. http://www.
Pet. Geosci. 24 (3), 323–334. https://doi.org/10.1144/petgeo2017-017. ams.sunysb.edu/lindquis/3dma/3dma_rock/3dma_rock.html.
Joekar Niasar, V., 2010. The Immiscibles: Capillarity Effects in Porous Media - Pore- Lindquist, W.B., Venkatarangan, A., 1999. Investigating 3D geometry of porous media
Network Modelling. Geologica Ultraiectina (318). PhD Thesis. Utrecht University from high resolution images. Phys. Chem. Earth A Solid Earth Geod. 24 (7), 593–599.
Retrieved from. http://dspace.library.uu.nl/handle/1874/42365. https://doi.org/10.1016/S1464-1895(99)00085-X.
Joekar-Niasar, V., Hassanizadeh, S.M., Dahle, H.K., 2010. Non-equilibrium effects in Liu, H., Kang, Q., Leonardi, C.R., Schmieschek, S., Narváez, A., Jones, B.D., Williams, J.R.,
capillarity and interfacial area in two-phase flow: dynamic pore-network modelling. Valocchi, A.J., Harting, J., 2016. Multiphase lattice Boltzmann simulations for porous
J. Fluid Mech. 655, 38–71. https://doi.org/10.1017/S0022112010000704. media applications. Comput. Geosci. 20 (4), 777–805. https://doi.org/10.1007/
Joekar-Niasar, V., Hassanizadeh, S.M., Leijnse, A., 2007. Insights into the relationships s10596-015-9542-3.
among capillary pressure, saturation, interfacial area and relative permeability using Liu, H., Valocchi, A.J., Werth, C., Kang, Q., Oostrom, M., 2014. Pore-scale simulation of
pore-network modeling. Transp. Porous Media 74 (2), 201–219. https://doi.org/10. liquid CO2 displacement of water using a two-phase lattice Boltzmann model. Adv.
1007/s11242-007-9191-7. Water Resour. 73, 144–158. https://doi.org/10.1016/j.advwatres.2014.07.010.
Kallel, W., van Dijke, M.I.J., Sorbie, K.S., Wood, R., Jiang, Z., Harland, S., 2016. Løvoll, G., Méheust, Y., Måløy, K.J., Aker, E., Schmittbuhl, J., 2005. Competition of
Modelling the effect of wettability distributions on oil recovery from microporous gravity, capillary and viscous forces during drainage in a two-dimensional porous
carbonate reservoirs. Adv. Water Resour. 95, 317–328. https://doi.org/10.1016/j. medium, a pore scale study. Energy 30 (6), 861–872. https://doi.org/10.1016/j.
advwatres.2015.05.025. energy.2004.03.100.
Kang, Q., Lichtner, P.C., Zhang, D., 2006. Lattice Boltzmann pore-scale model for mul- Lowry, M.I., Miller, C.T., 1995. Pore-scale modeling of nonwetting-phase residual in
ticomponent reactive transport in porous media. J. Geophys. Res.: Solid Earth 111 porous media. Water Resour. Res. 31 (3), 455–473. https://doi.org/10.1029/
(B5), B05203. https://doi.org/10.1029/2005JB003951. 94WR02849.
Kang, Q., Zhang, D., Chen, S., 2004. Immiscible displacement in a channel: simulations of Manwart, C., Aaltosalmi, U., Koponen, A., Hilfer, R., Timonen, J., 2002. Lattice-
fingering in two dimensions. Adv. Water Resour. 27 (1), 13–22. https://doi.org/10. Boltzmann and finite-difference simulations for the permeability for three-dimen-
1016/j.advwatres.2003.10.002. sional porous media. Phys. Rev. 66 (1), 016702. https://doi.org/10.1103/PhysRevE.
Karadimitriou, N.K., Hassanizadeh, S.M., 2012. A review of micromodels and their use in 66.016702.
two-phase flow studies. Vadose Zone J. 11 (3), 0. https://doi.org/10.2136/vzj2011. Martys, N.S., Chen, H., 1996. Simulation of multicomponent fluids in complex three-
0072. dimensional geometries by the lattice Boltzmann method. Phys. Rev. 53 (1), 743.
Karimpouli, S., Tahmasebi, P., 2016. Conditional reconstruction: an alternative strategy https://doi.org/10.1103/PhysRevE.53.743.
in digital rock physics. Geophysics 81 (4), D465–D477. https://doi.org/10.1190/ Mason, G., Mellor, D., 1995. Simulation of drainage and imbibition in a random packing
geo2015-0260.1. of equal spheres. J. Colloid Interface Sci. 176 (1), 214–225. https://doi.org/10.1006/
Kats, F.M., Egberts, P.J.P., 1999. Simulation of three-phase displacement mechanisms jcis.1995.0024.
using a 2D lattice-Boltzmann model. Transp. Porous Media 37 (1), 55–68. https:// McClure, J.E., Berrill, M.A., Gray, W.G., Miller, C.T., 2016. Influence of phase con-
doi.org/10.1023/A:1006502831641. nectivity on the relationship among capillary pressure, fluid saturation, and inter-
Keller, A.L., Zeidler, D., Kemen, T., 2014. High throughput data acquisition with a multi- facial area in two-fluid-phase porous medium systems. Phys. Rev. 94 (3), 033102.
beam SEM. In: Scanning Microscopies, vol. 9236. International Society for Optics and https://doi.org/10.1103/PhysRevE.94.033102.
Photonics, pp. 92360B. https://doi.org/10.1117/12.2069119. 2014. McClure, J.E., Pan, C., Adalsteinsson, D., Gray, W.G., Miller, C.T., 2004. Estimating in-
Kelly, S., El-Sobky, H., Torres-Verdín, C., Balhoff, M.T., 2016. Assessing the utility of FIB- terfacial areas resulting from lattice Boltzmann simulation of two-fluid-phase flow in
SEM images for shale digital rock physics. Adv. Water Resour. 95, 302–316. https:// a porous medium. In: Developments in Water Science, vol. 55. Elsevier, pp. 23–35.
doi.org/10.1016/j.advwatres.2015.06.010. https://doi.org/10.1016/S0167-5648(04)80034-0.
Kemen, T., Malloy, M., Thiel, B., Mikula, S., Denk, W., Dellemann, G., Zeidler, D., 2015. McClure, J.E., Prins, J.F., Miller, C.T., 2014. A novel heterogeneous algorithm to simulate
Further advancing the throughput of a multi-beam SEM. In: Metrology, Inspection, multiphase flow in porous media on multicore CPU–GPU systems. Comput. Phys.
and Process Control for Microlithography, vol. 9424. International Society for Optics Commun. 185 (7), 1865–1874. https://doi.org/10.1016/j.cpc.2014.03.012.
and Photonics, pp. 94241U. https://doi.org/10.1117/12.2188560. XXIX. McDougall, S.R., Sorbie, K.S., 1995. The impact of wettability on waterflooding: pore-
Khan, H., Gonzales, A., Prodanovic, M., Heidari, Z., Espinoza, D.N., 2018. Guelph scale simulation. SPE Reserv. Eng. 10 (03), 208–213. https://doi.org/10.2118/
Dolomite Characterization [Data Set]. Digital Rocks Portal. The University of Texas at 25271-PA.
Austinhttps://doi.org/10.17612/p78m25. Zolfaghari, A., Piri, M., 2017. Pore-scale network modeling of three-phase flow based on
Khan, H., Pope, G., 2018. Guelph Dolomite [Data Set]. https://doi.org/10.17612/ thermodynamically consistent threshold capillary pressures. II. Results. Transp.
p74t1m. Porous Media 116 (3), 1139–1165. https://doi.org/10.1007/s11242-016-0815-7.
Kim, D., Peters, C.A., Lindquist, W.B., 2011. Upscaling geochemical reaction rates ac- Mehmani, Y., Balhoff, M.T., 2015. Mesoscale and hybrid models of fluid flow and solute
companying acidic CO2-saturated brine flow in sandstone aquifers. Water Resour. transport. Rev. Mineral. Geochem. 80 (1), 433–459. https://doi.org/10.2138/rmg.
Res. 47 (1), W01505. https://doi.org/10.1029/2010WR009472. 2015.80.13.
Koplik, J., Lasseter, T.J., 1985. Two-phase flow in random network models of porous Mehmani, A., Mehmani, Y., Prodanović, M., Balhoff, M., 2015. A forward analysis on the
media. Soc. Pet. Eng. J. 25 (01), 89–100. https://doi.org/10.2118/11014-PA. applicability of tracer breakthrough profiles in revealing the pore structure of tight
Kovscek, A.R., Wong, H., Radke, C.J., 1993. A pore-level scenario for the development of gas sandstone and carbonate rocks. Water Resour. Res. 51 (6), 4751–4767. https://
mixed wettability in oil reservoirs. AIChE J. 39 (6), 1072–1085. https://doi.org/10. doi.org/10.1002/2015WR016948.
1002/aic.690390616. Mehmani, A., Prodanović, M., 2014a. The application of sorption hysteresis in nano-
Kozeny, J., 1927. Über kapillare Leitung des Wassers im Boden, vol. 136 Sitz. der Wien. petrophysics using multiscale multiphysics network models. Int. J. Coal Geol. 128,
Akad. Der Wissenschaften. 96–108. https://doi.org/10.1016/j.coal.2014.03.008.
Landry, C.J., Karpyn, Z.T., Ayala, O., 2014. Relative permeability of homogenous-wet and Mehmani, A., Prodanović, M., 2014b. The effect of microporosity on transport properties
mixed-wet porous media as determined by pore-scale lattice Boltzmann modeling. in porous media. Adv. Water Resour. 63, 104–119. https://doi.org/10.1016/j.
Water Resour. Res. 50 (5), 3672–3689. https://doi.org/10.1002/2013WR015148. advwatres.2013.10.009.
Latva-Kokko, M., Rothman, D.H., 2005. Diffusion properties of gradient-based lattice Mehmani, A., Prodanović, M., Javadpour, F., 2013. Multiscale, multiphysics network
Boltzmann models of immiscible fluids. Phys. Rev. 71 (5), 056702. https://doi.org/ modeling of shale matrix gas flows. Transp. Porous Media 99 (2), 377–390. https://
10.1103/PhysRevE.71.056702. doi.org/10.1007/s11242-013-0191-5.
Lee, Kashyap, Chu, 1994. Building skeleton models via 3-D medial surface Axis thinning Mehmani, A., Tokan-Lawal, A., Prodanović, M., Sheppard, A., 2011. The effect of mi-
algorithms. CVGIP Graph. Models Image Process. 56, 462–478. https://doi.org/10. croporosity on transport properties in tight reservoirs. Paper number 144384 In: SPE
1006/cgip.1994.1042. Americas Unconventional Gas Conference and Exhibition Proceedings. Society of
Lenormand, R., Touboul, E., Zarcone, C., 1988. Numerical models and experiments on Petroleum Engineers. https://doi.org/10.2118/144384-MS.
immiscible displacements in porous media. J. Fluid Mech. 189, 165–187. https://doi. Mehmani, Y., 2014. Modeling Single-phase Flow and Solute Transport across Scales
org/10.1017/S0022112088000953. (Doctoral Dissertation). The University of Texas at Austin.
Lenormand, R., Zarcone, C., 1984. Role of Roughness and Edges during Imbibition in Mehmani, Y., Balhoff, M.T., 2015. Eulerian network modeling of longitudinal dispersion.
Square Capillaries. Society of Petroleum Engineershttps://doi.org/10.2118/ Water Resour. Res. 51 (10), 8586–8606. https://doi.org/10.1002/2015WR017543.
13264-MS. Mehmani, Y., Oostrom, M., Balhoff, M.T., 2014. A streamline splitting pore-network
Lerdahl, T., Øren, P.-E., Bakke, S., 2000. A Predictive Network Model for Three-phase approach for computationally inexpensive and accurate simulation of transport in
Flow in Porous Media. Society of Petroleum Engineershttps://doi.org/10.2118/ porous media. Water Resour. Res. 50 (3), 2488–2517. https://doi.org/10.1002/
59311-MS. 2013WR014984.
Li, L., Peters, C.A., Celia, M.A., 2006. Upscaling geochemical reaction rates using pore- Mehmani, Y., Sun, T., Balhoff, M.T., Eichhubl, P., Bryant, S., 2012. Multiblock pore-scale
scale network modeling. Adv. Water Resour. 29 (9), 1351–1370. https://doi.org/10. modeling and upscaling of reactive transport: application to carbon sequestration.
1016/j.advwatres.2005.10.011. Transp. Porous Media 95 (2), 305–326. https://doi.org/10.1007/s11242-012-
Li, X., Zhang, Y., Wang, X., Ge, W., 2013. GPU-based numerical simulation of multi-phase 0044-7.
flow in porous media using multiple-relaxation-time lattice Boltzmann method. Mehmani, Y., Tchelepi, H.A., 2017. Minimum requirements for predictive pore-network
Chem. Eng. Sci. 102, 209–219. https://doi.org/10.1016/j.ces.2013.06.037. modeling of solute transport in micromodels. Adv. Water Resour. 108, 83–98.

29
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

https://doi.org/10.1016/j.advwatres.2017.07.014. imbibition in fractures. SPE J. 14 (3), 532–542. https://doi.org/10.2118/110448-PA.


Mehmani, Y., Tchelepi, H.A., 2018. Multiscale computation of pore-scale fluid dynamics: Prodanović, M., Bryant, S.L., 2006. A level set method for determining critical curvatures
single-phase flow. J. Comput. Phys. 375, 1469–1487. https://doi.org/10.1016/j.jcp. for drainage and imbibition. J. Colloid Interface Sci. 304 (2), 442–458. https://doi.
2018.08.045. org/10.1016/j.jcis.2006.08.048.
Mehmani, Y., Tchelepi, H.A., 2019. Multiscale formulation of two-phase flow at the pore Prodanovic, M., Esteva, M., Hanlon, M., Nanda, G., Agarwal, P., 2015. Digital rocks
scale. J. Comput. Phys. 389, 164–188. https://doi.org/10.1016/j.jcp.2019.03.035. portal. Digital Rocks Portal. https://doi.org/10.17612/P7CC7K.
Min, C., Gibou, F., 2007. A second order accurate level set method on non-graded Prodanović, M., Holder, J.T., Bryant, S.L., 2009. Coupling capillarity-controlled fluid
adaptive cartesian grids. J. Comput. Phys. 225 (1), 300–321. https://doi.org/10. displacement with unconsolidated sediment mechanics: grain scale fracture opening.
1016/j.jcp.2006.11.034. In: SPE Annual Technical Conference and Exhibition Proceedings, https://doi.org/10.
Mirabolghasemi, M., Prodanović, M., DiCarlo, D., Ji, H., 2015. Prediction of empirical 2118/124717-MS. (p. Paper number 124717).
properties using direct pore-scale simulation of straining through 3D micro- Prodanovic, M., Landry, C., Tokan-Lawal, A., Eichhubl, P., 2016. Niobrara Formation
tomography images of porous media. J. Hydrol. 529 (3), 768–778. https://doi.org/ Fracture [Data Set]. Digital Rocks Portal University of Texas at Austinhttps://doi.
10.1016/j.jhydrol.2015.08.016. org/10.17612/P7SG6Z.
Mogensen, K., Stenby, E.H., 1998. A dynamic two-phase pore-scale model of imbibition. Prodanović, M., Lindquist, W.B., Seright, R.S., 2006. Porous structure and fluid parti-
Transp. Porous Media 32 (3), 299–327. https://doi.org/10.1023/A:1006578721129. tioning in polyethylene cores from 3D X-ray microtomographic imaging. J. Colloid
Mohanty, K.K., 1982. Multiphase Flow in Porous Media: II. Pore-Level Modeling. Society Interface Sci. 298 (1), 282–297. https://doi.org/10.1016/j.jcis.2005.11.053.
of Petroleum Engineershttps://doi.org/10.2118/11018-MS. Prodanović, M., Mehmani, A., Sheppard, A.P., 2015. Imaged-based multiscale network
Mosser, L., Dubrule, O., Blunt, M.J., 2017. Reconstruction of three-dimensional porous modelling of microporosity in carbonates. Geol. Soc. Lond. Spec. Publ. 406 (1),
media using generative adversarial neural networks. Phys. Rev. 96 (4). https://doi. 95–113. https://doi.org/10.1144/SP406.9.
org/10.1103/PhysRevE.96.043309. Purcell, W.R., 1949. Capillary pressures - their measurement using mercury and the
Mousavi, M.A., Bryant, S.L., 2012. Connectivity of pore space as a control on two-phase calculation of permeability therefrom. J. Pet. Technol. 1 (2), 39–48. https://doi.org/
flow properties of tight-gas sandstones. Transp. Porous Media 94 (2), 537–554. 10.2118/949039-G.
https://doi.org/10.1007/s11242-012-0017-x. Raeesi, B., Piri, M., 2009. The effects of wettability and trapping on relationships between
Mousavi, M., Prodanović, M., Jacobi, D., 2012. New classification of carbonate rocks for interfacial area, capillary pressure and saturation in porous media: a pore-scale
process-based pore-scale modeling. SPE J. 18 (2), 243–263. https://doi.org/10.2118/ network modeling approach. J. Hydrol. 376 (3–4), 337–352. https://doi.org/10.
163073-PA. 1016/j.jhydrol.2009.07.060.
Muljadi, B.P., 2015. Estaillades Carbonate [Data Set]. Digital Rocks Portal. https://doi. Raeini, A.Q., Bijeljic, B., Blunt, M.J., 2014. Numerical modelling of sub-pore scale events
org/10.17612/P73W2C. in two-phase flow through porous media. Transp. Porous Media 101 (2), 191–213.
Nguyen, V.H., Sheppard, A.P., Knackstedt, M.A., Val Pinczewski, W., 2006. The effect of https://doi.org/10.1007/s11242-013-0239-6.
displacement rate on imbibition relative permeability and residual saturation. J. Pet. Raeini, A.Q., Bijeljic, B., Blunt, M.J., 2015. Modelling capillary trapping using finite-
Sci. Eng. 52 (1), 54–70. https://doi.org/10.1016/j.petrol.2006.03.020. volume simulation of two-phase flow directly on micro-CT images. Adv. Water
Nogues, J.P., Fitts, J.P., Celia, M.A., Peters, C.A., 2013. Permeability evolution due to Resour. 83, 102–110. https://doi.org/10.1016/j.advwatres.2015.05.008.
dissolution and precipitation of carbonates using reactive transport modeling in pore Raeini, A.Q., Blunt, M.J., Bijeljic, B., 2014. Direct simulations of two-phase flow on micro-
networks. Water Resour. Res. 49 (9), 6006–6021. https://doi.org/10.1002/wrcr. CT images of porous media and upscaling of pore-scale forces. Adv. Water Resour. 74,
20486. 116–126. https://doi.org/10.1016/j.advwatres.2014.08.012.
Oak, M.J., 1990. Three-Phase Relative Permeability of Water-Wet Berea. Society of Ramstad, T., Idowu, N., Nardi, C., Øren, P.-E., 2012. Relative permeability calculations
Petroleum Engineershttps://doi.org/10.2118/20183-MS. from two-phase flow simulations directly on digital images of porous rocks. Transp.
Okabe, H., Blunt, M.J., 2004. Prediction of permeability for porous media reconstructed Porous Media 94 (2), 487–504. https://doi.org/10.1007/s11242-011-9877-8.
using multiple-point statistics. Phys. Rev. 70 (6), 066135. https://doi.org/10.1103/ Ramstad, T., Øren, P.-E., Bakke, S., 2010. Simulation of two-phase flow in reservoir rocks
PhysRevE.70.066135. using a lattice Boltzmann method. SPE J. 15 (04), 917–927. https://doi.org/10.2118/
Okabe, H., Blunt, M.J., 2005. Pore space reconstruction using multiple-point statistics. J. 124617-PA.
Pet. Sci. Eng. 46 (1–2), 121–137. https://doi.org/10.1016/j.petrol.2004.08.002. Renardy, Y., Renardy, M., 2002. PROST: a parabolic reconstruction of surface tension for
Ollion, J., Cochennec, J., Loll, F., Escudé, C., Boudier, T., 2013. TANGO: a generic tool for the volume-of-fluid method. J. Comput. Phys. 183 (2), 400–421. https://doi.org/10.
high-throughput 3D image analysis for studying nuclear organization. Bioinformatics 1006/jcph.2002.7190.
29 (14), 1840–1841. https://doi.org/10.1093/bioinformatics/btt276. Reynolds, C.A., Menke, H., Andrew, M., Blunt, M.J., Krevor, S., 2017. Dynamic fluid
Øren, P.-E., Bakke, S., 2002. Process based reconstruction of sandstones and prediction of connectivity during steady-state multiphase flow in a sandstone. Proc. Natl. Acad. Sci.
transport properties. Transp. Porous Media 46 (2), 311–343. https://doi.org/10. 114 (31), 8187–8192. https://doi.org/10.1073/pnas.1702834114.
1023/A:1015031122338. Rhodes, M., Bijeljic, B., Blunt, M.J., 2009. A rigorous pore-to-field-scale simulation
Øren, P.-E., Bakke, S., 2003. Reconstruction of Berea sandstone and pore-scale modelling method for single-phase flow based on continuous-time random walks. SPE J. 14 (1),
of wettability effects. J. Pet. Sci. Eng. 39 (3–4), 177–199. https://doi.org/10.1016/ 88–94. https://doi.org/10.2118/106434-PA.
S0920-4105(03)00062-7. Ruspini, L.C., Farokhpoor, R., Øren, P.E., 2017. Pore-scale modeling of capillary trapping
Øren, P.-E., Bakke, S., Arntzen, O.J., 1998. Extending predictive capabilities to network in water-wet porous media: a new cooperative pore-body filling model. Adv. Water
models. SPE J. 3 (04), 324–336. https://doi.org/10.2118/52052-PA. Resour. 108 (Suppl. C), 1–14. https://doi.org/10.1016/j.advwatres.2017.07.008.
Pak, T., Butler, I.B., Geiger, S., van Dijke, M.I.J., Jiang, Z., Surmas, R., 2016. Multiscale Ryazanov, A.V., Sorbie, K.S., van Dijke, M.I.J., 2014. Structure of residual oil as a function
pore-network representation of heterogeneous carbonate rocks: multiscale re- of wettability using pore-network modelling. Adv. Water Resour. 63, 11–21. https://
presentation OF heterogeneous carbonates. Water Resour. Res. 52 (7), 5433–5441. doi.org/10.1016/j.advwatres.2013.09.012.
https://doi.org/10.1002/2016WR018719. Ryazanov, A.V., van Dijke, M.I.J., Sorbie, K.S., 2010. Pore-network prediction of residual
Pak, T., Butler, I.B., Geiger, S., van Dijke, M.I.J., Sorbie, K.S., 2015. Droplet fragmenta- oil saturation based on oil layer drainage in mixed-wet systems. In: SPE Improved Oil
tion: 3D imaging of a previously unidentified pore-scale process during multiphase Recovery Symposium. Society of Petroleum Engineers, Tulsa, Oklahoma, USA.
flow in porous media. Proc. Natl. Acad. Sci. 112 (7), 1947–1952. https://doi.org/10. https://doi.org/10.2118/129919-MS.
1073/pnas.1420202112. Sahimi, M., Hughes, B.D., Scriven, L.E., Ted Davis, H., 1986. Dispersion in flow through
Pan, C., Hilpert, M., Miller, C.T., 2004. Lattice-Boltzmann simulation of two-phase flow in porous media—I. One-phase flow. Chem. Eng. Sci. 41 (8), 2103–2122. https://doi.
porous media. Water Resour. Res. 40 (1). https://doi.org/10.1029/2003WR002120. org/10.1016/0009-2509(86)87128-7.
n/a–n/a. Scanziani, A., Singh, K., Blunt, M., 2018. Water-Wet Three-phase Flow Micro-CT
Pan, C., Luo, L.-S., Miller, C.T., 2006. An evaluation of lattice Boltzmann schemes for Tomograms [Data Set]. Digital Rocks Portal. https://doi.org/10.17612/p7ht11.
porous medium flow simulation. Comput. Fluid 35 (8), 898–909. https://doi.org/10. Schaap, M.G., Porter, M.L., Christensen, B.S.B., Wildenschild, D., 2007. Comparison of
1016/j.compfluid.2005.03.008. pressure-saturation characteristics derived from computed tomography and lattice
Payatakes, A.C., Ng, K.M., Flumerfelt, R.W., 1980. Oil ganglion dynamics during im- Boltzmann simulations. Water Resour. Res. 43 (12), W12S06. https://doi.org/10.
miscible displacement: model formulation. AIChE J. 26 (3), 430–443. https://doi. 1029/2006WR005730.
org/10.1002/aic.690260315. Scheibe, T.D., Murphy, E.M., Chen, X., Rice, A.K., Carroll, K.C., Palmer, B.J., Tartakovsky,
Payatakes, A.C., Tien, C., Turian, R.M., 1973. A new model for granular porous media: A.M., Battiato, I., Wood, B.D., 2015. An analysis platform for multiscale hydro-
Part I. Model formulation. AIChE J. 19 (1), 58–67. https://doi.org/10.1002/aic. geologic modeling with emphasis on hybrid multiscale methods. Gr. Water 53 (1),
690190110. 38–56. https://doi.org/10.1111/gwat.12179.
Peters, E.J., 2012. Advanced Petrophysics. Live Oak Book Company, Austin, TX. Scholle, P.A., Ulmer-Scholle, D.S., 2003. A Color Guide to the Petrography of Carbonate
Piri, M., Blunt, M.J., 2004. Three-phase threshold capillary pressures in noncircular ca- Rocks: Grains, Textures, Porosity, Diagenesis, AAPG Memoir 77. American
pillary tubes with different wettabilities including contact angle hysteresis. Phys. Rev. Association of Petroleum Geologists.
70 (6), 061603. https://doi.org/10.1103/PhysRevE.70.061603. Serra, J., 1982. Image Analysis and Mathematical Morphology, vol. 1 Retrieved from.
Piri, M., Blunt, M.J., 2005. Three-dimensional mixed-wet random pore-scale network http://cds.cern.ch/record/235415.
modeling of two- and three-phase flow in porous media. I. Model description. Phys. Shams, M., Raeini, A.Q., Blunt, M.J., Bijeljic, B., 2018. A numerical model of two-phase
Rev. 71 (2), 026301. https://doi.org/10.1103/PhysRevE.71.026301. flow at the micro-scale using the volume-of-fluid method. J. Comput. Phys. 357,
Popinet, S., Zaleski, S., 1999. A front‐tracking algorithm for accurate representation of 159–182.
surface tension. Int. J. Numer. Methods Fluids 30 (6), 775–793. https://doi.org/10. Shan, X., Chen, H., 1993. Lattice Boltzmann model for simulating flows with multiple
1002/(SICI)1097-0363(19990730)30:6%3C775::AID-FLD864%3E3.0.CO;2-%23. phases and components. Phys. Rev. 47 (3), 1815–1819. https://doi.org/10.1103/
Porter, M.L., Schaap, M.G., Wildenschild, D., 2009. Lattice-Boltzmann simulations of the PhysRevE.47.1815.
capillary pressure-saturation-interfacial area relationship for porous media. Adv. Sheppard, A.P., Sok, R.M., Averdunk, H., Robins, V.B., Ghous, A., 2006. Analysis of rock
Water Resour. 32 (11), 1632–1640. https://doi.org/10.1016/j.advwatres.2009.08. microstructure using high resolution x-ray tomography. In: Proceedings of the
009. International Symposium of the Society of Core Analysts, pp. SCA2006–S2026
Prodanovic, M., 2010. LSMPQS Software. Retrieved from. http://users.ices.utexas.edu/ (Trondheim, Norway).
masha/lsmpqs/index.html. Shin, H., Lindquist, W.B., Sahagian, D.L., Song, S.-R., 2005. Analysis of the vesicular
Prodanovic, M., Bryant, S., 2009. Physics-driven interface modeling for drainage and structure of basalts. Comput. Geosci. 31 (4), 473–487. https://doi.org/10.1016/j.

30
A. Mehmani, et al. Marine and Petroleum Geology 114 (2020) 104141

cageo.2004.10.013. simulations using the level set method. J. Comput. Phys. 228 (4), 1139–1156.
Silin, D.B., Jin, G., Patzek, T.W., 2003. Robust determination of the pore space mor- https://doi.org/10.1016/j.jcp.2008.10.032.
phology in sedimentary rocks. In: SPE Annual Technical Conference and Exhibition. Thompson, K.E., 2002. Pore-scale modeling of fluid transport in disordered fibrous ma-
Society of Petroleum Engineers, Denver, Colorado. https://doi.org/10.2118/ terials. AIChE J. 48 (7), 1369–1389. https://doi.org/10.1002/aic.690480703.
84296-MS. Thompson, K.E., Willson, C.S., Zhang, W., 2006. Quantitative computer reconstruction of
Singh, K., Blunt, M.J., 2018. High Resolution X-Ray Micro-tomography Datasets for In- particulate materials from microtomography images. Powder Technol. 163 (3),
Situ Effective Contact Angle Analysis in Carbonate Rocks [Data Set]. Digital 169–182. https://doi.org/10.1016/j.powtec.2005.12.016.
Rockshttps://doi.org/10.17612/P7D95F. Tomutsa, L., Silin, D.B., Radmilovic, V., 2007. Analysis of chalk petrophysical properties
Singh, K., Menke, H., Andrew, M., Lin, Q., Rau, C., Blunt, M.J., Bijeljic, B., 2017. by means of submicron-scale pore imaging and modeling. SPE Reserv. Eval. Eng. 10
Dynamics of snap-off and pore-filling events during two-phase fluid flow in perme- (3), 285–293. https://doi.org/10.2118/99558-PA.
able media. Sci. Rep. 7 (1), 5192. https://doi.org/10.1038/s41598-017-05204-4. Torskaya, T., Shabro, V., Torres-Verdín, C., Salazar-Tio, R., Revil, A., 2014. Grain shape
Singh, M., Mohanty, K.K., 2003. Dynamic modeling of drainage through three-dimen- effects on permeability, formation factor, and capillary pressure from pore-scale
sional porous materials. Chem. Eng. Sci. 58 (1), 1–18. https://doi.org/10.1016/ modeling. Transp. Porous Media 102 (1), 71–90. https://doi.org/10.1007/s11242-
S0009-2509(02)00438-4. 013-0262-7.
Skalinski, M., Kenter, J.A.M., 2015. Carbonate petrophysical rock typing: integrating Valvatne, P.H., Blunt, M.J., 2004. Predictive pore-scale modeling of two-phase flow in
geological attributes and petrophysical properties while linking with dynamic be- mixed wet media. Water Resour. Res. 40 (7), W07406. https://doi.org/10.1029/
haviour. Geol. Soc. Lond. Spec. Publ. 406 (1), 229–259. https://doi.org/10.1144/ 2003WR002627.
SP406.6. Van de Casteele, E., Van Dyck, D., Sijbers, J., Raman, E., 2002. An energy-based beam
Sok, R.M., Knackstedt, M.A., Sheppard, A.P., Pinczewski, W.V., Lindquist, W.B., hardening model in tomography. Phys. Med. Biol. 47 (23), 4181. https://doi.org/10.
Venkatarangan, A., Paterson, L., 2002. Direct and stochastic generation of network 1088/0031-9155/47/23/305.
models from tomographic images; effect of topology on residual saturations. Transp. van der Land, C., Wood, R., Wu, K., van Dijke, M.I.J., Jiang, Z., Corbett, P.W.M., Couples,
Porous Media 46 (2–3), 345–371. https://doi.org/10.1023/A:1015034924371. G., 2013. Modelling the permeability evolution of carbonate rocks. Mar. Pet. Geol.
Sok, R.M., Knackstedt, M.A., Varslot, T., Ghous, A., Latham, S., Sheppard, A.P., 2010. 48, 1–7. https://doi.org/10.1016/j.marpetgeo.2013.07.006.
Pore scale characterization of carbonates at multiple scales: integration of Micro-CT, van der Marck, S.C., Matsuura, T., Glas, J., 1997. Viscous and capillary pressures during
BSEM, and FIBSEM. Petrophysics 51 (06) SPWLA-2010-v51n6a1. drainage: network simulations and experiments. Phys. Rev. 56 (5), 5675–5687.
Sorbie, K.S., Clifford, P.J., 1991. The inclusion of molecular diffusion effects in the net- https://doi.org/10.1103/PhysRevE.56.5675.
work modelling of hydrodynamic dispersion in porous media. Chem. Eng. Sci. 46 Verma, R., Icardi, M., Prodanović, M., 2018. Effect of wettability on two-phase quasi-
(10), 2525–2542. https://doi.org/10.1016/0009-2509(91)80046-2. static displacement: validation of two pore scale modeling approaches. J. Contam.
Sorbie, K.S., Skauge, A., 2012. Can network modeling predict two-phase flow functions? Hydrol. https://doi.org/10.1016/j.jconhyd.2018.01.002.
Petrophysics 53 (6), 401–409. Victor, R.A., 2017. Multiscale, Image-Based Interpretation of Well Logs Acquired in a
Soulaine, C., Gjetvaj, F., Garing, C., Roman, S., Russian, A., Gouze, P., Tchelepi, H.A., Complex, Deep-Water Carbonate Reservoir (Ph.D. Thesis). The University of Texas at
2016. The impact of sub-resolution porosity of X-ray microtomography images on the Austin, Austin, TX, USA Retrieved from. https://www.pge.utexas.edu/images/pdfs/
permeability. Transp. Porous Media 113 (1), 227–243. https://doi.org/10.1007/ theses17/victordis.pdf.
s11242-016-0690-2. Victor, R.A., Prodanovic, M., 2017. Dual-Energy Medical CT in Carbonate Rocks [Data
Soulaine, C., Creux, P., Tchelepi, H.A., 2019. Micro-continuum framework for pore-scale Set]. Digital Rocks Portal. https://doi.org/10.17612/P74368.
multiphase fluid transport in shale formations. Transp. Porous Media 127 (1), Walsh, S.D., Burwinkle, H., Saar, M.O., 2009. A new partial-bounceback lattice-
85–112. https://doi.org/10.1007/s11242-018-1181-4. Boltzmann method for fluid flow through heterogeneous media. Comput. Geosci. 35
Spiteri, E.J., Juanes, R., Blunt, M.J., Orr, F.M., 2008. A new model of trapping and re- (6), 1186–1193.
lative permeability hysteresis for all wettability characteristics. SPE J. 13 (03), Whitaker, S., 1999. The Method of Volume Averaging. Retrieved from. http://public.
277–288. https://doi.org/10.2118/96448-PA. eblib.com/choice/publicfullrecord.aspx?p=3106854.
Sukop, M.C., Huang, H., Lin, C.L., Deo, M.D., Oh, K., Miller, J.D., 2008. Distribution of Wildenschild, D., Sheppard, A.P., 2013. X-ray imaging and analysis techniques for
multiphase fluids in porous media: comparison between lattice Boltzmann modeling quantifying pore-scale structure and processes in subsurface porous medium systems.
and micro-x-ray tomography. Phys. Rev. 77 (2), 026710. https://doi.org/10.1103/ Adv. Water Resour. 51, 217–246. https://doi.org/10.1016/j.advwatres.2012.07.018.
PhysRevE.77.026710. Wilkinson, D., 1986. Percolation effects in immiscible displacement. Phys. Rev. A 34 (2),
Sussman, M., 2005. A parallelized, adaptive algorithm for multiphase flows in general 1380–1391. https://doi.org/10.1103/PhysRevA.34.1380.
geometries. Comput. Struct. 83 (6–7), 435–444. https://doi.org/10.1016/j. Wilkinson, D., Willemsen, J.F., 1983. Invasion percolation: a new form of percolation
compstruc.2004.06.006. theory. J. Phys. A Math. Gen. 16 (14), 3365–3376. https://doi.org/10.1088/0305-
Sussman, M., Almgren, A.S., Bell, J.B., Colella, P., Howell, L.H., Welcome, M.L., 1999. An 4470/16/14/028.
adaptive level set approach for incompressible two-phase flows. J. Comput. Phys. 148 Wirth, R., 2009. Focused Ion Beam (FIB) combined with SEM and TEM: advanced ana-
(1), 81–124. https://doi.org/10.1006/jcph.1998.6106. lytical tools for studies of chemical composition, microstructure and crystal structure
Sussman, M., Puckett, E.G., 2000. A coupled level set and volume-of-fluid method for in geomaterials on a nanometre scale. Chem. Geol. 261 (3), 217–229. https://doi.
computing 3D and axisymmetric incompressible two-phase flows. J. Comput. Phys. org/10.1016/j.chemgeo.2008.05.019.
162 (2), 301–337. https://doi.org/10.1006/jcph.2000.6537. Wu, K., Dijke, M.I.J.V., Couples, G.D., Jiang, Z., Ma, J., Sorbie, K.S., Crawford, J., Young,
Sussman, M., Smereka, P., Osher, S., 1994. A level set approach for computing solutions I., Zhang, X., 2006. 3D stochastic modelling of heterogeneous porous media – ap-
to incompressible two-phase flow. J. Comput. Phys. 114 (1), 146–159. https://doi. plications to reservoir rocks. Transp. Porous Media 65 (3), 443–467. https://doi.org/
org/10.1006/jcph.1994.1155. 10.1007/s11242-006-0006-z.
Swift, M.R., Orlandini, E., Osborn, W.R., Yeomans, J.M., 1996. Lattice Boltzmann simu- Xu, R., Prodanović, M., 2018. Effect of pore geometry on nitrogen sorption isotherms
lations of liquid-gas and binary fluid systems. Phys. Rev. 54 (5), 5041–5052. https:// interpretation: a pore network modeling study. Fuel 225, 243–255. https://doi.org/
doi.org/10.1103/PhysRevE.54.5041. 10.1016/j.fuel.2018.03.143.
Tahmasebi, P., Sahimi, M., 2013. Cross-correlation function for accurate reconstruction of Yoon, H., Dewers, T.A., 2013. Nanopore structures, statistically representative elementary
heterogeneous media. Phys. Rev. Lett. 110 (7), 078002. https://doi.org/10.1103/ volumes, and transport properties of chalk. Geophys. Res. Lett. 40 (16), 4294–4298.
PhysRevLett.110.078002. https://doi.org/10.1002/grl.50803.
Tartakovsky, A.M., Meakin, P., 2006. Pore scale modeling of immiscible and miscible Zolfaghari, A., Piri, M., 2017. Pore-scale network modeling of three-phase flow based on
fluid flows using smoothed particle hydrodynamics. Adv. Water Resour. 29 (10), thermodynamically consistent threshold capillary pressures. I. Cusp formation and
1464–1478. https://doi.org/10.1016/j.advwatres.2005.11.014. collapse. Transp. Porous Media 116 (3), 1093–1137. https://doi.org/10.1007/
Tartakovsky, A.M., Meakin, P., Scheibe, T.D., Eichler West, R.M., 2007. Simulations of s11242-016-0814-8.
reactive transport and precipitation with smoothed particle hydrodynamics. J. Zhao, B., MacMinn, C.W., Juanes, R., 2016. Wettability control on multiphase flow in
Comput. Phys. 222 (2), 654–672. https://doi.org/10.1016/j.jcp.2006.08.013. patterned microfluidics. Proc. Natl. Acad. Sci. 113 (37), 10251–10256. https://doi.
Thane, C.G., 2006. Geometry and Topology of Model Sediments and Their Influence on org/10.1073/pnas.1603387113.
Sediment Properties (Thesis). Retrieved from. https://repositories.lib.utexas.edu/ Zhao, B., MacMinn, C.W., Primkulov, B.K., Chen, Y., Valocchi, A.J., Zhao, J., Kang, Q.,
handle/2152/35281. Bruning, K., McClure, J.E., Miller, C.T., Fakhari, A., 2019. Comprehensive compar-
Thömmes, G., Becker, J., Junk, M., Vaikuntam, A.K., Kehrwald, D., Klar, A., Steiner, K., ison of pore-scale models for multiphase flow in porous media. In: Proceedings of the
Wiegmann, A., 2009. A lattice Boltzmann method for immiscible multiphase flow National Academy of Sciences, https://doi.org/10.1073/pnas.1901619116.

31

S-ar putea să vă placă și