ArticlePDF Available

Point-cloud deep learning of porous media for permeability prediction

Authors:

Abstract and Figures

We propose a novel deep learning framework for predicting the permeability of porous media from their digital images. Unlike convolutional neural networks, instead of feeding the whole image volume as inputs to the network, we model the boundary between solid matrix and pore spaces as point clouds and feed them as inputs to a neural network based on the PointNet architecture. This approach overcomes the challenge of memory restriction of graphics processing units and its consequences on the choice of batch size and convergence. Compared to convolutional neural networks, the proposed deep learning methodology provides freedom to select larger batch sizes due to reducing significantly the size of network inputs. Specifically, we use the classification branch of PointNet and adjust it for a regression task. As a test case, two and three dimensional synthetic digital rock images are considered. We investigate the effect of different components of our neural network on its performance. We compare our deep learning strategy with a convolutional neural network from various perspectives, specifically for maximum possible batch size. We inspect the generalizability of our network by predicting the permeability of real-world rock samples as well as synthetic digital rocks that are statistically different from the samples used during training. The network predicts the permeability of digital rocks a few thousand times faster than a lattice Boltzmann solver with a high level of prediction accuracy.
Content may be subject to copyright.
Phys. Fluids 33, 097109 (2021); https://doi.org/10.1063/5.0063904 33, 097109
© 2021 Author(s).
Point-cloud deep learning of porous media
for permeability prediction
Cite as: Phys. Fluids 33, 097109 (2021); https://doi.org/10.1063/5.0063904
Submitted: 18 July 2021 . Accepted: 26 August 2021 . Published Online: 28 September 2021
Ali Kashefi and Tapan Mukerji
COLLECTIONS
This paper was selected as Featured
This paper was selected as Scilight
Point-cloud deep learning of porous media
for permeability prediction
Cite as: Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904
Submitted: 18 July 2021 .Accepted: 26 August 2021 .
Published Online: 28 September 2021
Ali Kashefi
1,a)
and Tapan Mukerji
2,b)
AFFILIATIONS
1
Department of Civil and Environmental Engineering, Stanford University, Stanford, California 94305, USA
2
Department of Energy Resources Engineering, Stanford University, Stanford, California 94305, USA
a)
Author to whom correspondence should be addressed: kashefi@stanford.edu
b)
Electronic mail: mukerji@stanford.edu
ABSTRACT
We propose a novel deep learning framework for predicting the permeability of porous media from their digital images. Unlike
convolutional neural networks, instead of feeding the whole image volume as inputs to the network, we model the boundary between solid
matrix and pore spaces as point clouds and feed them as inputs to a neural network based on the PointNet architecture. This approach
overcomes the challenge of memory restriction of graphics processing units and its consequences on the choice of batch size and
convergence. Compared to convolutional neural networks, the proposed deep learning methodology provides freedom to select larger batch
sizes due to reducing significantly the size of network inputs. Specifically, we use the classification branch of PointNet and adjust it for a
regression task. As a test case, two and three dimensional synthetic digital rock images are considered. We investigate the effect of different
components of our neural network on its performance. We compare our deep learning strategy with a convolutional neural network from
various perspectives, specifically for maximum possible batch size. We inspect the generalizability of our network by predicting the
permeability of real-world rock samples as well as synthetic digital rocks that are statistically different from the samples used during training.
The network predicts the permeability of digital rocks a few thousand times faster than a lattice Boltzmann solver with a high level of predic-
tion accuracy.
Published under an exclusive license by AIP Publishing. https://doi.org/10.1063/5.0063904
I. INTRODUCTION AND MOTIVATION
The importance of study of porous media in a wide range of sci-
entific and industrial fields such as digital rock physics,
1,2
membrane
systems,
3
geological carbon storage,
4
and medicine
5
in the past decades
has led to a growth in collection of pore-scale image data. Along with
pore-scale imaging, there has been a growth in the use of numerical
computation to assess physical and transport properties of porous
media based on the image data. Such a revolution in the age of data
has motivated the use of machine learning schemes as a data-driven
strategy to accelerate the computations for understanding the physical
properties of porous media. Among different machine learning techni-
ques, deep learning has been widely used in various applications for
the study of porous media. A few specific applications are rock image
segmentation
6–8
and predicting physical properties and geometrical
features such as permeability,
9–16
porosity,
9,17–20
effective diffusivity,
21
wave propagation velocities,
22
and fluid flow fields.
23,24
It is worth not-
ing that arguments and ideas proposed in this article are general and
usable for any desired porous media such as biological tissues and
ceramics; however, we restrict ourselves to the applications of rocks in
subsurface aquifers and petroleum reservoirs. Specifically, our focus in
the present article is on deep learning frameworks for predicting per-
meability from digital rock images.
Convolutional neural networks (CNNs) have been used exten-
sively to predict the permeability of digital rock images (see, e.g., Refs.
11,12,and14). In this setup, CNNs are trained on a set of labeled data
to learn a mapping from two (2D) or three dimensional (3D) digital
rock images to rock permeability. Generally speaking, a common chal-
lenge in using CNNs is the Graphics Processing Unit (GPU) memory
required for training CNNs.
25
This challenge is magnified in large,
deep, and three dimensional CNNs.
25
Limitation on the memory of
available GPU memory might lead to a restriction on the “batch size”
(see, e.g., Ref. 26 for the technical definition of “batch size”). Contrary
to the technique of stochastic gradient descent, the mini batch gradient
descent method accelerates the training procedure mainly by vectori-
zation. However, the associated batch size affects the performance of
deep neural networks.
27–30
Hence, it is vital to have freedom to choose
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-1
Published under an exclusive license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf
the optimal batch size. To overcome the above-mentioned challenge,
we propose a new machine learning architecture, which is based on
the deep learning of point cloud data. Next, we explain the key idea of
our methodology.
Mathematically, the permeability of a porous medium is a func-
tion of the velocity fields in the pore-space of the medium. The solu-
tion of the governing equations of fluid flow in porous media (e.g.,
reservoir rocks) is a function of the geometry of the grain–pore bound-
ary and the boundary conditions. Thus, if the geometry of the grain–
pore boundary can be used as an input representation to the neural
network, we do not need either the volume of the grain spaces or pore
spaces anymore. To reach this goal, for a given porous medium, we
only take the grain–pore boundary and represent it as a set of points,
constructing a point cloud (see Fig. 1). Points on the surface (in three
dimensional geometries) or on the edge (in two dimensional geome-
tries) of this cloud represent the geometry of the grain–pore bound-
aries. Representing digital rocks as sparse point clouds instead of full
two or three dimensional image pixels or voxels dramatically dimin-
ishes the size of memory required to be allocated on GPUs.
Additionally, it provides users with more freedom to select the batch
size.
Since we represent the boundary of the pore space as a point
cloud, a point-cloud-based deep learning framework is required. From
a computer science point of view, several architectures are available for
this purpose (see, e.g., Refs. 31–33). Among these options, PointNet
32
has been widely used for deep learning of point cloud data for classifi-
cation and segmentation of two and three dimensional objects in com-
puter vision and computer graphics (see, e.g., Refs. 34 and 35). Qi
et al.
32
first introduced PointNet in 2017, and the network has quickly
become popular for both industrial and academic applications such as
object detection,
35,36
shape reconstruction,
37
camera pose estimation,
37
and physical simulation.
38,39
To the best of our knowledge, PointNet
32
has been already used
twice for applications outside of the pure computer science areas. First,
the performance of PointNet
32
for predicting the velocity and pressure
fields of incompressible flows on irregular geometries has been exam-
ined by Kashefi et al.
38
Kashefi et al.
38
adjusted the PointNet architec-
ture
32
to predict the flow fields around a cylinder with various shapes
for its cross section and obtained an excellent to reasonable level of
accuracy. Additionally, they
38
demonstrated the generalizability of
their proposed neural network by predicting the velocity and pressure
fields on unseen category data such as multiple objects and airfoils (see
Figs. 13–19 of Ref. 38). Second, DeFever et al.
40
employed PointNet
32
to identify local structures in molecular simulations. These suc-
cesses
38,40
motivate us to utilize the PointNet
32
architecture and mod-
ify it for our own application. To accomplish this task, we use the
classification component of PointNet
32
and replace its cross-entropy
cost function by the mean squared error to establish an end-to-end
mapping from a point cloud (as input) to the corresponding perme-
ability (as output) framed as a regression problem. It is important to
mention that we utilize PointNet
32
for the first time for a regression
problem. Although our focus in this article is on permeability predic-
tion in porous media, our approach can be potentially used for any
other machine learning problems, where the output of interest is a real
number that is a function of the geometry of spatial domains. Further
details of our neural network are described in Sec. II C 1.
We assessthe prediction performance of the network, its sensitiv-
ity to different parameters and activation functions, and its computa-
tional efficiency in several ways. First, to evaluate prediction
performance, we report the coefficient of determination as well as the
maximum and minimum relative errors of the predicted permeability
with reference to the permeability calculated from a numerical solver
for a set of two and three dimensional porous medium geometries.
Second, to assess the sensitivity to different parameters, we discuss the
number of points in point clouds as a hyperparameter of the neural
network proposed in this article. We evaluate the effect of input and
feature transform blocks in PointNet
32
on the performance of the deep
learning framework. Additionally, we explore the influence of different
activation functions and different sizes of latent global feature on the
accuracy of the predicted permeability, and test the neural network
generalizability. Finally, we compute the speed-up factor obtained by
the proposed neural network compared to a conventional numerical
solver for flow simulation in pore spaces as well as compare the perfor-
mance of the point-cloud neural network with a regular CNN.
The rest of this paper is structured as follows. We describe the
governing equations of fluid flows in porous media and techniques for
FIG. 1. Schematic illustration of the algorithm for constructing point clouds: (a) voxel representation, 0 (red) and 1 (blue) indicate, respectively, pore and grain spaces, (b) pore–
grain space boundary identification, (c) point cloud representation; the green boundary is specified by a set of points.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-2
Published under an exclusive license by AIP Publishing
permeability computations using numerical solvers in Sec. II A.Data
generation for deep learning is explained in Sec. II B. We illustrate and
compare the architecture of our neural network with a CNN in Sec.
II C. Network training is illustrated in Sec. II D. An analysis of the net-
work performance for two dimensional geometries is provided in Sec.
III A. Prediction of the permeability in three dimensional porous
media is investigated in Sec. III B. Alternative approaches for perme-
ability prediction and potentials of our neural network in this regard
are discussed in Sec. III C.SectionIV summarizes and concludes the
study.
II. PROBLEM FORMULATION AND METHODOLOGY
A. Permeability computation in porous media
To compute the permeability of a porous medium, first we obtain
thevelocityfieldoffluidflowintheporespaceofthemedium.The
continuity and Navier–Stokes equations govern the dynamics of
single-phase incompressible flow within the pores of a porous medium
and are written as follows:
@q
@tþrðquÞ¼0inV;(1)
@ðquÞ
@tþrðquuÞþrplDu¼fin V;(2)
where uand pindicate, respectively, the velocity vector and absolute
pressure in the space of V, the fluid-filled pore space. The fluid density
and dynamic viscosity are shown by qand l, respectively. The vector
of external body force is indicated by f. We consider the porous
medium domains to be squares (in two dimensional spaces) or cubes
(in three dimensional spaces) with length Lalong each principal axis,
porosity of /, and spatial correlation length of l
c
. A pressure gradient
in the xdirection (dp/dx) is applied to stimulate the flow in the
medium. No flow boundary condition is enforced at the top and bot-
tom of the medium on the yzplanes. Periodic boundary conditions
are applied at the inflow and outflow velocity boundaries parallel to
the pressure gradient direction. A numerical solver based on the
Lattice Boltzmann Method (LBM) is used to obtain the steady state
solution to the governing equations. More details of the analysis are
discussed in Ref. 41. After calculating the flow velocity in the pore
space of the porous medium, the permeability in xdirection (k)is
obtained from Darcy’s law,
42
k¼ l
U
dp=dx ;(3)
where
Uis the mean velocity in the entire porous medium including
grain spaces. Note that Eq. (3) is only valid for low Reynolds numbers
(see, e.g., Ref. 43).
B. Data generation
To have a robust control on training data and investigate the
effect of different geometrical parameters such as porosity (/)andspa-
tial correlation length (l
c
) in porous media, we synthetically generate
our data set such that it represents a range of heterogeneity of reservoir
rocks. Similar approaches have been taken by Wu et al.
12
and Da
Wang et al.
23
To generate a synthetic binary (pore–grain) medium
with a targeted porosity (/) and spatial correlation (l
c
), a straightfor-
ward algorithm of truncated Gaussian simulation
44,45
is used by
truncating spatially correlated Gaussian random fields created by a
moving average filter applied to random uncorrelated Gaussian noise.
The algorithm is implemented as follows. First, we consider two and
three dimensional arrays, respectively, with the size of n
2
and n
3
.Next,
uncorrelated random variables with the standard normal distribution
are assigned to the array elements. Afterwards, Gaussian kernels with
different kernel sizes are applied as a filter introducing spatial correla-
tion. In the next stage, the numerical values of the arrays are normal-
ized in the range of [0, 1], and thresholded to give binary arrays with
desired ranges of porosity (i.e., see Figs. 2 and 3). Arrays with no corre-
lated fields are discarded. In this work, we set L¼ndx,wheredxis
the size of each pixel side and equal to 0.003 m. Concerning two
dimensional porous media, we set n¼128 and synthetically generate
data with three representative spatial correlation lengths (kernel of the
Gaussian filter) of 9, 17, and 33 pixels while considering the porosity
(/) in the range of [0.125, 0.25). We use 2600 data samples with a spa-
tial correlation length (l
c
) of 9 for training, validation, and test pur-
poses, while data with spatial correlation lengths (l
c
) of 17 and 33 are
used for the investigation of neural network generalizability. The mean
(and standard deviations) for the porosity and permeability of the
training and test set of the two dimensional porous media are as fol-
lows: training set porosity, 0.181 (0.03); test set porosity, 0.185 (0.04);
training set permeability, 121.62 mD (18.42 mD); test set permeability,
128.52 mD (20.95 mD). Concerning three dimensional porous media,
we generate data with n¼64 and spatial correlation length (kernel of
the Gaussian filter) of 17 pixels, while the porosity (/) in the range of
[0.125, 0.20) is selected. A total of 2175 samples are generated for use
in training, validation, and testing. The mean (and standard devia-
tions) for the porosity and permeability of the training and test set of
the three dimensional porous media are as follows: training set poros-
ity, 0.146 (0.04); test set porosity, 0.151 (0.05); training set permeabil-
ity, 67.12 mD (58.03 mD); test set permeability, 69.83 mD (61.12 mD).
We consider a real 3D CT-scan image of a rock sample to carry out
the generalizability level of our neural network. A set of Python codes
and batch files automates the process of generating synthetic data. The
LBM solver is run on all of the generated synthetic porous media to
get the corresponding permeabilities, thus creating a labeled dataset.
The next step is to define the neural network domain (V
NN
).
Indicating the grain-pore boundary by @V, then mathematically,
VNN @V.Infact,V
NN
must represent the geometry of the grain–-
pore boundary. Note that by keeping the physical properties of the
fluid (i.e., viscosity and density) and the boundary conditions fixed
over all the generated data, the solution of the governing equations is
only a function of the geometry of the boundary of the pore space @V,
and consequently V
NN
.V
NN
contains Npoints. The challenge is that
the number of pixels located on @Vvaries from one data sample to
another. Thus, Nis a hyperparameter in our deep learning framework.
We discuss the effect of the choice of Non our neural network perfor-
mance in Sec. III A.Figures 2 and 3depict, respectively, digital porous
medium images and their resulting point clouds (V
NN
)fortwoand
three dimensions. Note that our deep learning methodology is not lim-
ited to constructing V
NN
from digital images. One may use scattered
data obtained on unstructured finite element or finite volume grids to
establish V
NN
.
To accelerate the convergence of our neural network training and
equalize the contribution of each input component to the training of
neural network parameters (e.g., weights and bias), the input and
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-3
Published under an exclusive license by AIP Publishing
FIG. 3. Three dimensional digital porous medium images and their corresponding point-cloud representations; digital images and point clouds are used to train CNN and the
point-cloud neural network, respectively.
FIG. 2. Two dimensional digital porous medium images and their corresponding point-cloud representations; digital images and point clouds are used to train CNN and the
point-cloud neural network, respectively.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-4
Published under an exclusive license by AIP Publishing
output data are scaled in the range of [0, 1] using the maximum and
minimum values of each set of k,x,y,andz.Weindicatethescaledset
by k0;x0;y0,andz0. As an example, k0is computed as follows:
k0¼kminðkÞ
maxðkÞminðkÞ:(4)
x0;y0,andz0are computed similarly. Obviously, k0;x0;y0,andz0are
dimensionless.
C. Neural network architectures
1. Point-cloud neural network
Our neural network is mainly based on the PointNet
32
architec-
ture. In this subsection, we briefly describe the point-cloud neural net-
work. One may refer to Ref. 32 for further explanations. In this
subsection, the vectors and matrices of machine learning components
are shown by bold letters but not italic. This is to distinguish the
machine learning vectors and matrices from the physics-based ones.
The two main components are Multilayer Perceptron (MLP) and
Fully Connected (FC) layer. An MLP is constructed by several sequen-
tial FC layers. We use notation in the form of (A
1
,A
2
)toshowan
MLP with two layers, where A
1
and A
2
are, respectively, the size of the
first and second layer. Notations in the form of ðA1;A2;A3Þare simi-
larly defined. In the current study, the point-cloud neural network is
restricted to MLPs with two and three layers. We parameterize each
FC layer by a weight matrix Wand a bias vector b.ThesizeofanFC
layer indicates the number of rows in the corresponding matrix W.
Mathematically, a recursive function connects the input of ith FC layer
aito the output of i1th FC layer ai1such that
ai¼rðWiai1þbiÞ;(5)
where ris a nonlinear activation function. The activation function is
applied elementwise to each vector component.
Consider two sets Xand Y, the network inputs and the desired
target, respectively, with fxi2RdgN
i¼1and Y¼fyi2Rgnp
i¼1,
where dcorresponds to the spatial dimension (2 or 3) and n
p
is the
number of desired physical or geometrical quantities of interest as the
targets of the network. When predicting permeability alone, n
p
is 1.
We wish to design a neural network to map Xto Yby an operator f
such that fðXÞ. Two fundamental concepts need to be consid-
ered in the design. First, the output set Yis a function of the geometri-
cal features of the input set X. Thus, the operator fmust be able to
capture the geometrical features. Second, since the input set Xessen-
tially represents an unstructured and unordered point cloud, the oper-
ator fmust be invariant with respect to the order of input points xiof
the set X.ThePointNet
32
architecture provides these two critical fea-
tures. Hence, we approximate the operator fby a PointNet-based neu-
ral network that learns the mapping from Xto Ythrough a set of
labeled data described in Sec. II B.
The structure of the point-cloud neural network is depicted in
Fig. 4. As can be seen in Fig. 4, the network has two main branches:
one before and another after the global feature. The first branch enco-
des the geometrical feature of the input set Xin a latent global feature
with a vector of size 1024. The second branch decodes the latent global
feature to predict the permeability. Two Transform Nets (T-Nets) exist
in the first branch. The first T-Net transforms the input set Xinto an
implicit canonical space, while the second T-Net is used for an affine
transformation for alignment in the input set X. From a machine
learning perspective, T-Nets can be viewed as mini PointNets that
consist of an MLP component (64, 128, 1024) followed by a max pool-
ing operator to extract the underlying features. The feature can then
be decoded by two MLP components each with two layers (512, 256).
One may refer to Ref. 32 for further descriptions of T-Nets. In addition
to T-Nets, two MLP components contribute to the construction of the
first branch: the first (64, 64) and the second (64, 128, 1024).
Mathematically, PointNet
32
encodes the geometrical features of the
point set such that the latent code is independent of ordering over the
set of points. In other words, to aggregate information over the input
set X, a permutation invariant function such as maximum, minimum,
average, and summation is necessary. PointNet
32
uses the “max” func-
tion to handle it. We represent all the mathematical operations carried
outontheinputsetXjust before the max pooling operator by a func-
tion h. Thus, the latent global feature is established on the input set X
by a function gsuch that
FIG. 4. Structure of the proposed point-cloud neural network based on PointNet;
32
the network input is the point cloud representing the boundary line or boundary surface of
grain–pore spaces, respectively, for two and three dimensional porous media.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-5
Published under an exclusive license by AIP Publishing
gðXÞ  maxðhðx1Þ;;hðxNÞÞ:(6)
As can be observed in Fig. 4, an MLP component with three layers
(512, 256, 1) in the second branch is used to predict the permeability.
Note that all the MLP components in the first branch have shared
weights,whilethisisnotthecasefortheMLPcomponentinthe
second branch. This is another key feature of PointNet
32
to handle
unordered points in the set X, and this is why we use the single func-
tion hfor all the input points xiin Eq. (6). In fact, it does not matter
how the input set Xis constructed for feeding it to our neural network
as PointNet
32
treats all xiin a same manner due to the shared weights
of MLPs in the first branch. After each FC layer, a batch normaliza-
tion
46
operator is used. The activation function used for all the layers
is the Rectified Linear Unit (ReLU) defined as
rðcÞ¼maxð0;cÞ;(7)
except for the last layer where we employ a sigmoid function expressed
as
rðcÞ¼ 1
1þec:(8)
To close this subsection, we address a few points. First, we set
d¼3 for the permeability prediction in two dimensional pore spaces
by assigning zero values to the third axis. Alternatively, one may set
d¼2 for two dimensional porous media and adapt the size of network
layers accordingly. Second, we restrict our current study to the predic-
tion of the permeability (i.e., n
p
¼1); however, one may adjust n
p
for
the prediction of other quantities of interest such as porosity, average
pore size, and specific surface area (see, e.g., Ref. 17).
2. Convolutional neural networks
We briefly explain the architecture of the CNNs designed for pre-
dicting permeability from two and three dimensional digital rock
images. We skip describing the technical details used in this subsec-
tion, and we encourage potential audiences with interest in use of
CNNs for permeability prediction to read Sec. 2.3 of Ref. 11. Similar to
PointNet,
32
we need an encoder to extract the image features and a
decoder, which maps the learned features to the corresponding perme-
ability. We employ the encoder structure of DCGAN,
47
which is a
highly cited and successful unsupervised generative adversarial net-
work. Accordingly, ReLU activation function is used for all layers, and
no pooling layer is utilized. The number of filters starts with 16 and
doubles at each convolution layer, sequentially. All convolution layers
are set with no padding, a stride size of 2, and kernel size of (2, 2) and
(2, 2, 2), respectively, for the two and three dimensional CNNs, except
in the last layer of the three dimensional CNN, which has a kernel size
of (1, 1, 1). This is to enforce a latent global feature with the size of
1024 (see further discussions in Sec. II C 3). We use the PointNet
32
decoder for both the two dimensional (2D-CNN) and three dimen-
sional CNN (3D-CNN). As an example, Fig. 5 depicts the architecture
of 2D-CNN used in this study.
3. Comparison between PointNet and CNNs
A fair comparison between PointNet
32
and a CNN is not
straightforward. First, each of them is based on different underlying
mathematical and computational theories. Second, PointNet
32
has a
unique structure, whereas we can find many neural networks, which
are based on the convolution operation (see, e.g., Refs. 48 and 49)and
fall in the category of CNNs. Moreover, neural networks with the con-
volution operation are usually combined with other functions such as
max-pooling (see, e.g., Ref. 50), upsampling (see, e.g., Ref. 51), and
skip connection (see, e.g., Ref. 52). Thus, a variety of CNNs with differ-
ent performance can be implemented. With these in mind, we have
enforced two conditions for designing the CNN introduced in Sec.
II C 2 to make it similar to the PointNet
32
architecture as much as pos-
sible. First, the size of latent global feature of both 2D-CNN and
PointNet
32
is equal to 1024. Second, both networks use the same
decoder. This makes it easier to have a consistent comparison.
D. Training
The first step in the training process is to select an appropriate
cost function (or loss function). The mean squared error function has
been widely used in the area of deep learning of computational
mechanics (see, e.g., Ref. 53)aswellasinporousmediaapplications
(see, e.g., Refs. 11–13,23,and24). In the current study, we utilize this
function defined as
1
MX
M
i¼1
ðk0
i~
k0
iÞ2;(9)
where Mis the number of data in our training set. We label the perme-
ability (k0) obtained by the LBM solver as the “ground truth,” while we
denote the predicted permeability by ~
k0. After training, we rescale the
predicted permeability (~
k0) back into to the physical domain (~
k0)fora
FIG. 5. Structure of the 2D-CNN used for learning two dimensional porous media.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-6
Published under an exclusive license by AIP Publishing
post-processing analysis. We use the Adam
54
optimizer with hyper-
parameters of b1¼0:9;b2¼0:999, and ^
¼106.Tounderstand
the mathematical definition of b
1
,b
2
,and^
,onemayrefertoRef.54.
For two dimensional cases, our generated data are categorized into
three sets for training (2300 data), validation (150 data), and test (150
data) through a random selection process. Similarly for three dimen-
sional cases, we have three sets of training (1745 data), validation (215
data), and test (215 data). The validation data set is mainly used to
track the convergence rate of the training process and to avoid over-
fitting. A systematic procedure through a grid search is undertaken to
determine the network hyperparameters. Accordingly, the learning
rates of a¼0:07 for two dimensional cases and a¼0:1 for three
dimensional cases with an exponential decay with the rate of 0.1 pro-
vide the optimal choice based on the test cost (C). Using high learning
rates (a) in neural networks for the permeability prediction has been
reported by other researchers as well (e.g., see Fig. 10 of Ref. 14). We
use NVIDIA Tesla V100 graphics card with the memory clock rate of
1.41GHz and 24 Gigabytes of RAM for the training process. This pro-
cedure takes approximately 30 min and 2 h, respectively, for two and
three dimensional porous media. Note that we only provided the
details of training the point-cloud neural network in this subsection. A
similar procedure has been taken to obtain the highest possible perfor-
mance for the CNN discussed in Sec. IIC 2.
III. RESULTS AND DISCUSSION
A. Two dimensional porous media
The coefficient of determination, namely, R
2
score, is a com-
monlyusedmetrictoevaluatetheperformanceofpredictingtheper-
meability (see, e.g., Refs. 11,12,14,and15) and is defined as
R2¼1X
P
i¼1
ðki~
kiÞ2
X
P
i¼1
ðki
kÞ2
;(10)
where Pis the number of data in the test set and
kdenotes the mean
of the fkigP
i¼1set. We use this metric in the current work.
The first step in the analysis of our results is to discuss the choice
of N. As pointed out in Sec. II C 1,Nis a hyperparameter of our deep
learning framework. There is no restriction on N, and users of our
machine learning platform can tune it to search for the highest achiev-
able performance. For our current digital rock images, the number of
points located on the pore–grain boundaries varies between Nmin
¼673 and Nmax ¼1698. To construct Xwhen N¼Nmin,weran-
domly select Nmin points from each point cloud of our data set.
Similarly, to establish Xwhen N¼Nmax, we add some extra points to
point clouds by randomly repeating some of their own points to fill
them up to Nmax . Additionally, one may select Nsuch that
Nmin <N<Nmax . Furthermore, there is no restriction on the selec-
tions of Nmax <Nor N<Nmin although they do not seem reasonable
choices, unless one intends to reduce the network size due to a mem-
ory limitation by the choice of N<Nmin.Figures 6(a) and 6(b),
respectively, depict the examples of point cloud illustrations for the
choices of Nmin and Nmax and their corresponding R
2
score plots. As
can be realized from Figs. 6(a) and 6(b), selection of N¼Nmin results
in a higher R
2
score compared to N¼Nmax (0.962 22 vs 0.925 36).
From the above described algorithms, we argue that because the Xset
contains redundant data in the case of N¼Nmax,itmightcausea
deviation in the path of network learning for finding the minimum in
the space of the cost function. We also find the R
2
scores of 0.904 92
and 0.874 669 for N¼1000 and N¼1300, respectively.
Figure 6(c) exhibits the resulting prediction of permeability using
the 2D-CNN introduced in Sec. II C 2. Compared to our deep learning
strategy with N¼Nmin, a 6.714% decrease in the R
2
score is observed
(0.962 22 vs 0.897 61). A comprehensive comparison between the
point-cloud neural network and 2D-CNN is made in Table I.Based
on the information tabulated in Table I,2D-CNNexperienceshigher
minimum and maximum relative errors compared to the PointNet
based network. Additionally, the size of input vector in CNN increases
approximately by a factor of 9, leading to a higher GPU memory
requirement. More importantly, the maximum possible batch size on
our computational facilities for 2D-CNN is 1024, whereas the point-
cloud neural network is able to load all 2300 training data in one
epoch. It is conjectured that this is the main reason for a lower perfor-
mance of 2D-CNN compared to the point-cloud neural network.
Figures 7(a) and 7(b) illustrate the geometries with the minimum
relative errors for the point-cloud neural network and 2D-CNN, while
Figs. 7(c) and 7(d) exhibit the geometries with the maximum relative
errors for these networks, respectively. As can be inferred from Fig. 7,
these extremums happen in different geometries for these two net-
works. It means that each of these two networks has been optimized in
two different minima in the high dimensional space of the cost func-
tion. Note that we usually do not deal with convex optimization prob-
lems in the field of machine learning.
55
However, because both the
maximum and minimum relative errors of the point-cloud neural
network are smaller than the corresponding errors of 2D-CNN (see
Table I), we conclude that the point-cloud neural network is more suc-
cessful than 2D-CNN to solve the associated optimization problem.
Note that as discussed in Sec. II C, here we report the highest possible
performance obtained for each network by a grid search on their
hyperparameters. We emphasize on the fact that the goal of this
research paper is not to prove that the proposed network can
“definitely” gain a higher score than any existing CNN-based networks.
For instance, one may argue that one can adjust the 2D-CNN pro-
posed in Sec. II C 2 by making it deeper to reach a higher score com-
pared to our new neural network. Instead, we claim that the PointNet
based network with less training efforts and less memory allocations
still can compete and outperform CNN-based networks in many cases.
As explained in Sec. II B, we normalize the permeability in the
range of [0, 1] for training the network along with the sigmoid activa-
tion function [see Eq. (8)] in the last layer of the neural network to
cover that range. Our primary motivation to use this approach is that
Kashefi et al.
38
have taken the same procedure for predicting real con-
tinuous variables such as velocity and pressure. However, since the
permeability is a positive real number, another option would be to
keep the permeability in the physical domain and use the ReLU activa-
tion function [see Eq. (7)] in the last layer. This option has been used
by several researchers such as Hong and Liu
11
and Tembely et al.
14
We implement the latter option to compare these two strategies. The
outcome of using the ReLU function [see Eq. (7)]isillustratedinFig.
8(a). A comparison between Figs. 8(a) and 6(a) indicates a higher R
2
score for our current approach (i.e., using the sigmoid activation func-
tion). Note that the scatter in Figs. 6 and 8is quantified by the R
2
scores.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-7
Published under an exclusive license by AIP Publishing
Our next machine learning experiment addresses the effect of
input and feature transforms (see Fig. 4) on the accuracy of predicted
permeability. From a computer vision point of view, the input and
output transforms have two significant contributions to the shape
classification problems. Here, we briefly describe these two contribu-
tions at a high level. One may refer to the original PointNet
32
article
for a deeper discussion. First, these two transforms enhance the net-
work performance to identify rotated objects. For instance, a rotated
FIG. 6. Different input representations and their corresponding R
2
plots for (a) point-cloud neural network with N¼Nmin, (b) point-cloud neural network with N¼Nmax, and
(c) 2D-CNN.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-8
Published under an exclusive license by AIP Publishing
cat still needs to be classified as a cat by PointNet.
32
Second, using
these two transforms, the input point clouds are aligned to a canonical
space, and it leads to a more efficient global feature extraction using
the max-pooling operator (see Fig. 4). Concerning the application con-
sidered in this research article, the first contribution mentioned above
is not useful. It is mainly because of the fact that we do not rotate our
training data for data augmentation purposes. In other words, we are
interested in permeability along the x-axis [see Eq. (3)]andbyarigid
transformation of the digital rock, the corresponding permeability
changes in this direction. However, there would be a hope that the sec-
ond contribution of the transforms to the computer vision application
improves our results as well. To answer this question, we remove the
input and feature transform blocks from the point-cloud neural net-
work (see Fig. 4) to investigate its usefulness. Figure 8(b) shows the R
2
plot as a consequence of this modification. As can be observed in Fig.
8(b),theR
2
score is reduced to 0.915 27. Hence, we conclude that the
existence of these two transforms increases the network ability for
TABLE I. Comparison between the performance of the point-cloud neural network
and 2D-CNN for learning the permeability of two dimensional porous media.
Point-cloud
neural network 2D-CNN
R
2
score 0.962 22 0.897 61
Minimum relative error 0.002% 0.014%
Maximum relative error 5.365% 13.199%
Input vector size 2019 ðNmin 3Þ16 384
(128 128 images)
Number of trainable
parameters
2 415 763 3 459 601
Maximum possible
batch size (increasing by
a factor of 2)
Able to load all
2300 training data
in one epoch
1024
FIG. 7. Geometries with (a) minimum relative error for 2D-CNN, (b) minimum relative error for the point-cloud neural network, (c) maximum relative error for 2D-CNN, and (d)
maximum relative error for the point-cloud neural network.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-9
Published under an exclusive license by AIP Publishing
permeability prediction although the network still performs with a rel-
atively high level of accuracy in their absence.
The size of the latent global feature is a critical parameter of
PointNet.
32
Qi et al.
32
discussed the effect of this parameter on the
PointNet
32
performance for the object classification and segmentation
task. Kashefi et al.
38
also investigated the influence of the global feature
size for prediction of the velocity and pressure fields on unstructured
grids(seeTableIIofRef.38). It is important to mention that the origi-
nal PointNet
32
is designed with the global feature size of 1024 (see Fig.
2ofRef.32). Table II collects the R
2
scores of various global feature
sizes for the point-cloud neural network. According to Table II,the
highest performance is obtained for the size of 1024. Similar results are
reported by Qi et al.
32
and Kashefi et al.
38
Note that by changing the
global feature size, an adjustment in the size of the MLP right after
the latent global feature is necessary to maintain the global feature as
the main information bottleneck of the network structure.
We test the generalizability of the point-cloud neural network by
prediction of the permeability of synthetic digital rock images (150
data) with the spatial correlation length of l
c
¼17 and l
c
¼33, while
the network has only seen rock images with the correlation length of l
c
¼9 in the training process. Note that from a computer science point
of view, the generalizability of a neural network should be examined
on unseen data from unseen categories. Thus, measuring the network
performance on the test set cannot be interpreted as an indication of
network generalizability because the test set contains unseen data but
from seen categories. Table III demonstrates the outcome of this test.
For l
c
¼17, we only obtain a reasonable level of accuracy with the R
2
score of 0.673 02. However, for l
c
¼33, the R
2
score takes a negative
value with a maximum relative error of approximately 30%. The nega-
tive R
2
score indicates models that have worse predictions than a base-
line based on just the mean value. A similar investigation is conducted
for 2D-CNN, and a similar trend is experienced. However, a great
reduction in the R
2
score is observed according to Table III.Compared
to the point-cloud neural network performance, 2D-CNN experiences
higher maximum relative errors and lower minimum relative errors
based on the data tabulated in Table III. This observation demon-
strates that 2D-CNN has relatively high bias on some cases (those pre-
dicted by low relative errors) and relatively high variance on other
cases (those predicted by high relative errors). From this observation,
we can also conclude that 2D-CNN is less generalizable in comparison
with the point-cloud neural network. Similar results have been
reported by Hong and Liu.
11
Although their network
11
trained on
Coconino Sandstone achieved the R
2
score of 0.872 for the test set of
permeability in the x-direction, they could only obtain the R
2
score of
0.6623 for predicting the same quantity for Bentheim Sandstone.
The next topic to discuss is the speedup factors achieved by our
deep learning configuration. The point-cloud neural network estimates
the permeability of the test set (150 data) in approximately 6 s on the
GPU machine available in our computational resources. Computing
the permeability of these 150 data using the LBM code written in
Cþþ programing language takes on average 1350 s (approximately
23 min) on a single Intel(R) Core processor with the clock rate of
2.30 GHz. Consequently, the averaged achieved speedup factor is equal
to 225 compared to the numerical simulation. Note that the factors
FIG. 8. R
2
scores obtained (a) with the
ReLU activation function [see Eq. (7)]in
the last layer of the point-cloud neural net-
work and (b) without using the input and
feature transforms in the neural network
architecture (see Fig. 4).
TABLE II. R
2
score as well as minimum and maximum relative errors for two dimensional porous media for different sizes of the global feature with the choice of N¼Nmin ; the
FC size shows the size of different layers of the fully connected layer right after the global feature in the network (see Fig. 4).
Global feature size 128 256 512 1024 2048
FC size (128, 128, 1) (256, 128, 1) (512, 256, 1) (512, 256, 1) (512, 256, 1)
R
2
score 0.897 99 0.916 51 0.915 41 0.962 22 0.918 45
Minimum relative error 0.030% 0.032% 0.004% 0.002% 0.038%
Maximum relative error 19.688% 13.276% 12.294% 5.365% 13.592%
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-10
Published under an exclusive license by AIP Publishing
reported here are not absolute and strongly depend on the efficiency
of the LBM solver and the power of GPU and CPU (central processing
unit) used. It should also be noted that the LBM solver is an in-house
research code running on CPU alone. Modern commercial LBM codes
taking advantage of GPUs are expected to be much faster than the
code used in this work.
B. Three dimensional porous media
We construct the three dimensional point clouds (see Fig. 3)sim-
ilar to the procedure explained in Sec. III A with a choice of
Nmin ¼4003. Figure 9 compares the R
2
score obtained by the point-
cloud neural network with that achieved by 3D-CNN. Accordingly,
the proposed deep learning technology gains 1.565% higher accuracy
based on the metric of the coefficient of determination (0.99151 vs
0.975 99). Table IV compares these two neural network types from
other perspectives. As tabulated in Table IV, both the minimum and
maximum relative errors of 3D-CNN are higher than those obtained
by the neural network introduced in this study. In the point-cloud
neural network, the minimum and maximum relative errors occur for
porous media with permeability of 146.658 and 10.148 mD, respec-
tively. The number of trainable parameters of 3D-CNN is slightly
greater than the corresponding number in our deep learning
TABLE III. Comparison between the generalizability of the point-cloud neural net-
work and 2D-CNN, where they have never seen samples of porous media with spa-
tial correlation lengths of l
c
¼17 and l
c
¼33 during the training procedure.
Point-cloud neural network 2D-CNN
l
c
¼17 l
c
¼33 l
c
¼17 l
c
¼33
R
2
score 0.67302 0.917 50 0.358 59 1.231 18
Minimum
relative error
0.439% 0.393% 0.048% 0.062%
Maximum
relative error
13.178% 29.164% 66.971% 86.507%
FIG. 9. Different input representations for
three dimensional geometries and their
corresponding R
2
plots for (a) point-cloud
neural network and (b) 3D-CNN.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-11
Published under an exclusive license by AIP Publishing
framework. Note that the number of trainable parameters of the
point-cloud neural network is the same for both the two and three
dimensional cases (see Tables I and IV) and in fact is independent of
number of input points. This is simply because the MLP components
in the first branch of the point-cloud neural network use shared
weights as discussed in Sec. II C 1. Note that in contrast to the point-
cloud neural network, the number of trainable parameters is a func-
tion of input size in CNN architectures.
Based on the information provided in Table IV, the input of the
point-cloud neural network is a vector of size of 12009 (Nmin 3),
while this parameter is approximately 22 times greater for 3D-CNN
and is equal to 262 144 for 64 64 64 images. This fact leads to the
condition that maximum possible usable batch size of the point-cloud
neural network is 2048, whereas it is equal to 512 for 3D-CNN. We
further investigate the effect of batch size on the performance of the
point-cloud neural network and 3D-CNN as reported, respectively, in
Tables V and VI. As can be realized from Table V, the maximum R
2
score is obtained with the batch size of 1024 for the point-cloud neural
network, while we are not able to run 3D-CNN with the batch size of
1024 and 2048 due to the lack of sufficient GPU memory. According
to the data collected in Table VI, the batch size of 512 results in the
maximum R
2
score of 3D-CNN, while the performance of 3D-CNN
for the batch size of 1024 and 2048 is unknown to us. This is exactly
the issue of 3D-CNN addressed in Sec. I. In fact, it would be
completely possible that 3D-CNN with the batch size of 1024 or 2048
provides a higher R
2
score compared to when it is trained with the
batch size of 512; however, memory restriction on GPU prevents us
trying such experiments. Contrarily, such experiments are doable on
the point-cloud neural network and eventually we obtain higher accu-
racy with the batch size of 1024 compared to 512. We observe how
our new methodology overcomes the GPU memory limitation and
ends in a higher level of prediction accuracy compared to the CNN
methodology.
Concerning the achieved speedup factor, predicting the perme-
ability of the test set (215 data) approximately takes 9 s by the point-
cloud neural network, while the Cþþ LBM solver computes the
permeability of this set in 38700 s (approximately 11 h). Hence, the
point-cloud neural network, once trained, accelerates the permeability
computations on average by a factor of 4300. Again as mentioned
earlier, the LBM solver is an in-house research code running on CPU
alone. Modern commercial LBM codes taking advantage of GPUs are
expected to be much faster than the code used in this work.
To perform the generalizabilityof the point-cloud neural network
for three dimensional porous media, we inspect the performance of
the network on predicting the permeability of Berea sandstone sam-
ples (see, e.g., Ref. 2) as natural porous media, while the point-cloud
neural network are solely trained on the synthetic data. Figure 10
exhibits the voxel and point cloud representations of one of these sam-
ples. The point-cloud neural network obtains R
2
score of 0.70437 with,
respectively, the minimum and maximum relative errors of 2.735%
and 35.578% over eight samples. We observe that only a reasonable
accuracy level is gained because of two main reasons. First, although
the permeability of the natural samples is in the range of training data,
they have different spatial structure than data during the training pro-
cedure. Second, the number of points (N) in the clouds constructing
the boundary of pore spaces in the natural samples vary between 4388
and 8696; this is while we set N¼4003 for the network and it ends in
losing even further information about the correct structure of the nat-
ural samples and thus decreasing the R
2
score specifically for samples
with large numbers of N(compared to 4003). It is concluded that to
obtain higher R
2
scores, the network should be trained on similar nat-
ural samples or similar synthetic data from permeability, porosity, and
spatial correlation length perspectives. Note that the goal of such an
experiment is to test the generalizability of the network, meaning that
we quantitatively investigate how a deviation from one of these three
features negatively affects the accuracy of prediction; however, a
decrease in the R
2
score would be expected in advance. As discussed in
Sec. III A, we emphasize that the network must be asked to predict
unseen data from unseen categories in a generalizability test.
At the end of this subsection, we address three points. First, our
machine learning experiments shows that the contribution of input
and feature transforms (see Fig. 4) to increasing the accuracy of pre-
dicting the permeability of three dimensional point clouds is insignifi-
cant (less than 0.1%). Thus, one may optionally remove these two
transforms from the neural network to make it faster and lighter.
Second, an important feature of the point-cloud neural network is its
scalability. Depending on the number of points in the training point
clouds, one may make the network smaller or larger. Alternatively,
TABLE IV. Comparison between the performance of the point-cloud neural network
and 3D-CNN for learning the permeability of three dimensional porous media.
Point-cloud
neural network 3D-CNN
R
2
score 0.991 51 0.975 99
Minimum relative error 0.030% 0.279%
Maximum relative error 50.487% 57.745%
Input vector size 12 009
ðNmin 3Þ
262 144
(64 64 64 images)
Number of trainable
parameters
2 415 763 2 585 169
Maximum possible batch
size (increasing by a
factor of 2)
2048 512
TABLE V. Effect of batch size on the performance of the point-cloud neural network for predicting the permeability of three dimensional porous media.
Batch size 8 16 32 64 128 256 512 1024 2048
R
2
score 0.762 97 0.707 58 0.724 89 0.934 64 0.984 22 0.982 69 0.984 66 0.991 51 0.990 36
Minimum relative error (%) 0.013 0.032 0.060 0.224 0.019 0.058 0.066 0.030 0.031
Maximum relative error (%) 264.369 578.575 104.570 60.173 64.171 43.655 75.154 50.487 40.863
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-12
Published under an exclusive license by AIP Publishing
one may investigate the effect of the network size on its performance.
For example, the size of network can be reduced by scaling its MLPs
by a factor of 0.25, which leads to MPLs with sizes of (16, 16), (16, 32,
128), and (128, 64, 1), respectively, from left to right as shown in
Fig. 4. In this case, the number of trainable parameters of the network
without input and feature transforms decreases from 809601 to 51873.
Third, training the point-cloud neural network on a data set contain-
ing porous media with a great range of spatial correlation lengths is
challenging mainly because such set in practice leads to point-cloud
subsets with a comparatively big difference between N
max
and N
min
.
As discussed in Sec. III A,Nis a hyper-parameter that needs to be
tuned in the point-cloud neural network; however, when Nmax Nmin
is very large, the training process becomes demanding. In fact, by
selecting an Nclose to N
min
, the corresponding geometries of point
sets with NNmin are poorly represented. On the other hand, by
choosing an Nclose to N
max
, point sets with NNmax contain
unreasonably a great number of repeated points, which eventually
appear as redundant data to the point-cloud network and lead to
decreasing the network performance. Moreover, our machine learning
experiments show that a moderate N(e.g., arithmetic or geometric
mean of N
min
and N
max
) is not an ideal option as well. Hence, one of
our future study plans is to resolve these types of limitations from the
proposed point-cloud deep learning framework. It is conjectured that
such improvements could positively affect the generalizability of the
network as well.
C. Potentials for fluid flow field predictions
In this article, our main focus is the prediction of permeability
directly from the digital rock images. However, an alternative
approach for the permeability prediction is to first predict the entire
velocity fields in the pore spaces using a machine learning framework
and then compute the permeability from the predicted velocity field.
This approach has been so far taken by several researchers (see, e.g.,
Refs. 23 and 24). In this approach, a neural network is used as a
replacement of conventional numerical solvers to provide an end-to-
end mapping from the geometry of porous media to the fluid field of
interest in the pore space. Due to high complexity in geometries of
natural porous media, an efficient geometry representation in neural
networks is necessary. CNNs as deep learning tools have been so far
utilizedforpredictingthevelocityfieldsinporousmedia.Forinstance,
Santos et al.
24
used a three dimensional CNN but only to predict one
component of the velocity field (parallel to the direction of the applied
pressure gradient). Da Wang et al.
23
proposed a CNN based on U-
Net
56
to predict the velocity fields in two and three dimensional
porous media.
There are two common approaches to represent the geometry of
porous media in the case of using CNNs. The first approach is to mask
the pixels associated with the grains (see, e.g., Ref. 24). There are two
major shortcomings with this approach. The first one is that for each
numerical array (representing the porous media) a huge number of
pixels have to be masked, specifically for porous media with low
TABLE VI. Effect of batch size on the performance of 3D-CNN for predicting the permeability of three dimensional porous media; the cross symbol () indicates that training is
impossible due to limitation of GPU memory.
Batch size 8 16 32 64 128 256 512 1024 2048
R
2
score 0.734 34 0.682 99 0.927 04 0.932 44 0.958 03 0.968 23 0.975 99 
Minimum relative error (%) 0.096 0.300 0.079 0.320 0.007 0.097 0.279 
Maximum relative error (%) 116.177 80.346 80.026 73.766 77.470 79.800 57.745 
FIG. 10. Voxel and point cloud represen-
tations of one of Berea sandstone sam-
ples as a natural porous medium used for
exploring the generalizability of our deep
learning framework.
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-13
Published under an exclusive license by AIP Publishing
porosity. In fact, a considerable portion of computational capacity of
CNNs is wasted by taking this strategy. The second one is that for digi-
tal rock images with high resolutions, a huge (and uncommon) mem-
ory size on Graphics Processing Units (GPUs) is necessary in order to
load the entire domain of interest even with a batch size of one. This
issue becomes highlighted specifically for prediction of flow fields in
three dimensional porous media. For instance, this shortcoming has
been reported by Santos et al.
24
when it was impossible for them to
train their own CNN over the entire simulation domain. From a com-
puter vision point of view, a possible solution would be breaking the
entire domain into a few subdomains. Although it might resolve the
memory issue, it would bring other concerns and constrains for CNNs
training and the velocity field prediction. A full discussion on this issue
canbefoundinSec.3.1ofRef.24. The minor issue of this approach is
relevant to programing from a computer science point of view. An effi-
cient function needs to be written to handle the pixel-masking proce-
dure for a large number of training data, each one with a different
pattern of fluid flow channels. In this approach, CNNs learn over
training data that the fluid flow fields in pore spaces are a function of
the number and pattern of the masked pixels. It is important to men-
tion that this technique has been widely used in the area of fluid
mechanics for deep learning flows over bluff bodies or airfoils (see,
e.g., Ref. 57). The second approach is to label the input array using two
different digit numbers: one representing the pore spaces and another
representing the grain (solid) portions. Similarly, the corresponding
pixels of pore spaces in the output array take the numerical value of
the velocity field, whereas the corresponding pixels of grain spaces in
the output array take the same number as input. In comparison, with
the first approach, this scheme is easier to program and implement
and also consumes less wall time over each epoch of training. On the
other hand, the main difficulty with this approach is that CNNs not
only have to learn the fluid flow fields in pore spaces, but also have to
learn the geometry of grain spaces (see, e.g., Ref. 23). There are three
major shortcomings with this approach. The first and second issues
are similar to the first approach: wasting considerable computational
resources of CNN frameworks and requiring an unreasonable memory
size on GPUs even for batching one sample from data set. The third
major shortcoming is that machine learning experiments showed that
CNNs have troubles identifying the pore spaces vs the grain spaces,
specifically when the geometry of flow cluster becomes highly compli-
cated and the effective porosity of the rock decreases. For instance,
CNNs proposed by Da Wang et al.
23
predicted the velocity fields
(regardless the accuracy of its magnitude) in regions where the flow
does not even exist (i.e., in grain spaces). Note that there are a few
alternative methods to represent the pore spaces as the input of CNNs,
such as using the Euclidean distance transform function (see, e.g.,
Refs. 23 and 24), instead of using a single number. Although these
alternatives might increase the CNN performance, the fundamental
issues addressed here remains unchanged. In summary, both of these
two approaches suffer from modeling and involving the grain spaces
in a CNN deep learning framework. To resolve these issues, we suggest
only taking the pore space of a porous medium and representing it as
a set of points that constructs a point cloud. Point clouds represent,
respectively, the surface and volume of pore spaces of two and three
dimensional porous media. Consequently, one may use the segmenta-
tion component of PointNet
32
to establish an end-to-end mapping
between the spatial coordinates of each point of a cloud and the
numerical values of the velocity vector at that point. It is conjectured
that PointNetþþ
58
and Kpconv
33
would have higher performance in
comparison with PointNet
32
because they pay more attention to local
features of a given geometry.
Moreover, designing an efficient cost function in such problems
is critical (see, e.g., Sec. 2.3.2 of Ref. 24). Cost functions so far used are
mainly based on L
1
or L
2
norm error of the velocity fields.
23,24
To
enhance the performance of such neural networks and accuracy of the
velocity field prediction, we suggest two strategies. Both of these two
strategies are based on adding information of the flow governing equa-
tions to neural network cost functions. The first strategy leads to a
supervised deep learning approach, while the second one results in an
unsupervised (or semi-supervised) methodology. The first approach is
to add the residual of the continuity and Navier–Stokes equations
[Eqs. (1) and (2)] to the cost function. This approach has been carried
out in other research areas (see, e.g., Refs. 57 and 59). The second
approach is to use the technology of the Physics Informed Neural
Network (PINN).
60–62
In PINNs, the cost function is defined based on
the governing equations of the problem of interest as well as the
desired initial and boundary conditions, while there is no need of
labeled data for training. One may refer to Refs. 60–62 for a deeper
discussion on PINNs. Note that there is no mechanism in the current
version of PINNs to capture the variations in the geometry of problem
domains. In other words, the parameters (e.g., weights and bias) of
PINNs are not a function of the geometry of physical domains. Thus,
the combination of PointNet
32
(or other point-cloud based neural net-
works) with PINNs has the potential to resolve this issue.
IV. SUMMARY
In this study, we introduced a novel point-cloud based deep
learning configuration for permeability predictions of digital porous
media. We designed the architecture of this configuration according to
the classification branch of PointNet.
32
Taking the advantages of the
point-cloud based deep learning methodology, limitations on GPU
memory requirements were relaxed and selecting higher batch sizes
compared to CNNs became possible. It was mainly due to dramatically
diminishing the size of network inputs by only taking the boundary of
solid matrix and pore spaces in a porous medium via point cloud rep-
resentations, rather than taking its whole volume via voxel representa-
tions. Freedom in the choice of batch size provided the chance of
exploring a relatively wide range of batch sizes to obtain the highest
possible accuracy of the permeability prediction. We concentrated on
synthetic digital rocks as test cases. According to the metric of coeffi-
cient of determination, our deep learning technique achieved excellent
accuracy for the predicted permeability of both two and three dimen-
sional porous media. Compared to a numerical LBM solver, the point-
cloud neural network predicted the permeability of test set a few thou-
sand times faster. Finally, we discussed the generalizability of the
point-cloud neural network by examining it over two unseen catego-
ries: real-world samples and synthetic samples but with unseen spatial
correlation lengths.
ACKNOWLEDGMENTS
We acknowledge the sponsors of the Stanford Center for Earth
Resources Forecasting (SCERF) and support from Professor Steve
Graham, the Dean of the Stanford School of Earth, Energy and
Environmental Sciences. The work was funded by Shell-Stanford
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-14
Published under an exclusive license by AIP Publishing
collaborative project on Digital Rock Physics and the Army
Research Office Contract No. W911NF1810008. Some of the
computing for this project was performed on the Sherlock cluster.
We would like to thank Stanford University and the Stanford
Research Computing Center for providing computational resources
and support that contributed to these research results. Additionally,
we wish to thank the reviewers for their insightful comments.
DATA AVAILABILITY
The data that support the findings of this study are available
from the corresponding author upon reasonable request.
REFERENCES
1
H. Andr
a, N. Combaret, J. Dvorkin, E. Glatt, J. Han, M. Kabel, Y. Keehm, F.
Krzikalla, M. Lee, C. Madonna et al., “Digital rock physics benchmarks—Part I:
Imaging and segmentation,” Comput. Geosci. 50, 25–32 (2013).
2
H. Andr
a, N. Combaret, J. Dvorkin, E. Glatt, J. Han, M. Kabel, Y. Keehm, F.
Krzikalla, M. Lee, C. Madonna et al., “Digital rock physics benchmarks—Part
II: Computing effective properties,” Comput. Geosci. 50, 33–43 (2013).
3
M. Gruber, C. Johnson, C. Tang, M. Jensen, L. Yde, and C. H
elix-Nielsen,
“Computational fluid dynamics simulations of flow and concentration polari-
zation in forward osmosis membrane systems,” J. Membr. Sci. 379, 488–495
(2011).
4
M. J. Blunt, B. Bijeljic, H. Dong, O. Gharbi, S. Iglauer, P. Mostaghimi, A.
Paluszny, and C. Pentland, “Pore-scale imaging and modelling,” Adv. Water
Resour. 51, 197–216 (2013).
5
K. Khanafer, K. Cook, and A. Marafie, “The role of porous media in modeling
fluid flow within hollow fiber membranes of the total artificial lung,” J. Porous
Media 15, 113 (2012).
6
S. Karimpouli and P. Tahmasebi, “Segmentation of digital rock images using
deep convolutional autoencoder networks,” Comput. Geosci. 126, 142–150
(2019).
7
Y. D. Wang, M. Shabaninejad, R. T. Armstrong, and P. Mostaghimi, “Deep neural
networks for improving physical accuracy of 2D and 3D multi-mineral segmenta-
tion of rock micro-CT images,” Appl. Soft Comput. 104, 107185 (2021).
8
Y. Niu, P. Mostaghimi, M. Shabaninejad, P. Swietojanski, and R. T.
Armstrong, “Digital rock segmentation for petrophysical analysis with reduced
user bias using convolutional neural networks,” Water Resour. Res. 56,
e2019WR026597, https://doi.org/10.1029/2019WR026597 (2020).
9
K. M. Graczyk and M. Matyka, “Predicting porosity, permeability, and tortuos-
ity of porous media from images by deep learning,” Sci. Rep. 10, 21488 (2020).
10
A. Bhatt, “Reservoir properties from well logs using neural networks,” Ph.D.
thesis (Norwegian University of Science and Technology, 2002).
11
J. Hong and J. Liu, “Rapid estimation of permeability from digital rock using
3D convolutional neural network,” Comput. Geosci. 24, 1523–1539 (2020).
12
J. Wu, X. Yin, and H. Xiao, “Seeing permeability from images: Fast prediction
with convolutional neural networks,” Sci. Bull. 63, 1215–1222 (2018).
13
M. R
oding, Z. Ma, and S. Torquato, “Predicting permeability via statistical learn-
ing on higher-order microstructural information,” Sci. Rep. 10, 15239 (2020).
14
M. Tembely, A. M. AlSumaiti, and W. Alameri, “A deep learning perspective
on predicting permeability in porous media from network modeling to direct
simulation,” Comput. Geosci. 24, 1541–1556 (2020).
15
J. Tian, C. Qi, Y. Sun, Z. M. Yaseen, and B. T. Pham, “Permeability prediction
of porous media using a combination of computational fluid dynamics and
hybrid machine learning methods,” Eng. Comput. 2020, 1–17.
16
A. Zolotukhin and A. Gayubov, “Machine learning in reservoir permeability
prediction and modelling of fluid flow in porous media,” IOP Conf. Ser.: Mater.
Sci. Eng. 700, 012023 (2019).
17
N. Alqahtani, F. Alzubaidi, R. T. Armstrong, P. Swietojanski, and P.
Mostaghimi, “Machine learning for predicting properties of porous media from
2D x-ray images,” J. Pet. Sci. Eng. 184, 106514 (2020).
18
A. Rabbani, M. Babaei, R. Shams, Y. D. Wang, and T. Chung, “Deepore: A
deep learning workflow for rapid and comprehensive characterization of
porous materials,” Adv. Water Resour. 146, 103787 (2020).
19
F. Bordignon, L. Figueiredo, R. Exterkoetter, B. B. Rodrigues, and M. Correia,
“Deep learning for grain size and porosity distributions estimation on micro-
CT images,” in Proceedings of the 16th International Congress of the Brazilian
Geophysical Society & Expogef (2019).
20
V. H
ebert, T. Porcher, V. Planes, M. L
eger, A. Alperovich, B. Goldluecke, O.
Rodriguez, and S. Youssef, “Digital core repository coupled with machine
learning as a tool to classify and assess petrophysical rock properties,” in E3S
Web Conf. 146, 01003 (2020).
21
H. Wu, W.-Z. Fang, Q. Kang, W.-Q. Tao, and R. Qiao, “Predicting effective dif-
fusivity of porous media from images by deep learning,” Sci. Rep. 9, 20387
(2019).
22
S. Karimpouli and P. Tahmasebi, “Image-based velocity estimation of rock
using convolutional neural networks,” Neural Networks 111, 89–97 (2019).
23
Y. Da Wang, T. Chung, R. T. Armstrong, and P. Mostaghimi, “ML-LBM:
Machine learning aided flow simulation in porous media,” arXiv:2004.11675
(2020).
24
J. E. Santos, D. Xu, H. Jo, C. J. Landry, M. Prodanovic´, and M. J. Pyrcz,
“PoreFlow-Net: A 3D convolutional neural network to predict fluid flow
through porous media,” Adv. Water Resour. 138, 103539 (2020).
25
S. Li, X. Shen, Y. Dou, S. Ni, J. Xu, K. Yang, Q. Wang, and X. Niu, “A novel
memory-scheduling strategy for large convolutional neural network on
memory-limited devices,” Comput. Intell. Neurosci. 2019, 4328653.
26
I. Goodfellow, Y. Bengio, and A. Courville, Deep Learning (MIT Press, 2016).
27
N. S. Keskar, D. Mudigere, J. Nocedal, M. Smelyanskiy, and P. T. P. Tang, “On
large-batch training for deep learning: Generalization gap and sharp minima,”
arXiv:1609.04836 (2016).
28
I. Kandel and M. Castelli, “The effect of batch size on the generalizability of the
convolutional neural networks on a histopathology dataset,” ICT Express 6,
312–315 (2020).
29
D. Masters and C. Luschi, “Revisiting small batch training for deep neural
networks,” arXiv:1804.07612 (2018).
30
Y. Bengio, “Practical recommendations for gradient-based training of deep
architectures,” in Neural Networks: Tricks of the Trade (Springer, 2012), pp.
437–478.
31
Y. Wang, Y. Sun, Z. Liu, S. E. Sarma, M. M. Bronstein, and J. M. Solomon,
“Dynamic graph CNN for learning on point clouds,” ACM Trans. Graph. 38 ,
146 (2019).
32
C. R. Qi, H. Su, K. Mo, and L. J. Guibas, “PointNet: Deep learning on point sets
for 3D classification and segmentation,” in Proceedings of the IEEE Conference
on Computer Vision and Pattern Recognition (IEEE, 2017), pp. 652–660.
33
H. Thomas, C. R. Qi, J.-E. Deschaud, B. Marcotegui, F. Goulette, and L. J.
Guibas, “KPConv: Flexible and deformable convolution for point clouds,” in
Proceedings of the IEEE/CVF International Conference on Computer Vision
(IEEE, 2019), pp. 6411–6420.
34
C. R. Qi, O. Litany, K. He, and L. J. Guibas, “Deep hough voting for 3D object
detection in point clouds,” in proceedings of the IEEE/CVF International
Conference on Computer Vision (IEEE, 2019), pp. 9277–9286.
35
C. R. Qi, W. Liu, C. Wu, H. Su, and L. J. Guibas, “Frustum pointNets for 3D
object detection from RGB-D Data,” in Proceedings of the IEEE Conference on
Computer Vision and Pattern Recognition (IEEE, 2018), pp. 918–927.
36
X. Liu, C. R. Qi, and L. J. Guibas, “FlowNet3D: Learning scene flow in 3D point
clouds,” in Proceedings of the IEEE/CVF Conference on Computer Vision and
Pattern Recognition (IEEE, 2019), pp. 529–537.
37
D. Rempe, T. Birdal, Y. Zhao, Z. Gojcic, S. Sridhar, and L. J. Guibas, “CaSPR:
Learning canonical spatiotemporal point cloud representations,”
arXiv:2008.02792 (2020).
38
A. Kashefi, D. Rempe, and L. J. Guibas, “A point-cloud deep learning frame-
work for prediction of fluid flow fields on irregular geometries,” Phys. Fluids
33, 027104 (2021).
39
D. Rempe, S. Sridhar, H. Wang, and L. J. Guibas, “Learning generalizable final-
state dynamics of 3D rigid objects,” in IEEE Conference on Computer Vision
and Pattern Recognition Workshops,CVPR Workshops,Long Beach, CA, USA,
16–20 June, 2019 (Computer Vision Foundation/IEEE, 2019), pp. 17–20.
40
R. S. DeFever, C. Targonski, S. W. Hall, M. C. Smith, and S. Sarupria, “A gen-
eralized deep learning approach for local structure identification in molecular
simulations,” Chem. Sci. 10, 7503–7515 (2019).
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-15
Published under an exclusive license by AIP Publishing
41
Y. Keehm, T. Mukerji, and A. Nur, “Permeability prediction from thin sections:
3D reconstruction and lattice-Boltzmann flow simulation,” Geophys. Res. Lett.
31, L04606, https://doi.org/10.1029/2003GL018761 (2004).
42
H. Darcy, Les Fontaines Publiques de la Ville de Dijon: Exposition et
Application (Victor Dalmont, 1856).
43
A. Eshghinejadfard, L. Dar
oczy, G. Janiga, and D. Th
evenin, “Calculation of
the permeability in porous media using the lattice Boltzmann method,” Int. J.
Heat Fluid Flow 62, 93–103 (2016).
44
C. Lantu
ejoul, Geostatistical Simulation: Models and Algorithms (Springer
Science & Business Media, 2013).
45
W. Xu, A. Journel et al., “GTSIM: Gaussian truncated simulations reservoir
units in a W. Texas carbonate field,” SPE Paper No. 27412, 1993.
46
S. Ioffe and C. Szegedy, “Batch normalization: Accelerating deep network train-
ing by reducing internal covariate shift,” in International Conference on
Machine Learning (PMLR, 2015), pp. 448–456.
47
Y. Yu, Z. Gong, P. Zhong, and J. Shan, “Unsupervised representation learning
with deep convolutional neural network for remote sensing images,” in
International Conference on Image and Graphics (Springer, 2017), pp. 97–108.
48
A. Ajit, K. Acharya, and A. Samanta, “A review of convolutional neural
networks,” in International Conference on Emerging Trends in Information
Technology and Engineering (ic-ETITE) (IEEE, 2020), pp. 1–5.
49
A. Khan, A. Sohail, U. Zahoora, and A. S. Qureshi, “A survey of the recent
architectures of deep convolutional neural networks,” Artif. Intell. Rev. 53,
5455–5516 (2020).
50
J. Nagi, F. Ducatelle, G. A. Di Caro, D. Cires¸an, U. Meier, A. Giusti, F. Nagi, J.
Schmidhuber, and L. M. Gambardella, “Max-pooling convolutional neural net-
works for vision-based hand gesture recognition,” in IEEE International
Conference on Signal and Image Processing Applications (ICSIPA) (IEEE,
2011), pp. 342–347.
51
P. Wang, P. Chen, Y. Yuan, D. Liu, Z. Huang, X. Hou, and G. Cottrell,
“Understanding convolution for semantic segmentation,” in IEEE Winter
Conference on Applications of Computer Vision (WACV) (IEEE, 2018), pp.
1451–1460.
52
J. Yamanaka, S. Kuwashima, and T. Kurita, “Fast and accurate image super res-
olution by deep CNN with skip connection and network in network,” in
International Conference on Neural Information Processing (Springer, 2017),
pp. 217–225.
53
V. Sekar, Q. Jiang, C. Shu, and B. C. Khoo, “Fast flow field prediction over air-
foils using deep learning approach,” Phys. Fluids 31, 057103 (2019).
54
D. P. Kingma and J. Ba, “Adam: A method for stochastic optimization,”
arXiv:1412.6980 (2014).
55
P. Jain and P. Kar, “Non-convex optimization for machine learning,”
arXiv:1712.07897 (2017).
56
O. Ronneberger, P. Fischer, and T. Brox, “U-Net: Convolutional networks for
biomedical image segmentation,” in International Conference on Medical
Image Computing and Computer-Assisted Intervention (Springer, 2015), pp.
234–241.
57
S. Bhatnagar, Y. Afshar, S. Pan, K. Duraisamy, and S. Kaushik, “Prediction of
aerodynamic flow fields using convolutional neural networks,” Comput. Mech.
64, 525–545 (2019).
58
C. R. Qi, L. Yi, H. Su, and L. J. Guibas, “PointNetþþ: Deep hierarchical feature
learning on point sets in a metric space,” in Advances in Neural Information
Processing Systems 30: Annual Conference on Neural Information Processing
Systems,Long Beach, CA, USA, 4–9 December, 2017, edited by I. Guyon, U. von
Luxburg, S. Bengio, H. M. Wallach, R. Fergus, S. V. N. Vishwanathan, and R.
Garnett (Curran Associates, Inc., 2017), pp. 5099–5108.
59
A. Subramaniam, M. L. Wong, R. D. Borker, S. Nimmagadda, and S. K. Lele,
“Turbulence enrichment using physics-informed generative adversarial
networks,” arXiv:2003.01907 (2020).
60
A. D. Jagtap, E. Kharazmi, and G. E. Karniadakis, “Conservative physics-
informed neural networks on discrete domains for conservation laws:
Applications to forward and inverse problems,” Comput. Methods Appl.
Mech. Eng. 365, 113028 (2020).
61
Z. Mao, A. D. Jagtap, and G. E. Karniadakis, “Physics-informed neural net-
works for high-speed flows,” Comput. Methods Appl. Mech. Eng. 360, 112789
(2020).
62
M. Raissi, P. Perdikaris, and G. E. Karniadakis, “Physics-informed neural net-
works: A deep learning framework for solving forward and inverse problems
involving nonlinear partial differential equations,” J. Comput. Phys. 378,
686–707 (2019).
Physics of Fluids ARTICLE scitation.org/journal/phf
Phys. Fluids 33, 097109 (2021); doi: 10.1063/5.0063904 33, 097109-16
Published under an exclusive license by AIP Publishing
... Fluids flow in the granular aggregates can be investigated by experiments on the tests of transport parameters like permeability [4,5] which characterizes the hydrodynamic properties of the mass flow in the porous media [6]. The permeability can be measured by Darcy test which uses two constant pressure heads of a fluid to generate steady flow in the porous media and then the permeability can be calculated based on the flow rate, pressure gradient and sample sizes [5]. ...
... Fluids flow in the granular aggregates can be investigated by experiments on the tests of transport parameters like permeability [4,5] which characterizes the hydrodynamic properties of the mass flow in the porous media [6]. The permeability can be measured by Darcy test which uses two constant pressure heads of a fluid to generate steady flow in the porous media and then the permeability can be calculated based on the flow rate, pressure gradient and sample sizes [5]. For convenience, the falling-head method can also be used in the granular packing though improvements are needed for the high flow rate of larger-size aggregates packing [7]. ...
... The simulation mainly involves two steps, i.e., collision and streaming. After the simulation reaches to convergency, the permeability can be obtained by the Darcy's law [5,34]: ...
Preprint
Full-text available
Properties of fluids flow in granular aggregates are important for the design of pervious infrastructures used to alleviate urban water-logging problems. Here in this work, five groups of aggregates packing with similar average porosities but varying particle sizes were scanned by a high-energy X-ray computed tomography (X-CT) facility. The structures of the packings were reconstructed. Porosities were calculated and compared with those measured by the volume and mass of infilled water in the packing. Then pore networks were extracted and analyzed. Simulations of fluids flow in the packings were performed by using a lattice Boltzmann method (LBM) with BGK (Bhatnagar-Gross-Krook) collision model in the pore-network domain of the packings. Results showed wall effect on the porosity of aggregates packing was significant and the influence increased with the aggregate sizes. In addition, Poisson law and power law can be used to fit the coordination number and coordination volume of the packing's pore network, respectively. Moreover, the mass flow rates of fluids in the aggregates were affected by the porosities. On the two-dimensional slices, the mass flow rate decreased when the slice porosity increased. But for the three-dimensional blocks, the average mass flow rate increased with the volume porosity. And the permeability of the aggregates packing showed correlating change trend with the average pore diameter and fitting parameters of coordination volumes, when the sizes of aggregates changed. Though the limitation of merging interfaces causing fluctuation and discontinuity on micro parameters of fluid flow existed, the methods and results here may provide knowledge and insights for numerical simulations and optimal design of aggregate-based materials.
... In recent years, with the explosive growth of data and rapid development of computer hardware, deep learning algorithms, represented by neural networks, have made significant advances in areas such as computer vision and image processing . Deep learning has been widely applied in various applications of porous media research, such as permeability estimation (WU et al., 2018) (RABBANI et al., 2020 (ISHOLA and VILCáEZ, 2022) (Kashefi and Mukerji, 2021), porosity characterization (WU et al., 2018) (RABBANI et al., 2020) (ISHOLA and VILCáEZ, 2022), flow field prediction (Takbiri et al., 2022) (Takbiri-Borujeni et al., 2020), and colloidal deposition (Marcato et al., 2021), among others. Numerous studies have shown that data-driven deep learning methods have great potential for predicting flow parameters in porous media. ...
... Numerous studies have shown that data-driven deep learning methods have great potential for predicting flow parameters in porous media. Taking pure data-driven machine learning as an example, Keshefi et al. (Kashefi and Mukerji, 2021) utilized point cloud neural networks to obtain the geometric shapes of solid skeletons and pore boundaries, representing porous media as a set of points rather than complete porous media voxels. This approach enables the prediction of permeability and flow fields in porous media using point clouds. ...
... A prediction at an arbitrary time requires iteratively feeding past ML predictions into the model. Kashefi and Mukerji [20] used neural networks to predict the permeability from 2D or 3D point clouds of the boundary surface of porous media. These studies all had very specific applications in mind, and our work will build upon these methods to show how to approach a general hydrodynamics problem. ...
Preprint
Full-text available
We are interested in the computational study of shock hydrodynamics, i.e. problems involving compressible solids, liquids, and gases that undergo large deformation. These problems are dynamic and nonlinear and can exhibit complex instabilities. Due to advances in high performance computing it is possible to parameterize a hydrodynamic problem and perform a computational study yielding $\mathcal{O}\left({\rm TB}\right)$ of simulation state data. We present an interactive machine learning tool that can be used to compress, browse, and interpolate these large simulation datasets. This tool allows computational scientists and researchers to quickly visualize "what-if" situations, perform sensitivity analyses, and optimize complex hydrodynamic experiments.
... Finally, it is not straightforward to embed the complexity of varying degrees of confinement effects in the ML model. While much progress has been made in emulating transport in porous media using ML [6][7][8][9][10][11][12][13][14][15], models in the literature are not general enough to be used as a generic tool. ...
Article
Full-text available
Modeling effective transport properties of 3D porous media, such as permeability, at multiple scales is challenging as a result of the combined complexity of the pore structures and fluid physics—in particular, confinement effects which vary across the nanoscale to the microscale. While numerical simulation is possible, the computational cost is prohibitive for realistic domains, which are large and complex. Although machine learning (ML) models have been proposed to circumvent simulation, none so far has simultaneously accounted for heterogeneous 3D structures, fluid confinement effects, and multiple simulation resolutions. By utilizing numerous computer science techniques to improve the scalability of training, we have for the first time developed a general flow model that accounts for the pore-structure and corresponding physical phenomena at scales from Angstrom to the micrometer. Using synthetic computational domains for training, our ML model exhibits strong performance (R ² = 0.9) when tested on extremely diverse real domains at multiple scales.
... This requirement for extensive Graphics Processing Unit (GPU) memory not only demands resources but also affects training accuracy by constraining the batch size, a critical parameter for model optimization [25][26][27]. To address some of these limitations, PointNet architectures were introduced, which represent the microstructures with a point cloud to predict the permeability of porous media [28]. Despite its more economic use of GPU memory compared to CNNs, PointNet also has its limitations. ...
Article
This study presents a Graph Neural Networks (GNNs)-based approach for predicting the effective elastic moduli of rocks from their digital CT-scan images. We use the Mapper algorithm to transform 3D digital rock images into graph datasets, encapsulating essential geometrical information. These graphs, after training, prove effective in predicting elastic moduli. Our GNN model shows robust predictive capabilities across various graph sizes derived from various subcube dimensions. Not only does it perform well on the test dataset, but it also maintains high prediction accuracy for unseen rocks and unexplored subcube sizes. Comparative analysis with Convolutional Neural Networks (CNNs) reveals the superior performance of GNNs in predicting unseen rock properties. Moreover, the graph representation of microstructures significantly reduces GPU memory requirements (compared to the grid representation for CNNs), enabling greater flexibility in the batch size selection. This work demonstrates the potential of GNN models in enhancing the prediction accuracy of rock properties and boosting the efficiency of digital rock analysis.
... 11 Additionally, the inherent susceptibility of soil structure to changes makes it challenging to obtain complete soil samples for permeability studies. 12 Consequently, some scholars aim to clarify the percolation mechanism 13 and establish the relationship between pore characteristics and permeability through theoretical analysis, 14 experiments, 15 simulation, 16 and artificial intelligence, 3,17 to facilitate the engineering practice. ...
Article
The fluid transport in porous media is a critical property for oil and gas exploitation, construction engineering, and environmental protection. It is profoundly influenced by pore geometry and mineral properties. Currently, the Kozeny–Carman equation serves as the permeability prediction equation for porous media, established on the circular pores model. However, it fails to fully account for the impact of pore shape and mineral properties of the soil, leading to significant deviations between predicted and measured soil permeability results. In this paper, based on scanning electron microscope image and mercury intrusion porosimetry, the pores were divided into circular pores and narrow slit pores according to the ratios of pore area and circumference. Then, the quantitative expression of the two types of pores and their connectivity and tortuosity were given, and the circular and narrow slit composite pore model was used to describe the soil pore. Subsequently, the electrostatic potential of pore water was calculated by the Poisson–Boltzmann equation to consider the adsorption effect of minerals on pore water. Combined with the Navier–Stokes equation, the permeability prediction equation considering pore geometry, pore connectivity, and tortuosity and mineral properties was established. Finally, the experimental results illustrated that the theoretical prediction results were in good agreement with the experimental results. The proposed permeability prediction equation proves valuable for assessing and predicting the fluid transport in porous media.
... This requirement for extensive Graphics Processing Unit (GPU) memory not only demands resources but also affects training accuracy by constraining the batch size, a critical parameter for model optimization (Smith et al., 2017;He et al., 2019;Kandel and Castelli, 2020). To address some of these limitations, PointNet architectures were introduced, which represent the microstructures with a point cloud to predict the permeability of porous media (Kashefi and Mukerji, 2021). Despite its more economic use of GPU memory compared to CNNs, PointNet also has its limitations. ...
Preprint
Full-text available
This study presents a Graph Neural Networks (GNNs)-based approach for predicting the effective elastic moduli of rocks from their digital CT-scan images. We use the Mapper algorithm to transform 3D digital rock images into graph datasets, encapsulating essential geometrical information. These graphs, after training, prove effective in predicting elastic moduli. Our GNN model shows robust predictive capabilities across various graph sizes derived from various sub-cube dimensions. Not only does it perform well on the test dataset, but it also maintains high prediction accuracy for unseen rocks and unexplored subcube sizes. Comparative analysis with Convolutional Neural Networks (CNNs) reveals the superior performance of GNNs in predicting unseen rock properties. Moreover, the graph representation of microstructures significantly reduces GPU memory requirements (compared to the grid representation for CNNs), enabling greater flexibility in the batch size selection. This work demonstrates the potential of GNN models in enhancing the prediction accuracy of rock properties and boosting the efficiency of digital rock analysis.
Article
Full-text available
Fourier neural operators (FNOs) are invariant with respect to the size of input images, and thus images with any size can be fed into FNO-based frameworks without any modification of network architectures, in contrast to traditional convolutional neural networks. Leveraging the advantage of FNOs, we propose a novel deep-learning framework for classifying images with varying sizes. Particularly, we simultaneously train the proposed network on multi-sized images. As a practical application, we consider the problem of predicting the label (e.g., permeability) of three-dimensional digital porous media. To construct the framework, an intuitive approach is to connect FNO layers to a classifier using adaptive max pooling. First, we show that this approach is only effective for porous media with fixed sizes, whereas it fails for porous media of varying sizes. To overcome this limitation, we introduce our approach: instead of using adaptive max pooling, we use static max pooling with the size of channel width of FNO layers. Since the channel width of the FNO layers is independent of the input image size, the introduced framework can handle multi-sized images during training. We show the effectiveness of the introduced framework and compare its performance with the intuitive approach through the example of the classification of three-dimensional digital porous media of varying sizes.
Conference Paper
Full-text available
Generative Adversarial Networks (GANs) have been widely used for generatingphoto-realistic images. In this work, we develop physics-informed meth-ods for generative enrichment of turbulence. We incorporate a physics-informedlearning approach to minimize the residuals of the governing equations for thegenerated data. We analyze two physics-informed models including a GAN model,and show that they outperform tricubic interpolation. We also show that usingphysics-informed learning can significantly improve the model’s ability to gener-ate data that satisfies the physical constraints. Finally, we analyze the generatedenriched data to show that it is able to recover statistical metrics of the flow fieldincluding energy metrics.
Article
Full-text available
We present a novel deep learning framework for flow field predictions in irregular domains when the solution is a function of the geometry of either the domain or objects inside the domain. Grid vertices in a computational fluid dynamics (CFD) domain are viewed as point clouds and used as inputs to a neural network based on the PointNet architecture, which learns an end-to-end mapping between spatial positions and CFD quantities. Using our approach, (i) the network inherits desirable features of unstructured meshes (e.g., fine and coarse point spacing near the object surface and in the far field, respectively), which minimizes network training cost; (ii) object geometry is accurately represented through vertices located on object boundaries, which maintains boundary smoothness and allows the network to detect small changes between geometries and (iii) no data interpolation is utilized for creating training data; thus accuracy of the CFD data is preserved. None of these features are achievable by extant methods based on projecting scattered CFD data into Cartesian grids and then using regular convolutional neural networks. Incompressible laminar steady flow past a cylinder with various shapes for its cross section is considered. The mass and momentum of predicted fields are conserved. We test the generalizability of our network by predicting the flow around multiple objects as well as an airfoil, even though only single objects and no airfoils are observed during training. The network predicts the flow fields hundreds of times faster than our conventional CFD solver, while maintaining excellent to reasonable accuracy.
Article
Full-text available
Segmentation of 3D micro-Computed Tomographic (uCT) images of rock samples is essential for further Digital Rock Physics (DRP) analysis, however, conventional methods such as thresholding and watershed segmentation are susceptible to user-bias. Deep Convolutional Neural Networks (CNNs) have produced accurate pixelwise semantic (multi-category) segmentation results with natural images and uCT rock images, however, physical accuracy is not well documented. The performance of 4 CNN architectures is tested for 2D and 3D cases in 10 configurations. Manually segmented uCT images of Mt. Simon Sandstone guided by QEMSCANs are treated as ground truth and used as training and validation data, with a high voxelwise accuracy (over 99%) achieved. Downstream analysis is used to validate physical accuracy. The topology of each mineral is measured, the pore space absolute permeability and single/mixed wetting multiphase flow is modelled with direct simulation. These physical measures show high variance, with models that achieve 95%+ in voxelwise accuracy possessing permeabilities and connectivities orders of magnitude off. A network architecture is introduced as a hybrid fusion of U-Net and ResNet, combining short and long skip connections in a Network-in-Network configuration, which overall outperforms U-Net and ResNet variants in some minerals, while outperforming SegNet in all minerals in voxelwise and physical accuracy measures. The network architecture and the dataset volume fractions influence accuracy trade-off since sparsely occurring minerals are over-segmented by lower accuracy networks such as SegNet at the expense of under-segmenting other minerals which can be alleviated with loss weighting. This is an especially important consideration when training a physically accurate model for segmentation.
Article
Full-text available
Abstract Convolutional neural networks (CNN) are utilized to encode the relation between initial configurations of obstacles and three fundamental quantities in porous media: porosity ( $$\varphi$$ φ ), permeability (k), and tortuosity (T). The two-dimensional systems with obstacles are considered. The fluid flow through a porous medium is simulated with the lattice Boltzmann method. The analysis has been performed for the systems with $$\varphi \in (0.37,0.99)$$ φ ∈ ( 0.37 , 0.99 ) which covers five orders of magnitude a span for permeability $$k \in (0.78, 2.1\times 10^5)$$ k ∈ ( 0.78 , 2.1 × 10 5 ) and tortuosity $$T \in (1.03,2.74)$$ T ∈ ( 1.03 , 2.74 ) . It is shown that the CNNs can be used to predict the porosity, permeability, and tortuosity with good accuracy. With the usage of the CNN models, the relation between T and $$\varphi$$ φ has been obtained and compared with the empirical estimate.
Article
Full-text available
DeePore is a deep learning workflow for rapid estimation of a wide range of porous material properties based on the binarized micro–tomography images. By combining naturally occurring porous textures we generated 17,700 semi–real 3–D micro–structures of porous geo–materials with size of 2563 voxels and 30 physical properties of each sample are calculated using physical simulations on the corresponding pore network models. Next, a designed feed–forward convolutional neural network (CNN) is trained based on the dataset to estimate several morphological, hydraulic, electrical, and mechanical characteristics of the porous material in a fraction of a second. In order to fine–tune the CNN design, we tested 9 different training scenarios and selected the one with the highest average coefficient of determination (R2) equal to 0.885 for 1418 testing samples. Additionally, 3 independent synthetic images as well as 3 realistic tomography images have been tested using the proposed method and results are compared with pore network modelling and experimental data, respectively. Tested absolute permeabilities had around 13% relative error compared to the experimental data which is noticeable considering the accuracy of the direct numerical simulation methods such as Lattice Boltzmann and Finite Volume. The workflow is compatible with any physical size of the images due to its dimensionless approach and can be used to characterize large–scale 3–D images by averaging the model outputs for a sliding window that scans the whole geometry.
Article
Full-text available
Quantitative structure-property relationships are crucial for the understanding and prediction of the physical properties of complex materials. For fluid flow in porous materials, characterizing the geometry of the pore microstructure facilitates prediction of permeability, a key property that has been extensively studied in material science, geophysics and chemical engineering. In this work, we study the predictability of different structural descriptors via both linear regressions and neural networks. A large data set of 30,000 virtual, porous microstructures of different types, including both granular and continuous solid phases, is created for this end. We compute permeabilities of these structures using the lattice Boltzmann method, and characterize the pore space geometry using one-point correlation functions (porosity, specific surface), two-point surface-surface, surface-void, and void-void correlation functions, as well as the geodesic tortuosity as an implicit descriptor. Then, we study the prediction of the permeability using different combinations of these descriptors. We obtain significant improvements of performance when compared to a Kozeny-Carman regression with only lowest-order descriptors (porosity and specific surface). We find that combining all three two-point correlation functions and tortuosity provides the best prediction of permeability, with the void-void correlation function being the most informative individual descriptor. Moreover, the combination of porosity, specific surface, and geodesic tortuosity provides very good predictive performance. This shows that higher-order correlation functions are extremely useful for forming a general model for predicting physical properties of complex materials. Additionally, our results suggest that artificial neural networks are superior to the more conventional regression methods for establishing quantitative structure-property relationships. We make the data and code used publicly available to facilitate further development of permeability prediction methods.
Article
Full-text available
Predicting the petrophysical properties of rock samples using micro-CT images has gained significant attention recently. However, an accurate and an efficient numerical tool is still lacking. After investigating three numerical techniques, (i) pore network modeling (PNM), (ii) the finite volume method (FVM), and (iii) the lattice Boltzmann method (LBM), a workflow based on machine learning is established for fast and accurate prediction of permeability directly from 3D micro-CT images. We use more than 1100 samples scanned at high resolution and extract the relevant features from these samples for use in a supervised learning algorithm. The approach takes advantage of the efficient computation provided by PNM and the accuracy of the LBM to quickly and accurately estimate rock permeability. The relevant features derived from PNM and image analysis are fed into a supervised machine learning model and a deep neural network to compute the permeability in an end-to-end regression scheme. Within a supervised learning framework, machine and deep learning algorithms based on linear regression, gradient boosting, and physics-informed convolutional neural networks (CNNs) are applied to predict the petrophysical properties of porous rock from 3D micro-CT images. We have performed the sensitivity analysis on the feature importance, hyperparameters, and different learning algorithms to make a prediction. Values of R2 scores up to 88% and 91% are achieved using machine learning regression models and the deep learning approach, respectively. Remarkably, a significant gain in computation time—approximately 3 orders of magnitude—is achieved by applied machine learning compared with the LBM. Finally, the study highlights the critical role played by feature engineering in predicting petrophysical properties using deep learning.
Article
Full-text available
Many hyperparameters have to be tuned to have a robust convolutional neural network that will be able to accurately classify images. One of the most important hyperparameters is the batch size, which is the number of images used to train a single forward and backward pass. In this study, the effect of batch size on the performance of convolutional neural networks and the impact of learning rates will be studied for image classification, specifically for medical images. To train the network faster, a VGG16 network with ImageNet weights was used in this experiment. Our results concluded that a higher batch size doesn’t usually achieve high accuracy, and the learning rate and the optimizer used will have a significant impact as well. Lowering the learning rate and decreasing the batch size will allow the network to train better, especially in the case of fine-tuning.
Article
Full-text available
We propose a conservative physics-informed neural network (cPINN) on discrete domains for nonlinear conservation laws. Here, the term discrete domain represents the discrete sub-domains obtained after division of the computational domain, where PINN is applied and the conservation property of cPINN is obtained by enforcing the flux continuity in the strong form along the sub-domain interfaces. In case of hyperbolic conservation laws, the convective flux contributes at the interfaces, whereas in case of viscous conservation laws, both convective and diffusive fluxes contribute. Apart from the flux continuity condition, an average solution (given by two different neural networks) is also enforced at the common interface between two sub-domains. One can also employ a deep neural network in the domain, where the solution may have complex structure, whereas a shallow neural network can be used in the sub-domains with relatively simple and smooth solutions. Another advantage of the proposed method is the additional freedom it gives in terms of the choice of optimization algorithm and the various training parameters like residual points, activation function, width and depth of the network etc. Various forms of errors involved in cPINN such as optimization, generalization and approximation errors and their sources are discussed briefly. In cPINN, locally adaptive activation functions are used, hence training the model faster compared to its fixed counterparts. Both, forward and inverse problems are solved using the proposed method. Various test cases ranging from scalar nonlinear conservation laws like Burgers, Korteweg–de Vries (KdV) equations to systems of conservation laws, like compressible Euler equations are solved. The lid-driven cavity test case governed by incompressible Navier–Stokes equation is also solved and the results are compared against a benchmark solution. The proposed method enjoys the property of domain decomposition with separate neural networks in each sub-domain, and it efficiently lends itself to parallelized computation, where each sub-domain can be assigned to a different computational node.