stochastic shale permeability matching: three-dimensional characterization … · 2020. 6. 17. ·...

12
Stochastic shale permeability matching: Three-dimensional characterization and modeling Pejman Tahmasebi a,b, , Farzam Javadpour a , Muhammad Sahimi b a Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, 78713 Austin, TX, USA b Mork Family Department of Chemical Engineering and Materials Science, University of Southern California, Los Angeles, CA 90089-1211, USA abstract article info Article history: Received 20 June 2016 Received in revised form 22 August 2016 Accepted 26 August 2016 Available online 29 August 2016 The recently-discovered promise of shale-gas reservoirs necessitates their characterization. As a rst step, two- dimensional (2D) scanning electron microscope (SEM) imaging can provide excellent view of the complexity as- sociated with such reservoirs, and has been used in recent years to study organic materials, clays, minerals, and the pores in shales. Other important information, such as the connectivity of the pores in 3D and such macroscop- ic properties as the permeability are difcult to infer from 2D SEM images. Newer techniques, such as focused ion-beam SEM (FIB-SEM) have been utilized to overcome the shortcomings. The extremely small sample size and the costs associated with the FIB-SEM technique limit, however, wide use of the FIB-SEM in characterization of shale samples. An alternative approach is to use the recently developed advanced algorithms for 3D recon- struction that use one or a few 2D samples. Such methods may not, however, be able to fully reproduce the shale permeability. To accurately reproduce the permeability, we propose a new method based on a combination of the Metropolis-Hastings and the genetic algorithms. The new method learns from its own previously generat- ed realizations of the shale and produces models that match the existing permeability data. The method is vali- dated with the measured permeability for an actual 3D shale sample. It generates an ensemble of stochastic realizations that honor the permeability data, which may then be used for more accurate characterization of shale gas reservoir and analysis of their pore network. Published by Elsevier B.V. Keywords: Shale-gas Characterization Stochastic modeling SEM image Shale permeability 1. Introduction Due to their special intrinsic physical characteristics, such as multi- modal nano-scale structures and complex features, shale formations are of much interest in various studies in earth science, including geol- ogy, water resources, oil/gas reservoirs, CO 2 sequestration, and many other related subsurface systems and phenomena. Owing to their im- mense hydrocarbon content as unconventional reservoirs, shales have recently become even more important. The size of the pores in shales is a distinct aspect of such formations that gives rise to complex mor- phology and connected ow paths. Such multiscale pore structures have considerable inuence on the methodology and tools for studying shale reservoirs in terms of uid ow, imaging, pore network modeling, and interpretation of the data (Javadpour, 2009; Sahimi, 2011). Permeability is perhaps the most critical property for evaluating shale gas reservoirs. A precise evaluation of the permeability helps one to un- derstand how the pore networks and important petrophysical parame- ters in unconventional reservoirs differ from those in the conventional ones. Unlike conventional reservoirs, gas ow in shale gas reservoirs is de- termined by various properties, including the multiscale structure of the reservoirs, their nano-porosity, and special ow mechanisms. Therefore, a considerable number of samples are needed to understand the variabil- ity and complexity of the features that determine the permeability. Accurate 2D and 3D imaging is a prominent tool for better under- standing of the complexities of shale gas reservoirs (Javadpour et al., 2012; Loucks et al., 2012). Such images provide excellent inside look into the depositional environment, mineralogy, maturation, thermal condition, total organic carbon (TOC), strain/stress properties, porosity, permeability, and the pore network of the formation. Thus, deeper stud- ies of such reservoirs inevitably require a large number of samples using two- or three-dimensional (3D) images. Due to its use of various signals, high-resolution focused ion-beam scanning electron microscope (FIB- SEM) method (Lemmens et al., 2011) offers images that contain very useful information about the morphology and composition of shale samples (Loucks et al., 2012). Such images also provide a means for ac- curate evaluation of the complex connectivity in the pore network of shale reservoirs. Such key parameters as the TOC, mineralogy, porosity, permeability, and other petrophysical properties are best evaluated through the use of such images. The necessity of using several 3D sam- ples to represent the heterogeneity and complexities in shale gas reser- voirs was investigated by Kelly et al. (2015), who concluded that a vast number of samples are needed to improve the likelihood of a reliable as- sessment of the various petrophysical parameters. Obtaining such a large number of 3D images is, however, neither feasible nor practical. International Journal of Coal Geology 165 (2016) 231242 Corresponding author at: Department of Petroleum Engineering, University of Wyoming, Laramie, WY 82071, USA. E-mail address: [email protected] (P. Tahmasebi). http://dx.doi.org/10.1016/j.coal.2016.08.024 0166-5162/Published by Elsevier B.V. Contents lists available at ScienceDirect International Journal of Coal Geology journal homepage: www.elsevier.com/locate/ijcoalgeo

Upload: others

Post on 18-Sep-2020

9 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

International Journal of Coal Geology 165 (2016) 231–242

Contents lists available at ScienceDirect

International Journal of Coal Geology

j ourna l homepage: www.e lsev ie r .com/ locate / i j coa lgeo

Stochastic shale permeability matching: Three-dimensionalcharacterization and modeling

Pejman Tahmasebi a,b,⁎, Farzam Javadpour a, Muhammad Sahimi b

a Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, 78713 Austin, TX, USAb Mork Family Department of Chemical Engineering and Materials Science, University of Southern California, Los Angeles, CA 90089-1211, USA

⁎ Corresponding author at: Department of PetroleuWyoming, Laramie, WY 82071, USA.

E-mail address: [email protected] (P. Tahmasebi).

http://dx.doi.org/10.1016/j.coal.2016.08.0240166-5162/Published by Elsevier B.V.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 20 June 2016Received in revised form 22 August 2016Accepted 26 August 2016Available online 29 August 2016

The recently-discovered promise of shale-gas reservoirs necessitates their characterization. As a first step, two-dimensional (2D) scanning electronmicroscope (SEM) imaging can provide excellent view of the complexity as-sociated with such reservoirs, and has been used in recent years to study organic materials, clays, minerals, andthe pores in shales. Other important information, such as the connectivity of the pores in 3D and suchmacroscop-ic properties as the permeability are difficult to infer from 2D SEM images. Newer techniques, such as focusedion-beam SEM (FIB-SEM) have been utilized to overcome the shortcomings. The extremely small sample sizeand the costs associatedwith the FIB-SEM technique limit, however, wide use of the FIB-SEM in characterizationof shale samples. An alternative approach is to use the recently developed advanced algorithms for 3D recon-struction that use one or a few 2D samples. Such methods may not, however, be able to fully reproduce theshale permeability. To accurately reproduce the permeability, we propose a newmethod based on a combinationof theMetropolis-Hastings and the genetic algorithms. The newmethod learns from its own previously generat-ed realizations of the shale and produces models that match the existing permeability data. The method is vali-dated with the measured permeability for an actual 3D shale sample. It generates an ensemble of stochasticrealizations that honor the permeability data, which may then be used for more accurate characterization ofshale gas reservoir and analysis of their pore network.

Published by Elsevier B.V.

Keywords:Shale-gasCharacterizationStochastic modelingSEM imageShale permeability

1. Introduction

Due to their special intrinsic physical characteristics, such as multi-modal nano-scale structures and complex features, shale formationsare of much interest in various studies in earth science, including geol-ogy, water resources, oil/gas reservoirs, CO2 sequestration, and manyother related subsurface systems and phenomena. Owing to their im-mense hydrocarbon content as unconventional reservoirs, shales haverecently become even more important. The size of the pores in shalesis a distinct aspect of such formations that gives rise to complex mor-phology and connected flow paths. Such multiscale pore structureshave considerable influence on the methodology and tools for studyingshale reservoirs in terms of fluid flow, imaging, pore networkmodeling,and interpretation of the data (Javadpour, 2009; Sahimi, 2011).

Permeability is perhaps themost critical property for evaluating shalegas reservoirs. A precise evaluation of the permeability helps one to un-derstand how the pore networks and important petrophysical parame-ters in unconventional reservoirs differ from those in the conventionalones. Unlike conventional reservoirs, gasflow in shale gas reservoirs is de-termined by various properties, including the multiscale structure of the

m Engineering, University of

reservoirs, their nano-porosity, and special flow mechanisms. Therefore,a considerable number of samples are needed to understand the variabil-ity and complexity of the features that determine the permeability.

Accurate 2D and 3D imaging is a prominent tool for better under-standing of the complexities of shale gas reservoirs (Javadpour et al.,2012; Loucks et al., 2012). Such images provide excellent inside lookinto the depositional environment, mineralogy, maturation, thermalcondition, total organic carbon (TOC), strain/stress properties, porosity,permeability, and the pore network of the formation. Thus, deeper stud-ies of such reservoirs inevitably require a large number of samples usingtwo- or three-dimensional (3D) images. Due to its use of various signals,high-resolution focused ion-beam scanning electron microscope (FIB-SEM) method (Lemmens et al., 2011) offers images that contain veryuseful information about the morphology and composition of shalesamples (Loucks et al., 2012). Such images also provide a means for ac-curate evaluation of the complex connectivity in the pore network ofshale reservoirs. Such key parameters as the TOC, mineralogy, porosity,permeability, and other petrophysical properties are best evaluatedthrough the use of such images. The necessity of using several 3D sam-ples to represent the heterogeneity and complexities in shale gas reser-voirs was investigated by Kelly et al. (2015), who concluded that a vastnumber of samples are needed to improve the likelihood of a reliable as-sessment of the various petrophysical parameters. Obtaining such alarge number of 3D images is, however, neither feasible nor practical.

Page 2: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

Nomenclature

AUX Auxiliary dataCD Covariance of experimental dataCm Covariance of shale samples parametersCCF Cross-correlation functionCCSIM Cross-correlation based simulationdexp Experimental dataDT(u) Data event as position udJS Distance between two models using the JS methodDI Digital imageE Euclidean distanceG Simulation gridGA Genetic algorithmJS Jensen-ShannonM Porous media modelmcand Candidate modelmprior Prior modelMCMC Markov Chain Monte CarloMH Metropolis-HastingsmutStep Flag for mutationOL Overlap regionP Size in x directionQ Size in y directionR Size in z directionS Least-squares misfit functionSD Soft dataSEM Scanning electron microscopeT TemplateTOC Total organic carbonμ Mean or expectation of the distributionρ Acceptance probabilityσ Standard deviationω Weight

232 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

On the other hand, 2D images can beprovidedwith ease and at a low cost.Because of the obstacles involvedwith obtaining 3D images, Tahmasebi etal. (2012, 2015a) proposed a new approach based on taking one or more2D images and building a 3Dmodel stochastically. The models generatedby their approach exhibit petrophysical and flow properties similar tothose directly observed and measured in actual samples. Their methodproduces an ensemble of 3D realizations that provide acceptable approx-imation of the same properties in the 2D image(s). We should point outthat, in addition to our method, there are several other important tech-niques that can reproduce the connectivities at small sales (see, for exam-ple, Adler et al., 1990; Roberts and Teubner, 1995; Yeong and Torquato,1998a, 1998b; Jiao et al., 2009; Mehmani et al., 2013; Gerke andKarsanina, 2015; Gao et al., 2015). But, they have never been tested forreconstructing models of shale reservoirs.

Aside from recent progress in producing various equi-probable 3Drealizations of shales (Tahmasebi et al., 2015a), reproduction of accuratepermeability for complex structures remains an essentially open prob-lem (Veselý et al., 2015; Gao et al., 2015). Indeed, it is perhaps excessiveexpectation of reproducing the existing experimental data using thecurrent 3D stochastic models, when they are generated from one orvery few 2D images. Intuitively, we know that 2D images can hardlyconvey, for example, a flow property that depends on the spatial poreconnectivity in 3D. Furthermore, since one cannot study all the core

data and samples due to the limitations of cost and time-consuming na-ture of the task, stochastic modeling can generate many samples, quan-tify the uncertainties, and put forward a range of variability. Clearly, theinitial data - here the input 2D images - should represent the variabilityin the structures, in which case the stochastic methods produce variouspossible realizations in a matter of a few CPU seconds.

Building a 3Dmodel of porousmedia based on limited data has beeninvestigated extensively. Suchmodelsmay be divided into three groups.In one group, called object-based techniques, the pore and grain struc-tures are treated as a set of objects that are defined based on the priorknowledge of the pore space (Pyrcz and Deutsch, 2014). Clearly, suchtechniques are not applicable to random structures. Instead, the secondgroup, called pixel-based methodologies, can be used effectively to pro-duce the shapes that are hard to fit to a specific regular object(Strebelle, 2002; Okabe and Blunt, 2004). Unlike the object-basedmethods, however, these techniques are unable to produce very realis-tic pore structures. Thus, onemust resort to the third class of such tech-niques, called process-based models (Bryant and Blunt, 1992; Coelho etal., 1997; Biswal et al., 1999, 2007; Øren and Bakke, 2002) that producevery realistic models by mimicking the real processes that form the po-rousmedia. They are not, however, general enough as they are designedfor specific sedimentary conditions appropriate for a given reservoirand, thus, need to be defined separately for every new type of reservoir.Furthermore, they also require highly intensive computations.

As a very effective alternative, a combination of the object- andpixel-basedmethods can producemore realisticmorphology for porousmedia (Tahmasebi et al., 2012). In other words, the strength of pixel-based method for producing complex pore space and object-basedmethods for realistic models are integrated. For this aim, one of the re-cent techniques in this class of methods (Tahmasebi et al., 2015a) is im-plemented in the present study.

None of the previously described methodologies is capable of repro-ducing accurately the experimental data for the permeability and electri-cal conductivity. They are able to produce the spatial structure of the porespace of porous media, but reproducing the data for the flow and trans-port properties is not guaranteed. This is the reason why the stochasticmethods produce an ensemble of realizations and, consequently, thepetrophysical properties that are used for uncertainty quantification.They produce a range of, for example, permeability, instead of one singlevalue. In this paper, a new iterative stochasticmethod for reproducing thepermeability from 2D SEM images is proposed. Using concepts from thegenetic algorithm (GA) and data mining, the new method uses various2D images from different parts of a shale gas reservoir in order to itera-tively improve the estimation of the permeability. To this end, the accept-ed realizations are pooled together to make a set of soft data (SD) onwhich the subsequent models are conditioned. The pool of the realiza-tions and the SD are updated progressively, building upon the previous-ly-accepted realizations. As with the optimization problems, the initialguesses (i.e. models)may produce results very different from the existingexperimental data. At the later steps, however, as the pool of the realiza-tions and the SD are updated, the calculated attributes draw closer to theactual measured experimental data. Thus, this paper presents a frame-work by which the experimental properties can be reproduced more orless exactly.

The rest of this paper is organized as follows. In the next section wediscuss the methodology that we propose in this paper, including an al-gorithm that was recently suggested and utilized in the present paper,and how it is integrated with the GA. The method is then validated bytwo examples, a 2D synthetic model of shales and a 3D example basedon an actual shale sample.

2. Methodology

Major obstacles to the development of shale reservoirs are their costly analysis and studies that point to the need for the development of moreefficient technologies. As mentioned earlier, a basic and common type of study is conducting extensive analysis of various aspects of a large numberof samples, including theirmineralogy and someessential petrophysical and laboratory properties. Permeability is a key property that is evaluated for

Page 3: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

233P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

providing a realistic assessment of fluid flow in shale formations. Therefore, the permeability data are usually available formany samples of shale gasreservoirs.

At the same time, detailed imaging provides a direct view of a shale reservoir. High-quality 3D images are usually not affordable, however, andthus there is always a trade-off between the cost and amount of information that can be acquired. Generally speaking, most methods of predictingthe permeability are based on empirical equations that try to bridge the experimental data and the petrophysical parameters extracted from the SEMimages. Such techniques usually lead to a single equation that tries tomodel the complexity and heterogeneity of shale reservoirs, which is, however,not realistic. In addition, such equationsusually differ for different reservoirs, and cannot utilize all the information gathered, including the qualitative(cross-sectionmaps, SEM images, etc.) and quantitative (core data, experimental results, etc.) data. Thus, important questions arise: (i) how can suchdata be integrated, as the quantitative experimental data for, for example, the permeability and the SEM images are entirely different, and (ii) howcan themaximumamount of information be extracted from the SEM images and be anchored to the quantitative data, or vice versa?Wewill addressthe two questions and present a method that reduces the computation cost, using an efficient approach.

The methodology proposed in this paper is cast on a recently-developed method called the cross correlation-based simulation (CCSIM). TheCCSIM algorithm (Tahmasebi et al., 2012) is an efficient stochastic approach that utilizes 2D/3D images and generates 3D models or realizations ofa given porous formation. Then, the algorithm is integratedwith an optimization and training framework to generate an ensemble of 3D realizationsof shale that reproduce the experimental results. The realizations are used to investigate the flow properties of complex pore networks in shale-gassamples.

2.1. The CCSIM algorithm

The CCSIMalgorithm is based on high-order statistics and, unlike the existingmethods of shale characterization, extracts themaximumamount ofinformation from an image or data set in order to reconstruct the shale samples. Indeed, lower-order statistical methods, such as covariance-basedapproaches, are unable to characterize the complex patterns and, consequently, cannot generate accurate models. Complex and multiscale patternsin unconventional reservoirs requiremethods that can reproduce such features. Aside fromobtainingmore information, however, it is also importantto uncover the physical processes within the reservoirs that have contributed to the formation of their morphology and, thus, are also of great im-portance for accurate characterization of shales. The understanding that such physical processes and complexities produce are then translatedinto a digital image (DI) that is widely available from many shale samples.

Briefly, the CCSIM algorithm is similar to a concept in problem-solving inwhich amodel is generated using the similarity between the pieces (i.e.,the patterns) in theDI. In otherwords, sincewe aremore interested in the uncertainty that exist in any such data due to lack of complete information,an exact match between the patterns in the DI and the model is not the goal of the CCSIM, rather generating multiple stochastic realizations of thesame DI is pursued.

Mathematically speaking, most of the high-order statistical methods are based on the Euclidean distance, i.e. a function that characterizes the dif-ferences between the DI and a sample of the data taken from the DI. Computations with such a distance function is, however, is very costly. This mo-tivated Tahmasebi et al. (2012) to propose a new methodology based on a cross-correlation function (CCF) for which the computation takes verylittle CPU time. For a complex DI represented by an image, the CCF yieldsmuchmore accurate realizations of the image, because it takes into accountits higher-order statistics. Physically, the CCF provides a measure of similarity to a dataset in a realization of the DI.

We describe the CCF and the CCSIMalgorithm for 2D systems. Its extension to 3Dmedia is straightforward. Let G represent the computational gridused in themodeling. The templates—the blocks of G— bywhich the DI is reconstructed are denoted by T, while DT(u) is the data event at position uin T. Every two neighboring blocks share an overlap (OL) region. Use of “data event” instead of “data” is preferred because during the modeling, DT

changes as the algorithm progresses. Finally, the OL regions represent the overlap segment between neighboring templates. Suppose that DI (x,y)represents the datum at point (x,y) of a DI of size Tx × Ty, with x ∈ {0,…,Tx−1} and y ∈ {0,…,Ty−1}. Examining the DI, we focus on a portionDT(u) of size OLx × OLy and regenerate it based on the data such that it matches the corresponding portion in the DI. The matching is based onthe CCF, defined by

C i; jð Þ ¼XOLx−1

x¼0

XOLy−1

y¼0

DI xþ i; jð ÞDT x; yð Þ ð1Þ

WhereDT(x,y)is the data event at point (x,y). Eq. (1) indicates that the desired position of (i,j)—the bestmatchwith the DI—is one thatmaximizesCði; jÞ. If the DI is not too large - typically images smaller than 700× 700 pixels - all the computations are carried out in the spatial domain (Tahmasebiet al., 2012; Tahmasebi and Sahimi 2016a,b). When the DI is very large, the calculations of the CCF, which is between the DI and DT, in the Fourierspace reduce the CPU time very significantly (Tahmasebi et al., 2012). The CCSIM algorithm is illustrated in Fig. 1, which shows how a new patternis inserted in G.

2.2. Three-dimensional reconstruction

The CCSIM, in its original form, can be used for direct 2D-to-2D and 3D-to-3Dmodeling. For example, it can generate a 3Dmodel using a 3DDI. Inwhat follows, we describe an extension of the CCSIM algorithm that uses a few 2D DIs to reconstruct a 3D model. Intuitively, 3D (p × q × r) porousmedia may be thought of as a series of r 2D planes of size (p × q) that are stacked together in a given direction. In most stationary porous media(Tahmasebi and Sahimi, 2015a), one can identify a “vertical” connectivity perpendicular to the 2D images that are used. Thus, one can developmodels of 3D porous media by reconstructing many interdependent 2D layers (planes) using the CCSIM algorithm and the given DI, and thenstack them together to form the final 3D model. To preserve the external continuity, we first reconstruct the orthogonal external frames using theCCSIM and the DI, and then fill the internal part of the system layer by layer. In this method, the first layer is reconstructed using the DI (in fact,we can use the 2D DI itself as the first layer), and reconstruct the rest layer by layer using the conditional CCSIM, one in which some of the datamust be honored exactly. This means, in the present context, that each time we reconstruct a new layer, some of the data in the previously-recon-structed layer must be honored exactly in order to preserve the connectivity between the two layers. Thus, to preserve the continuity of the system,the first layer must be conditioned to its four external vertical edges. Then, to reconstruct the second layer, it is conditioned to both the external

Page 4: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

(a)

(b)

(c)

OL vertical OL horizontal

Candidate pattern

Zoom-in view (before implanting the candidate pattern)

Optimal region Optimal pattern

Zoom-in view (after implanting the candidate pattern)

Simulation grid (G)

Fig. 1. —(a) The DI in which the organic matters and their corresponding pore spaces are shown by dark gray and black, respectively. (b) The process in which a candidate pattern isinterlocked into G. The overlap regions are selected first (the vertical and horizontal bands; OLvertical and OLhorizontal). As a rule of thumb, 1/4 - 1/5 of the size of the DT is selected forthe overlap dimensions in both vertical and horizontal directions. Then, the CCF is computed and the candidate pattern is identified and inserted in the grid. (c) One completed realization.

234 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

frames and the previous layer (in this instance, the first layer). Conditioning in this case is performed through point data that are extracted from theprevious layer. Such point data are identified using the Shannon entropy based on the entropy of the previous layer. These steps are iterated until theuppermost predefined layer is reached and the 3D model is complete (Tahmasebi et al., 2015a). It should be noted that using just one single 2D DIimplies isotropy assumption in the porousmedium. Anisotropy can still be considered if one usesmore images in different directions. In suchmodel-ing, each 2D image represents a specific directional anisotropy. For the sake of simplicity, isotropy assumption is made in this study.

Page 5: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

235P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

Our previous work (Tahmasebi et al., 2012) indicated that, though highly efficient and in many cases very accurate, the method might generatesome artifacts if the DIs are very complex and host multiscale features. To alleviate this problem and increase the quality of the realizations, we in-tegrate an iterative algorithm with the CCSIM in such a way that it removes the unwanted patchiness and discontinuities (Tahmasebi et al., 2015;Tahmasebi and Sahimi, 2015b). First, a 3D realization is generated using the CCSIM algorithm that may contain some artefacts and patchiness. Thepatchiness is identified using the mismatches between the overlaps when two of them are interlocked in the 3Dmodel. Then, based on the initiallygenerated realization, the voxels of the 3D simulation grid are visited one-by-one. Three perpendicular 2D planes of size (T× T) are extracted at eachvoxel. Then, the similarity between each extracted plane and theDI is evaluated using the CCF. A set of candidate patterns are then identified for eachplane. Next, the final value of the visiting voxel is defined based on the Jensen-Shannon equation (see below). To this end, the histograms of the 3Dmodel for all of the candidate patterns are constructed and their distances to the DI are calculated. Finally, the value of the pattern thatminimizes theaforementioned distance is selected and inserted in the 3D model. This algorithm iteratively removes the patchiness and artefact in the original 3Dmodel (Tahmasebi et al., 2015b).

2.3. Workflow

Several factors affect performance of shales and its forecasting. Contributing to the uncertainty is lack of extensive data and accurate analysis ofthe existing samples. Improving our understanding of such factors reduces investment risks, leading to more accurate forecasts. Thus, having a set ofacceptable realizations of the shale samples that have the same permeability as the actual ones provides a reliable view of such complex reservoirs.Another advantage of having satisfactory realizations is their usefulness in assessing the uncertainty and the connectivity of the macroscopic porenetwork that strongly affects gas production from shales. Analyzing a large number of realizations is not, however, feasible, and indeed, is very costly.

The various data sets available for shale reservoirs make it difficult to establish a generic workflow that can efficiently integrate them. For exam-ple, core and seismic data in such reservoirs provide valuable information about the porosity, mineralogy, petrophysical properties, and large-scalestructures. Themultivariate CCSIM can be used to integrate such data with the shale-gasmodels. Integration of the experimental data for such prop-erties as the permeability and electrical conductivity requires, however, an iterative process that includes generating various samples and evaluatingthe important parameters using forward modeling. The iterative process is very demanding computationally, however, and is not similar to and assystematic as an optimization algorithm. In addition, producing a large number of realizations and forward modeling is tedious, even computation-ally prohibitive, taking days or weeks. Therefore, keeping low the number of required realizations and, consequently, the number of forward simu-lations is critical. Moreover, due to the nonlinear relationship between the shale samples and the responses of their forward modeling, differentmodels can produce very similar responses, creating uncertainty.

The objective here — to produce many realizations and then iteratively alter them to better match the experimental data for the directly-mea-sured permeability and electrical conductivity — can be viewed as an inverse problem in which several realizations are first generated and then afew forward simulations are performed (Tarantola, 2005). One useful approach for this objective is through Bayes' rule by which the uncertaintyis quantified using a posterior distribution of acceptable realizations. The posterior distribution of the shale realizations may be thought of as theproduct of a likelihood function that measures the similarity between the responses of forward modeling and the experimental data. The existinginformation on the shale samples' parameters is presented by the DIs. Note that Bayes' formulation allows simultaneous integration of differenttypes of data. Mathematically speaking, a distribution can be rebuilt when the entire space is explored, a process that can be thought of as generatingnumerous realizations that fall within the posterior distribution. Clearly, sampling complex, multiscale, and non-Gaussian properties of shales is notstraightforward. Thus, another objective of this study is developing an efficient and “intelligent” sampling method for such posterior distribution.

To achieve our objective, we split the problem into a series of iterative processes that minimize the differences between the experimental andcalculated data, while the generated realizations preserve the defined parameters of the existing information contained in the DIs. A combinationof Markov Chain Monte Carlo (MCMC) (Robert and Casella, 2013), the genetic algorithm (GA) as a stochastic optimization tool, and a training algo-rithm converge to the global minimum of an “energy function” that indicates convergence to the true model. The combined techniques reduce theCPU time of the MCMC by reducing its high rejection rate. The final algorithm is efficient and the acceptance rate of the generated realizations in-creases gradually, while the CPU time decreases. Let us describe our method.

As stated, we use Bayes' rule to update data on the shale properties based on the experimental data dexp, such as the effective permeability. It isalso assumed that the existing information is integrated into the shale realizations that are generated using the CCSIM. Thus, in a Bayesian context,the posterior probability is expressed by

f mjd exp� � ¼ f d expjm

� �f mð Þ

f d exp� � ð1Þ

or, more simply,

f m d exp��� �

∝ f d exp mj� �

f mð Þ ð2Þ

where f(m |dexp) is the posterior probability distribution, f(dexp|m) is the likelihood function, and f(m) indicates the prior probability distribution forthe porous media. Solving Eqs. (1) and (2) entails drawing a considerable number of samples from the posterior distribution, which can be donewhen the other two components, the prior f(m) and the likelihood f(dexp|m), are known. Sampling is not, however, straightforward, and for this rea-son the MCMC as an efficient sampling technique is employed to draw samples from f(dexp|m).

In some cases, however, onemay also take experimental errors into account. Assuming a Gaussian distribution for the prior model and the error,Eq. (2) is rewritten as:

f mjd exp� �

∝k exp −S mð Þ½ � ð3Þ

Page 6: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

236 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

where k is a normalization constant, and S(m) is the least-squares misfit function, defined by:

S mð Þ ¼ 12

f mð Þ−d exp� �tC−1

D f mð Þ−d exp� �þ m−mprior

� �tC−1m m−mprior

� �h ið4Þ

where CD is the covariance of the experimental data, f(m) is the response of the generated shale realizationm that must be obtained using forwardmodeling,mprior is the prior model or original DI, and Cm is the covariance of shale samples' parameters. Note that the 3Dmodel should be first gen-erated since the original DI is in 2D. This part is further discussed in the next section.

The MCMC is a technique for sequentially drawing realizations m from a probability distribution. The result after a number of iterations isused as a sample of the target distribution f(m |dexp). A new realization is drawn based on the previously drawn medium in such a way that itsequilibrium probability is the target distribution. The most important advantage of this method is its improvement of the equilibrium distri-bution at each step as it converges to the target distribution. This method is used in this study because the target distribution, f(m |dexp), is notavailable, and the MCMC can be used to build or approximate it. Here, however, due to its efficiency, the Metropolis-Hastings (MH) algorithm,a member of the MCMC family, is used for sampling and approximating the target distribution (Metropolis et al., 1953; Hastings, 1970). Withthe MH method, we begin with generating one realization using the prior distribution data. Then, a candidate realization from the proposeddistribution is generated as the next sample, and its acceptance probability is calculated using the proposed distribution and the full joint prob-ability density.

Clearly, theMHalgorithm is computationally expensive, requiring thousands of realizations to build the target distribution. The forwardmodelingis also excessively time consuming. Thus, the candidate realizations must be produced in a way that accelerates the convergence and increases theacceptance rate. For this objective,we use the CCSIMalgorithmwith its advantage in terms of using auxiliary data, i.e. soft data. Similar to the learningrules of neural networks, all the accepted realizations are placed in a pool fromwhich the algorithm learns about the previous realizations and theirresponses, whichmakes it more productive. In addition, other concepts, such asmutation used in the GA, are also integrated. Mutation in the GA is acorrective step: After the realizations are generated, it is possible that they exhibit some kind of “defect”, which exhibits itself by producing an ob-jective function, to beminimized,with larger numerical value than expected. Such “defects” in the realizations that are, for example, unusual patternsthat do not match the data are known as mutation.

We use the multivariate CCSIM algorithm, which requires the DI, soft data (SD), and an auxiliary variable (AUX). Briefly, the CCSIM uses the SD,which in essence constitutes a probability map, as the conditioning data. The SD creates a continuous and smooth map that is different from theexisting SEM images. Therefore, the SD and the DI must be linked in order to condition the output model. To construct such a link, we need moredata in the form of an auxiliarymap, which is generated using an inverse CCSIM, i.e. one in whichwe exchange the DI with the SD, and use the latteras the DI. In particular, the AUX data are the output of the CCSIMwhen the DI and the SD are switched. The algorithm commences by generating 3Drealizations using the given 2D DIs. Next, forwardmodeling is performed, e.g. simulation of flow and electrical conduction. For each new realization,the computed responses in the presence of an error are examined using the following equation, by which the probability of accepting any new re-alization compared to the previously-accepted one is tested:

ρ mn;m n−1ð Þ� � ¼ min 1;

exp −12

f mnð Þ−d exp� �tC−1

D f mnð Þ−d exp� �þ mn−mprior

� �tC−1m mn−mprior

� �� �� �

exp −12

f m n−1ð Þ� �

−d exp� �tC−1

D f m n−1ð Þ� �

−d exp� �þ m n−1ð Þ−mprior

� �tC−1m m n−1ð Þ−mprior:

� �� �� �8>><>>:

9>>=>>; ð5Þ

The candidate realization is accepted with probability ρ or rejected with probability 1−ρ. The newmodel, n, is accepted if it provides a more ac-curate response than the previously accepted realization, n− 1. If it is accepted, the new 3Dmodel is added to the pool of the realizations. Each ac-cepted realization is given a weight using the following equation,

ωi ¼ exp − f mið Þ−d exp� �tC−1

D f mið Þ−d exp� �þ mi−mprior

� �tC−1m mi−mprior

� �� �h ið6Þ

The weight defines the proximity of the permeability and conductivity of a stochastic realization to the correspondingmeasured data. For exam-ple, the first few accepted realizations are given a lowerweight, as the subsequently accepted ones provide a better match. As mentioned, some con-cepts of the GA, such asmutation, are also used to avoid trapping in a localminimum. Thus, the givenweights for the SD are changed using a Gaussiandistribution after some random steps. This provides a chance for exploring more scenarios, which is similar to the mutation operation in the GA.Assigning a weight can, however, be thought of as the crossover step in the GA in the sense that it combines the available realizations - the so-called“parent” in the GA - to make a new realization - the “children” in the GA. At the next step, the SD is updated using the new pool andweights. Clearly,the realizationswith higherweights constitute a greater part of the SD. Accordingly, the AUXdata are also updated using the new SD and the selectedrealizations. The DI is updated using the newly accepted pool of the realizations to ensure that the most informative models are used as the DI (seebelow). The steps are repeated until a predefined number of the accepted realizations converge.

Selecting the most accurate realizationm, or a set of them, directly controls the quality of the generated 3D realizations, and the rate of conver-gence to a completely plausiblemodel. One can use the 2DDIs to build the 3D realizations, but this increases the CPU time. Thus, asmentioned earlier,we propose an alternative process that uses a multiple-point histogram to construct a pattern distribution. Then, the distance d of each model m totheDI is quantified using the Jensen-Shannon (JS) divergence (Endres and Schindelin, 2003),which represents amethod formeasuring the similaritybetween two probability distributions. This data mining approach is the average of two Kullback-Leibler divergences,

dJS DIjjmð Þ ¼ 12∫þ∞−∞DIi log

DIimi

� �dxþ 1

2∫þ∞−∞mi log

mi

DIi

� �dx ð7Þ

and is always positive. Finally, a set of realizations that best maximizes the pattern distribution is selected; see Fig. 2. Altogether, by using the abovedescribed smart sampling, the number of samples that should be generated and, consequently, evaluated decreases dramatically. The algorithm issummarized in Algorithm 1.

Page 7: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

Pat.

f

Pat.

f

Pat.

f

Pat.

f

Pat.

f

Pat.

f

Pat.

fFig. 2.—Selecting the most accurate pattern distribution, or a set of them, as the DI. In this case the most accurate model is the fourth from the left. However, a combination of the second andsixth models yields the most accurate pattern distribution. The objective is to identify those models that, if combined, offer a distribution that contains most of the pattern's configurations.

237P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

3. Results and discussion

The algorithm proposed in this paper may be used when several 2DDIs and various types of experimental data are available. To better dem-onstrate the describedmethod, however, we used a single 2Dgray-scaleDI and the measured permeability data. It should be noted that the ini-tial gray-scale images were first segmented into a binary image forrepresenting thepores and grains (Andrä et al., 2013) for the permeabil-ity calculations. In other words, the segmentation is performed on thegenerated stochastic models. We then used a simple thresholdingscheme to discretize the images, which was used in the computations.It should be noted that other segmentation methods, such as the water-shed and clustering-based methods, were also implemented(Karimpouli and Tahmasebi, 2016). In this study, however, the manualthresholding scheme was used as the segmentation method. For this

Algorithm 1. Digital stochastic shale characterization.

1: Initialize minit→CCSIM(DI)2: Initialize AUXinit→movingAverage(minit)3: Initialize SDinit→mean(DI)4: Initialize mutStep5: u � U½0;1�6: for iteration n = 1, 2, …, do7: generate mcand→q(m(n)|SD,AUX,DI)8:

ρðmcand ;mðn−1ÞÞ ¼ ;min

(1; ;expð−1

2ðð f ðmnÞ−dexpÞtC−1D ð f ðmnÞ−dexpÞÞÞ

expð−12ðð f ðmðn−1Þ Þ−dexp ÞtC−1

D ð f ðmðn−1Þ Þ−dexp ÞÞÞ

)

9: if ubρ then10: j++11: m(n)←mcand

12: if (n % mutStep==0)13:

pðxjμ;σÞ ¼ 1σ

ffiffiffiffi2π

p � e−ðx−μÞ22σ2

14: end if15: else16: m(n)←m(n−1)

17: end if18: Update the SD19: Update the AUX20: Calculate the JS divergence and update the DI as illustrated in Fig. 221: end for

*x ,μ ,σ represent the mean, standard deviation and variance.

aim, based on our knowledge of pore topology of shale samples, we de-fined a series of proper thresholds by which the gray scale images wereconverted into binary media. Due to using one single image as theinput, the demarcated thresholds were utilized for the rest of modeling.

Before generating the stochastic pool of the 3D realizations, onemust determine the optimal sample size, namely, the so-called repre-sentative element volume (REV). Various methodologies for similarshale samples are discussed elsewhere (Tahmasebi et al., 2016). For ex-ample, using a very small sample generally results in increased calculat-ed permeability. On the other hand, by increasing the sample size thecalculated permeability tends to converge to a stationary value. Thus,the selected size of the stochastic realization should not be too smallto overestimate the permeability and, in a similar fashion, it shouldnot be very large to increase the computational time. Furthermore, nu-merous lower- and higher-order tests, including variogram reproduc-tion, marginal distribution and flow properties were carried out in ourprevious studies and they all produced good agreementwith the exper-imental and original datasets (Tahmasebi et al., 2016).

3.1. Two-dimensional synthetic shale

The proposed algorithm begins by using the supplied DI, AUX, andSD. For the initial iterations, the marginal probability of the DI is

Fig. 3. —(a) Reference (true) model generated using the CCSIM. (b) An initial stochasticrealization. The size of the images in (a) and (b) are 150 × 150 pixels (5 × 5 μm2).

Page 8: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

Fig. 5. —(a) The SD. (b) AUX data corresponding to the DI in Fig. 1(a).

238 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

assigned to both AUX and SD. Then, using the provided data, the CCSIMproduces a stochastic model; Fig. 3(b). Forward modeling – fluid flowsimulation to compute the permeability - is carried out. In this study,the Stokes' equation (Leal, 2007) is solved numerically for calculatingthe effective permeability:

μ∇2u−∇pþ f ¼ 0

∇ � u ¼ 0

8<: ð8Þ

where p is the pressure of thefluid, u is its velocity, and μ is the viscosity,taken to be 0.001 Pa·s. We assumed that the fluid is Newtonian and in-compressible, and the flow is creeping under steady-state conditions.The downstream and upstream pressures were assumed to be 100 kPaand 130 kPa, respectively—equivalent to a pressure gradient of 3 kPa/nm. The permeability was computed using a finite-volume schemebased on the Avizo, which simplifies the computations of flow field.After the flow field was computed, Darcy's law was used to computethe effective permeability.

The error of the computed permeability is comparedwith that of theprevious realization. The current model is accepted if its error is lessthan the previous realization's, and rejected otherwise. A weightingstep, as indicated by Eq. (6), is added to the algorithm and an appropri-ate weight is assigned to each accepted realization. After a few itera-tions, the accepted realizations create a pool, the ensemble average ofwhich reveals very useful information. A few members of the pool areshown in Fig. 4. Then, one uses such realizations and put the algorithmin a training workflow to produce other realizations with smaller errorin the absolute permeabilities. To do so, the ensemble average map ofthe accepted realizations is used as the SD that conditions the subse-quent realizations. The SD here represents the probability of occurrenceof organic matter. Using the algorithm, we update the model pool, SD,and AUX data; see Fig. 5. There is no need to update the DI aswe alreadyhave a comprehensive 2D image. This step is further discussed in thenext example.

Onemay argue that the future realizationsmaymimic themost suc-cessful (i.e. matched) previous realization with a small perturbation,since the algorithm is conditioned to the SD. To avoid such a problem,a useful concept of the GA, the mutation step, is customized here.Thus, the weight of the accepted realization is changed after specific in-tervals. In fact, the SDs are generated using a combination of theweightsand accepted realizations. For this reason, a Gaussian distribution isused to reassign the weights. Indeed, mutation breaks the previousprobability in the SD and discovers more scenarios. In other words, asthe SD leads the algorithm to converge on a shale-sample model, muta-tion provides anopportunity for exploration and testingmoremodels inthe search space.

The results of the proposed method are summarized in Fig. 6 inwhich a sequence of various stochastic models along with their corre-sponding SD and AUX data are depicted. As can be seen, the initial SDand, consequently, the AUX data represent a very smooth map causedby the high uncertainty in the initial models. As the algorithmproceeds,

Fig. 4.—Three realizations generated by Algorithm 1.

however, the SD and AUX take on more structure and become lesssmooth, and gradually converge to the model represented by the refer-ence case in Fig. 3(a). It is worthmentioning that after a certain numberof iterations, the generated stochastic models (the SD and AUX data) donot change significantly. As is shown in Fig. 6, in this example 200 real-izations were generated, each of which took 2 CPU s. Therefore, the al-gorithm was stopped when the variance of a few consecutive modelswas close to zero.

3.2. Three-dimensional real shale-gas system

The second example represents a more complex case in which a 2Dimage is used to construct 3D realizations. A series of experiments werecarried out on the sample. Nitrogen sorption (Clarkson et al., 2013, Hu etal., 2015) was measured on a powder size of 297–841 μm using acleaned sample and under no confined stress. Then, the helium porositytest (Javadpour and Ettehadtavakkol, 2015) was implemented at 30 °Cthat yielded a porosity of 3.9 ± 0.3%. Finally, pulse decay permeabilitywas performed at a confined stress of 13.8MPa at the same temperaturethat yielded an absolute permeability of 400 nD (Bhandari et al., 2015).Thus, a 3D realizationwas generated using the CCSIM algorithm; see Fig.7. Then, forward modeling in the form of flow calculations was carriedout to estimate the absolute permeability. Next, the error, the differencebetween the computed permeability and the experimental value, wascalculated. The generated realization is accepted if the result computedby Eq. (5) is less than 1 and rejected otherwise. Each new realization isaccepted if its error (in the permeability) is less than that of the previ-ously accepted realization. The accepted realizations are all placed intoa model pool and used for various purposes. The database is then usedto produce 3D SDs that are used to guide the algorithm and conditionnew realizations. As mentioned earlier, however, to differentiate the ef-fect of the accepted realizations on the SD, a weight is assigned to eachof them. The realizations with lower weights do not contribute much tothe SD and, consequently, to the subsequent 3D models.

Then, the SDs are generated using the accepted realizations and theirrelative weights. The number of the accepted realizations increases as

The sizes of the images are the same as in Fig. 3.

Page 9: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

Iteration

1 10 80 100 200

Stochastic models

SD

AUX

Fig. 6.—Gradual alteration of the realizations by the proposed algorithm. The iteration number is increased from left to right, showing how the stochasticmodel and the SD converge to thereference case of Fig. 3(a). The smoothness of the AUX data decreases as the SD become more accurate.

239P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

the algorithm proceeds, yielding more accurate conditioning data, i.e.the SD. Furthermore, the AUX data must be updated accordingly. Thus,the initial rough SD and AUX become more accurate as the algorithmconverges and the models' computed permeabilities draw closer tothe experimental data.

Two methods are suggested in this paper for generating 3D realiza-tions. One is using the initial 2D DIs, generating a 3D model, and pro-ceeding to other steps of the workflow as before. The computation

SD2D DI

+

j > λ ?

No

Yes

AUX

+

Update the AUX

End

Update the DI(s)

Build a 3D model

Pat.

f

Fig. 7.—The result produced by the proposedmethod inAlgorithm1 for the 3D example. Thewomaximum number of the realizations.

time of themethod is very high, however, due to the direct 2D-to-3D re-construction. The second approach is to generate a few acceptable 3Drealizations, select a set of the most informative ones, and use them asthe DI. Thus, 2D DIs are replaced with 3D ones, and the workflow be-comes 3D-to-3D modeling, which decreases the computation timetremendously.

The use of iteration in this study is illustrated in Fig. 8. The SD pos-sesses a very smooth structure at the beginning, indicating that

A stochastic model(m)

Forward modelingi++

Yes

NoUse the current SD

j++

Update the SD

1 2 j

Mutation?

Yes

No

Update the

< 1

rkflow shows the implementation of themethod step by step.λ represents the predefined

Page 10: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

Iteration

1 147 285

Fig. 8.—Various 3Dmodels and their corresponding SD. The SD represents the probability of occurrence of the organic matter in the model, and iteration 285 shows the locations wheresuch features most probably exist. The size of the images is 600 × 600 × 600 pixels (9 × 9 × 9 μm3).

240 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

organic matter is distributed widely. The absolute permeabilitychanges from 1200 to 500 to 401 nD after 1, 147, and 285 iterations(realizations), respectively. Altogether, the CCSIMmodel, patchinessremoval and histogram matching took 25 CPU s per realization. Itshould be noted that the permeability calculations for the 3D sam-ples were performed on graphical processing units (GPU), which re-sult in a very fast evaluation. In this example, the permeabilitycomputations took about 4 GPU min on a GeForce GTX 750TiNVIDA card. As the algorithm proceeds, the SD and, consequently,the AUX data become more accurate and less smooth. For example,the last 3D SD contains the most probable locations of the pore net-works in the 3D sample. Note that the proposed method, regardless

0

0.002

0.004

0.006

0.008

0.01

0.012

1 10 100

Pro

babi

lity

Pore Radius (nm)

Fig. 9. —Comparison of pore-throat size distribution of the original sample (black) andthose of several 3D realizations (gray).

of the pore-size distribution of the original sample, can be used forany multi-scale structures that are very common in shale formations(Lemmens and Richards, 2013).

For the sake of comparison, pore-throat size of the original sample iscompared with the produced realizations (Münch and Holzer, 2008;Sahimi, 2011). First, an equivalent binary image B(p) is producedbased on the generated realization, in which all the pores' locationsare identified, P⊂B. The grain location is detonated by P, where P∪P¼B. The distance map d(p) =min|p′−p | is calculated, which representsthe closest distance of the pore space to its grain boundaries. The maprepresents the possible locations of the pores where the centers of allspheres of radius of rs are located. The comparison is shown in Fig. 9. Al-though the realizations do not reproduce the exact pore-throat size dis-tribution of the original sample, they represent a cloud of variationaround the true distribution. This variation indicates the fact that sever-al samples with different pore-throat size distributions can producesimilar absolute permeabilities. A bimodal distribution of the pore-throat size is evident in Fig. 9, indicating a complex distribution ofpore space in shale reservoirs (Shultz, 2015). Furthermore, the fractionof the organic matters and their associated porosity are also calculatedfor both the original and the realizations. The average of original mate-rials in the original sample and 3D realizationswere 19.50% and 21.27%,respectively. The porosity associated with each of the original sampleand generated realizations was also computed to be 2.55% and 2.70%,accordingly. Furthermore, the TOC were calculated to be 55% and 56%in the original and the realizations, correspondingly, both indicatingvery good agreement with the data.

4. Conclusions

Producing equi-probable 3D shale realizations validated with mea-sured permeability is of high importance. In our study, we introduced

Page 11: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

241P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

a new iterative method that takes a few 2D images of a complex shalesample and reconstructs its 3D model stochastically. Such reconstruc-tion, which uses physical rules derived from the given DIs, produces re-alizations that contain the pore networks in the actual samples. Due toits direct usage of the SEM images in the stochastic realizations, the re-constructed 3Dmodels forecast more accurately the TOC of shale reser-voirs than any of the previous methods.

As is well known, stochasticmethodsmust be conditioned further toproduce certain important properties. For example, experimental datafor such properties as flow modeling and electrical conductivity can beintegrated to reduce the uncertainty and producemore realistic realiza-tions. Our approach for reconstructing 3D samples uses a high-orderstatistical approach represented by a cross-correlation function that ef-ficiently integrates the experimental data. The method is based on theMetropolis-Hasting approach and the genetic algorithm, which arecast on an iterative learning concept in such a way that the previous it-erations produce the next realizations.

Finally, the approach proposed in this paper can be used to betterunderstand the complexity of shale reservoirs that produce gas. Themethod generates realization of shales thatmatch themeasuredperme-ability without using full 3D imaging. Clearly, providing a number ofsuch realizations improves our knowledge of the pore-network connec-tivity of shales. The method may also be used with various types of ex-perimental data, such as the permeability, electrical conductivity.Clearly, more experimental data reduce the uncertainty in the realiza-tions. The proposed method in this paper, regardless of the pore struc-tures, can be used for any type of natural porous media with multi-scale structures.

Acknowledgements

Work at USCwas supported by Research Partnership to Secure Ener-gy for America (RPSEA) (53-4504-1115), and at the University of Texasat Austin by the NanoGeosciences Laboratory andMudrock Systems Re-search Laboratory (MSRL) consortium. Dr. R. Loucks provided the SEMimages. Publicationwas authorized by the Director, Bureau of EconomicGeology. All data and digital content in this manuscript can be accessedby sending an e-mail to [email protected].

References

Adler, P.M., Jacquin, C.G., Quiblier, J.A., 1990. Flow in simulated porous media. Int.J. Multiphase Flow 16 (4), 691–712. http://dx.doi.org/10.1016/0301-9322(90)90025-E.

Andrä, H., Combaret, N., Dvorkin, J., Glatt, E., Han, J., Kabel, M., Keehm, Y., Krzikalla, F., Lee,M., Madonna, C., Marsh, M., 2013. Digital rock physics benchmarks—Part I: Imagingand segmentation. Comput. Geosci. 50, 25–32. http://dx.doi.org/10.1016/j.cageo.2012.09.005.

Bhandari, A.R., Flemings, P.B., Polito, P.J., Cronin, M.B., Bryant, S.L., 2015. Anisotropy andstress dependence of permeability in the Barnett shale. Transp. Porous Media 108(2), 393–411. http://dx.doi.org/10.1007/s11242-015-0482-0.

Biswal, B., Manwart, C., Hilfer, R., Bakke, S., Øren, P.E., 1999. Quantitative analysis of exper-imental and synthetic microstructures for sedimentary rock. Physica A 273 (3),452–475. http://dx.doi.org/10.1016/S0378-4371(99)00248-4.

Biswal, B., Øren, P.E., Held, R.J., Bakke, S., Hilfer, R., 2007. Stochastic multiscale model forcarbonate rocks. Phys. Rev. E 75 (6). http://dx.doi.org/10.1103/PhysRevE.75.061303(061303).

Bryant, S., Blunt, M., 1992. Prediction of relative permeability in simple porous media.Phys. Rev. A 46 (4), 2004. http://dx.doi.org/10.1103/PhysRevA.46.2004.

Clarkson, C.R., Solano, N., Bustin, R.M., Bustin, A.M.M., Chalmers, G.R.L., He, L.,Melnichenko, Y.B., Radliński, A.P., Blach, T.P., 2013. Pore structure characterizationof North American shale gas reservoirs using USANS/SANS, gas adsorption, and mer-cury intrusion. Fuel 103, 606–616. http://dx.doi.org/10.1016/j.fuel.2012.06.119.

Coelho, D., Thovert, J.F., Adler, P.M., 1997. Geometrical and transport properties of randompackings of spheres and aspherical particles. Phys. Rev. E 55 (2), 1959. http://dx.doi.org/10.1103/PhysRevE.55.1959.

Endres, D.M., Schindelin, J.E., 2003. A new metric for probability distributions. IEEE Trans.Inf. Theory 49 (7), 1858–1860. http://dx.doi.org/10.1109/TIT.2003.813506.

Gao, M., He, X., Teng, Q., Zuo, C., Chen, D., 2015. Reconstruction of three-dimensional po-rous media from a single two-dimensional image using three-step sampling. Phys.Rev. E 91 (1), 013308. http://dx.doi.org/10.1103/PhysRevE.91.013308.

Gerke, K.M., Karsanina, M.V., 2015. Improving stochastic reconstructions by weightingcorrelation functions in an objective function. EPL (Europhysics Letters) 111 (5),56002. http://dx.doi.org/10.1209/0295-5075/111/56002.

Hastings, W.K., 1970. Monte Carlo sampling methods using Markov chains and their ap-plications. Biometrika 57 (1), 97–109. http://dx.doi.org/10.1093/biomet/57.1.97.

Hu, H., Zhang, T., Wiggins-Camacho, J.D., Ellis, G.S., Lewan, M.D., Zhang, X., 2015. Experi-mental investigation of changes in methane adsorption of bitumen-free Woodfordshale with thermal maturation induced by hydrous pyrolysis. Mar. Pet. Geol. 59,114–128. http://dx.doi.org/10.1016/j.marpetgeo.2014.07.029.

Javadpour, F., 2009. Nanopores and apparent permeability of gas flow in mudrocks(shales and siltstone). J. Can. Pet. Technol. 48 (8), 16–21. http://dx.doi.org/10.2118/09-08-16-DA.

Javadpour, F., Ettehadtavakkol, A., 2015. Gas transport processes in shale. In: Rezaee, R.(Ed.), Fundamentals of Gas Shale Reservoirs. John Wiley & Sons Inc., In,pp. 245–266 http://dx.doi.org/10.1002/9781119039228.ch11.

Javadpour, F., Moravvej, M., Amrein, M., 2012. Atomic force microscopy: a new tool forgas-shale characterization. J. Can. Pet. Technol. 51 (4), 236–243. http://dx.doi.org/10.2118/161015-PA.

Jiao, Y., Stillinger, F.H., Torquato, S., 2009. A superior descriptor of random textures and itspredictive capacity. Proc. Natl. Acad. Sci. 106 (42), 17634–17639. http://dx.doi.org/10.1073/pnas.0905919106.

Karimpouli, S., Tahmasebi, P., 2016. Conditional reconstruction: an alternative strategy indigital rock physics. Geophysics 81 (4), D465–D477. http://dx.doi.org/10.1190/geo2015-0260.1.

Kelly, S., El-Sobky, H., Torres-Verdín, C., Balhoff, M.T., 2015. Assessing the utility of FIB-SEM images for shale digital rock physics. Adv. Water Resour. 18 (8). http://dx.doi.org/10.1016/j.advwatres.2015.06.010.

Leal, L.G., 2007. Advanced Transport Phenomena: Fluid Mechanics and Convective Trans-port Processes. Cambridge University Press.

Lemmens, H., Richards, D., 2013. 4 Multiscale imaging of shale samples in the scanningelectron microscope. American Association of Petroleum Geologists Memoir 102:Electron Microscopy of Shale Hydrocarbon Reservoirs, pp. 27–35 http://dx.doi.org/10.1306/13391702M1023582.

Lemmens, H.J., Butcher, A.R., Botha, P.W.S.K., 2011. FIB/SEM and SEM/EDX: a new dawnfor the SEM in the Core Lab? Petrophysics 52 (06), 452–456.

Loucks, R.G., Reed, R.M., Ruppel, S.C., Hammes, U., 2012. Spectrum of pore types and net-works in mudrocks and a descriptive classification for matrix-related mudrock pores.AAPG Bull. 96 (6), 1071–1098. http://dx.doi.org/10.1306/08171111061.

Mehmani, A., Prodanović, M., Javadpour, F., 2013. Multiscale, multiphysics networkmodeling of shale matrix gas flows. Transp. Porous Media 99 (2), 377–390. http://dx.doi.org/10.1007/s11242-013-0191-5.

Metropolis, N., Rosenbluth, A.W., Rosenbluth, M.N., Teller, A.H., Teller, E., 1953. Equationof state calculations by fast computing machines. J. Chem. Phys. 21 (6), 1087–1092.http://dx.doi.org/10.1063/1.1699114.

Münch, B., Holzer, L., 2008. Contradicting geometrical concepts in pore size analysisattained with electron microscopy and mercury intrusion. J. Am. Ceram. Soc. 91(12), 4059–4067. http://dx.doi.org/10.1111/j.1551-2916.2008.02736.x.

Okabe, H., Blunt, M.J., 2004. Prediction of permeability for porous media reconstructedusing multiple-point statistics. Phys. Rev. E 70 (6), 066135. http://dx.doi.org/10.1103/PhysRevE.70.066135.

Øren, P.E., Bakke, S., 2002. Process based reconstruction of sandstones and prediction oftransport properties. Transp. Porous Media 46 (2-3), 311–343. http://dx.doi.org/10.1023/A:1015031122338.

Pyrcz, M.J., Deutsch, C.V., 2014. Geostatistical Reservoir Modeling. Oxford Universitypress.

Robert, C., Casella, G., 2013. Monte Carlo Statistical Methods. second edition. Springer,New York.

Roberts, A.P., Teubner, M., 1995. Transport properties of heterogeneous materials derivedfrom Gaussian random fields: bounds and simulation. Phys. Rev. E 51 (5), 4141.http://dx.doi.org/10.1103/PhysRevE.51.4141.

Sahimi, M., 2011. Flow and Transport in Porous Media and Fractured Rock: From ClassicalMethods to Modern Approaches. second edition. Wiley-VCH, Weinheim, Germany.

Shultz, J.M., 2015. Pore Network Variation Identified through NMR Analysis: Eagle FordGroup, South Texas. University of Texas at Austin, USA (DOI: 2152/32583., MScThesis).

Strebelle, S., 2002. Conditional simulation of complex geological structures using multi-ple-point statistics. Math. Geol. 34 (1), 1–21. http://dx.doi.org/10.1023/A:1014009426274.

Tahmasebi, P., Hezarkhani, A., Sahimi, M., 2012. Multiple-point geostatistical modelingbased on the cross-correlation functions. Comput. Geosci. 16 (3), 779–797. http://dx.doi.org/10.1007/s10596-012-9287-1.

Tahmasebi, P., Sahimi, M., 2015a. Reconstruction of nonstationary disordered materialsand media: Watershed transform and cross-correlation function. Phys. Rev. E 91(3), 032401. http://dx.doi.org/10.1103/PhysRevE.91.032401.

Tahmasebi, P., Sahimi, M., 2015b. Geostatistical simulation and reconstruction of porousmedia by a cross-correlation function and integration of hard and soft data. Transp.Porous Med. 107 (3), 871–905. http://dx.doi.org/10.1007/s11242-015-0471-3.

Tahmasebi, P., Javadpour, F., Sahimi, M., 2015a. Three-dimensional stochastic characteri-zation of shale SEM images. Transp. Porous Media 110 (3), 521–531. http://dx.doi.org/10.1007/s11242-015-0570-1.

Tahmasebi, P., Javadpour, F., Sahimi, M., 2015b. Multiscale and multiresolution modelingof shales and their flow and morphological properties. Sci. Rep. 5, 16373. http://dx.doi.org/10.1016/10.1038/srep16373.

Tahmasebi, P., Sahimi, M., 2016a. Enhancing multiple‐point geostatistical modeling: 1.Graph theory and pattern adjustment. Water Resour. Res. 52, 2074–2098. http://dx.doi.org/10.1002/2015WR017806.

Page 12: Stochastic shale permeability matching: Three-dimensional characterization … · 2020. 6. 17. · Stochastic shale permeability matching: Three-dimensional characterization and modeling

242 P. Tahmasebi et al. / International Journal of Coal Geology 165 (2016) 231–242

Tahmasebi, P., Sahimi, M., 2016b. Enhancing multiple‐point geostatistical modeling: 2. It-erative simulation and multiple distance function. Water Resour. Res. 52,2099–20122. http://dx.doi.org/10.1002/2015WR017807.

Tahmasebi, P., Javadpour, F., Sahimi, M., Piri, M., 2016. Multiscale study for stochasticcharacterization of shale samples. Adv. Water Resour. 89, 91–103. http://dx.doi.org/10.1016/j.advwatres.2016.01.008.

Tarantola, A., 2005. Inverse Problem Theory and Methods for Model Parameter Estima-tion. Society for Industrial and Applied Mathematics (SIAM), Paris, France.

Veselý, M., Bultreys, T., Peksa, M., Lang, J., Cnudde, V., Van Hoorebeke, L.L., Kočiřík, M.,Hejtmánek, V., Šolcová, O., Soukup, K., Gerke, K., 2015. Prediction and evaluation of

time-dependent effective self-diffusivity of water and other effective transport prop-erties associated with reconstructed porous solids. Transp. Porous Media 110 (1),81–111. http://dx.doi.org/10.1007/s11242-015-0557-y.

Yeong, C.L.Y., Torquato, S., 1998a. Reconstructing randommedia. Phys. Rev. E 57 (1), 495.http://dx.doi.org/10.1103/PhysRevE.57.495.

Yeong, C.L.Y., Torquato, S., 1998b. Reconstructing random media. II. Three-dimensionalmedia from two-dimensional cuts. Phys. Rev. E 58 (1), 224. http://dx.doi.org/10.1103/PhysRevE.58.224.