Journal Pre-proof Computing heterogeneous core sample velocity using Digital Rock Physics: A multiscale approach Sadegh Karimpouli, Asra Faraji, Martin Balcewicz, Erik H. Saenger PII:
S0098-3004(19)30728-9
DOI:
https://doi.org/10.1016/j.cageo.2019.104378
Reference:
CAGEO 104378
To appear in:
Computers and Geosciences
Received Date: 1 August 2019 Revised Date:
15 November 2019
Accepted Date: 18 November 2019
Please cite this article as: Karimpouli, S., Faraji, A., Balcewicz, M., Saenger, E.H., Computing heterogeneous core sample velocity using Digital Rock Physics: A multiscale approach, Computers and Geosciences (2020), doi: https://doi.org/10.1016/j.cageo.2019.104378. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. © 2019 Published by Elsevier Ltd.
1
Computing Heterogeneous Core Sample Velocity Using Using Digital Rock
2
Physics: A Multiscale Approach†
3
Sadegh Karimpoulia*, Asra Farajia, Martin Balcewicz b,c and Erik H. Saenger b,c
4
a. Mining Engineering Group, Faculty of Engineering, University of Zanjan, Zanjan, Iran.
5
* Corresponding author’s Email:
[email protected]
6
b. International Geothermal Centre, Bochum University of Applied Sciences, Germany. Email:
7
[email protected]
8
c. Ruhr-University Bochum, Germany. E-mail:
[email protected]
9
Abstract
10
Digital Rock Physics (DRP) is an effective approach to compute physical properties of rock using
11
high-resolution 3D images. Although micro-scale structures are well studied in DRP, micro
12
computed tomography (µCT) scanners are still not widely available and imaging could be both
13
expensive and time consuming. Moreover, there is always a trade-off between image resolution
14
and sample size. The later issue is crucial especially in heterogeneous samples such as
15
carbonates. In this study, we propose a multiscale procedure and computed wave velocity of
16
three heterogeneous travertine (calcite) samples with the exact core size of 55 mm diameter.
17
This procedure is conducted using relatively cheap and widely available imaging tools, namely
18
medical CT scanner and conventional microscope. In a low-resolution step, 3D CT-images with
19
200 µm resolution are obtained and used to compute porosity cubes by a three-phase †
Sadegh Karimpouli conceived the problem, developed the idea, performed the coding and contributed to analyzing the results and writing the paper. Asra Faraji prepared core samples, conducted laboratory analysis in Iran and performed image analysis and computation. Martin Balcewicz conducted laboratory analysis in Germany and contributed to writing the paper. Erik Saenger performed computation and contributed to writing the paper.
1
20
segmentation (mineral, pore and sub-resolution). In a high-resolution step, elastic moduli are
21
computed using 2D microscopic images with 1 µm resolution and converted into 3D properties
22
using a 2D-to-3D DRP method. The obtained micro-trends of effective elastic moduli are
23
analytically approximated using the modified Differential Effective Medium (DEM) model. Low-
24
resolution porosity cubes and high-resolution micro-trends are combined to obtain 3D cubes of
25
elastic moduli and density. Finally, a dynamic Finite Difference Method (FDM) is used to
26
propagate P- and S-waves through these samples for velocity computations. Comparison of lab
27
and computed results showed that the errors of P- and S-wave velocities vary from 3.4 % to 7 %
28
and from 6.7 % to 11.1 %, respectively.
29
Keyword: Keyword: Digital Rock Physics (DRP); Heterogeneous carbonate; Multiscale; Core sample velocity
30 31
1. Introduction
32
Knowledge about the wave velocity through the rock is critical tool supporting reservoir
33
evaluation. Velocity and, generally, elastic properties of rock are the main parameters relating
34
seismic data with rock parameters measured in borehole and laboratory. Rock properties are
35
quantified in pore scale using, for example, Digital Rock Physics (DRP) method, which uses high-
36
resolution images of rock and simulates physical processes such as wave propagation to
37
compute velocity. Upscaling of such quantities into well, laboratory (i.e., core) and/or seismic
38
scale is still controversial and, therefore, a direct evaluation of such quantities in a larger scale
39
than pore scale is highly desired. DRP been developed in the last two decades mostly due to
40
emerging of micro X-Ray computed tomography (µCT) scanners and focused ion beam scanning
41
electron microscopy (FIB-SEM) (Andrä et al., 2013a, 2013b). High-resolution representation of
2
42
rock microstructures enabled the DRP to describe the role of complex pore geometry on
43
effective parameters of rock. The standard workflow of DRP is based on three main steps (Andrä
44
et al., 2013a, 2013b; Blunt et al., 2013): (1) imaging, (2) segmentation and (3) numerical
45
computation. The rock sample is imaged using, for example, a µCT scanner which provides a 3D
46
gray-scale image. The reconstructed 3D image is based on a large number of 2D projections in
47
different angles (Mathews et al., 2017). The brightness of image depends on the attenuation
48
coefficient of the material and is strongly linked to density. This means, bright and dark voxels
49
could be representative of mineral and pore phases, respectively. The process of labeling each
50
voxel into different phases is called image segmentation. It is a very critical step in DRP due to its
51
direct effects on pore space structure and, thus, the numerical results. Therefore, several
52
segmentation methods have been introduced including thresholding (Otsu, 1975; Sezgin, 2004),
53
region growing (Beucher and Meyer, 1992; Kass et al., 1988), hybrid (Andrä et al., 2013a;
54
Ramandi et al., 2016) and Deep Learning (DL)-based methods (Karimpouli and Tahmasebi, 2019).
55
Segmentation is even the most critical step when sample fractures and discontinuities are
56
supposed to be delineated (Deb et al., 2008; Karimpouli et al., 2019; Kemeny and Post, 2003;
57
Muller et al., 1992). After image segmentation, proper physical properties could be assigned to
58
each phase and physical processes are numerically computed to simulate elastic wave
59
propagation, electrical current and fluid flow for evaluation of effective elasticity, wave velocity,
60
resistivity and permeability of the sample.
61
Although DRP is an effective method for capturing micro-scale characteristics of rock, there are
62
some limitations. Main limitations are (1) the high-technology µCT scanners are not widely
63
available, (2) imaging is expensive and could be time consuming if higher resolutions are desired,
3
64
(3) sample size is small (mostly in mm order), and (4) a high capacity in terms of computer
65
memory and CPU usage is needed for numerical simulations. The first two issues opened a new
66
avenue through methods, which use 2D this section images to compute 3D properties of rock
67
(Berrezueta et al., 2019; Karimpouli et al., 2018b; Saxena et al., 2017; Saxena and Mavko, 2016).
68
The two last issues are more critical especially in heterogeneous samples such as carbonates,
69
where the sample contains different pores from nano to macroscale. This means, micro porosity
70
could not be resolved by a low-resolution image, but high-resolution image may not be a
71
representative element volume (REV) due to the small field-of-view of the sample (Hebert et al.,
72
2015; Knackstedt et al., 2009). In fact, there is a trade-off between sample size and image
73
resolution. This is one reason why digital models appear to perform stiffer with higher velocities
74
than the measured values of similar samples in laboratory (Arns et al., 2005; Saenger et al.,
75
2011). This difference could also be due to grain-to-grain modulus, crystal deflects and in grain
76
micro-fractures. To overcome this issue and characterize the heterogeneity in different scales,
77
multi-scale imaging is a possible approach (Devarapalli et al., 2017). Although some researchers
78
found similar trends between rock physical parameters in different scales of carbonates (Dvorkin
79
et al., 2011; Saenger et al., 2016), a multi-scale study could be a more reliable approach since it
80
integrates all information from different scales (Hebert et al., 2015). Knackstedt et al. (2009)
81
used thin sections and detected that unresolved pixels in their CT images are due to cemented
82
micrite crystals. They used Self-Consistent (SC) theory to compute elastic moduli of microporous
83
regions and computed more reasonable velocities comparing to experiments. Jouini et al. (2015)
84
proposed a multi-scale approach to extrapolate results from high-resolution (0.3 µm) images
85
into a core scale image with 38 mm diameter. Computed effective moduli by Jouini et al. (2015)
4
86
are overestimated regarding to experimental measurements. They assume nanoscale pores may
87
cause this overestimation, which could be further resolved using nano-tomography. Faisal et al.
88
(2017) tried to directly simulate effective elastic properties of Silurian Dolomite on the core scale
89
(with 38 mm diameter). Their computations in different resolutions showed that effective elastic
90
properties depend not only on the porosity value, but also on microstructure details (we simply
91
call it pore shape). They extracted small non-overlapping cubes from core scale image with
92
different sizes and computed effective elastic moduli of the original sample using elastic moduli
93
of all small cubes. Faisal et al. (2017) showed, that estimations of computed elastic moduli
94
strongly depend on the REVs size, if a constant resolution is set. For instance, a sample with an
95
appropriate REV size results in better estimations than several samples with sizes smaller than
96
the determined REV size. They recently extended their study on similar samples using a multi-
97
scale approach (Farhana Faisal et al., 2019), where images with three resolutions were used: 13
98
µm (low-resolution), 1 µm (medium-resolution), and 0.12 µm (high-resolution). To resolve the
99
sub-resolution parts of the low-resolution image, they prepared limited numbers of images with
100
medium-resolution. They also used some high-resolution FIB-SEM images to resolve nanoscale
101
pores of the medium images. Effective moduli of small images were computed, and their
102
average values were used as the moduli of the sub-resolution parts of the medium-resolution
103
image. Similarly, the average properties of medium-resolution images were used for sub-
104
resolution parts of the low-resolution image. Results showed reasonable improvements, where
105
overestimated errors were reduced from 8.9 to 4.5 % for P-wave velocity and from 11.9 to 7.8 %
106
for S-wave velocity. Although, these works represented valuable results, they neglected the
107
effect of pore shapes. In fact, in two latter researches, Faisal et al. (2017, 2019) determined that
5
108
microstructural details like pore shape and structure play an important role, but they suppressed
109
this effect by choosing an average value. We suggest instead of using the average properties for
110
the unresolved part, microscale trends of elastic moduli should be studied and used in
111
macroscale images. As a similar work, Karimpouli et al. (2018a) used such an idea to resolve the
112
sub-resolution carbonate cement in a cemented sandstone. Due to limitation of the available
113
data, they used a random procedure for porosity value of sub-resolution pixels. Subsequently,
114
they found two negative effects: (1) the spatial distribution of sub-resolution porosity is
115
neglected and (2) the overall properties of sub-resolution pixels approaches to average
116
properties of the medium (similar to Faisal et al. (2017, 2019)). This is why elastic moduli were
117
also overestimated in their study. Unlike to that work, we here demonstrate, this idea could lead
118
to better results if a reliable procedure is used for both porosity evaluation of the sub-resolution
119
pixels and assigning proper elastic properties based on elastic property-porosity trends in
120
microscale.
121
High resolution imaging requires small subsamples, if µCT or FIB-SEM methods are applied.
122
Preparing subsamples results into numerous recordings which are expensive, time consuming,
123
and not widely available. Therefore, we aim to directly compute velocity of a core sample (e.g.,
124
with about 5 cm diameter and 10 cm length) and tackle the multiscale problem using
125
inexpensive and widely spread imaging tools. Our methodology is subdivided into three steps.
126
First, we use a medical CT scanner, which is both cheap and accessible for the macro-imaging.
127
The limitation for this technology is pore size. The size of the largest pores in the sample must be
128
at least two-times more than the CT resolution. In a second step, we use 2D microscopic image
129
and try to convert 2D results into 3D characteristics using alternative DRP methods (Saxena and
6
130
Mavko, 2016). Finally, we merge high resolution 3D characteristics with lower resolution medical
131
CT images to calculate effective properties of the rock.
132
The rest of the paper is organized as follows. In section 2, core samples and their lab
133
measurements are introduced. Then, the multiscale procedure is explained in section 3. Section
134
4 presents computations and results and, finally, section 5 summarizes the limitations and
135
advantages of the methodology.
136 137
2. Core samples
138
In this study, we use a travertine (limestone) sample from west of Zanjan, Iran. Travertine layers
139
were formed by calcareous hot springs originated from tertiary volcanic intrusions. It is a highly
140
heterogeneous sample with pores sizes varying from cm to even nm. This lithology has been
141
chosen due to macro- and microscopic heterogeneity, and to demonstrate our multiscale
142
procedure using medical CT scanners. Three core samples, namely S1, S2, and S3, were drilled
143
with 5.5 cm diameter and 10 cm length, cut plane-parallel and polished (Fig. 1). After
144
preparation all samples were oven-dried at 60 °C for 48 hours. As it is observed in this figure,
145
each sample contains large (about 1-2 cm) and micro pores (less than 1 µm). Spatial distribution
146
of pore spaces is also variable from well distributed (S2) to poorly distributed (S3) sample.
147
Chemical analysis (Table 1) showed that more than 99 % composition of travertine sample is
148
calcite. Physical characteristics of these samples were measured in the laboratory and
149
summarized in Table 2. Porosimetry was conducted using helium injection method and wave
150
velocity was measured by ultrasonic pulse velocity test at ambient pressure-temperature
7
151
conditions in the laboratory of Research Institute of Petroleum of Industry (RIPI), Iran. Wave
152
velocity was also measured in laboratory of International Geothermal Center (GZB), Germany.
153
Ultrasonic measurements were conducted by similar set-ups at RIPI and GZB. For example, in
154
Germany, a table-top solution was used. The GZB set-up consists mainly of three parts. A
155
waveform generator, two identical ultrasonic sensors (transmitter/receiver) and an oscilloscope.
156
The external waveform generator (Panametrix EPOCH 650) creates a rectangular signal (400 V)
157
with a controlled central frequency (1 MHz), which causes the piezoelectric transducer to emit
158
mechanical pulses into the sample. At the other sample's end, the receiver converts the arriving
159
mechanical pulses (waveforms) back into electric signals. The digital oscilloscope (Picoscope
160
5444B) records the electric waveforms. P- and S-wave arrival signals were determined by manual
161
picking. Effective travel times for each sample was ensured by correcting travel times in the set-
162
up. P- and S-wave velocities were detected for dry state.
163 164
3. Multiscale procedure
165
As it is illustrated in Fig. 2, the heterogeneous travertine sample contains pore spaces with a size
166
from cm to µm or even nm. This means, in a low-resolution image, macro-pores are detected,
167
and sub-resolution micro-pores are missed. In a similar manner, a high-resolution image resolves
168
micro-pores, but macro-pores are missed due to small field-of-view.
169
Our suggested multiscale procedure is visually explained in Fig. 3. In this procedure, we use two
170
high- and low-resolution images, but one can extend this procedure with more images in
171
different resolutions. Large size pores of travertine made it reasonable for us to use medical CT
172
scanners for capturing the low-resolution images with all relevant features. The largest cube
8
173
from the center of the core is extracted as the digital image of each sample. The size of this cube
174
is relatively small (e.g. less than 5003 voxels) even for a sample in cm scale. This is a significant
175
matter from a computational standpoint, where numerical computations could be also
176
conducted very cheap. In this step, a three-phase segmentation (Jouini et al., 2015) is used to
177
divide the CT-image into a porosity cube. In this cube, porosity of each voxel is either zero, if it is
178
a mineral phase, 100 %, if it is a pore phase, or a value between 0 to 100 % depending on gray-
179
scale level of the sub-resolution voxel. Then, material-specific properties (i.e. elastic moduli and
180
density) are assigned to each phase. For the mineral and pore phases with 0 and 100 %,
181
corresponding mineral and fluid properties (e.g. air in dry condition) were applied. However, the
182
question is: “Which values should be assigned to sub-resolution voxels?” A widespread method is
183
to use average values computed from high-resolution images with micro-scale pores (Faisal et
184
al., 2017; Farhana Faisal et al., 2019; Jouini et al., 2015), which suppresses the effects of micro-
185
scale pore shape and value. Previous studies in carbonates showed, that the effects of pore
186
shape could be as important as pore values for evaluation of velocity (Anselmetti and Eberli,
187
1993; Eberli et al., 2003; Karimpouli et al., 2013). This means, effective properties of sub-
188
resolution voxels should be obtained using real trends, which are behavior of one physical
189
property versus another property of rock (e.g. bulk modulus versus porosity). Since we obtain
190
these trends in micro-scale, we call them “micro-trends”, which are affected by exact micro-pore
191
values and structures.
192
To find micro-trends, high-resolution images are needed. These images could be captured either
193
by µCT scanners or FIB-SEM, but for the sake of a simple and cost-effective procedure, we use
194
2D microscopic images. High-resolution images are captured from polished thin sections (Fig. 3).
9
195
The main point is, that the size of high-resolution images should be equal to one low-resolution
196
voxel length. Generally, it is not necessary that the size of high-resolution image is exactly equal
197
to the voxel-size of the low resolution image, but the best state is theoretically when that are
198
equal.
199
In this case, these images are segmented into two phases, namely mineral and pore. After
200
segmenting, 2D effective elastic properties could be computed by either finite element (FEM) or
201
finite difference methods (FDM) codes. The solver related uncertainty is in order of 1/10th of
202
input mineral moduli (Saxena et al., 2019) and could be ignored compared with other major
203
controls such as image resolution. To convert 2D into 3D effective properties, several different
204
DRP methods could be used (Karimpouli et al. 2018b; Karimpouli and Tahmasebi 2016; Saxena
205
and Mavko 2016). Analytical effective models such as the Weighted Average Boundary (WAB)
206
model or the Differential Effective Medium (DEM) model describe micro-trends. Since these
207
models are fitted onto effective moduli computed by exact micro-scale structures of the sample,
208
they enact the effects of pore shape and value simultaneously.
209
Micro-trends can be used to assign to sub-resolution voxels of the low-resolution porosity cube
210
appropriate effective elastic properties (Fig. 3). This leads to cubes of elastic moduli and density
211
from which the effective velocities are computed. In practice, FEM code (Garboczi and Day,
212
1995) can be applied if there are less than 100 phases in the image (Garboczi, 1998), however, in
213
this procedure each voxel is considered as one phase due to its individual properties. This means,
214
for a 3D cube with voxel numbers in order of millions, FEM code is impractical. The used dynamic
215
FDM code (Saenger et al., 2000; Saenger and Shapiro, 2002) does not struggle with allocation
216
limitations.
10
217
3.1. Effective elastic models
218
Estimation of effective elastic moduli of a mixture of grains and pores deals with the volume
219
fraction, elastic moduli, and geometrical details of each phase. By specifying volume fraction and
220
elastic moduli, we can predict upper (stiffer) and lower (softer) material bounds (Mavko et al.,
221
2009). Hashin and Shtrikman (1963) represented the best bounds for an isotropic linear elastic
222
composite known as Hashin-Shtrikman bounds. Dvorkin and Nur (1996) introduced stiff and soft
223
sand models for cemented and uncemented sands, but were also applied for general trends of
224
carbonates (Andrä et al., 2013b). Bounding models reveal the effective elastic properties of two
225
endmembers, but for a sample located between two bounds, Weighted Average Bounding
226
(WAB) model could be used. To consider the third parameter, geometrical details, the methods
227
based on Self-consistent (SC) or Differential Effective Medium (DEM) theory could be used.
228
These models are based on pore shape and generate reliable effective elastic moduli according
229
to their fields of applications (Wang et al., 2015; Xu and Payne, 2009).
230
3.1.1. Differential Effective Medium (DEM)
231
In DEM theory, two-phase composites are modeled by considering one host (or mineral) phase
232
and incrementally adding guest (or pore) phase. When porosity is zero, mineral is assumed as
233
the matrix or host phase. Porosity with known geometrical details or pore shape is added
234
incrementally until a target porosity value is obtained. In each step, effective elastic moduli are
235
computed and used as the matrix elastic moduli for the next step (Mavko et al., 2009). The DEM
236
equations are: 1 − ∗ = − ∗ ∗ ,
(1)
11
1 − 237 238 239
∗ = − ∗ ∗ ,
(2)
where ∗ and ∗ are effective elastic moduli with initial conditions ∗ = and ∗ = ,
where and are the bulk and shear moduli of the host material (mineral), and are the
bulk and shear moduli of the incrementally added inclusions, is the concentration of guest
240
phase (porosity) and ∗ and ∗ are geometrical factors of guest phase (for more details see
241
Mavko et al. (2009)). Geometrical factors determine the effects of guest inclusions with different
242
materials and shapes. The computational details are explained in Appendix A. For a two-phase
243
medium, as in our case, geometrical factors are obtained according to only the shape of
244
inclusions. Each inclusion is assumed as an ellipsoid and the corresponding ratio of small to large
245
radii, known as aspect ratio, is used as the shape factor to compute geometrical factors. Effective
246
elastic moduli are highly affected by the aspect ratio of pore spaces (Berryman, 1980). In a
247
randomly oriented model, it is stiff model if the aspect ratio is one (spherical pores), and it
248
becomes softer if the aspect ratio is reduced (ellipsoidal pores). Where the aspect ratio is very
249
small, pore space is like a fracture and a weak model is obtained. In DEM model, porosity can
250
increase up to 100 % if solid phase is connected. However, in rock samples, there is a critical
251
porosity ( ), where solid matrix falls apart. Mukerji et al. (1995) proposed modified DEM and
252
assumed percolation behavior of the rock at critical porosity as the guest phase. Therefore, in
253
modified DEM, denotes the concentration of the critical phase in the matrix.
254 255
4. Results
256
4.1. Low-resolution step
12
257
All core samples are imaged by a medical CT scanner, which are shown in Fig. 4. The largest cube
258
is extracted from each image called “CT-image” (Fig. 4). For each sample, the size of the CT
259
image is 200×200×475 voxels, where edge length of each voxel is 200 µm. According to image
260
resolution all pores with a size of less than 200 µm are not resolved and the containing voxels
261
represent a gray-scale value depending on the amount of occupied grains and pores.
262
To have an insight about porosity and pore structure of the samples, large-scale pores were
263
divided using a two-phase segmentation by selecting a rough threshold value. Fig. 4 illustrates
264
the pore structure and distribution through 3D space of the samples. First hand computations
265
(i.e., two phase segmentation) showed that porosity of samples is roughly about 15 % for
266
samples S1 and S2, and 10 % for S3. These values are not consistent with laboratory
267
measurements (Table 2), referring 9.8 % for S1 and S2, and 7.4 % for S3, respectively. Fig. 4
268
reveals that there are many large and isolated pores with no connection with connected pore
269
system, which are not measured during helium-injection porosimetry. According to the nature of
270
travertines, when hot spring waters rich in calcium carbonate and dissolved carbon dioxide are
271
exposed to Earth’s surface, carbon dioxide is outgassed and calcite (and occasionally aragonite)
272
may precipitate from supersaturated solution of carbonate calcium (Drysdale, 1999). Gasification
273
could lead to isolated pores with any size from cm to nm. Distribution of such pores could also
274
be very heterogeneous due to variable precipitation process of travertines.
275
To explore if the CT-image is larger than the Representative Volume Element (RVE), in each
276
sample, several cubes with different sizes were extracted from the center of CT-image and
277
corresponding porosity were computed. Porosity variation with image size is plotted in Fig. 5.
13
278
According to this plot, almost in all sample, a cube with a size of 1503 voxels could be considered
279
as RVE.
280
Since the resolution of CT-images is low, two-phase segmentation could not produce reliable
281
results. For tacking sub-resolution porosity into account, three-phase segmentation is used (Fig.
282
6) (Jouini et al., 2015). This method uses two threshold values ( and in Fig. 6) and
283
divides all voxels into three phases: mineral, pore and sub-resolution phases. The corresponding
284
porosity of mineral and pore phases are 0 and 100 %, but the porosity of sub-resolution phase is
285
computed by a simple linear interpolation between two threshold values (Fig. 6). However, a
286
more reliable way for computing porosity of sub-resolution phase is to use a high-resolution
287
image. Supposing, if all micro-structures are resolved in such an image, an empirical relation is
288
obtained by correlating micro-porosity and gray-scale values (Ramandi et al., 2016). In this study,
289
we used the linear interpolation. One can use K-means clustering (MacQueen, 1967) or bi-level
290
segmentation (Ridler and Calvard, 1978) methods for finding threshold values, but Jouini et al.
291
(2015) set them manually for more accurate results. We also used a porosity inverse method
292
(Yao et al., 2009) to set the threshold values. To this end, porosity of each sample is obtained
293
using bulk and grain density measurements, and threshold values are set accordingly. Porosity
294
values were computed as 15.04 %, 14.29 % and 10.95 % for S1, S2, and S3, respectively. The
295
porosity cubes could be used to assign corresponding values of elastic moduli and density to
296
each phase. We used properties of calcite and air for mineral and pore phases (Table 3), but for
297
the sub-resolution phase micro-trends are needed, which are obtained using high-resolution
298
scale.
299
4.2. High-resolution step
14
300
In this step, four polished thin sections were prepared and imaged using conventional optical
301
microscopes. These sections were not select from the cores, which means our high-resolution
302
images are not registered to low-resolution ones. The size of high-resolution images is set to 200
303
µm with a resolution of 1 µm/pixel, which is equal to low-resolution voxel length (200 µm/voxel).
304
Since 2D imaging and corresponding numerical processing are fast, 200 individual 2D high-
305
resolution microscopic images were captured (Fig. 7). The microscope settings and
306
corresponding imaging software were arranged to show high contrast. Reflective light was
307
applied to indicate pores as black spots and minerals as grey areas. This makes it easy to
308
segment them using a simple thresholding. We used Otsu thresholding (Otsu, 1975) to segment
309
the grain and pore phases. Results of segmentation are shown in Fig. 7.
310
To compute effective elastic properties from 2D images of rock samples, the Finite Element code
311
by Garboczi and Day (1995) was used, which is based on solving the Hooke’s law equations of
312
linear elasticity. Results are shown in Fig. 8. These results are in 2D and, before any further
313
process, they should be converted into 3D properties. By using 2D information to predict 3D rock
314
properties we assume a similar heterogeneity level in all three spatial directions. Among all
315
alternative DRP methods, we used the method by Saxena and Mavko (2016), which converts
316
elastic properties from 2D into 3D using the following equations:
=
"
=
"
317 318
,
(3)
,
(4)
" "
!
#
where and are bulk and shear moduli of three- and two-dimensional models (3% and 2%) and mineral ('). ( and (" are constant powers, which should be adjusted in any case. Saxena
15
319 320
and Mavko (2016) proposed ( = (" = 0.6 for carbonate samples. Also, we found ( =
0.6, (" = 0.5 for real carbonate reservoir rocks (Karimpouli et al., 2018a). In this case, we used
321
an average value of core measured velocity and determined ( = (" = 0.7 are reasonable
322
values. Results of 3D effective elastic moduli of high-resolution images are shown in Fig. 7. These
323
results reveal the microscopic properties of the sample, which are affected by real pore shapes
324
and structures.
325
To find micro-trends, effective elastic model based on DEM equations (Eq. 1, 2) were used. In the
326
DEM model, bulk and shear aspect ratios (needed for ∗ and ∗ in Eq. 1, 2) should be
327
optimized for obtaining the best fit. The criterion for best fitting is Mean Square Error (/0): /0 =
328
∑= 4>?234567 839:7; <
(5)
,
where / @ and /A@B are effective elastic moduli computed from images (we mean the 3D
329
values) and calculated by models and C is the number of images. In modified DEM, two aspect
330
ratios are obtained for bulk and elastic moduli. The optimum aspect ratios were obtained as 0.74
331
and 0.21 with MSE values of 1072 and 1982 (GPa)2 for modified DEM models of bulk and shear
332
moduli. These micro-trends are shown in Fig. 7. In this figure, porosity axes are plotted up to 60
333
% porosity, however, the micro-trends extend up to 100 % porosity where they are 0.
334
4.3. Combined high- and low-resolution step
335
The micro-trends (DEM models) obtained in the last step could be used to assign proper
336
properties to each voxel of low-resolution image based on the porosity cube. For each voxel, if
337
porosity is 0 or 100 %, properties of calcite or air (Table 3) are assigned, respectively. For other
338
porosity values, properties are obtained from the modified DEM trends and assigned into the
16
339
corresponding voxels, which leads to low-resolution cubes of elastic moduli. Since porosity of
340
each voxel is known, the following relation is used to compute cube of density: DEFBG = DH + 1 − D ,
(6)
341
where DH and D are density of fluid (here air) and mineral.
342
Finally, the dynamic pulse propagation method (e.g. Saenger and Shapiro (2002)), which is based
343
on a staggered-grid FDM, is used to compute P- and S-wave velocities using cubes of elastic and
344
density properties. The heterogeneous sample constructed by these cubes is embedded in a
345
homogeneous elastic region with elastic properties of calcite. A plane-source wave is generated
346
and propagated numerically through the sample using a periodic boundary condition. The time-
347
delay of the peak amplitude of the mean plane wave could be used to estimate effective P- and
348
S-wave velocities. Results for velocity computations are plotted in Fig. 9 and summarized in Table
349
4.
350 351
5. Discussions
352
According to our results (Table 4), the mean percentage errors (Δ %) of P-wave velocity, defined
353
as the difference between lab and computed velocity, are 3.4 % and 7 % for RIPI and GZB
354
laboratories. The errors of S-wave velocity are 11.1 % and 6.7 % for RIPI and GZB laboratories.
355
The smallest errors of the alternative multiscale studies (Faisal et al., 2017; Farhana Faisal et al.,
356
2019; Jouini et al., 2015) are 4.5 and 7.8 % for P- and S-wave velocity. Comparing to these
357
studies, which are expensive in both imaging tools and computation points of view, our results
358
are reasonable and solid.
17
359
There are several parameters, which impact the results. For example, medical CT is a proper tool
360
for porous carbonates such as travertine, but it could not be used for other reservoir rocks with
361
pore sizes smaller than about 200 µm, for instance, most of sandstones and compacted
362
carbonates. However, the proposed multiscale procedure (Fig. 3) could be used by other
363
techniques on different scales. It means, instead of using images in mm and µm scales such as
364
what we had in our case, one can use some images in µm and nm scales with similar multiscale
365
procedure. This procedure could also be extended in more than two steps. The main points for
366
such a procedure are: (1) the image size of each step should be equal to the voxel size of the last
367
step, and (2) segmentation of all steps are three-phase except for the highest-resolution step,
368
which is two-phase segmentation.
369
We limited ourselves to use 2D microscopic images in the high-resolution step, because we
370
wanted to find the most economical but still accurate way. Using 3D images in this step could
371
lead to more reliable results depending on rock type and heterogeneity. In our study, we use a
372
huge number of images (i.e., 200) for the case of carbonates. The resulting rock physical trend
373
should be representative even for this relative heterogeneous rock type. In addition, these
374
images were not registered. The authors suppose that being registered does not have a drastic
375
effect on final results, however, we suggest high-resolution images to be registered if smaller
376
errors are desired.
377
Finally, the key point of this multiscale study is using micro-trends instead of an average value for
378
sub-resolution pixels. The effective properties model, used to fit the micro-trends, could be
379
variant in different samples. Since they are computed from the real micro-scale world of the
380
sample, they are affected by the of pore shapes of the sample. It means, the effects of pore
18
381
shape, which is considered as one of the most important rock physics parameters in carbonates,
382
are implied in numerical computation, which leads to more realistic results.
383 384
6. Conclusion
385
In this study, three core samples with 5.5 cm diameter and 10 cm length from a highly
386
heterogeneous travertine were selected for velocity computation using DRP. We introduced a
387
multiscale procedure using relatively cheap and widely available imaging tools. In the low-
388
resolution step, we used a medical CT tool for 3D imaging, which generates images with the size
389
reasonable for a fast-numerical computation. Since micro-pores are not resolved in this scale,
390
three-phase segmentation was used to divide CT-images into mineral, pore and sub-resolution
391
voxels. To assign physical properties to each voxel, elastic properties of calcite and air properties
392
were used for voxels with mineral and pore phases. However, for sub-resolution voxels, we used
393
two threshold values and predicted porosity using a linear interpolation. Corresponding
394
properties to each porosity were obtained from micro-trends of the sample. In the high-
395
resolution step, conventional optical microscopes were applied to capture 2D images in order to
396
show the ability of alternative DRP methods. Micro-trends were modeled by WAB and modified
397
DEM models, but DEM was better fitted with smaller MSE for both bulk and shear moduli. Since
398
these trends capture micro-scale pore shape and structure, they highly improve the estimation
399
of effective elastic moduli. Using these micro-trends, cubes of effective elastic moduli and
400
density were computed based on low-resolution CT-image and porosity cube. Finally, a FDM
401
code was applied to compute effective wave velocity. Our results showed that the errors of P-
19
402
and S-wave velocity computations are 3.4 % and 11.1 % for RIPI measurements and they are 7 %
403
and 6.7 % for GZB measurements.
404
The proposed procedure was used for a sample with large pores. Though, it could be
405
implemented for samples with pore sizes smaller than the resolution of medical CT scanners. In
406
this case, images with smaller scales are needed. This procedure could also be extended to more
407
than two scale levels. For a routine use of such a multi-scale approach some basic calibration
408
steps are necessary as we discussed in this paper.
409 410
References
411
Andrä, H., Combaret, N., Dvorkin, J., Glatt, E., Han, J., Kabel, M., Keehm, Y., Krzikalla, F., Lee, M.,
412
Madonna, C., Marsh, M., Mukerji, T., Saenger, E.H., Sain, R., Saxena, N., Ricker, S.,
413
Wiegmann, A., Zhan, X., 2013a. Digital rock physics benchmarks-Part I: Imaging and
414
segmentation. Comput. Geosci. 50, 25–32. https://doi.org/10.1016/j.cageo.2012.09.005
415
Andrä, H., Combaret, N., Dvorkin, J., Glatt, E., Han, J., Kabel, M., Keehm, Y., Krzikalla, F., Lee, M.,
416
Madonna, C., Marsh, M., Mukerji, T., Saenger, E.H., Sain, R., Saxena, N., Ricker, S.,
417
Wiegmann, A., Zhan, X., 2013b. Digital rock physics benchmarks-part II: Computing effective
418
properties. Comput. Geosci. 50, 33–43. https://doi.org/10.1016/j.cageo.2012.09.008
419
Anselmetti, F.S., Eberli, G.P., 1993. Controls on sonic velocity in carbonates. Pure Appl. Geophys.
420
PAGEOPH 141, 287–323. https://doi.org/10.1007/BF00998333
421
Arns, C.H., Bauget, F., Limaye, A., Sakellariou, A., Senden, T., Sheppard, A., Sok, R.M., Pinczewski,
422
V., Bakke, S., Berge, L.I., Oren, P., Knackstedt, M., 2005. Pore-Scale Characterization of
423
Carbonates Using X-Ray Microtomography. SPE J. 10, 26–29.
20
424 425
https://doi.org/10.2118/90368-PA Berrezueta, E., Domínguez-Cuesta, M.J., Rodríguez-Rey, Á., 2019. Semi-automated procedure of
426
digitalization and study of rock thin section porosity applying optical image analysis tools.
427
Comput. Geosci. 124, 14–26. https://doi.org/10.1016/j.cageo.2018.12.009
428 429 430 431 432 433 434
Berryman, J.G., 1980. Long-wavelength propagation in composite elastic media II. Ellipsoidal inclusions. J. Acoust. Soc. Am. 68, 1820–1831. Beucher, S., Meyer, F., 1992. The morphological approach to segmentation: the watershed transformation. Opt. Eng. YORK-MARCEL DEKKER Inc. 34, 433–433. Blunt, M.J., Bijeljic, B., Dong, H., Gharbi, O., Iglauer, S., Mostaghimi, P., Paluszny, A., Pentland, C., 2013. Pore-scale imaging and modelling. Adv. Water Resour. 51, 197–216. Deb, D., Hariharan, S., Rao, U.M., Ryu, C.H., 2008. Automatic detection and analysis of
435
discontinuity geometry of rock mass from digital images. Comput. Geosci. 34, 115–126.
436
https://doi.org/10.1016/j.cageo.2007.03.007
437
Devarapalli, R.S., Islam, A., Faisal, T.F., Sassi, M., Jouiad, M., 2017. Micro-CT and FIB–SEM imaging
438
and pore structure characterization of dolomite rock at multiple scales. Arab. J. Geosci. 10,
439
361. https://doi.org/10.1007/s12517-017-3120-z
440
Drysdale, R.N., 1999. The sedimentological significance of hydropsychid caddis-fly larvae (order;
441
Trichoptera) in a travertine-depositing stream; Louie Creek, Northwest Queensland,
442
Australia. J. Sediment. Res. 69, 145–150. https://doi.org/10.2110/jsr.69.145
443 444 445
Dvorkin, J., Derzhi, N., Diaz, E., Fang, Q., 2011. Relevance of computational rock physics. Geophysics 76, E141–E153. https://doi.org/10.1190/geo2010-0352.1 Dvorkin, J., Nur, A., 1996. Elasticity of high-porosity sandstones: Theory for two North Sea data
21
446 447
sets. GEOPHYSICS 61, 1363–1370. https://doi.org/10.1190/1.1444059 Eberli, G.P., Baechle, G.T., Anselmetti, F.S., Incze, M.L., 2003. Factors controlling elastic
448
properties in carbonate sediments and rocks. Lead. Edge 22, 654–660.
449
https://doi.org/10.1190/1.1599691
450
Faisal, T.F., Awedalkarim, A., Chevalier, S., Jouini, M.S., Sassi, M., 2017. Direct scale comparison
451
of numerical linear elastic moduli with acoustic experiments for carbonate rock X-ray CT
452
scanned at multi-resolutions. J. Pet. Sci. Eng. 152, 653–663.
453
https://doi.org/10.1016/j.petrol.2017.01.025
454
Farhana Faisal, T., Islam, A., Jouini, M.S., Devarapalli, R.S., Jouiad, M., Sassi, M., Faisal, T.F., Islam,
455
A., Jouini, M.S., Devarapalli, R.S., Jouiad, M., Sassi, M., 2019. Numerical prediction of
456
carbonate elastic properties based on multi-scale imaging. Geomech. Energy Environ.
457
100125.
458
Garboczi, E.J., 1998. Finite Element and Finite Difference Programs for Computing the Linear
459
Electric and Elastic Properties of Digital Images of Random Materials | NIST. NIST
460
Interagency/Internal Rep. - 6269.
461
Garboczi, E.J., Day, A.R., 1995. An algorithm for computing the effective linear elastic properties
462
of heterogeneous materials: three-dimensional results for composites with equal phase
463
Poisson ratios. J. Mech. Phys. Solids 43, 1349–1362. https://doi.org/10.1016/0022-
464
5096(95)00050-S
465 466 467
Hashin, Z., Shtrikman, S., 1963. A variational approach to the theory of the elastic behaviour of multiphase materials. J. Mech. Phys. Solids 11, 127–140. Hebert, V., Garing, C., Luquot, L., Pezard, P.A., Gouze, P., 2015. Multi-scale X-ray tomography
22
468
analysis of carbonate porosity. Geol. Soc. London, Spec. Publ. 406, 61–79.
469
https://doi.org/10.1144/SP406.12
470
Jouini, M.S., Vega, S., Al-Ratrout, A., Al-Ratrout, A., 2015. Numerical estimation of carbonate rock
471
properties using multiscale images. Geophys. Prospect. 63, 405–421.
472
https://doi.org/10.1111/1365-2478.12156
473
Karimpouli, S., Hassani, H., Nabi-Bidhendi, M., Khoshdel, H., Malehmir, A., 2013. Application of
474
probabilistic facies prediction and estimation of rock physics parameters in a carbonate
475
reservoir from Iran. J. Geophys. Eng. 10, 1–14. https://doi.org/10.1088/1742-
476
2132/10/1/015008
477
Karimpouli, S., Khoshlesan, S., Saenger, E.H., Koochi, H.H., 2018a. Application of alternative
478
digital rock physics methods in a real case study: a challenge between clean and cemented
479
samples. Geophys. Prospect. Accepted, Published Online. https://doi.org/10.1111/1365-
480
2478.12611
481 482 483 484 485
Karimpouli, S., Tahmasebi, P., 2019. Segmentation of Digital Rock Images Using Deep Convolutional Autoencoder Networks. Comput. Geosci. 126, 142–150. Karimpouli, S., Tahmasebi, P., 2016. Conditional reconstruction: An alternative strategy in digital rock physics. GEOPHYSICS 81, D465–D477. https://doi.org/10.1190/geo2015-0260.1 Karimpouli, S., Tahmasebi, P., Saenger, E.H., 2019. Coal Cleat/Fracture Segmentation Using
486
Convolutional Neural Networks. Nat. Resour. Res. https://doi.org/10.1007/s11053-019-
487
09536-y
488 489
Karimpouli, S., Tahmasebi, P., Saenger, E.H., 2018b. Estimating 3D elastic moduli of rock from 2D thin-section images using differential effective medium theory. GEOPHYSICS 83, MR211–
23
490 491 492 493
MR219. https://doi.org/10.1190/geo2017-0504.1 Kass, M., Witkin, A., Terzopoulos, D., 1988. Snakes: Active contour models. Int. J. Comput. Vis. 1, 321–331. https://doi.org/10.1007/BF00133570 Kemeny, J., Post, R., 2003. Estimating three-dimensional rock discontinuity orientation from
494
digital images of fracture traces. Comput. Geosci. 29, 65–77.
495
https://doi.org/10.1016/S0098-3004(02)00106-1
496
Knackstedt, M.A., Latham, S., Madadi, M., Sheppard, A., Varslot, T., Arns, C., 2009. Digital rock
497
physics: 3D imaging of core material and correlations to acoustic and flow properties. Lead.
498
Edge 28, 28–33. https://doi.org/10.1190/1.3064143
499
MacQueen, J.B., 1967. Some Methods for classification and Analysis of Multivariate
500
Observations, in: Proceedings of 5th Berkeley Symposium on Mathematical Statistics and
501
Probability. University of California Press, pp. 281–297.
502
Mathews, J.P., Campbell, Q.P., Xu, H., Halleck, P., 2017. A review of the application of X-ray
503
computed tomography to the study of coal. Fuel 209, 10–24.
504
https://doi.org/10.1016/J.FUEL.2017.07.079
505
Mavko, G., Mukerji, T., Dvorkin, J., 2009. The rock physics handbook: Tools for seismic analysis of
506
porous media. Cambridge university press.
507
https://doi.org/http://dx.doi.org/10.1017/CBO9780511626753
508
Mukerji, T., Berryman, J., Mavko, G., Berge, P., 1995. Differential effective medium modeling of
509
rock elastic moduli with critical porosity constraints. Geophys. Res. Lett. 22, 555–558.
510
https://doi.org/10.1029/95GL00164
511
Muller, J.J., Schrauber, H., Müller, J.J., Schrauber, H., 1992. The inertia-equivalent ellipsoid: a link
24
512
between atomic structure and low-resolution models of small globular proteins determined
513
by small-angle X-ray scattering. J. Appl. Crystallogr. 25, 181–191.
514
https://doi.org/10.1107/S0021889891011421
515
Otsu, N., 1975. A threshold selection method from gray-level histograms. Automatica 11, 23–27.
516
Ramandi, H.L., Armstrong, R.T., Mostaghimi, P., 2016. Micro-CT image calibration to improve
517
fracture aperture measurement. Case Stud. Nondestruct. Test. Eval.
518
Ridler, T.W., Calvard, S., 1978. Picture Thresholding Using an Iterative Selection Method. IEEE
519
Trans. Syst. Man. Cybern. 8, 630–632. https://doi.org/10.1109/TSMC.1978.4310039
520
Saenger, E.H., Enzmann, F., Keehm, Y., Steeb, H., 2011. Digital rock physics: Effect of fluid
521
viscosity on effective elastic properties. J. Appl. Geophys. 74, 236–241.
522
https://doi.org/10.1016/j.jappgeo.2011.06.001
523
Saenger, E.H., Gold, N., Shapiro, S.A., 2000. Modeling the propagation of elastic waves using a
524
modified finite-difference grid. Wave Motion 31, 77–92. https://doi.org/10.1016/S0165-
525
2125(99)00023-2
526
Saenger, E.H., Shapiro, S.A., 2002. Effective velocities in fractured media: a numerical study using
527
the rotated staggered finite-difference grid. Geophys. Prospect. 50, 183–194.
528
https://doi.org/10.1046/j.1365-2478.2002.00309.x
529
Saenger, E.H., Vialle, S., Lebedev, M., Uribe, D., Osorno, M., Duda, M., Steeb, H., 2016. Digital
530
carbonate rock physics. Solid Earth 7, 1185–1197. https://doi.org/10.5194/se-7-1185-2016
531
Saxena, N., Hofmann, R., Hows, A., Saenger, E.H., Duranti, L., Stefani, J., Wiegmann, A., Kerimov,
532
A., Kabel, M., 2019. Rock compressibility from microcomputed tomography images:
533
Controls on digital rock simulations. GEOPHYSICS 84, WA127–WA139.
25
534 535
https://doi.org/10.1190/geo2018-0499.1 Saxena, N., Mavko, G., 2016. Estimating elastic moduli of rocks from thin sections: Digital rock
536
study of 3D properties from 2D images. Comput. Geosci. 88, 9–21.
537
https://doi.org/10.1016/j.cageo.2015.12.008
538
Saxena, N., Mavko, G., Hofmann, R., Srisutthiyakorn, N., 2017. Estimating permeability from thin
539
sections without reconstruction: Digital rock study of 3D properties from 2D images.
540
Comput. Geosci. 102, 79–99. https://doi.org/10.1016/j.cageo.2017.02.014
541 542 543
Sezgin, M., 2004. Survey over image thresholding techniques and quantitative performance evaluation. J. Electron. Imaging 13, 146–168. Wang, Z., Wang, R., Li, T., Qiu, H., Wang, F., 2015. Pore-scale modeling of pore structure effects
544
on P-wave scattering attenuation in dry rocks. PLoS One 10, e0126941.
545
https://doi.org/10.1371/journal.pone.0126941
546
Xu, S., Payne, M.A., 2009. Modeling elastic properties in carbonate rocks. Lead. Edge 28, 66–74.
547
Yao, Y., Liu, D., Che, Y., Tang, D., Tang, S., Huang, W., 2009. Non-destructive characterization of
548
coal samples from China using microfocus X-ray computed tomography. Int. J. Coal Geol.
549
80, 113–123. https://doi.org/10.1016/J.COAL.2009.08.001
550 551 552 553
Appendix A‒ A‒ Geometric factors ∗ and ∗ For dry ellipsoidal pore inclusions, the geometric coefficients are given by the following (Berryman, 1980; Mavko et al., 2009): ∗ = L MNN , K
(A-1)
26
∗ = O MNN − L MNN , K
554 555
K
(A-2)
where the tensor MNGB relates the uniform strain field to the strain field within the ellipsoidal
556
inclusion . Berryman (1980) gave the scalar formulation required for calculating ∗ and ∗
557
as follows: MNN =
LP? P
,
(A-3)
MNN = MNN + K L
Q
P
+
K
PR
+
,
PR PS TPU PV TPW PX P PR
(A-4)
YK = 1 + Z [ \ + ] − ^ \ + ] − `, L
(A-5)
YK = 1 + Z [ \ + ] − ^ \ + ] − `, L
(A-6)
L
L
Q
Q
L Q
O Q
L
O
Q
Q
_ _
YQ = 1 + Z [ 1 + \ + ] − ^3\ + 5] ` + a3 − 4^ + ZZ + 3a 3 − L
K
Q
K
Q
Q
4^ \ + ] − ^ \ − ] + 2] Q ,
(A-7)
YL = 1 + Z [1 − \ + ] + ^\ + ]`,
(A-8)
Y_ = 1 + _ Z\ + 3] − ^ \ − ] ,
(A-9)
L Q
K
YO = Z [−\ + ^ \ + ] − L` + a]3 − 4^ , _
(A-10)
Yc = 1 + Z1 + \ − ^\ + ] + a1 − ]3 − 4^,
(A-11)
Yd = 2 + Z3\ + 9] − ^ 3\ + 5] + a]3 − 4^ ,
(A-12)
Yf = Z [1 − 2^ + \^ − 1 + ]5^ − 3` + a1 − ]3 − 4^,
(A-13)
K _
K Q
K Q
Yg = Z^ − 1\ − ^] + a]3 − 4^ , Z =
"6
"h
(A-14)
− 1,
(A-15)
27
a = L − " , K 6
^ = ] = \ =
h
K8Qih
QK8ih
l
"6
(A-16)
,
(A-17)
h
8K
mmQ − 1 − cosh8K m , m > 1 ?
k [ cos 8K m − m1 − mQ ` , m < 1 j K8
K8
?
,
(A-18)
3] − 2,
(A-19)
558 559 560 561 562
Table 1. 1. Chemical analysis of the travertine sample. Loss of Ignition (LOI) is the percentage of
563
volatile substances escaping by strongly heating. Since the major concentration goes back to CaO
564
and other oxides are less than 1 %, it can be concluded that the sample is completely composed
565
of calcite. Composition
SiO2
Al2O3
BaO
Value (%)
0.19
0.05
<0.05 55.42 <0.05 <0.05
Composition
MnO
Na2O
P2O5
SO3
TiO2
<0.05 <0.05 <0.05
0.25
<0.05 43.55
Value (%) 566 567 568 569
28
CaO
Fe2O3
K2O
LOI
MgO 0.54
570 571 572 573 574 575 576 577 578 579 580
Table 2. 2. Laboratory measurements of physical properties of core samples measured in Research
581
Institute of Petroleum of Industry (RIPI) and International Geothermal Center (GZB). Length
Diameter
Sample -3
(10 m)
-3
(10 m)
Bulk
Grain
density
density
Weight
VP (m/s)
Porosity
-3
(10 Kg)
VS (m/s)
(%) 3
3
(Kg/m )
(Kg/m )
RIPI
GZB
RIPI
GZB
S1
100.70
54.33
540.99
2320
2720
9.8
5204
5820
2629
2700
S2
100.97
54.48
546.71
2320
2700
9.8
5385
5835
2751
2770
S3
100.37
54.47
566.56
2420
2720
7.4
5530
6012
2811
2990
582 583 584 585 586 587 29
588 589 590 591 592 593 594 595 596 597 Table 3. 3. Properties of mineral and pore phases (Mavko et al., 2009).
598 Phase
Bulk modulus (GPa)
Shear modulus (GPa)
Density (kg/m3)
Calcite
76.8
32
2710
Air
0
0
0
599 600 601 602 603 604 605 606 30
607 608 609 610 611 612 613 614 615 616 Table 4. Results for velocity computations.
617 Sample
Computed VP
Δ % VP-RIPI
Δ % VP-GZB
Computed VS
Δ % VS-RIPI
Δ % VS-GZB
S1
5503
5.7
6.9
2974
13.1
8.7
S2
5436
0.9
8.2
3036
10.3
8.5
S3
5727
3.5
5.8
3087
9.8
2.7
3.4
7.0
-
11.1
6.7
Average 618 619 620 621 622 623
31
624 625 626 627 628 629 630 631 632 633 634
Figure captions:
635
Fig. 1. Core samples used in this study.
636
Fig. 2. 2. An illustration of heterogeneous travertine sample with macro and micro pores.
637
Fig. 3. 3. Multiscale procedure proposed in this study.
638
Fig. 4. 4. Core samples (S1, S2, and S3) and corresponding CT-images extracting as the largest cube.
639
Pore structures obtained via bimodal thresholding with a rough threshold value.
640
Fig. 5. RVE analysis for three samples. A plateau is observed above 1503 voxels for almost all
641
samples.
642
Fig. 6. Computation of porosity cube by three-phase segmentation of the CT-image.
643
Fig. 7. Four examples of 2D high-resolution microscopic and corresponding segmented images.
644
Fig. 8. Bulk and shear moduli of 2D images (circles) and corresponding 3D models (square).
645
Micro-trends based on modified DEM are also shown by dash lines.
32
646
Fig. 9. (a) Lab and computed velocity versus porosity plot and (b) lab versus computed velocity
647
plot.
33
Highlights • • • •
Velocity of a core sample (with 55 mm diameter) is computed via a multiscale procedure. Cost-effective and widely available imaging tools are used. Microscopic 2D images are used to enact 3D effects of pore type in microscale. Velocity computations are highly accurate compared to laboratory tests.
Declaration of interests The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.