A semi-empirical relation between static and dynamic ...

25 downloads 0 Views 2MB Size Report
May 21, 2017 - Martin, R. J., Price, R. H., Boyd, P. J. and Noel, J. S., 1993, Unconfined compression ... Morales, R. H., Marcinew, R. P., 1993, Fracturing of ...
Accepted Manuscript A semi-empirical relation between static and dynamic elastic modulus Mohammad Reza Asef, Mohsen Farrokhrouz PII:

S0920-4105(17)30544-2

DOI:

10.1016/j.petrol.2017.06.055

Reference:

PETROL 4067

To appear in:

Journal of Petroleum Science and Engineering

Received Date: 18 September 2016 Revised Date:

21 May 2017

Accepted Date: 19 June 2017

Please cite this article as: Asef, M.R., Farrokhrouz, M., A semi-empirical relation between static and dynamic elastic modulus, Journal of Petroleum Science and Engineering (2017), doi: 10.1016/ j.petrol.2017.06.055. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.

RI PT

ACCEPTED MANUSCRIPT

SC

A semi-empirical relation between static and dynamic elastic modulus

M AN U

Mohammad Reza Asef: Kharzami University, Faculty of Earth Sciences, Applied Geology Department, Mofateh Ave. No 49, Tehran 15719-14911, Iran

Mohsen Farrokhrouz: South Zagros Oil Company, ICOFC, No. 30, West Roodsar St., Hafez St., Tehran, Iran

AC C

EP

TE

D

Corresponding author: Kharzami University , Faculty of Earth Sciences, Applied Geology Department Mofateh Ave. No 49. Tehran 15719-14911. Iran. E-mail: [email protected]

ACCEPTED MANUSCRIPT

ABSTRACT

2

Researchers have long investigated the difference between static and dynamic

3

Young’s modulus in terms of values and application. In field experiences, dynamic

4

Young’s modulus can often be calculated using logging data, while core data for

5

static experiments in the lab are scarce. On the other hand, geomechanical analyses

6

are traditionally based on static values. Therefore, it is essential to estimate static

7

Young’s modulus based on dynamic data for petroleum reservoir case studies.

8

A number of empirical equations have been suggested to estimate static Young’s

9

modulus based on dynamic Young’s modulus. However, the validity of these

10

relations are limited and local. In this paper, 10 suggested relations were evaluated

11

against an experimental dataset from different geographic locations. The validity of

12

each relation was evaluated, and the results confirmed the dependency of these

13

equations to the original datasets.

14

Based on the concept of effective Young's modulus, we suggested a relationship to

15

predict the value of static Young's modulus from dynamic Young's modulus and

16

porosity data. The approach was performed for different rock types, and the

17

predictions were found to be very accurate. This research explored different views

18

on this aspect and facilitated better decision making to determine the static

19

Young’s modulus of rocks in the case where only dynamic data were available.

AC C

EP

TE D

M AN U

SC

RI PT

1

ACCEPTED MANUSCRIPT

1 2

A semi-empirical relation between static and dynamic elastic modulus

3

5

RI PT

4

1. Introduction

Elastic Young's modulus is a general rock mechanical parameter that is usually

7

measured in all geomechanical projects. This parameter is important because it is a

8

direct representative of material stiffness: the lower the elastic modulus value, the

9

lower the material stiffness.

M AN U

SC

6

Elastic constants (Young’s modulus and Poisson’s ratio) are considered to be

11

fundamental mechanical properties of rock material that are required to be

12

measured for analysis and design of engineering projects. Additionally, they are

13

extensively used in analytical equations and modeling techniques to predict the

14

stress–strain behavior of rock subjected to various loading conditions.

15

Two methods are commonly used to measure the elastic modulus of rock material:

16

static and dynamic. Static Young’s modulus (Es) is commonly obtained as the

17

tangent of the stress-strain curve at 50% of the maximum strength of rock core

18

specimens. Dynamic Young’s modulus (Ed) can be calculated knowing the rock

19

density as well as compressional and shear wave velocities. Measurement of static

AC C

EP

TE D

10

ACCEPTED MANUSCRIPT

Young’s modulus requires a destructive testing approach in the lab, where

21

relatively large static loads are applied to rock core specimens. Dynamic Young's

22

modulus, however, is obtained from non-destructive tests, where the velocities of

23

propagation of elastic waves (seismic or acoustic) are measured. It is interesting to

24

note that acoustic wave data, namely, compressional and shear wave velocities (Vp

25

and Vs), can be obtained both in the lab and at the project site from logs. When an

26

acoustic wave propagates through a porous medium, the deformation of the grains

27

is elastic. Since core samples are not always available, laboratory measurements

28

are more expensive. Therefore, dynamic measurements are often more accessible.

29

In addition, using Vp and Vs log data makes it possible to obtain dynamic values

30

for the whole volume of encountered rock, while static tests are performed on a

31

sample of limited length. However, there is a meaningful difference between static

32

and dynamic Young’s moduli.

33

Several studies have shown that static and dynamic properties are different. The

34

ratio between dynamic and static moduli was found to range between 1 to 20

35

(Wang, 2000). Low ratios occur for stiff rocks, and higher ratios are obtained for

36

softer sediments. For a solid material like steel, the ratio is 1 (Weast, 1986).

37

Therefore, correlating these two values has been an important issue from very early

38

times. Many descriptions have been proposed to explain this difference, ranging

39

from strain amplitude effects to viscoelastic behavior. Zisman (1933) provided a

AC C

EP

TE D

M AN U

SC

RI PT

20

ACCEPTED MANUSCRIPT

qualitative explanation of the consistently higher magnitude of Ed compared to Es.

41

He suggested that the wave pulse or packet of energy passing through rock upon

42

entering a cavity (crack, pore) suffers a loss of energy (due to reflection and

43

refraction) at the fluid/rock and rock/fluid interfaces. A small amount of wave

44

energy is transmitted across the boundary and a large amount is scattered around

45

the cavities. Therefore, this difference was explained as static measurements being

46

more influenced by the presence of fractures, cracks, cavities, planes of weakness,

47

or foliation (Ide, 1936; Sutherland, 1962; Coon, 1968). Investigation of such

48

differences is still an active area of research to understand the various contributing

49

parameters and to better interpret the mechanical properties from wave velocity

50

measurements (Asef and Najibi, 2013; Najibi and Asef, 2014). In this paper, a

51

comprehensive study of suggested prediction equations is presented and the range

52

of predicted values based on different equations is compared. Afterwards, a new

53

approach is implemented according to the poroelastic nature of rocks, especially

54

sedimentary rocks, to present theoretical and practical views on the conversion of a

55

dynamic modulus to a static modulus.

57

SC

M AN U

TE D

EP

AC C

56

RI PT

40

2. Previously Suggested Relations 2.1 General Equation Form

ACCEPTED MANUSCRIPT

58

The general form of previously suggested equations for predicting static Young’s

59

modulus is as follows:

60

E s = aE d + b

61

where ‘a’ and ‘b’ are the constants that are highly dependent on the rock type. A

62

list of suggested equations for predicting static Young’s modulus is shown in Table

63

1. As stated before, dynamic Young’s modulus is larger than the corresponding

64

static modulus. The difference is large for weak rocks and is reduced with

65

increases of confinement (King, 1970). The static modulus is affected by existing

66

micro-cracks, while the dynamic modulus appears to be less affected by micro-

67

cracks (Tutuncu and Sharma, 1992; Morales and Marcinew, 1993; Yale and

68

Jamieson, 1994). However, since dynamic Young’s modulus is calculated by wave

69

velocities, some parameters, such as porosity, can influence the dynamic modulus

70

to a large extent (Asef et al, 2010).

71

Table 1

72

Suggested equations for prediction of static Young’s modulus

AC C

EP

TE D

M AN U

SC

RI PT

(1)

Equation

Unit

R2 *

Rock Type

Reference

(2)

Es =1.26Ed − 29.5

GPa

0.82

Igneous and metamorphic rocks

King (1983)

(3)

Es = 0.74Ed − 0.82

GPa

0.84

Different Rock Types

Eissa and Kazi (1988)

No.

Es =1.05Ed − 3.16

GPa

0.99

Sedimentary rocks

Christaras et al (1994)

(5)

Es = 0.83Ed

GPa

---

Concrete

Neville (1995)

(6)

Es =1.25Ed −19

GPa

---

Structural Design

CP110 (1972)

(7)

Es = 0.2807Ed

GPa

0.60

Sedimentary rocks

Brautigam (1998)

(8)

Es = 0.8069Ed − 29.5

GPa

0.92

Composite Resin

Helvatjoglu et al (2006)

(9)

Es = 0.7707Ed − 5854

MPa

0.96

Mokovciakova (2003)

(10)

E s = e (0.0477 Ed )

GPa

0.72

Different Rock Types

Fahimifar and Soroush (2003)

(11)

Ed = 0.5087Es − 6 ×106

psi

0.60

Sandstone

Al-Tahini (2003)

RI PT

(4)

SC

ACCEPTED MANUSCRIPT

M AN U

Different Rock Types

*R2 reported in the original references

74

2.2 Prediction Verification

75

A large amount of experimental data were gathered and classified into four

76

categories: limestone, sandstone, shale, and tuff. Some of these data have been

77

previously published by others, and some data were obtained experimentally by the

78

authors. Among these data, 19 limestone values were obtained from Al-Shayea

79

(2004) and Farquhar et al (1994) and 38 values were obtained from the current

80

study. Additionally, 15 shale values were reported by Lashkaripour (1996), and 16

81

shale values were from the current study. Al-Tahini (2003) published some

82

sandstone data from deep core samples in an oil field, and 15 other specimens were

AC C

EP

TE D

73

ACCEPTED MANUSCRIPT

tested in the current study. The scatter of the gathered data versus different

84

equations is illustrated in Figure 1.

85

As is seen in Figure 1, the data points are highly scattered versus the predicted

86

equations. This analysis also revealed that large errors imply predictions that are

87

far from reality. Some of the equations give negative values for Young’s modulus,

88

which is obviously not true. Some of them may be suitable for a specific rock type,

89

while others give highly erroneous values. Altogether, application of the prediction

90

equations listed in Table 1 requires calibration of the constant values, and any

91

calibration means a new correlation. Accordingly, in this research, the

92

methodology suggested for predicting Young's modulus of any porous material

93

(ceramics, biological tissues, foams and other sintered materials) was selected for

94

natural porous material (rock and/or soil). The mechanical theory behind this

95

methodology is introduced in poromechanics and, in this paper, it is modified into

96

poroelasticity.

AC C

EP

TE D

M AN U

SC

RI PT

83

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

TE D

97

Figure 1: Scatter of data points versus different suggested equations listed in Table

99

1

3. New Approach for estimation of the static modulus

AC C

100

EP

98

101

Within the last few years, Phani and Niyogi (1987) have proposed a power-law

102

empirical relationship for predicting porous material properties over an entire

103

range of porosity (Equation 12):

104

 ϕ E = E 0 1 − ϕ c 

   

f

(12)

ACCEPTED MANUSCRIPT

where E is effective Young’s modulus of porous material with porosity φ, E0 is

106

Young’s modulus of solid material, φc is the porosity at which the effective

107

Young’s modulus becomes zero, and ‘f’ is the parameter dependent on the grain

108

morphology and pore geometry of porous material (Phani and Niyogi, 1987).

109

Later, it was noted that fitting experimental data to Equation (12) often gives φc = 1

110

(Wagh et al. 1991). Some other studies suggest that either φc = 1 (Boccaccini et al.

111

1995; Matikas et al. 1997; Maitra and Phani, 1994) or that the power-law

112

relationship is converted to a linear model when f ≡ 1 (Lam et al. 1994). In the

113

latter case, φc is considered to be the initial powder porosity. There have been some

114

studies that state that the theory behind Equation (12) can be justified by

115

percolation theory for the behavior of Young’s and shear moduli of porous

116

material (Kovacik, 1999). Porous materials (including rocks) are generally

117

prepared from powders or particles of different sizes and shapes. During the

118

consolidation process, a range of porosities can be achieved based on tectonic

119

parameters, such as temperature, external pressure or time. As particles bump into

120

each other, the compaction process starts and creation and growth of the necks

121

between particles leads to a decrease in porosity. Subsequently, pore channel

122

closure results in the elimination of the pores. According to percolation theory,

123

there is a critical volume fraction ‘nc,’ called the percolation threshold, at which a

124

continual solid phase is formed that spans the whole system. At and above this

AC C

EP

TE D

M AN U

SC

RI PT

105

ACCEPTED MANUSCRIPT

threshold, there is a relationship between the volume fractions of the solid material

126

and other geometrical, physical and mechanical properties, as shown by Equation

127

(13):

128

Π ∝ (n − n c )f

RI PT

125

for n ≥ n c

(13)

where Π is the property under investigation (it could be any parameter, such as E,

130

Vp, Vs), ‘n’ is the volume fraction of the solid material, and ‘f’ is the critical

131

exponential that is predicted for the investigated property. Another result of

132

percolation theory is that the value of the critical exponential (f) does not depend

133

on the structure or geometrical properties of the system, but only depends on the

134

dimension of the problem (Stauffer and Aharony, 1992). Moreover, the value of

135

the percolation threshold (nc) significantly depends on the structure. Therefore,

136

experimental values are reported in the range of 0.06 - 60 vol% of the three-

137

dimensional structure (Schueler et al. 1997; Balberg and Binenbaum, 1987). If

138

porosity is used instead of the volume fraction of the powder (as it is known that φ

139

= 1 - n), another form of percolation theory is obtained:

M AN U

TE D

EP

AC C

140

SC

129

Π ∝ (ϕ c − ϕ )f

for ϕ ≤ ϕc

141

Applying the boundary conditions of Π = Π0 at φ = 0 leads to Equation (15):

142

ϕ −ϕ   Π = Π 0  c   ϕc 

(14)

f

for ϕ ≤ ϕ c

(15)

ACCEPTED MANUSCRIPT

This is identical to Equation (12). Percolation theory predicts that for Young’s

144

modulus in a 3D structure, ‘f = 2.1’ (Sahimi, 1994).

145

Kovacik (1999) calculated the value of ‘f’ for different sintered materials and

146

obtained various values depending on the shape of the powders. Table 2 shows the

147

results of his studies.

148

Table 2

149

The values of ‘f’ in different sintered materials (Kovacik, 1999) Sintered material

Porosity Range

E0 (GPa)

SC

RI PT

143

Th2O powder (2µm)

0 – 0.33

263.9 ± 3.4

1.22 ± 0.10

Th2O powder (4µm)

0 – 0.39

264.2 ± 3.9

1.36 ± 0.17

Si3N4 doped with CeO2

0 – 0.38

284.1 ± 4.3

1.13 ± 0.12

Sintered Iron

0 – 0.28

213.2 ± 3.1

1.19 ± 0.28

150

M AN U

TE D

α-Al2O3 Powder

0 – 0.39

f

399.6 ± 7.7

1.25 ± 0.13

In the present study, the same approach was followed to predict static Young’s

152

modulus from dynamic Young’s modulus. We also assumed that φc = 1 and that

153

the new proposed model for Young’s modulus of rock was:

154

E = E 0 (1 − ϕ )f

155

To generate a more simple equation by applying mathematical series, the general

156

form of Equation (16) is nearly equal to the following form (see also Appendix I):

157

E ≈ E 0 [1 − fϕ ]

AC C

EP

151

(16)

(17)

ACCEPTED MANUSCRIPT

Considering (static or dynamic) Young’s modulus as a function of E0 and porosity

159

(as in Equation 17) and bearing in mind the general form of the conversion of static

160

Young’s modulus to dynamic Young’s modulus (Equation (1)), the new proposed

161

equation is Equation (18):

162

 E s = aE d + b  ⇒  E s = a' [E d0 (1 − fϕ )] + b' ( ) ϕ E E 1 f = − 0   

163

where ‘a'’ and ‘b'’ are constant values. This semi-empirical approach was

164

employed for the above-mentioned dataset that contained different rock types. The

165

values of the constants and correlation coefficients are shown in Table 3. In this

166

Table, the adjusted R2 (R2 adjusted for the number of independent variables in the

167

model) and predicted R2 (that reflects how well the model will predict future data)

168

were presented to adjust R2 for unrepresentative statistical values. It should be

169

clarified that the R2, adjusted R2, and predicted R2 values shown in Table 3 are in

170

reasonable agreement with each other. Adjusted R2 is especially important in this

171

research because if unnecessary variables are included, R2 can be misleadingly

172

high. For all data sets, adjusted R2 is close to R2, which implies that both variables

173

Ed and ϕ significantly contribute to the prediction equation.

RI PT

158

AC C

EP

TE D

M AN U

SC

(18)

174

Further statistical investigation showed that the constant ‘f’ could be considered to

175

be unity, as in Lam et al (1994), within an acceptable error range. Therefore, the

176

final form is Equation (19):

ACCEPTED MANUSCRIPT

177

E s = a' E d (1 − ϕ ) + b'

(19)

178

Table 3

180

Statistical parameters for different rock types according to Equation (19) Rock

No. of

a'

b'

Normal

Adjusted

Predicted

Data

Type

data

R2

R2

R2

a'

b'

Price et al (1994)

Tuff

12

0.83091

-0.48656

0.8712

0.8658

0.8518

0.0001

0.0001

Martin et al

Tuff

26

2.16135

-49.14988

0.7868

0.7654

0.6234

0.0001

0.0001

38

0.90255

-2.07400

0.8621

0.8583

0.8463

0.0001

0.0001

Shale

38

0.61345

7.26414

0.7870

0.7811

0.7620

0.0001

0.0001

Shale

14

0.48898

-0.59017

0.8528

0.8405

0.8068

0.0001

0.0001

52

0.93913

-3.69662

0.8946

0.8925

0.8879

0.0001

0.0001

All Tuff Yale and

M AN U

(1993)

Jamieson (1994) Lashkaripour (1996)

Al-Tahini (2003)

Sand

Morales and

Sand

Marcinew (1993)

p-value

30

0.99581

-11.10263

0.8561

0.8510

0.8362

0.0001

0.0001

9

0.89071

-5.15572

0.9451

0.9372

0.8850

0.0001

0.0001

39

0.92596

-7.48396

0.8949

0.8921

0.8831

0.0001

0.0001

EP

All Sand

TE D

All Shale

SC

Reference

RI PT

179

Lime

15

0.64897

8.90570

0.9396

0.9350

0.9268

0.0001

0.0001

Author 1

Lime

17

1.04057

-7.60778

0.9514

0.9481

0.9400

0.0001

0.0001

Author 2

Lime

8

0.93596

-4.95594

0.9604

0.9538

0.9314

0.0001

0.0001

Yale and

Carbonates

22

0.66762

8.86661

0.8917

0.8863

0.8747

0.0001

0.0001

All Lime

62

0.90804

-2.07677

0.9391

0.9380

0.9311

0.0001

0.0001

All Data

191

0.89561

-2.84374

0.9141

0.9136

0.9114

0.0001

0.0001

Yale and

AC C

Jamieson (1994)

Jamieson (1994)

181

ACCEPTED MANUSCRIPT

As can clearly be concluded from Table 3, in Equation (19), a' is often a coefficient

183

that is less than unity and b' is generally a negative value depending on the range

184

and normal distribution of the input values. All of the p-values presented in Table 3

185

state that there is only a 0.01% chance that this model could occur due to noise.

186

Considering the optimization of the constant parameters, a general model for all

187

rock types is described by Equation (20), with a 16.1% average error (estimated

188

static Young’s modulus with ± 4.5 GPa accuracy) and 4.14 standard error:

189

E s = 0.88Ed (1 − ϕ ) − 3.7

190

Figure 2 illustrates the scatter of the predicted Es values against the observed

191

values based on Equation (20) for all data on a 45° symmetry line. If the predicted

192

values match the observed values, all of the points will lie on the 45° line. The

193

fairly symmetric spread of points along the 45° line in this figure visually confirms

194

the verification of the prediction.

AC C

EP

TE D

M AN U

SC

RI PT

182

(20)

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

195 196

Figure 2: Estimated Es versus Measured Es using Equation (20)

In the next step we observed that constant b' in Equation (20), which is a negative

198

value, improved the prediction when it was related to porosity. Therefore, we

199

further improved and optimized Equation (20) into Equation (21), and R2 increased

200

to 0.92, the average error decreased to 15.2 and the standard error decreased from

201

4.14 to 3.89:

202

E s = E d (1 − ϕ ) − 3lnφ

203

AC C

EP

197

(21)

ACCEPTED MANUSCRIPT

Conclusion

205

Comprehensive statistical and theoretical studies regarding the relationship

206

between static and dynamic Young's moduli for different rock types, including

207

sandstone, shale, limestone, and tuff, revealed that porosity plays an important

208

role. We borrowed the concept of effective Young’s modulus to propose the

209

impact of porosity on the stiffness of rock material. Therefore, a new approach for

210

estimating static Young’s modulus from dynamic Young’s modulus was

211

suggested, while previous general forms of the prediction equations were also

212

considered, simultaneously. Accordingly, using dynamic Young's modulus and

213

porosity, static Young's modulus can be estimated with an acceptable error range

214

of 15% and correlation coefficient of 0.9. This is a significant improvement for

215

predicting the static properties of rock material from petrophysical data in deep

216

underground explorations.

217

References

218 219

Al- Tahini, A., 2003, The effect of cementation on the mechanical properties for Jauf reservoir at Saudi Arabia, Msc Thesis, University of Oklahoma.

220 221

Al-Shayea, N. A., 2004, Effects of testing methods and conditions on the elastic properties of limestone rock, Engineering Geology, Vol. 74, pp. 139–156, DOI: DOI: 10.1016/j.enggeo.2004.03.007.

222 223

Asef, M. R., Farrokhrouz, M., Haghi, H., 2010, Evaluation of P-wave and S-wave Correlations, Proceedings of 28th IUGG Conference on Mathematical Geophysics, Pisa, p. 75.

224 225 226

Asef, M. R., Najibi, A. R., 2013, The effect of confining pressure on elastic wave velocities and dynamic to static Young’s modulus ratio. Geophysics, 78, 3, P. D135–D142. URL: library.seg.org/doi/abs /10.1190/geo2012-0279.1.

AC C

EP

TE D

M AN U

SC

RI PT

204

ACCEPTED MANUSCRIPT

Balberg, I., Binenbaum, N., 1987, Invariant properties of the percolation thresholds in the soft-core–hardcore transition, Physical Review A., Vol. 35, pp. 5174-5177, DOI: 10.1103/ PhysRevA.35.5174.

229 230 231

Boccaccini, A. R., Ondracek, G., Mombello, E., 1995, Determination of stress concentration factors in porous materials, Journal of Materials Science Letters, Vol. 15, pp. 534-536, DOI: 10.1007/BF00275423.

232 233

Brautigam, T., Knochel, A., Lehne, M., 1998, Prognosis of uniaxial compressive strength and stiffness of rocks based on point load and ultrasonic tests, Otto-Graf-Journal, Vol. 9, pp. 61-79.

234 235 236

Christaras, B., Auger, F., Mosse, E., 1994, Determination of the moduli of elasticity of rocks: Comparison of the ultrasonic velocity and mechanical resonance frequency methods with direct static methods, Journal of Materials and Structures, Vol. 27, pp. 222-228, DOI: 10.1007/BF02473036.

237 238

Coon, R., 1968, Correlation of engineering behavior with the classification of in-situ rock, PhD Thesis, University of Illinois.

239

CP110, 1972, Code of Practice for the Structural Use of Concrete, British Standards Institution, London.

240 241

Eissa, E.A., Kazi, A., 1988, Relation Between Static and Dynamic Young’s Moduli of Rocks, Journal of Rock Mechanics and Mining Sciences and Geomechanic Abstract, Vol. 25, No. 6, pp. 479-482.

242 243 244

Fahimifar, A., Soroush, H., 2003, Evaluation of some physical and mechanical properties of rocks using ultrasonic pulse technique and presenting equations between dynamic and static elastic constants, South African Institute of Mining and Metallurgy–Technology roadmap for rock mechanics, ISRM.

245 246 247

Farquhar, R. A., Somerville, J. M., Smart, B. G. D., 1994, Porosity as a geomechanical indicator: An application of core and log data and rock mechanics, SPE 28853, presented at the European Petroleum Conference, London, DOI: 10.2118/28853-MS.

248 249 250 251

Helvatjoglu, A. M., Papadogiannisa, Y., Lakesb, R. S., Dionysopoulosa, P., Papadogiannisa, D., 2006, Dynamic and static elastic moduli of packable and flowable composite resins and their development after initial photo curing, Journal of Dental Materials, Vol. 22, pp. 450–459, DOI: 10.1016/j.dental.2005.04.038.

252 253

Ide, J. M., 1936, Comparison of statically and dynamically determined Young’s modulus of rocks, Proceedings of National Academy of science of U.S.A., Vol. 22, pp. 81-92.

254 255

King, M.S., 1970, Static aid dynamic elastic moduli of rocks under pressure, Proceedings of the 11th US Rock Mechanics Symposium, Society of Mining Engineers of AIME.

256 257 258

King, M.S., 1983, Static and dynamic elastic properties of rocks form the Canadian shield, Technical note, International Journal of Rock Mechanics and Mining Sciences and Geomechanical Abstracts, Vol. 20, No. 5, pp. 237– 241, DOI: 10.1016/0148-9062(84)90037-8.

259 260

Kovacik, J., 1999, Correlation between Young’s modulus and porosity in porous materials, Journal of Materials Science, Vol. 18, pp. 1007-1010, DOI: 10.1023/A:1006669914946.

AC C

EP

TE D

M AN U

SC

RI PT

227 228

ACCEPTED MANUSCRIPT

Lam, D. C., Lange, F. F., Evans, A. G., 1994, Mechanical Properties of Partially Dense Alumina Produced from Powder Compacts, Journal of American Ceramics Society, Vol. 77, No. 8, pp. 21132117.

264 265

Lashkaripour, G.R., 1996, The Relationship between Index and Mechanical Properties of Shales: A Database Approach, PhD thesis, University of New Castle upon Tyne, UK.

266 267

Maitra, A. K., Phani, K. K., 1994, Ultrasonic evaluation of elastic parameters of sintered powder compacts, Journal od Materials Science, Vol. 29, pp. 4415-4419, DOI: 10.1007/BF00376263.

268 269 270

Martin, R. J., Price, R. H., Boyd, P. J. and Noel, J. S., 1993, Unconfined compression experiments on Topopah Spring Member Tuff at 22°C and a strain rate of 10-9 s-1: data report, SAND92-1810 Report, Yuka Mountain Site Characterization Project.

271 272

Matikas, T. E., Karpur, P., Shamasundar, S., 1997, Measurement of the dynamic elastic moduli of porous compacts, Journal of Materials Science, Vol. 32, pp. 1099-1103, DOI: 10.1023/A:1018503127949.

273 274

Mokovciakova, A., Pandula, B., 2003, Study of the relation between the static and dynamic moduli of rocks, Journal of METABK, Vol. 42, No. 1, p. 37.

275 276 277

Morales, R. H., Marcinew, R. P., 1993, Fracturing of High-Permeability Formations: Mechanical Properties Correlations, SPE 26561, 68th Annual Technical Conference end Exhibition of the Society of Petroleum Engineers, Houston, Texas, DOI: /10.2118/26561-MS.

278 279

Najibi, A. R., Asef, M. R., 2014, Prediction of seismic-wave velocities in rock at various confining pressures based on unconfined data. Geophysics, v. 79, p. D235-D242, doi:10.1190/geo2013-0349.1.

280

Neville, A.M., 1995, Properties of Concrete, Longman Publications, 4th ed., UK.

281 282

Phani, K. K., Niyogi, S. K., 1987, Young’s modulus of porous brittle solids, Journal of Materials Science, Vol. 22, pp. 257-263, DOI: 10.1007/BF01160581.

283 284 285

Price, R. H., Boyd, P. J., Noel, J. S. and Martin, R. J., 1994, Relation between static and dynamic rock properties in welded and non-welded tuff, Association for Information and Image Management, Report No. 301/587-8202.

286

Sahimi, M., 1994, Applications of Percolation Theory, Taylor and Francis, London, p. 185.

287 288 289

Schueler, R., Petermann, J., Schulte, K., Wentzel, H. P., 1997, Agglomeration and electrical percolation behavior of carbon black dispersed in epoxy resin, Journal of Applied Polymer Science, Vol. 63, pp. 1741–1746. DOI: 10.1002/(SICI)1097-4628(19970328)63:133.0.CO;2-G.

290

Stauffer, D., Aharony, A., 1992, Introduction to Percolation Theory, 2nd ed., Taylor and Francis, London.

291 292

Sutherland, R.B., 1962, Some dynamic and static properties of rock, Proceedings of the 5th Symposium on Rock Mechanics, Minneapolis, pp. 473– 490.

293 294

Tutuncu, A. N., Sharma, M. M., 1992, Relating Static and Ultrasonic Laboratory Measurements to Acoustic Log Measurements in Tight Gas Sands, SPE 24689, 67th Annual Technical Conference and

AC C

EP

TE D

M AN U

SC

RI PT

261 262 263

ACCEPTED MANUSCRIPT

295 296

Exhibition of the 10.2118/24689-MS

Society

of

Petroleum

Engineers,

Washington

DC,

DOI:

Wagh, A. S., Poeppel, R. B., Singh, J. P., 1991, Open pore description of mechanical properties of ceramics, Journal of Materials Science, Vol. 26, Issue 14, pp. 3862-3868, DOI: 10.1007/BF01184983.

299 300 301

Wang, Z., 2000, Dynamic versus static elastic properties of Reservoir rocks, In: Wang, Z. and Nur, A. (ed), Journal of Seismic and acoustic velocities in reservoir rocks. Society of Exploration Geophysicists, Tulsa, Vol. 19, pp. 531-539.

302 303

Weast, R. C., 1986, Handbook of Chemistry and Physics, Chemical Rubber Publishing Company, 66th edition, Boca Raton, Florida, pp. 23-26.

304 305 306

Yale, D. P., Jamieson, W. H., 1994, Static and Dynamic Rock Mechanical Properties in the Hugoton and Panoma Fields, Kansas, SPE 27939, SPE Mid-Continent Gas Symposium, Texas, DOI: 10.2118/27939-MS

307 308 309

Zisman, W. A., 1933, A comparison of the statically and seismologically determined elastic constants of rock, Proceedings of National Academy of Science of USA., Vol. 19, No. 7, pp. 680– 686. http://www.jstor.org/stable/86139.

310

Appendix I:

M AN U

SC

RI PT

297 298

311

Let g(x) be a function and ‘a’ some point from the interior of its domain. Assume

313

that g(x) has derivatives of all orders at ‘a.’ Then, we define its Taylor series at ‘a’

314

by the following formula:

315

T(x) =





316

317

g (k) (a ) (x − a )k k!

AC C

k =0

EP

TE D

312

(A-1)

In the case of a = 0, the series simplifies to the McLaurin series form: T(x) =



∑ k =0

' " "' g (k) (0 ) k (x ) = 1 + g (0) x + g (0) x 2 + g (0) x 3 + ... k! 1! 2! 3!

318

In Equation (5), the function form can be converted as follows:

319

g(x) = (1 − ϕ )f

⇒ g(x) = 1 +

f (− 1) − f (f − 1) 2 ϕ+ ϕ + ... 1! 2!

(A-2)

(A-3)

ACCEPTED MANUSCRIPT

0 < ϕ < 1,

φ2 and higher exponential forms are negligible in

320

As porosity is

321

comparison with the first order of φ, the final equation is:

322

E ≈ E 0 [1 − fϕ ]

RI PT

(A-4)

AC C

EP

TE D

M AN U

SC

323

ACCEPTED MANUSCRIPT

Highlights: - Empirical equation based on two input parameters (dynamic Young's modulus, porosity)

RI PT

- Theoretical backgrounds of porous media were used to explain the meaning of equation

AC C

EP

TE D

M AN U

SC

- We successfully added data from different authors to our own data

Suggest Documents