Next Article in Journal
Homogeneity and Trend Analysis of Climatic Variables in Cap-Bon Region of Tunisia
Previous Article in Journal
Chiral Metasurfaces: A Review of the Fundamentals and Research Advances
Previous Article in Special Issue
Evaluation of Mechanical Performance of a New Disc Spring-Cable Counter Pressure Shock Absorber
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Laboratory Study of Effective Stress Coefficient for Saturated Claystone

1
School of Resources and Environmental Engineering, Hefei University of Technology, Hefei 230009, China
2
Research and Development Center on Technology and Equipment for Energy Conservation and Environmental Protection of Highway Transport, Anhui Transport Consulting & Design Institute Co., Ltd., Hefei 230088, China
3
State Key Laboratory of Geomechanics and Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Sciences, Wuhan 430071, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2023, 13(19), 10592; https://doi.org/10.3390/app131910592
Submission received: 5 August 2023 / Revised: 11 September 2023 / Accepted: 19 September 2023 / Published: 22 September 2023

Abstract

:
Claystone is potentially the main rock formation for the deep geological disposal of high-level radioactive nuclear waste. A major factor that affects the deformation of the host medium is effective stress. Therefore, studying the effective stress principle of claystone is essential for a stability analysis of waste disposal facilities. Consolidated drained (CD) tests were carried out on claystone samples to study their effective stress principle in this paper. Firstly, two samples were saturated under a specified confining pressure and pore pressure for about one month. Secondly, the confining pressure and pore pressure were increased to a specified value simultaneously and then reverted to the previous stress state (the deformations of the samples were recorded during the whole process). Different incremental combinations of the confining pressure and pore pressure were tried at this step. Finally, the effective stress coefficients of the samples were obtained through a back analysis. Furthermore, some potential influencing factors (the neutral stress and loading rate) of the effective stress coefficient were also studied through additional tests. Some interesting results are worth mentioning: (1) the effective stress coefficient of claystone is close to one; (2) the neutral stress and loading rate may have little effect on the effective stress coefficient of claystone.

1. Introduction

As a potential host medium for the deep geological disposal of radioactive nuclear waste, claystone is found under highly complex hydro-mechanical (HM) conditions [1]. Therefore, studying the stress–strain laws of claystone under HM conditions is essential for the stability analysis of radioactive waste disposal repositories. Because of its important role in the stress–strain laws of a geomaterial, the effective stress principle of claystone was studied in this paper.
Effective stress was first proposed by Terzaghi [2], who considered that pore pressure could counteract total stress. The effective stress principle satisfies the following form:
σ e = σ σ p
where σe is the effective stress, σ is the total stress, and σp is the pore pressure.
However, Biot [3] found that Terzaghi’s effective stress principle was unsuitable for porous media with low permeability and believed that pore pressure could not offset the total pressure in a 1:1 ratio. A modified effective stress principle was proposed as follows [3]:
σ e = σ η σ p
where η is the effective stress coefficient of a geomaterial.
Since Biot [3] modified the effective stress principle, numerous studies have investigated it in different soils and rocks [4,5,6,7,8]. Table 1 presents a summary of the various equations used in the literature for the effective stress principle.
The theoretical derivation and empirical formula are the two main expressions of the effective stress principle. The advantage of the theoretical derivation is that an equation that is more consistent with the physical principles can be obtained, but the disadvantage is that some physical quantities in the equation are difficult to measure. The accuracy of the empirical formula is related to the measured physical parameter (the permeability coefficient is favored), which is related to the effective stress coefficient. The adverse effects caused by water pressure changes (the permeability test requires a head difference between the upstream and downstream of the sample) in the permeability test are difficult to evaluate. In this study, the CD tests were conducted to measure the effective stress coefficient of claystone. The samples were firstly saturated under an in situ stress state for about one month. Then the effective stress coefficients were checked by changing the confining pressure and pore pressure in specific increments (the deformations of the samples were recorded during the whole process). Different increments of pore pressure and confining pressure were tried during the effective stress coefficient checking procedures, and an accurate conclusion was obtained through a back analysis (the effective stress coefficient was 0.991 for sample one and 0.995 for sample two). At last, some factors potentially influencing the effective stress principle of claystone were tested and discussed.

2. Experimental Testing

2.1. Experimental Device

The CD tests were performed on a parallel-linkage triaxial testing machine designed by the Institute of Rock and Soil Mechanics, Chinese Academy of Sciences. This machine can implement the mechanical tests simultaneously for two samples with the same confining pressures, the same pore pressures, and different axial pressures. The double-linkage triaxial testing machine is shown in Figure 1a, and the schematic diagram of the increase/decrease in the specimen’s confining, axial, and pore pressure is shown in Figure 1b. The deformations of the samples were measured by the LVDT (Linear Variable Differential Transformer), and the accuracy of the LVDT achieved was 10−7 m.

2.2. Samples’ Preparation and Basic Properties

The adequate claystone samples in this study were trimmed from blocks extracted at a depth of 223 m [24] (the effective stress is about 2.5 MPa). The detailed procedure of the sample preparation is illustrated in Figure 2 (the laboratory temperature is kept constant during the process of the sample preparation). Although we have tried our best to minimize the impact on the sample during the sampling process, the possible impact of unloading (the sample is reduced from in situ stress to atmospheric pressure after coring) cannot be avoided.
The initial physical characteristics of the samples were measured and displayed, as shown in Table 2.
The claystone samples in this paper were relatively homogeneous (Figure 2b). Yu et al. [1] revealed the micro-structure and composition of the same claystone through scanning electron microscopic (SEM) tests and X-ray diffraction tests. In detail, the particles, most of which are clay mineral crystal particles with pores and are in contact from edge to face or face to face, are bending flaky structures [1]. The main composition of them are shown in Table 3 [1].

2.3. Test Procedures

The experimental procedures are as follows:
  • A confining pressure of 0.1 MPa was applied to establish the initial system stabilization (about 24 h).
  • The confining pressure was increased to 2.5 MPa at a low rate of 1.68 kPa/min. The volume changes were monitored by a measuring system.
  • The water was thus injected into the samples from the top and bottom. Then, the pore pressure and confining pressure were simultaneously increased by 0.2 MPa at the rate of 1.68 kPa/min.
  • Simultaneous increases of both the confining pressure to 3.5 MPa and pore pressure to 1.0 MPa were applied. The loading rate was controlled at 0.2 kPa/min.
  • Then Skempton’s coefficient B was checked to determine the saturation degree of the samples by the following method:
    • Closing the drainage valve.
    • Increasing confining pressure of 0.2 MPa.
    • Measuring the increase in the corresponding pore pressure.
    • Decreasing the confining pressure down to 3.5 MPa and calculating Skempton’s coefficient B.
    • Checking Skempton’s coefficient B repeatedly until it was greater than 0.85.
  • Keeping the drainage valve open, the effective stress coefficient of claystone was tested in line with the following steps:
    • Increasing the confining pressure by n MPa (n = 0, 0.1, 0.14, 0.16, 0.18, 0.2, 0.22), and the pore pressure by 0.2 MPa simultaneously. Then the confining pressure and pore pressure were restored to the previous state (σc = 3.5 MPa, σp = 1.0 MPa).
    • Observing and recording the deformation of samples.
In addition to the above test of the effective stress coefficient, this study also attempted to explore the influence of other potential factors (the neutral stresses and loading rate) on the effective stress coefficient. Different loading rates (0.01 kPa/s, 0.02 kPa/s, and 0.04 kPa/s) and a new pressure increment (Δσc = Δσp = 0.5 MPa) were tried in the additional tests.
The variations of confining pressure and pore pressure are illustrated in Figure 3, and the controlled laboratory temperature is shown in Figure 4.

3. Results

As shown in Figure 4, the experiments were divided into two stages: the saturation stage at the beginning and the stage of effective stress coefficient determination (the test stage). The following sections will detail the experimental results of each stage.

3.1. Results of the Saturation Stage

After the water was injected into the samples for about one month, Skempton’s coefficient B was checked to determine the saturation degree. The results are shown in Table 4 and Figure 5.
Usually, Skempton’s coefficient B is below one [25,26]. Therefore, it was assumed that claystone was saturated as long as Skempton’s coefficient B greater than 0.85. Therefore, sample one and sample two were considered to be saturated. It is worth mentioning the following: (1) the claystone studied in this paper can reach saturation in about 20 days [27]; (2) the saturation of the samples were determined after all the tests, and the results showed that the samples were saturated.

3.2. Results of the Test Stage

The test stage was also divided into two stages (Figure 3): the aim of Stage one was to determine the effective stress coefficient, and the aim of Stage two was to study the effect of the potential influencing factors (the neutral stress and loading rate) on the effective stress coefficient.

3.2.1. Results of Stage One

During Stage one, the pore pressure was always increased by 0.2 MPa at a rate of 0.01 kPa/s. However, different increments were tried for the confining pressure. Meanwhile, different increasing rates were used for the confining pressure to make the confining pressure and pore pressure reach their peak values simultaneously. The variations of the pressures and strains during Stage one are recorded in Figure 6 (negative values represent expansion, whereas positive values represent compression).
It is well known that effective stress acts on the skeleton directly. When the effective stress changes, the skeleton will undergo corresponding deformation, which is manifested as when the effective stress decreases, the skeleton of the claystone will expand outwards, and the internal pores will be filled with more water; when the effective stress increases, the framework of the claystone will be compressed, and the pore water will be squeezed out.
An obvious law summarized by the test data is illustrated in Figure 6:
(1)
The samples undergo a significant expansion when the increment of pore pressure Δσp = 0.2, 0.2, 0.2, 0.2, 0.2 MPa is greater than the increment of confining pressure Δσc = 0, 0.1, 0.14, 0.16, 0.18 MPa; This means that the effective stress is reduced under this confining and pore pressure condition.
(2)
Conversely, when Δσp = 0.2 MPa is less than Δσc = 0.22 MPa, these samples are compressed; This means that the effective stress is increased under this confining and pore pressure condition.
(3)
The deformation of the samples are hardly observed when Δσp = 0.2 MPa and Δσc = 0.2 MPa are equal; This means that the effective stress is constant under this confining and pore pressure condition.

3.2.2. Results of Stage Two

At this stage, the increments of pore pressure and confining pressure were 0.5 MPa, and the loading/unloading rate was 0.01 kPa/s during the first cycle; the increments of pore pressure and confining pressure during the second and last cycles were the same as the first, but the loading/unloading rates were 0.02 kPa/s and 0.04 kPa/s, respectively. The variations in the pressures and strains during the whole of Stage two are recorded in Figure 7.
The results showed the following:
(1)
There were slight fluctuations at the end of the loading and unloading. The reason for this phenomenon is that it was impossible to increase or decrease the pore pressure and confining pressure completely simultaneously.
(2)
The sudden jitter of the strain under stable stress conditions between the first and second cycles was due to a temperature fluctuation (Figure 4).
(3)
The deformations of the samples were all not obvious when the pore pressure and confining pressure were both increased or decreased by 0.5 MPa at a rate of 0.01 kPa/s, 0.02 kPa/s, and 0.04 kPa/s, because pore water could dissipate in time and no excess pore pressure would occur.

4. Discussion on the Effective Stress Principle for Saturated Claystone

4.1. The Value of Effective Stress Coefficient for Claystone

The incremental form of Equation (2) can be written as follow:
Δ σ e = Δ σ c η Δ σ p
where Δσe is the increment of effective stress; Δσc is the increment of confining pressure; Δσp is the increment of pore pressure.
The largest linear elastic deformation of the samples during Stage one can be obtained from Figure 6 (Table 5).
(1)
When Δσp = 0.2 MPa and Δσc = 0.18 MPa, the samples are dilatant, meaning Δσe < 0 under this stress condition.
(2)
When Δσp = 0.2 MPa and Δσc = 0.22 MPa, the samples are compressed, meaning Δσe > 0 under this stress condition.
Substituting the above results into Equation (3) gives the following:
0.18 0.2 η < 0 0.22 0.2 η > 0 0.9 < η < 1.1
Berryman [11] presented that the effective stress coefficient of homogeneous rock was not expected to exceed one. Thus, Equation (4) evolves into 0.9 < η < 1.0 .
(3)
When Δσp = 0.2 MPa and Δσc = 0.2 MPa, the deformation of samples are not obvious, meaning Δσe ≈ 0 under this condition.
Then the following equation is deduced:
0.2 0.2 η 0 η 1
The above conclusions indicate that η is a value greater than 0.9 and less than but close to 1.
The effective stress coefficient will be further analyzed according to the above conclusions. The relationship between the stress increments and axial strain increments can be expressed by a generalized Hooke’s law:
Δ ε z = 1 E v [ Δ σ z e v v h ( Δ σ x e + Δ σ y e ) ]
where Δ ε z in the axial strain increment; Δ σ x , Δ σ y , and Δ σ z are the effective stress increments in x, y, and z direction; E v is the vertical elastic modulus; v v h is the Poison ratio in the horizontal and vertical plane.
According to the results of the stress (Figure 6) and strain (Table 5) variations, the back analysis was performed to determine the effective stress coefficient. The FMINSEARCH function of MATLAB (version number: 8.3) was selected for the optimization calculation. The optimization calculation process is shown in Figure 8, and the objective function of the back analysis can be expressed as the following:
i = 1 n Δ ε z i 1 E v [ Δ σ z i e v v h ( Δ σ x i e + Δ σ y i e ) ] 2 min
To ensure the accuracy of this back analysis, the range of physical parameters is shown in Table 6.
The analysis results showed that the effective stress coefficient of claystone was 0.991 for sample one and 0.995 for sample two.

4.2. Discussion on the Influencing Factors of Effective Stress Coefficient

The effective stress coefficient of claystone was probably related to the testing conditions.
The results of Section 3.2.2 showed that the deformations of the samples were not obvious when Δσc = Δσp = 0.5 MPa under different loading rates (0.01 kPa/s, 0.02 kPa/s, 0.04 kPa/s), indicating that the effective stress coefficient of claystone was still close to 1 even if different pressure increments and loading rates were taken. Therefore, the following conclusion can be drawn:
(1)
Whether the testing condition was Δσc = Δσp = 0.2 MPa or Δσc = Δσp = 0.5 MPa, the results were the same. It means that the neutral stresses may have little effect on the determination of the effective stress coefficient for claystone.
(2)
The results with different loading rates showed an inconspicuous deformation. We may conclude that as long as the loading rates are low enough so that the pore pressure can be dissipated time-efficiently, it will not affect the effective stress coefficient of claystone.

5. Conclusions

The effective stress principle of claystone was studied using the CD tests in this paper.
  • Several loading–unloading cycles were conducted on the saturated claystone samples, and the deformation during this procedure was recorded. The results showed that both of the two samples were dilatant when the increments of pore pressure were greater than the increments of confining pressure; conversely, these samples were compressed when the increments of pore pressure were less than the increments of confining pressure; finally, the deformation of the sample was hardly observed when the increments of pore pressure and confining pressure were equal. The conclusion is that the effective stress coefficient of claystone is close to one.
  • The relationship between effective stress and strain was analyzed to determine the effective stress coefficient of claystone. The results showed that the coefficient was 0.991 for sample one and 0.995 for sample two.
  • The relationship between the effective stress coefficient of claystone and the potential influencing factors (the neutral stress and loading rate) were studied. The results show that the neutral stress and loading rate may have little effect on the effective stress coefficient of claystone.

Author Contributions

Conceptualization, F.L. and W.C.; methodology, F.L.; validation, F.L., W.C. and Z.W.; formal analysis, F.L.; investigation, F.L.; resources, H.Y.; data curation, M.L.; writing—original draft preparation, F.L.; writing—review and editing, W.C. and Z.Z.; visualization, F.L.; supervision, F.Z.; project administration, W.C.; funding acquisition, W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by The National Natural Science Foundation of China, grant numbers 51879258 and 42377179.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request.

Conflicts of Interest

Authors Fanfan Li, Zhigang Wu, Ming Li and Zhifeng Zhang were employed by the company ATCDI (Anhui Transport Consulting & Design Institute CO., LTD). The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could beconstrued as a potential conflict of interest.

References

  1. Yu, H.D.; Chen, W.Z.; Jia, S.P.; Cao, J.J.; Li, X.L. Experimental study on the hydro-mechanical behavior of Boom clay. Int. J. Rock Mech. Min. Sci. 2012, 53, 159–165. [Google Scholar] [CrossRef]
  2. Ghabezloo, S.; Sulem, J.; Guédon, S.; Martineau, F. Effective stress law for the permeability of a limestone. Int. J. Rock Mech. Min. Sci. 2009, 46, 297–306. [Google Scholar] [CrossRef]
  3. Biot, M.A. General theory of three-dimensional consolidation. J. Appl. Phys. 1941, 12, 155–164. [Google Scholar] [CrossRef]
  4. Skempton, A.W. Selected Papers on Soil Mechanic; Thomas Telford Publishing: London, UK, 1984. [Google Scholar]
  5. Taghizadeh, R.; Goshtasbi, K.; Manshad, A.K.; Ahangari, K. Geomechanical and thermal reservoir simulation during steam flooding. Struct. Eng. Mech. 2018, 66, 505–513. [Google Scholar] [CrossRef]
  6. Sonmezer, Y.B. Energy-based evaluation of liquefaction potential of uniform sands. Geomech. Eng. 2019, 17, 145–156. [Google Scholar] [CrossRef]
  7. Lee, H.; Oh, T.M.; Park, C. Analysis of permeability in rock fracture with effective stress at deep depth. Geomech. Eng. 2020, 22, 375–384. [Google Scholar] [CrossRef]
  8. Guerriero, V.; Mazzoli, S. Theory of effective stress in soil and rock and implications for fracturing processes: A review. Geosciences 2021, 11, 119. [Google Scholar] [CrossRef]
  9. Nur, A.; Byerlee, J.D. An exact effective stress law for elastic deformation of rock with fluids. J. Geophys. Res. 1971, 76, 6414–6419. [Google Scholar] [CrossRef]
  10. Zoback, M.D.; Byerlee, J.D. Permeability and effective stress. AAPG Bull. 1975, 59, 154–158. [Google Scholar] [CrossRef]
  11. Berryman, J.G. Effective stress for transport properties of inhomogeneous porous rock. J. Geophys. Res. Solid Earth 1992, 97, 17409–17424. [Google Scholar] [CrossRef]
  12. Tuncay, K.; Corapcioglu, M.Y. Effective stress principle for saturated fractured porous media. Water Resour. Res. 1995, 31, 3103–3106. [Google Scholar] [CrossRef]
  13. George, J.D.; Barakat, M.A. The change in effective stress associated with shrinkage from gas desorption in coal. Int. J. Coal Geol. 2001, 45, 105–113. [Google Scholar] [CrossRef]
  14. Kwon, O.; Kronenberg, A.K.; Gangi, A.F.; Johnson, B. Permeability of Wilcox shale and its effective pressure law. J. Geophys. Res. Solid Earth 2001, 106, 19339–19353. [Google Scholar] [CrossRef]
  15. Tsuji, T.; Tokuyama, H.; Pisani, P.C.; Moore, G. Effective stress and pore pressure in the Nankai accretionary prism off the Muroto Peninsula, southwestern Japan. J. Geophys. Res. Solid Earth 2008, 113, 401. [Google Scholar] [CrossRef]
  16. Bagherieh, A.R.; Khalili, N.; Habibagahi, G.; Ghahramani, A. Drying response and effective stress in a double porosity aggregated soil. Eng. Geol. 2009, 105, 44–50. [Google Scholar] [CrossRef]
  17. Nowamooz, H.; Ho, X.N.; Chazallon, C.; Hornych, P. The effective stress concept in the cyclic mechanical behavior of a natural compacted sand. Eng. Geol. 2013, 152, 67–76. [Google Scholar] [CrossRef]
  18. Konrad, J.M.; Lebeau, M. Capillary-based effective stress formulation for predicting shear strength of unsaturated soils. Can. Geotech. J. 2015, 52, 2067–2076. [Google Scholar] [CrossRef]
  19. Zhang, C.L. Examination of effective stress in clay rock. J. Rock Mech. Geotech. Eng. 2016, 9, 479–489. [Google Scholar] [CrossRef]
  20. Saurabh, S.; Harpalani, S. The effective stress law for stress-sensitive transversely isotropic rocks. Int. J. Rock Mech. Min. Sci. 2018, 101, 69–77. [Google Scholar] [CrossRef]
  21. Ma, J.; Querci, L.; Hattendorf, B.; Saar, M.; Kong, X.Z. The effect of mineral dissolution on the effective stress law for permeability in a tight sandstone. Geophys. Res. Lett. 2020, 47, e2020GL088346. [Google Scholar] [CrossRef]
  22. Civan, F. Effective-stress coefficients of porous rocks involving shocks and loading/unloading hysteresis. SPE J. 2021, 26, 44–67. [Google Scholar] [CrossRef]
  23. Zhao, Z.; Chen, S.; Chen, Y.; Yang, Q. On the effective stress coefficient of single rough rock fractures. Int. J. Rock Mech. Min. Sci. 2021, 137, 104556. [Google Scholar] [CrossRef]
  24. Dehandschutter, B.; Vandycke, S.; Sintubin, M.; Vandenberghe, N.; Wouters, L. Brittle fractures and ductile shear bands in argillaceous sediments: Inferences from Oligocene Boom Clay (Belgium). J. Struct. Geol. 2005, 27, 1095–1112. [Google Scholar] [CrossRef]
  25. Mesri, G.; Adachi, G.; Ullrich, C.R. Pore-pressure response in rock to undrained change in all-round stress. Géotechnique 1976, 26, 317–330. [Google Scholar] [CrossRef]
  26. Dropek, R.K.; Johnson, J.N.; Walsh, J.B. The influence of pore pressure on the mechanical properties of Kayenta sandstone. J. Geophys. Res. Solid Earth 1978, 83, 2817–2824. [Google Scholar] [CrossRef]
  27. Yu, H.D.; Chen, W.Z.; Gong, Z.; Ma, Y.S.; Chen, G.J.; Li, X.L. Influence of temperature on the hydro-mechanical behavior of Boom Clay. Int. J. Rock Mech. Min. Sci. 2018, 108, 189–197. [Google Scholar] [CrossRef]
  28. Chen, G.J.; Sillen, X.; Verstricht, J.; Li, X.L. ATLAS III in situ heating test in boom clay: Field data, observation and interpretation. Comput. Geotech. 2011, 38, 683–696. [Google Scholar] [CrossRef]
Figure 1. Double-linkage triaxial testing machine: (a) actual test equipment; (b) test equipment schematic diagram.
Figure 1. Double-linkage triaxial testing machine: (a) actual test equipment; (b) test equipment schematic diagram.
Applsci 13 10592 g001
Figure 2. The detailed procedure of hand-trimming: (a) claystone core; (b) pre-cut prism; (c) grinding the sample; (d) the sample; (e) sealing the sample.
Figure 2. The detailed procedure of hand-trimming: (a) claystone core; (b) pre-cut prism; (c) grinding the sample; (d) the sample; (e) sealing the sample.
Applsci 13 10592 g002
Figure 3. A detailed diagram of the test procedure.
Figure 3. A detailed diagram of the test procedure.
Applsci 13 10592 g003
Figure 4. The temperature of the pressure chamber.
Figure 4. The temperature of the pressure chamber.
Applsci 13 10592 g004
Figure 5. The pressure curve during the saturation testing: (a) the pressure curve during the saturation; (b) testing the saturation for the first time; (c) testing the saturation for the second time.
Figure 5. The pressure curve during the saturation testing: (a) the pressure curve during the saturation; (b) testing the saturation for the first time; (c) testing the saturation for the second time.
Applsci 13 10592 g005aApplsci 13 10592 g005b
Figure 6. The pressures and strains curve during Stage one: (a) sample one; (b) sample two.
Figure 6. The pressures and strains curve during Stage one: (a) sample one; (b) sample two.
Applsci 13 10592 g006
Figure 7. The pressures and strains curve during Stage two: (a) sample one; (b) sample two.
Figure 7. The pressures and strains curve during Stage two: (a) sample one; (b) sample two.
Applsci 13 10592 g007
Figure 8. Optimization calculation flow diagram.
Figure 8. Optimization calculation flow diagram.
Applsci 13 10592 g008
Table 1. Effective stress principle of references.
Table 1. Effective stress principle of references.
ReferencesEffective StressRelated Research Work
Nur and Byerlee [9]Saturated: η = 1 – (K/Ks);
K is the bulk moduli of rock; Ks is the bulk moduli of grain.
The rock strain was quantified based on this effective stress principle, and its validity was verified via compression test results of sandstone and granite.
Zoback and Byerlee [10]Saturated: A variable greater than 1 and varies with stress.The phenomenon that pore pressure has a greater impact on permeability than confining pressure was found according to the permeability test results of sandstone under different confining and pore pressures.
Berryman [11]Saturated: η = K K ( 1 ) K ( 2 ) η ( 1 ) η ( 2 ) + θ ;
K is the bulk modulus of drained porous frame; K(1) and K(2) are the drained frame moduli of porous constituents (1) and (2); η(1) and η(2) are the effective stress coefficient of porous constituents (1) and (2); θ is the relative change in the effective stress coefficient for a two-component porous medium.
The effective stress coefficients of a two-component porous medium are deduced based on the existing research work, and the phenomenon of Zoback and Byerlee [10] is explained based on this conclusion.
Tuncay and Corapcioglu [12]Saturated: σ e = σ η 1 σ p f η 2 σ p p ; η 1 = 1 α f K f r K f r m 1 ; η 2 = K f r K s 1 α f K f r K f r m ;
η1 is the effective stress coefficient for the fractures; η2 is the effective stress coefficient for the pores; σ p f is the pore pressure in the fractures; σ p p is the pore pressure in the pores; α f is the volume fraction of the fractures; K s is the bulk moduli of grain; K f r is the drained bulk modulus of the fractured porous medium; K f r m is the drained bulk modulus of the nonfractured porous medium.
An effective stress principle for saturated fractured porous media is proposed based on the assumption of linear elasticity.
George and Barakat [13]Saturated: η = 0.71;A series of loading–unloading (both total stress and pore pressure) cycles applied by gas pressure were performed on the coal specimen, and the effective stress coefficient was obtained.
Kwon et al. [14]Saturated: η = log k / σ p log k / σ ;
k is the permeability.
The effective stress coefficient of illite-rich shale was determined according to the transient pulse test results of saturated samples.
Tsuji et al. [15]Saturated: Terzaghi’s effective stress principle.A theoretical relationship between the acoustic velocity and mean effective stress of the Nankai accretionary prism was calculated using DEM (Differential Effective Medium) theory and the aspect ratio spectrum of pore space, and Terzaghi’s effective stress principle was introduced to analyze the pore pressure distribution of Nankai trough.
Bagherieh et al. [16]Unsaturated: σ e = σ α m χ m u m w + 1 χ m u m a I α M χ M u M w + 1 χ M u M a I χ m = S m e u m a u m w Ω f o r S m S m e 1 f o r S m S m e χ M = S M e u M a u M w Ω f o r S M S M e 1 f o r S M S M e ;
u m w , u m a , u M w , and u M a are the micropore water, micropore air, macro pore water, and macro pore air pressures; α m and α M are the conventional effective stress coefficients of saturated double-porous media; S m e is the suction value separating saturated from unsaturated conditions in the micropores; S M e is the matric suction value separating saturated from unsaturated conditions in the macro pores; Ω is the material parameter.
Drying and one-dimensional consolidation tests are performed on initially saturated samples of the kaolin (double-porosity compacted soil) at different net stresses, and the test results were accurately predicted using the new effective stress principle.
Ghabezloo et al. [2]Saturated: η = c σ d + d ;
c and d are the fitting parameters; σ d is the differential stress.
The effective stress law for the permeability of limestone is studied by drained hydrostatic compression and constant-head permeability tests, and the results show that the effective stress coefficient was linearly related to differential stress.
Nowamooz et al. [17]Unsaturated: σ = σ e + σ a + η σ w σ a η = tan ϕ b tan ϕ ;
σ w is the pore water pressure; σ a is the pore air pressure; ϕ b is the angle indicating the increase in the matrix suction’s shear stress function; ϕ is the lowest matrix suction.
The effective stress law of compacted natural clay sand in Missillac was obtained using the direct shear test and the soil water retention curve (SWRC), and its validity was verified by simulating the resilient behavior of repeated-load triaxial tests.
Konrad and Lebeau [18]Unsaturated: σ = σ e + σ a + η σ w σ a η = A S r + e * B 2 cos θ ;
A S r is the fraction of the fracture surface wetted by water; e * is the maximum local aperture currently occupied by water; B is the total perimeter of water menisci between two fracture walls; θ is the contact angle of the fracture water.
The effective stress equation is derived for partially saturated rough-walled fractures with any aperture distributions based on the capillary law, and the nonlinearity of the effective stress parameter versus the saturation degree curves is found to be mainly determined by the surface roughness (or the coefficient of variation) rather than the mean aperture of the fractures.
Zhang [19]Saturated: σ A = σ i s A s + σ l A l + σ p A p ;
σ A is the total normal force externally applied on the surface A ; σ i s A s is the total normal force acting on the solid– solid contact area, A s , with the local total stress σ i s ; σ l A l is the total normal repulsive force acting in the water-film section, A l , between clay particles, with the local total pressure σ l ; σ p A p is the pressure σ p acting on the surface A p in the large pores occupied by free water.
The effective stress in a dense clay–water system is transferred through both the adsorbed interparticle porewater in narrow pores and the solid–solid contact between non-clay mineral grains. This concept has been widely validated by various kinds of experiments performed on the COX and OPA claystones.
Saurabh and Harpalani [20]Unsaturated: σ 11 = C 11 ε 11 + C 12 ε 22 + C 12 ε 33 η 1 p c 1 η 1 s m a 1 p σ 22 = C 12 ε 11 + C 11 ε 22 + C 13 ε 33 η 1 p c 1 η 1 s m a 1 p σ 33 = C 13 ε 11 + C 13 ε 22 + C 33 ε 33 η 3 p c 1 η 3 s m a 3 p σ 23 = 2 C 44 ε 23 ; σ 31 = 2 C 44 ε 31 ; σ 12 = 2 C 11 C 12 2 ε 12 η 1 = C 13 d ε 3 d p d σ h d p d s m a 1 p d p 1 d s m a 1 p d p ; η 1 = C 33 d ε 3 d p d s m a 3 p d p 1 d s m a 1 p d p ;
σ i j , ε i j and C i j are the total stress, compliance matrix, and strain along the i j direction, respectively; η 1 and η 3 are the effective stress coefficients of transversely isotropic media; p is the pore pressure; p c is the pore pressure in the cleat system of media; s m a p is a pressure-dependent quantity coupling the sorption-based stress and strain in a microporous media.
The estimated values of the effective stress coefficients in both the vertical and horizontal directions are different, varying with pressure for methane depletion, and a conceptual physical model of effective stress, which considered absorption, was proposed.
Ma et al. [21]Saturated: σ e = a σ p 2 a σ + b σ p + σ ;
a and b are the fitting parameters.
A series of flow-through experiments consisting of three continuous stages (pre-reaction stage, reaction stage, and post-reaction stage) were conducted, and the cross-plot method was introduced to determine the effective law of tight sandstone with mineral dissoluble mineral.
Civan [22]Saturated: η 1 η = a K ϕ b ;
a and b are the fitting parameters; K is the intrinsic permeability of porous rock; ϕ is the porosity of porous formation.
The modified power-law equation yields a physically meaningful correlation because it successfully satisfies the low-end- and high-end-limit values of the effective stress coefficient and also provides a better quality match of the available experimental data than the semilogarithmic equation and the popular basic power-law equation.
Zhao et al. [23]Saturated: η = 1 1 + σ σ p k n 0 u n max σ σ p 1 σ < σ p ;
k n 0 is the initial normal stiffness; u n max is the maximum fracture closure.
A new effective stress coefficient model for single rough water-bearing fractures is proposed in terms of initial normal stiffness and maximum normal closure, and it was verified by laboratory and in situ experimental data.
Table 2. Initial condition for the tested samples.
Table 2. Initial condition for the tested samples.
NumberHeight (mm)Diameter (mm)Density (g/cm3)Dry Density (g/cm3)Water
Content
Void RatioInitial Saturation
Sample one76.2638.362.021.6820.68%0.6489.06%
Sample two75.8137.812.031.6919.46%0.6380.28%
Table 3. The main composition of the claystone [1].
Table 3. The main composition of the claystone [1].
Mineral CompositionIlliteKaoliniteIllite-Smectite
Mix Layer
ChloriteQuartzCalcite,
Dolomite
Feldspar
Proportion (%)151520251564
Table 4. The Skempton’s coefficient B for the samples.
Table 4. The Skempton’s coefficient B for the samples.
Sample NumberFirst TimeSecond Time
Sample one0.9130.925
Sample two0.9070.914
Table 5. The largest linear elastic deformation of the two samples during each cycle.
Table 5. The largest linear elastic deformation of the two samples during each cycle.
The Number of CyclesDeformation of Sample OneDeformation of Sample Two
Axial StrainRadial StrainVolumetric StrainAxial StrainRadial StrainVolumetric Strain
The first−4.95−0.81−6.57−5.24−0.73−6.70
The second−2.36−0.23−2.82−2.46−0.30−3.06
The third−1.36−0.07−1.50−1.49−0.17−1.83
The fourth−0.89−0.07−1.03−1.01−0.10−1.21
The fifth−0.45−0.02−0.49−0.51−0.03−0.57
The sixth≈0≈0≈0≈0≈0≈0
The last0.460.080.620.470.110.69
Note: Negative values represent expansion; positive values represent compression; the unit of these strains are 10−4.
Table 6. The range of physical parameters before back analysis [28].
Table 6. The range of physical parameters before back analysis [28].
Physical ParametersEv/MPaVvhη
Range300–8000.1–0.30.9–1.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, F.; Chen, W.; Wu, Z.; Yu, H.; Li, M.; Zhang, Z.; Zha, F. Laboratory Study of Effective Stress Coefficient for Saturated Claystone. Appl. Sci. 2023, 13, 10592. https://doi.org/10.3390/app131910592

AMA Style

Li F, Chen W, Wu Z, Yu H, Li M, Zhang Z, Zha F. Laboratory Study of Effective Stress Coefficient for Saturated Claystone. Applied Sciences. 2023; 13(19):10592. https://doi.org/10.3390/app131910592

Chicago/Turabian Style

Li, Fanfan, Weizhong Chen, Zhigang Wu, Hongdan Yu, Ming Li, Zhifeng Zhang, and Fusheng Zha. 2023. "Laboratory Study of Effective Stress Coefficient for Saturated Claystone" Applied Sciences 13, no. 19: 10592. https://doi.org/10.3390/app131910592

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop