Next Article in Journal
Development and Simulation of a Type of Four-Shaft ECVT for a Hybrid Electric Vehicle
Previous Article in Journal
An Automatic Switched-Capacitor Cell Balancing Circuit for Series-Connected Battery Strings
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Numerical Investigation of Influence of In-Situ Stress Ratio, Injection Rate and Fluid Viscosity on Hydraulic Fracture Propagation Using a Distinct Element Approach

1
Key Laboratory of Shale Gas and Geoengineering, Institute of Geology and Geophysics, Chinese Academy of Science, Beijing 100029, China
2
College of Earth Science, University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Energies 2016, 9(3), 140; https://doi.org/10.3390/en9030140
Submission received: 21 November 2015 / Revised: 3 February 2016 / Accepted: 16 February 2016 / Published: 27 February 2016

Abstract

:
Numerical simulation is very useful for understanding the hydraulic fracturing mechanism. In this paper, we simulate the hydraulic fracturing using the distinct element approach, to investigate the effect of some critical parameters on hydraulic fracturing characteristics. The breakdown pressure obtained by the distinct element approach is consistent with the analytical solution. This indicates that the distinct element approach is feasible on modeling the hydraulic fracturing. We independently examine the influence of in-situ stress ratio, injection rate and fluid viscosity on hydraulic fracturing. We further emphasize the relationship between these three factors and their contributions to the hydraulic fracturing. With the increase of stress ratio, the fracture aperture increases almost linearly; with the increase of injection rate and fluid viscosity, the fracture aperture and breakdown pressure increase obviously. A low value of product of injection rate and fluid viscosity (i.e., ) will lead to narrow fracture aperture, low breakdown pressure, and complex or dispersional hydraulic fractures. A high value of would lead wide fracture aperture, high breakdown pressure, and simple hydraulic fractures (e.g., straight or wing shape). With low viscosity fluid, the hydraulic fracture geometry is not sensitive to stress ratio, and thus becomes a complex fracture network.

1. Introduction

Hydraulic fracturing technology is widely used in the unconventional energy field [1,2,3]. In recent years, hydraulic fracturing technology has played a critical role in the shale gas revolution [4,5]. The in-situ stress ratio, injection rate, and fluid viscosity all have a crucial effect on the hydraulic fracturing results. King [6] reviewed the 30-year experience in hydraulic fracturing of shale in North America, and showed that the desired hydraulic-fracturing effect depends on not only rock mechanics in the shale reservoir, but also on in-situ stress state and operational parameters. Some researchers [7,8,9] have investigated the influence of injection rate on hydraulic fractures through laboratory experiments, which indicated that under the same experimental conditions, with the increase of injection rate, the branches of the hydraulic fracture would decrease, and the complex fracture network turns into a single straight fracture. Ishida et al. [10,11,12,13], Fan and Zhang [9], and Hou et al. [14], have performed laboratory experiments on the effect of fluid viscosity. Warpinski et al. [15] have analyzed the microseismic monitoring data in hydraulic fracturing sites. These studies indicated that under the condition of low fluid viscosity (such as the case of supercritical CO2), the hydraulic fracture becomes flexed with multiple branches and develops a complex fracture network. With the increase of fluid viscosity (such as the case of a gel), the branches disappear and hydraulic fracture turns into a simple wing shape. We note that in these studies, only Ishida et al. [10,11,12,13] used the supercritical CO2 and liquid CO2 as fracturing fluid of which the viscosity is smaller than 1 mPa·s. Supercritical CO2 is important but difficult to use in hydraulic fracturing [16], thus, we need to perform systematic studies to enhance the theoretical basis for unconventional resource production. Meanwhile, the relationship between stress ratio, injection rate, and fluid viscosity should be further investigated and analyzed.
Numerical simulation [17,18,19,20,21,22,23,24,25,26,27,28] would avoid some of the randomness in laboratory experiments and could reflect the effect of critical parameters on hydraulic fracturing characteristics. Many works have been carried out to investigate the influence of some important parameters on hydraulic fracturing by numerical simulation. Wangen [29] established a two-dimensional numerical model based on the finite element method and further analyzed the bottom-hole pressure and the number of failed elements under different permeability conditions. It was found that under conditions of low permeability, the bottom-hole pressure rises more quickly and more elements fail. Carrier and Granet [30] developed a cohesive zone to govern the fracture propagation in the finite element analysis framework. Carrier and Granet’s numerical results about injection pressure and fracture aperture are very consistent with analytical solutions under a zero fluid lag assumption. Zhang et al. [31] set up a 3D nonlinear fluid-mechanics coupling model, by using finite element software ABAQUS. The field data from the Daqing (in China) oilfield were used in numerical computations, and the evolution of the bottomhole pressure in the simulation was coincident with the corresponding data from the field measurements. Subsequently, the influence of some critical parameters on hydraulic fracturing characteristics were investigated. Using the finite difference software Flac3D, Zhou and Hou [32] established a model in which the leak-off could be simulated in both rock matrix and fractures. They verify the modeling approach by comparing the borehole pressure between numerical and laboratory results. They further applied this approach on a tight gas sandstone reservoir from the Northern German Basin. In addition, some commercial software based on a continuum were launched for design and prediction of hydraulic fracturing, such as StimPlan, FRANC3D, and GOHFER. However, the relationships between stress ratio, injection rate, and fluid viscosity have not been comprehensively investigated, and the function of supercritical CO2 with low fluid viscosity is still not clear. Meanwhile, previous studies conducted with discontinuum numerical method have rarely been reported.
In this paper, the Universal Distinct Element Code (UDEC) [33], a distinct element method (DEM) [34], was used to simulate hydraulic fracture initiation and propagation. The numerical solution obtained is consistent with the analytical solution of the breakdown pressure, which indicates that the distinct element method is feasible for modeling the hydraulic fracturing. Then, based on our numerical model, the influence of stress ratio, injection rate, and fluid viscosity on hydraulic fracturing are independently investgated by tuning one while retaining the other parameters. Finally, the relationships between the above three parameters are analyzed.

2. Numerical Model and Computational Scheme

2.1. Algorithms Introduction

The DEM is a numerical simulation method which is based on a discontinuum approach. It was put forward by Cundall [34] in 1971, and now has become an effective method for dealing with geotechnical problems. The DEM allows finite displacements and rotations of discrete bodies including complete detachment, and recognizes new contacts automatically as the calculation progresses [35]. With the DEM software Particle Flow Code (PFC), some works have been carried out to simulate the hydraulic fracturing characteristics [36,37,38,39]. However, there are some disadvantages in simulating hydraulic fracturing using PFC. The PFC assumes that each element acts as a rounded particle; thus, the PFC could not well describe the contact relationships between particles. In other words, the PFC does not satisfy the failure criterion, for example, the Mohr-Coulumb failure criterion. In addition, the PFC would generate huge amounts of particles when simulating hydraulic fracturing on a large scale, which results in an extremely huge computational cost. In contrast, the UDEC is immune to these disadvantages because it assumes that each particle is polygonal. The particle, or block, in the UDEC is impermeable, and the fluid flows only in the contacts between blocks, which is consistent with the physical characteristics of the extremely low permeability of the shale matrix. Using UDEC, Choi [40] proposed several methods for obtaining shut-in pressure from the ambiguous shut-in curves, under various remote stress regimes and various rock properties. Zangeneh et al. [41] investigated the interaction between the natural fractures and hydraulic fracture, and the results show that the natural fractures influence significantly the size, orientation, and path of the hydraulic fracture.
In the UDEC, the research object is discretized into rigid or deformable blocks by one group or several groups of discontinuity. If the blocks are deformable, they will be subdivided into a mesh of finite-difference elements, and each element responds according to a prescribed linear or nonlinear stress-strain law. These rigid or deformable blocks are considered to be impermeable to the flow of fluid restricted to the contacts. For edge-to-edge contacts in fluid-mechanical analyses, the calculation of flow rate q obeys the cubic law [33]:
q = a 3 12 μ Δ p L
where q is the flow volume between contacts per unit time, a is the joint hydraulic aperture, μ is the dynamic viscosity of fluid, Δ p is the pressure difference between the domains, and L is the contact length.
The contact conductivity depends on deformation and displacement, which is in turn influenced by the fluid pressure. Figure 1 shows the relationship between aperture and normal stress. Given the hydraulic aperture at zero normal stress azero, the minimum allowable joint hydraulic aperture (residual aperture) ares, and the maximum allowable joint hydraulic aperture amax, we can simulate the hydraulic fracture. When the hydraulic aperture exceeds azero, or, the effective stress acting on the joint is 0, the hydraulic fracture is assumed to be open and will propagate along the contacts between blocks.

2.2. Model Setup and Parameters Acquisition

A two dimensional model was built with UDEC, to investigate the hydraulic fracturing characteristics of laboratory-scale shale. According to laboratory hydraulic fracturing experiments [7,14,42,43], the model dimension was confirmed to be 300 mm × 300 mm. There is a circular hole that is 20 mm in diameter at the center of the model (Figure 2), from which the fluid was injected at a constant rate. The fluid is considered as incompressible with transient flow. According to references [40,41], an irregular distribution of elements leads to an irregular hydraulic fracture. This would disturb the influence of critical factors we investigated here. To eliminate the influence of the irregular shape, the model is discretized into 900 squares with 10 mm side length. In view of the elastic-brittle mechanical characteristic of rock, the blocks are considered to be elastic to reflect the deformation characteristic of the model. The contacts are considered to be elastic-plastic with Mohr-Coulomb failure criterion. When interacting with natural fracture, hydraulic fracture will turn to be asymmetrical and complicated. Therefore, the influence of natural fractures on hydraulic fracturing was not considered in our model. As the research object is intact rock, the contact strength parameters are assumed to be high (say close to the intact rock strength), which makes the mechanical response of the numerical model close to the intact shale. Hydraulic fracture will propagate along the contacts between blocks, and the discrete block may deform and move.
To investigate the laboratory mechanical and physical properties and provide accurate parameters for numerical modeling of hydraulic fracturing, shale specimens were collected from the outcrops of the Longmaxi Formation in Shizhu County (in China). The values of some parameters in Table 1 are rounded based on above laboratory experiments.
The contact strength parameters are set to be high to make the mechanical response of the numerical model close to the intact rock. The normal stiffness of contacts, kn, was calculated [44]:
k n = n [ k + ( 4 / 3 ) g Δ Z m i n ] ,    1 n 10
where k is bulk modulus, g is shear modulus, and Δ Z min is the smallest width of the zone adjoining the contact.
The operational parameters, injection rate, and fluid viscosity are shown in Table 2. The stress boundary condition in the numerical simulation is that the major horizontal stress was set equal to 15 MPa (σH = 15 MPa), and the minor horizontal stress was set equal to 10 MPa (σh = 10 MPa).

2.3. Simulation Schemes

Based on the above configurations of the model, the initiation and propagation of hydraulic fracture for the numerical model was analyzed using UDEC, and the obtained numerical solution of the breakdown pressure was compared with the analytical solution. Then, the independent influence of stress ratio, injection rate, and fluid viscosity on hydraulic fracturing characteristics was investigated within reasonable ranges. Finally, the relation between above three parameters was analyzed.

3. Simulation and Verification

3.1. Initiation and Propagation of Hydraulic Fracture

When fluid is first injected in the model, injection pressure rises linearly with time, as shown in Figure 3. When the running time equaled 1.5 s and the injection pressure reached 9.81 MPa, two hydraulic fractures initiated simultaneously on both sides of the hole (the grids were hidden). This indicates that the breakdown pressure should be 9.81 MPa. With the failure of contacts between blocks, the injection pressure was suddenly released and fluid leakage occurred because the aperture grew, which leads to an obvious pressure drop at 2 s. At 6 s (Figure 4), the hydraulic fracture further propagated. Every extension of the hydraulic fracture coupled with a corresponding pressure drop.
The breakdown pressure Pb was calculated [45]:
P b = 3 σ h - σ H - α ( 1 - 2 υ ) / ( 1 - υ ) + T 2 - α ( 1 - 2 υ ) / ( 1 - υ ) ,    0 α 1
where α is the poroelastic coefficient, υ is the Poisson ratio, and T is the rock tensile strength. We test α = 0.15 and 0.24 to calculate the breakdown pressure, respectively. These two cases (i.e., α = 0.15 and 0.24) can be considered as analytical solution, which were compared with our numerical solution to verify the reliability of the numerical simulation (Figure 5). Stress ratio K was defined to be σHh, where σH is the major horizontal stress and σh is the minor horizontal stress.
The relation between breakdown pressure and K was showed in Figure 5, where we kept σh = 10 MPa for simplicity. As shown in Figure 5, with the increase of stress ratio, the analytical and numerical solutions of breakdown pressure decreased, and the numerical solution decreases much more slowly. It is worth noting that when K is between 1.3 and 1.6, the numerical solution is close to the analytical solution. Therefore, the stress ratio for the basic model was set to 1.5, to assure the accuracy and reliability of the numerical simulation.

3.2. Influence of Stress Ratio

In this section, the influence of stress ratio on hydraulic fracturing characteristics was investigated on the condition that the running time equaled to 15 s and other parameters in the basic model remained unchanged. As shown in Figure 6, when the stress ratio K = 1.1~1.2 (σh is kept 10 MPa), the fractures were divergent and there were small branch cracks on both sides of the main hydraulic fracture. When the stress ratio K = 1.3, the main hydraulic fracture propagated, and most branch cracks disappeared. When the stress ratio K ≥ 1.4, only the main hydraulic fracture existed, and the branch cracks totally disappeared. The main reason for this result is that hydraulic fracture usually propagates along the direction perpendicular to minimum principal stress. When the stress ratio K is close to 1, i.e., the stress field is nearly isotropic, the branches of hydraulic fracture will appear. Whereas, with the increase of stress ratio K, the stress anisotropy is more and more significant; thus, those branch cracks would disappear.
Figure 7 shows the relation between hydraulic fracture aperture and stress ratio. Obviously, the aperture increases with increasing K. The increase of stress ratio would cause the decrease of breakdown pressure; thus, the earlier generated hydraulic fracture would lead to a much wider aperture for the same amount of time.

3.3. Influence of Injection Rate

In this section, the influence of injection rate on hydraulic fracturing characteristics was investigated on the condition that running time equaled to 15 s and the other parameters in the basic model remained unchanged. As shown in Figure 8, when Q = 8 × 10−6 m3/s/m, the branch crack initiated. When Q ≥ 9 × 10−6 m3/s/m, branch cracks disappear and the main hydraulic fracture extends significantly with an increasing injection rate. The lower injection rate results in the more lateral fluid leak-off, which makes a much smaller branch crack. As shown in Figure 9a, the hydraulic fracture aperture is also affected by injection rate, and the aperture would increase obviously with increasing injection rate. Figure 9b shows that with the increasing injection rate, the breakdown pressure increases almost linearly. When the injection rate is high, the loading condition would approach to dynamic procedure; thus, the breakdown pressure needed for hydraulic fracture initiation is higher. This phenomenon is consistent with the rock uniaxial compression tests.
Some researchers have investigated the influence of injection rate on hydraulic fractures taking water [7], viscasil [8], and gel [9] as fracturing fluid. These studies indicated similar conclusions that under the same experiment conditions, with the increase of injection rate, the branches of the hydraulic fracture decrease, and the complex fracture network turns into a single straight fracture.

3.4. Influence of Fluid Viscosity

In this section, the influence of fluid viscosity on hydraulic fracturing characteristics was investigated on the condition that running time equaled to 10 s and other parameters in the basic model remained unchanged. As shown in Figure 10 and Figure 11a, the length and aperture of hydraulic fracture increases significantly with increasing viscosity. The main reason is that the higher fluid viscosity would diminish the lateral fluid leak-off, which noticeably elevates the fluid volume needed for generating straight fracture. With the increasing fluid viscosity, breakdown pressure increases almost linearly, as shown in Figure 11b. This is because higher viscosity diminishes the fluid leak-off, which is equivalent to elevating the fluid volume needed for fracturing within the same time. According to the analysis of Section 3.3, the above reason leads to higher breakdown pressure.
Some researchers, through laboratory hydraulic fracturing experiments [8,9,10,11,12,13,14] and site microseismic monitoring data [15], arrived at the similar conclusions that under the condition of low fluid viscosity (such as the case of supercritical CO2), the hydraulic fracture becomes flexed and with multiple branches, and develops a complex fracture network; with the increase of fluid viscosity (such as the case of a gel), the branches disappear and the hydraulic fracture becomes a simple wing shape.
According to [12,13], the viscosity of supercritical CO2 and liquid CO2 are 0.05 mPa·s and 0.1 mPa·s, respectively. Under the same condition, we simulated the effect of these two fluids on hydraulic fracturing, and no fracture appeared. This result indicates that when the injection rate is small enough, a low viscosity fluid could not break the rock. Based on the above numerical simulation and previous laboratory experiment results [9,14], as shown in Figure 12, we could know that when the fracturing fluid viscosity is too low, the injection rate could be elevated to produce hydraulic fracture. We verified the reliability of the numerical model, through comparison between numerical simulation, analytical solution, and experiment results shown in Section 3.1, Section 3.2, Section 3.3 and Section 3.4.

4. Relationship between in-situ Stress Ratio, Injection Rate and Fluid Viscosity

4.1. Relationship between Injection Rate and Fluid Viscosity

Rearranging Equation (1), we obtain:
q μ = a 3 12 Δ p L
It is obvious that when the flow rate q or fluid viscosity μ is increasing, the influence is equal on the right side of Equation (4), that is, either would lead to an increase of both the aperture a and the pressure Δ p . Because there is a positive correlation between injection rate Q (a parameter in software UDEC) and flow rate q, we could replace q with Q. The product Qμ is considered to be a whole in this section, to investigate its influence on hydraulic fracturing characteristic.
In Section 3.4, we did not see the hydraulic fracture using the fluid viscosity μ = 0.05 mPa·s and 0.1 mPa·s. Thus, in this section, we take Q = 100 × 10−6 and 50 × 10−6 m3/s/m to have Qμ = 5 × 10−6 mN; meanwhile, when fixing the fluid viscosity at 1 and 2 mPa·s, we adjust the injection rate to have Qμ = 10 × 10−6 mN, as shown in Table 3. For the convenience of comparison, we retain the stress ratio at K = 1.5 and the time at 10 s to verify the influence of the Qμ on the hydraulic fracturing.
Figure 13 and Figure 14 show that the hydraulic fracture becomes flexed and complex with multiple branches when Qμ is small, and the corresponding fracture aperture as well as the breakdown pressure are small. In contrast, the branches disappear and hydraulic fracture turns to be a simple wing shape. Therefore, a small value of Qμ would be helpful for generating complex fracturing network.

4.2. Relationship between Stress Ratio and Fluid Viscosity

As shown in Section 4.1, we can obtain hydraulic fractures if we increase the injection rate using supercritical CO2 with fluid viscosity of 0.05 mPa·s. Thus, we verify the sensitivity of the low viscosity fluid on the stress ratio by only tuning the stress ratio while retaining the other parameters.
With the stress ratio of K = 1.1, the hydraulic fracturing does not occur; thus, the results are not shown in Figure 15 and Figure 16.
As shown in Figure 15, for low viscosity fluid (e.g., supercritical CO2), the geometry of the hydraulic fractures are not sensitive to stress ratio, and still turns to be flexed and complex, which is quite different to those shown in Figure 6. This result indicates that in the formation with remarkable anisotropy stress, taking supercritical CO2 as fracturing fluid is in favor of hydraulic fracturing. The stress ratio still influences the breakdown pressure of the rock, as shown in Figure 16b).

5. Conclusions

Based on the shale mechanical parameters from laboratory tests, we set up the basic configuration of a numerical model, which was used to verify the reliability of the distinct element code UDEC on simulating the hydraulic fracturing. We mainly discussed the influence of critical parameters on hydraulic fracturing. The results are summarized as follows:
(1)
The increasing injection rate and fluid viscosity have the same influence on hydraulic fracturing characteristics. With the increasing injection rate and fluid viscosity, fracture aperture and breakdown pressure obviously increase.
(2)
The hydraulic fracture becomes flexed and complex with multiple branches when the value of Qμ is low, and the corresponding fracture aperture as well as the breakdown pressure are small. In contrast, the branches disappear and hydraulic fracture becomes a simple wing shape, and the corresponding fracture aperture as well as the breakdown pressure are high.
(3)
With low viscosity fluid, the hydraulic fracture geometry is not sensitive to stress ratio, and becomes a complex fracture network.

Acknowledgments

This research is supported by funding from the Strategic Priority Research Program of the Chinese Academy of Sciences (Grants XDB10030000, XDB10030300 and XDB10050400) and the National Natural Science Foundation of China (Grants 41502294).

Author Contributions

Bo Zhang and Xiao Li designed the theoretical framework; Bo Zhang and Zhaobin Zhang wrote the code of the program and designed the numerical simulation. Yu Wang, Yanfang Wu and Yusong Wu compared the results.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Legarth, B.; Huenges, E.; Zimmermann, G. Hydraulic fracturing in a sedimentary geothermal reservoir: Results and implications. Int. J. Rock Mech. Min. Sci. 2005, 42, 1028–1041. [Google Scholar] [CrossRef]
  2. Freeman, C.M.; Moridis, G.; Ilk, D.; Blasingame, T.A. A numerical study of performance for tight gas and shale gas reservoir systems. J. Petrol. Sci. Eng. 2013, 108, 22–39. [Google Scholar] [CrossRef]
  3. Zhou, F.D.; Chen, Z.X.; Rahman, S.S. Effect of hydraulic fracture extension into sandstone on coalbed methane production. J. Nat. Gas Sci. Eng. 2015, 22, 459–467. [Google Scholar] [CrossRef]
  4. Clark, C.; Burnham, A.; Harto, C.; Horner, R. Hydraulic Fracturing and Shale Gas Production: Technology, Impacts, and Regulation; Argonne National Laboratory: Chicago, IL, USA, 2013. [Google Scholar]
  5. Healy, D. Hydraulic Fracturing or “Fracking”: A Short Summary of Current Knowledge and Potential Environmental Impacts; University of Aberdeen: Aberdeen, UK, 2012. [Google Scholar]
  6. King, G.E. Thirty years of gas shale fracturing: What have we learned? In Proceedings of the SPE Annual Technical Conference and Exhibition, Florence, Italy, 19–22 September 2010.
  7. Guo, T.; Zhang, S.; Qu, Z.; Zhou, T.; Xiao, Y.; Gao, J. Experimental study of hydraulic fracturing for shale by stimulated reservoir volume. Fuel 2014, 128, 373–380. [Google Scholar] [CrossRef]
  8. De Pater, C.J.; Beugelsdijk, L.J.L. Experiments and numerical simulation of hydraulic fracturing in naturally fractured rock. In Proceedings of the 40th U.S. Symposium on Rock Mechanics, Anchorage, AK, USA, 25–29 June 2005.
  9. Fan, T.; Zhang, G. Influence of injection rate and fracturing fluid viscosity on hydraulic fracture geometry in coal. J. China Univ. Petrol. 2014, 38, 117–123. [Google Scholar]
  10. Ishida, T.; Aoyagi, K.; Niwa, T.; Chen, Y.; Murata, S.; Chen, Q.; Nakayama, Y. Acoustic emission monitoring of hydraulic fracturing laboratory experiment with supercritical and liquid CO2. J. Geophys. Res. 2012, 39. [Google Scholar] [CrossRef]
  11. Bennour, Z.; Ishida, T.; Nagaya, Y.; Nara, Y.; Chen, Y.; Chen, Q.; Nakayama, Y.; Sekine, K.; Nagano, Y. Fracture development and mechanism in shale cores by viscous oil, water and L-CO2 injection. In Proceedings of the 48th US Rock Mechanics/Geomechanics Symposium, Minneapolis, MN, USA, 1–4 June 2014.
  12. Chen, Y.; Nagaya, Y.; Ishida, T. Observations of fractures induced by hydraulic fracturing in anisotropic granite. Rock Mech. Rock Eng. 2015, 48, 1455–1461. [Google Scholar] [CrossRef]
  13. Bennour, Z.; Ishida, T.; Nagaya, Y.; Chen, Y.; Nara, Y.; Chen, Q.; Sekine, K.; Nagano, Y. Crack extension in hydraulic fracturing of shale cores using viscous oil, water, and liquid carbon dioxide. Rock Mech. Rock Eng. 2015, 48, 1463–1473. [Google Scholar] [CrossRef] [Green Version]
  14. Hou, B.; Chen, M.; Li, Z.; Wang, Y.; Diao, C. Propagation area evaluation of hydraulic fracture networks in shale gas reservoirs. Pet. Explor. Dev. 2014, 41, 833–838. [Google Scholar] [CrossRef]
  15. Warpinski, N.R.; Kramm, R.C.; Heinze, J.R.; Kennedy, C. Microseismic mapping techniques in the Barnett Shale. In Proceedings of the SPE Annual Technical Conference and Exhibition, Dallas, TX, USA, 9–12 October 2005.
  16. Middleton, R.; Carey, J.; Currier, R.; Hyman, J.; Kang, Q.; Karra, S.; Jiménez-Martínez, J.; Porter, M.; Viswanathan, H. Shale gas and non-aqueous fracturing fluids: Opportunities and challenges for supercritical CO2. Appl. Energy 2015, 147, 500–509. [Google Scholar] [CrossRef]
  17. Areias, P.; Rabczuk, T.; Dias-da-Costa, D. Element-wise fracture algorithm based on rotation of edges. Eng. Fract. Mech. 2013, 110, 113–137. [Google Scholar] [CrossRef]
  18. Areias, P.; Rabczuk, T. Finite strain fracture of plates and shells with configurational forces and edge rotation. Int. J. Numer. Meth. Eng. 2013, 94, 1099–1122. [Google Scholar] [CrossRef]
  19. Zhuang, X.; Augarde, C.; Mathisen, K. Fracture modelling using meshless methods and level sets in 3D: Framework and modelling. Int. J. Numer. Meth. Eng. 2012, 92, 969–998. [Google Scholar] [CrossRef]
  20. Zhuang, X.; Cai, Y.; Augarde, C. A meshless sub-region radial point interpolation method for accurate calculation of crack tip fields. Theor. Appl. Fract. Mech. 2014, 69, 118–125. [Google Scholar] [CrossRef] [Green Version]
  21. Rabczuk, T.; Belytschko, T. Cracking particles: A simplified meshfree method for arbitrary evolving cracks. Int. J. Numer. Meth. Eng. 2004, 61, 2316–2343. [Google Scholar] [CrossRef]
  22. Rabczuk, T.; Belytschko, T. A three dimensional large deformation meshfree method for arbitrary evolving cracks. Comput. Methods Appl. Mech. Eng. 2007, 196, 2777–2799. [Google Scholar] [CrossRef]
  23. Chau-Dinh, T.; Zi, G.; Lee, P.S.; Song, J.H.; Rabczuk, T. Phantom-node method for shell models with arbitrary cracks. Comput. Struct. 2012, 92–93, 242–256. [Google Scholar] [CrossRef]
  24. Vu-Bac, N.; Nguyen-Xuan, H.; Chen, L.; Lee, C.K.; Zi, G.; Zhuang, X.; Liu, G.R.; Rabczuk, T. A phantom-node method with edge-based strain smoothing for linear elastic fracture mechanics. J. Appl. Math. 2013, 2013. [Google Scholar] [CrossRef]
  25. Ghorashi, S.; Valizadeh, N.; Mohammadi, S.; Rabczuk, T. T-spline based XIGA for fracture analysis of orthotropic media. Comput. Struct. 2015, 147, 138–146. [Google Scholar] [CrossRef]
  26. Nguyen-Thanh, N.; Valizadeh, N.; Nguyen, M.N.; Nguyen-Xuan, H.; Zhuang, X.; Areias, P.; Zi, G.; Bazilevs, Y.; de Lorenzis, L.; Rabczuk, T. An extended isogeometric thin shell analysis based on Kirchho-Love theory. Comput. Methods Appl. Mech. Eng. 2015, 284, 265–291. [Google Scholar] [CrossRef]
  27. Rabczuk, T.; Gracie, R.; Song, J.H.; Belytschko, T. Immersed particle method for fluid-structure interaction. Int. J. Numer. Meth. Eng. 2010, 81, 48–71. [Google Scholar] [CrossRef]
  28. Zhuang, X.; Huang, R.; Rabczuk, T.; Liang, C. A coupled thermo-hydro-mechanical model of jointed hard rock for compressed air energy storage. Math. Probl. Eng. 2014, 2014. [Google Scholar] [CrossRef]
  29. Wangen, M. Finite element modeling of hydraulic fracturing on a reservoir scale in 2D. J. Pet. Sci. Eng. 2011, 77, 274–285. [Google Scholar] [CrossRef]
  30. Carrier, B.; Granet, S. Numerical modeling of hydraulic fracture problem in permeable medium using cohesive zone model. Eng. Fract. Mech. 2012, 79, 312–328. [Google Scholar] [CrossRef] [Green Version]
  31. Zhang, G.M.; Liu, H.; Zhang, J.; Wu, H.A.; Wang, X.X. Three-dimensional finite element simulation and parametric study for horizontal well hydraulic fracture. J. Pet. Sci. Eng. 2010, 72, 310–317. [Google Scholar] [CrossRef]
  32. Zhou, L.; Hou, M.Z. A numerical simulation on rules of crack extension during hydraulic fracturing in shale gas reservoir. Int. J. Rock Mech. Min. Sci. 2012, 60, 370–380. [Google Scholar]
  33. Itasca Consulting Group, Inc. UDEC (Universal Distinct Element Code), Version 4.0; Itasca Consulting Group, Inc.: Minneapolis, MN, USA, 2004. [Google Scholar]
  34. Cundall, P.A. A computer model for simulating progressive large scale movements in blocky rock systems. In Proceedings of the Symposium of the International Society of Rock Mechanics, Nancy, France, 1971.
  35. Nagel, N.; Gil, I.; Sanchez-nagel, M. Simulating hydraulic fracturing in real fractured rock-overcoming the limits of pseudo3D models. In Proceedings of the SPE Hydraulic Fracturing Technology Conference and Exhibition, The Woodlands, TX, USA, 24–26 January 2011.
  36. Al-Busaidi, A.; Hazzard, J.F.; Young, R.P. Distinct element modeling of hydraulically fractured Lac du Bonnet granite. J. Geophys. Res. 2005, 110. [Google Scholar] [CrossRef]
  37. Gil, I.; Damjanac, B.; Nagel, N.; Guo, Q.X. Geomechanical evaluation of solids injection. In Proceedings of the 44th US Rock Mechanics symposium, Salt Lake City, UT, USA, 27–30 June 2010.
  38. Damjanac, B.; Gil, I.; Pierce, M.; Sanchez-Nagel, M.A.; Van, A.A.; McLennan, J. A new approach to hydraulic fracturing modeling in naturally fractured reservoirs. In Proceedings of the 44th US Rock Mechanics symposium, Salt Lake City, UT, USA, 27–30 June 2010.
  39. Sarmadivaleh, M. Experimental and Numerical Study of Interaction of a Pre-Existing Natural Interface and an Induced Hydraulic Fracture. Ph.D. Thesis, Curtin University, Perth, Australia, 2012. [Google Scholar]
  40. Choi, S. Interpretation of shut-in pressure in hydrofracturing pressure-time records using numerical modeling. Int. J. Rock Mech. Min. Sci. 2012, 50, 29–39. [Google Scholar] [CrossRef]
  41. Zangeneh, N.; Eberhardt, E.; Bustin, R. Investigation of the influence of natural fractures and in situ stress on hydraulic fracture propagation using a distinct-element approach. Can. Geotech. J. 2015, 52, 926–946. [Google Scholar] [CrossRef]
  42. Li, Y.P.; Wang, Y.H.; Dingm, Y.H.; Wang, X. Physical and numerical simulation of hydraulic fracture initiation and propagation in gas shale. In Proceedings of the SPE Unconventional Resources Conference and Exhibition-Asia Pacific, Brisbane, Australia, 11–13 November 2013.
  43. Bahorich, B.; Olson, J.E.; Holder, J. Examining the effect of cemented natural fractures on hydraulic fracture propagation in hydrostone block experiments. In Proceedings of the SPE Annual Technical Conference and Exhibition, San Antonio, TX, USA, 8–10 October 2012.
  44. Cho, N.; Martin, C.D.; Sego, D.C. A clumped particle model for rock. Int. J. Rock Mech. Min. Sci. 2007, 44, 997–1010. [Google Scholar] [CrossRef]
  45. Fjær, E.; Holt, R.M.; Horsrud, P. Petroleum Related Rock Mechanics; Elsevier: Amsterdam, The Netherlands, 2008; p. 374. [Google Scholar]
Figure 1. The relationship between fracture aperture and normal stress.
Figure 1. The relationship between fracture aperture and normal stress.
Energies 09 00140 g001
Figure 2. Numerical model.
Figure 2. Numerical model.
Energies 09 00140 g002
Figure 3. The hydraulic fracturing characteristics at the time of 2 s. (a) fracture geometry; (b) injection pressure curve; (c) distribution of pore pressure.
Figure 3. The hydraulic fracturing characteristics at the time of 2 s. (a) fracture geometry; (b) injection pressure curve; (c) distribution of pore pressure.
Energies 09 00140 g003aEnergies 09 00140 g003b
Figure 4. The hydraulic fracturing characteristics at 6 s. (a) Fracture geometry; (b) Injection pressure curve; (c) Distribution of pore pressure.
Figure 4. The hydraulic fracturing characteristics at 6 s. (a) Fracture geometry; (b) Injection pressure curve; (c) Distribution of pore pressure.
Energies 09 00140 g004aEnergies 09 00140 g004b
Figure 5. Comparison of breakdown pressure under different stress ratios between the analytical and numerical solutions.
Figure 5. Comparison of breakdown pressure under different stress ratios between the analytical and numerical solutions.
Energies 09 00140 g005
Figure 6. The relationship between hydraulic fracture geometry and stress ratio. (a) K = 1.1; (b) K = 1.2; (c) K = 1.3; (d) K = 1.4; (e) K = 1.5; (f) K = 1.6; (g) K = 1.7; (h) K = 1.8.
Figure 6. The relationship between hydraulic fracture geometry and stress ratio. (a) K = 1.1; (b) K = 1.2; (c) K = 1.3; (d) K = 1.4; (e) K = 1.5; (f) K = 1.6; (g) K = 1.7; (h) K = 1.8.
Energies 09 00140 g006
Figure 7. The relationship between maximum aperture and stress ratio.
Figure 7. The relationship between maximum aperture and stress ratio.
Energies 09 00140 g007
Figure 8. The relationship between hydraulic fracture geometry and injection rate. (a) Q = 8E-06m3/s/m; (b) Q = 9E-06m3/s/m; (c) Q = 10E-06m3/s/m; (d) Q = 11E-06m3/s/m; (e) Q = 12E-06m3/s/m; (f) Q = 13E-06m3/s/m.
Figure 8. The relationship between hydraulic fracture geometry and injection rate. (a) Q = 8E-06m3/s/m; (b) Q = 9E-06m3/s/m; (c) Q = 10E-06m3/s/m; (d) Q = 11E-06m3/s/m; (e) Q = 12E-06m3/s/m; (f) Q = 13E-06m3/s/m.
Energies 09 00140 g008aEnergies 09 00140 g008b
Figure 9. The relationship between hydraulic fracturing characteristics and injection rate. (a) The relationship between maximum aperture and injection rate; (b) The relationship between breakdown pressure and injection rate.
Figure 9. The relationship between hydraulic fracturing characteristics and injection rate. (a) The relationship between maximum aperture and injection rate; (b) The relationship between breakdown pressure and injection rate.
Energies 09 00140 g009aEnergies 09 00140 g009b
Figure 10. The relationship between hydraulic fracture geometry and injection fluid viscosity. (a) μ = 0.5 mPa·s; (b) μ = 1 mPa·s; (c) μ = 2 mPa·s; (d) μ = 3 mPa·s; (e) μ = 4 mPa·s; (f) μ=5 mPa·s.
Figure 10. The relationship between hydraulic fracture geometry and injection fluid viscosity. (a) μ = 0.5 mPa·s; (b) μ = 1 mPa·s; (c) μ = 2 mPa·s; (d) μ = 3 mPa·s; (e) μ = 4 mPa·s; (f) μ=5 mPa·s.
Energies 09 00140 g010
Figure 11. The relationship between hydraulic fracturing characteristics and injection fluid viscosity. (a) The relationship between maximum aperture and fluid viscosity; (b) The relationship between breakdown pressure and fluid viscosity.
Figure 11. The relationship between hydraulic fracturing characteristics and injection fluid viscosity. (a) The relationship between maximum aperture and fluid viscosity; (b) The relationship between breakdown pressure and fluid viscosity.
Energies 09 00140 g011aEnergies 09 00140 g011b
Figure 12. The relationship between injection rate and fluid viscosity.
Figure 12. The relationship between injection rate and fluid viscosity.
Energies 09 00140 g012
Figure 13. The relationship between hydraulic fracture geometry and Qμ.
Figure 13. The relationship between hydraulic fracture geometry and Qμ.
Energies 09 00140 g013
Figure 14. The relationship between Qμ and hydraulic fracturing characteristics. (a) relationship between Qμ and the fracture aperture; (b) relationship between Qμ and the breakdown pressure.
Figure 14. The relationship between Qμ and hydraulic fracturing characteristics. (a) relationship between Qμ and the fracture aperture; (b) relationship between Qμ and the breakdown pressure.
Energies 09 00140 g014aEnergies 09 00140 g014b
Figure 15. The relationship between hydraulic fracture geometry and stress ratio. (a) K = 1.3; (b) K = 1.5; (c) K = 1.7.
Figure 15. The relationship between hydraulic fracture geometry and stress ratio. (a) K = 1.3; (b) K = 1.5; (c) K = 1.7.
Energies 09 00140 g015
Figure 16. The relationship between stress ratio and hydraulic fracturing characteristics. (a) relationship between stress ratio and the fracture aperture; (b) relationship between stress ratio and the breakdown pressure.
Figure 16. The relationship between stress ratio and hydraulic fracturing characteristics. (a) relationship between stress ratio and the fracture aperture; (b) relationship between stress ratio and the breakdown pressure.
Energies 09 00140 g016aEnergies 09 00140 g016b
Table 1. Block and contact parameters in numerical model.
Table 1. Block and contact parameters in numerical model.
Block and Contact Parameters
Cohesion/MPa30Normal stiffness/GPa2500
Friction angle/°30Shear stiffness/GPa1000
Tensile strength/MPa3Minimum allowable aperture/m2 × 10−6
Elastic modulus/GPa27Zero normal stress aperture/m1 × 10−5
Poisson rate0.25Maximum allowable aperture/m2 × 10−4
Density/(g/cm3)2.6
Table 2. Operational parameters.
Table 2. Operational parameters.
Injection Rate/10−6 m3/s/mFluid Viscosity/mPa·s
101
Table 3. Different combination of the injection rate and viscosity.
Table 3. Different combination of the injection rate and viscosity.
Qμ/10−6mNInjection Rate, Q/10−6 m3/s/mFluid Viscosity, μ/mPa·s
51000.05
5500.1
10101
1052

Share and Cite

MDPI and ACS Style

Zhang, B.; Li, X.; Zhang, Z.; Wu, Y.; Wu, Y.; Wang, Y. Numerical Investigation of Influence of In-Situ Stress Ratio, Injection Rate and Fluid Viscosity on Hydraulic Fracture Propagation Using a Distinct Element Approach. Energies 2016, 9, 140. https://doi.org/10.3390/en9030140

AMA Style

Zhang B, Li X, Zhang Z, Wu Y, Wu Y, Wang Y. Numerical Investigation of Influence of In-Situ Stress Ratio, Injection Rate and Fluid Viscosity on Hydraulic Fracture Propagation Using a Distinct Element Approach. Energies. 2016; 9(3):140. https://doi.org/10.3390/en9030140

Chicago/Turabian Style

Zhang, Bo, Xiao Li, Zhaobin Zhang, Yanfang Wu, Yusong Wu, and Yu Wang. 2016. "Numerical Investigation of Influence of In-Situ Stress Ratio, Injection Rate and Fluid Viscosity on Hydraulic Fracture Propagation Using a Distinct Element Approach" Energies 9, no. 3: 140. https://doi.org/10.3390/en9030140

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