Skip to main content
Log in

An approximate solution for a plane strain hydraulic fracture that accounts for fracture toughness, fluid viscosity, and leak-off

  • Original Paper
  • Published:
International Journal of Fracture Aims and scope Submit manuscript

Abstract

The goal of this paper is to develop an approximate solution for a propagating plane strain hydraulic fracture, whose behavior is determined by a combined interplay of fluid viscosity, fracture toughness, and fluid leak-off. The approximation is constructed by assuming that the fracture behavior is primarily determined by the three-process (viscosity, toughness, and leak-off) multiscale tip asymptotics and the global fluid volume balance. First, the limiting regimes of propagation of the solution are considered, that can be reduced to an explicit form. Thereafter, applicability regions of the limiting solutions are investigated and transitions from one limiting solution to another are analyzed. To quantify the error of the constructed approximate solution, its predictions are compared to a reference numerical solution. Results indicate that the approximation is able to predict hydraulic fracture parameters for all limiting and transition regimes with an error of under one percent. Consequently, this development can be used to obtain a rapid solution for a plane strain hydraulic fracture with leak-off, which can be used for quick estimations of fracture geometry or as a reference solution to evaluate accuracy of more advanced hydraulic fracture simulators.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Abou-Sayed A, Andrews D, Buhidma I (1989) Evaluation of oily waste injection below the permafrost in prudhoe bay field. In: Proceedings of the California regional meetings. CA, Society of Petroleum Engineers. Richardson, Bakersfield, pp 129–142

  • Adachi J, Siebrits E, Peirce A, Desroches J (2007) Computer simulation of hydraulic fractures. Int J Rock Mech Min Sci 44:739–757

    Article  Google Scholar 

  • Adachi J, Detournay E (2002) Self-similar solution of a plane-strain fracture driven by a power-law fluid. Int J Numer Anal Methods Geomech 26:579–604

    Article  Google Scholar 

  • Adachi JI, Detournay E (2008) Plane-strain propagation of a hydraulic fracture in a permeable rock. Eng Fract Mech 75:4666–4694

    Article  Google Scholar 

  • Bunger A, Detournay E, Garagash D (2005) Toughness-dominated hydraulic fracture with leak-off. Int J Fract 134:175–190

    Article  Google Scholar 

  • Bunger A, Detournay E (2007) Early time solution for a penny-shaped hydraulic fracture. ASCE J Eng Mech 133:175–190

    Article  Google Scholar 

  • Crouch S, Starfield A (1983) Boundary element methods in solid mechanics. George Allen and Unwin, London

    Google Scholar 

  • Desroches J, Detournay E, Lenoach B, Papanastasiou P, Pearson J, Thiercelin M, Cheng AD (1994) The crack tip region in hydraulic fracturing. Proc R Soc Lond A 447:39–48

    Article  Google Scholar 

  • Detournay E (2004) Propagation regimes of fluid-driven fractures in impermeable rocks. Int J Geomech 4:35–45

    Article  Google Scholar 

  • Detournay E (2016) Mechanics of hydraulic fractures. Annu Rev Fluid Mech 48(31):139

    Google Scholar 

  • Detournay E, Garagash D (2003) The tip region of a fluid-driven fracture in a permeable elastic solid. J Fluid Mech 494:1–32

    Article  Google Scholar 

  • Dontsov E (2016a) An approximate solution for a penny-shaped hydraulic fracture that accounts for fracture toughness, fluid viscosity, and leak-off. R Soc Open Sci 3(160):737

    Google Scholar 

  • Dontsov E (2016b) Propagation regimes of buoyancy-driven hydraulic fractures with solidification. J Fluid Mech 797:1–28

    Article  Google Scholar 

  • Dontsov E (2016c) Tip region of a hydraulic fracture driven by a laminar-to-turbulent fluid flow. J Fluid Mech 797:R2

    Article  Google Scholar 

  • Dontsov E, Peirce A (2016) Implementing a universal tip asymptotic solution into an implicit level set algorithm (ILSA) for multiple parallel hydraulic fractures. In: Proceedings of the 50th US rock mechanics symposium, Houston, TX, ARMA-2016-268. American Rock Mechanics Association, Houston

  • Dontsov E, Peirce A (2015) A non-singular integral equation formulation to analyze multiscale behaviour in semi-infinite hydraulic fractures. J Fluid Mech 781:R1

    Article  Google Scholar 

  • Dontsov E, Peirce A (2017) A multiscale implicit level set algorithm (ILSA) to model hydraulic fracture propagation incorporating combined viscous, toughness, and leak-off asymptotics. Comput Methods Appl Mech Eng 313:53–84

    Article  Google Scholar 

  • Economides M, Nolte K (eds) (2000) Reservoir stimulation, 3rd edn. Wiley, Chichester

    Google Scholar 

  • Frank U, Barkley N (2005) Remediation of low permeability subsurface formations by fracturing enhancements of soil vapor extraction. J Hazard Mater 40:191–201

    Article  Google Scholar 

  • Garagash D (2006) Plane-strain propagation of a fluid-driven fracture during injection and shut-in: asymptotics of large toughness. Eng Fract Mech 73:456–481

    Article  Google Scholar 

  • Garagash D, Detournay E, Adachi J (2011) Multiscale tip asymptotics in hydraulic fracture with leak-off. J Fluid Mech 669:260–297

    Article  Google Scholar 

  • Garagash D, Detournay E (2000) The tip region of a fluid-driven fracture in an elastic medium. J Appl Mech 67:183–192

    Article  Google Scholar 

  • Garagash D, Detournay E (2005) Plane-strain propagation of a fluid-driven fracture: small toughness solution. ASME J Appl Mech 72:916–928

    Article  Google Scholar 

  • Hills D, Kelly P, Dai D, Korsunsky A (1996) Solution of crack problems, the distributed dislocation technique, solid mechanics and its applications, vol 44. Kluwer Academic Publisher, Dordrecht

    Google Scholar 

  • Hu J, Garagash D (2010) Plane-strain propagation of a fluid-driven crack in a permeable rock with fracture toughness. J Eng Mech 136:1152–1166

    Article  Google Scholar 

  • Jeffrey R, Mills K (2000) Hydraulic fracturing applied to inducing longwall coal mine goaf falls. Pacific Rocks 2000. Balkema, Rotterdam, pp 423–430

    Google Scholar 

  • Lenoach B (1995) The crack tip solution for hydraulic fracturing in a permeable solid. J Mech Phys Solids 43:1025–1043

    Article  Google Scholar 

  • Lister JR (1990) Buoyancy-driven fluid fracture: the effects of material toughness and of low-viscosity precursors. J Fluid Mech 210:263–280

    Article  Google Scholar 

  • Lister J, Kerr R (1991) Fluid-mechanical models of crack propagation and their application to magma transport in dykes. J Geophys Res 96:10,049–10,077

    Article  Google Scholar 

  • Madyarova M (2003) Fluid-driven penny-shaped fracture in elastic medium. Master’s thesis, University of Minnesota

  • Peirce A (2016) Implicit level set algorithms for modelling hydraulic fracture propagation. Phil Trans R Soc A 374(20150):423. doi:10.1098/rsta.2015.0423

    Google Scholar 

  • Peirce A, Detournay E (2008) An implicit level set method for modeling hydraulically driven fractures. Comput Methods Appl Mech Eng 197:2858–2885

    Article  Google Scholar 

  • Rice J (1968) Mathematical analysis in the mechanics of fracture. In: Liebowitz H (ed) Fracture: an advanced treatise, Chap 3, vol II. Academic Press, New York, pp 191–311

    Google Scholar 

  • Roper S, Lister JR (2005) Buoyancy-driven crack propagation from an over-pressured source. J Fluid Mech 536:79–98

    Article  Google Scholar 

  • Roper S, Lister JR (2007) Buoyancy-driven crack propagation: the limit of large fracture toughness. J Fluid Mech 580:359–380

    Article  Google Scholar 

  • Rubin A (1995) Propagation of magma-filled cracks. Annu Rev Earth Planet 23:287–336

    Article  Google Scholar 

  • Savitski A, Detournay E (2002) Propagation of a fluid-driven penny-shaped fracture in an impermeable rock: asymptotic solutions. Int J Solids Struct 39:6311–6337

    Article  Google Scholar 

  • Spence D, Turcotte D (1985) Magma-driven propagation of cracks. J Geophys Res 90:575–580

    Article  Google Scholar 

  • Tsai V, Rice J (2010) A model for turbulent hydraulic fracture and application to crack propagation at glacier beds. J Geophys Res 115(F03):007

    Google Scholar 

  • Weng X (2015) Modeling of complex hydraulic fractures in naturally fractured formation. J Unconv Oil Gas Res 9:114–135

    Article  Google Scholar 

Download references

Acknowledgements

Start-up funds provided by the University of Houston are greatly acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. V. Dontsov.

Appendices

Appendix 1: Functions \(g_\delta ({\hat{K}},{\hat{C}})\) and \(\varDelta ({\hat{K}},{\hat{C}})\)

This appendix provides expressions for the functions \(g_\delta \left( \hat{K},{\hat{C}}\right) \) and \(\varDelta \left( {\hat{K}},{\hat{C}}\right) \) that are used in the paper to approximate the solution for a semi-infinite hydraulic fracture. Note that these functions were obtained in (Dontsov and Peirce 2015, 2017).

With the reference to the scaling (12) and the fact that \(\dot{l}=\alpha l/t\), the function f can be introduced as

$$\begin{aligned} {\hat{s}}= & {} \dfrac{1}{3C_1(\delta ) }\left[ 1-{\hat{K}}^3-\frac{3}{2}{\hat{C}}{\hat{b}} \left( 1-{\hat{K}}^2\right) +3 {\hat{C}}^2{\hat{b}}^2\left( 1-{\hat{K}}\right) \right. \nonumber \\&\left. -\,3{\hat{C}}^3{\hat{b}}^3\ln \left( \dfrac{{\hat{C}}{\hat{b}}+1}{{\hat{C}}\hat{b}+{\hat{K}}}\right) \right] \equiv f\left( {\hat{K}},{\hat{C}}{\hat{b}},C_1\right) , \end{aligned}$$
(62)

where \({\hat{b}}={C_2(\delta )}/{C_1(\delta )}\) and

$$\begin{aligned} C_1(\delta )= & {} \dfrac{4(1-2\delta )}{\delta (1-\delta )} \tan \left( \pi \delta \right) ,\nonumber \\ C_2(\delta )= & {} \dfrac{16(1-3\delta )}{3\delta (2-3\delta )} \tan \left( \dfrac{3\pi }{2}\delta \right) . \end{aligned}$$
(63)

The zeroth-order approximation for the solution can be written as

$$\begin{aligned} {\hat{s}}=f\left( {\hat{K}},\dfrac{3\beta _{{\tilde{m}}}^4}{4\beta _m^3}\hat{C},\frac{\beta _m^3}{3}\right) \equiv g_0\left( {\hat{K}}, {\hat{C}}\right) , \end{aligned}$$
(64)

where \(\beta _{{\tilde{m}}}={4}/{\left( 15^{1/4}\left( \sqrt{2}-1\right) ^{1/4}\right) }\) and \(\beta _{m}=2^{1/3}3^{5/6}\). As mentioned in (Dontsov and Peirce 2015), the solution varies spatially as \(w_a(s)\propto s^{{\bar{\delta }}}\), where \({\bar{\delta }} =\tfrac{1}{2}(1+\delta )\) and the power \(\delta \) is given by

$$\begin{aligned} \delta =\dfrac{\beta _m^3 }{3} \left( 1+\dfrac{3\beta _{\tilde{m}}^4}{4\beta _m^3}{\hat{C}}\right) g_0\left( {\hat{K}}, {\hat{C}}\right) \equiv \varDelta \left( {\hat{K}},{\hat{C}}\right) ,\nonumber \\ \end{aligned}$$
(65)

which defines the function \(\varDelta ({\hat{K}},{\hat{C}})\) and leads to the relation \({\bar{\delta }} =\tfrac{1}{2}\left( 1+\varDelta \left( {\hat{K}},{\hat{C}}\right) \right) \). By substituting (65) into (62), the \(\delta \)-corrected solution (12) can be written as

$$\begin{aligned} {\hat{s}}=f\left( {\hat{K}},{\hat{C}}{\hat{b}}\left( \varDelta ({\hat{K}},\hat{C})\right) ,C_1\left( \varDelta ({\hat{K}},{\hat{C}})\right) \right) \equiv g_\delta ({\hat{K}}, {\hat{C}}),\nonumber \\ \end{aligned}$$
(66)

which defines the function \(g_\delta ({\hat{K}}, {\hat{C}})\).

Appendix 2: Numerical scheme

To construct the numerical scheme, Eq. (2) is rewritten using \(\xi =x/l(t)\) and the scaling (47)–(49) as

$$\begin{aligned}&\dfrac{\partial \varOmega }{\partial \tau }- \dfrac{\xi V}{\gamma }\dfrac{\partial \varOmega }{\partial \xi } -\dfrac{1}{\gamma ^2}\dfrac{\partial }{\partial \xi }\left( \varOmega ^3\dfrac{\partial \varPi }{\partial \xi } \right) \nonumber \\&\quad +\dfrac{1}{\sqrt{\tau -\tau _0(\gamma \xi )}}=\dfrac{1}{\gamma }\delta (\xi ),\nonumber \\&\qquad \varOmega \rightarrow K_m\gamma ^{-1/2} (1-\xi )^{1/2},~~~ \xi \rightarrow 1, \end{aligned}$$
(67)

where \(V={\dot{\gamma }}\), while the elasticity Eq. (3) is reduced to

$$\begin{aligned} \varPi (\xi ,\tau )= & {} -\dfrac{1}{2\pi \gamma }\int _0^1 M(\xi ,s) \dfrac{\partial \varOmega (s,\tau )}{\partial s}\,{\hbox {d}}s,\nonumber \\ M(\xi ,s)= & {} \dfrac{\xi }{\xi ^2-s^2}. \end{aligned}$$
(68)

The spatial coordinate \(\xi \) is discretized as \(\xi _j=(\tfrac{1}{2}+j) \varDelta \xi \), \(j=1\ldots N\), in which case \(\xi _1=\tfrac{1}{2}\varDelta \xi \) and \(\xi _{N}=1-\tfrac{1}{2}\varDelta \xi \), and the temporal coordinate \(\tau \) is discretized uniformly on a logarithmic scale. Piecewise constant approximation for \(\varOmega \) is used, in which case \(\varOmega ^i_j=\varOmega (\xi _j,\tau _i)\) and the vector \(\varvec{\varOmega }^i\) represents an array of values of \(\varOmega ^i_j\) for all j. In this situation, the elasticity Eq. (68) is discretized as

$$\begin{aligned} \varvec{\varPi }^i= & {} \varvec{C} \varvec{\varOmega }^i,\nonumber \\ C_{mn}= & {} \dfrac{1}{2\pi \gamma ^i} \left[ M\left( \xi _m,\xi _n+\tfrac{1}{2}\varDelta \xi \right) -M\left( \xi _m,\xi _n-\tfrac{1}{2}\varDelta \xi \right) \right] .\nonumber \\ \end{aligned}$$
(69)

Fluid balance (67), on the other hand, is discretized using backward time differencing

$$\begin{aligned} \dfrac{ \varvec{\varOmega }^i - \varvec{\varOmega }^{i-1}}{\varDelta \tau }=\varvec{B} \varvec{\varOmega }^i+\varvec{A}\left( \varvec{\varOmega }^i\right) \varvec{\varPi }^i+\varvec{S}^i, \end{aligned}$$
(70)

where the term that captures the moving mesh is discretized as

$$\begin{aligned} \left[ \varvec{B} \varvec{\varOmega }^i\right] _j= & {} \dfrac{\xi _j V}{\gamma }\dfrac{{\varOmega }_{j+1}^i-{\varOmega }^i_{j-1}}{2\varDelta \xi },\qquad j=2\ldots N-1,\\ \left[ \varvec{B} \varvec{\varOmega }^i\right] _1= & {} \dfrac{\xi _1 V}{\gamma }\dfrac{{\varOmega }_{2}^i-{\varOmega }^i_{1}}{2\varDelta \xi }, \\ \left[ \varvec{B} \varvec{\varOmega }^i\right] _N= & {} -\dfrac{\xi _N V}{\gamma }\dfrac{\varOmega ^i_N+{\varOmega }^i_{N-1}}{2\varDelta \xi }, \end{aligned}$$

the lubrication term is discretized as

$$\begin{aligned} \left[ \varvec{A}\left( \varvec{\varOmega }^i\right) \varvec{\varPi }^i\right] _j= & {} \dfrac{1}{\gamma ^2 \varDelta \xi }\left[ \left( \varOmega _{j+1/2}^i\right) ^3\dfrac{\varPi ^i_{j+1}-\varPi ^i_{j}}{\varDelta \xi }\right. \\&\left. - \left( \varOmega _{j-1/2}^i\right) ^3\dfrac{\varPi ^i_{j}-\varPi ^i_{j-1}}{\varDelta \xi } \right] ,~~ j=2\ldots N-1,\\ \left[ \varvec{A}\left( \varvec{\varOmega }^i\right) \varvec{\varPi }^i\right] _1= & {} \dfrac{\left( \varOmega _{3/2}^i\right) ^3}{\gamma ^2 \varDelta \xi }\dfrac{\varPi ^i_{2}-\varPi ^i_{1}}{\varDelta \xi } ,\\ \left[ \varvec{A}\left( \varvec{\varOmega }^i\right) \varvec{\varPi }^i\right] _N= & {} -\dfrac{ \left( \varOmega _{N-1/2}^i\right) ^3}{\gamma ^2 \varDelta \xi }\dfrac{\varPi ^i_{N}-\varPi ^i_{N-1}}{\varDelta \xi } , \end{aligned}$$

where the widths at the mid points \(j\pm 1/2\) are calculated as an average between the corresponding values of the widths, and the source/leak-off term is

$$\begin{aligned} {S}_j^i= & {} -\dfrac{2}{\varDelta \tau } \left[ \sqrt{\tau _i+\varDelta \tau -\tau _0(\gamma \xi _j)} - \sqrt{\tau _i-\tau _0\left( \gamma \xi _j\right) } \right] \nonumber \\&+\dfrac{\delta _{1j}}{2\gamma \varDelta \xi }, \end{aligned}$$

where \(j=1\ldots N-1\) and \(\delta _{1j}\) is the Kronecker delta. The values of \(\left[ \varvec{B} \varvec{\varOmega }^i\right] _N\) and \(\left[ \varvec{A}( \varvec{\varOmega }^i) \varvec{\varPi }^i\right] _N\) are obtained by integration of (67) over the last element and using the no flux condition at the tip.

Similar to (Dontsov and Peirce 2017; Peirce 2016), in order to capture the multiscale behavior near the fracture tip, the following propagation condition is used

$$\begin{aligned} \varOmega ^i_{N-1}=\varOmega _a\left( \tfrac{3}{2}\gamma \varDelta \xi \right) , \end{aligned}$$

where \(\varOmega _a\) is the scaled tip asymptotic solution. The latter equation implies that the numerical solution follows the asymptotic solution from the penultimate element to the tip, and allows one to determine the propagation velocity V. To successfully use this condition, pressure at the tip element \(\varPi ^i_N\) is treated as an unknown. Since the tip asymptotic solution satisfies (12), it follows that

$$\begin{aligned} {\hat{s}}= & {} g_\delta \left( {\hat{K}},{\hat{C}}\right) ,\qquad {\hat{K}}=\dfrac{K_m d^{1/2}}{\varOmega ^i_{N-1}},\nonumber \\ {\hat{C}}= & {} \dfrac{2 d^{3/2} }{\left( \varOmega ^i_{N-1}\right) ^{5/2} \hat{s}^{1/2}},\qquad V=\dfrac{{\hat{s}} \,\left( \varOmega ^i_{N-1}\right) ^3}{d^2}, \end{aligned}$$
(71)

where \(d=\tfrac{3}{2}\gamma \varDelta \xi \) signifies the distance from the center of the penultimate element to the tip. The above equation is solved for \({\hat{s}}\) using Newton’s method, and the velocity of propagation is calculated. Since the solution in the tip element follows the asymptotic solution, it is possible to determine its average opening as

$$\begin{aligned} \varOmega ^i_{N}=\left( \dfrac{2}{3}\right) ^{(3+\delta )/2}\, \dfrac{2\, \varOmega ^i_{N-1} }{3+\delta } \dfrac{d}{\varDelta \xi }, \end{aligned}$$
(72)

where it is used that \(\varOmega _s\propto d^{(1+\delta )/2}\) and \(\delta =\varDelta ({\hat{K}},{\hat{C}})\). The factor 2 / 3 ensures that the calculation is performed only within the tip element (and does not continue to the middle of the penultimate element). Leak-off in the tip element is calculated by taking \(\tau -\tau _0=\gamma (1-\xi )/V\), in which case

$$\begin{aligned} S^i_{N}=-2 \sqrt{\dfrac{V}{\gamma \varDelta \xi }}. \end{aligned}$$
(73)

The numerical scheme consists of solving (69)–(73) for \(\varOmega ^i_j\) (for \(j=1\ldots N-1\)) and \(\varPi ^i_N\), which is done iteratively for each time step. The fracture length is then updated as \(\gamma ^i=\gamma ^{i-1}+V\varDelta \tau \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dontsov, E.V. An approximate solution for a plane strain hydraulic fracture that accounts for fracture toughness, fluid viscosity, and leak-off. Int J Fract 205, 221–237 (2017). https://doi.org/10.1007/s10704-017-0192-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10704-017-0192-4

Keywords

Navigation