Skip to main content
Log in

3D model of transversal fracture propagation from a cavity caused by Herschel–Bulkley fluid injection

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

Abstract

The paper presents an extension of authors’ previous model for a 3D hydraulic fracture with Newtonian fluid, which aims to account for the Herschel–Bulkley fluid rheology and to study the associated effects. This fluid rheology model is the most suitable for description of modern complex fracturing fluids, in particular, for description of foamed fluids that have been successfully utilized recently as fracturing fluids in tight and ultra-tight unconventional formations with high clay contents. Another advantage of using Herschel–Bulkley rheological law in the hydraulic fracture model consists in its generality as its particular cases allow describing the behavior of the majority of non-Newtonian fluids employed in hydraulic fracturing. Except the Herschel–Bulkley fluid flow model the considered model of hydraulic fracturing includes the model of the rock stress state. It is based on the elastic equilibrium equations that are solved by the dual boundary element method. Also the hydraulic fracturing model contains the new mixed mode propagation criterion, which states that the fracture should propagate in the direction in which mode \({{\mathrm{\mathrm {II}}}}\) and mode \({{\mathrm{\mathrm {III}}}}\) stress intensity factors both vanish. Since it is not possible to make both modes zero simultaneously the criterion proposes a functional that depends on both modes and is minimized along the fracture front in order to obtain the direction of propagation. Solution for Herschel–Bulkley fluid flow in a channel is presented in detail, and the numerical algorithm is described. The developed model has been verified against some reference solutions and sensitivity of fracture geometry to rheological fluid parameters has been studied to some extent.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18
Fig. 19
Fig. 20

Similar content being viewed by others

References

  • Abass HH, Brumley JL, Venditto JJ et al (1994) Oriented perforations-a rock mechanics view. In: SPE annual technical conference and exhibition

  • Adachi JI, 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(16):4666–4694

    Article  Google Scholar 

  • Aliabadi MH (2002) The boundary element method, applications in solids and structures, vol 2. Wiley, Chichester

    Google Scholar 

  • Aud WW, Wright TB, Cipolla CL, Harkrider JD (1994) The effect of viscosity on near-wellbore tortuosity and premature screenouts. In: SPE annual technical conference and exhibition, New Orleans, Louisiana. Society of Petroleum Engineers

  • Barati R, Liang J-T (2014) A review of fracturing fluid systems used for hydraulic fracturing of oil and gas wells. J Appl Polym Sci 131(16):1–11

    Article  Google Scholar 

  • Barati R, Hutchins RD, Friedel T, Ayoub JA, Dessinges M-N, England KW (2009) Fracture impact of yield stress and fracture-face damage on production with a three-phase 2D model. SPE Product Oper 24(02):336–345 (SPE-111457-PA)

    Article  Google Scholar 

  • Bunger AP, Detournay E (2007) Early-time solution for a radial hydraulic fracture. J Eng Mech 133(5):534–540

    Article  Google Scholar 

  • Bunger AP, Detournay E, Garagash DI (2005) Toughness-dominated hydraulic fracture with leak-off. Int J Fract 134(2):175–190

    Article  Google Scholar 

  • Chen J-T, Hong H-K (1999) Review of dual boundary element methods with emphasis on hypersingular integrals and divergent series. Appl Mech Rev 52(1):17–33

    Article  Google Scholar 

  • Cherny SG, Lapin VN (2016) 3D model of hydraulic fracture with Herschel–Bulkley compressible fluid pumping. Procedia Struct Integr 2:2479–2486

    Article  Google Scholar 

  • Cherny S, Chirkov D, Lapin V, Muranov A, Bannikov D, Miller M, Willberg D, Medvedev O, Alekseenko O (2009) Two-dimensional modeling of the near-wellbore fracture tortuosity effect. Int J Rock Mech Min Sci 46(6):992–1000

    Article  Google Scholar 

  • Cherny S, Lapin V, Esipov D, Kuranakov D, Avdyushenko A, Lyutov A, Karnakov P (2016) Simulating fully 3D non-planar evolution of hydraulic fractures. Int J Fract 201(2):181–211. https://doi.org/10.1007/s10704-016-0122-x

    Article  Google Scholar 

  • Cleary MP, Johnson DE, Kogsbøll H-H, Owens KA, Perry KF, de Pater CJ, Stachel A, Schmidt H, Mauro T (1993) Field implementation of proppant slugs to avoid premature screen-out of hydraulic fractures with adequate proppant concentration. In: Low permeability reservoirs symposium, Denver, Colorado. Society of Petroleum Engineers

  • Cooke ML, Pollard DD (1996) Fracture propagation paths under mixed mode loading within rectangular blocks of polymethyl methacrylate. J Geophys Res 101(B2):3387–3400

    Article  Google Scholar 

  • Detournay E (2004) Propagation regimes of fluid-driven fractures in impermeable rocks. Int J Geomech 4(1):35–45

    Article  Google Scholar 

  • Detournay E (2016) Mechanics of hydraulic fractures. Annu Rev Fluid Mech 48(1):311–339

    Article  Google Scholar 

  • Dontsov EV (2016) An approximate solution for a penny-shaped hydraulic fracture that accounts for fracture toughness, fluid viscosity and leak-off. R Soc Open Sci 3(12):160737

    Article  Google Scholar 

  • Dontsov EV, Kresse O (2018) A semi-infinite hydraulic fracture with leak-off driven by a power-law fluid. J Fluid Mech 837:210–229

    Article  Google Scholar 

  • Economides MJ, Nolte KG (2000) Reservoir stimulation, 3rd edn. Wiley, Chichester

    Google Scholar 

  • Esipov DV, Kuranakov DS, Lapin VN, Cherny SG (2014) Mathematical models of hydraulic fracturing of a reservoir. Comput Technol 19(2):33–61 (in Russian)

    Google Scholar 

  • Garagash DI (2006) Transient solution for a plane-strain fracture driven by a shear-thinning, power-law fluid. Int J Numer Anal Methods Geomech 30(14):1439–1475

    Article  Google Scholar 

  • Garagash D, Detournay E (2000) The tip region of a fluid-driven fracture in an elastic medium. J Appl Mech 67(1):183–192

    Article  Google Scholar 

  • Herschel WH, Bulkley R (1926) Konsistenzmessungen von gummi-benzollösungen. Kolloid-Zeitschrift 39(4):291–300

    Article  Google Scholar 

  • Hong H-K, Chen J-T (1988) Derivations of integral equations of elasticity. J Eng Mech 114(6):1028–1044

    Article  Google Scholar 

  • Kauzlarich JJ, Greenwood JA (1972) Elastohydrodynamic lubrication with Herschel–Bulkley model greases. ASLE Trans 15(4):269–277

    Article  Google Scholar 

  • Kuranakov DS, Esipov DV, Lapin VN, Cherny SG (2016) Modification of the boundary element method for computation of three-dimensional fields of strain-stress state of cavities with cracks. Eng Fract Mech 153:302–318

    Article  Google Scholar 

  • Linkov A (2015) Bench-mark solution for a penny-shaped hydraulic fracture driven by a thinning fluid. ArXiv e-prints arXiv:1508.07968

  • Mi Y, Aliabadi MH (1992) Dual boundary element method for three-dimensional fracture mechanics analysis. Eng Anal Bound Elem 10(2):161–171

    Article  Google Scholar 

  • Mi Y, Aliabadi MH (1994) Three-dimensional crack growth simulation using BEM. Comput Struct 52(5):871–878

    Article  Google Scholar 

  • Mitsoulis E (2007) Flows of viscoplastic materials: models and computations. In: Rheology Reviews 2007. British Society of Rheology

  • Nuismer RJ (1975) An energy release rate criterion for mixed mode fracture. Int J Fract 11(2):245–250

    Article  Google Scholar 

  • Ouyang S, Carey GF, Yew CH (1997) An adaptive finite element scheme for hydraulic fracturing with proppant transport. Int J Numer Methods Fluids 24:645–670

    Article  Google Scholar 

  • Pereira JPA (2010) Generalized finite element methods for three-dimensional crack growth simulations. Ph.D. thesis, Department of Civil and Environmental Engineering, University of Illinois, Urbana-Champaign, p 221

  • Rungamornrat J, Wheeler MF, Mear MF (2005) Coupling of fracture/non-newtonian flow for simulating nonplanar evolution of hydraulic fractures. In: SPE annual technical conference and exhibition, SPE-96968-MS

  • Savitski AA, Detournay E (2002) Propagation of a penny-shaped fluid-driven fracture in an impermeable rock: asymptotic solutions. Int J Solids Struct 39(26):6311–6337

    Article  Google Scholar 

  • Shokin Yu, Cherny S, Esipov D, Lapin V, Lyutov A, Kuranakov D (2015) Three-dimensional model of fracture propagation from the cavity caused by quasi-static load or viscous fluid pumping. Commun Comput Inf Sci. https://doi.org/10.1007/978-3-319-25058-8-15

    Google Scholar 

  • Sousa JL, Carter BJ, Ingraffea AR (1993) Numerical simulation of 3D hydraulic fracture using newtonian and power-law fluids. Int J Rock Mech Min Sci Geomech Abstr 30(7):1265–1271

    Article  Google Scholar 

Download references

Acknowledgements

The authors acknowledge the financial support of this research by the Russian Science Foundation (Grant No. 17-71-20139).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sergey Cherny.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 A.1 Herschel–Bulkley fluid model

Let us consider the most general form of the equations of incompressible HB-fluid flow in the 3D case. It includes the continuity equation

$$\begin{aligned} \nabla \cdot {{\mathrm{\mathbf v}}}= 0 \end{aligned}$$
(43)

and the momentum equation

$$\begin{aligned} \rho \frac{D {{\mathrm{\mathbf v}}}}{D t} = \nabla \cdot {\mathbb {P}}. \end{aligned}$$
(44)

In (43), (44) \({{\mathrm{\mathbf v}}}= \{ v_i \}\) is the fluid velocity vector and \({\mathbb {P}} = \{ p_{ij} \}\) is the total stress tensor divided into two parts

$$\begin{aligned} {\mathbb {P}} = -p {\mathbb {E}} + {\mathbb {T}}, \end{aligned}$$
(45)

where p is the scalar called the hydrodynamic pressure, \({\mathbb {T}} = \{ \tau _{ij} \}\) is the viscous stress tensor, and \({\mathbb {E}} = diag(1,1,1)\) is the unit (or identity) tensor. The viscous stress tensor \({\mathbb {T}}\) is connected with the strain rate tensor \({\mathbb {D}}=\{ D_{ij} \}\) by the constitutive relations

$$\begin{aligned} {\mathbb {T}}= & {} \eta {\mathbb {D}}, \quad \text {for} \ T \geqslant \tau _0, \end{aligned}$$
(46)
$$\begin{aligned} {\mathbb {D}}= & {} 0, \quad \text {for} \ T < \tau _0. \end{aligned}$$
(47)

In (46), (47) the components of the tensor \({\mathbb {D}}\) are

$$\begin{aligned} D_{ij} = \frac{\partial v_i}{\partial x_j} + \frac{\partial v_j}{\partial x_i} \end{aligned}$$
(48)

and \(\eta \) is the viscosity function given by the Herschel–Bulkley rheology model

$$\begin{aligned} \eta (D) = K D^{n-1} + \frac{\tau _0}{D}. \end{aligned}$$
(49)

Here, K is the flow consistency factor; n is the flow behavior index that governs the degree of shear thinning or thickening; \(\tau _0\) is the yield stress. T and D denote the second invariants of the respective tensors given by the following formulas

$$\begin{aligned} T = \sqrt{\frac{1}{2} \tau _{ij} \tau _{ij}} \quad \text {and} \quad D = \sqrt{\frac{1}{2} D_{ij} D_{ij}}. \end{aligned}$$
(50)

The index form of Eqs. (43) and (44) is the following:

$$\begin{aligned}&\frac{\partial v_i}{\partial x_i} = 0, \end{aligned}$$
(51)
$$\begin{aligned}&\rho \left[ \frac{\partial v_i}{\partial t} + \frac{\partial (v_i v_j)}{\partial x_j} \right] = - \frac{\partial p}{\partial x_i} \nonumber \\&\quad + \frac{\partial }{\partial x_j} \left[ \eta (D) \left( \frac{\partial v_i}{\partial x_j} + \frac{\partial v_j}{\partial x_i} \right) \right] . \end{aligned}$$
(52)

1.2 A.2 Equations for 2D HB fluid flow in 3D fracture

The fracture width W is much smaller than its longitudinal size. Therefore, the flow inside the fracture can be considered locally as the flow in a thin channel between two parallel plates. Without loss of generality let’s assume that the axis \(x_1, x_2\) of the local coordinate system lie in the tangent plane to the fracture surface and the axis \(x_3\) is orthogonal to the fracture surface. Then the transversal fluid velocity component can be assumed small (\(v_3 \approx 0\)) as compared to its longitudinal components. Also the derivatives of the fluid velocity components in the longitudinal directions \({\partial }/ {\partial x_1}\), \({\partial }/{\partial x_2}\) are small in comparison with their derivatives in the transverse direction \({\partial }/{\partial x_3}\). The fluid pressure and its consistency factor are considered constant in the transverse direction. The time derivatives in Eq. (52) are disregarded. The non-stationarity of the fracture propagation model is conditioned by the continuity Eq. (51) through the relationship between the fracture width W and the transversal fluid velocity \(v_3\)

$$\begin{aligned} \frac{\partial W}{\partial t} = v_3. \end{aligned}$$
(53)

Under the assumptions made above, Eqs. (51) and (52) can be simplified. Omitting small terms in Eq. (52), one can obtain for \(i=1,2\)

$$\begin{aligned} \frac{\partial p}{\partial x_1} = \frac{\partial }{\partial x_3}\left[ \eta (D) \frac{\partial v_1}{\partial x_3} \right] ,\frac{\partial p}{\partial x_2} = \frac{\partial }{\partial x_3}\left[ \eta (D) \frac{\partial v_2}{\partial x_3} \right] . \nonumber \\ \end{aligned}$$
(54)

Equation (52) for \(i=3\) degenerates under the stated assumptions. The integration of Eq. (54) over \(x_3\) gives

$$\begin{aligned} \eta (D) \frac{\partial v_i}{\partial x_3} = \frac{\partial p}{\partial x_i} + A_i, \ \ \end{aligned}$$
(55)

where \(A_i\) are some constant values.

Further, the cases of Newtonian and Herschel–Bulkley fluids are discussed separately. For Newtonian fluid, we have \(\eta (D) = \mu = const\), \(\tau _0=0\) and then from Eq. (55) it follows that

$$\begin{aligned} v_i = \frac{x_3^2}{2\mu } \frac{\partial p}{\partial x_i} + A_i \frac{x_3}{2\mu } + B_i. \ \ \end{aligned}$$
(56)

Taking into account the boundary conditions at \(x_3=0\) and \(x_3=W\)

$$\begin{aligned} v_i=0, i=1,2 \end{aligned}$$
(57)

one can obtain

$$\begin{aligned} v_i = -x_3 \frac{W-x_3}{2\mu } \frac{\partial p}{\partial x_i}. \end{aligned}$$
(58)

From the expressions for fluid fluxes

$$\begin{aligned} q_i = \int _{0}^{W} v_i d x_3, \ \ i = 1,2 \end{aligned}$$
(59)

and Eq. (58) we can get the equations connecting the derivatives of the fluid pressure with the fluxes

$$\begin{aligned} q_i = - \frac{W^3}{12\mu } \frac{\partial p}{\partial x_i}, \end{aligned}$$
(60)

Equations (60) coincide with the ones used in Shokin et al. (2015) and Cherny et al. (2016).

Using the Equations (58) and (60) one can obtain shear rate value

$$\begin{aligned} {\dot{\gamma }}= \frac{\partial v_1}{\partial x_3} = \frac{6q_i}{W^2}. \end{aligned}$$
(61)

For HB-fluid, the viscosity function \(\eta (D)\) depends on the velocity components that are functions of \(x_3\). For integrating Equations (55), let us move to a coordinate system, in which the direction of \(x_1\)-axis coincides with the direction of the velocity vector \({\mathbf {u}}\). In this coordinate system, we have

$$\begin{aligned} \eta (D) = K \left( \frac{\partial v_1}{\partial x_3} \right) ^{n-1} + \tau _0 \left( \frac{\partial v_1}{\partial x_3} \right) ^{-1} . \end{aligned}$$
(62)

Integrating Equation (55) and taking into account the boundary conditions (57) and the expressions for the viscosity function (62), one can obtain

$$\begin{aligned} v_1= & {} -\frac{n K^{-1/n}}{ (n+1)} \left( \frac{\partial p}{\partial x_1} \right) ^{1/n}\nonumber \\&\left( (0.5W-z_{\tau })^{1+1/n}-v_{d}(x_3)^{1+1/n}\right) , \end{aligned}$$
(63)

where

$$\begin{aligned} v_d(x_3) = {\left\{ \begin{array}{ll} 0.5W- x_3 - z_{\tau }, &{} x_3<0.5W-z_{\tau } \\ 0, &{} 0.5W-z_{\tau } \le x_3 \le 0.5W+z_{\tau } \\ x_3 - 0.5W - z_{\tau }, &{} x_3>0.5W+z_{\tau } \end{array}\right. }, \ z_{\tau } = \tau _0 \left| \frac{\partial p}{\partial x_1}\right| ^{-1} . \end{aligned}$$

The latter formula for \(v_d(x_3)\) takes into account the typical velocity profile of the HB fluid flow between two parallel plates as shown in Fig. 21. This formula is correct if the pressure gradient is enough to overcome the yield stress

$$\begin{aligned} \left| \frac{\partial p}{\partial x_1}\right| > \frac{2 \tau _0}{W}. \end{aligned}$$
(64)

Otherwise the fluid is motionless. Similarly, one can obtain the expression for the second velocity component \(v_2\). Now the expressions for the fluid fluxes can be written in the following form:

$$\begin{aligned} q_i= & {} - \frac{n}{(4n+2) (2K)^{1/n}} W^{2+1/n} \left( \frac{\partial p}{\partial x_i} \right) ^{1/n}\nonumber \\&\left( 1-\frac{2z_{\tau }}{W} \right) ^{1+1/n} \left( 1+\frac{2z_{\tau }}{W}\frac{n}{n+1} \right) . \end{aligned}$$
(65)
Fig. 21
figure 21

Typical velocity profile of HB fluid flow between two parallel plates

Extracting the first degree of the pressure derivatives \({\partial p}/{\partial x_i}\) in the right-hand side of Equation (65) gives

$$\begin{aligned} q_i = - \frac{W^3}{12\eta _a} \frac{\partial p}{\partial x_i}, \end{aligned}$$
(66)

where \(\eta _{a}\) is an apparent viscosity. It can be expressed in terms of the pressure derivatives \({\partial p}/{\partial x_i}\) as well as in terms of the fluxes.

The expression of the apparent viscosity through the pressure derivatives can be obtained from the Equation (63)

$$\begin{aligned} \eta _{a}= & {} \eta _{p} = \frac{(2K)^{1/n}(2n+1)}{6n} \left| \frac{\partial p}{\partial x_1}\right| ^{(n-1)/n}\nonumber \\&W^{(n-1)/n} + \frac{(4n+2)2^{1/n} \tau _0}{3 n \left| \frac{\partial p}{\partial x_1}\right| ^{1/n} W^{1/n}}. \end{aligned}$$
(67)

Rewriting Equation (67) in another coordinate system results in the substitution of the pressure derivative \({\partial p}/{\partial x_i}\) by the pressure gradient \(\nabla p\), which is invariant relative to coordinate transformations

$$\begin{aligned} \eta _a= & {} \eta _p(\nabla p) = \frac{(2K)^{1/n}(2n+1)}{6n} (W |\nabla p|)^{(n-1)/n}\nonumber \\&+ \frac{(4n+2)2^{1/n} \tau _0}{3 n (W |\nabla p|)^{1/n} }, \end{aligned}$$
(68)

where

$$\begin{aligned} |\nabla p| = \left( \left( \frac{\partial p}{\partial x_1}\right) ^2 + \left( \frac{\partial p}{\partial x_2}\right) ^2 \right) ^{1/2}. \end{aligned}$$

To get the formula for the apparent viscosity in terms of the fluid fluxes, let us sum the squares of the right-hand sides of Equation (65) for \(q_1\) and \(q_2\) and express the pressure gradient \(|\nabla p|\) in terms of the fluid flow vector module \(|\mathbf{q}| = (q_1^2 + q_2^2)^{1/2}\), which is invariant relative to coordinate transformations similar to \(|\nabla p|\). Then we obtain

$$\begin{aligned} \eta _{q} = \frac{K}{6} \left( \frac{4n+2}{n}\right) ^n \left( \frac{W^2}{|\mathbf{q}|} \right) ^{1-n} + \frac{2n+1}{3(n+1)}\frac{\tau _0 W^2}{|\mathbf{q}|} . \nonumber \\ \end{aligned}$$
(69)

Formula (69) is a 2D generalization to the case of Herschel–Bulkley rheology of the expressions obtained earlier by other researchers for a Power-law fluid. For the case of the Power-law fluid (\(\tau _0 \equiv 0\)) the Equation (66) with the apparent viscosity expressed by formula (69) has been used in Sousa et al. (1993), Garagash (2006), and the expression (68) has been employed in Ouyang et al. (1997) and Rungamornrat et al. (2005). Here as well as in Cherny and Lapin (2016) formula (68) is chosen because then there is no need to calculate the fluid flux while the fluid flow equations are solving.

Note that like in case of Newtonian fluid one can obtain the value of shear rate using the Equations (63), (66) (69), but the calculations are quite more complicated

$$\begin{aligned} {\dot{\gamma }}= \frac{\partial v_1}{\partial x_3} = \frac{1}{K^{1/n}} \left[ K \left( \frac{4n+2}{n}\right) ^n \left( \frac{|{\mathbf {q}}|}{W^2} \right) ^n + \frac{3n+1}{n+1} \tau _0 \right] ^{1/n}. \nonumber \\ \end{aligned}$$
(70)

Integrating equation (51) over \(x_3\) and taking into account relationships (53) and (59), one can obtain the equation for the fracture width W

$$\begin{aligned} \frac{\partial W}{\partial t} + \frac{\partial q_1}{\partial x_1} + \frac{\partial q_2}{\partial x_2} = 0. \end{aligned}$$
(71)

The substitution of the expression for the fluxes (66) into Equation (71) gives the equation for the fluid pressure

$$\begin{aligned} \frac{\partial }{\partial x_1} \left( \frac{W^{3}}{12 \eta _{a}} \frac{\partial p}{\partial x_1} \right) + \frac{\partial }{\partial x_2} \left( \frac{W^{3}}{12 \eta _{a}} \frac{\partial p}{\partial x_2} \right) = \frac{\partial W}{\partial t}. \end{aligned}$$
(72)

One should note that one of the major difficulties while modeling the fluid flows with non-zero yield stress (such as Bingham of Herschel–Bulkley fluids) is the necessity to trace the boundary between the region of non-zero strain rate (46), where the Eqs. (49), (52), (62), (65)–(66), (67), (68), (69), (72) are applied and the region (47) where the fluid flow should be treated as the rigid body motion (\(D_{ij}=0\)). For the modeling of the regions with the zero and non-zero strain rate tensor (yielded and unyielded regions) within the framework of the same equations various modifications of the expression for the viscosity (49), are used, and they prevent its degeneration when \(D\rightarrow 0\) (Mitsoulis 2007).

In the problems of fluid flow in the hydraulic fracture the search for the boundary of the unyielded region is easier than in the problems of fluid flow in the regions with the given boundaries. The unyielded region appears in each point of the fracture (\(x_1, x_2\)) in the middle of its cross-section (between the fracture sides \(0<x_3<W\)). The boundaries of the region are calculated explicitly, while solving the problem of fluid flow between two parallel plates (65). This solutions is used to obtain equation (72) As it can be concluded from (64) the pressure gradient is bounded from zero. The area where strain rate tensor is zero can fill the whole cross section, then fluid at this crack point (\( x_1, x_2 \)) stops, and equation (72) is not valid. But the calculations made in Sect. 5.3, show that such a situation can not be realized in the simulation of the fractures considered. This means that at all points of the fracture the fluid flows, and the pressure gradient always exceeds \( \frac{2 \tau _0}{W} \). This allows us to use equation (72) with the expression for viscosity (68) without additional modifications.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cherny, S., Lapin, V., Kuranakov, D. et al. 3D model of transversal fracture propagation from a cavity caused by Herschel–Bulkley fluid injection. Int J Fract 212, 15–40 (2018). https://doi.org/10.1007/s10704-018-0289-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10704-018-0289-4

Keywords

Navigation