Skip to main content
Top
Published in: Rock Mechanics and Rock Engineering 4/2024

Open Access 18-01-2024 | Original Paper

2D Phase-Field Modelling of Hydraulic Fracturing Affected by Cemented Natural Fractures Embedded in Saturated Poroelastic Rocks

Authors: Nima Sarmadi, Mohaddeseh Mousavi Nezhad, Quentin J. Fisher

Published in: Rock Mechanics and Rock Engineering | Issue 4/2024

Activate our intelligent search to find suitable subject content or patents.

search-config
loading …

Abstract

The interaction between a propagating hydraulic fracture (HF) and a pre-existing natural fracture (NF) embedded in saturated poroelastic rock formations is studied numerically in 2D plane–strain configurations. In this study, the phase-field method is further developed to be employed for modelling the HF propagation and the evolution of tensile and shear failure in geo-materials as gradient-type diffusive damaged zones. The shear slippage and dilation mechanisms inside the cemented NF are modelled using a Mohr–Coulomb–Griffith failure criterion that fitted in the framework of the phase-field fracture using appropriate energy functionals. The most important factors controlling the HF–NF interaction outcome are the approaching angle, differential in-situ stress, and hydro-mechanical characteristics of the NF. It is found out that higher tensile and shear strengths of the cemented NF are in favour of the crossing outcome when the differential stress is high enough to mobilise the resisting shear stresses against the slippage. Small hydraulic aperture (low hydraulic conductivity) for the NF is also in favour of the crossing outcome which helps to restrict the pressurised region local to the HF tip, lowering the possibility of shear slippage in the NF and the HF’s diversion. It is also concluded that the injection rate and the viscosity of fracturing fluid are operative factors to be adjusted for increasing the chance of crossing, a critical element for successful operation of hydraulic fracturing for effective use of subsurface energy resources.
Notes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abbreviations
\({\sigma }_{\text{H}}\)
Maximum horizontal in-situ stress
\({\sigma }_{\text{h}}\)
Minimum horizontal in-situ stress
\({\sigma }_{\text{v}}\)
Vertical in-situ stress (due to gravity)
\({\sigma }_{\text{t}}\)
Tensile strength of the material
\({\sigma }_{\text{c}}\)
Compressive strength of the material
\(b\)
Blanton’s parameter (Eq. 1)
\(c\)
Mohr–Coulomb Cohesion
\(\varphi\)
Mohr–Coulomb friction angle
\({\mu }_{\text{int}}\)
NF’s Coefficient of friction
\(\beta\)
HF–NF Approaching angle
\(d\)
Damage (phase-field) parameter
\({G}_{c}\)
Critical energy release rate
\({K}_{Ic}\)
Mode-I fracture toughness
\({l}_{0}\)
Length-scale parameter
\({{\varvec{m}}}^{\text{r}}\)
Fluid mass flow rate in space
\(m\)
Fluid mass moving through the body
\({\varvec{q}}\)
Fluid volume rate on the boundaries
\(\lambda\)
1St Lame parameter
\(\mu\)
2Nd Lame parameter (shear modulus)
\(E\)
Elastic Young’s modulus
\(\alpha\)
Biot’s coefficient
\({M}_{\text{b}}\)
Biot’s modulus
\(\xi\)
Volumetric fluid content term
\(t\)
Time
\({\varvec{\varepsilon}}\)
Strain tensor of the solid skeleton
\({\varepsilon }_{\text{v}}\)
Volumetric strain of the solid
\({\varvec{\sigma}}\)
Cauchy total stress tensor
\({{\varvec{\sigma}}}^{\prime}\)
Cauchy effective stress tensor
\(p\)
Pore fluid pressure
\(K\)
Drained bulk modulus of the rock
\({K}_{\text{s}}\)
Bulk modulus of the solid constituent
\({K}_{\text{u}}\)
Undrained bulk modulus of the rock
\({{\varvec{v}}}^{\text{r}}\)
Relative superficial velocity
\({\varvec{k}}\)
Permeability tensor (units: mD or m2)
\(\widehat{\mathcal{H}}\)
Maximum energy history term
\({\varepsilon }_{\text{t}}\)
Critical tensile strain at the rupture
\({{\varvec{\varepsilon}}}_{D}\)
Deviatoric part of the strain tensor
\({\sigma }_{\text{i}}{\prime}\)
Effective principal stresses (i = 1, 2, 3)
\({\varepsilon }_{\text{i}}\)
Principal strains (i = 1, 2, 3)
\({{\varvec{T}}}^{*}\)
Traction forces on the boundaries
\({\varvec{g}}\)
Vector of gravity
\(\Upsilon\)
Time integration term
\({H}_{\text{el}}\)
Element size
\(w\)
Fracture (HF) width
\({\mu }_{\text{f}}\)
Fluid dynamic viscosity
\({Q}_{\text{inj}}\)
Fluid injection rate applied in the HF
\({P}_{\text{HF}}\)
Fluid pressure inside the HF
\({l}_{\text{HF}}\)
Half-length of the HF (KGD)

1 Introduction

The hydraulic fracturing technique is usually employed in low-permeable geological formations such as shale, granite, and sandstone which may contain oil and gas resources or may be targeted for exploiting the geothermal energy or trapping and storage of hydrocarbons. One of the main goals of the hydraulic fracturing operations is to develop a vast fracture network throughout the reservoir to increase the overall permeability of the reservoir for easing the industrial exploitation process. In real-world applications, driving the hydraulic fractures in the desired direction can be challenging because of the existence of randomly distributed natural discontinuities which may cause unintended diversion or arrest of the propagating hydraulic fracture (HF). Microseismic technology has proven that pre-existing faults might be activated in the form of shear slippage during the hydrofracking operations (Pine and Batchelor 1984). Mine-back observations have also indicated that a single natural fracture (NF) can affect a propagating HF in three general ways: termination, diversion, and offsetting (Warpinski and Teufel 1987).
The role of a NF as a mechanical discontinuity or a flow conduit depends on its inherent properties, such as the openness, length, mechanical strength, spatial arrangement, and the degree of internal cementation (Gale et al. 2014). Stochastic analyses have shown that spatial heterogeneity of macropores in geo-materials can contribute to developing pathways and micro-cracks for groundwater flow (Mousavi Nezhad et al. 2011; Nezhad and Javadi 2011), which can potentially affect the hydraulic fracture operations in larger scales. In shale reservoirs, NFs are usually found as embedded inclusions of clay or coal materials which add to the hydro-mechanical heterogeneity of the system (Warpinski et al. 1993). Depending on the pre-historic conditions, the degree of cementation, mineral constituents of the geological formations, and the smears of dissimilar material in the host rock may act as a flow conduit or a sealed barrier against the flow of fluid in subsurface reservoirs (Fisher and Knipe 2001). Therefore, the effect of permeability and hydraulic aperture of the NFs on the propagation of the HF cannot be ignored in studying the HF–NF interactions; it has also been addressed in some laboratory-scale experimental works (Jeffrey et al. 2009; Kear et al. 2017).
Analytical and numerical approaches have been developed to predict the outcome of the HF–NF interaction ever since the effect of NFs on the HF propagation path through experiments and field observations have been understood. From a mechanistic point of view, the conditions under which the HF is expected to cross the NF were first formulated analytically with regard to the following factors in 2D plane–strain case: the approaching angle (\(\beta\) in Fig. 1a), the in-situ stress field (\({\sigma }_{\text{v}},{\sigma }_{\text{H}},{\sigma }_{\text{h}}\)), tensile strength of the rock (\({\sigma }_{\text{t}}\)), fluid pressure inside the HF (\({P}_{\text{HF}}\)), and the coefficient of friction (\({\mu }_{\text{int}}\)) between the faces of the NF (Blanton 1986; Warpinski and Teufel 1987; Renshaw and Pollard 1995). The overburden stress \({\sigma }_{\text{v}}\) is the vertical in-situ stress in the direction of gravity, and \({\sigma }_{\text{H}},{\sigma }_{\text{h}}\) are the maximum and minimum horizontal in-situ stresses, respectively. In this study, transverse hydraulic fracturing is considered; the HF propagates horizontally in a plane that is perpendicular to the horizontal wellbore (Michael et al. 2018). In the 3D configuration of transverse hydraulic fracturing operations, the horizontal wellbore is planted parallel to the direction of the least in-situ stresses (i.e., \({\sigma }_{\text{h}}\)), and in a depth at which the overburden stress is less than the maximum horizontal in-situ stress (\({\sigma }_{\text{H}}>{\sigma }_{\text{v}}>{\sigma }_{\text{h}}\)). Therefore, the HF tends to develop horizontally rather than vertically in the reservoir rock. In our 2D numerical analyses, the plane of interest is horizontal, and therefore, the influential in-situ stresses are \({\sigma }_{\text{H}}\) and \({\sigma }_{\text{h}}\). The factors influencing the interactive behaviour of the fractures are specified in Fig. 1a. The foundation of analytical formulations for predicting the HF–NF interaction outcome is based on two possible scenarios: (i) internal shear slippage of the NF which is affected by the HF front and in-situ stresses, and (ii) re-initiation of the HF along the NF. Blanton (1986) proposed an analytical formulation based on some of the factors mentioned above for predicting the HF crossing scenario for a NF with internal shear strength. This criterion is based on satisfying the condition for breaking down the bulk rock along the NF after initial arrest of the HF; the crossing scenario happens if
$$\left({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\right)>\frac{{\sigma }_{\text{t}}}{\text{cos}\left(2\beta \right)-b\text{sin}\left(2\beta \right)},$$
(1)
where \({\sigma }_{\text{H}}\) and \({\sigma }_{\text{h}}\) are the maximum and minimum horizontal stresses, and \(b\) is a parameter which depends on the amount of fluid in the NF and the roughness of NF’s internal faces. Warpinski and Teufel (1987) considered the shear slippage of the NF as a deterministic factor in predicting crossing or arrest behaviour of the HF. They used the linear Coulomb’s frictional law to check the occurrence of shear slippage inside the NF. Renshaw and Pollard (1995) accounted for the singular stresses at the crack-tip blunted by an orthogonal NF with frictional faces. They hypothesised that the HF is arrested initially by one face of the NF; then, shear stresses are transferred through the NF to re-initiate the HF on the opposite face of the discontinuity. Renshaw and Pollard’s formulation accounts for inherent characteristics of the NF which are related to the inner-face roughness and cohesion of the filler materials and is written as
$$3{\mu }_{\text{int}}{\sigma }_{\text{H}}^{2}+\left({c}_{\text{int}}-\left({\sigma }_{\text{t}}+{\sigma }_{\text{h}}\right)\left(1+{\mu }_{\text{int}}\right)\right){\sigma }_{\text{H}}-{c}_{\text{int}}\left({\sigma }_{\text{t}}+{\sigma }_{\text{h}}\right)\ge 0,$$
(2)
where \({c}_{\text{int}}\) and \({\mu }_{\text{int}}\) are the cohesion and the coefficient of friction of the NF (interface). The above inequality is plotted in Fig. 1b for the case of a rock with tensile strength of \({\sigma }_{\text{t}}\)=8MPa, and it is evident that increasing both \({c}_{\text{int}}\) and \({\mu }_{\text{int}}\) is in favour of the crossing outcome. Capturing the shear slippage within the NF is essential for a reliable modelling of the fracture interaction problem. The slippage (shear failure) of the NF can significantly affect the state of effective stresses and, therefore, change the direction of HF propagation in the vicinity of the NF.
For numerical modelling of the interaction between a propagating HF and a cemented NF, a wide range of methods have been used. Physical continuity inside the cemented NF makes the continuum damage methods appropriate for capturing the stress field nicely throughout the fractured (damaged) zones (Dahi Taleghani et al. 2016). In these methods, the region of weakness (cemented NF) is usually modelled by assigning softer material properties compared to the host rock (Cooke and Underwood 2001; Dyskin and Caballero 2009; Rahman et al. 2009). Modelling the HF crossing through cemented NFs is mainly connected with the tensile and shear failure of the embedded weak region, whereas modelling open frictional NFs is more associated with the consideration of normal and shear forces on the contacting faces of the NF, where discrete models perform better (Xie et al. 2016; Shi et al. 2017; Khoei et al. 2018). In this study, we use the continuum-based phase-field method to study the problem of HF–NF interaction. Modelling the HF propagation using the phase-field method goes back to the work of Bourdin et al. (2012) who defined an energy functional based on the fluid pressure and the displacement jump within the fracture and considered the discontinuity as a gradient-type diffusive damaged zone. The bulk energy stored in the saturated porous continuum provides the driving force for the evolution of damage throughout the continuum. Later, the general variational formulation of the phase-field fracture was implemented in hydro-mechanical models which rest on Biot's theory of poroelastic coupling in water-saturated porous media (Biot 1941; Biot and Willis 1957) by other research groups (Mikelić et al. 2015; Wilson and Landis 2016; Heider and Markert 2017; Mauthe and Miehe 2017; Sarmadi and Mousavi Nezhad 2023).
The phase-field method has performed well in modelling fracture propagation in heterogeneous materials (Mousavi Nezhad et al. 2018a, b) and in bi-material solids with adhesive interfaces and local weaknesses (Verhoosel and de Borst 2013; Nguyen et al. 2016) because of its inherent versatility to be implemented in the finite-element analysis. The first attempts to model the interface failure using the phase-field method was founded based on the energetic form of the cohesive zone element method. The energetical approach to model phase-field fractures was first adapted to the cohesive discrete fracture by Verhoosel and de Borst (2013) who defined a new form of the surface energy functional dependent on the displacement jump in the direction orthogonal to the diffusive fracture. In another work on modelling the interface between aggregates and the cement paste in concrete materials (Nguyen et al. 2016), a multi-phase-field formulation for the propagation of diffusive fractures was developed, where the interface was formulated by a secondary phase-field interacting with the primary phase-field fracture in the bulk. Paggi and Reinoso (2017) adapted a mixed formulation of the phase-field fracture and cohesive zone element and studied the problem of crack impinging on an interface embedded in the bulk medium specifically to forecast possible deflection and penetration of the fracture in bi-material elastic media, depending on the degree of dissimilarity between the values of fracture toughness, the length-scale, and tensile strength of the interface and bulk regions. The effect of the assumed width of the interface region embedded in the bulk elastic medium in the context of the phase-field fracture method has been studied by Hansen-Dörr et al. (2019, 2020), who recommended to use the compensation factors for correcting the values of the critical energy release rate and elastic material properties of the interface region based on the width of the interface and the chosen length-scale parameter. A similar study was conducted by Yoshioka et al. (2021) to apply corrections on the value of the fracture toughness used in the model for a finite-width NF that intersects a propagating HF. The recommendations and correction factors proposed by Hansen-Dörr et al. (2020) are analytically sound, so they are implemented in the numerical model developed in this study to reduce the dependency of the propagation behaviour of the blunted fracture on the input parameters.
Using the phase-field method for modelling fractures as gradient-type diffusive damaged zones, some research works have studied the HF–NF interaction problem by simply assigning softer material properties to the NF region compared to the bulk rock. Alotaibi et al. (2020) modelled the propagation of the HF in layered poroelastic media, where the width of the layer dissimilar to the bulk is significant. They concluded that the deflection and crossing behaviour of the HF facing the layer with dissimilar material properties is related to the ratio of the assigned critical energy release rate of the layer to that of the bulk as well as the inclination of the layer. In a work by Yi et al. (2020), where the phase-field method is employed, the evolution of damage within the NF (the interface) is only allowed to happen if the maximum principal strain of the NF elements exceed a user-defined critical value; this can only observe the tensile failure of the interface and disregards the possible shear slippage. Lepillier et al. (2020) simply assumed smaller values of the fracture toughness for the NF compared to that of the bulk while not considering any change of the hydraulic conductivity for the region of the NF. Despite the importance of the possible shear failure (slippage) within the NF which was understood from the analytical studies, the consideration of shear slippage inside the NF has not been addressed well in the studies that have employed the phase-field method. The transition from tensile cracking to shear cracking in rocks subjected to the confining compressive stress field has been observed in the experiments (Ramsey and Chester 2004) and in the field (Ferrill et al. 2012). Accordingly, some research works can be found in the context of phase-field method that address such phenomena in modelling the failure of rock-like materials via separating mode-I and mode-II fracturing processes (Zhang et al. 2017) and defining energy functionals with regard to compressive shear stresses (Zhou et al. 2019; Fei and Choo 2020). In the context of phase-field method, modelling mode-I or mode-II fracturing in the material depends on the implementation of correct formulations of tensile and deviatoric energy contributions (Amor et al. 2009). For this purpose, the phase-field method is developed to model the shear failure as damaged zones over a continuous finite-element mesh structure based on the Mohr–Coulomb–Griffith failure model (Jaeger et al. 2009).
We aim to identify the controlling parameters on the outcome of the interaction between a propagating HF and an embedded cemented NF in the saturated porous rock using our developed numerical code which is capable of modelling mixed-mode tensile and shear failure as gradient-type diffusive damaged zones over the finite elements. The fracture driving force for the evolution of damaged zones due to shearing is formulated using the deviatoric part of the strain energy if the state of shear stress exceeds the Mohr–Coulomb shear failure envelope. In this paper, the cemented NF is referred to the long-standing pre-existing cracks/discontinuities in the bulk mass of the rock which have formed during geological periods. We only analyse the cemented NF within which some extents of material continuity can be found; analysing the effects of open NFs (pre-existing cracks) is out of the scope of this study. The effect of hydraulic conductivity of the NF region on the outcome of HF–NF interaction and on the possible shear slippage of the NF is specifically evaluated. The internal tensile strength and cohesion of the NF region are the main controlling parameters in the model for capturing the dilation of the NF because of the fluid penetration. The developed model is verified with respect to analytical solutions and experimental data to ensure its reliability, effectivity, and applicability in modelling the fracture interaction mechanisms. To identify the controlling factors in the HF–NF problem, we conduct an extensive parametric study by varying the approaching angle, in-situ stresses, fluid injection rate, and hydro-mechanical characteristics of the cemented NF.

2 Methodology

In this section, mathematical formulations to model the propagation of the diffusive HF in water-saturated poroelastic media are elaborated first, and the formulations are further developed to model the mixed-mode tensile and shear failure in geo-materials as damaged zones.

2.1 Phase-Field Diffusive Fracture (Gradient-Type Damaged Zone)

Griffith’s hypothesis on the cracking phenomenon implies that new surfaces within the volume of a deforming body are formed as a result of the release of the stored strain energy due to the external loading. In fact, the released energy is consumed to form new surfaces, so the fracture propagation problem can be mathematically defined as searching for an equilibrium state of energy in a fracturing body (Griffith 1921). The energy contribution of the new surfaces (\({\Psi }_{\text{surface}}\)) within the volume of the body can be formulated as \({\int }_{\Gamma }{G}_{c} dA\), where \({G}_{c}\) is so-called the critical energy release rate. The domain of this integral (\(\Gamma\)) is mathematically unknown, because fractures form anonymously throughout the body. Therefore, Bourdin et al. (2000) re-formulated \({\Psi }_{\text{surface}}\) as an integral over the volume of the body (\(\Omega\)) by assuming the sharp discontinuity as a gradient-type diffusive damaged zone, represented by the phase-field parameter (\(d\)), and introducing the surface energy density functional \(\gamma \left(d,\nabla d\right)\) in the following manner (Ambrosio and Tortorelli 1990):
$${\Psi }_{\text{surface}}\left(d\right)={G}_{c}{\int }_{\Omega }\gamma \left(d,\nabla d\right) dV={G}_{c}{\int }_{\Omega }\left[\frac{{l}_{0}}{2}{\left|\nabla d\right|}^{2}+\frac{1}{2{l}_{0}}{\left(1-d\right)}^{2}\right]dV.$$
(3)
In the above equation, \(\nabla\) is the gradient operator in Cartesian coordinates, \(d\) is the so-called damage (phase-field) parameter, and \({l}_{0}\) is length-scale parameter which characterises the width of the diffusive fracture. The value of damage parameter varies between \(d\) = 0 for fully damaged region and \(d\)=1 associated with the intact region, which is governed by a function shown in Fig. 2.
The fracture propagation can be predicted by searching for the equilibrium in the internal potential energy of the body; the regularised functional which needs to be minimised for tracking the fracture propagation is formulated as (Bourdin et al. 2000)
$$\Pi \left({\varvec{\varepsilon}},d\right)=\underbrace{{\int }_{\Omega }g\left(d\right) {\Psi }_{\text{bulk}}\left({\varvec{\varepsilon}}\right)dV}_{\text{degraded bulk strain energy}}+\underbrace{{G}_{c}{\int }_{\Omega }\left[\frac{{l}_{0}}{2}{\left|\nabla d\right|}^{2}+\frac{1}{2{l}_{0}}{\left(1-d\right)}^{2}\right]dV}_\text{surface energy of the cracks},$$
(4)
where \({\Psi }_{\text{bulk}}\left({\varvec{\varepsilon}}\right)\) is the strain energy stored in the bulk material (undamaged state), and \(g\left(d\right)={d}^{2}+{\kappa }_{\epsilon }\) is the degradation function based on the damage field which will be discussed later. In the process of energy minimisation, as the fractures propagate and new surfaces form throughout \(\Omega\), the first integral in the above equation decreases because of the effect of the development of damage throughout the body (i.e., decreasing effect of the degradation function when \(d\to 0\)), while the second integral grows constantly as the damaged zones spread in the continuum.

2.2 Coupled Hydro-Mechanical Formulation for Modelling Saturated Poroelastic Rocks

Consider a physical body \(\Omega\) in Euclidean space \({\mathbb{R}}^{3}\) as a saturated porous medium consisted of two phases, the porous solid skeleton with mass density \({\rho }_{\text{s}}\) and the fluid phase with mass density \({\rho }_{\text{f}}\) in the pores. The initial porosity for a volume element \(dV\) is defined as
$${\varphi }_{0}^{\text{f}}=\frac{d{V}_{\text{f}}}{dV},$$
(5)
where \(d{V}_{\text{f}}=dV-d{V}_{\text{s}}\) is the volume of the fluid (equal to the volume of pores in the saturated condition). The initial mass of the unit volume \(dV\) is equal to \({m}_{0}={\rho }_{\text{s}}{\varphi }_{0}^{\text{s}}+{\rho }_{\text{f}}{\varphi }_{0}^{\text{f}}\). When the domain \(\Omega\) is subjected to the hydro-mechanical loading, including the fluid injection and the exertion of mechanical forces on the external boundaries, both fluid and solid undergo motion. Herein, the movement of the fluid phase is formulated with respect to the deformation of the solid skeleton, since the developed formulations are based upon a Lagrangian description of the continuum. In the process of deformation in time \(t\), the rate of the fluid mass \(m\) flowing into or out of the unit volume of the element is defined as
$$\frac{{D}^{\text{s}}}{Dt}\left(m\right)=-\nabla \cdot {{\varvec{m}}}^{\text{r}},$$
(6)
where \({D}^{\text{s}}\left(\cdot \right)/Dt\) operator is the Lagrangian time derivative of the fluid mass with respect to the solid motion, and the vector \({{\varvec{m}}}^{\text{r}}\) is the fluid mass flow rate in the space moving throughout the porous medium, defined as
$${{\varvec{m}}}^{\text{r}}=\left({\rho }_{\text{f}}\right){\varvec{q}},$$
(7)
where \({\varvec{q}}\) is the fluid volume rate affecting on the boundaries of the porous body, as shown in Fig. 3. Hydro-mechanical response of water-saturated porous rock is associated with two mechanisms: (i) the dilation of the solid skeleton due to an increase in the pore fluid pressure, and (ii) the increase in the pore fluid pressure when the rock is under compression (Detournay and Cheng 1993). For modelling such behaviour, Biot’s theory of poroelastic coupling in saturated porous rock is employed, assuming incompressible fluid and linear elastic behaviour for the solid skeleton (Biot and Willis 1957). The total Cauchy stress tensor \({\varvec{\sigma}}\), the strain energy stored in the solid skeleton \({\Psi }_{\text{eff}}\left({\varvec{\varepsilon}}\right)\), and the energy stored in the pore fluid phase \({\Psi }_{\text{fluid}}\left({\varvec{\varepsilon}},\xi \right)\) are defined, respectively, as
$${\varvec{\sigma}}={{\varvec{\sigma}}}^{\prime}-\alpha p{\varvec{I}},$$
(8)
$${\Psi }_{\text{eff}}\left({\varvec{\varepsilon}}\right)=\frac{\lambda }{2}{\left({\varepsilon }_{\text{v}}\right)}^{2}+\mu \text{tr}\left[{{\varvec{\varepsilon}}}^{2}\right],$$
(9)
$${\Psi }_{\text{fluid}}\left({\varvec{\varepsilon}},\xi \right) =\frac{{\alpha }^{2}{M}_{\text{b}}}{2}{\left({\varepsilon }_{\text{v}}\right)}^{2}-\alpha {M}_{\text{b}}\xi \left({\varepsilon }_{\text{v}}\right)+\frac{{M}_{\text{b}}}{2}{\xi }^{2},$$
(10)
where the strain tensor is calculated from the deformation field (\({\varvec{u}}\)) as \({\varvec{\varepsilon}}=1/2\left(\nabla {\varvec{u}}+{\nabla }^{\text{T}}{\varvec{u}}\right)\), and \({\varepsilon }_{\text{v}}=\text{tr}\left[{\varvec{\varepsilon}}\right]\). We have based our model on the infinitesimal strain theory; this assumption is valid when the deformations are very small due to a high stiffness in the rock and low confining stresses. Several studies have developed phase-field hydraulic fracture models in poroelastic media based on linear elasticity (Mauthe and Miehe 2017; Heider and Markert 2017; Zhou et al. 2018), similar to the model developed herein. In higher magnitudes of the confining pressure, rock’s mechanical response becomes nonlinear (Choo and Sun 2018), and deformations become significant when the stiffness decreases; therefore, employing the kinematics of large deformations (finite strains) would be essential (Miehe and Mauthe 2016). The tensor \({{\varvec{\sigma}}}{\prime}=\partial {\Psi }_{\text{eff}}\left({\varvec{\varepsilon}}\right)/\partial{\varvec{\varepsilon}}\) is so-called the effective Cauchy stress tensor is a part of \({\varvec{\sigma}}\) that only applies on the solid skeleton structure. The pore fluid pressure is defined as \(p=\partial {\Psi }_{\text{fluid}}\left({\varvec{\varepsilon}},\xi \right)/\partial \xi ={M}_{\text{b}}\left(\xi -\alpha {\varepsilon }_{\text{v}}\right)\), \(\xi\) is the volumetric fluid production term (or so-called fluid content), and \(d\xi =dm/{\rho }_{\text{f}}\) is the variation of the fluid volume per total unit volume \(dV\) due to fluid mass transport (Detournay and Cheng 1993). The pore fluid pressure (\(p\)) is considered as a positive value, which is incorporated in the total Cauchy stress tensor \({\varvec{\sigma}}\) (Eq. 8) in a way that negative value for the hydrostatic part of the stress would imply a compressive stress state. In the above equations, \(\alpha\) is the Biot’s coefficient defined as (Biot and Willis 1957)
$$\alpha =1-K/{K}_{\text{s}},$$
(11)
where \(K\) is the drained bulk modulus of the porous rock and \({K}_{\text{s}}\) is the bulk modulus of the solid constituents (i.e., grains). Several laboratorial methods can be found in the literature to measure Biot’s coefficient (Franquet and Abass 1999). Biot’s modulus \({M}_{\text{b}}\) is defined as \({M}_{\text{b}}={K}_{\text{s}}^{2}\left({K}_{\text{u}}-K\right)/{\left({K}_{\text{s}}-K\right)}^{2}\), where \({K}_{\text{u}}\) is the undrained bulk modulus of the porous rock and can be measured experimentally (Detournay and Cheng 1993).

2.2.1 Governing Equations of the Hydro-Mechanical Model (u–p Formulation)

In this section, the governing equations to form the coupled u–p formulation are introduced, and the essential and natural boundary conditions are depicted in Fig. 3. The fractures are expected to form throughout the saturated porous body, so the degradation function must be applied on the effective stresses in the solid skeleton, equally meaning that the total Cauchy stress tensor is defined as \({\varvec{\sigma}}=g\left(d\right){{\varvec{\sigma}}}{\prime}-\alpha p{\varvec{I}}\). Thus, the balance of linear momentum reads as
$$\nabla \cdot \left(g\left(d\right){{\varvec{\sigma}}}{\prime}-\alpha p{\varvec{I}}\right)+\left({m}_{0}+m\right){\varvec{g}}=0.$$
(12)
The total mass of the saturated porous body must be conserved in the process of deformation and the flow of fluid in to/out of the domain, so the total mass balance equation reads as
$$\alpha {\dot{\varepsilon }}_{\text{v}}+\nabla \cdot {{\varvec{v}}}^{\text{r}}+\dot{p}/{M}_{\text{b}}=0,$$
(13)
where \({{\varvec{v}}}^{\text{r}}={\varphi }^{\text{f}}\left({{\varvec{v}}}_{\text{f}}-{{\varvec{v}}}_{\text{s}}\right)\) is the relative superficial velocity of the fluid phase by definition, \({{\varvec{v}}}_{\text{f}}\) and \({{\varvec{v}}}_{\text{s}}=\partial {\varvec{u}}/\partial t\) are the intrinsic velocities of the fluid and solid phases, respectively, and \({\dot{\varepsilon }}_{\text{v}}\) is the volumetric strain rate of the solid skeleton. Darcy’s law is used to model the fluid flow throughout the porous body, based on which the velocity field \({{\varvec{v}}}^{\text{r}}\) in Eq. (13) can be substituted by
$${{\varvec{v}}}^{\text{r}}=-{\varvec{k}}\cdot \left(\nabla p-{\rho }_{\text{f}}{\varvec{g}}\right),$$
(14)
where \(p\) is the pore fluid pressure field, \({\varvec{k}}=\left(\overline{k }/{\mu }_{\text{f}}\right){\varvec{I}}\) is the symmetric permeability tensor, \({\mu }_{\text{f}}\) is the dynamic viscosity of the fluid, and \(\overline{k }\) is the intrinsic permeability of the porous material. Modelling the fluid flow inside the HF will be discussed in Sect. 2.4.

2.3 Finite-Element Weak Forms to Model the HF Propagation in Poroelastic Rocks

The finite-element weak forms are derived by minimising the total energy with respect to arbitrary damage, displacements, and pore fluid pressure. The total potential energy functional \(\Pi \left({\varvec{\varepsilon}},\xi ,d\right)\) is consisted of the effective elastic strain energy of the fractured solid skeleton \({\Psi }_{\text{eff}}\left({\varvec{\varepsilon}},d\right)\), the energy of the pressurised fluid inside the pores \({\Psi }_{\text{fluid}}\left({\varvec{\varepsilon}},\xi \right)\), and the surface energy \({\Psi }_{\text{surface}}\left(d\right)\) which is the contribution of the newly propagated fractures within the volume (\(\Omega\)). Employing a continuous Galerkin method and applying the natural boundary conditions, shown in Fig. 3, the weak form of the balance of linear momentum is derived by finding a stationary position of \(\delta\Pi \left({\varvec{\varepsilon}},\xi ,\overline{d }\right)=0\) with respect to arbitrary displacement field \({\varvec{\eta}}\), while the damage field is assumed to be fixed (\(\overline{d }\)):
$${\int }_{\Omega }\nabla{\varvec{\eta}}:g\left(\overline{d }\right)\left[{\mathbb{C}}\right]:\nabla {\varvec{u}} dV-{\int }_{\Omega }\alpha p\left(\nabla \cdot{\varvec{\eta}}\right)\boldsymbol{ }dV-{\int }_{\Omega }{\varvec{\eta}}\cdot \rho {\varvec{g}}\boldsymbol{ }d\text{V}-{\int }_{{\partial\Omega }_{{\varvec{\sigma}}}}{\varvec{\eta}}\cdot {{\varvec{T}}}^{*}d\text{A}=0.$$
(15)
In the above equation, \({\mathbb{C}}\) is the fourth-order elastic stiffness tensor, \({{\varvec{T}}}^{*}\) is the vector of traction forces applied on the boundaries of the body (see Fig. 3), \(g\left(\overline{d }\right)={\overline{d} }^{2}+{\kappa }_{\epsilon }\) is the degradation function, as defined by (Bourdin et al. 2000; Miehe et al. 2010b), for a fixed state of damage in the body, and \({\kappa }_{\varepsilon }\ll 1\) is the regularisation parameter. The second weak form completing the u–p formulation is obtained from the total mass conservation law (Eq. 13) and applying the Galerkin weighting function \(\psi\) and natural boundary conditions shown in Fig. 3:
$${\int }_{\Omega }\alpha \left(\nabla \cdot \dot{{\varvec{u}}}\right)\psi d\text{V}-{\int }_{\Omega }\nabla \psi \cdot {{\varvec{v}}}^{r}dV+\frac{1}{{M}_{\text{b}}}{\int }_{\Omega }\psi \left(\dot{p}\right)dV-{\int }_{{\partial\Omega }_{{\varvec{q}}}}\psi {\varvec{q}}d\text{a}=0.$$
(16)
Obeying the staggered approach in solving for the damage field, both displacement and pore fluid pressure fields are assumed to be fixed (\(\overline{{\varvec{u}} },\overline{p }\)) throughout the domain, to eventually solve for the damage field (phase-field parameter \(d\)). The weak form of the damage evolution model is obtained by finding a stationary position of \(\delta\Pi \left(\overline{{\varvec{u}} },\overline{p },d\right)=0\) with respect to the arbitrary damage \(\omega\) as
$${\int }_{\Omega }2\widehat{\mathcal{H}}\left(d\right)\omega d\text{V}+{G}_{c}{l}_{0}{\int }_{\Omega }\nabla d\cdot \nabla \omega d\text{V}-{G}_{c}{\int }_{\Omega }\frac{\left(1-d\right)}{{l}_{0}}\omega d\text{V}=0,$$
(17)
where \(\widehat{\mathcal{H}}\) is a history field, defined as the maximum energy acquired by a volume element during the history of deformation to satisfy the fracture irreversibility condition (Miehe et al. 2010b). The history field reads as
$$\widehat{\mathcal{H}} =\underset{t\in \left[{t}_{0},{t}_{\text{i}}\right]}{\text{max}}{\Psi }_{\text{eff}}\left(\overline{{\varvec{u}} },t\right),$$
(18)
where \({t}_{0}\) and \({t}_{\text{i}}\) are the initial and current times of the deformation history. The method for defining the contribution of the effective bulk energy in calculating the value of \(\widehat{\mathcal{H}}\) can be chosen differently, based on the physical failure behaviour of the material. The most widely used criteria on defining \(\widehat{\mathcal{H}}\) have been introduced by (i) Amor et al. (2009) based on splitting the strain tensor into volumetric and deviatoric parts, and (ii) Miehe et al. (2010b) based on the spectral decomposition of the strain tensor into tensile and compressive parts. These formulations are written in the following neglecting the operator \(\underset{t\in \left[{t}_{0},{t}_{\text{i}}\right]}{\text{max}}\left[\cdot \right]\) for the sake of simplicity:
$${\widehat{\mathcal{H}}}_{\text{Amor}}=\frac{K}{2}{\left({\langle \text{tr}\boldsymbol{ }\left[{\varvec{\varepsilon}}\right]\rangle }_{+}\right)}^{2}+\mu {{\varvec{\varepsilon}}}_{D}\cdot {{\varvec{\varepsilon}}}_{D},$$
(19)
$${\widehat{\mathcal{H}}}_{\text{Miehe}}=\frac{\lambda }{2}{\left({\langle \text{tr }\left[{\varvec{\varepsilon}}\right]\rangle }_{+}\right)}^{2}+\mu \text{tr}\boldsymbol{ }\left[{\left({\left[{\varvec{\varepsilon}}\right]}^{+}\right)}^{2}\right].$$
(20)
In the above equations, \(\lambda ,\mu\) are the elastic Lame parameters and \(K\) is the bulk modulus of porous rock, \({\langle x\rangle }_{\pm }=\left(x\pm \left|x\right|\right)/2\), and \({{\varvec{\varepsilon}}}_{D}={\varvec{\varepsilon}}-\left(\text{tr}\left[{\varvec{\varepsilon}}\right]/3\right){\varvec{I}}\) is the deviatoric part of the strain. The positive part of the strain tensor \({\left[{\varvec{\varepsilon}}\right]}^{+}\) is calculated as stated in Eq. (33), based on the spectral decomposition of the strain tensor \({\varvec{\varepsilon}}\). It is worth mentioning that the method of defining \(\widehat{\mathcal{H}}\) also affects the calculation of the effective stresses within the solid skeleton, see Sect. 2.5. For the case of \({\widehat{\mathcal{H}}}_{\text{Miehe}}\), the degradation function \(g\left(d\right)\) only affects the elements under tension and not the ones under compression. For the case of \({\widehat{\mathcal{H}}}_{\text{Amor}}\), the negative part of the volumetric stress remains unaffected by the degradation function \(g\left(d\right)\). As a result, contacting compressive stresses remain in place within the damaged zone, which is a better physical representative of the fracturing phenomenon especially in mode-II failure (Amor et al. 2009).

2.4 Fluid Flow Within the HF and Throughout the Porous Rock

For modelling the fluid flow within the HF, Poiseuille’s law representing the flow between parallel plates is postulated, so the fluid velocity field is formulated as
$${{\varvec{v}}}^{\text{crack}}=\frac{{w}^{2}}{12{\mu }_{\text{f}}}\nabla p,$$
(21)
where \(w\) is the fracture (HF) width. Since the HF is modelled as a gradient-type damaged zone over a continuous finite-element mesh, there will be no discrete boundary considered for calculating the HF’s width when employing the Poiseuille’s law. Equation (21) is only used to model the fluid flow inside the HF, so one needs to specify the elements which are guaranteed to be within the fully damaged zone. Herein, we recommend using a strain-based criterion, Eqs. (22) and (23), that can identify whether the critical strain state in the element has exceeded the critical tensile strain of the material or not; Poiseuille’s law is only employed over these elements. To illustrate these critical elements within the fully damaged zone, a typical phase-field fracture in mode-I opening status, including the relevant displacement and the maximum principal strain fields, is depicted in Fig. 4. It is shown that the fully damaged zone can be restricted to the critical elements for which the maximum principal strain \({\varepsilon }_{\text{p}1}\) is significantly higher compared to the adjacent elements. To maintain the consistency of the hydro-mechanical model (u–p formulation) for modelling the fluid flow in both HF and the intact porous rock, the permeability tensor \({\varvec{k}}\), used in Eq. (14), should be modified as
$${\varvec{k}}= {{\varvec{k}}}_{\text{intact}}+H\left({\varvec{\varepsilon}},\boldsymbol{ }{\varepsilon }_{\text{t}}\right)\left(\frac{{w}^{2}}{12{\mu }_{\text{f}}}{\varvec{I}}-{{\varvec{k}}}_{\text{intact}}\right),$$
(22)
where
$$H\left({\varvec{\varepsilon}},\boldsymbol{ }{\varepsilon }_{\text{t}}\right)=\left\{\begin{array}{cc}1& \quad \text{if}\,\, \sqrt{{\sum }_{\text{i}=1}^{3}{\varepsilon }_{\text{i}}^{2}}\ge {\varepsilon }_{\text{t}}\\ 0& \quad \text{if} \,\,\sqrt{{\sum }_{\text{i}=1}^{3}{\varepsilon }_{\text{i}}^{2}}<{\varepsilon }_{\text{t}}\end{array}\right..$$
(23)
In the above equations, \(H\left({\varvec{\varepsilon}},\boldsymbol{ }{\varepsilon }_{\text{t}}\right)\) is a jump function, \({\varepsilon }_{\text{i}}\) are the principal strains calculated from spectral decomposition of the strain tensor \({\varvec{\varepsilon}}\), refer to Eq. (33), and \({{\varvec{k}}}_{\text{intact}}=\left(\overline{k }/{\mu }_{\text{f}}\right){\varvec{I}}\) is defined for the intact porous rock with no damage. The critical tensile strain of the material is \({\varepsilon }_{\text{t}}={\sigma }_{\text{t}}/E\), which if exceeded, the element is encountered as “critical” and is ensured to be within the fully damaged zone.
Our method of capturing the critical elements based on exceeding the critical tensile strain limit \({\varepsilon }_{\text{t}}\), gives a physical meaning to the evolution of permeability for the elements inside the diffusive damaged zone. Calculation of the HF’s width (\(w\)) is affected by the element size (\({H}_{\text{el}}\)) and the strain tensor \(\left[{\varvec{\varepsilon}}\right]\), as stated in Eq. (24). According to Fig. 4, for a triangular mesh structure, the minimum \({H}_{\text{el}}\) in the vicinity of the fracture cannot be larger than \({l}_{0}/4\) to guarantee the existence of adequate elements capturing the fully damaged zone and to ensure a more accurate calculation of \(w\) (Sarmadi and Mousavi Nezhad 2023). Satisfying the condition \({H}_{\text{el}}\le {l}_{0}/4\) for the damaged zone is compatible with the recommendations of (Bourdin et al. 2000; Miehe et al. 2010b) on choosing the element size \({H}_{\text{el}}\) to ensure the mesh-independency of the answer. For calculating \(w\), we follow the procedure proposed by Mauthe and Miehe (2017), as mentioned below, albeit with further recommendations on choosing the value of \({L}_{\perp }\):
$$w=\frac{\nabla \left(d\right) \left[{\varvec{\varepsilon}}\right] {\nabla }^{T}\left(d\right)}{{\left|\nabla d\right|}^{2}}{L}_{\perp }.$$
(24)
In the relationship above, \({L}_{\perp }\) is called the characteristic length which is measured in the direction perpendicular to the orientation of the diffusive fracture and is typically considered equal to \({H}_{\text{el}}\) (Mauthe and Miehe 2017). By comparing the distribution of the principal strain fields for two different mesh configurations in Fig. 4, we realise that having a finer mesh may result in appearing more than one element within the fracture walls. As a result, we recommend choosing \({L}_{\perp }\) equal to \({l}_{0}/4\) while keeping the element size around the HF-tip \({H}_{\text{el}}\le {l}_{0}/4\) by performing a simple h-refinement on the coarser elements. It is worth mentioning that the length-scale parameter \({l}_{0}\) can be chosen with regard to the physical parameters, such as the fracture toughness and tensile strength of the material (Amor et al. 2009; Miehe et al. 2015; Sarmadi et al. 2023).

2.5 Mixed-Mode Cracking in Geo-materials

In this section, we develop a mixed-mode formulation to calculate the crack driving force accounting for both tensile and shearing failure modes. This approach rests upon the Mohr–Coulomb–Griffith failure model (Jaeger et al. 2009) to account for tensile and compressive shear failure due to the accumulated stresses in the continuum. The final purpose is to use this formulation in modelling shear slippage of NFs which can affect the propagation path of the HF. The strong form of the damage evolution model can be derived from Eq. (17) as
$$2\frac{\widehat{\mathcal{H}}}{{G}_{c}}\left(d\right)-\left({l}_{0}\right)\Delta d-\frac{\left(1-d\right)}{{l}_{0}}=0,$$
(25)
where \(\Delta\) is the Laplacian operator, and the term \(\widehat{\mathcal{H}}/{G}_{c}\), so-called the crack driving force, is specifically responsible for the evolution of damage throughout the solid skeleton. In geo-materials, the critical energy release rate for mode-II cracking is usually greater than the one for mode-I. Zhang et al. (2017) modified the crack driving force by introducing a hybrid format of Eq. (25) as follows:
$$2\left(\frac{{\widehat{\mathcal{H}}}_{1}}{{G}_{c}^{I}}+\frac{{\widehat{\mathcal{H}}}_{2}}{{G}_{c}^{II}}\right)\left(d\right)-\left({l}_{0}\right)\Delta d-\frac{\left(1-d\right)}{{l}_{0}}=0,$$
(26)
where \({G}_{c}^{I}\) and \({G}_{c}^{II}\) are the critical energy release rates for mode-I and mode-II fracture, respectively; also, \({\widehat{\mathcal{H}}}_{1}=\lambda /2{\left({\langle \text{tr}\left[{\varvec{\varepsilon}}\right]\rangle }_{+}\right)}^{2}\) and \({\widehat{\mathcal{H}}}_{2}=\mu \text{tr}\left[{\left({\left[{\varvec{\varepsilon}}\right]}^{+}\right)}^{2}\right]\) are defined based on the spectral decomposition of the strain tensor. This approach still accounts for the elements under tension only, while the shear cracking (mode-II) can happen for the elements under compression in rock-like materials. A comparison between the cracking behaviour of limestone and a type of glass (polymethyl methacrylate) in single-flaw square plates subjected to compression has been done by Ingraffea and Heuze (1980), see Fig. 5, which confirms having a different crack pattern in rocks. Calculating the crack driving force based on Miehe’s approach replicates a crack pattern similar to the one shown in Fig. 5c. Miehe’s approach on splitting the energy based on spectral decomposition of the strain tensor can predict the crack pattern in glass, since the brittle tensile failure is the dominant mode. A more advanced criterion based on a statistical damage model was developed by Zhou et al. (2019) to form the crack driving force (\(\widehat{\mathcal{H}}/{G}_{c}\)) based on the elastic parameters, internal angle of friction \(\varphi\), and the internal cohesion \(c\). The modification by Zhou et al. (2019) enforces the dominancy of the shearing mode in the damage evolution, while the work of Zhang et al. (2017) still tends to forecast tensile cracks. In modelling the cracking phenomenon in geo-materials, both tensile and shear failure might occur depending on the tensile and shear strength limits. To predict the correct crack patterns in geo-materials using the phase-field fracture model, which is essentially a continuum damage model (Marigo et al. 2016), we construct the crack driving force based on the idea of mixed-mode cracking, Eq. (26), but use two different formulations to calculate \({\widehat{\mathcal{H}}}_{1}\) and \({\widehat{\mathcal{H}}}_{2}\). The maximum principal stress criterion with the critical tensile strength threshold (\({\sigma }_{\text{t}}\)), introduced in (Miehe et al. 2015), is employed for calculating \({\widehat{\mathcal{H}}}_{1}\) and is in charge of driving the damaged zone in the opening mode-I:
$${\widehat{\mathcal{H}}}_{1}={\left\langle \sum_{i=1}^{3}\frac{{\langle {\sigma }_{\text{i}}^{\prime}\rangle }_{+}^{2}}{{\sigma }_{\text{t}}^{2}}-1\right\rangle }_{+}\frac{{\sigma }_{\text{t}}^{2}}{2{E}^{\prime}}.$$
(27)
In the above equation, \({\sigma }_{\text{i}}^{\prime}\) are the principal effective Cauchy stresses obtained via spectral decomposition of the undegraded Cauchy effective stress tensor, and \({E}^{\prime}\) is the effective elastic young modulus. For the compressive shear failure of geo-materials, Mohr–Coulomb–Griffith failure criterion (Jaeger et al. 2009) is implemented in the model to calculate the contribution of \({\widehat{\mathcal{H}}}_{2}\), see Eq. (28). If the stress state exceeds the Mohr–Coulomb shear failure envelope, \({\widehat{\mathcal{H}}}_{2}\) would take the value of the deviatoric part of the strain energy, defined by Amor et al. (2009):
$${\widehat{\mathcal{H}}}_{2}=H\left({\sigma }_{\text{i}}^{\prime}\right)\cdot \left(\mu {{\varvec{\varepsilon}}}_{D}\cdot {{\varvec{\varepsilon}}}_{D}\right),$$
(28)
$$H\left({\sigma }_{\text{i}}^{\prime},c,\varphi \right)=\left\{\begin{array}{cc}1& \frac{1}{2}\left|{\sigma }_{\text{i}}^{\prime}-{\sigma }_{\text{j}}^{\prime}\right|>\frac{1}{2}\left|{\sigma }_{\text{i}}^{\prime}+{\sigma }_{\text{j}}^{\prime}\right|\text{sin}\varphi +\left(c\right)\text{cos}\varphi \\ 0& \text{otherwise}\end{array}.\right.$$
(29)
Equation (29) introduces \(H\left({\sigma }_{\text{i}}^{\prime},c,\varphi \right)\), a jump function that depends on the principal stresses and checks if the Mohr–Coulomb shear failure condition (\(i,\text{j}=\text{1,2},3\)) is met. Parameters \(c\) and \(\varphi\) are cohesion and internal friction angle, and \({{\varvec{\varepsilon}}}_{D}\) is the deviatoric part of the strain tensor. Finally, the finite-element weak form of the damage evolution model, which was previously introduced in Eq. (17), is replaced by the following:
$${\int }_{\Omega }2\left(\frac{{\widehat{\mathcal{H}}}_{1}}{{G}_{c}^{I}}+\frac{{\widehat{\mathcal{H}}}_{2}}{{G}_{c}^{II}}\right)\cdot \left(d\right)\omega d\text{V}+{l}_{0}{\int }_{\Omega }\nabla d\cdot \nabla \omega d\text{V}-{\int }_{\Omega }\frac{\left(1-d\right)}{{l}_{0}}\omega d\text{V}=0.$$
(30)
The effective stress tensor is defined in Eqs. (31) and (32), depending on whether the failure occurs in tensile or shear mode, respectively:
$${{\varvec{\sigma}}}^{\prime}=g\left(d\right)\left[\lambda {\langle \text{tr}\left[{\varvec{\varepsilon}}\right]\rangle }_{+}\left[{\varvec{I}}\right]+2\mu {\left[{\varvec{\varepsilon}}\right]}^{+}\right]+\left[\lambda {\langle \text{tr}\left[{\varvec{\varepsilon}}\right]\rangle }_{-}\left[{\varvec{I}}\right]+2\mu {\left[{\varvec{\varepsilon}}\right]}^{-}\right],$$
(31)
$${{\varvec{\sigma}}}^{{{\prime}}}=K{\langle \text{tr}\left[{\varvec{\varepsilon}}\right]\rangle }_{-}\left[{\varvec{I}}\right]+g\left(d\right)\left[K{\langle \text{tr}\left[{\varvec{\varepsilon}}\right]\rangle }_{+}\left[{\varvec{I}}\right]+2\mu {{\varvec{\varepsilon}}}_{D}\right].$$
(32)
For the failed elements under compressive shearing, the procedure proposed by Amor et al. (2009) is followed using Eq. (32). The volumetric part of the effective stress remains unaffected by the degradation function \(g\left(d\right)\) to maintain the normal compressive stresses within the fully damaged zone. In tensile failure, however, the formulation of Miehe et al. (2010a) is employed for obtaining effective stresses, i.e., Eq. (31). Tensors \({\left[{\varvec{\varepsilon}}\right]}^{\pm }\) are obtained from the spectral decomposition of \({\varvec{\varepsilon}}\) as
$${\left[{\varvec{\varepsilon}}\right]}^{\pm }=\sum_{\text{i}=1}^{3}{\langle {\varepsilon }_{\text{i}}\rangle }_{\pm }{{\varvec{n}}}_{\text{i}}\otimes {{\varvec{n}}}_{\text{i}},$$
(33)
where \({\varepsilon }_{\text{i}}\) are the principal strains and \({{\varvec{n}}}_{\text{i}}\) are the relative principal directions. In the next section, the presented model will be verified by comparing the results to the KGD analytical solution of HF propagation and experimental data on the shear failure in rock-like materials.

3 Numerical Algorithm and the Method Verification

3.1 Linearised Weak Forms of the Coupled u–p Formulation

The Newton–Raphson method is employed to solve for the displacement and fluid pressure fields in an iterative manner. After discretising Eq. (16) in time domain, the weak form of the mass conservation law reads as
$$-{\int }_{\Omega }\psi \left(\nabla \cdot \delta {\varvec{u}}\right)d\text{V}-\frac{1}{{M}_{\text{b}}}{\int }_{\Omega }\psi \cdot \left(\delta p\right)d\text{V}+\left(1-\Upsilon \right)\Delta t{\int }_{\Omega }\nabla \psi \cdot {{\varvec{v}}}_{\text{n}}^{r} d\text{V}+\Upsilon \Delta t{\int }_{\Omega }\nabla \psi \cdot {{\varvec{v}}}_{\text{n}+1}^{r} Jd\text{V}+\left(1-\Upsilon \right)\Delta t{\int }_{{\partial\Omega }_{\mathbf{q}}}\psi \cdot {\varvec{Q}} d\text{A}+\Upsilon \Delta t{\int }_{{\partial\Omega }_{\mathbf{q}}}\psi \cdot {{\varvec{Q}}}_{\text{n}} d\text{A}=0,$$
(34)
where \(\Upsilon\) is the time-integration parameter, \(\Delta t={t}_{\text{n}+1}-{t}_{\text{n}}\), \(\delta p={p}_{\text{n}+1}-{p}_{\text{n}}\) and \(\delta {\varvec{u}}={{\varvec{u}}}_{\text{n}+1}-{{\varvec{u}}}_{\text{n}}\). Note that for calculating the velocity fields \({{\varvec{v}}}_{\text{n}+1}^{r}\) and \({{\varvec{v}}}_{\text{n}}^{r}\), Eq. (14) is employed, albeit with incorporating the correction of the permeability tensor \({\varvec{k}}\) mentioned in Eq. (22) into the model. The weak form of the balance of linear momentum reads as
$${\int }_{\Omega }\nabla{\varvec{\eta}}:\left[{\mathbb{C}}\left({{\varvec{\sigma}}}_{\text{n}}^{\prime},{\overline{d} }_{\text{n}}\right)\right]:\nabla \left(\delta {\varvec{u}}\right) dV-{\int }_{\Omega }\alpha \left(\nabla \cdot{\varvec{\eta}}\right)\cdot \left(\delta p\right)dV+{\int }_{\Omega }\left[{{\varvec{\sigma}}}_{\text{n}}^{\prime}-\alpha {p}_{\text{n}}{\varvec{I}}\right]:\left[{\nabla }^{T}\left(\delta {\varvec{u}}\right)\cdot \nabla{\varvec{\zeta}}\right]dV-{\int }_{\Omega }{\varvec{\eta}}\cdot \rho {\varvec{g}}\boldsymbol{ }d\text{V}-{\int }_{{\partial\Omega }_{{\varvec{\sigma}}}}{\varvec{\eta}}\cdot {{\varvec{T}}}^{*}d\text{A}=0,$$
(35)
where the degraded elasticity tensor \({\mathbb{C}}\left({{\varvec{\sigma}}}_{\text{n}}^{\prime},{\overline{d} }_{\text{n}}\right)\) is updated based on the current state of damage in the body at time-step \({t}_{\text{n}}\). Notice that the effective stress tensor \({{\varvec{\sigma}}}_{\text{n}}^{\prime}\) is calculated using Eqs. (31) or (32) depending on the failure’s mode. The numerical algorithm, elaborated in Fig. 6, to solve for the variables using finite-element method in an iterative manner is implemented in a MATLAB-based code for the case of 2D plane–strain configurations. We choose an implicit approach in all the simulations by setting the time integration parameter as \(\Upsilon\)=1. A simple h-refinement strategy is employed to re-fine first-order 3-noded elements by adding an extra node on the longest boundary of the element using the built-in MATLAB function “refineMesh” in the PDE toolbox. The algorithm goes through several cycles of refinement until the condition \({H}_{\text{el}}<{l}_{0}/4\) is satisfied for the elements in which the nodal values of damage are (\(d<0.7\)). It is noticed that the coupled u–p formulation and the damage evolution model are solved after each cycle of performing mesh refinement over the domain.

3.2 Verification of the HF Propagation Model with the Analytical KGD Solution

To simplify the inherent complexities of the HF propagation in reservoirs, the analytical KGD solution was proposed and developed in (Zheltov 1955; Geertsma and De Klerk 1969) to calculate the fluid pressure, width, and length of the HF based on the equilibrium of viscous forces of a Newtonian fluid between two parallel surfaces. The KGD solution proposes the following equations, assuming an elliptical cavity within an elastic continuum in plane–strain configuration:
$${l}_{\text{HF}}\left(t\right)=0.539 \sqrt[6]{E^{\prime}{Q}_{\text{inj}}^{3}/{\mu }_{\text{f}}}\left(\sqrt[3]{{t}^{2}}\right),$$
(36)
$${w}_{\text{HF}}\left(t\right)=2.36 \sqrt[6]{{\mu }_{\text{f}}{Q}_{\text{inj}}^{3}/E^{\prime}}\left(\sqrt[3]{t}\right),$$
(37)
$${P}_{\text{HF}}\left(t\right)-{\sigma }_{\text{min}}=1.09 \sqrt[3]{{\mu }_{\text{f}}{{E}^{\prime}}^{2}} \left(\sqrt[3]{t}\right).$$
(38)
In the above solution, \({l}_{\text{HF}}\left(t\right)\) is the HF’s half-length, and \({P}_{\text{HF}}\left(t\right)\) is the fluid pressure inside the HF during the propagation with respect to time, and the HF's width \({w}_{\text{HF}}\left(t\right)\) is the ellipse’s diameter on the axis perpendicular to the direction of propagation. The analytical KGD solution has been commonly used to validate the numerical simulation of the HF propagation using discrete fracture methods (Dahi Taleghani 2009; Xie et al. 2016). To compare the numerical results of our phase-field HF model to the KGD solution, we use the material properties in Table 1, similar to the values chosen in (Xie et al. 2016), and the boundary value problem is shown in Fig. 7. The critical energy release rate in our simulations is chosen based on LEFM-based relationship \({G}_{c}={K}_{\text{IC}}^{2}/E^{\prime}\).
Table 1
Material properties used for the KGD model (Xie et al. 2016)
Parameter
Name
Value
unit
\(E\)
Young’s modulus
20
\(\text{GPa}\)
\(\nu\)
Poisson’s ratio
0.25
\({K}_{IC}\)
Mode-I fracture toughness
1
\(\text{MPa}.\sqrt{m}\)
\(Q\)
Fluid injection rate, 2D plane strain
\(2\times {10}^{-4}\)
\({\text{m}}^{3}/\left(\text{m}\cdot \text{s}\right)\)
\(\overline{k }\)
Intrinsic permeability of the bulk
10
\(\text{mD}\)
\({\mu }_{\text{f}}\)
Fluid dynamic viscosity
\(1\times {10}^{-3}\)
\(\text{Pa}.\text{s}\)
To highlight the dependency of the results on the length-scale parameter (\({l}_{0}\)), simulations are done for three different values of \({l}_{0}\) equal to 0.02, 0.03, and 0.04 m, and the results are compared to the KGD solution in Figs. 8 and 9. The validity of the phase-field method for the HF simulation can be seen in Fig. 9a, where the propagation speed is compatible with the solution obtained from the analytical KGD solution. The reduction of fluid pressure inside the HF with the propagation, shown in Fig. 8b, is reasonably matched with the analytical KGD solution, albeit with significant discrepancies for larger values of \({l}_{0}\). According to Fig. 8a, the HF’s width is acceptably in accordance with the KGD solution. Considering the results in Figs. 8 and 9, it can be concluded that reducing the length-scale parameter \({l}_{0}\) (e.g., \({l}_{0}=0.02\)) results in outputting the numerical results that better match the KGD solution. This conclusion is compatible with the theoretical proof made by (Bellettini and Coscia 1994), implying that when \({l}_{0}\to 0\), the elasticity solution of sharp fracture is retrieved. In modelling HF using the phase-field method, the choice of element size also affects the characteristics of the HF, specifically the fracture width (w). Increasing the element size (\({H}_{\text{el}}\)) can result in an overestimate in the calculated value of \(w\) using Eq. (24), which can be inferred from Fig. 10. To clarify the effect of \({H}_{\text{el}}\) on the calculation of \(w\), three simulations are borrowed from Sect. 3.2 and run by setting the element size to \({H}_{\text{el}}\) equal to \({l}_{0}/6\), \(\frac{{l}_{0}}{4},\) and \({l}_{0}/2\). The calculated values of \(w\) for the critical (fully damaged) elements throughout the domain are shown in Fig. 10 for three cases of \({H}_{\text{el}}\) in one specific snapshot of the simulations at \(t\)=1 s. It is seen in Fig. 10c that the calculated values of \(w\) in some elements for the case of \({H}_{\text{el}}={l}_{0}/2\) are significantly higher than the plotted values in the other two cases in Fig. 10a, b; the expected HF’s width at \(t\)=1 s can be seen in Fig. 8a, which matches the majority of \(w\) values plotted in the histograms related to the cases of \({H}_{\text{el}}\) equal to \({l}_{0}/6\), and \({l}_{0}/4\) (Fig. 10a, b). Thus, setting the \({H}_{\text{el}}\) larger than \({l}_{0}/4\) can potentially cause an overestimate in calculating the \(w\) in some of the critical elements.

3.3 Verification of the Damage Model Due to Shear Failure

The results of the formulations presented in Sect. 2.4 to model shear failure is tested by comparing the outputs of this model to the experimental data and the numerical results of the particle flow code of the failure of single-flaw concrete specimens under compression, as documented in (Jin et al. 2017). The boundary value problem is identical to Fig. 5a with \(L=50\text{mm}\), \(H=100\text{mm}\), \(a=10\text{mm}\), and \(b=0.5\text{mm}\). Material properties are the same as those presented in (Jin et al. 2017) which are listed in Table 2. The tensile strength \({\sigma }_{\text{t}}\) is calculated with regard to the compressive strength \({\sigma }_{\text{c}}\) and the internal friction angle \(\varphi\) using the relationship \({\sigma }_{\text{t}}=\left(1-\text{sin}\varphi \right){\sigma }_{\text{c}} /\left(1+\text{sin}\varphi \right)\), recommended by Labuz and Zang (2012). The length-scale parameter is chosen as \({l}_{0}=1 \text{mm}\) by calibrating the model for the case \(\beta =60^\circ\), and it is kept the same for other cases of \(\beta\). The crack patterns taken from our analyses are compared to the experimental observations for two geometries with \(\beta =0^\circ\) and \(\beta =60^\circ\) in Fig. 11. The deformed geometries in Fig. 11c, d, g, and h show that compressive shear failure has been modelled in a realistic way without any interpenetration of two pieces of the broken samples along the diffusive fractures. In fact, the developed model can effectively capture the existing compressive stresses that have developed within the damaged zones. The strain–stress response of the single-flaw specimens with different inclination angles taken from our numerical simulations show a strong compatibility with the experimental data and the results of the discrete modelling of fracture using the particle-flow code (Jin et al. 2017), see Fig. 12a. However, the experimental strain–stress curves show significant discrepancies from the numerical ones which can be related to the potential heterogeneities and the existence of micro-cracks in the concrete specimens. The effect of micro-cracks can be seen in the experimental stress–strain curve with a low tangent stiffness at the beginning of loading, and its gradual increase by the application of the load. The inherent heterogeneity and existence of micro-cracks cannot be captured neither in our numerical model nor in the numerical analysis of Jin et al. (2017). Experimental data indicates that the captured maximum uniaxial compressive stress (UCS) grows as the inclination angle of the embedded flaw in the concrete specimen increases (Jin et al. 2017). Our numerical results in Fig. 12b show the same trend and confirm the reliability of our developed phase-field model in modelling shear failure in geo-materials.
Table 2
Material properties used for uniaxial compressive tests on single-flaw specimens (Jin et al. 2017)
Parameter
Name
Value
Unit
\(E\)
Young’s modulus
2.5
\(\text{GPa}\)
\(\nu\)
Poisson’s ratio
0.20
\({\sigma }_{\text{c}}\)
Uniaxial compressive strength
27.21
\(\text{MPa}\)
\(\varphi\)
Internal friction angle
30
degrees
\(c\)
Internal cohesion
1
\(\text{MPa}\)
\({\sigma }_{\text{t}}\)
Uniaxial tensile strength
9.08
\(\text{MPa}\)
\({G}_{c}^{I}, {G}_{c}^{II}\)
Critical energy release rates
1200, 2400
\(\text{N}/\text{m}\)
\({l}_{0}\)
Length-scale parameter
0.001
\(\text{m}\)

3.4 Shear Slippage Inside the Cemented NF

A series of semi-circular bending (SCB) tests have been done by Wang et al. (2018) on hydrostone samples to mimic the propagation of the HF intersected by a cemented NF filled by plaster. The material properties of hydrostone and plaster are detailed in Table 3. In this section, we try to examine the capacity of our model in predicting the mixed-mode tensile and shear failure inside the NF by comparing our results to the data provided in (Wang et al. 2018). We model the interaction between the propagating fracture and the cemented NF without the effect of fluid and compare the outcome to the experimental observations. The SCB test has been introduced by Chong and Kuruppu (1984) as a versatile method to measure mode-I fracture toughness (\({K}_{IC}\)) in rocks based on the maximum load capacity (\({P}_{\text{max}}\)) of the specimen using the relationship:
$${K}_{Ic}=\frac{{Y}_{I}^{\prime}{P}_{\text{max}}\sqrt{\pi a}}{2BR},$$
(39)
where \(B\) and \(R\) are the thickness and radius of the semi-circular specimen, \(a\) is the length of the initial notch in the centre. The value \({Y}_{I}^{\prime}\) is the non-dimensional stress intensity factor, calculated based on the relationships proposed by Lim et al. (1994) and Kuruppu et al. (2014) depending on the geometry. Having chosen the same geometry (a specific \({Y}_{I}^{\prime}\)) for both analytical solutions and the numerical simulations, numerically predicted values of (\({P}_{\text{max}}/B\)) with respect to (\(a/R\)) are compared to the analytical answers of (Lim et al. 1994; Kuruppu et al. 2014) in Fig. 13a. This comparison is done for three different values of \({l}_{0}\) due to the inherent sensitivity of the phase-field method to this parameter (Gironacci et al. 2018). The strong agreement between the predictions of the developed phase-field model and the analytical solutions ensures the reliability of the model in predicting the fracture propagation and the brittle behaviour of the hydrostone in the first place.
Table 3
Material properties of hydrostone and plaster in the SCB tests, reported in (Wang et al. 2018)
Material
\(E\) (\(GPa\))
\(\nu\)
\({\sigma }_{\text{t}}\) (\(MPa\))
\({\sigma }_{\text{c}}\) (\(MPa\))
\({K}_{\text{IC}}\) (\(MPa\sqrt{m}\))
\({G}_{c}={K}_{IC}^{2}/E^{\prime}\) (\(N/m\))
Hydrostone
6.18
0.32
4.82
35.75
0.42
25.6
Plaster
1.84
0.32
2.07
9.17
0.19
17.6
Thereafter, our developed model, presented in Sect. 2.5, is tested by modelling the hydrostone samples containing plaster inclusions representing a cemented NF. Using the geometry shown in Fig. 14a with the dimensions \(R\)=20mm, \(S\)=32mm, \(a\)=4mm, \({w}_{\text{NF}}\)=0.4mm, and \({L}_{\text{NF}}\)=16mm, the material properties presented in Table 3, and assuming \(\varphi =\left({{\sigma }_{\text{c}}-\sigma }_{\text{t}}\right)/\left({{\sigma }_{\text{c}}+\sigma }_{\text{t}}\right)=40^\circ\) and \(c\)=800kPa for the plaster inclusion (cemented NF), the interaction between the propagating vertical fracture and the pre-existing NF are simulated. Increasing the width and the inclination of the embedded NF reduces the value of the fracture toughness obtained from the simulations, according to Fig. 14b, since the overall stiffness of the sample has decreased. This conclusion, based on the ensuing numerical results, is in line with the experimental data. According to Fig. 15, it is found out that reducing the angle of approach (increasing the inclination of the NF) reduces is not in favour of the crossing outcome. The dominant scenarios seen for three approaching angles \(\beta\)=\(90^\circ\), \(\beta\)=\(60^\circ\), and \(\beta\)=\(30^\circ\) are crossing, offset crossing, and diversion, respectively, which are in line with the experimental observations in (Wang et al. 2018). As the vertically propagating fracture approaches an inclined NF, more elements inside the NF region undergo shear failure, because the shearing stresses exceed Mohr–Coulomb strength envelope. As the approaching angle decreases, the shearing mode becomes dominant and causes the development of damage along the NF.

4 Interaction Between the HF and Cemented NFs

In this section, we focus on the limited-length cemented NFs embedded in the saturated poroelastic rocks. To choose realistic values for the material properties and to compare the ensuing numerical results to the experimental observations, the boundary value problem and the material properties, which are listed in Table 4 and presented in (Kear et al. 2017), are used for the simulations. Square domains containing a single inclined NF are subjected to confining stresses (\({\sigma }_{\text{H}}\) and \({\sigma }_{\text{h}}\)) and the fluid injection rate of \({Q}_{\text{inj}}\) into the centre, as introduced in (Kear et al. 2017) and shown schematically in Fig. 16. Due to symmetry, half of the square domain with appropriate boundary conditions, shown in Fig. 16, is modelled to study the HF–NF interaction problem. The elastic modulus and fracture toughness of the interface (NF) region (\({E}_{\text{int}}\), \({K}_{Ic}^{\text{int}}\)) are chosen relatively smaller than those of the bulk rock to imply a degree of weakness in the domain. Although, these values are set to be not too smaller than that of the bulk rock (\({E}_{\text{int}}\)=0.5 \({E}_{\text{bulk}}\) and \({K}_{Ic}^{\text{int}}\)=0.5 \({K}_{Ic}^{\text{bulk}}\)). This assumption helps to maintain the continuity of the computational domain and is compatible with the assumptions undertaken by similar studies in this context (Rahman et al. 2009; Wang et al. 2018; Alotaibi et al. 2020; Sun et al. 2022; Zhang et al. 2022).
Table 4
Material properties used for the HF–NF interaction problem
Parameter
Name
Value
unit
\({E}_{\text{bulk}},{E}_{\text{int}}\)
Young’s modulus of rock and the NF
19, 9.5
\(\text{GPa}\)
\(\nu\)
Poisson’s ratio
0.24
\({K}_{Ic}^{\text{bulk}},{K}_{Ic}^{\text{int}}\)
fracture toughness of the rock and NF
2.4, 1.2
\(\text{MPa}\sqrt{\text{m}}\)
\(\overline{k }\)
intrinsic permeability of the rock
0.013
\(\text{mD}\)
\({\sigma }_{t}^{\text{bulk}}\)
tensile strength of the rock
7.7
\(\text{MPa}\)
\({\overline{k} }_{\text{int}}\)
intrinsic permeability of the NF
0.1–100
\(mD\)
\({\varphi }_{\text{int}}\)
internal friction angle of the NF
10
\({c}_{\text{int}}\)
cohesion of the NF
100–1000
\(\text{kPa}\)
\({\sigma }_{t}^{\text{int}}\)
Tensile strength of the cemented NF
0.25–4
\(\text{MPa}\)
\(K\)
Drained bulk modulus of the rock
1
\(\text{GPa}\)
\({\mu }_{\text{f}}\)
Dynamic viscosity of the fluid
0.89E-3
\(\text{Pa}.\text{s}\)
\({l}_{0}\)
Length-scale parameter
5
\(\text{mm}\)
Two major factors that contribute to increasing the chance of crossing have been identified by the experimental study of (Kear et al. 2017). By increasing the differential stress (\({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\)) and the approaching angle (\(\beta \to {90}^{^\circ }\)), the HF will be more expected to cross the NF. To identify the most important controlling factors in the HF–NF interaction problem, we conduct a comprehensive parametric study by observing the interaction outcomes while varying the following factors: (i) the approaching angle \(\beta\), (ii) differential stress (\({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\)), (iii) tensile strength and cohesion of the NF, and (iv) the hydraulic conductivity of the interface (\({\overline{k} }_{\text{int}}\)). We separate the observations of the interaction outcome taken from the numerical analyses into crossing and diversion cases, although the HF may cross the NF and continue propagation in the original direction, albeit with an offset from the initial point of arrest. The offset crossing will be encountered as crossing if the offset length is less than \({L}_{\text{NF}}/4\) and the HF maintains the original direction of propagation after the point of intersection. For the case of diversion, the HF continues to propagate along the NF and the original direction is ignored, since the most favourable path has been changed because of the change in the orientation of the effective principal stresses due to the shear failure of the NF.
For illustrative purposes and to show the crossing and diversion outcome results, the BVP in Fig. 16 is modelled for two different values of \(\beta\) and \({\overline{k} }_{\text{int}}\) while keeping the same stiffness properties for the NF (\({c}_{\text{int}}\)=1000kPa and \({\sigma }_{t}^{\text{int}}\)=4MPa). The interaction outcome is changed from diversion (Fig. 17b) to crossing (Fig. 17a) by reducing the NF permeability (\({\overline{k} }_{\text{int}}\)) from 100 mD (milli-Darcy) to 0.1 mD (\(1 \text{m}D=9.869\times {10}^{-16} {\text{m}}^{2}\)). An increase in the hydraulic conductivity of the NF is not in favour of the crossing outcome. On the other hand, increasing the approaching angle from \({50}^{^\circ }\) to \({60}^{^\circ }\) while keeping the NF permeability equal to \({\overline{k} }_{\text{int}}\)=100 mD can cause the HF to cross the NF, compare Fig. 17b–c. It can still be seen that the damage has evolved along the NF for the case of \(\beta\)=\({60}^{^\circ }\) but the shear slippage has not occurred fully, so the overall direction of the maximum effective principal stresses remains horizontal in favour of the crossing scenario. In what follows, the input parameters are varied extensively to conduct a comprehensive parametric study to identify the most important factors affecting the outcome of the HF–NF interaction problem.

4.1 Approaching Angle and Differential In-Situ Stress

It can be attested from the past studies (Xie et al. 2016; Zhang et al. 2020; Sun et al. 2022) that the approaching angle is the most important factor controlling the interaction outcome, as the crossing outcome usually occurs for nearly orthogonal intersections (Warpinski and Teufel 1987) and diversion is usually the case when the HF approaches a highly inclined NF. According to Fig. 18a, the interaction outcome is always diversion when \(\beta \le 30^\circ\) regardless of the NF’s strength (i.e., \({c}_{\text{int}}\) and \({\sigma }_{\text{t}}^{\text{int}}\)). For intermediate approaching angles (\(40^\circ <\beta <70^\circ\)), increasing the differential stress (\({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\)) would be in favour of the crossing outcome. The interaction outcomes for the case of having high cohesion and high tensile strength of the NF for two different values of NF permeability (\({\overline{k} }_{\text{int}}\)=0.1 mD and \({\overline{k} }_{\text{int}}\)=100 mD) are shown in scatter plots in Fig. 18b, c. Our results are compared to the analytical criterion of (Blanton 1986) on specifying the interaction outcome based on the approaching angle and the differential stress. In this criterion, according to Eq. (1), parameter \(b\) depends on the hydro-mechanical characteristics of the NF, the amount of fluid penetrated the NF. It can be concluded that \(b\to \infty\) if no slippage is expected to happen, in which case the crossing outcome would only occur (Blanton 1986). By increasing the NF permeability (\({\overline{k} }_{\text{int}}\)), a larger area of the NF become pressurised by the fracturing fluid which makes the shear slippage more likely to occur because of reducing the effective resisting shear stresses inside the NF. Therefore, by increasing the NF permeability (\({\overline{k} }_{\text{int}}\)), it would make sense to choose a smaller value for \(b\) to be used in Blanton’s criterion for the sake of comparing our numerical results to this analytical reference. For relatively low values of the differential stress (e.g., 0.5 MPa and 1 MPa), the numbers of the crossing outcomes reduce significantly, especially in the case of having a high-permeable NF. It is also understood that the numerical predictions of crossing/diversion outcome better match the analytical predictions for the higher values of differential stress \({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\). It is evident from Fig. 18b, c that increasing both the approaching angle and the differential stress would help the HF to cross the NF.

4.2 Hydraulic Conductivity (Permeability) of the Embedded NF

The NF permeability (\({\overline{k} }_{\text{int}}\)) is dependent on the hydraulic aperture, the degree of cementation inside the NF, and the confining in-situ stresses (Kear et al. 2017). Having a small hydraulic aperture for the NF is in favour of the HF crossing the NF (Xie et al. 2016), although other factors such as differential stresses and approaching angle have a bigger impact on the interaction outcome. We identify the permeability NF region to be the third important factor affecting the HF–NF interaction outcome. Capturing the crossing outcome becomes more likely by reducing the NF permeability, because the fracturing fluid gets less chance to affect a large area around the NF; therefore, the NF is less vulnerable to shear slippage, which is the main cause for the HF to divert into the NF’s direction. A comparison between Fig. 17a and b confirms that a lower value of \({\overline{k} }_{\text{int}}\) keeps the fluid pressurised region locally around the HF tip, avoiding the diversion. Figure 19 depicts the positive effect of lowering the NF permeability on increasing the possibility of crossing for different values of the approaching angle and the differential stress. Our results are also compatible with the experimental data of (Zhou et al. 2008; Kear et al. 2017) affirming that a low value of the NF’s aperture (NF’s permeability) makes the crossing scenario the dominant mode. The effect of NF permeability appears to play a more significant role when the differential stress is low.

4.3 Tensile and Shear Strength of the Cemented NF

A high degree of cementation in the NF region increases the tensile and shear strength of the NF, which is expected to be in favour of the crossing outcome. In Fig. 20, Crossing/No-Crossing is plotted with respect to different values of the approaching angle and the tensile strength \({\sigma }_{t}^{\text{int}}\) for various cases of the differential stress while keeping a constant value for the NF permeability (\({\overline{k} }_{\text{int}}\)=25 mD). For the case of (\({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\)=2 MPa), increasing the NF strength has no effect on the crossing outcome, and the approaching angle remains the dominant factor. However, the NF’s tensile strength (\({\sigma }_{t}^{\text{int}}\)) comes into play as an important factor as the differential stress increases. The reason of capturing this behaviour can be related to the vulnerability of the NF against shear slippage. The NF tends to slip when touched by the HF, so a lower tensile strength of the cemented NF will worsen the situation and leads to a full diversion of the HF into the NF. Having a mechanically stronger NF is in favour of the crossing outcome in intermediate approaching angles especially when the differential stress (\({\sigma }_{\text{H}}-{\sigma }_{\text{h}}\)) is high.

4.4 The Effect of Fluid Injection Rate

The parametric studies presented in previous sections were all conducted under a constant fluid injection rate of \({Q}_{\text{inj}}\)=200 m3/s. However, the combination of the fluid injection rate and viscosity can affect the HF–NF interaction and create complex crack-patterns in naturally fractured reservoirs (Chuprakov et al. 2014). Here, we repeat the simulations with a different injection rate of \({Q}_{\text{inj}}\)=500 m3/s, and the relevant results are shown in Fig. 21. Increasing the fluid injection rate seems not to be in favour of HF crossing a high-permeable NF. Increasing \({Q}_{\text{inj}}\) only helps the HF crossing when the NF permeability is low enough, so that the NF will not be affected by the shear slippage mechanism. In more permeable and low-cohesive NFs, increasing the fluid injection rate will significantly reduce the possibility of crossing outcome by causing shear slippage along the NF and changing the stress field around the NF especially in intermediate approaching angles (\(40^\circ <\beta <70^\circ\)). As a result, the effect of fluid injection rate on the interaction outcome must be evaluated with respect to other parameters of the NF, such as the degree of cementation, hydraulic aperture, and the tensile and shear strength of the cemented NF.

5 Conclusions

In this study, the phase-field method was employed to model the HF propagation and its interaction with cemented NFs embedded in saturated poroelastic rocks. In studying the interaction between the HF and an embedded NF, one of the most important mechanisms to be considered is the shear slippage inside the NF, which causes the HF to be arrested by or diverted into the NF. For this reason, we first developed a numerical framework to model the HF and shear failure in geo-materials simultaneously as gradient-type diffusive damaged zones using the phase-field method and the well-known Mohr–Coulomb–Griffith failure model. The developed numerical model was verified by comparing the results to the experimental data on the failure of rock-like specimens under compressive shear stresses and to the KGD analytical solution on predicting the HF propagation. One of the main concerns in modelling phase-field fractures as damaged zones is related to the choice of the length-scale parameter in the phase-field method. We found out that the choice of this parameter can significantly affect the hydraulic fracturing characteristics such as the evolution of fluid pressure inside the HF and the crack-width in a way that the results do not match nicely with the analytical KGD solution. In this regard, discrete-based methods for modelling fractures perform better, because the HF is modelled as a discontinuity (Xie et al. 2016; Huang et al. 2023); however, discrete-based methods lack in modelling the fluid exchange between the HF and the surrounding porous medium. Our in-house numerical code for modelling the HF is equipped with a mesh refinement unit which can detect the critical elements in the vicinity of damaged zones to ensure a correct prediction of stress field and fluid flow over the elements. The proposed numerical model is satisfactorily capable of modelling the HF propagation through pre-existing NFs by predicting various interaction scenarios.
Using the developed numerical model, we investigated the most important factors controlling the outcome of the interaction between a propagating HF and a pre-existing cemented NF, which is considered as a region of weakness embedded in the bulk rock. When the HF meets the NF, the NF may fail under a combination of tensile and shear stresses. Based on several factors, such as tensile and shear strengths of the NF, the degree of cementation, and the hydraulic conductivity of the NF region, two general scenarios (i.e., crossing and diversion) were detected for the propagating HF. We conducted an extensive parametric study on the HF–NF interaction by simulating a representative BVP, and the conclusions are listed in the following.
  • The main controlling parameters are identified to be the approaching angle and the differential stress, which is in line with the most analytical, experimental, and numerical studies in the literature relevant to this problem.
  • The hydro-mechanical characteristics of the NF, such as tensile and shear strengths as well as the permeability, play important roles for intermediate approaching angles (\(40^\circ <\beta <70^\circ\)). Crossing and diversion are always the dominant scenarios for very high and very low approaching angles, respectively.
  • The hydraulic conductivity of the NF region (NF permeability) is identified as the third most important factor affecting the interaction outcome, since it controls the extent of the fluid pressurisation of the NF causing the shear slippage. A lower value of the NF permeability generally helps the HF to cross the NF even in low approaching angles (high inclination of the NF with respect to the orientation of the HF).
  • Increasing shear and tensile strengths of the NF region are in favour of the crossing outcome if the differential stress is high enough to mobilise shear strength affected by the normal stress on the plane of NF.
  • The analytical criterion of Blanton (1986) was found to be fitted well to our numerical results in terms of predicting the crossing outcome. The effect of variation of the NF permeability could be understood as it is compatible with Blanton’s assumptions on the possibility of shear slippage along the NF because of the accumulation of the fluid pressure and reducing the resisting shear stresses inside the NF.
  • Increasing the fluid injection rate is in favour of the crossing outcome for low-permeable and highly cemented NFs. For the case of high-permeable NFs, increasing the fluid injection rate worsens the situation for crossing outcome because of the rapid accumulation of the fluid pressure inside the NF, leading to shear slippage within the NF and the HF’s diversion.

6 Areas of Improvement

Natural discontinuities in reservoir rocks are randomly distributed and are found in clusters of joints and fissures; modelling the hydro-mechanical behaviour of rocks containing these features requires the use of other methods, such as the discrete fracture network (Lei et al. 2017). Considering the problem in 2D and modelling a single NF in intact rock (instead of fractured rock) are encountered as the limitations of our work; however, we were mainly focused on identifying the controlling factors such as the in-situ stresses and hydro-mechanical characteristics affecting the crossing behaviour of the HF. Developing this study to 3D analysis would be beneficial in terms of capturing the in-situ stresses more realistically (Adachi et al. 2010). The hydromechanical response of rocks under hydraulic fracturing operations can vary depending on the temperature (Kumari et al. 2018; Zhou et al. 2018; Cheng et al. 2021) and cyclic loadings (Zhao et al. 2018). Hence, developing the poroelastic material to poro-viscoelastic (Shen et al. 2019; Song et al. 2021), considering thermo-hydromechanics (Li et al. 2016), and capturing ductile behaviour of rocks (Choo and Sun 2018) can be counted as the potential areas of research to expand this current study. The versatility of the phase-field method in the finite-element implementation makes it easier for developing the presented HF model to capture the rupture characteristics of the rock under realistic environmental effects.

Acknowledgements

Authors acknowledge the financial support provided by the European Commission-Marie Curie Research Staff Exchange programme (ID 872607). Q.J. Fisher acknowledges funding from NERC for the SHAPE-UK project with the grant number NE/R017565/1.

Declarations

Conflict of Interest

The authors declare that they have no competing financial interests or personal relationships that could have appeared to influence the piece of research presented in this study.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literature
go back to reference Adachi JI, Detournay E, Peirce AP (2010) Analysis of the classical pseudo-3D model for hydraulic fracture with equilibrium height growth across stress barriers. Int J Rock Mech Min Sci 47(4):625–639 Adachi JI, Detournay E, Peirce AP (2010) Analysis of the classical pseudo-3D model for hydraulic fracture with equilibrium height growth across stress barriers. Int J Rock Mech Min Sci 47(4):625–639
go back to reference Alotaibi TE, Landis CM and AlTammar MJ (2020) Phase-field modeling of hydraulic fracture propagation in mechanically heterogeneous formations. International Petroleum Technology Conference. Dhahran, Kingdom of Saudi Arabia, International Petroleum Technology Conference: 15 Alotaibi TE, Landis CM and AlTammar MJ (2020) Phase-field modeling of hydraulic fracture propagation in mechanically heterogeneous formations. International Petroleum Technology Conference. Dhahran, Kingdom of Saudi Arabia, International Petroleum Technology Conference: 15
go back to reference Ambrosio L, Tortorelli VM (1990) Approximation of functional depending on jumps by elliptic functional via t-convergence. Commun Pure Appl Math 43(8):999–1036 Ambrosio L, Tortorelli VM (1990) Approximation of functional depending on jumps by elliptic functional via t-convergence. Commun Pure Appl Math 43(8):999–1036
go back to reference Amor H, Marigo J-J, Maurini C (2009) Regularized formulation of the variational brittle fracture with unilateral contact: numerical experiments. J Mech Phys Solids 57(8):1209–1229 Amor H, Marigo J-J, Maurini C (2009) Regularized formulation of the variational brittle fracture with unilateral contact: numerical experiments. J Mech Phys Solids 57(8):1209–1229
go back to reference Bellettini G, Coscia A (1994) Discrete approximation of a free discontinuity problem. Numer Funct Anal Optim 15(3–4):201–224 Bellettini G, Coscia A (1994) Discrete approximation of a free discontinuity problem. Numer Funct Anal Optim 15(3–4):201–224
go back to reference Biot MA (1941) General theory of three-dimensional consolidation. J Appl Phys 12(2):155–164 Biot MA (1941) General theory of three-dimensional consolidation. J Appl Phys 12(2):155–164
go back to reference Biot MA, Willis D (1957) The elastic coefficients of the theory of consolidation. J Appl Mech 24:594–601 Biot MA, Willis D (1957) The elastic coefficients of the theory of consolidation. J Appl Mech 24:594–601
go back to reference Blanton TL (1986) Propagation of hydraulically and dynamically induced fractures in naturally fractured reservoirs. SPE Unconventional Gas Technology Symposium. Louisville, Kentucky, Society of Petroleum Engineers: 15 Blanton TL (1986) Propagation of hydraulically and dynamically induced fractures in naturally fractured reservoirs. SPE Unconventional Gas Technology Symposium. Louisville, Kentucky, Society of Petroleum Engineers: 15
go back to reference Bourdin B, Francfort GA, Marigo JJ (2000) Numerical experiments in revisited brittle fracture. J Mech Phys Solids 48(4):797–826 Bourdin B, Francfort GA, Marigo JJ (2000) Numerical experiments in revisited brittle fracture. J Mech Phys Solids 48(4):797–826
go back to reference Bourdin B, Chukwudozie CP and Yoshioka K (2012) A variational approach to the numerical simulation of hydraulic fracturing. SPE Annual Technical Conference and Exhibition. San Antonio, Texas, USA, Society of Petroleum Engineers: 9 Bourdin B, Chukwudozie CP and Yoshioka K (2012) A variational approach to the numerical simulation of hydraulic fracturing. SPE Annual Technical Conference and Exhibition. San Antonio, Texas, USA, Society of Petroleum Engineers: 9
go back to reference Cheng Y, Zhang Y, Yu Z, Hu Z, Ma Y, Yang Y (2021) Experimental and numerical studies on hydraulic fracturing characteristics with different injection flow rates in granite geothermal reservoir. Energy Sci Eng 9(1):142–168 Cheng Y, Zhang Y, Yu Z, Hu Z, Ma Y, Yang Y (2021) Experimental and numerical studies on hydraulic fracturing characteristics with different injection flow rates in granite geothermal reservoir. Energy Sci Eng 9(1):142–168
go back to reference Chong KP, Kuruppu MD (1984) New specimen for fracture toughness determination for rock and other materials. Int J Fract 26(2):R59–R62 Chong KP, Kuruppu MD (1984) New specimen for fracture toughness determination for rock and other materials. Int J Fract 26(2):R59–R62
go back to reference Choo J, Sun WC (2018) Coupled phase-field and plasticity modeling of geological materials: from brittle fracture to ductile flow. Comput Methods Appl Mech Eng 330:1–32 Choo J, Sun WC (2018) Coupled phase-field and plasticity modeling of geological materials: from brittle fracture to ductile flow. Comput Methods Appl Mech Eng 330:1–32
go back to reference Chuprakov D, Melchaeva O, Prioul R (2014) Injection-sensitive mechanics of hydraulic fracture interaction with discontinuities. Rock Mech Rock Eng 47(5):1625–1640 Chuprakov D, Melchaeva O, Prioul R (2014) Injection-sensitive mechanics of hydraulic fracture interaction with discontinuities. Rock Mech Rock Eng 47(5):1625–1640
go back to reference Cooke ML, Underwood CA (2001) Fracture termination and step-over at bedding interfaces due to frictional slip and interface opening. J Struct Geol 23(2):223–238 Cooke ML, Underwood CA (2001) Fracture termination and step-over at bedding interfaces due to frictional slip and interface opening. J Struct Geol 23(2):223–238
go back to reference Dahi Taleghani A, Gonzalez M, Shojaei A (2016) Overview of numerical models for interactions between hydraulic fractures and natural fractures: challenges and limitations. Comput Geotech 71:361–368 Dahi Taleghani A, Gonzalez M, Shojaei A (2016) Overview of numerical models for interactions between hydraulic fractures and natural fractures: challenges and limitations. Comput Geotech 71:361–368
go back to reference Taleghani AD (2009) Analysis of hydraulic fracture propagation in fractured reservoirs: an improved model for the interaction between induced and natural fractures. The University of Texas at Austin, Chicago Taleghani AD (2009) Analysis of hydraulic fracture propagation in fractured reservoirs: an improved model for the interaction between induced and natural fractures. The University of Texas at Austin, Chicago
go back to reference Detournay E, Cheng AHD (1993) 5 - Fundamentals of Poroelasticity. Analysis and design methods. C. Fairhurst. Oxford, Pergamon, pp 113–171 Detournay E, Cheng AHD (1993) 5 - Fundamentals of Poroelasticity. Analysis and design methods. C. Fairhurst. Oxford, Pergamon, pp 113–171
go back to reference Dyskin AV, Caballero A (2009) Orthogonal crack approaching an interface. Eng Fract Mech 76(16):2476–2485 Dyskin AV, Caballero A (2009) Orthogonal crack approaching an interface. Eng Fract Mech 76(16):2476–2485
go back to reference Fei F, Choo J (2020) A phase-field model of frictional shear fracture in geologic materials. Comput Methods Appl Mech Eng 369:113265 Fei F, Choo J (2020) A phase-field model of frictional shear fracture in geologic materials. Comput Methods Appl Mech Eng 369:113265
go back to reference Ferrill DA, McGinnis RN, Morris AP, Smart KJ (2012) Hybrid failure: Field evidence and influence on fault refraction. J Struct Geol 42:140–150 Ferrill DA, McGinnis RN, Morris AP, Smart KJ (2012) Hybrid failure: Field evidence and influence on fault refraction. J Struct Geol 42:140–150
go back to reference Fisher QJ, Knipe RJ (2001) The permeability of faults within siliciclastic petroleum reservoirs of the North Sea and Norwegian Continental Shelf. Mar Pet Geol 18(10):1063–1081 Fisher QJ, Knipe RJ (2001) The permeability of faults within siliciclastic petroleum reservoirs of the North Sea and Norwegian Continental Shelf. Mar Pet Geol 18(10):1063–1081
go back to reference Franquet JA and Abass HH (1999) Experimental evaluation of Biot's poroelastic parameter: three different methods. Vail Rocks 1999, The 37th U.S. Symposium on Rock Mechanics (USRMS) Franquet JA and Abass HH (1999) Experimental evaluation of Biot's poroelastic parameter: three different methods. Vail Rocks 1999, The 37th U.S. Symposium on Rock Mechanics (USRMS)
go back to reference Gale JFW, Laubach SE, Olson JE, Eichhubl P, Fall A (2014) Natural fractures in shale: a review and new observations. AAPG Bull 98(11):2165–2216 Gale JFW, Laubach SE, Olson JE, Eichhubl P, Fall A (2014) Natural fractures in shale: a review and new observations. AAPG Bull 98(11):2165–2216
go back to reference Geertsma J, De Klerk F (1969) A rapid method of predicting width and extent of hydraulically induced fractures. J Petrol Technol 21(12):1571–1581 Geertsma J, De Klerk F (1969) A rapid method of predicting width and extent of hydraulically induced fractures. J Petrol Technol 21(12):1571–1581
go back to reference Gironacci E, Mousavi Nezhad M, Rezania M, Lancioni G (2018) A non-local probabilistic method for modeling of crack propagation. Int J Mech Sci 144:897–908 Gironacci E, Mousavi Nezhad M, Rezania M, Lancioni G (2018) A non-local probabilistic method for modeling of crack propagation. Int J Mech Sci 144:897–908
go back to reference Griffith AA (1921) VI. The phenomena of rupture and flow in solids. Philos Trans R Soc Lond 221(582–593):163–198 Griffith AA (1921) VI. The phenomena of rupture and flow in solids. Philos Trans R Soc Lond 221(582–593):163–198
go back to reference Hansen-Dörr AC, de Borst R, Hennig P, Kästner M (2019) Phase-field modelling of interface failure in brittle materials. Comput Methods Appl Mech Eng 346:25–42 Hansen-Dörr AC, de Borst R, Hennig P, Kästner M (2019) Phase-field modelling of interface failure in brittle materials. Comput Methods Appl Mech Eng 346:25–42
go back to reference Hansen-Dörr AC, Dammaß F, de Borst R, Kästner M (2020) Phase-field modeling of crack branching and deflection in heterogeneous media. Eng Fract Mech 232:107004 Hansen-Dörr AC, Dammaß F, de Borst R, Kästner M (2020) Phase-field modeling of crack branching and deflection in heterogeneous media. Eng Fract Mech 232:107004
go back to reference Heider Y, Markert B (2017) A phase-field modeling approach of hydraulic fracture in saturated porous media. Mech Res Commun 80:38–46 Heider Y, Markert B (2017) A phase-field modeling approach of hydraulic fracture in saturated porous media. Mech Res Commun 80:38–46
go back to reference Ingraffea AR, Heuze FE (1980) Finite element models for rock fracture mechanics. Int J Numer Anal Meth Geomech 4(1):25–43 Ingraffea AR, Heuze FE (1980) Finite element models for rock fracture mechanics. Int J Numer Anal Meth Geomech 4(1):25–43
go back to reference Jaeger JC, Cook NG, Zimmerman R (2009) Fundamentals of rock mechanics. John Wiley & Sons Jaeger JC, Cook NG, Zimmerman R (2009) Fundamentals of rock mechanics. John Wiley & Sons
go back to reference Jeffrey RG, Zhang X and Thiercelin MJ (2009) Hydraulic fracture offsetting in naturally fractures reservoirs: quantifying a long-recognized process. SPE Hydraulic Fracturing Technology Conference Jeffrey RG, Zhang X and Thiercelin MJ (2009) Hydraulic fracture offsetting in naturally fractures reservoirs: quantifying a long-recognized process. SPE Hydraulic Fracturing Technology Conference
go back to reference Jin J, Cao P, Chen Y, Pu C, Mao D, Fan X (2017) Influence of single flaw on the failure process and energy mechanics of rock-like material. Comput Geotech 86:150–162 Jin J, Cao P, Chen Y, Pu C, Mao D, Fan X (2017) Influence of single flaw on the failure process and energy mechanics of rock-like material. Comput Geotech 86:150–162
go back to reference Kear J, Kasperczyk D, Zhang X, Jeffrey RG, Chuprakov D and Prioul R (2017) 2D Experimental and numerical results for hydraulic fractures interacting with orthogonal and inclined discontinuities. 51st U.S. Rock Mechanics/Geomechanics Symposium. San Francisco, California, USA, American Rock Mechanics Association: 12 Kear J, Kasperczyk D, Zhang X, Jeffrey RG, Chuprakov D and Prioul R (2017) 2D Experimental and numerical results for hydraulic fractures interacting with orthogonal and inclined discontinuities. 51st U.S. Rock Mechanics/Geomechanics Symposium. San Francisco, California, USA, American Rock Mechanics Association: 12
go back to reference Khoei AR, Vahab M, Hirmand M (2018) An enriched–FEM technique for numerical simulation of interacting discontinuities in naturally fractured porous media. Comput Methods Appl Mech Eng 331:197–231 Khoei AR, Vahab M, Hirmand M (2018) An enriched–FEM technique for numerical simulation of interacting discontinuities in naturally fractured porous media. Comput Methods Appl Mech Eng 331:197–231
go back to reference Kumari WGP, Ranjith PG, Perera MSA, Li X, Li LH, Chen BK, De Silva VRS (2018) Hydraulic fracturing under high temperature and pressure conditions with micro CT applications: geothermal energy from hot dry rocks. Fuel 230:138–154 Kumari WGP, Ranjith PG, Perera MSA, Li X, Li LH, Chen BK, De Silva VRS (2018) Hydraulic fracturing under high temperature and pressure conditions with micro CT applications: geothermal energy from hot dry rocks. Fuel 230:138–154
go back to reference Kuruppu MD, Obara Y, Ayatollahi MR, Chong KP, Funatsu T (2014) ISRM-suggested method for determining the mode I static fracture toughness using semi-circular bend specimen. Rock Mech Rock Eng 47(1):267–274 Kuruppu MD, Obara Y, Ayatollahi MR, Chong KP, Funatsu T (2014) ISRM-suggested method for determining the mode I static fracture toughness using semi-circular bend specimen. Rock Mech Rock Eng 47(1):267–274
go back to reference Labuz JF, Zang A (2012) Mohr–Coulomb failure criterion. Rock Mech Rock Eng 45(6):975–979 Labuz JF, Zang A (2012) Mohr–Coulomb failure criterion. Rock Mech Rock Eng 45(6):975–979
go back to reference Lei Q, Latham JP, Tsang CF (2017) The use of discrete fracture networks for modelling coupled geomechanical and hydrological behaviour of fractured rocks. Comput Geotech 85:151–176 Lei Q, Latham JP, Tsang CF (2017) The use of discrete fracture networks for modelling coupled geomechanical and hydrological behaviour of fractured rocks. Comput Geotech 85:151–176
go back to reference Lepillier B, Yoshioka K, Parisio F, Bakker R, Bruhn D (2020) Variational phase-field modeling of hydraulic fracture interaction with natural fractures and application to enhanced geothermal systems. J Geophys Res 125(7):e2020JB019856 Lepillier B, Yoshioka K, Parisio F, Bakker R, Bruhn D (2020) Variational phase-field modeling of hydraulic fracture interaction with natural fractures and application to enhanced geothermal systems. J Geophys Res 125(7):e2020JB019856
go back to reference Li S, Li X, Zhang D (2016) A fully coupled thermo-hydro-mechanical, three-dimensional model for hydraulic stimulation treatments. J Nat Gas Sci Eng 34:64–84 Li S, Li X, Zhang D (2016) A fully coupled thermo-hydro-mechanical, three-dimensional model for hydraulic stimulation treatments. J Nat Gas Sci Eng 34:64–84
go back to reference Lim IL, Johnston IW, Choi SK, Boland JN (1994) Fracture testing of a soft rock with semi-circular specimens under three-point bending. Part 1—mode I. Int J Rock Mech Min Sci Geomech Abstracts 31(3):185–197 Lim IL, Johnston IW, Choi SK, Boland JN (1994) Fracture testing of a soft rock with semi-circular specimens under three-point bending. Part 1—mode I. Int J Rock Mech Min Sci Geomech Abstracts 31(3):185–197
go back to reference Marigo J-J, Maurini C, Pham K (2016) An overview of the modelling of fracture by gradient damage models. Meccanica 51(12):3107–3128 Marigo J-J, Maurini C, Pham K (2016) An overview of the modelling of fracture by gradient damage models. Meccanica 51(12):3107–3128
go back to reference Mauthe S, Miehe C (2017) Hydraulic fracture in poro-hydro-elastic media. Mech Res Commun 80:69–83 Mauthe S, Miehe C (2017) Hydraulic fracture in poro-hydro-elastic media. Mech Res Commun 80:69–83
go back to reference Michael A, Olson JE and Balhoff MT (2018) Analysis of hydraulic fracture initiation from perforated horizontal wellbores. ARMA US Rock Mechanics/Geomechanics Symposium. ARMA Michael A, Olson JE and Balhoff MT (2018) Analysis of hydraulic fracture initiation from perforated horizontal wellbores. ARMA US Rock Mechanics/Geomechanics Symposium. ARMA
go back to reference Miehe C, Mauthe S (2016) Phase field modeling of fracture in multi-physics problems. Part III. Crack driving forces in hydro-poro-elasticity and hydraulic fracturing of fluid-saturated porous media. Comput Methods Appl Mech Eng 304:619–655 Miehe C, Mauthe S (2016) Phase field modeling of fracture in multi-physics problems. Part III. Crack driving forces in hydro-poro-elasticity and hydraulic fracturing of fluid-saturated porous media. Comput Methods Appl Mech Eng 304:619–655
go back to reference Miehe C, Hofacker M, Welschinger F (2010a) A phase field model for rate-independent crack propagation: robust algorithmic implementation based on operator splits. Comput Methods Appl Mech Eng 199(45):2765–2778 Miehe C, Hofacker M, Welschinger F (2010a) A phase field model for rate-independent crack propagation: robust algorithmic implementation based on operator splits. Comput Methods Appl Mech Eng 199(45):2765–2778
go back to reference Miehe C, Welschinger F, Hofacker M (2010b) Thermodynamically consistent phase-field models of fracture: variational principles and multi-field FE implementations. Int J Numer Meth Eng 83(10):1273–1311 Miehe C, Welschinger F, Hofacker M (2010b) Thermodynamically consistent phase-field models of fracture: variational principles and multi-field FE implementations. Int J Numer Meth Eng 83(10):1273–1311
go back to reference Miehe C, Schänzel L-M, Ulmer H (2015) Phase field modeling of fracture in multi-physics problems. Part I. Balance of crack surface and failure criteria for brittle crack propagation in thermo-elastic solids. Comput Methods Appl Mech Eng 294:449–485 Miehe C, Schänzel L-M, Ulmer H (2015) Phase field modeling of fracture in multi-physics problems. Part I. Balance of crack surface and failure criteria for brittle crack propagation in thermo-elastic solids. Comput Methods Appl Mech Eng 294:449–485
go back to reference Mikelić A, Wheeler MF, Wick T (2015) A phase-field method for propagating fluid-filled fractures coupled to a surrounding porous medium. Multiscale Model Simul 13(1):367–398 Mikelić A, Wheeler MF, Wick T (2015) A phase-field method for propagating fluid-filled fractures coupled to a surrounding porous medium. Multiscale Model Simul 13(1):367–398
go back to reference Mousavi Nezhad M, Javadi AA, Rezania M (2011) Modeling of contaminant transport in soils considering the effects of micro- and macro-heterogeneity. J Hydrol 404(3):332–338 Mousavi Nezhad M, Javadi AA, Rezania M (2011) Modeling of contaminant transport in soils considering the effects of micro- and macro-heterogeneity. J Hydrol 404(3):332–338
go back to reference Mousavi Nezhad M, Fisher QJ, Gironacci E, Rezania M (2018a) Experimental study and numerical modeling of fracture propagation in shale rocks during Brazilian disk test. Rock Mech Rock Eng 51(6):1755–1775 Mousavi Nezhad M, Fisher QJ, Gironacci E, Rezania M (2018a) Experimental study and numerical modeling of fracture propagation in shale rocks during Brazilian disk test. Rock Mech Rock Eng 51(6):1755–1775
go back to reference Mousavi Nezhad M, Gironacci E, Rezania M, Khalili N (2018b) Stochastic modelling of crack propagation in materials with random properties using isometric mapping for dimensionality reduction of nonlinear data sets. Int J Numer Meth Eng 113(4):656–680 Mousavi Nezhad M, Gironacci E, Rezania M, Khalili N (2018b) Stochastic modelling of crack propagation in materials with random properties using isometric mapping for dimensionality reduction of nonlinear data sets. Int J Numer Meth Eng 113(4):656–680
go back to reference Nezhad MM, Javadi AA (2011) Stochastic finite-element approach to quantify and reduce uncertainty in pollutant transport modeling. J Hazardous Toxic Radioact Waste 15(3):208–215 Nezhad MM, Javadi AA (2011) Stochastic finite-element approach to quantify and reduce uncertainty in pollutant transport modeling. J Hazardous Toxic Radioact Waste 15(3):208–215
go back to reference Nguyen TT, Yvonnet J, Zhu QZ, Bornert M, Chateau C (2016) A phase-field method for computational modeling of interfacial damage interacting with crack propagation in realistic microstructures obtained by microtomography. Comput Methods Appl Mech Eng 312:567–595 Nguyen TT, Yvonnet J, Zhu QZ, Bornert M, Chateau C (2016) A phase-field method for computational modeling of interfacial damage interacting with crack propagation in realistic microstructures obtained by microtomography. Comput Methods Appl Mech Eng 312:567–595
go back to reference Paggi M, Reinoso J (2017) Revisiting the problem of a crack impinging on an interface: a modeling framework for the interaction between the phase field approach for brittle fracture and the interface cohesive zone model. Comput Methods Appl Mech Eng 321:145–172 Paggi M, Reinoso J (2017) Revisiting the problem of a crack impinging on an interface: a modeling framework for the interaction between the phase field approach for brittle fracture and the interface cohesive zone model. Comput Methods Appl Mech Eng 321:145–172
go back to reference Pine RJ, Batchelor AS (1984) Downward migration of shearing in jointed rock during hydraulic injections. Int J Rock Mech Min Sci Geomech Abstracts 21:249–263 Pine RJ, Batchelor AS (1984) Downward migration of shearing in jointed rock during hydraulic injections. Int J Rock Mech Min Sci Geomech Abstracts 21:249–263
go back to reference Rahman MM, Aghighi MA, Rahman SS and Ravoof SA (2009) Interaction between induced hydraulic fracture and pre-existing natural fracture in a poro-elastic environment: effect of pore pressure change and the orientation of natural fractures. Asia Pacific Oil and Gas Conference & Exhibition. Jakarta, Indonesia, Society of Petroleum Engineers: 8 Rahman MM, Aghighi MA, Rahman SS and Ravoof SA (2009) Interaction between induced hydraulic fracture and pre-existing natural fracture in a poro-elastic environment: effect of pore pressure change and the orientation of natural fractures. Asia Pacific Oil and Gas Conference & Exhibition. Jakarta, Indonesia, Society of Petroleum Engineers: 8
go back to reference Ramsey JM, Chester FM (2004) Hybrid fracture and the transition from extension fracture to shear fracture. Nature 428(6978):63–66 Ramsey JM, Chester FM (2004) Hybrid fracture and the transition from extension fracture to shear fracture. Nature 428(6978):63–66
go back to reference Renshaw CE, Pollard DD (1995) An experimentally verified criterion for propagation across unbounded frictional interfaces in brittle, linear elastic materials. Int J Rock Mech Min Sci Geomech Abstracts 32(3):237–249 Renshaw CE, Pollard DD (1995) An experimentally verified criterion for propagation across unbounded frictional interfaces in brittle, linear elastic materials. Int J Rock Mech Min Sci Geomech Abstracts 32(3):237–249
go back to reference Sarmadi N, Mousavi Nezhad M (2023) Phase-field modelling of fluid driven fracture propagation in poroelastic materials considering the impact of inertial flow within the fractures. Int J Rock Mech Min Sci 169:105444 Sarmadi N, Mousavi Nezhad M (2023) Phase-field modelling of fluid driven fracture propagation in poroelastic materials considering the impact of inertial flow within the fractures. Int J Rock Mech Min Sci 169:105444
go back to reference Shen R, Waisman H, Guo L (2019) Fracture of viscoelastic solids modeled with a modified phase field method. Comput Methods Appl Mech Eng 346:862–890 Shen R, Waisman H, Guo L (2019) Fracture of viscoelastic solids modeled with a modified phase field method. Comput Methods Appl Mech Eng 346:862–890
go back to reference Shi F, Wang X, Liu C, Liu H, Wu H (2017) An XFEM-based method with reduction technique for modeling hydraulic fracture propagation in formations containing frictional natural fractures. Eng Fract Mech 173:64–90 Shi F, Wang X, Liu C, Liu H, Wu H (2017) An XFEM-based method with reduction technique for modeling hydraulic fracture propagation in formations containing frictional natural fractures. Eng Fract Mech 173:64–90
go back to reference Song H, Liang Z, Chen Z, Rahman SS (2021) Numerical modelling of hydraulic fracture propagation in poro-viscoelastic formation. J Petrol Sci Eng 196:107640 Song H, Liang Z, Chen Z, Rahman SS (2021) Numerical modelling of hydraulic fracture propagation in poro-viscoelastic formation. J Petrol Sci Eng 196:107640
go back to reference Sun T, Zeng Q, Xing H (2022) A quantitative model to predict hydraulic fracture propagating across cemented natural fracture. J Petrol Sci Eng 208:109595 Sun T, Zeng Q, Xing H (2022) A quantitative model to predict hydraulic fracture propagating across cemented natural fracture. J Petrol Sci Eng 208:109595
go back to reference Verhoosel CV, de Borst R (2013) A phase-field model for cohesive fracture. Int J Numer Meth Eng 96(1):43–62 Verhoosel CV, de Borst R (2013) A phase-field model for cohesive fracture. Int J Numer Meth Eng 96(1):43–62
go back to reference Wang W, Olson JE, Prodanović M, Schultz RA (2018) Interaction between cemented natural fractures and hydraulic fractures assessed by experiments and numerical simulations. J Petrol Sci Eng 167:506–516 Wang W, Olson JE, Prodanović M, Schultz RA (2018) Interaction between cemented natural fractures and hydraulic fractures assessed by experiments and numerical simulations. J Petrol Sci Eng 167:506–516
go back to reference Warpinski NR, Teufel LW (1987) Influence of Geologic Discontinuities on Hydraulic Fracture Propagation (includes associated papers 17011 and 17074). J Petrol Technol 39(02):209–220 Warpinski NR, Teufel LW (1987) Influence of Geologic Discontinuities on Hydraulic Fracture Propagation (includes associated papers 17011 and 17074). J Petrol Technol 39(02):209–220
go back to reference Warpinski NR, Lorenz JC, Branagan PT, Myal FR, Gall BL (1993) Examination of a cored hydraulic fracture in a deep gas well. SPE Prod Facil 8(03):150–158 Warpinski NR, Lorenz JC, Branagan PT, Myal FR, Gall BL (1993) Examination of a cored hydraulic fracture in a deep gas well. SPE Prod Facil 8(03):150–158
go back to reference Wilson ZA, Landis CM (2016) Phase-field modeling of hydraulic fracture. J Mech Phys Solids 96:264–290 Wilson ZA, Landis CM (2016) Phase-field modeling of hydraulic fracture. J Mech Phys Solids 96:264–290
go back to reference Xie L, Min K-B, Shen B (2016) Simulation of hydraulic fracturing and its interactions with a pre-existing fracture using displacement discontinuity method. J Nat Gas Sci Eng 36:1284–1294 Xie L, Min K-B, Shen B (2016) Simulation of hydraulic fracturing and its interactions with a pre-existing fracture using displacement discontinuity method. J Nat Gas Sci Eng 36:1284–1294
go back to reference Yi L-P, Li X-G, Yang Z-Z, Yang C-X (2020) Phase field modeling of hydraulic fracturing in porous media formation with natural fracture. Eng Fract Mech 236:107206 Yi L-P, Li X-G, Yang Z-Z, Yang C-X (2020) Phase field modeling of hydraulic fracturing in porous media formation with natural fracture. Eng Fract Mech 236:107206
go back to reference Yoshioka K, Mollaali M, Kolditz O (2021) Variational phase-field fracture modeling with interfaces. Comput Methods Appl Mech Eng 384:113951 Yoshioka K, Mollaali M, Kolditz O (2021) Variational phase-field fracture modeling with interfaces. Comput Methods Appl Mech Eng 384:113951
go back to reference Zhang X, Sloan SW, Vignes C, Sheng D (2017) A modification of the phase-field model for mixed mode crack propagation in rock-like materials. Comput Methods Appl Mech Eng 322:123–136 Zhang X, Sloan SW, Vignes C, Sheng D (2017) A modification of the phase-field model for mixed mode crack propagation in rock-like materials. Comput Methods Appl Mech Eng 322:123–136
go back to reference Zhang Q, Zhang X-P, Zhang H, Ji P-Q, Wu S, Peng J (2020) Study of interaction mechanisms between hydraulic fracture and weak plane with different strengths and widths using the bonded-particle model based on moment tensors. Eng Fract Mech 225:106813 Zhang Q, Zhang X-P, Zhang H, Ji P-Q, Wu S, Peng J (2020) Study of interaction mechanisms between hydraulic fracture and weak plane with different strengths and widths using the bonded-particle model based on moment tensors. Eng Fract Mech 225:106813
go back to reference Zhang J, Yu H, Wang Q, Lv C, Liu C, Shi F, Wu H (2022) Hydraulic fracture propagation at weak interfaces between contrasting layers in shale using XFEM with energy-based criterion. J Nat Gas Sci Eng 101:104502 Zhang J, Yu H, Wang Q, Lv C, Liu C, Shi F, Wu H (2022) Hydraulic fracture propagation at weak interfaces between contrasting layers in shale using XFEM with energy-based criterion. J Nat Gas Sci Eng 101:104502
go back to reference Zheltov AK (1955) 3. Formation of vertical fractures by means of highly viscous liquid. 4th world petroleum congress. World Petroleum Congress Zheltov AK (1955) 3. Formation of vertical fractures by means of highly viscous liquid. 4th world petroleum congress. World Petroleum Congress
go back to reference Zhou J, Chen M, Jin Y, Zhang G-Q (2008) Analysis of fracture propagation behavior and fracture geometry using a tri-axial fracturing system in naturally fractured reservoirs. Int J Rock Mech Min Sci 45(7):1143–1152 Zhou J, Chen M, Jin Y, Zhang G-Q (2008) Analysis of fracture propagation behavior and fracture geometry using a tri-axial fracturing system in naturally fractured reservoirs. Int J Rock Mech Min Sci 45(7):1143–1152
go back to reference Zhou C, Wan Z, Zhang Y, Gu B (2018) Experimental study on hydraulic fracturing of granite under thermal shock. Geothermics 71:146–155 Zhou C, Wan Z, Zhang Y, Gu B (2018) Experimental study on hydraulic fracturing of granite under thermal shock. Geothermics 71:146–155
go back to reference Zhou S, Zhuang X, Rabczuk T (2019) Phase field modeling of brittle compressive-shear fractures in rock-like materials: A new driving force and a hybrid formulation. Comput Methods Appl Mech Eng 355:729–752 Zhou S, Zhuang X, Rabczuk T (2019) Phase field modeling of brittle compressive-shear fractures in rock-like materials: A new driving force and a hybrid formulation. Comput Methods Appl Mech Eng 355:729–752
Metadata
Title
2D Phase-Field Modelling of Hydraulic Fracturing Affected by Cemented Natural Fractures Embedded in Saturated Poroelastic Rocks
Authors
Nima Sarmadi
Mohaddeseh Mousavi Nezhad
Quentin J. Fisher
Publication date
18-01-2024
Publisher
Springer Vienna
Published in
Rock Mechanics and Rock Engineering / Issue 4/2024
Print ISSN: 0723-2632
Electronic ISSN: 1434-453X
DOI
https://doi.org/10.1007/s00603-023-03621-8

Other articles of this Issue 4/2024

Rock Mechanics and Rock Engineering 4/2024 Go to the issue