2.1.2. Softening and Hardening Models
This method simulates the weakening (softening) or strengthening (hardening) behavior of rock masses by using plasticity theories as they experience deformation. In this approach, the nonlinear transition from peak to residual strength is simulated, a feature that is captured by neither elastoplastic nor perfectly plastic models.
Hardening Models
Hardening defines how a material’s resistance and strength against deformation increases by plastic deformation. In other words, this phenomenon describes how the yield stress of a material changes with plastic deformation. The model is categorized into four parts: isotropic hardening, kinematic hardening, mixed hardening, and anisotropic hardening, illustrated below in depth.
Isotropic hardening assumes the material expands uniformly. In this regard, in isotropic hardening models, during the plastic deformation, the yield surface expands uniformly in the stress space shown in Equation (3), where
is the current yield stress,
is the initial yield stress,
is the isotropic hardening modulus, and
is the equivalent plastic strain [
4].
Kinematic hardening is another type of hardening model that translates the yield surface in the stress space without changing its size or shape. Equation (4) demonstrates the translation of yield surface to stress space where
is the back stress tensor,
is the kinematic hardening modulus, and
is the plastic strain tensor [
1].
Additionally, the mix hardening model combines isotropic and kinematic hardening models to account for both expansion and translation of the yield surface and for both expansion and translation of the yield surface into stress space. Finally, the anisotropic hardening model is introduced, which can be described by various models depending on the structure of the material and the loading conditions [
14].
Softening Models
Softening rules, on the other hand, refer to the reduction in material strength and stiffness after a peak stress has been reached, resulting in an increased tendency of failure. In this regard, softening is a critical concept in understanding a possible failure of the material. This concept is divided into three categories: strain softening, damage softening, and viscoplastic softening. Strain softening characterizes the initiation and propagation of cracks or other localized defects, while stress decreases by incremental strain after a material has been yielded and deformed plastically, as illustrated in
Figure 1. This figure shows the stress–strain relationships for the rock samples tested, showcasing the critical points of hardening and softening during the loading process.
The basic softening model consists of current stress
, peak stress
,
as the softening modulus, current strain as
, and
as the current strain, shown in Equation (5) [
6].
Additionally, damage softening involves progressive deterioration of material under load. Therefore, it is often modeled using damage mechanics. Damage softening models incorporate an accumulation of microstructural defects such as voids or microcracks, leading to macroscopic softening behavior.
The last definition is the viscoplastic models, which describe the rate-dependent behavior of materials that exhibit both viscous and plastic deformations, such as polymers or certain metals, at high temperatures.
Furthermore, the SDR method developed by Kalos et al. [
5] illustrates this category by focusing on reduction strength due to accumulated strain, offering insights into long-term stability prediction.
Strain-Controlled Strength Degradation
In 2017, Kalos et al. developed the SDR, which represents an advancement in constitutive modeling by using a continuum mechanics incremental plasticity rate-independent constitutive model designed to account for structure and strength degradation due to the accumulation of plastic shear strain [
5].
In [
5], the SDR is built upon a generalized Hoek–Brown structure envelope in a stress environment characterized by a pressure-dependent curved failure envelope. This approach (SDR model) addresses limitations in the classical plasticity model that predict a large linear elastic domain up to the peak strength, perfect plastic strength without hardening or softening, and non-zero dilatancy at large strains.
The SDR model surpasses classical plasticity models by incorporating:
- (1)
Nonlinear and irreversible stress paths.
- (2)
Non-associated flow rule.
- (3)
Structure degradation mechanism.
The nonlinear and irreversible stress path within the structure envelope (SE) simulates the nonlinear behavior of the weak rock mass up to the peak strength and under stress reversal conditions. In addition, it incorporates the non-associated flow rule in which the direction of plastic strain deformation does not match the direction of the stress vector causing the deformation. In this regard, this method was used to better control dilatancy especially at larger shear strains. The last method is the structure degradation mechanism to predict post-peak strain softening for brittle rock masses or strain hardening for ductile rock masses [
5].
While the SDR model primarily is developed for weak rock masses, it can be used for simulating hard rocks by calibrating its parameters to reflect higher strength and stiffness characteristics typical for hard rock formations. This adaptability is centered on adjusting variables such as yield strength, stiffness reduction, and structural degradation coefficients that accurately simulate hard rock behavior [
12]. Therefore, the SDR model is flexible for a range of rocks upon suitable parameter calibration and validation, making it more reliable for predictions.
Stress Path and Structure Envelope (SE)
The SDR model has differentiated between a “plastic” path that alters the material structure and a “non-plastic” path, which is nonlinear, but does not change the structure of the material. Moreover, the SE is defined as a stress space that includes plastic stresses on the envelope, leading to changes in structure. In contrast, non-plastic stress paths are considered inside the SE leading to nonlinear behavior without altering structure. The fundamental equation defining the structure envelope (SE) of the SDR model is crucial to understanding the constitutive model of rock mass simulation under different loading conditions, is defined in Equation (6). The model is a Hoek–Brown (HB) curve open surface stress in a stress hyperspace consisting of the isotropic axis
and a five-dimensional deviatoric hyperplane. The key variables include the stress component along the
x-axis
, the Hoek–Brown constant for rock masses
, cohesion
, and the parameters
, which adjust according to rock condition and typically represent a scaling factor that represents the yield criterion for different rock masses. Additionally, uniaxial compressive strength
and F
represent the failure condition function, which is equal to zero in the failure condition considered [
5].
The SDR model allows for the gradual evolution of the structure parameters from their initial to final values with accumulated plastic shear strain, modeling strain hardening or softening depending on the evolution “friction angle” and the reduction in tensile strength with plastic strain, the evolution of which is “cohesion”. In this regard, the model includes hardening rules to control the evolution of structure variables with the incremental plastic strain. These rules can be applied to the hardening of the structure variable
, which affects the shape of the axial curve and shear strain, as well as cohesion
. When cohesion reaches its final value, then the material constant is equal to the accumulated plastic shear strain. Kalos et al. depicts the softening branches’ dependence on the plastic shear strain and demonstrates the model simulating the degradation of rock mass strength with increased plastic deformation, highlighting the importance of considering accumulated strain strength degradation due to accumulated plastic shear strain in material response [
5]. In addition, the same paper shows how the structure variable
can affect the shape of the shear strain and axial strain curve. Consequently, this effect demonstrates the model’s ability to account for different material behavior, including strain hardening and softening, by adjusting the
variable.
Following this formula, the authors described other variables to capture the deformation of rock masses under different loading conditions. They described how the stiffness reduction variables evolve with the stress state inside the SE. The stiffness reduction variables include
and
, which control the nonlinearity and irreversibility of the material as a response to different loading conditions, and
and
represent the final or steady-state value of internal variables
and
after significant plastic deformation.
and
are initial values of internal variables representing the state of material before significant plastic deformation.
controls the variables
and
and evolves as the material undergoes plastic deformation.
is the reference stress function used to normalize the stress function
in Equations (7) and (8).
This analysis explores the efficacy of a model in predicting stress behaviors within a structural element by contrasting its outputs under nonlinear and linear scenarios. It examines the model’s response to variations in stress related to volume and shape changes, termed volumetric and deviatoric stiffness, respectively, particularly under conditions where the material does not revert to its original form (non-plastic paths). Constants θ and ξ are employed within the model to maintain consistency in simulation parameters. This comparative approach demonstrates the model’s capability to accurately forecast material behaviors under varying stress conditions.
Hardening Rules
The model decomposes total strain into their plastic and elastic components to describe the relationship between total elastic and plastic strain increments, which is foundational for understanding how plastic deformation accumulates in rock mass. Equation (9) shows the basic of separation of the total strain increment (
) into its elastic (
) and plastic increment (
), which can be decomposed into volumetric and deviatoric (shear) in Equation (10).
Flow Rule and Plastic Modulus
The model employs a flow rule to define plastic strain increment along with plastic paths by a potential tensor (
) with isotropic and deviatoric components
, respectively), which gives directional information on the plastic strain increment (
) and scalar plastic multiplier that controls the magnitude of the plastic strain increment (
) shown in Equation (11). In addition, this model uses a non-associated flow rule for the volumetric components only, allowing the model to predict zero dilatancy at large accumulated shear strains, since dilatancy indicates a volume change in granular materials when they are subjected to shear deformation [
9].
2.1.3. Elastoplastic Constitutive Models
Elastoplastic models are essential in understanding the behavior of rock masses that exhibit elastic or plastic deformations under load. This methodology is for analyzing compressible and incompressible materials or other materials that show different behaviors in tension and compression. The elastoplastic models constitute elastic, yield, and plastic criteria before showing hardening or softening behaviors, illustrated in
Figure 2 and
Figure 3, showing different elastoplastic nonlinear models. These figure illustrates the strain–stress behavior of material under loading, focusing on different modeling approaches for strain hardening and perfect plasticity. The figure depicts three possible paths that the material may follow depending on the constitutive model chosen: bilinear with perfect plasticity, bilinear with strain hardening, and multilinear strain hardening.
There are crucial criteria for developing a constitutive model for the onset of yielding and plastic deformation. There are several yield criteria for various materials and conditions, including the von Mises yield criterion, Tresca yield criterion, Mohr–Coulomb criterion, Drucker–Prager criterion, and Hill’s criterion.
The von Mises yield criterion in Equation (12), known as
flow theory, suggests that yielding occurs when the energy distortion reaches a critical value that is the same as:
The Tresca yield criterion, known as the maximum shear stress criterion, proposes that yielding occurs when maximum shear stress in material reaches the maximum shear stress in simple tension tests. Equation (13) is an extended version of the von Mises yield criterion that includes other ductile materials besides metals:
where
and
are the principal stresses, the difference between them represents the maximum shear stress, and
indicates the yield stress. The Tresca criteria is usually useful for materials that fail under shear stresses, such as brittle materials or metal at lower temperatures.
In addition, there is the Mohr–Coulomb yield criterion, which represents the 2D principal stress state as a series of straight lines, indicating the dependency on cohesion and internal friction angle (ϕ). This method in Equation (14)
The last one is the Drucker–Prager yield criterion, which is a 3D yield criterion in the form of a smooth cone around the hydrostatic axis approximating both von Mises and Mohr–Coulomb criteria, shown below in Equation (15). In this equation
is the invariant of the stress tensor, defined as sum of the principal stresses.
is the second invariant of the deviatoric stress tensor, which is a measure of the shear stress in the material.
is often used in yield criteria like the Drucker–Prager or von Mises criterion, and
is the constant that represents the yield criterion.
In general, von Mises and Drucker–Prager criteria are employed for 3D spaces creating a smooth surface around the pressure or hydrostatic axis. On the other hand, the Tresca criterion visualizes a hexagon in a deviatoric plane, indicating that yield can occur at a specific shear stress level. Moreover, the Mohr–Coulomb criterion is depicted through a 2D plot of normal stress versus shear stress, an indication of failure under combined stress conditions.
2.1.4. Ubiquitous-Joint Models
Ubiquitous-joint models are used to describe the mechanical behavior of rock masses that contain joints and faults, designed to account for weak (joints) of rock masses. These models predict the behavior of the rock mass material by incorporating their mechanical property matrices, like rock and joints. Additionally, the accounts for the anisotropic nature of rock masses and the mechanical response influenced by joint orientation and strength. It can predict failure along joints without the intact rock itself failing.
The equations for ubiquitous-joint models typically involve a combination of following components.
Strain–Stress Relationship
The general stress–strain relationship for rock mass can be expressed using the stress tensor
, the stiffness matrix of the rock mass, which incorporates the mechanical properties of both the intact rock and joints
, and the strain tensor
[
17], as illustrated in Equation below:
Joint Orientation and Strength
Here,
represents the stress contribution due to the joints and
is the Dirac delta function, which activates only along the joint planes with normal
[
18].
Anisotropy Consideration
Here
is the stiffness matrix for the intact rock, and the
is the contribution from each joint [
19].
Failure Criteria
Failure alongside the joints is often predicted using a criterion such as Mohr–Coulomb or Barton–Bandis, adapted to account for joint properties:
where
is the shear stress along the joint,
is the cohesion of the joint,
is the cohesion of the joint, and
is the friction angle of the joint [
20].
Composite Yield Criterion
For cases where failure can occur in both the intact rock and along the joints, a composite yield criterion may be used:
where
and
represent yield functions for the intact rock and the joints, respectively [
21].
Damage Evolution
Damage Mechanics Models
Damage mechanics models focus on the progressive deterioration of materials under different loading conditions. These models consider the effects of microstructural defects such as cracks, voids, and other inclusions that progressively deteriorate the material’s properties.
The damage variable includes a scalar or tensor, which represents the state of the damage in a material, typically ranging from 0 (undamaged material) to 1 (complete failure). There are some equations that encapsulate the core idea of the principle of continuous damage mechanics. Equation (21) represents damage as a reduction in material stiffness, where
is the current modulus of the damaged material, and
is the young modulus of undamaged material. Meanwhile, there is an effective stress concept in Equation (22), which adjusts the nominal stress (
) accounting for the presence of damage, where
is the damage variable [
22].
Another simplified model is the energy equivalence principle, which suggests that the work done on a damaged material is equivalent to the work that would be done on undamaged material, where
and
are all the strains in the damaged and equivalent undamaged material, respectively. This principle provides a basis for determining how damage affects a material’s stiffness and strength, serving as a foundation for formulating damage evolution laws:
Continuous Damage Mechanics
Continuous Damage Mechanics (CDM) involves a theoretical framework and empirical observations used to describe the damage and failure processes in jointed rock masses under triaxial loading conditions. This method evaluates the impact of stress states on rock structural damage to predict the failure of the material due to the evolution of microlevel defects, such as cracks, voids, and any inclusions that accumulate over time and stress states.
The damage evolution law is the other principle under continuous damage mechanics that describes how damage progresses toward failure under different loading conditions. This principle ensures predicting the lifetime of material and structures, as it describes how quickly damage progresses toward failure under different loading conditions. A generic formula is depicted where the rate of damage evolution is a function of the stress state and the current damage state
[
6]:
The thermodynamics of irreversible processes is another principle that CDM often relies on to formulate constitutive relations of damaged material by using the second law of thermodynamics and the concept of dissipative energy to model damage evolution. Indeed, the principle is that the damage mechanics models are thermodynamically consistent, allowing for the prediction of material behavior under complex loading and environmental conditions. The model relates the change in free energy
of a system to the work done
, and heat exchange
, where
is temperature and
is the change in entropy [
23]:
Continuous damage mechanics models damage within a continuum mechanics framework, considering the material as a continuous media despite microlevel cracks. The balance of momentum in continuum mechanics is expressed by the following equation, where σ is the stress tensor,
is the body of the force vector,
is the density, and
is the acceleration vector [
24].
Finally, the last principle under CDM is a generalized model showing damage evolution as a function of temperature
, current damage
, and strain
, illustrating the coupling with thermal processes. Recognizing and modeling this interaction is critical for predicting material behavior in real-world conditions, especially in high-temperature applications or aggressive environments [
25].
Indeed, CDM considers the progressive nature of damage, which can lead to eventual failure, which is a powerful tool to model and predict material behavior under stress.
Chen et al. [
26] developed a constitutive model synthesizing Weibull statistical damage theory and fracture mechanics to analyze the behavior of jointed rock masses under triaxial compression.The foundational concept of the model is a geometric configuration of the joints within a rock mass, including spacing, orientation, and size, which are critical inputs that influence the initial damage estimation.
Furthermore, Chen et al. evaluated Equation (28) for the initial damage (
) done by joints before external pressure is applied, focusing on the importance of the rock’s initial condition When pressure is applied, the rock sustains additional damage, requiring further calculations to combine the initial damage with stress-induced damage. This illustrates how the rock’s condition evolves under pressure, incorporating joint and load-induced damage (Equation (29)). Additionally, Equation (30) integrates various damage components, including the progression from initial to total damage. The terms,
and
account for stress-induced damage normalized by joint-induced damage, while
reflects initial damage normalized by stress-induced damage.
As a result, it can be seen in Equations (31) and (33) that the model’s parameters are related to joint geomaterial and the mechanical characteristic parameters, such as confining pressure, peak strength, and the elastic modulus of the jointed rock mass. The parameter reflects the average strength level of the rock mass, and a higher suggests a higher strength.
Additionally, the theory was put into practice to show that the model can predict what happens inside the rock with different joint setups under various pressures, demonstrating its ability to handle complex real-world conditions. Furthermore, according to Equation (30) and available data, the damage evolution curve with different joint sets was evaluated.
The model was validated through comparing its predictions with experimental data obtained from triaxial compression test conducted on jointed rock masses. The result suggested that the model can predict the failure mechanism in real world scenarios.
The key strength of the model is its ability to consider complex behavior of rock, including anisotropy and damage progression. However, the model’s validation was limited to two degrees of fracture and considered only two only two-dimensional stress. This may not fully capture rock behavior under various environmental conditions, such as different initial fracture levels or more complex stress environments. The model could benefit from incorporating empirical data and practical testing to improve its accuracy in predicting real-world scenarios.
2.1.5. In Situ Load-Testing Interpretation
This methodology aims to interpret the behavior of the rock directly by using the data from the load test performed directly on the rock masses. By doing so, it provides valuable insights into the properties of weak rock masse such ass stiffness and strength, which are crucial for safety analysis.
Several types of in situ tests for weak rock masses are commonly used to evaluate the mechanical properties of the rock in its natural environment.
Plate Load Test
The plate load test is an in-situ testing method used to determine the bearing capacity of the ground and likely settlement. This test is mostly used for designing foundations, especially when structures are going to be built on soil that is not perfectly rigid. This test involves loading a steel plate placed at the foundation level measuring the settlements corresponding to each load increment [
10].
The data observed from this test are plotted in load-versus-settlement curves. These curves are analyzed to determine the ultimate bearing capacity of soil, the safe load-bearing capacity, and the modulus of the subgrade reaction, which is a measure of soil stiffness.
Pressure meter test
The pressure meter test (PMT) is an in situ geotechnical test used to measure ground deformation. It involves inserting a probe into the ground and inflating it to exert pressure on the surrounding soil or rock mass. The test measures the ground’s response to this pressure, thereby providing data on rock mass stiffness, strength, and in situ stress conditions [
11].
Dilatometer Test
The dilatometer test (DMT) is an in-situ test to assess weak rock properties. This method inserts a flat dilatometer blade into the ground at the test depth. The blade has a flexible membrane on one side that can be expanded horizontally into the soil. Furthermore, by applying controlled pressure to the membrane and measuring the soil’s displacement or dilation, weak rock parameters such as elasticity, lateral earth pressure coefficient, and the stratigraphy profile can be determined. This information is crucial for predicting settlement, designing foundations, and other geotechnical purposes [
25].
Flat Jack
The flat jack test is another in situ test involving the insertion of a thin hydraulic jack into a small precut slot in the tested material. Once inserted, the jack is slowly pressurized, causing controlled deformation. This test helps evaluate the stress state of rock masses without causing significant damage to them. The flat jack test setup is illustrated with the actuator system and concrete test blocks.
Goodman Jack Test
The Goodman jack test is widely used to measure the deformation modulus of rock masses in the field. This method involves using a device called a Goodman jack, which is inserted into a bore hole drilled into the rock while hydraulic pressure is applied to the jack, causing it to expand against the borehole walls. Consequently, the deformation of the rock in response to the pressure is measured, resulting in the calculation of the rock deformation modulus [
27].
Asem et al. used the in-situ load test dataset to develop a model that could elucidate the behavior of a drilled shaft in a weak rock mass. The methodology involves creating two types of datasets to develop a model that can predict peak side resistance, tip resistance, and load transfer function for rock sockets in weak rocks, as well as a weak rock mass deformation predictive model. The model was validated by comparing predictions made against real-world observations from in situ loaded tests. However, due to limited data and lack of model documentation quality, the model’s effectiveness and applicability was not justified.
Various load tests were carried out to create a side resistance and base resistance database. The side resistance database provides information about peak side resistance, and stress–shear displacement for the sidewall directly through in situ instrumentation. The base side database focuses on load-transfer functions for shaft bases, properties of intact rock, properties of soft rock masses, base resistance, initial normal stiffness, and drilled shaft geometry. Hence, the methodology aims to provide a comprehensive database on load–displacement responses and rock mass properties to address previous studies’ limitations.
Furthermore, the database is employed to assess existing predictive models for axial resistance, and settlement of the rock foundations. This involves analyzing the model’s ability to consider the variability in rock properties and its impact on foundations reliability.
For design purposes, Young’s modulus
was estimated from a linear regression analysis by [
28,
29] of results from an in-situ plate test, where
is the unconfined compressive strength (UCS) of the rock, which represents the maximum and axial compressive stress that a rock specimen can withstand under no confinement. The existing predictive base model includes a basic rock mass estimation modulus from unconfined compressive strength in a linear model:
For empirical side resistance models, Rosenberg and Journeaux [
30] are among the first ones to provide a model for estimating peak side resistance based on the unconfined compressive strength of the rock. However, they did not come up with a mathematical formula. In this regard, Kulhawy [
31] found the best-fit equation for their method, which constitutes the peak side resistance
, in situ confining pressure
, and unconfined compressive strength
:
Asem provided a summary of the existing base resistance
formula taken from an in situ test database to compare his features and existing assumptions of Rowe and Armitage in Equation (35) [
28], Zhang and Einstein in Equation (36), and Stark et al. [
32] and Baghdady [
33] in Equation (37) iincorporates
which is a dimensionless correction factor that accounts for the effects of rock mass conditions such as presence of discontinuities, weathering, or other factors that reduce the rock strength.
Furthermore, the variation in available theories for base resistance (
) and comparison with measured fracture initiation pressure
is shown in
Figure 4, which helps illustrate the theoretical underpinnings against the measured data, aiming at the model’s effectiveness in predicting base resistance in weak rock masses [
34].
Besides base resistance, Carter and Kulhawy (1988) [
30] presented a method for end resistance of foundation based on the Hoek–Brown failure criterion for jointed rock masses and developed foundation placed on the surface of rock masses (
Figure 5):