Skip to main content
Top
Published in:

Open Access 2025 | OriginalPaper | Chapter

2. Probability Analysis Method of Seismic Response for Earth-Rockfill Dams

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

search-config
loading …

Abstract

Seismic response probabilistic analysis for earth-rock dams is a crucial step in performance-based seismic safety evaluation. It represents a significant transition from deterministic analysis to stochastic analysis. To conduct such an analysis, it's essential to thoroughly consider the uncertainties associated with the seismic response of earth-rock dams and select appropriate and effective probabilistic analysis methods. In the following, the uncertainties present in the seismic response of earth-rock dams and the primary probabilistic analysis methods will be briefly outlined. Additionally, the theoretical foundation and solution procedures of these methods will be briefly introduced. This will establish the theoretical groundwork for subsequent analyses of stochastic dynamic responses and performance-based seismic safety evaluations for high concrete faced rockfill dams.
Seismic response probabilistic analysis for earth-rock dams is a crucial step in performance-based seismic safety evaluation. It represents a significant transition from deterministic analysis to stochastic analysis. To conduct such an analysis, it's essential to thoroughly consider the uncertainties associated with the seismic response of earth-rock dams and select appropriate and effective probabilistic analysis methods. In the following, the uncertainties present in the seismic response of earth-rock dams and the primary probabilistic analysis methods will be briefly outlined. Additionally, the theoretical foundation and solution procedures of these methods will be briefly introduced. This will establish the theoretical groundwork for subsequent analyses of stochastic dynamic responses and performance-based seismic safety evaluations for high concrete faced rockfill dams.

2.1 Uncertainties in the Seismic Response of Earth-Rock Dams

In geotechnical engineering, uncertainties primarily include objective uncertainty and subjective uncertainty. Objective uncertainty is essentially determined by factors such as geotechnical engineering loads, soil parameters, various construction environments, and conditions. On the other hand, subjective uncertainty arises due to the limited human understanding of geotechnical engineering analysis and simulation. This includes aspects like computational models, assumption conditions, and simplifications. Considering the specific context of high concrete faced rockfill dams, two main uncertainty factors are primarily taken into account: the stochastic nature of seismic motion and the uncertainty in the parameters of the rockfill materials.

2.1.1 Randomness of Ground Motion

A considerable number of actual seismic records have indicated the significant uncertainty in seismic motion characteristics. Therefore, a very limited number of computational samples can hardly capture the influence of various factors related to seismic motion stochastic characteristics on the response of earth-rock dam structures, such as the frequency content, amplitude variations, peak values, duration, and arrangement order of different-amplitude pulses within the seismic motion. As early as the 1990s, scholars began exploring the stochastic dynamic response and reliability of earth-rock dams under stochastic seismic excitation. For example, Yu et al. (1993) generated 100 artificial earthquake waves through random simulation, introduced the concept of a state point, employed dynamic programming to locate the most probable sliding surface, and calculated the probability of permanent displacement of earth-rock dams. Chen et al. (1995) used a power function to describe the variation of average shear modulus with dam height and proposed a simplified method for analyzing the stochastic seismic response of heterogeneous earth dams based on a one-dimensional shear beam model. Liu (1996) and Liu et al. (1996) simulated the seismic process as a stationary Gaussian filtered white noise process, utilized the equivalent nodal force method, conducted random seismic response analysis, established an analysis method for the average permanent deformation failure probability of earth-rock dams, and developed a nonlinear random response and dynamic reliability analysis method based on the theory of random vibration and the virtual excitation method. They validated the method's rationality through numerical examples related to earth-rock dams. Shao et al. (1999) simulated the seismic action process as a zero-mean stationary Gaussian process, conducted random seismic response analysis to calculate the stochastic dynamic response of earth-rock dams, and employed the Hook-Jeeves search method to locate the most dangerous sliding plane and the minimum safety factor in the mean sense. This was compared with shake table model tests to validate the effectiveness and rationality of the approach. Wang et al. (2006) proposed a simple seismic motion model that simulates stationary random processes based on stochastic process theory. They investigated the seismic response characteristics of an actual homogeneous earth dam under random load excitation.

2.1.2 Uncertainty of Rockfill Material Parameters

Earth-rock dams are constructed using natural materials, and their properties vary naturally, making them complex with diverse physical and mechanical characteristics. Additionally, their gradation ranges widely, and on-site construction is primarily controlled by the void ratio. It's challenging to precisely control gradation, leading to significant variability and uncertainty in deformation and strength parameters of the dam materials (Wichtmann and Triantafyllidis 2013). However, both deformation and strength parameters have a substantial impact on the numerical analysis results of seismic dynamic responses for high-faced rockfill dams. Currently, in the seismic probabilistic analysis of earth-rock dams, most efforts have been directed towards the uncertainty of seismic motion, with fewer studies investigating the influence of uncertainty in dam construction materials on dam dynamic responses. Zhang and Liu (1994) collected physical and mechanical test data for construction materials from 95 dam engineering projects in China. They mainly established a statistical database based on c and φ parameters and developed a comprehensive probability statistical analysis program, resulting in many valuable outcomes. Wu et al. (1991) considered the uncertainty of rockfill mass, dynamic modulus, and damping. They combined frequency domain analysis and perturbation method in structural dynamic response analysis to study the variation of dam slope displacement. Sanchez et al. (2014) utilized the Karhunen–Loeve expansion method to simulate the stochastic fields of rockfill and dam foundation materials, and solved for the random slope displacement of an earth-rock dam. Kartal et al. (2010) using the example of the Torul rockfill dam, applied an improved response surface method to consider the uncertainty of material and geometric properties of panels and rockfill. They explored the crack resistance and compression reliability of panels with different thicknesses under seismic actions. Wang et al. (2013) combined the mechanical parameter samples of dam construction materials using orthogonal design. Using dam crest settlement as an evaluation indicator, they studied the vulnerability of earth-rock dams. Yang and Zhu (2016) proposed a combined technique of stochastic field simulation and finite element method in the stochastic finite element method. They discussed the effects of spatial uncertainty in dry density, void ratio, coefficient of non-uniformity, average particle size, Duncan-Chang E-B model modulus coefficient, and initial friction angle on the random seismic response and permanent deformation of earth-rock dams.

2.2 Probabilistic Analysis Method

Probability analysis is an effective approach for addressing uncertainty in seismic engineering of earth-rock dams and forms the foundation for the performance-based seismic safety assessment of high concrete faced rockfill dams. Traditional engineering structural probability analysis methods mainly include the first-order second-moment method, the Monte Carlo method, and the response surface method, while other methods are mostly improvements or developments based on these categories. Additionally, for probability analysis based on random vibrations, the Generalized Probability Density Evolution (GPDE) method is a recently developed innovative approach. Below, we will elucidate their concepts and developments from aspects including failure probability definition, first-order second-moment method, Monte Carlo method, and response surface method. Furthermore, we will emphasize the Generalized Probability Density Evolution method, which serves as a probability analysis tool in this paper.

2.2.1 Failure Probability Definition

Generally speaking, the factors influencing the structural performance requirements can be represented by two random comprehensive variables, namely the comprehensive effect S of the structure and the comprehensive resistance R of the structure. Therefore, the functional expression of a structure reaching a certain limit resistance can be quantified as:
$$ Z = R - S $$
(2.1)
Therefore, Z is also a random variable. When Z > 0, the structural response is within the safe region, indicating that the functional safety requirements are satisfied, and this quantified indicator is represented by Pf. When Z < 0, the structural response is outside the safe region, signifying structural functional failure, and this quantified indicator is represented by Pf. Since the factors influencing S and R are a series of more fundamental random variables (such as the strength parameters of earth-rock dams, seismic loads, displacement responses, etc.), let these fundamental random variables be X1, X2, …, Xn. Then, the general form of the functional expression can be represented as:
$$ Z = g(X_{1} ,X_{2} , \ldots ,X_{n} ) \, $$
(2.2)
The limit state equation is:
$$ Z = g(X_{1} ,X_{2} , \ldots ,X_{n} ) = 0 \, $$
(2.3)

2.2.2 First-Order Second-Moment Method

The basic principle of the First-order Second-moment Method (FOSM) (Hasofer and Lind 1974) is to expand the limit state function at a specific point using a Taylor series, considering the uncertainty of random variable distributions. The linearized first-order term is selected, and the mean and standard deviation of the random variables are used to calculate reliability indices. Expanding the limit state function at the point \(X_{0i} {(}i = 1,2, \, \ldots , \, n{) }\) using a Taylor series, we have:
$$ \begin{aligned} Z & = g(X_{01} ,X_{02} ,...,X_{0n} ) + \sum\limits_{i = 1}^{n} {(X_{i} } - X_{0i} )\left( {\frac{\partial g}{{\partial X_{i} }}} \right)_{{X_{0} }} \\ & \quad + \sum\limits_{i = 1}^{n} {\frac{{(X_{i} - X_{0i} )^{2} }}{2}\left( {\frac{{\partial^{2} g}}{{\partial X_{i}^{2} }}} \right)_{{X_{0} }} } + \cdots \\ \end{aligned} $$
(2.4)
In order to derive the linear limit state equation, only the first-order term is approximated:
$$ Z \approx g(X_{01} ,X_{02} ,...,X_{0n} ) + \sum\limits_{i = 1}^{n} {(X_{i} } - X_{0i} )\left( {\frac{\partial g}{{\partial X_{i} }}} \right)_{{X_{0} }} $$
(2.5)
\(\left( {\frac{\partial g}{{\partial X_{i} }}} \right)_{{X_{0} }}\) represents the value of that derivative at point \(X_{0i} (i = 1,2, \, ..., \, n) \, \). Equation (2.5) is a commonly used formula for linearizing the functional expression in reliability analysis.
From the above fundamental principles, it can be seen that the first-order second-moment method, including its various improved forms, is straightforward and computationally convenient. However, for complex nonlinear problems, its accuracy is difficult to ensure. Convergence of iterations is often problematic, and it is no longer applicable to engineering structures that frequently involve implicit functions or complex state functions. In conclusion, it is challenging to apply this method to probabilistic analysis of large and complex nonlinear structures.

2.2.3 Monte Carlo Method

The Monte Carlo method is a numerical computation approach that involves generating random samples of variables based on different probability distribution characteristics. These random variable samples are then used as inputs to obtain samples of a functional function. By counting the number of samples that result in failure or destruction, the method estimates the probability of failure. This approach has a clear conceptual framework, is easy to use, and finds wide application in the field of probability analysis. As the number of simulations increases, the accuracy of the Monte Carlo method improves, but the computational complexity also increases significantly. This makes it challenging to apply to practical engineering scenarios, especially for large-scale and strongly nonlinear projects such as high earth-rock dams. Nonetheless, it is often used to validate the accuracy of other new probability analysis methods.
To enhance the efficiency of numerical simulation methods, various techniques have been developed based on the Monte Carlo foundation, such as importance sampling, subset simulation, line sampling, directional sampling, and Latin hypercube sampling. While these methods improve computational efficiency, they require the determination of design points for limit state functions to obtain important density functions. However, obtaining design points for limit states can be difficult in practical engineering, especially for complex projects, presenting significant challenges to widespread adoption. The estimated probability of failure obtained using the Monte Carlo method can be summarized briefly as follows:
$$ P_{\text{f}} = \frac{{n_{\text{f}} }}{n} $$
(2.6)
in which, nf represents the number of samples in the failure region, and n is the total number of samples.

2.2.4 Response Surface Methodology

For large and complex structures, the limit state function is often implicit. The response surface methodology (RSM) involves constructing a simple explicit function that progressively approximates the actual implicit (or explicit) limit state function. This simplification aids in the calculation of probabilistic reliability and has gained increasing attention in recent years. In 1951, Box and Wilson (1951) first introduced the response surface methodology, focusing on how to use statistical methods to obtain an explicit function for approximating a complex implicit function. Wong (1985) was the first to apply the response surface methodology to analyze the reliability of slope stability. Subsequently, numerous scholars both domestically and internationally have further expanded the research related to this methodology. Based on random variables, the form is as follows:
$$ Y^{\prime} = g^{\prime}(X) = a + X^{T} B + X^{T} CX $$
(2.7)
Determining the constant a and constant matrices B and C requires an adequate number of sample points.
Regarding the reliability analysis of structures, the associated random variables are often numerous. To determine the unknown coefficients in Eq. (2.7), a substantial number of sample points need to be analyzed and computed. This can impact the computational efficiency of the response surface methodology. In order to achieve both sufficient computational accuracy and improved efficiency, in 1990, Bucher and Bourgund (1990) introduced the quadratic response surface function, which has been widely utilized. Its form is as follows:
$$ g^{\prime}(X) = a + \sum\limits_{i = 1}^{n} {b_{i} X_{i} } + \sum\limits_{i = 1}^{n} {c_{i} X_{i}^{2} } $$
(2.8)
In the equation: coefficients a, bi, and ci need to be determined by obtaining a sufficient number of equations using 2n + 1 sample points. However, since the interaction terms do not appear in the response surface function, the number of sample points required for determining the response surface function decreases. The specific calculation method is as follows:
The first step involves considering the mean as the central point and selecting sample points within the interval \({(}m_{\text{X}} - f_{{\sigma_{\text{X}} }} , \, m_{\text{X}} + f_{{\sigma_{\text{X}} }} {)}\). Literature suggests that a suitable value for f is in the range of 1–3. Utilizing the chosen sample points yields 2n + 1 values of the function g(X). Subsequently, the unknown coefficients in the response surface function can be calculated. Once the response surface function is obtained, approximate values for the design check points XD on the limit state surface can be calculated.
The second step involves selecting a new center point, XM, which can be chosen from the line connecting the mean point mX and XD, while ensuring the validity of the limit equation, \(g{(}X{)} = 0\), namely:
$$ X_{M} = m_{X} + \left( {X_{D} - m_{X} } \right)\frac{{g\left( {m_{X} } \right)}}{{g\left( {m_{X} } \right) - g\left( {X_{D} } \right)}} $$
(2.9)
This selection of a new center point aims to ensure that the original limit state surface's information is encompassed by the chosen sample points as much as possible.
The third step involves considering XM as the center point to select a new set of sample points and then repeating the process from the first step. This will yield numerical values for the design check points on the limit state surface and related reliability indicators. The entire procedure requires solving for the values of 4n + 3 functions g(X). Figure 2.1 illustrates the schematic diagram of the entire approximation process.
However, in traditional quadratic response surface methods, when computing nonlinear functional functions, there are often issues with convergence failure and significant errors. Many scholars have proposed improved response surface methods, as well as intelligent response surface methods like Kriging (Luo et al. 2012), Artificial Neural Networks (ANN) (Cho 2009), Radial Basis Functions (RBF) (Deng et al. 2005), and Support Vector Machines (SVM) (Zhao 2008), to enhance the accuracy of structural probability reliability analysis. However, this also increases the complexity of the solution process. Additionally, for highly nonlinear structures such as high earth-rock dams, the effectiveness of sample training for these intelligent methods might not be ideal, and the improvement in accuracy for failure probability estimation might not be significant. It is important to note that first-order second-moment methods and response surface methods are generally used for the probability analysis of static systems. They are less frequently used in dynamic systems or are limited to simple structural probability analysis. Moreover, they cannot adequately capture the stochastic dynamic response process.
In conclusion, the first-order second-moment method, although conceptually simple and computationally convenient, often lacks accuracy and struggles with convergence for complex problems. It may not guarantee precision and is unsuitable for solving engineering structural probability with implicit functions or complex state functions. The Monte Carlo method is versatile, but it's computationally inefficient and challenging to apply to complex real-world engineering problems. However, it can be used to validate the accuracy of other probability analysis methods. The response surface method is primarily employed for solving probability analysis problems with implicit state functions, but it has drawbacks such as convergence failures and significant errors. Some of its improvement methods often suffer from suboptimal sample training results and increased complexity in the solution process.
It is important to emphasize that the first-order second-moment method and the response surface method are generally used for the probability analysis of static systems. They are less frequently used in dynamic systems or are limited to simple structural probability analysis, and they cannot capture the stochastic dynamic response process. In recent years, the probability density evolution method based on stochastic vibration theory has made significant contributions to the field of structural probability analysis. It decouples the solution from complex state functions by solving physical equations and probability density evolution equations. It can incorporate all probability information about structural response and has been successfully applied to structural probability analysis. This approach has played a prominent role in the development and application of structural reliability probability theory.

2.3 Generalized Probability Density Evolution Method

In the 1940s, American scholar Housner (1947) conducted relevant research on the stochastic nature of earthquake ground motion, drawing the attention of seismic researchers from various countries. This marked the beginning of research into stochastic dynamic analysis of earthquakes, including the study of random vibration theory for structural earthquake responses. It found widespread applications in aerospace, mechanical, civil, bridge, and marine engineering fields. Random vibration involves treating input ground motion and resulting structural responses as random processes. Applying stochastic dynamic theory yields statistical characteristics of structural responses, enabling the estimation of failure probabilities. Over time, scholars both domestically and internationally have conducted extensive research based on power spectral analysis, moment evolution, and FPK equation. The study of linear systems within stochastic vibration theory has gradually matured, with improved computational efficiency, and it has been extensively applied in engineering. For instance, Zhu (1993) derived the steady-state solution of the FPK equation based on Hamilton's theory, Lin and Zhong. (1998) proposed the pseudo excitation method, and Fujimura and Kiureghian (2007) introduced the truncated equivalent linear method. Several stationary random earthquake motion models, such as the K-T model (Kanai 1957), C-P model (Clough 1993), S-O model (Ou and Niu 1990), and D-C model (Du and Chen 1994), have also been proposed or developed, yielding satisfactory results.
The aforementioned methods only address the randomness of structural loads. When considering the randomness of structural parameters, the classical stochastic vibration theory encounters significant challenges. On the other hand, earthquakes are typically composed of initiation, main shock, and decay phases. Strictly speaking, they should be considered non-stationary excitation processes. However, extensive research indicates that there is a distinct difference in seismic response between non-stationary and stationary stochastic ground motion models for the same nonlinear structure. Stationary stochastic ground motion models often underestimate structural dynamic responses. Furthermore, for highly nonlinear structures like earth-rock dams under seismic loads, traditional response spectrum analysis based on power spectral models is no longer applicable. Instead, refined simulation of random dynamic responses using seismic acceleration time-history analysis is needed.
Therefore, there is a need for the development and adoption of stochastic dynamic analysis methods tailored to highly nonlinear structures and non-stationary seismic motion models. Currently, some scholars have proposed methods such as stochastic simulation, random perturbation methods, and orthogonal polynomial expansion theory to solve the stochastic dynamic responses of nonlinear structures. However, these methods only yield approximations of second-order statistics and fail to provide complete probability information about structural responses (Liu 2013). For dynamic response analysis of structures, it's essential to consider the influence of various stochastic factors as comprehensively as possible. The stochastic vibration analysis of structures should encompass random dynamic responses and ultimately serve structural probability analysis.
Since 2003, Li and Chen from Tongji University (2003, 2010, 2017) have developed the concept of probability density evolution based on the fundamental idea of probability density evolution, known as the Generalized Probability Density Evolution Method (GPDEM). Starting from the state equation, this method utilizes the stochastic event characterization based on the principle of probability conservation to derive a decoupled generalized probability density evolution equation. It combines techniques such as probabilistic space point selection, deterministic structural numerical analysis, and finite difference methods to perform nonlinear stochastic vibration analysis of structural responses and dynamic reliability probability analysis. Moreover, the probability density function encompasses all stochastic factors in the system, laying a solid foundation for studying reliability probability analysis and uncertainty propagation in complex structures. This approach provides a more accurate description of structural dynamic behavior and has yielded positive results in seismic safety analysis of large and complex structures such as bridges, dams, and aqueducts (Liu et al. 2013, 2014; Liu and Fang 2012).

2.3.1 The Generalized Probability Density Evolution Equation

As is well known, external excitations, system parameters, and initial conditions of structural dynamic responses are all characterized by randomness. Structural dynamic responses can be regarded as stochastic processes, and their statistical probability characteristics are entirely determined by the aforementioned sources of randomness. Therefore, based on the principle of probability conservation, by establishing the probability density evolution rules from the source random factors to the target random responses, it is possible to solve the stochastic dynamic processes and reliability probabilities of complex nonlinear structures. This allows for a comprehensive understanding of their seismic performance under different levels of earthquake excitation.
Based on the knowledge of structural dynamics, the equation of motion for an n-degree-of-freedom system under dynamic loading excitation can be represented as follows:
$$\overline{{\mathbf{M}}} ({\varvec{\Theta}}){\mathbf{\ddot{X}}}(t) + {\mathbf{C}}({\varvec{\Theta} }){\mathbf{\ddot{X}}}({\text{t}} ) + {\mathbf{K}}({\varvec{\Theta} }){\mathbf{X}}(t) = -\overline{{\mathbf{M}}} {\mathbf{\ddot{X}}}_{{\mathbf{g}}} ({\varvec{\Theta} },t) $$
(2.10)
where, \(\overline{M}\) C, and K represent the effective mass, damping, and stiffness matrices of the structure, with their fundamental parameters possibly exhibiting randomness; \({\mathbf{\ddot{X}}}({\text{t}} )\), \({\dot{\mathbf{X}}}({\text{t}} )\) and \({\mathbf{X}}({\text{t}} )\) are the acceleration, velocity, and displacement vectors of the structural response, respectively; \({\mathbf{\ddot{X}}}_{\text{g}} ({\varvec{\Theta} ,}\;{\text{t}} )\) is the stochastic dynamic excitation process; Θ is the random vector within the entire system, and the solution of Eq. (3.​1) uniquely and continuously depends on Θ. For convenience, the solution of Eq. (2.10) can be written as:
$$ {\mathbf{X}}({{t}}) = {\mathbf{H}}({{\varvec{\Theta}}},t) $$
(2.11)
where, H = (H1, H2, …, Hn)T. The velocity time history can be expressed as:
$$ {\mathbf{\ddot{X}}}({{t}}) = {\mathbf{h}}({{\varvec{\Theta}}},t) $$
(2.12)
where, h = (h1, h2, …, hn)T.
Therefore, more generally, any physical quantity of response in the structural system, such as deformation, displacement, velocity, acceleration, and stress, uniquely and continuously depend on Θ. Denote the interested physical quantity as Z = (Z1, Z2, …, Zm)T, then:
$$ \mathbf{Z}({{t}}) = \mathbf{H}_{{Z}} (\varvec{\Theta} ,\;{{t}}) $$
(2.13)
$$ \dot{\mathbf{Z}}({{t}}) = \mathbf{h}_{{Z}} (\varvec{\Theta} ,\;{\text{t}}) $$
(2.14)
where, h = (hz, 1, hz, 2, …, hZ,m)T.
Clearly, Eq. (2.13) can also be regarded as a stochastic dynamic system, where the source random factors are entirely described by Θ. Considering the augmented system formed by (Z, Θ), since all the probabilistic factors are encompassed, it represents a conservative probabilistic system. The joint probability density function (PDF) of (Z, Θ) can be denoted as p(z, θ, t). According to the principle of probability conservation (Chen and Li 2009), the following equation can be obtained:
$$ \frac{\text{D}}{\text{D}t}\int_{{\Omega_{t} \times \Omega_{{{\varvec{\Theta}}}} }} {p_{{Z{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)} \text{d}{\varvec{z}}\text{d}{\varvec{\theta}} = 0 $$
(2.15)
After undergoing certain mathematical manipulations (El Hami and Radi 2016), the above equation can be transformed into:
$$ \frac{{\partial p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)}}{\partial t} + \sum\limits_{l = 1}^{m} {h_{{{\varvec{z}},l}} ({\varvec{\theta}}, \, t)} \frac{{\partial p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)}}{{\partial z_{l} }} = 0 $$
(2.16)
Combining Eq. (2.14), a more explicit conclusion can be drawn:
$$ \frac{{\partial p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)}}{\partial t} + \sum\limits_{l = 1}^{m} {\dot{Z}_{1} ({\varvec{\theta}}, \, t)} \frac{{\partial p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)}}{{\partial z_{l} }} = 0 $$
(2.17)
Therefore, the joint probability density function of Z(t), denoted as pZ(z, t), is:
$$ p{\varvec{z}}({\varvec{z}},\;t) = \int_{{\Omega_{{{\varvec{\Theta}}}} }} {p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)} d{\varvec{\theta}} $$
(2.18)
When considering only a specific response physical quantity, Eq. (2.16) degenerates into a one-dimensional partial differential equation, namely:
$$ \frac{{\partial p{\varvec{z}}({\varvec{z}},\;{\varvec{\theta}},\;t)}}{\partial t} + \dot{\varvec{Z}} (\varvec{\theta} , \, t)\frac{{\partial p_{{{\varvec{z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)}}{\partial z} = 0 $$
(2.19)
The dimensionality m of Eq. (2.16) is independent of the original physical system's degree of freedom n. Regardless of whether the source random factors originate from initial conditions, structural parameters, or external excitations, the governing equation takes the form of Eq. (2.16). Therefore, Eq. (2.16) is referred to as the Generalized Probability Density Evolution Equation, sometimes simply referred to as the Generalized Density Evolution Equation (GDEE). From the above derivation, it is evident that the physical law revealed by the GDEE indicates that, during the evolution process of a general dynamic system, the rate of change of the joint probability density function distribution of the generalized displacement (which can represent actual displacement, stress, deformation, etc.) with respect to time is proportional to the rate of change of the generalized displacement. The proportionality coefficient is determined by the instantaneous generalized velocity, implying that the process of probability density evolution follows strict physical laws. The initial and boundary conditions for Eq. (2.16) are as follows:
$$ p_{{{\varvec{Z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)|_{{t = t_{0} }} = \delta ({\varvec{z}} - {\varvec{z}}_{0} )p_{{{\varvec{\Theta}}}} ({\varvec{\theta}}) $$
(2.20)
$$ p_{{{\varvec{Z}}{{\varvec{\Theta}}}}} ({\varvec{z}},\;{\varvec{\theta}},\;t)|_{{z_{j} \to \pm \infty }} = 0,\;\;j = 1,\;2,\; \ldots ,\;m $$
(2.21)
where z0 represents the deterministic initial value. Hence, solving the GDEE is a combination of solving physical equations and solving the Probability Density Evolution Equation. For some simpler problems, analytical solutions can be obtained using methods like characteristic lines. However, for the majority of engineering practical problems, numerical solutions are needed. From Eq. (2.17), it's apparent that this is a linear partial differential equation. To obtain a numerical solution for this partial differential equation, the coefficients of the equation need to be obtained, which are the derivatives of the physical quantity under consideration when {Θ = θ}. These derivatives can be acquired from the solutions of Eqs. (2.10) and (2.14). Therefore, the numerical implementation of the probability density evolution theory can follow the steps below, as shown in Fig. 2.2. The process of solving the GDEE can be divided into the following four steps:
(1)
Probabilistic space point selection and assignment of probabilities: In the distribution space ΩΘ of the fundamental random variable Θ, a set of discrete representative points θq (q = 1, 2, …, nsel) is selected using methods like number theory, quasi-Monte Carlo, spherical simplex, and GF-deviation, among others. Here, nsel represents the number of discrete representative points. Meanwhile, the assigned probability \(P_{\text{q}} = \int_{{V_{\text{q}} }} {p_{{{\varvec{\Theta}}}} (\;{\varvec{\theta}})} d{\varvec{\theta}}\) for each representative point is determined. Vq denotes the representative volume.
 
(2)
Deterministic dynamic system solution: For each given Θ = θq, solve the physical Eqs. (2.10) and (2.14) to obtain the time derivatives (velocities) \(\dot{Z}_{j} ({\varvec{\theta}}_{q} , \, t_{\text{m}} )\) (j = 1, 2, …, m) of the desired physical quantities. It's worth noting that when the random parameters are determined through probabilistic space point selection, the differential equations of the stochastic dynamic system are transformed into a set of deterministic dynamic equations. For engineering structures, these equations can be solved using various numerical simulation methods such as finite element method and finite difference method.
 
(3)
Solving the GDEE: After the first step of discrete representative point selection and assignment of probabilities, the obtained GDEE is shown below:
$$ \frac{{\partial {\text{p}}_{{{\mathbf{Z\Theta }}}} ({\varvec{z}},\;{\varvec{\theta}}_{\text{q}} ,\;{\text{t}} )}}{{\partial {\text{t}} }} + \sum\limits_{{j = {1}}}^{m} {\mathop {Z_{\text{j}} }\limits^{ \cdot } ({\varvec{\theta}}_{\text{q}} , \, {\text{t}} )} \frac{{\partial {\text{p}}_{{{\mathbf{Z\Theta }}}} ({\varvec{z}},\;{\varvec{\theta}}_{\text{q}} ,\;{\text{t}} )}}{{\partial z_{\text{j}} }} = 0 $$
(2.22)
The corresponding initial conditions become:
$$ p_{{\mathbf{Z}\varvec{\Theta} }} ({\varvec{z}},\;{\varvec{\theta}}_{\text{q}} ,\;t)|_{{t = t_{0} }} = \delta ({\varvec{z}} - {\varvec{z}}_{0} )P_{\text{q}} $$
(2.23)
Substituting the expression, \(\dot{Z}_{j} ({{\varvec{\uptheta}}}_{q} , \, t_{m} )\), obtained in the second step into Eqs. (2.22) and (2.23), and utilizing certain numerical methods such as finite difference method, the partial differential equation can be solved to obtain its numerical solution.
 
(4)
Cumulative summation: By cumulatively summing up all the discrete numerical solutions, \(p_{{{\mathbf{Z\Theta }}}} ({\varvec{z}},\;{\varvec{\theta}}_{\text{q}} ,\;t)\), obtained above, the numerical solution of \(p_{{\varvec{Z}}} (z,\;t)\) can be obtained.
$$ p_{{\varvec{Z}}} \left( {z,\;t} \right) = \sum\limits_{q = 1}^{{n_{sel} }} {p_{{{{\varvec{Z}\varvec{\Theta} }}}} \left( {z,\;{\varvec{\theta}}_{\text{q}} ,\;t} \right)} $$
(2.24)
 
As seen, the solution of the Generalized Probability Density Evolution process is essentially based on the principle of probability conservation. It transforms the stochastic dynamic system into a series of deterministic physical equations that possess inherent probabilistic connections. Through the GDEE, it acquires the probabilistic information about the system's physical state. This process combines the solutions of a series of deterministic dynamic systems and the solution of the GDEE. This exactly embodies the fundamental idea that the evolution of probability density functions depends on the evolution mechanism of the physical system's state. Moreover, the solution of the dynamic system and the solution of the GDEE are decoupled, avoiding the need for repetitive iterations, or solving implicit functions as in traditional reliability probability methods. This approach demonstrates good applicability for complex nonlinear dynamic systems.

2.3.2 The Selection Method of Discrete Representative Points in Probability Space

The selection of discrete representative points in probabilistic space is one of the crucial techniques in the application of the probability density evolution method. Assuming a multidimensional probabilistic space ΩΘ contains a set of discrete representative points θq = (θ1, q, θ2, q, …, θs, q), where q = 1, 2, …, nsel. For each representative point, if we take the Voronoi region as the representative volume (as shown in Fig. 2.3), the probability within this representative volume corresponds to the assigned probability of that representative point:
$$ P_{q} = \Pr \left\{ {{\varvec{\varTheta}}\in V_{q} } \right\} = \int\limits_{{V_{q} }} {P_{{\varvec{\varTheta}}} } ({\varvec{\theta}}){\text{d}}{{\varvec{\uptheta}}}\,\,\,\,q = 1,2, \ldots ,n_{{{\text{sel}}}} $$
(2.25)
$$ \mathop {{\text{lim}}}\limits_{{r_{{{\text{cv}}}} \to 0}} \tilde{p}{}_{{{\varvec{\Theta}}}}({{\varvec{\uptheta}}}) = p{}_{{{\varvec{\Theta}}}}({{\varvec{\uptheta}}}) $$
(2.26)
where rcv is the covering radius of the point set \({\mathcal{P}}_{{{\text{sel}}}} = \left\{ {{\varvec{\theta}}_{\text{q}} = (\theta_{{1,\text{q}}} ,\theta_{{2,\text{q}}} , \cdot \cdot \cdot ,\theta_{{\text{s},\text{q}}} );\;q = 1,2, \cdot \cdot \cdot ,n_{{\text{sel}}} } \right\}\). It can be easily deduced that:
$$ \begin{aligned} \int\limits_{{\Omega_{{{\varvec{\Theta}}}} }} {p_{{{\varvec{\Theta}}}} } ({{\varvec{\uptheta}}}){\text{d}}{{\varvec{\uptheta}}} & = \int\limits_{{\Omega_{{{\varvec{\Theta}}}} }} {\tilde{p}_{{{\varvec{\Theta}}}} } ({{\varvec{\uptheta}}}){\text{d}}{{\varvec{\uptheta}}} = \sum\limits_{q = 1}^{{n_{sel} }} {P_{q} } = \sum\limits_{q = 1}^{{n_{sel} }} {\int\limits_{{V_{q} }} {p_{{{\varvec{\Theta}}}} ({\varvec{\theta}}){\text{d}}{{\varvec{\uptheta}}}} } \\ & = \int\limits_{{U_{q = 1}^{{n_{sel} }} V_{q} }} {p_{{{\varvec{\Theta}}}} ({\varvec{\theta}}){\text{d}}{{\varvec{\uptheta}}}} = 1 \\ \end{aligned} $$
(2.27)
This is the compatibility condition of assigned probabilities. In the probabilistic space, the selection of discrete fundamental point sets should have minimal deviations; based on the probability distribution type, suitable transformations are applied to the fundamental point set to minimize the F-deviation. The formula for calculating the F-deviation is:
$$ {\mathcal{D}}_{{\mathcal{F}}} (n,{\mathcal{P}}) = \mathop {\sup }\limits_{{x{ \in }R^{\text{s}} }} {|\mathcal{F}}_{\text{n}} (x) - {\mathcal{F}}(x)| $$
(2.28)
$$ {\mathcal{F}}_{\text{n}} (x) = \frac{1}{n}\sum\limits_{q = 1}^{n} {I\left\{ {x_{\text{q}} \le x} \right\}} $$
(2.29)
where \({\mathcal{F}}(x)\) is the joint distribution function and \({\mathcal{F}}_{n} (x)\) is the empirical distribution function of the representative point set. \(I\left\{ \cdot \right\}\) is the indicator function. Considering the assigned probabilities of the representative point set, the F-deviation in Eq. (2.29) can be further modified as:
$$ {\mathcal{F}}_{n} (x) = \sum\limits_{q = 1}^{n} {P_{q} \cdot } I\left\{ {x_{q} \le x} \right\} $$
(2.30)
Typically, discrete representative point selection can be achieved using methods such as the sphere-cutting method, lattice point method, number-theoretic methods, and the GF-deviation-based optimization method, significantly reducing the number of selected points. Here, we will briefly introduce the number-theoretic point selection method and the GF-deviation-based point set optimization method used in this paper.
(1)
Number-Theoretic Method: Let there be a vector composed of a set of integers (n, Q1, Q2, …, Qs), and according to the following formula, a set of points can be generated in an s-dimensional space:
$$ \widehat{x}_{j,k} = (2kQ_{j} - 1)\bmod (2n);\quad j = 1,2, \ldots ,s;\,\,k = 1,2, \ldots ,n $$
$$ x_{j,\;k} = \frac{{\widehat{x}_{j,\;k} }}{2n} $$
(2.31)
where \( {\text{mod}} (\cdot)\) represents the remainder after division. The above equation is equivalent to:
$$ x_{j,k} = \frac{{2kQ_{j} - 1}}{2n} - {\text{int}} \left( {\frac{{2kQ_{j} - 1}}{2n}} \right); \quad j = 1,2, \ldots ,s;\,\,k = 1,2, \ldots ,n $$
(2.32)
where \({\text{int}}( \cdot )\) represents the integer part of the expression inside the brackets, and n is the total number of points in the number-theoretic point set, \({\mathcal{P}}_{{\text{NTM}}}\).
Clearly, the numbers in Eqs. (2.31) and (2.32) satisfy:
$$ 0 < x_{j,\;k} < 1,j = 1,2, \ldots ,s;\;\;k = 1,2, \ldots ,n $$
(2.33)
By appropriately selecting the integer vector (n, Q1, Q2, …, Qs), it is possible to generate a point set using Eq. (2.33) with a smaller deviation, such as the point set generation method proposed by Hua and Wang (1981).
 
(2)
GF-Deviation Method: Chen et al. (2016) developed a GF-deviation minimization point set optimization method for non-uniform and non-normal multidimensional distributions. It is mainly achieved in two steps. Step 1: Generate an initial point set using the Sobol sequence (Radović et al. 1996), then rearrange the point set to minimize the GF-deviation. The initial point sets xq = (xq,1, xq,2, …, xq,i), obtained from Sobol point sets uq = (uq,1, uq,2, …, uq,i) (q = 1,2, …, n; i represents the ith random variable) over the unit hypercube, are given by the following expressions:
$$ x_{{{\text{m}},i}} = F_{i}^{ - 1} \left( {u_{{{\text{m}},i}} } \right) $$
(2.34)
 
In the expression, \(F_{i}^{ - 1} ( \cdot )\) is the inverse cumulative distribution function of the ith random variable. Step 2: Transform each random variable so that the assigned probabilities of the sets of n points become mutually close. The assigned probabilities \(x_{q}^{ * } = (x_{q,\;1}^{ * } ,\;x_{q,\;2}^{ * } ,\; \ldots \;,\;x_{{q, i}}^{ * } )\) points are estimated:
$$ x_{m, \, i}^{ * } = F_{i}^{ - 1} \left( {\sum\limits_{q = 1}^{n} {\frac{1}{n} \cdot I\{ x_{q, \, i} < x_{m, \, i} \} + \frac{1}{2} \cdot \frac{1}{n}} } \right) $$
(2.35)
To reduce the GF bias, the following transformation is applied:
$$ x_{m,i}^{**} = F_{i}^{ - 1} \left( {\sum\limits_{q = 1}^{n} {p_{q} \cdot I\{ x_{q, \, i}^{*} < x_{m, \, i}^{*} \} + \frac{1}{2} \cdot p_{m} } } \right) $$
(2.36)
Finally, \(x_{q}^{ * * } = (x_{q,\;1}^{ * * } ,x_{q,\;2}^{ * * } , \ldots ,x_{q,\;i}^{ * * } )\) represents the representative set of points used.

2.3.3 Numerical Solution Methods

In some simpler cases, analytical solutions for the probability density evolution equation can be obtained. However, for complex multi-degree-of-freedom structural solutions, numerical methods are typically employed, such as various forms of finite difference methods and finite element methods. Here, we briefly introduce the Total Variation Diminishing (TVD) finite difference method and the Streamline Upwind Petrov–Galerkin (SUPG) finite element method used in this paper.
(1)
Total Variation Diminishing (TVD) finite difference method.
 
To solve the GDEE, the Total Variation Diminishing (TVD) finite difference method is often employed. Equation (2.21) can be discretized as follows:
$$ \begin{aligned} p_{{\text{j}}}^{{\left( {{\text{k + }}1} \right)}} & =p_{{\text{j}}}^{{\left( {\text{k}} \right)}} - \frac{1}{2}\left( {\uplambda a{ - | \uplambda }\text{a}|} \right)\Delta p_{{{\text{j}} + \frac{1}{2}}}^{{\left( {\text{k}} \right)}} - \frac{1}{2}\left( {\uplambda a{ + | \uplambda }\text{a}|} \right)p_{{{\text{j}} - \frac{1}{2}}}^{{\left( {\text{k}} \right)}} \\ &\quad - \frac{1}{2}\left( {{| \uplambda }\text{a}| -\uplambda ^{2} a^{2} } \right)\left( {\uppsi _{{{\text{j}} + \frac{1}{2}}}\Delta p_{{{\text{j}} + \frac{1}{2}}}^{{\left( {\text{k}} \right)}} -\uppsi _{{{\text{j}} - \frac{1}{2}}}\Delta p_{{{\text{j}} - \frac{1}{2}}}^{{\left( {\text{k}} \right)}} } \right) \\ \end{aligned} $$
(2.37)
where:
$$ {r_{{\text{j} + \frac{1}{2}}}^{ + } = \frac{{\Delta p_{{\text{j} + \frac{3}{2}}}^{{(\text{k})}} }}{{\Delta p_{{\text{j} + \frac{1}{2}}}^{{(\text{k})}} }} = \frac{{\Delta p_{{\text{j} + 2}}^{{(\text{k})}} -\Delta p_{{\text{j} + 1}}^{{(\text{k})}} }}{{\Delta p_{{\text{j} + 1}}^{{(\text{k})}} { - \Delta }\text{p}_{\text{j}}^{{(\text{k})}} }},r_{{\text{j} + \frac{1}{2}}}^{ - } = \frac{{\Delta p_{{\text{j} - \frac{1}{2}}}^{{(\text{k})}} }}{{\Delta p_{{\text{j} + \frac{1}{2}}}^{{(\text{k})}} }} = \frac{{\Delta p_{\text{j}}^{{(\text{k})}} -\Delta p_{{\text{j} - 1}}^{{(\text{k})}} }}{{\Delta p_{{\text{j} + 1}}^{{(\text{k})}} - \Delta }p_{\text{j}}^{{(\text{k})}} }} $$
(2.38)
Introducing the expression for flux limiter:
$$\uppsi _{{\text{j} + \frac{1}{2}}} (r_{{\text{j} + \frac{1}{2}}}^{ + } ,r_{{\text{j} + \frac{1}{2}}}^{ - } ) = u( - a)\uppsi _{0} (r_{{\text{j} + \frac{1}{2}}}^{ + } ) + u(a)\uppsi _{0} (r_{{\text{j} + \frac{1}{2}}}^{ - } ) $$
(2.39)
In the equation, \(u( \cdot )\) represents the Heaviside function, and \(\psi_{0} (r) = \max (0,\min (2r,1),\min (r,2))\). Research indicates that the difference scheme (Eq. 2.38) possesses the TVD (Total Variation Diminishing) property:
$$ {\text{TV}}[p( \cdot ,\;t_{2} )] \le {\text{TV}}[p( \cdot ,\;t_{1} )] \le {\text{TV}}[p( \cdot ,\;t_{0} )],\quad t_{2} > t_{1} > t_{0} , $$
(2.40)
where the total variation is:
$$ {\text{TV}}[p( \cdot ,t)] = \int\limits_{ - \infty }^{\infty } {|\text{ }\frac{\partial p(x,t)}{{\partial x}}|} dx $$
(2.41)
For the case of discrete curves:
$$ {\text{TV}}\left( {p_{ \cdot }^{(k)} } \right) = \sum\limits_{j = - \infty }^{\infty } {\left| {p_{j + 1}^{k} - p_{j}^{k} } \right|} $$
(2.42)
(2)
Streamline Upwind Petrov–Galerkin (SUPG) finite element method.
 
The SUPG finite element method possesses better convergence properties and can obtain the probability density function at boundary locations more effectively. Based on similarity, Eq. (2.21) can be expressed as follows:
$$ {\mathbf{w}} \cdot \nabla p = 0 $$
(2.43)
where w = (1, α(x)) represents the velocity field, α(x) is the flux. and \(\Omega =\Omega_{\text{u}} \times \Omega_{\text{x}} \subset R^{2}\) denotes the computational domain of the finite element partition, where \( e = 1,2, \ldots ,N_{{\text{el}}}\). As the domain is two-dimensional, we can discretize it using quadrilateral elements.
According to reference (Elman et al. 2014), we can define the trial solution space as, where \({\mathcal{V}} = \left\{ {\left. p \right|p \in {\mathcal{H}}^{1} \left( \Omega \right),p = g\;\;on \, \partial \Omega } \right\}\) is the Sobolev space. We also define the space: \({\mathcal{V}}_{0} = \left\{ {\left. \Psi \right|\Psi \in {\mathcal{H}}^{1} \left( \Omega \right),\Psi = 0\;\;on \, \partial \Omega } \right\}\).The fundamental idea of Petrov–Galerkin approximation is to specify a weak formulation where the space of test (weighting) functions is different from that of the trial solution. More specifically, the test space is spanned by functions of the form given in Eq. (2.44):
$$ \widetilde{\psi }=\psi { + }\tau {\mathbf{w}}\nabla \psi $$
(2.44)
where \(\Psi \in {\mathcal{V}}_{0}\) is the Galerkin-type weighting function, τ is a coefficient, and element e is given by the following expression:
$$ \tau_{\text{e}} = \frac{{a\lambda_{\text{e}} }}{{2\left| {{\mathbf{w}}_{\text{e}} } \right|}} $$
(2.45)
In the above expression, e represents the characteristic length of element \( \tau_{\text{e}} = \min \frac{{\lambda_{\text{x}} }}{\cos \vartheta }, \frac{{\lambda_{\text{u}} }}{\sin \vartheta }\), λx and λu denote the lengths of the rectangle in the x and u directions, respectively, where \(\vartheta =\arctan \left( {\left| {\frac{{\omega_{\text{u}} }}{{\omega_{\text{x}} }}} \right|} \right)\). The following equations will be denoted by subscript h to indicate the discrete finite element problem. The weak form of Eq. (2.43) is:
$$ \int\limits_{\Omega } {\psi^{\text{h}} } {\mathbf{w}}\nabla p^{\text{h}} d\Omega + \sum\limits_{e = 1}^{{n_{{\text{el}}} }} {\int\limits_{{\Omega_{\text{e}} }} \tau } {\mathbf{w}}\nabla \psi^{\text{h}} {\mathbf{w}}\nabla p^{\text{h}} d\Omega_{\text{e}} =0 $$
(2.46)
To further enhance the accuracy of the method, we incorporate the approach proposed in (Hughes et al. 1986) into Eq. (2.46), resulting in:
$$ \begin{aligned} \int\limits_{\Omega } {\psi^{\text{h}} } {\mathbf{w}}\nabla p^{\text{h}} d\Omega & + \sum\limits_{e = 1}^{{n_{{\text{el}}} }} {\int\limits_{{\Omega_{\text{e}} }} {\tau_{1} } } {\mathbf{w}}\nabla \psi^{\text{h}} {\mathbf{w}}\nabla p^{\text{h}} d\Omega_{\text{e}} \\ & + \sum\limits_{e = 1}^{{n_{{\text{el}}} }} {\int\limits_{{\Omega_{\text{e}} }} {\tau_{2} } } {\mathbf{w}}\nabla \psi^{\text{h}} {\mathbf{w}}_{||} \nabla p^{\text{h}} d\Omega_{\text{e}} = 0 \\ \end{aligned} $$
(2.47)
Here, w|| represents the projection of w onto the values on \(\nabla p^{\text{h}}\):
$$ {\mathbf{w}}_{||} = \left\{ {\begin{array}{*{20}c} {\frac{{{\mathbf{w}}\nabla p^{\text{h}} }}{{\left| {\nabla p^{\text{h}} } \right|_{2}^{2} }},\quad {\text{if}}\,\,\nabla p^{\text{h}} \ne 0} \\ {0\quad {\text{if}}\,\,\nabla p^{\text{h}} = 0} \\ \end{array} } \right. $$
(2.48)

2.4 Non-Stationary Stochastic Seismic Motion Model

Seismic motion significantly influences the seismic response of structures, exhibiting pronounced randomness in both intensity and frequency. However, most of the current research is based on deterministic analysis methods. Therefore, it is essential to employ stochastic dynamic theory based on dynamic time history analysis to explore the seismic response of structures under stochastic seismic excitations. Utilizing the numerical solution process of the generalized probability density evolution method, the probability density evolution equation discretizes the probability space constituted by stochastic factors. As a result, the number of acceleration time history samples is determined by the quantity of discretized representative points. It is evident that the non-stationary stochastic seismic motion model forms the foundation for analyzing the random seismic response of engineering structures and seismic reliability probability using the generalized probability density evolution method.

2.4.1 Improved Clough-Penzien Power Spectral Model

Large-scale structures, especially earth-rock dams, are quite intricate. Under seismic actions, they exhibit nonlinear and even strongly nonlinear effects. Nonlinear dynamic analysis is necessary. The traditional response spectrum method is no longer applicable, and seismic acceleration time history analysis must be employed to comprehensively understand the seismic response process. Conventional artificial synthesis methods for stochastic seismic acceleration time histories seldom consider the influence of non-stationary characteristics of seismic motion, which is unreasonable. Therefore, it is appropriate to use frequency-dependent non-stationary power spectral models to synthesize seismic acceleration time histories. In recent years, some researchers have introduced frequency-dependent non-stationary power spectral models into the field of hydraulic structures to study their effects on dam bodies. However, this area has been less explored in the context of earth-rock dams, especially high-panel block dams. Nonetheless, based on the characteristics of high-panel block dams, the study of non-stationary seismic stochastic dynamic excitation is one of the effective paths for the performance-based seismic safety assessment of high-panel block dams. In 1996, Deodatis proposed an evolving power spectral model for fully non-stationary seismic acceleration time histories based on the stationary Clough-Penzien power spectral model. In 2011, Cacciola and Deodatis improved the model. The bilateral evolving power spectral density function can be constructed using the following equation:
$$ \begin{aligned} S_{{{\ddot{\text{X}g}}}} (t,\;\omega ) & = A^{2} (t)\frac{{\omega _{{\text{g}}}^{4} (t) + 4\xi _{{\text{g}}}^{{\text{2}}} (t)\omega _{{\text{g}}}^{{\text{2}}} (t)\omega ^{2} }}{{[\omega ^{2} - \omega _{{\text{g}}}^{2} (t)]^{2} + 4\xi _{{\text{g}}}^{{\text{2}}} (t)\omega _{{\text{g}}}^{{\text{2}}} (t)\omega ^{2} }} \\ & \cdot \frac{{\omega ^{4} }}{{[\omega ^{2} - \omega _{{\text{f}}}^{{\text{2}}} (t)]^{2} + 4\xi _{{\text{f}}}^{{\text{2}}} (t)\omega _{{\text{f}}}^{{\text{2}}} (t)\omega ^{2} }}S_{0} (t) \\ \end{aligned} $$
(2.49)
In the equation: A(t) is the intensity modulation function, recommended to be taken as (Cacciola and Deodatis 2011):
$$ A(t)=\left[ {\frac{t}{c}\exp \left( {1 - \frac{t}{c}} \right)} \right]^{\text{d}} $$
(2.50)
In the equation, c represents the time at which the peak ground acceleration (PGA) of the seismic motion occurs, and in this paper, it is set to 4 s; d is a parameter controlling the shape of A(t), and in this paper, it is set to 2. In the modulation function of the evolving power spectral density, the following parameters reflect its frequency-dependent non-stationary characteristics:
$$ \omega_{{\text{g}}} (t) = \omega_{0} - a\frac{t}{T},\xi_{{\text{g}}} (t) = \xi_{0} + b\frac{t}{T} $$
(2.51)
$$ \omega_{{\text{f}}} (t) = 0.1\omega_{\text{g}} (t),\,\xi_{{\text{f}}} (t) = \xi_{\text{g}} (t) $$
(2.52)
In the above equation, ω0 and ξ0 are the initial circular frequency and initial damping ratio, which can be determined by site characteristics. In this paper, they are taken as 25 and 0.45 rad/s, respectively. a and b are parameters determined based on site characteristics and seismic category. Their values are 3.5 and 0.3 rad/s, respectively. T represents the duration of the seismic acceleration time history, and its value varies depending on the site. Typically, for Site Classes I0, I1, II, III and IV, the values are taken as 12s, 15s, 20s, 25s, and 30s respectively. The spectral parameters that reflect spectral intensity can be expressed as:
$$ S_{0} (t) = \frac{{\overline{a}_{\max }^{2} }}{{\gamma^{2} \pi \omega_{\text{g}} (t)[2\xi_{\text{g}} (t) + 1/(2\xi_{\text{g}} (t))]}} $$
(2.53)
where \(\overline{a}_{\max }\) is the mean of the peak ground acceleration (PGA) of the seismic motion, and γ is the equivalent peak factor, taken as Eq. (2.7).

2.4.2 Random Seismic Generation Based on Spectral Representation-Stochastic Process

Building upon the generalized Clough-Penzien power spectral model, this paper employs the concept of stochastic processes to achieve the generation of non-stationary seismic motion using a spectral representation based on hydraulic seismic design codes. By combining this approach with the generalized probability density evolution method, the paper conducts refined stochastic dynamic response and seismic reliability analysis for high-panel block dams. Typically, non-stationary seismic acceleration random processes with zero mean can be generated using the following formula (Ou and Wang 1998):
$$ \varvec{\ddot{X}}_{g} (t) = \sum\limits_{k = 1}^{N} {\sqrt {2S_{{\ddot{\text{X}}_{{\text{g}}} }} (t,\omega_{\text{k}} )\Delta \omega } } \left[ {\cos (\omega_{\text{k}} t){\varvec{X}}_{\text{k}} + \sin (\omega_{k} t){\varvec{Y}}_{\text{k}} } \right] $$
(2.54)
In the equation, ωk = kΔω, and \({\text{S}}_{{\ddot{X}_{{\text{g}}} }}\) is the bilateral evolving power spectral density function. At the frequency ω = 0, it should satisfy:
$$ {\text{S}}_{{\ddot{X}_{\text{g}} }} (t,\;\omega_{0} ) = {\text{S}}_{{\ddot{X}_{\text{g}} }} (t,\;0) = 0 $$
(2.55)
In Eq. (2.54), \(\{ X_{{\text{k}}} ,Y_{{\text{k}}} \}\) (k = 1, 2, …, N) are standard orthogonal random variables that satisfy the following fundamental conditions:
$$ E[X_{\text{k}} ] = E[Y_{\text{k}} ] = 0 $$
(2.56)
$$ E[X_{\text{j}} Y_{\text{k}} ] = 0,E\left[ {X_{\text{j}} X_{\text{k}} } \right] = E\left[ {Y_{\text{j}} Y_{\text{k}} } \right] = \delta_{{\text{jk}}} $$
(2.57)
In the equation, E [⋅] represents the mathematical expectation, and δjk is the Kronecker delta. The relative error of the mean for simulating non-stationary seismic acceleration processes can be expressed as:
$$ \varepsilon {(}N{)} = 1 - \frac{{\int\limits_{0}^{{\omega_{\text{u}} }} {\int\limits_{0}^{T} {S_{{\mathop X\limits^{ \cdot \cdot }_{\text{g}} }} (t,\omega )} } \text{d}t\text{d}\omega }}{{\int\limits_{0}^{\infty } {\int\limits_{0}^{T} {S_{{\mathop X\limits^{ \cdot \cdot }_{\text{g}} }} (t,\omega )} } \text{d}t\text{d}\omega }} $$
(2.58)
In the equation: ωu = NΔω represents the truncation frequency, T is the duration of the non-stationary seismic acceleration process. Typically, the relative error of the mean ε(N) ≤ 1.0 is less than or equal to 1.0. In this paper, Δω is set to 0.15 rad/s, and the truncation term N is taken as 1600.
Then, based on the concept of stochastic processes, the random function expression of the standard orthogonal random vector \(\{ X_{{\text{k}}} ,Y_{{\text{k}}} \}\) can be constructed. Assuming that any two sets of standard orthogonal random vectors \(\overline{X}_{n}\) and \(\overline{{Y_{n} }}\) (n = 1, 2, …, N) are functions of two independent random variables Θ1 and Θ2, the random function can be denoted as:
$$ \overline{{X_{n} }} = {\text{cas}}\left( {n\Theta_{1} } \right)\,\,\overline{Y}_{n} = {\text{cas}}\left( {n\Theta_{2} } \right) $$
(2.59)
In the equation, \(cas({\text{x}}) = \cos ({\text{x}}) + \sin ({\text{x}})\) is the Hartley orthogonal basis function. The fundamental random variables Θ1 and Θ2 are uniformly distributed and mutually independent in the interval [0, 2π], which can usually be obtained using number-theoretical methods. After a certain deterministic mapping, \(\{ X_{{\text{k}}} ,Y_{{\text{k}}} \}\) becomes the standard orthogonal basis random variable needed for Eq. (2.54) and is uniquely determined.
Taking Site I1 as an example, an acceleration time history with a duration of 15s is generated. The discretized representative points Θ1,i and Θ2,i are obtained using number-theoretical methods. To reduce fitting errors, the evolving power spectral density can often be corrected using the following equation:
$$ S_{{{\ddot{\text{X}}}_{{\text{g}}} }} (t,\omega )|_{{\text{m + 1}}} = \left\{ {\begin{array}{*{20}l} {S_{{{\ddot{\text{X}}}_{{\text{g}}} }} (t,\omega ),{\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} 0 < \omega \le \omega_{{{\text{c}}{\kern 1pt} }} {\kern 1pt} {\kern 1pt} } \\ {S_{{{\ddot{\text{X}}}_{{\text{g}}} }} (t,\omega )|_{{\text{m}}} \frac{{RSA^{{\text{T}}} (\omega ,\zeta )^{2} }}{{RSA^{{\text{S}}} (\omega ,\zeta )^{2} |_{{\text{m}}} }}{\kern 1pt} ,{\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} \omega > \omega_{{{\text{c}}{\kern 1pt} }} {\kern 1pt} {\kern 1pt} {\kern 1pt} {\kern 1pt} } \\ \end{array} } \right. $$
(2.60)
\(S_{{{\ddot{\text{X}}}_{{\text{g}}} }} (t,\omega )|_{{{\text{m + }}1}}\) and \(S_{{{\ddot{\text{X}}}_{{\text{g}}} }} (t,\omega )|_{{\text{m}}}\) are the evolving power spectra obtained after the m + 1 and m iterations, respectively. \(RSA^{{\text{S}}} (\omega ,\zeta )^{2} |_{{\text{m}}}\) is the average response spectrum obtained in the m iteration. \(RSA^{{\text{T}}} (\omega ,\zeta )^{2}\) is the target spectrum, often based on seismic design codes. Here, ω = 2π/T0, T0 is the natural period of vibration. ζ is the damping ratio, typically taken as 0.05 for earth-rock dams. ωc is the cutoff frequency, which can be set as 1.57 rad/s. By substituting Eq. (2.60) into Eq. (2.54), a set of representative acceleration time histories can be generated. Generally, after a few iterations, the required accuracy can be achieved. To provide a detailed description of the differences between the generated acceleration time histories and the target values, the following relative error controls are commonly employed:
$$ \varepsilon_{m} = \frac{1}{{N_{m} }} \cdot \sum\limits_{k = 1}^{{N_{m} }} {\left| {\frac{{X(t_{k} ) - \overline{X} (t_{k} )}}{{Y(t_{k} )}}} \right|} $$
(2.61)
$$ \varepsilon_{s} = \frac{1}{{N_{m} }} \cdot \sum\limits_{k = 1}^{{N_{m} }} {\left| {\frac{{Y(t_{k} ) - \overline{Y} (t_{k} )}}{{Y(t_{k} )}}} \right|} $$
(2.62)
$$ \varepsilon_{r} = \frac{1}{{N_{r} }} \cdot \sum\limits_{k = 1}^{{N_{r} }} {\left| {\frac{{R(T_{0, \, k} ) - \overline{R} (T_{0, \, k} )}}{{R(T_{0, \, k} )}}} \right|} $$
(2.63)
In the equation, εm represents the average relative error of the mean acceleration time history; εs is the average relative error of the standard deviation of the acceleration time history; εr is the average relative error of the response spectrum mean. X(tk) and Y(tk) are the mean and standard deviation of the target values at the kth control point. \(\overline{X} (t_{k} )\) is the sample mean of the kth control point, and \(\overline{Y} (t_{k} )\) is the sample standard deviation. Nm = 1500 is the number of time history discrete points. R(T0, k) is the response spectrum obtained based on seismic design codes, such as hydraulic seismic design codes. \(\overline{R} (T_{{{0, }k}} )\) is the response spectrum of the generated samples. Nr = 400 is the number of discrete points for natural period. When generating 233 sets of acceleration time histories, the resulting errors are 4.8% for εm, 4.2% for εs, and 3.4% for εr. Figure 2.4 provides a comparison between the generated seismic motion sample mean, standard deviation, and response spectrum, and the target values (with \(\overline{a}_{\max } = 2\)). It can be observed that the fit is very good.

2.5 Dynamic Reliability Probability Analysis

In many cases, the statistics based on random dynamic response can describe the seismic response information of the structure and judge whether it is safe or not. However, based on a certain failure or failure criterion, it is more reasonable and more intuitive to give the reliability probability information of the system. It is also the main purpose of stochastic dynamics research to quantitatively evaluate the safety of structural systems from the perspective of failure probability, and it is also an important part of performance-based seismic safety evaluation of high concrete face rockfill dams. The dynamic reliability of engineering structures usually includes the reliability of first-passage failure and cumulative damage failure, which can be obtained by constructing a virtual stochastic process and solving the corresponding generalized probability density evolution equation.
For stochastic dynamical systems, within a given time interval [0, T], the extreme value or cumulative value depends on the random vector Θ. Taking the extreme value distribution X(t) as an example, the extreme value under seismic action can be expressed as:
$$ {\mathbf{Y}}_{{\mathbf{X}}} = \max \text{ }\left( {\left| {{\mathbf{H}}_{{\mathbf{X}}} ({\varvec{\varTheta}},\;T)} \right|,\;{\mathbf{t}} \in \left[ {0,\;{\mathbf{T}}} \right]} \right) $$
(2.64)
For a given Θ, YX exists and is unique, so there is
$$ {\mathbf{Y}}_{{\mathbf{X}}} = {\mathbf{W}}_{{\mathbf{X}}} ({{\varvec{\Theta}}}\text{,}\;T) $$
(2.65)
Therefore, a virtual stochastic process can be constructed:
$$ {\mathbf{Q}}_{{\mathbf{X}}} (\tau ) = {\mathbf{Y}}_{{{\varvec{\Theta}}}} \tau = {\mathbf{W}}_{{\mathbf{X}}} ({{\varvec{\Theta}}},\;T)\tau $$
(2.66)
where, τ is a virtual time parameter. Obviously there are:
$$ {\mathbf{Q}}_{{\mathbf{X}}} (\tau )|_{\tau = 0} = 0,{\mathbf{Y}}_{{\mathbf{X}}} = {\mathbf{Q}}_{{\mathbf{X}}} (\tau ){|}_{\tau = 1} $$
(2.67)
Derivation of Eq. (2.67) with respect to τ, then
$$ \dot{\varvec{Q}}_{{\mathbf{X}}} = \frac{{\partial {\varvec{Q}}_{{\mathbf{X}}} }}{\partial \tau } = {\varvec{W}}_{{\mathbf{X}}} ({\varvec{\varTheta}},\;T) $$
(2.68)
Thus, (Q(τ), ϴ) forms a conservative probability system, and the joint probability density equation pQΘ(q, θ, τ) can be written as:
$$ \frac{{\partial {\text{p}}_{{\text{Q}{{\varvec{\Theta}}}}} ({\mathbf{q}},\;{{\varvec{\uptheta}}},\;\tau )}}{\partial \tau } + W({{\varvec{\uptheta}}}, \, {\text{T}} )\frac{{\partial p_{{\text{Q}{{\varvec{\Theta}}}}} ({\mathbf{q}},\;{{\varvec{\uptheta}}},\;\tau )}}{{\partial {\text{q}} }} = 0 $$
(2.69)
Finally, the generalized probability density evolution equation method is used to solve the Eq. (2.69) to obtain the corresponding reliability or failure probability.

2.6 Verification and Application of the Examples

To validate the computational efficiency and accuracy of the Generalized Probability Density Evolution Method (GPDEM) for stochastic dynamic problems, especially in the stochastic analysis and probability assessment of nonlinear and complex engineering structures, this section will conduct verification in two aspects: seismic load randomness and parameter randomness. This will be achieved by utilizing non-stationary stochastic seismic ground motion models and the GF-bias method to generate acceleration time history samples and high-dimensional random parameter samples. These samples will be compared with analytical solutions, the Duffing oscillator, and multi-layered soil-rock slopes to obtain second-order statistical quantities and probability information. Finally, the applicability of the method for earth-rock dams will be primarily demonstrated through the verification on a panel-stacked rockfill dam.

2.6.1 Verification Based on Analytical Solution

This section verifies the accuracy and efficiency of GPDEM based on an undamped single-degree-of-freedom (SDOF) system. The equation of free vibration is as follows:
$$ \ddot{u}(t) + \omega^{2} u(t) = 0 $$
(2.70)
where the initial conditions are as follows:
$$ u(t)|_{t = 0} = x_{0} ,\dot{u}(t)|_{t = 0} = \dot{x}_{0} $$
(2.71)
The natural frequency ω is a random parameter uniformly distributed in the interval \(\left[ {\frac{5\pi }{4},\;\frac{7\pi }{4}} \right]\), with initial values of \(x_{0} = 1.0\text{m}\) and \(\dot{x}_{0} = 0.0\text{m}\), respectively. When \(\dot{x}_{0} = 0.0\text{m}\), the displacement solution for the dynamic system Eq. (2.70) is:
$$ X(t) = x_{0} \cos (\omega t) $$
(2.72)
Clearly, X(t) is a stochastic process. Ultimately, the probability density function (PDF) of the displacement and the second-order statistical time history (mean and standard deviation) can be obtained, as shown in Figs. 2.5 and 2.6. The excellent fit between the analytical solution and GPDEM demonstrates the high accuracy of this method.

2.6.2 Verification Based on Duffing Equation

To verify the accuracy and efficiency of employing the GPDEM method in solving the stochastic processes and reliability of nonlinear dynamic systems under random seismic effects, further investigation is conducted based on the typical nonlinear vibration system, the Duffing oscillator (Sekar and Narayanan 1994). The response of the Duffing oscillator under random seismic effects can be expressed as follows:
$$ \ddot{x}(t) + 2\zeta_{0} a_{0} \dot{x}(t) + a_{0}^{2} \left[ {x(t) + \mu x^{3} (t)} \right] = - \ddot{U}_{g} (t) $$
(2.73)
Here, where a0 = 2.0 rad/s, ζ0 = 0.05 and μ = 200 m represent the natural frequency, damping ratio, and nonlinearity coefficient, respectively. Finally, through comparison with the second-order statistical values obtained using the Monte Carlo Method (MCM) (Fig. 2.7) and the cumulative distribution functions (CDFs) at different time instances (Fig. 2.8), the efficiency and accuracy of GPDEM in dealing with stochastic processes and reliability of nonlinear structural responses under stochastic seismic excitations are verified.

2.6.3 Verification Based on Stochastic Dynamic and Probabilitistic Analysis of Multilayer Slopes

High-dimensional random parameter samples were generated using the GF-bias optimization-based point selection method. By combining this approach with the GPDEM, a probabilistic analysis of multi-layered soil-rock engineering slopes considering the randomness of material parameters was performed. By comparing the results with those obtained using traditional Monte Carlo simulation methods, this approach proves to be effective in generating samples and conducting structural reliability analysis involving dozens or even scores of random variables. It demonstrates high efficiency and reliability. The probabilistic analysis employed deterministic seismic motion, with acceleration time histories shown in Fig. 2.9, where the peak ground acceleration PGA = 0.2g.
(1)
Case 1: Slope Model with Six Random Parameter
 
The first example involves a multi-layered embankment, as described in reference (Reale et al. 2016). Figure 2.10 shows its finite element mesh, consisting of two layers of fill material, topped with a 10 m-deep layer of hard silty soil. The parameter details are provided in Table 2.1. The mesh comprises 3985 elements and 4107 nodes. All random parameters are assumed to follow independent normal distributions, totaling six random variables. Monte Carlo simulation was performed 10,000 times, while GPDEM simulation was carried out 600 times.
Table 2.1
Statistical values of soil parameters in case study 1
Layer
Bulk density/(KN/m3)
Young's modulus/MPa
Poisson's ratio
Cohesion/KPa
Angle of friction/°
Mean
Coefficient of Variation
Mean
Coefficient of Variation
Layer 1
18.0
20
0.3
10
0.1
28
0.05
Layer 2
18.5
20
0.3
8
0.15
29
0.1
Layer 3
20.0
20
0.3
5
0.2
36
0.1
Figure 2.11 presents the stochastic process of safety factors, including a comparison of means and standard deviations. Figure 2.12 showcases probability density functions and cumulative distribution functions at several representative time instances. Figure 2.13 illustrates the probability density function (PDF) and cumulative distribution function (CDF) of the minimum safety factor. The reliability based on the Monte Carlo Method (MCM) and GPDEM is 0.8159 and 0.8173, respectively. It can be observed that the two results are in good agreement, yet the GPDEM demonstrates higher efficiency.
(2)
Case 2: Eight Random Parameter Slopes
 
The second example involves a complex four-layer soil profile, detailed in reference (Zolfaghari et al. 2005). Figure 2.14 depicts its finite element mesh. There are a total of eight random parameters, each following an independent lognormal distribution, as specified in Table 2.2. Monte Carlo simulation was conducted 10,000 times, while GPDEM simulation was performed 800 times.
Table 2.2
Statistical values of soil parameters in case study 2
Layer
Bulk density/(KN/m3)
Young's modulus/MPa
Poisson's ratio
Cohesion/KPa
Angle of friction/°
Mean
Coefficient of variation
Mean
Coefficient of variation
Layer 1
19.0
20
0.3
18
0.5
16
0.3
Layer 2
19.0
20
0.3
20
0.5
29
0.3
Layer 3
19.0
20
0.3
12
0.3
36
0.2
Layer 4
19.0
20
0.3
20
0.5
36
0.3
Figure 2.15 shows the stochastic process of safety factors, including a comparison of means and standard deviations. Figure 2.16 presents the probability density functions and cumulative distribution functions at several representative time instances. Figure 2.17 illustrates the probability density function and cumulative distribution function of the minimum safety factor. The reliabilities based on the MCM and GPDEM are 0.6701 and 0.6675, respectively.

2.6.4 Verification Based on Stochastic Dynamic and Probabilistic Analysis of CFRD

To further verify the effectiveness and efficiency of the GPDEM combined with the non-stationary seismic generation method for random dynamic simulation and probabilistic analysis in highly nonlinear structural panel-stacked rockfill dams, this section contrasts the results with those obtained using traditional MCM. It examines their influence on the dam crest acceleration, panel stress stochastic dynamic response, and associated probabilities.
The finite element mesh of the concrete panel-stacked rockfill dam is shown in Fig. 2.18. In the calculations, the elements use quadrilateral 4-node isoparametric elements, with a total of 3777 elements and 3795 nodes. The computations were conducted on a PC machine with an Intel i7 8-core CPU and 32GB of RAM. The dynamic calculation for a single sample takes about 3 min, while the stable calculation takes around 30 s. The dynamic water pressure is applied using the Westergaard additional mass method (Westergaard 1933). The considered panel-stacked rockfill dam has a crest width of 25 m, a height of 245 m, an upstream dam slope of 1:1.5, an upper downstream ramp slope of 1:1.7, and a lower downstream ramp slope of 1:1.4. The dam body consists of panels and five stacking areas: stacking area A, stacking area B, stacking area C, transition zone, and cushion zone. The reservoir water level is 225 m.
The static analysis of the rockfill material uses the Duncan E-B nonlinear elastic model (Duncan and Chang 1970), and Table 2.3 provides the parameters for the static model. The dynamic analysis employs the Hardin-Drnevich equivalent linear viscoelastic model (Hardin and Drnevich 1972), with parameters as presented in Table 2.4. The peak ground acceleration of the seismic motion is adjusted to PGA = 0.4g, and a total of 377 and 5000 seismic records are generated using the number theoretic method and the Monte Carlo method, denoted as 377 GPDEM and 5000 MCM, respectively. Through a series of finite element dynamic time history analyses and solutions of probability density evolution equations, the dam crest acceleration and panel stress stochastic dynamic response second-order statistical values, as well as probability information, can be obtained.
Table 2.3
Parameters for Duncan E-B model
Material
ρ/(kg/m3)
K
n
Rf
Kb
m
φ0/(°)
Δφ/(°)
Rockfill A
2150
1109
0.24
0.64
420
0.26
49.8
7.2
Rockfill B
2100
800
0.32
0.64
490
0.30
49.8
7.2
Rockfill C
2170
980
0.26
0.79
400
0.31
50
8.2
Transition
2222
1250
0.31
0.78
500
0.16
53.5
10.7
Cushion
2258
1200
0.30
0.75
680
0.15
54.4
10.6
Table 2.4
Parameters for Hardin-Drnevich model
Material
K
n
ν
Rockfill A
2660
0.444
0.33
Rockfill B
3115
0.396
0.33
Rockfill C
4997
0.298
0.33
Transition
3223
0.455
0.40
Cushion
3828
0.345
0.40
The comparison of mean and standard deviation time histories of dam crest acceleration and panel dynamic stress (tensile stress is positive) based on the Generalized Probability Density Evolution Method (GPDEM) and the Monte Carlo Method (MCM) is illustrated in Fig. 2.19. By comparing 377 GPDEM samples with 5000 MCM samples, the mean and standard deviation obtained from 377 GPDEM closely match those from 5000 MCM. This demonstrates the effectiveness and accuracy of the GPDEM method, which requires fewer finite element samples and computation time.
However, the fluctuations in the mean and standard deviation time histories also indicate significant differences in various physical quantities for different seismic responses. These differences arise from the stochastic nature of seismic motion. Therefore, a comprehensive analysis of the seismic response of high concrete faced rockfill dams is needed from the perspective of random vibration. Nevertheless, it's worth noting that there is still a slight difference between the results of 377 GPDEM samples and 5000 MCM samples, as seen from the second-order statistical values.
A series of deterministic seismic response analyses were used to obtain the physical quantities of acceleration and panel stress, which were then inserted into the GPDEM equation. The probability information was obtained using the Total Variation Diminishing (TVD) scheme of finite difference method. Figure 2.20 shows the cumulative distribution functions at two representative time instances. It can be observed that the cumulative distribution functions obtained from 377 GPDEM samples and 5000 MCM samples fit well, confirming the high precision and effectiveness of GPDEM. It's also noticeable that the cumulative distribution functions change over time, indicating the significant influence of seismic motion on the seismic response of panel-stacked rockfill dams under the coupling effect of nonlinear rockfill behavior and stochastic seismic excitation. Therefore, the seismic response of high concrete faced rockfill dams requires analysis from a stochastic perspective.
Figure 2.21 presents the cumulative distribution functions of maximum acceleration and maximum stress obtained based on the virtual random process and solving the GDEE using the SUPG scheme of the finite element method. The comparison between the results of 377 GPDEM samples and 5000 MCM samples further validates the accuracy and efficiency of the GPDEM method. Additionally, due to the influence of the probability density evolution process, an improved level of accuracy is achieved.
By comparing the failure probabilities or reliabilities, as well as the mean and standard deviation of responses, the results obtained from several hundred GPDEM simulations are within the same order of magnitude as those obtained from tens of thousands of Monte Carlo simulations. However, the efficiency of GPDEM is dozens or even several tens of times higher than the Monte Carlo method. This indicates the effectiveness and efficiency of this approach for large-scale geotechnical engineering stochastic and probabilistic analysis. In conclusion, considering the high accuracy and efficiency of the Generalized Probability Density Evolution Method, along with its strong theoretical foundation, it can be effectively employed for performance-based seismic safety assessment of high concrete faced rockfill dams, with the expectation of achieving favorable results.

2.7 Seismic Fragility Analysis

Seismic vulnerability analysis is a crucial step in the next-generation performance-based seismic design and a natural extension of the performance-based design philosophy. Seismic vulnerability typically refers to the probability of a structure reaching a certain state of damage under a known seismic intensity. It quantitatively expresses the structural seismic performance from a probabilistic perspective, reflecting the probabilistic relationship between seismic intensity and the extent of structural damage.
There are four main methods for conducting structural seismic vulnerability analysis: judgment methods, empirical methods, experimental methods, and numerical analysis methods. Judgment methods involve a broad assessment and judgment of different types of structural damage within a region based on the expertise of experts and engineers. The concept was introduced by American scholar Whiteman in 1973, who proposed the Damage Probability Matrix (DPM) method. Empirical methods use a large amount of observed structural damage data to predict the probability of various levels of damage occurring under different seismic intensity levels, resulting in empirical vulnerability curves.
Experimental methods establish physical structural models and study vulnerability through extensive experimental testing. However, these methods are often limited by sample quantity, laboratory and equipment conditions. Considering the seismic conditions of dams, especially high concrete faced rockfill dams, the first three methods are generally less applicable. Damages to high concrete faced rockfill dams are rare, and the complexity of dam construction materials, loading conditions, and boundary conditions makes it difficult to simulate realistic damage through laboratory experiments.
Therefore, numerical analysis methods should become an effective approach for vulnerability analysis, including that of high concrete faced rockfill dams. Numerical analysis is widely used in the field of structural seismic performance research. It involves constructing numerical analysis models using finite element methods, selecting actual or artificially generated seismic motions, and conducting numerous numerical simulations to obtain the seismic response of structures. This information is then used to derive vulnerability analysis curves. While this method is widely used in large civil engineering projects such as buildings, bridges, concrete dams, and nuclear power structures, its application to earth-rock dams, especially high concrete faced rockfill dams, has been limited. The detailed analysis process can be represented by Fig. 2.22.
The results of seismic vulnerability analysis for structures are typically presented in two ways: vulnerability curves and vulnerability matrices. Figure 2.23 illustrates a typical form of vulnerability curve for structures, and Table 2.5 presents a typical vulnerability matrix obtained by Liang et al. through the analysis of dam damage data from the Wenchuan earthquake in Mianyang City.
Table 2.5
Damage probability matrix for earth-rock dams in Mianyang
Seismic damage condition
Loss ratio range/%
Median loss ratio/%
The probability of dam failure at different intensity levels/%
VI
VII
VIII
IX
Intact
0–10
5
49.77
29.17
1.75
0
Moderate risk situation
10–20
15
39.73
44.55
62.81
33.33
High risk situation
20–50
35
10.05
22.76
27.02
44.44
Dam breach hazard
50–70
60
0.46
3.53
8.42
22.22
Earthquake destruction
70–100
85
0
0
0
0

2.8 Conclusion

This chapter provides an overview of uncertainty factors and some probabilistic analysis methods in seismic response analysis of earth-rock dams. These include traditional methods like the first and second-order moment method, Monte Carlo method, and response surface method. The chapter also covers the recently popular research topics, such as the probability density evolution method based on stochastic vibration theory and non-stationary stochastic seismic motion models. Additionally, it introduces the vulnerability analysis process suitable for performance-based seismic safety assessment of high concrete faced rockfill dams.
From the above research, it's evident that there is limited work related to probabilistic analysis of earth-rock dams, particularly using stochastic dynamic time history analysis methods, and rare studies focused on high concrete faced rockfill dams. The uncertainty factors in earth-rock dams often consider stochastic seismic excitations and often overlook the influence of dam construction materials, which should not be neglected. Few studies fully consider the stochastic process and dynamic probability of seismic response in earth-rock dams.
Furthermore, this chapter elaborately explains the Generalized Probability Density Evolution Method and its application and solving process, as well as the process of generating stochastic seismic motion and high-dimensional random sample parameters. It establishes a non-stationary stochastic seismic motion model based on the latest seismic design code spectrum, and discretely generates seismic acceleration sample time histories with rich probabilistic characteristics. Through various equivalent linear stochastic dynamic and probabilistic analyses including analytical solutions, Duffing oscillator, multi-layered soil slopes, and panel-stacked rockfill dams, the combination of stochastic seismic motion and high-dimensional random parameter generation methods with the Generalized Probability Density Evolution Method is validated. This approach not only exhibits high efficiency but also guarantees a high level of accuracy in analyzing the random dynamic response and probabilistic behavior of complex geological and engineering structures. This paves the way for subsequent research in the field of elastoplastic stochastic dynamic analysis and probabilistic assessment for high concrete faced rockfill dams, as well as for performance-based seismic safety evaluations, laying a solid foundation.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), 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 license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's Creative Commons license 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.
Literature
go back to reference Box GEP, Wilson KB (1951) On the experimental attainment of optimum conditions. J Roy Stat Soc: Ser B (methodol) 13:1–38MathSciNet Box GEP, Wilson KB (1951) On the experimental attainment of optimum conditions. J Roy Stat Soc: Ser B (methodol) 13:1–38MathSciNet
go back to reference Bucher CG, Bourgund U (1990) A fast and efficient response surface approach for structural reliability problems. Struct Saf 7:57–66 Bucher CG, Bourgund U (1990) A fast and efficient response surface approach for structural reliability problems. Struct Saf 7:57–66
go back to reference Cacciola P, Deodatis G (2011) A method for generating fully non-stationary and spectrum-compatible ground motion vector processes. Soil Dyn Earthq Eng 31:351–360 Cacciola P, Deodatis G (2011) A method for generating fully non-stationary and spectrum-compatible ground motion vector processes. Soil Dyn Earthq Eng 31:351–360
go back to reference Chen JB, Li J (2014) Probability density evolution method for stochastic seismic response and reliability of structures. Eng Mechan 31:1–10 Chen JB, Li J (2014) Probability density evolution method for stochastic seismic response and reliability of structures. Eng Mechan 31:1–10
go back to reference Chen GX, Xie JF, Zhang KX (1995) One-dimensional random seismic response analysis method for heterogeneous earth dams. Eng Reliab 35–38 Chen GX, Xie JF, Zhang KX (1995) One-dimensional random seismic response analysis method for heterogeneous earth dams. Eng Reliab 35–38
go back to reference Chen JB, Li J (2009) A note on the principle of preservation of probability and probability density evolution equation. Probab Eng Mech 24:51–59 Chen JB, Li J (2009) A note on the principle of preservation of probability and probability density evolution equation. Probab Eng Mech 24:51–59
go back to reference Chen JB, Yang JY, Li J (2016) A GF-discrepancy for point selection in stochastic seismic response analysis of structures with uncertain parameters. Struct Saf 59:20–31 Chen JB, Yang JY, Li J (2016) A GF-discrepancy for point selection in stochastic seismic response analysis of structures with uncertain parameters. Struct Saf 59:20–31
go back to reference Cho SE (2009) Probabilistic stability analyses of slopes using the ANN-based response surface. Comput Geotech 36:787–797 Cho SE (2009) Probabilistic stability analyses of slopes using the ANN-based response surface. Comput Geotech 36:787–797
go back to reference Clough RW (1993) Dynamics of structures, 2nd edn. McGraw-Hill, New York Clough RW (1993) Dynamics of structures, 2nd edn. McGraw-Hill, New York
go back to reference Das PK, Zheng Y (2000) Cumulative formation of response surface and its use in reliability analysis. Probab Eng Mech 15:309–315 Das PK, Zheng Y (2000) Cumulative formation of response surface and its use in reliability analysis. Probab Eng Mech 15:309–315
go back to reference Deng J, Gu DS, Li XB (2005) Structural reliability analysis for implicit performance functions using artificial neural network. Struct Saf 27:25–48 Deng J, Gu DS, Li XB (2005) Structural reliability analysis for implicit performance functions using artificial neural network. Struct Saf 27:25–48
go back to reference Du XL, Chen HQ (1994) Random simulation and its parameter determination method of earthquake ground motion. Earthq Eng Eng Vib 1–5 Du XL, Chen HQ (1994) Random simulation and its parameter determination method of earthquake ground motion. Earthq Eng Eng Vib 1–5
go back to reference Duncan JM, Chang CY (1970) Nonlinear Anal Stress Strain Soils. J Soil Mechan Found Div 96:1629–1653 Duncan JM, Chang CY (1970) Nonlinear Anal Stress Strain Soils. J Soil Mechan Found Div 96:1629–1653
go back to reference El Hami A, Radi B (2016) Stochastic dynamics of structures. Wiley El Hami A, Radi B (2016) Stochastic dynamics of structures. Wiley
go back to reference Elman H, Silvester D, Wathen A (2014) Finite elements and fast iterative solvers: with applications in incompressible fluid dynamics, 2nd edn. Oxford University Press Elman H, Silvester D, Wathen A (2014) Finite elements and fast iterative solvers: with applications in incompressible fluid dynamics, 2nd edn. Oxford University Press
go back to reference Fujimura K, Der Kiureghian A (2007) Tail-equivalent linearization method for nonlinear random vibration. Probab Eng Mech 22:63–76 Fujimura K, Der Kiureghian A (2007) Tail-equivalent linearization method for nonlinear random vibration. Probab Eng Mech 22:63–76
go back to reference Haian L (2013) Research on seismic damage prediction and rapid evaluation method of earth-rock dam. Int J Earthq Dyn 44–45 Haian L (2013) Research on seismic damage prediction and rapid evaluation method of earth-rock dam. Int J Earthq Dyn 44–45
go back to reference Hardin BO, Drnevich VP (1972) Shear modulus and damping in soils: design equations and curves. J Soil Mechan Found Div 98:667–692 Hardin BO, Drnevich VP (1972) Shear modulus and damping in soils: design equations and curves. J Soil Mechan Found Div 98:667–692
go back to reference Hasofer A, Lind N (1974) Exact and invariant second-moment code format. J Eng Mechan Div-ASCE 100:111–121 Hasofer A, Lind N (1974) Exact and invariant second-moment code format. J Eng Mechan Div-ASCE 100:111–121
go back to reference Housner GW (1947) Characteristics of strong-motion earthquakes. Bull Seismol Soc Am 37 Housner GW (1947) Characteristics of strong-motion earthquakes. Bull Seismol Soc Am 37
go back to reference Hughes TJR, Franca LP, Balestra M (1986) A new finite element formulation for computational fluid dynamics: V. circumventing the Babuška-Brezzi condition: a stable Petrov-Galerkin formulation of the Stokes problem accommodating equal-order interpolations. Comput Method Appl Mechan Eng 59:85–99 Hughes TJR, Franca LP, Balestra M (1986) A new finite element formulation for computational fluid dynamics: V. circumventing the Babuška-Brezzi condition: a stable Petrov-Galerkin formulation of the Stokes problem accommodating equal-order interpolations. Comput Method Appl Mechan Eng 59:85–99
go back to reference Kanai K (1957) Semi-empirical formula for the seismic characteristics of ground. Trans Archit Inst Jpn 57(1):281–284 Kanai K (1957) Semi-empirical formula for the seismic characteristics of ground. Trans Archit Inst Jpn 57(1):281–284
go back to reference Kartal ME, Bayraktar A, Başağa HB (2010) Seismic failure probability of concrete slab on CFR dams with welded and friction contacts by response surface method. Soil Dyn Earthq Eng 30:1383–1399 Kartal ME, Bayraktar A, Başağa HB (2010) Seismic failure probability of concrete slab on CFR dams with welded and friction contacts by response surface method. Soil Dyn Earthq Eng 30:1383–1399
go back to reference Kaymaz I, McMahon CA (2005) A response surface method based on weighted regression for structural reliability analysis. Probab Eng Mech 20:11–17 Kaymaz I, McMahon CA (2005) A response surface method based on weighted regression for structural reliability analysis. Probab Eng Mech 20:11–17
go back to reference Keng HL, Yuan W (1981) Applications of number theory to numerical analysis. Heidelberg, Berlin Keng HL, Yuan W (1981) Applications of number theory to numerical analysis. Heidelberg, Berlin
go back to reference Li J, Chen JB (2003) Probability density evolution method for analysis of stochastic structural dynamic response. Theor Appl Mechan 437–442 Li J, Chen JB (2003) Probability density evolution method for analysis of stochastic structural dynamic response. Theor Appl Mechan 437–442
go back to reference Li J, Chen JB (2010) Advances in the research on probability density evolution equations of stochastic dynamical systems. Adv Mechan 40:170–188 Li J, Chen JB (2010) Advances in the research on probability density evolution equations of stochastic dynamical systems. Adv Mechan 40:170–188
go back to reference Li J, Chen JB (2017) Some new advances in the probability density evolution method. Appl Math Mechan 38:32–43+2 Li J, Chen JB (2017) Some new advances in the probability density evolution method. Appl Math Mechan 38:32–43+2
go back to reference Lin JH, Zhong WX (1998) Some notes on FEM and structural random response analysis. Chin J Comput Mechan 94–100 Lin JH, Zhong WX (1998) Some notes on FEM and structural random response analysis. Chin J Comput Mechan 94–100
go back to reference Liu HL (1996) Permanent deformation of foundation and embankment dam due to stochastic seismic excitation. Chin J Geotech Eng 19–27 Liu HL (1996) Permanent deformation of foundation and embankment dam due to stochastic seismic excitation. Chin J Geotech Eng 19–27
go back to reference Liu LB (2013) Reliability analysis of nonlinear stochastic structures based on probability density evolution (Master's thesis). Dalian University of Technology Liu LB (2013) Reliability analysis of nonlinear stochastic structures based on probability density evolution (Master's thesis). Dalian University of Technology
go back to reference Liu ZJ, Fang X (2012) Stochastic earthquake response and seismic reliability analysis of large-scale aqueduct structures. J Yangtze River Sci Res Inst 29:77–81 Liu ZJ, Fang X (2012) Stochastic earthquake response and seismic reliability analysis of large-scale aqueduct structures. J Yangtze River Sci Res Inst 29:77–81
go back to reference Liu HL, Lu ZQ, Qian JH (1996) Nonlinear random response and dynamic reliability analysis of earth-rock dam. J Hohai Univ 105–109 Liu HL, Lu ZQ, Qian JH (1996) Nonlinear random response and dynamic reliability analysis of earth-rock dam. J Hohai Univ 105–109
go back to reference Liu ZJ, Lei YL, Fang X (2013) Probability density evolution method-based random seismic response analysis of cable-stayed bridges. Chin Civil Eng J 46:226–232 Liu ZJ, Lei YL, Fang X (2013) Probability density evolution method-based random seismic response analysis of cable-stayed bridges. Chin Civil Eng J 46:226–232
go back to reference Liu ZJ, Zeng B, Zhou YH et al (2014) Probabilistic model of ground motion processes and seismic dynamic reliability analysis of the gravity dam. J Hydraul Eng 45:1066–1074 Liu ZJ, Zeng B, Zhou YH et al (2014) Probabilistic model of ground motion processes and seismic dynamic reliability analysis of the gravity dam. J Hydraul Eng 45:1066–1074
go back to reference Luo XF, Li X, Zhou J et al (2012) A Kriging-based hybrid optimization algorithm for slope reliability analysis. Struct Saf 34:401–406 Luo XF, Li X, Zhou J et al (2012) A Kriging-based hybrid optimization algorithm for slope reliability analysis. Struct Saf 34:401–406
go back to reference Ou JP, Niu DT (1990) Parameters in the random process models of earthquake ground motion and their effects on the response of structures. J Harb Univ Civil Eng Archit 24–34 Ou JP, Niu DT (1990) Parameters in the random process models of earthquake ground motion and their effects on the response of structures. J Harb Univ Civil Eng Archit 24–34
go back to reference Ou J, Wang G (1998) Random vibration of structures. Higher Education Press, Beijing Ou J, Wang G (1998) Random vibration of structures. Higher Education Press, Beijing
go back to reference Radović I, Sobol’ IM, Tichy RF (1996) Quasi-Monte Carlo methods for numerical integration: comparison of different low discrepancy sequences. Monte Carlo Methods Appl 2:1–14 Radović I, Sobol’ IM, Tichy RF (1996) Quasi-Monte Carlo methods for numerical integration: comparison of different low discrepancy sequences. Monte Carlo Methods Appl 2:1–14
go back to reference Reale C, Xue J, Gavin K (2016) System reliability of slopes using multimodal optimization. Géotechnique 66:413–423 Reale C, Xue J, Gavin K (2016) System reliability of slopes using multimodal optimization. Géotechnique 66:413–423
go back to reference Sanchez Lizarraga H, Lai CG (2014) Effects of spatial variability of soil properties on the seismic response of an embankment dam. Soil Dyn Earthq Eng 64:113–128 Sanchez Lizarraga H, Lai CG (2014) Effects of spatial variability of soil properties on the seismic response of an embankment dam. Soil Dyn Earthq Eng 64:113–128
go back to reference Sekar P, Narayanan S (1994) Periodic and chaotic motions of a square prism in cross-flow. J Sound Vib 170:1–24 Sekar P, Narayanan S (1994) Periodic and chaotic motions of a square prism in cross-flow. J Sound Vib 170:1–24
go back to reference Shao LT, Tang HX, Kong XJ et al (1999) Finite element analysis for slope stability of earth-rock dam under the action of stochastic seismic. J Hydraul Eng 66–71 Shao LT, Tang HX, Kong XJ et al (1999) Finite element analysis for slope stability of earth-rock dam under the action of stochastic seismic. J Hydraul Eng 66–71
go back to reference Wang DB, Liu HL, Yu T et al (2013) Seismic fragility analysis for earth-rock dams based on deformation. Chin J Geotech Eng 35:814–819 Wang DB, Liu HL, Yu T et al (2013) Seismic fragility analysis for earth-rock dams based on deformation. Chin J Geotech Eng 35:814–819
go back to reference Wang ZH, Liu HL, Chen GX (2006) Study on stationary stochastic seismic motion model and Stochastic Earthquake response of earth-rock dam. J Disas Preven Mitig Eng 389–394 Wang ZH, Liu HL, Chen GX (2006) Study on stationary stochastic seismic motion model and Stochastic Earthquake response of earth-rock dam. J Disas Preven Mitig Eng 389–394
go back to reference Westergaard HM (1933) Water pressures on dams during earthquakes. Trans Am Soc Civ Eng 98:418–433 Westergaard HM (1933) Water pressures on dams during earthquakes. Trans Am Soc Civ Eng 98:418–433
go back to reference Wichtmann T, Triantafyllidis T (2013) Effect of uniformity coefficient on G/Gmax and damping ratio of uniform to well-graded quartz sands. J Geotech Geoenviron Eng 139:59–72 Wichtmann T, Triantafyllidis T (2013) Effect of uniformity coefficient on G/Gmax and damping ratio of uniform to well-graded quartz sands. J Geotech Geoenviron Eng 139:59–72
go back to reference Wong FS (1985) Slope reliability and response surface method. J Geotech Eng 111:32–53 Wong FS (1985) Slope reliability and response surface method. J Geotech Eng 111:32–53
go back to reference Wu ZG (1991) Seismic response of soil layers with random material parameters. J Hydraul Eng 1–6 Wu ZG (1991) Seismic response of soil layers with random material parameters. J Hydraul Eng 1–6
go back to reference Yang G, Zhu S (2016) Seismic response of rockfill dams considering spatial variability of rockfill materials via random finite element method. Chin J Geotech Eng 38:1822–1832 Yang G, Zhu S (2016) Seismic response of rockfill dams considering spatial variability of rockfill materials via random finite element method. Chin J Geotech Eng 38:1822–1832
go back to reference Yu SD, Du JH, Xiong XR (1993) The stochastic dynamic reliability analysis of earth dam. J Yangtze River Sci Res Inst 68–72 Yu SD, Du JH, Xiong XR (1993) The stochastic dynamic reliability analysis of earth dam. J Yangtze River Sci Res Inst 68–72
go back to reference Zhang GW, Liu LY (1994) Basic parameter probability and correlation analysis for earth-rock dam materials. Water Resour Hydropower Eng 12–16 Zhang GW, Liu LY (1994) Basic parameter probability and correlation analysis for earth-rock dam materials. Water Resour Hydropower Eng 12–16
go back to reference Zhao HB (2008) Slope reliability analysis using a support vector machine. Comput Geotech 35:459–467 Zhao HB (2008) Slope reliability analysis using a support vector machine. Comput Geotech 35:459–467
go back to reference Zhu WQ (1993) Stationary solutions of stochastically excited dissipative Hamiltonian systems. Chin J Theor Appl Mechan 676–684 Zhu WQ (1993) Stationary solutions of stochastically excited dissipative Hamiltonian systems. Chin J Theor Appl Mechan 676–684
go back to reference Zolfaghari AR, Heath AC, McCombie PF (2005) Simple genetic algorithm search for critical non-circular failure surface in slope stability analysis. Comput Geotech 32:139–152 Zolfaghari AR, Heath AC, McCombie PF (2005) Simple genetic algorithm search for critical non-circular failure surface in slope stability analysis. Comput Geotech 32:139–152
Metadata
Title
Probability Analysis Method of Seismic Response for Earth-Rockfill Dams
Authors
Bin Xu
Rui Pang
Copyright Year
2025
Publisher
Springer Nature Singapore
DOI
https://doi.org/10.1007/978-981-97-7198-1_2