Home Geological earthquake simulations generated by kinematic heterogeneous energy-based method: Self-arrested ruptures and asperity criterion
Article Open Access

Geological earthquake simulations generated by kinematic heterogeneous energy-based method: Self-arrested ruptures and asperity criterion

  • Patricio Venegas-Aravena EMAIL logo
Published/Copyright: August 21, 2023
Become an author with De Gruyter Brill

Abstract

The lack of clarity regarding slip distribution within heterogeneous rupture areas has a significant impact on characterizing the seismic source and the role of heterogeneities in determining ground motion. One approach to understand the rupture process is through dynamic simulations, which require substantial computational resources, thereby limiting our comprehension of seismic rupture processes. Consequently, there is a need for methods that efficiently describe the spatial complexities of seismic rupture in a realistic manner. To address this, the statistics of real self-arrested ruptures that conform to the asperity criterion are investigated. This research demonstrates that power law distributions can describe the final slip statistics. Regarding the computational efficiency, a simple heterogeneous energy-based (HE-B) method is proposed. The HE-B method is characterized by the spatial correlation between the rupture parameters, such as the final slip or the rupture velocity, and the distribution of residual energy which determines the zones where the rupture could occur. In addition, the HE-B method defines the rupture area in those zones of the fault where the coupling function exceeds the energy required for rupture initiation. Therefore, the size of the earthquake is directly influenced by the distribution of coupling within faults. This method also leads to the successful reproduction of the statistical characteristics of final slip and generates slip rates that match the kinematic behavior of seismic sources. Notably, this kinematic rupture simulation produces seismic moment rates characterized by f 1 and f 2 spectra with a double corner frequency. Finally, it is observed that the maximum fracture energy value within the ruptured area is strongly correlated with both the magnitude and peak seismic moment rate. Thus, by employing this method, realistic rupture scenarios can be generated efficiently, enabling the study of spatial correlations among rupture parameters, ground motion simulations, and quantification of seismic hazard.

1 Introduction

The study of spatial and temporal properties of seismic source parameters allows us to assess the radiated energy, amplitude, and frequency of ground motion [14]. Two of the most relevant source parameters are the slip rate u ̇ and final slip u f . This is due to the former constraining the radiated energy while the latter controls the spatial variability of self-arrested ruptures.

The slip rate u ̇ is proportional to the seismic moment rate M ̇ 0 which implies that sudden increase in u ̇ generates a burst of released seismic energy [5]. Thus, the consideration of realistic u ̇ is of uttermost importance because its proper consideration in heterogeneous faults implies better predictions of strong ground motion [68]. One way to describe u ̇ properly is by means of the Yoffe function [9]. Despite a few problems, such as indetermination, it has been proposed that the truncated version of the Yoffe function is consistent with earthquake dynamics for finite slip-weakening distances [10].

Second, the spatial domain that constrains where u ̇ evolves could be determined by the distribution of u f . Nevertheless, the proper distribution of u f is still a matter of debate. For example, several authors assume that u f is well described by stochastic von Karman correlation function in dynamic simulations [11,12]. Contrarily, other studies show that the complexity of u f can be well described by power law (fractal) distributions [13]. Due to the fact that one of the main features of fractal model is the lack of any characteristic length, and as von Karman model is characterized by it, the latter is commonly used in simulations [14]. Nonetheless, the usage of von Karman in detriment of fractal model is based on oversimplified distribution of u f or non-physical ruptures that cover the domain totally [14,15,16]. In contrast, the spatial complexity of u f from real earthquakes is associated more with the self-similarity of fractal distributions [17,18,19]. Besides, the self-similarity has also been found in the roughness of exhumed faults [20], in experiments describing the relative displacement of rock samples [21], stress states in systems of faults [22], or in the scale invariant path of cracks [23]. Therefore, self-similarity not only describes u f but is also a characteristic of various rupture parameters, such as mechanical properties, friction, shear resistance, fault proportions, source frequency content, initial stress concentration, changes in the b-value, transition to crack-like rupture, or rupture velocity scaling [2434].

Special attention should be given to the zones that represent large slips, referred to as asperities, as they may also exhibit a distribution following power laws [35]. As the asperity zone accounts for about a quarter to a third of the total rupture area (asperity criterion) [36,37], it is expected that the total u f could be described as a power law as well. Thus, the abovementioned constraints support the idea that u f could be better described as power laws for realistic self-arrested ruptures that match the asperity criteria. Nevertheless, the non-consideration of these spatial heterogeneities leads to simulations of ruptures on flat surfaces which are considerably sensitive to the initial conditions [38]. This increases the difficulty to obtain simulations of self-arrested ruptures characterized by self-similar properties. Some computational simulations of dynamic ruptures have solved the problem of obtaining self-arrested ruptures. Nonetheless, the sensitivity of the initial conditions makes it difficult to prescribe any rupture magnitude. This implies that large self-arrested ruptures tend to be rare. For example, Ide and Aochi [39] have shown that the generation of a single self-arrested rupture of large magnitude requires thousands of small magnitude ruptures. Then, the generation of realistic self-arrested ruptures of specific prescribed magnitudes is also required.

In order to incorporate the abovementioned phenomena, Section 2 analyses the basis statistics of fractal distributions that match the asperity criteria. Section 3 develops an efficient model to generate infinite synthetic ruptures that are statistically equivalent to the dynamic self-arrested scenarios. The incorporation of truncated Yoffe’s function is also given in Section 3. Statistical analysis of the energy-based method constrained with fractal final slip for several examples is described in Section 4. Discussion and conclusion are provided in Sections 5 and 6, respectively.

2 Properties of random power law surfaces

As the u f distributions could be regarded as fractal, it is worthwhile to describe the statistics of two random normalized fractal surfaces that can be created by using the methodology developed by Chen and Yang [40]. Note that this method allows for the generation of fractal surfaces where patches of maximum values are not necessarily located in the same spatial regions. This indicates that the final distribution of slip can have more than one asperity. The first one (Figure 1a) represents the case when the rupture reaches the domain boundaries (runaway rupture) and where its histogram is shown in Figure 1b. The second one (Figure 1c) represents the case of ruptures that are arrested before any section of the rupture front reaches the boundaries (self-arrested ruptures) and its histogram is shown in Figure 1d. The truncated version of the runaway case involves considering values between 0.5 and 1 (a.u.) from the original runaway case. Any values below 0.5 (a.u.) are set as zero, which is visualized as dark blue areas in Figure 1c. This selection leads to a concentration of positive final slip values in the central region of the domain, without extending to the domain’s boundaries. Note that a.u. stands for arbitrary units as they simplify calculations and allow for relative comparisons between variables. For both cases, the fractal dimension D is 2.2. The minimum value for the runaway-like and self-arrested-like cases are 0.0044 and 0.5 (a.u.), respectively. The histogram of the truncated fractal surface (red solid line) is compared to the runaway-like (black solid line) in Figure 1b. The histogram of the runaway-like surface is obtained by considering 20 bins from 0.0044 up to 1 (a.u.) while the truncated surfaces also consider 20 bins but from 0.5 to 1 (a.u.). The difference in the truncation is manifested as a short horizontal axis (red solid line between dashed magenta vertical lines in Figure 1b). If the horizontal values of the truncated histogram are scaled, that is, expand the range of values from (0.5, 1) to (0.0044, 1) the histogram in Figure 1d is obtained (red solid line). Note that these types of curves resemble second-order power law functions, which are defined as N a x b + c , where N represents the histogram and a , b , and c are constants that correspond to the best fit of the histogram. This can be seen as a dashed blue line in Figure 1d. The area of large values of Figure 1c can be compared to that used for real earthquakes. For instance, according to Somerville et al. [36], the zone of large slip known as asperity corresponds to an area between fourth and third of the total ruptured area. The definition of large slip corresponds to those places where the slip is larger than 1.5 times the average slip. By considering that the truncated and scaled fractal distribution of Figure 1c resembles the final slip distribution, the area that resembles the asperity corresponds to 23.9% of the total area while that resembling the large slip is larger than 0.5343 (a.u.). By considering the power law, the total ruptured area is proportional to the number of points N of the histogram. That is, A = N δ A 0 , where δ A 0 is a constant. Thus, the ratio between the total ( A T ) and asperity ( A asp ) area is as follows:

(1) A asp A T = B 0 ( u max b + 1 u min b + 1 ) + C ( u max   u min ) B 0 ( u max b + 1 u asp b + 1 ) + C ( u max u asp ) ,

where u asp = 1.5 u ̅ and B 0 = a / ( b + 1 ) (Figure 1e). By using the values of a, b, and c (Figure 1d), the ratio represents 22 % of the total ruptured area which is similar to that described by Somerville et al. [36]. Thus, the second order power law describes a basic feature of ruptures. Then, a physical method for obtaining ruptures statistically equivalent to that shown in Figure 1 is required.

Figure 1 
               Difference of power law statistics as to whether it is a runaway or self-arrested case by using a fractal surface generated by the code developed by Chen and Yang [40]. (a) 2D Power law normalized distribution. This scenario resembles the runaway ruptures type because it covers the entire domain. (b) Histogram of the distribution shown in (a) (black solid line) and Figure c (red solid line). (c) Self-arrested type is characterized by a distribution that does not reach boundaries. This distribution corresponds to the truncated version of the original distribution shown in (a). That is, only considering those values larger than 0.5 a.u., the second color bar shows the scaled values. Then, scaling the horizontal axis of the truncated histogram (red solid line shown in (b)). This scale reveals that ∼25% of the area is within the zone of large values (>1.5 times the average). (d) Histogram of the distribution for self-arrested rupture shown in (c). The best fit (blue solid line) for the distribution that resembles the self-arrested case which is a second-order power law. (e) schematic representation of the second-order power law and the area that covers the asperity (A
                  asp).
Figure 1

Difference of power law statistics as to whether it is a runaway or self-arrested case by using a fractal surface generated by the code developed by Chen and Yang [40]. (a) 2D Power law normalized distribution. This scenario resembles the runaway ruptures type because it covers the entire domain. (b) Histogram of the distribution shown in (a) (black solid line) and Figure c (red solid line). (c) Self-arrested type is characterized by a distribution that does not reach boundaries. This distribution corresponds to the truncated version of the original distribution shown in (a). That is, only considering those values larger than 0.5 a.u., the second color bar shows the scaled values. Then, scaling the horizontal axis of the truncated histogram (red solid line shown in (b)). This scale reveals that ∼25% of the area is within the zone of large values (>1.5 times the average). (d) Histogram of the distribution for self-arrested rupture shown in (c). The best fit (blue solid line) for the distribution that resembles the self-arrested case which is a second-order power law. (e) schematic representation of the second-order power law and the area that covers the asperity (A asp).

3 Heterogeneous energy-based (HE-B) method

Dynamic ruptures evolve from no slip up to fractally distributed final slip. Thus, it is relevant to find out the manner to describe the transition from the initial to final state. One way to do this is by considering the kinematic source time functions. For instance, the slip rate function has been described by using the Yoffe function [9]. Nonetheless, the kinetic parameters should be constrained by considering realistic scenarios. One of them could be regarded as the energy-based method [41]. This method considers that the residual energy E res is the result of the available energy Δ W 0 minus the fracture energy G C for linear slip weakening law [41].

(2) E res ( x ) = Δ W 0 ( x ) G C ( x ) ,

where x represents any points within the fault. In addition, those places within the fault where positive values of E res are found are also more prone to generate ruptures. That is, E res ( x ) > 0 implies slip and E res ( x ) 0 implies no slip.

The available energy Δ W 0 depends on a function that describes the heterogeneous initial stress states S 0 ( x ) [41,42]. Then, the available energy can be written as Δ W 0 S 0 . The available energy could also be affected by the existence of asperities which generates large stress coupling C [43]. This implies that the available energy can be regarded as:

(3) Δ W 0 S 0 C ,

where C corresponds to the coupling function that incorporates the influence of large slip deficit within faults. Note that the function C describes the mechanical coupling which corresponds to the areas where the stress is larger than the surrounded area [44]. The fracture energy G C means the energy per unit of area required to spread the rupture even further [45,46] and it could be regarded as a material property which is proportional to the geometrical irregularities within faults [47]. In addition, there is ample evidence indicating that the fracture energy is dependent on the material used. This has been demonstrated by various studies on rock samples which have shown that the fracture energy varies based on whether the material is brittle, ductile, or composite in nature [4850]. Thus, each fault is characterized by their own heterogeneous values of G C . When the other dissipation mechanisms are also included, this energy is known as apparent fracture energy and tends to be related to the final slip [51]. As this geometry of faults has been described as power laws [20] and G C is related to the final slip despite the initial stress states [52], it means that the fracture energy could transfer its spatial properties to u f . That is, the final slip is also a function of the fracture energy and its spatial properties ( u f F ( G C ) ). Examples of S 0 , C , Δ W 0 , G C , and E res distributions per unit of area can be found in Figure 2a–e. The G C and S 0 distributions are power laws generated by the indications shown in the study by Chen and Yang [40]. Note that the maximum of G C is 10 9 J / m 2 because this is the expected order of magnitude obtained for largest earthquake with slip larger than 10 m (Figure 3 in Ke et al. [51] and Figure 5 in Nielsen et al. [52]). The domain size of the distribution is 1,000 × 200 and resembles a fault of 100 and 20 km of length and depth, respectively, where d x = d y = 0.1  km. Note that S 0 has been filtered at its shallower and deeper sections (blue colors in Figure 2a). This allows to obtain no rupture at these depths and allows to define a seismogenic zone (orange to red shades in Figure 2a). The coupling functions arise from heterogeneous stresses within the plates, making it impossible to create them analytically. However, since the coupling distributions of real faults tend to be smoothed, a method is proposed to randomly generate them. That is, the coupling function has been obtained by using the sum of seven bimodal Gaussian distributions where the standard deviation depends on the seismic moment M 0 ( σ = σ ( M 0 ) ). The center of each distribution is randomly located within a specific rectangle of length L M 0 2 / 5 and width W L β [29], where β is a constant. Thus, the coupling function is defined by the prescribed seismic moment of reference. In Figure 2b, an example of the distribution of C for a reference dimension L and W equivalent to a magnitude of 6.8 can be observed. The white dot in Figure 2b indicates the maximum coupling function. The available energy is shown in Figure 2c, and the residual energy is shown in Figure 2e. Here the negative values of E res has been set as zero. Note that the distribution of E res tends to zero at its boundaries. The final slip also tends to be zero at the rupture area boundaries for self-arrested ruptures [53]. This spatial feature shared between E res and u f implies that u f should also depend on E res ( u f F ( G C , E res ) ). Additionally, the rupture velocity v r is positively correlated to the peak slip rate u ̇ p while a weak correlation between u f and u ̇ p have been found [5458]. This also implies that a certain dependence between u f and v r could exist. This leads to the following equation:

(4) u f = F ( G C , E res , v r ) .

Figure 2 
               Example of the generation of HE-B model. (a) The background stress states 
                     
                        
                        
                           
                              
                                 S
                              
                              
                                 0
                              
                           
                        
                        {S}_{0}
                     
                   is generated by a random fractal surface [40] and filtered by a function that forces zero values at the upper and lower boundaries. (b) The coupling factor 
                     
                        
                        
                           C
                        
                        C
                     
                   is considered as a set of normalized bimodal Gaussian distribution where its center is located within a rectangle of width and large governed by the Leonard scaling relations [29]. (c) The available energy 
                     
                        
                        
                           Δ
                           
                              
                                 W
                              
                              
                                 0
                              
                           
                        
                        \text{Δ}{W}_{0}
                     
                   corresponds to the multiplication of 
                     
                        
                        
                           
                              
                                 S
                              
                              
                                 0
                              
                           
                        
                        {S}_{0}
                     
                   and 
                     
                        
                        
                           C
                        
                        C
                     
                  . (d) Another normalized random fractal surface is generated. This represents the fracture energy. The difference between 
                     
                        
                        
                           Δ
                           
                              
                                 W
                              
                              
                                 0
                              
                           
                        
                        \text{Δ}{W}_{0}
                     
                   and 
                     
                        
                        
                           
                              
                                 G
                              
                              
                                 C
                              
                           
                        
                        {G}_{C}
                     
                   is shown in (e). The negative values are not considered. (f) The image depicts the gradient of residual energy 
                     
                        
                        
                           
                              ∣
                              
                                 ∇
                                 
                                    
                                       E
                                    
                                    
                                       res
                                    
                                 
                              
                              ∣
                           
                        
                        | \nabla {E}^{\text{res}}| 
                     
                  . (g) The rupture velocity obtained by using equation (5) for 
                     
                        
                        
                           
                              
                                 a
                              
                              
                                 0
                              
                           
                           =
                           
                              
                                 b
                              
                              
                                 0
                              
                           
                           =
                           −
                           
                              
                                 c
                              
                              
                                 0
                              
                           
                           =
                           1
                        
                        {a}_{0}={b}_{0}=-{c}_{0}=1
                     
                  . The obtained rupture time by using equation (6) is shown in figure (h). Note that the magnitude order of the energy per unit of surface values has been taken from real earthquakes [52] and the coupling function is defined between 0 and 1.
Figure 2

Example of the generation of HE-B model. (a) The background stress states S 0 is generated by a random fractal surface [40] and filtered by a function that forces zero values at the upper and lower boundaries. (b) The coupling factor C is considered as a set of normalized bimodal Gaussian distribution where its center is located within a rectangle of width and large governed by the Leonard scaling relations [29]. (c) The available energy Δ W 0 corresponds to the multiplication of S 0 and C . (d) Another normalized random fractal surface is generated. This represents the fracture energy. The difference between Δ W 0 and G C is shown in (e). The negative values are not considered. (f) The image depicts the gradient of residual energy E res . (g) The rupture velocity obtained by using equation (5) for a 0 = b 0 = c 0 = 1 . The obtained rupture time by using equation (6) is shown in figure (h). Note that the magnitude order of the energy per unit of surface values has been taken from real earthquakes [52] and the coupling function is defined between 0 and 1.

Figure 3 
               Different snapshots that show how the slip rate evolves within the fault and in time. (a) The image shows that the rupture begins at a specific location (maximum 
                     
                        
                        
                           
                              
                                 E
                              
                              
                                 res
                              
                           
                        
                        {E}^{\text{res}}
                     
                  ). That is at a length and depth of ∼20 and ∼10 km, respectively. (b), (c), and (d) display the times when the rupture behaves as an elliptically-shaped pulse-like rupture. The rupture tends to accelerate to the right in (e). This is because the rupture has reached the place where the maximum coupling is located (length and depth of ∼40 and ∼10 km, respectively). The burst of slip rate is still present in (f). (g) and (h) show how the rupture tends to arrest.
Figure 3

Different snapshots that show how the slip rate evolves within the fault and in time. (a) The image shows that the rupture begins at a specific location (maximum E res ). That is at a length and depth of ∼20 and ∼10 km, respectively. (b), (c), and (d) display the times when the rupture behaves as an elliptically-shaped pulse-like rupture. The rupture tends to accelerate to the right in (e). This is because the rupture has reached the place where the maximum coupling is located (length and depth of ∼40 and ∼10 km, respectively). The burst of slip rate is still present in (f). (g) and (h) show how the rupture tends to arrest.

The v r also should tend to zero when it arrives at the boundaries of E res . This implies that a direct spatial relation between v r and E res could also exist. In order to incorporate the spatial variability of E res , the rupture velocity could be written as follows:

(5) v r = a 0 + b 0 E res + c 0 | E res | ,

where a 0 , b 0 , and c 0 are constants. Note that the sign of c 0 should be negative as the rupture onset is set at the place where the largest E res is found (blue star in Figure 2). This implies that v r could only decrease towards the rupture boundaries. The gradient of E res is shown in Figure 2f while the rupture velocity is shown in Figure 2g for a 0 = b 0 = 1 and c 0 = 1 . The initiation of rupture is determined by the rupture time t 0 [59], which is commonly related to the rupture velocity v r through the expression v r 1 / t 0 . Therefore, by integrating this expression, the rupture time can be obtained using the following equation:

(6) t 0 = d 0 + e 0 t ( r ) + f 0 1 v r d x ,

where d 0 , e 0 , and f 0 are constants. The function t ( r ) is the initial rupture time at each point within the fault for circular of elliptical ruptures and represent the homogeneous case. This implies that equation (6) could be regarded as a homogeneous rupture time plus the heterogeneities concentrated in the third term that could accelerate or delay the rupture time. For example, the rupture can easily spread if G C is small, while zones with large G C could take more time to rupture. In other words, G C could rush or delay the onset of the rupture. For radial (circular or elliptical) homogeneous rupture, the second term of equation (6) can be written as e 0 t ( r ) e 0 r . Note that small values of f 0 could generate homogeneous ruptures. In contrast, high values of f 0 can cause certain areas of the fault to initiate rupture before the arrival of the rupture front. Thus, the proper election of this factor could ensure that the heterogeneous rupture front velocity is lower than the super shear regime [6062]. For example, the factor f 0 can be constrained by the time required for the rupture to spread from the nucleation point to the furthest point of the ruptured area. This could be achieved by f 0 R L / ( 0.8 v s ) , where R L is the domain length and v s is the shear velocity ( 3 , 464 m / s ) (Table 1). The example of t 0 can be seen in Figure 2h.

Table 1

Global physical values used for the simulations

Variable Symbol Value Units
S-wave velocity v s 3 , 464 m / s
P-wave velocity v p 6 , 000 m / s
Density ρ 2 , 670 kg / m 3
Shear modulus μ 3.2038 × 10 10 kg / ( ms 2 )
Element length x d x 100 m
Element length y d y 100 m
Time step d t 0.0289 s
Triangular windows t s 0.0433 s
Fault length l x 100 km
Fault width l y 20 km
Kinematic parameters a 0 , b 0 , c 0 1 km / s
d 0 1 s
e 0 0.5 ( 9 M W 43 ) s
f 0 L ref / ( 0.5 v s ) s
g 0 0.5 / τ ¯ R Dimensionless
Reference width w ref L ref C 1 1 / β km
Reference length L ref ( M 0 ref ) 2 / 5 km
Reference Seismic moment M 0 ref 10 7 × 10 ( 1.5 ( M Wref + 10.7 ) ) Nm
Factor F 0 1 sm 4 / J 2
Fractal dimension D 2.2 Dimensionless

Additional considerations should be made in order to obtain realistic kinematic ruptures. For instance, the rupture could be regarded to begin at the maximum value of E res . This consideration set the origin of r . In addition, G C is proportional to the final slip u f which is proportionally related to the rise time τ R [55]. That is: τ R G C . The final time t f at any point in the fault is defined as t 0 + g 0 × τ R , where g 0 is a dimensionless factor that can dilate or constrict the rupture front. That is, small (large) values of g 0 could describe pulse-like (crack-like) ruptures. Note that g 0 has been constrained by 1 / τ ¯ R and implies that large τ R can be reduced by its average in order to obtain values of g 0 1  s which can be regarded as pulse-like ruptures (Table 1) [63]. The slip rate is described by the truncated Yoffe function which corresponds to the convolution between the normal Yoffe function and a triangular function of width t s [10]. The amplitude of the truncated Yoffe’s function is normalized and then integrated in order to obtain the slip in terms of time at each point within the fault. Then, the slip at each point is set as proportional to the equation as follows:

(7) u f = F 0 × E res × G C × v r ,

where F 0 is a constant. Note that equation (7) corresponds to the simplest version of equation (4). Finally, the slip is differentiated in order to obtain the time series of slip rate for each point in the fault. An example of the slip rate distribution can be found in Figure 3a–h. Here the slip rate is shown for different times. The parameters used are shown in Table 1. The rupture begins where E res is maximum. This is shown in Figure 3a where the rupture is concentrated close to the length and depth, 20 and 10 km, respectively. This can be seen as a blue star in Figure 2e. The first stage (Figure 3b–d) shows the ruptured front is mainly elliptically shaped as it was set. Nevertheless, the heterogeneities are relevant for Figure 3e, and f. This is because the rupture is entering the zone where the maximum coupling function is found (length and depth of ∼40 and ∼10 km, respectively, in Figure 3e, and f and in the white point in Figure 2b). The arrest stage is shown in Figure 3g and h where the slip rate tends to zero.

The obtained final slip can be seen in Figure 4a. This final slip matches the asperity criteria because a portion of the rupture area releases most of the total seismic moment. Specifically, 29.1% of the total area (area within the black solid line) releases 70.9% of the total seismic moment. It is possible to observe that the histogram of u f (solid black line in Figure 4b) also tends to be similar to the second-order power laws (red solid line Figure 4b). The seismic moment rate is shown in Figure 4c. It can be seen that the maximum M ̇ 0 is found when t ∼ 12 s. This corresponds to the time when the right side of the ruptured front enters the area characterized by a large coupling (white point in Figure 2b). This can roughly be seen in Figure 3e for the slip rate where a patch of yellow to orange shades cover an area close to a length of 40–47 km at a depth of 9–14 km. Figure 4c shows the impulsive increases in M ̇ 0 for t 10  s (vertical orange dashed line). The duration of this burst of released seismic moment lasts ∼5 s (Figure 4c). Figure 4d shows the frequency amplitude of M ̇ 0 . This spectrum is characterized by a double corner frequency ( f Cc 1 = 0.019 Hz and f c 2 = 0.21 Hz ). The decay f 1 is found between f c 1 and f c 2 and f 2 is found for f > f c 2 . Note that the frequency f 0.068 Hz represents the projection of the spectrum with a decay of f 2 towards the plateau level. This feature of double corner frequency has also been observed for real earthquakes [64,65]. These corner frequencies are related to the total rupture time ∼35 s ( Δ t 1 in Figure 4c) and the time required to rupture a coupled zone ( Δ t 2 5  s). In terms of frequency, this is 1 Δ t 1 0.02 Hz and 1 Δ t 2 0.2 Hz which are the corner frequencies shown in Figure 4d. Note that no consideration has directly been taken in order to constrain the decay of this spectrum. That is, the spectrum appears naturally from the considered heterogeneities and rupture size.

Figure 4 
               Statistical analysis of the example shown in Figures 2 and 3. (a) The final slip shows that the rupture does not reach the domain’s boundaries. The percentage of asperity is ∼29% and releases ∼70% of the total seismic moment. (b) The statistic of the obtained 
                     
                        
                        
                           
                              
                                 u
                              
                              
                                 f
                              
                           
                        
                        {u}_{\text{f}}
                     
                   reveals the second-order power law tendency. (c) The seismic moment rate 
                     
                        
                        
                           
                              
                                 
                                    M
                                    ̇
                                 
                              
                              
                                 0
                              
                           
                        
                        {\dot{M}}_{0}
                     
                   is also obtained. The maximum value is found when 
                     
                        
                        
                           t
                           ∼
                           12
                           
                           s
                        
                        t\sim 12\hspace{.25em}\text{s}
                     
                  . That is the time when the rupture front enters the main coupling (Figure 3e and f). (d) Spectrum of 
                     
                        
                        
                           
                              
                                 
                                    M
                                    ̇
                                 
                              
                              
                                 0
                              
                           
                        
                        {\dot{M}}_{0}
                     
                  . This is characterized by a plateau (yellow segmented line) and a double corner frequency. The decay is 
                     
                        
                        
                           
                              
                                 f
                              
                              
                                 −
                                 1
                              
                           
                        
                        {f}^{-1}
                     
                   (blue segmented line) and 
                     
                        
                        
                           
                              
                                 f
                              
                              
                                 −
                                 2
                              
                           
                        
                        {f}^{-2}
                     
                   (red segmented line).
Figure 4

Statistical analysis of the example shown in Figures 2 and 3. (a) The final slip shows that the rupture does not reach the domain’s boundaries. The percentage of asperity is ∼29% and releases ∼70% of the total seismic moment. (b) The statistic of the obtained u f reveals the second-order power law tendency. (c) The seismic moment rate M ̇ 0 is also obtained. The maximum value is found when t 12 s . That is the time when the rupture front enters the main coupling (Figure 3e and f). (d) Spectrum of M ̇ 0 . This is characterized by a plateau (yellow segmented line) and a double corner frequency. The decay is f 1 (blue segmented line) and f 2 (red segmented line).

4 Examples

Let us consider two sets of kinematic simulations to account for the effects of changes in fault’s heterogeneities. The first set is that when the distribution of G CC and S 0 are the same and different distributions of C are considered. The second set is when C and S 0 are the same and changes come from the distribution of G C .

4.1 Changes in the coupling distribution

The different magnitude examples can be done by considering an initial magnitude of reference M W ref . This allows to set the boundaries where the center of each bimodal Gaussian function can be located. Specifically, the boundaries are determined by the Leonard relation [29] which established the scaling between the width and length of rupture in terms of a specific magnitude (Table 1). Note that the standard deviation σ also could be regarded as a function of the magnitude ( σ = σ ( M W ) ). Some examples of different distributions of C are shown in Figure 5a–j. Particularly, the increase in the reference magnitude generates large scaling relations between width and large (Figure 5a, c, e, g, and i) which generates large, ruptured areas (Figure 5b, d, f, h, and j). The full results are shown in Tables 2 and 3. On average, the ruptured area that can be regarded as an asperity is 25.6303% and releases 63.0439% of the total seismic moment (Table 2). This confirms that the spatial properties of u f are self-similar regardless of the distribution of C . Regarding the scaling of frequency spectrum, Figure 6a shows how the plateau increases with the magnitude as well as the corner frequencies decrease. Decay found in these spectra is similar to that found in Section 3. That is, there is a decay of f 1 between the two corner frequencies and a decay of f 2 for the high frequency content. In reference to the distribution of final slip shown in Figure 6b, it is evident that they display a self-similar characteristic. This implies that each distribution can be described by a second-order power law ( a u f b + c ), irrespective of the earthquake magnitude. The varying shades from magenta to black in Figure 6b indicate earthquakes of magnitudes close to the range of 5–7. The seismic moment rate is shown in Figure 7a. Both, the peak M ̇ 0 and rupture duration depends on the magnitude. Large magnitude earthquakes (black shades in Figure 7a) tend to have large max M ̇ 0 and last for more than 25 s. This correlation between max M ̇ 0 and M W can be seen in Figure 7b. In addition, this figure also shows that max M ̇ 0 , M W , and the maximum fracture energy within the ruptured area ( max G C ) are well correlated among them. That is, correlations larger than 0.85. An almost perfect correlation exists between the maximum frequency f max and the average fracture energy ( G ¯ C ) as it is shown in Figure 7c. However, no correlation is found between G ¯ C and M W and between f max and M W . This implies that there are differences between the physical interpretation of G ¯ C and max G C . Besides, f max does not depend on the earthquake magnitude. Figure 7d shows the positive correlation between the area of asperity and the amount of seismic moment that is being released for asperity. The larger the area, the larger the released seismic moment. In contrast, no correlation is found between the latter and former to f max . That is, f max does not depend on the magnitude, the asperity area, and the seismic moment released by it.

Figure 5 
                  Example of increasing heterogeneous magnitude. The full details of the results is listed in Tables 2 and 3. It can be seen that the reference magnitude generates a random coupling function that satisfies the scaling relation proposed by Leonard [29]. This implies that each 
                        
                           
                           
                              
                                 
                                    E
                                 
                                 
                                    res
                                 
                              
                           
                           {E}^{\text{res}}
                        
                      is restricted to a specific area and generates a rupture of similar magnitude. (a) Displays the coupling functions for a reference magnitude of Mw5, while (b) shows the resulting final slip. Similarly, (c), (e), (g), and (i) display the couplings for reference earthquakes with magnitudes of 5.5, 6.0, 6.5, and 7.0 respectively. The final slips for each of these coupling functions are shown in (d), (f), (h), and (j) respectively.
Figure 5

Example of increasing heterogeneous magnitude. The full details of the results is listed in Tables 2 and 3. It can be seen that the reference magnitude generates a random coupling function that satisfies the scaling relation proposed by Leonard [29]. This implies that each E res is restricted to a specific area and generates a rupture of similar magnitude. (a) Displays the coupling functions for a reference magnitude of Mw5, while (b) shows the resulting final slip. Similarly, (c), (e), (g), and (i) display the couplings for reference earthquakes with magnitudes of 5.5, 6.0, 6.5, and 7.0 respectively. The final slips for each of these coupling functions are shown in (d), (f), (h), and (j) respectively.

Table 2

Values obtained by different simulations of different magnitudes

M Wref M W u som   ( m ) Asperity  ( % ) M 0   ( % ) a b c
5.0 5.0519 0.1862 24.4611 63.6338 1600.1427 −0.0170 −1614.8813
5.1 5.1103 0.1982 25.5094 63.5403 665.4257 −0.0424 −679.9147
5.2 5.0997 0.1958 26.8170 62.2115 −640.9314 0.0519 623.7332
5.3 5.1963 0.2190 24.5649 59.1244 −683.2097 0.0601 665.5470
5.4 5.3888 0.2716 24.7119 59.3898 4948.3241 −0.0137 −4943.67243
5.5 5.4056 0.2780 25.0309 59.6445 −1937.4294 0.0343 1929.6274
5.6 5.5770 0.3380 25.2357 60.5781 −4680.2396 0.0214 4694.5254
5.7 5.6702 0.3759 25.3745 61.9355 −1205212.6361 0.0001 1205240.3304
5.8 5.8517 0.4618 25.7776 62.6430 −19169.1590 0.0093 19244.7341
5.9 5.9808 0.5357 25.6975 62.6764 165862.0483 −0.0014 −165720.8613
6.0 6.0114 0.5547 24.6447 59.8050 −9725.3506 0.0281 99071.4110
6.1 6.1643 0.6624 27.0784 62.0645 −7185.3387 0.0492 7445.8838
6.2 6.2925 0.7654 25.1499 60.2559 −23764.3912 0.0225 24345.8120
6.3 6.4236 0.8916 24.9902 62.5688 −46234.8098 0.0131 46837.8276
6.4 6.5612 1.0399 24.3248 61.3599 164746.2798 −0.0058 −163456.1167
6.5 6.6942 1.2126 26.0065 65.3109 25741.0613 −0.0420 −24332.0990
6.6 6.8592 1.4651 25.4644 66.4560 23729.0286 −0.0626 −21611.2308
6.7 6.9052 1.5466 24.7817 61.6741 −136758.5092 0.0130 139540.9435
6.8* 7.0814 1.8926 29.0826 71.7570 15015.9276 −0.1293 −11921.6992
6.9 7.1841 2.1309 26.5846 68.8305 51295.4564 −0.0557 −46383.1040
7.0 7.1973 2.1651 26.9484 68.4622 38802.5636 −0.0721 −33986.6888
Average 25.6303 63.0439

These results are obtained by changing the coupling function and letting the other parameters remain the same. Besides the reference magnitude M Wref is used to constrain Δ W 0 , while the other results listed are the real magnitude M W , the Somerville average slip is defined as the area where u f is larger than 1.5 × u ¯ f [36], percentage of rupture area that corresponds to the asperity, amount of seismic moment released by the asperity and the second-order power law that fit the histogram for each rupture (a, b and c). The *means that this is the simulation used as example in Section 3.

Table 3

Other results obtained by different simulations of different magnitudes (same simulations as those presented in Table 2)

M Wref M W u ̇ p   ( m / s ) g 0   f max   ( Hz ) G ¯ C   × 10 8 J m 2 G Cmax × 10 8 J m 2 e 0   ( s )
5.0 5.0519 2.2139 1.0185 8.9552 5.8913 6.9557 1
5.1 5.1103 2.4226 1.1491 7.9372 5.2216 6.7075 1.4500
5.2 5.0997 2.2840 0.8135 11.2113 7.3754 8.3420 1.9000
5.3 5.1963 2.5343 0.7384 12.3519 8.1259 9.1527 2.3500
5.4 5.3888 4.4793 0.9942 9.1739 6.0351 6.9125 2.8000
5.5 5.4056 3.4795 0.7546 12.0857 7.9507 9.1527 3.2500
5.6 5.5770 4.9094 0.8662 10.5293 6.9268 8.3879 3.7000
5.7 5.6702 5.2705 0.9275 9.8335 6.4691 8.6006 4.1500
5.8 5.8517 5.9096 0.9729 9.3749 6.1674 8.0444 4.6000
5.9 5.9808 6.9127 0.9648 9.4533 6.2190 7.7943 5.0500
6.0 6.0114 7.8100 0.9561 9.5391 6.2754 9.2076 5.5000
6.1 6.1643 9.7900 0.9028 10.1028 6.6462 10.3750 5.9500
6.2 6.2925 12.9623 0.9120 10.0000 6.5786 9.1767 6.4000
6.3 6.4236 9.2945 0.8301 10.9877 7.2283 10.1323 6.8500
6.4 6.5612 17.5137 0.8888 10.2613 6.7505 9.5637 7.3000
6.5 6.6942 11.5334 0.8027 11.3621 7.4747 10.9065 7.7500
6.6 6.8592 16.3985 0.8879 10.2718 6.7574 10.0888 8.2000
6.7 6.9052 19.2114 0.9074 10.0508 6.6120 11.4204 8.6500
6.8* 7.0814 15.7329 0.8766 10.4049 6.8450 11.5286 9.1000
6.9 7.1841 16.5019 0.8976 10.1614 6.6848 12.0000 9.5500
7.0 7.1973 18.7485 0.9036 10.0930 6.6398 11.7488 10

This table shows the maximum slip rate of the rupture u ̇ p , g 0 , the maximum frequency f max , the average fracture energy G ¯ C , the maximum fracture energy G Cmax , and e 0 .

Figure 6 
                  (a) The frequency spectra and (b) the power law histogram for ruptures shown in Tables 2 and 3. The only change is in the coupling function. Color gradient means the magnitude. Magenta is for those magnitude close to 5 and the black means magnitude close to 7.2. It can be seen that the basic properties of real earthquakes are obtained. For instance, the amplitude of the plateau is large for large earthquakes and its corner frequency tend to be smaller (a). (b) the histogram for each rupture. The larger the earthquake, the larger the final slip and the number of points (area). However, the same second-order power law is obtained regardless of the magnitude.
Figure 6

(a) The frequency spectra and (b) the power law histogram for ruptures shown in Tables 2 and 3. The only change is in the coupling function. Color gradient means the magnitude. Magenta is for those magnitude close to 5 and the black means magnitude close to 7.2. It can be seen that the basic properties of real earthquakes are obtained. For instance, the amplitude of the plateau is large for large earthquakes and its corner frequency tend to be smaller (a). (b) the histogram for each rupture. The larger the earthquake, the larger the final slip and the number of points (area). However, the same second-order power law is obtained regardless of the magnitude.

Figure 7 
                  Seismic moment rate and correlation among different rupture parameters. (a) Seismic moment rate of different magnitude ruptures where the ruptures are the same that are described in Tables 2 and 3. Ruptures only differ in the coupling function. Large 
                        
                           
                           
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           {\dot{M}}_{0}
                        
                      (∼7.2) tends to be in black shade and smaller 
                        
                           
                           
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           {\dot{M}}_{0}
                        
                      (∼5) is in magenta shade. (b) Positive correlation among peak seismic moment rate (
                        
                           
                           
                              max
                              
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           \max \hspace{.25em}{\dot{M}}_{0}
                        
                     ), maximum fracture energy (
                        
                           
                           
                              max
                              
                              
                                 
                                    G
                                 
                                 
                                    C
                                 
                              
                           
                           \max \hspace{.25em}{G}_{\text{C}}
                        
                     ), and magnitude (
                        
                           
                           
                              
                                 
                                    M
                                 
                                 
                                    W
                                 
                              
                           
                           {M}_{\text{W}}
                        
                     ). (c) The maximum frequency (
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    max
                                 
                              
                           
                           {f}_{\text{max}}
                        
                     ) is correlated to the average fracture energy (
                        
                           
                           
                              
                                 
                                    
                                       G
                                       ¯
                                    
                                 
                                 
                                    C
                                 
                              
                           
                           {\bar{G}}_{\text{C}}
                        
                     ) but none of them are correlated to the magnitude. (d) The percentage of ruptured area that can be regarded as an asperity tend to be positively correlated to the amount of the released seismic moment. None of them are correlated to the maximum frequency.
Figure 7

Seismic moment rate and correlation among different rupture parameters. (a) Seismic moment rate of different magnitude ruptures where the ruptures are the same that are described in Tables 2 and 3. Ruptures only differ in the coupling function. Large M ̇ 0 (∼7.2) tends to be in black shade and smaller M ̇ 0 (∼5) is in magenta shade. (b) Positive correlation among peak seismic moment rate ( max M ̇ 0 ), maximum fracture energy ( max G C ), and magnitude ( M W ). (c) The maximum frequency ( f max ) is correlated to the average fracture energy ( G ¯ C ) but none of them are correlated to the magnitude. (d) The percentage of ruptured area that can be regarded as an asperity tend to be positively correlated to the amount of the released seismic moment. None of them are correlated to the maximum frequency.

4.2 Changes in the fracture energy distribution

The next set of examples considers that the changes are only performed in the distribution of G C . That is, the same distribution of S 0 and C are used. The results are shown in Tables 4 and 5. Here it is possible to observe that all the cases turned out to be ruptures with similar magnitude (∼7). Figure 8 shows some examples. Note that the number of simulations is located at the upper left corner of each plot in Figure 8 and the full results of the labeled examples can be seen in Tables 4 and 5. It can be seen that all the final slip distributions tend to be quite similar among them. In fact, the asperity area and the seismic moment released there is almost the same for all the cases (average of 28.6848 and 68.9061%, respectively, as shown in Table 4). This indicates that the parameter that drives the size of earthquakes is the coupling function. In addition, note that the maximum coupling (white point in Figure 8), initial rupture point (blue star in Figure 8), and maximum slip (magenta triangle in Figure 8) may not be all located in the same place.

Table 4

The same parameters described in Table 2 are presented. However, these 20 new simulations have the same coupling function and S 0 distribution but different fracture energy distributions G C among them. While the coupling function uses a magnitude 7 M W earthquake as a reference, earthquakes of similar magnitude to the reference are generated regardless of the G C distribution

Number # M W u som   ( m ) Asperity  ( % ) M 0   ( % ) a b c
1 7.0814 1.8926 29.0826 71.7570 15015.9276 −0.1293 −11921.6992
2 7.1004 1.9392 29.6045 69.4418 35906.5182 −0.0559 −32579.4075
3 7.0583 1.8404 27.9023 69.6037 19477.0796 −0.1054 −16217.7947
4 7.0209 1.7679 28.7229 66.8259 34023.6423 −0.0522 −31205.1155
5 7.0735 1.8747 29.2123 71.9647 10651.2396 −0.1716 −7718.0547
6 7.0745 1.8784 27.5368 68.9274 19379.2027 −0.1044 −16145.9520
7 7.1004 1.9392 29.6045 69.4418 35906.5198 −0.0559 −32579.4090
8 7.0597 1.8488 28.9776 69.3402 21215.8819 −0.0854 −18309.9258
9 7.0977 1.9333 29.0871 68.3850 110937.0692 −0.0208 −107090.9096
10 7.0525 1.8294 27.5905 68.7998 25189.6916 −0.0793 −22026.6679
11 7.0827 1.8897 28.3338 73.4226 11593.7537 −0.1666 −8582.2291
12 7.0593 1.8503 29.8987 68.0159 110799.0800 −0.0172 −107663.9327
13 7.0888 1.9127 27.0004 67.6688 42955.4155 −0.0485 −39575.5718
14 7.0652 1.8588 29.0743 70.5901 15105.8194 −0.1177 −12265.9039
15 7.0847 1.9017 29.8438 68.0794 48611.6106 −0.0429 −45130.5152
16 7.0588 1.8463 28.3138 68.9105 24810.5266 −0.0748 −21842.2727
17 7.0747 1.8857 27.6533 65.7931 −58819.5243 0.0364 62400.5053
18 7.0528 1.8380 28.5759 67.1862 78934.4536 −0.0236 −75905.1812
19 7.0663 1.8680 29.6638 67.8802 1640425.8229 −0.0013 −1637030.0899
20 7.0636 1.8620 28.4546 66.0260 −57646.2064 0.0374 61267.2608
21 7.0884 1.9110 28.2483 68.9678 44698.7289 −0.0519 −40877.8246
Average 7.0716 1.8747 28.6848 68.9061 106151.0600 −0.0634 −102904.7948

Nevertheless, the fracture energy is changed. This implies that the coupling function and S 0 distribution are the same. The changes in G C generate earthquakes of similar magnitude ( M W 7 ).

Table 5

Same results as that shown in Table 3 but for the examples described in Table 4

Number # M W u ̇ p   ( m / s ) g 0   f max (Hz) G ¯ C   × 10 8 J m 2 G Cmax × 10 8 J m 2
1 7.0814 15.7329 0.8766 10.4049 6.8450 11.5286
2 7.1004 16.7908 1.1711 12.9430 5.1233 8.7511
3 7.0583 13.6012 0.7832 11.4797 7.6608 11.8212
4 7.0209 13.5083 0.7622 8.8000 7.8715 12.0000
5 7.0735 15.4551 0.8202 13.5270 7.3152 12.0000
6 7.0745 15.3344 0.8776 13.5853 6.8372 11.6546
7 7.1004 17.2497 1.1711 12.9430 5.1233 8.7511
8 7.0597 15.6780 0.8515 20.6447 7.0461 11.2354
9 7.0977 16.2587 1.1177 17.8148 5.3681 8.7329
10 7.0525 12.9551 0.8238 19.5617 7.2832 12.0000
11 7.0827 16.6316 0.7622 45.7667 7.8723 12.0000
12 7.0593 16.3859 1.0190 27.1925 5.8880 9.8759
13 7.0888 15.2155 1.1151 25.6441 5.3806 8.4600
14 7.0652 15.3827 0.8254 15.4226 7.2690 11.1246
15 7.0847 16.8932 1.1785 13.7856 5.0912 9.3091
16 7.0588 13.9613 0.8431 11.4010 7.1163 11.3021
17 7.0747 13.7973 1.0753 21.2124 5.5796 9.8109
18 7.0528 13.3773 0.8097 10.4060 7.4103 12.0000
19 7.0663 14.4638 0.8599 17.6595 6.9773 11.8272
20 7.0636 12.4038 0.8883 16.7349 6.7544 10.6858
21 7.0884 13.3394 0.8572 19.7810 6.9995 12.0000
Figure 8 
                  (a)-(j) showcase examples of ruptures appearing in Tables 4 and 5 when different fractal distributions of fracture energy are used alongside 
                        
                           
                           
                              C
                           
                           C
                        
                      and 
                        
                           
                           
                              
                                 
                                    S
                                 
                                 
                                    0
                                 
                              
                           
                           {S}_{0}
                        
                      distributions that are the same for all cases. (a), (c), (e), (g), and (i) depict the distributions of fracture energy for simulations numbered 2, 6, 11, 16, and 21 (Tables 4 and 5). (b), (d), (f), (h), and (j) display their respective final slips. Note that the change in 
                        
                           
                           
                              
                                 
                                    G
                                 
                                 
                                    C
                                 
                              
                           
                           {G}_{C}
                        
                      results in locations of maximum coupling (white point), rupture onset (blue star), and maximum displacement potentially differing, even though the shapes of the ruptures remain similar among them.
Figure 8

(a)-(j) showcase examples of ruptures appearing in Tables 4 and 5 when different fractal distributions of fracture energy are used alongside C and S 0 distributions that are the same for all cases. (a), (c), (e), (g), and (i) depict the distributions of fracture energy for simulations numbered 2, 6, 11, 16, and 21 (Tables 4 and 5). (b), (d), (f), (h), and (j) display their respective final slips. Note that the change in G C results in locations of maximum coupling (white point), rupture onset (blue star), and maximum displacement potentially differing, even though the shapes of the ruptures remain similar among them.

The change in G C generates frequency content that is almost identical among all the cases (Figure 9a). That is, the same double corner frequency and the double spectrum decay ( f 1 and f 2 ). Similarities are also found in the final slip distribution (Figure 9b). The change in G C does not affect the basic statistics of ruptures. Minor differences can be seen in the seismic moment rate (Figure 10a). It can be observed that there are slight differences in M ̇ 0 due to the change in G C . Nevertheless, the main shape of M ̇ 0 is shared among all the cases. Despite these minor changes, there are some evident changes in the correlation among rupture parameters. For instance, there are weak correlations among max G C , max M ̇ 0 , and M W (Figure 10b) while the correlation between f max and G ¯ C vanished (Figure 10c). Here a weak negative correlation is found between G ¯ C and M W . Figure 10d shows that there is no correlation among the asperity area, the seismic moment released, and f max .

Figure 9 
                  (a) Frequency spectrum of the earthquakes listed in Tables 4 and 5. The spectrum behaves similarly across all the cases. That is, the double corner frequency determined by the double decay (
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    −
                                    1
                                 
                              
                           
                           {f}^{-1}
                        
                      and 
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    −
                                    2
                                 
                              
                           
                           {f}^{-2}
                        
                     ). (b) The histogram reveals that the statistics of final slip remain identical among the cases regardless of the fracture energy distribution.
Figure 9

(a) Frequency spectrum of the earthquakes listed in Tables 4 and 5. The spectrum behaves similarly across all the cases. That is, the double corner frequency determined by the double decay ( f 1 and f 2 ). (b) The histogram reveals that the statistics of final slip remain identical among the cases regardless of the fracture energy distribution.

Figure 10 
                  (a) Seismic moment rate. Regardless of the fracture energy distribution, the main shape of 
                        
                           
                           
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           {\dot{M}}_{0}
                        
                      remains almost the same. (b) A weak negative correlation exists among 
                        
                           
                           
                              max
                              
                              
                                 
                                    G
                                 
                                 
                                    C
                                 
                              
                           
                           \max \hspace{.25em}{G}_{\text{C}}
                        
                      and 
                        
                           
                           
                              max
                              
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           \max \hspace{.25em}{\dot{M}}_{0}
                        
                      (∼ −0.38). A weak positive (negative) correlation exists between 
                        
                           
                           
                              max
                              
                              
                                 
                                    
                                       M
                                       ̇
                                    
                                 
                                 
                                    0
                                 
                              
                           
                           \max \hspace{.25em}{\dot{M}}_{0}
                        
                      and 
                        
                           
                           
                              
                                 
                                    M
                                 
                                 
                                    W
                                 
                              
                           
                           {M}_{\text{W}}
                        
                      (
                        
                           
                           
                              max
                              
                              
                                 
                                    G
                                 
                                 
                                    C
                                 
                              
                           
                           \max \hspace{.25em}{G}_{\text{C}}
                        
                      and 
                        
                           
                           
                              
                                 
                                    M
                                 
                                 
                                    W
                                 
                              
                           
                           {M}_{\text{W}}
                        
                     ). (c) No correlation is found between 
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    max
                                 
                              
                           
                           {f}_{\text{max}}
                        
                      and 
                        
                           
                           
                              
                                 
                                    
                                       G
                                       ¯
                                    
                                 
                                 
                                    C
                                 
                              
                           
                           {\bar{G}}_{C}
                        
                      and between 
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    max
                                 
                              
                           
                           {f}_{\text{max}}
                        
                      and 
                        
                           
                           
                              
                                 
                                    M
                                 
                                 
                                    W
                                 
                              
                           
                           {M}_{\text{W}}
                        
                     . A weak negative correlation is found between 
                        
                           
                           
                              
                                 
                                    
                                       G
                                       ¯
                                    
                                 
                                 
                                    C
                                 
                              
                           
                           {\bar{G}}_{\text{C}}
                        
                      and 
                        
                           
                           
                              
                                 
                                    M
                                 
                                 
                                    W
                                 
                              
                           
                           {M}_{\text{W}}
                        
                     . (d) There is no correlation among the asperity percentage, seismic moment released for that asperity, and 
                        
                           
                           
                              
                                 
                                    f
                                 
                                 
                                    max
                                 
                              
                           
                           {f}_{\text{max}}
                        
                     .
Figure 10

(a) Seismic moment rate. Regardless of the fracture energy distribution, the main shape of M ̇ 0 remains almost the same. (b) A weak negative correlation exists among max G C and max M ̇ 0 (∼ −0.38). A weak positive (negative) correlation exists between max M ̇ 0 and M W ( max G C and M W ). (c) No correlation is found between f max and G ¯ C and between f max and M W . A weak negative correlation is found between G ¯ C and M W . (d) There is no correlation among the asperity percentage, seismic moment released for that asperity, and f max .

5 Summary and discussions

5.1 Examining realistic rupture generation using the HE-B model

Realistic simulations are needed in order to obtain new information regarding the rupture process. That is why simulations must be tested with properties that can be found in real earthquakes. For example, the asperity criterion found by Somerville et al. [36] established that the portion of the fault with large slip (1.5 times the average slip) radiates more than half of the total seismic energy. Note that, for real earthquakes, there could be several separated asperities [66,67]. The analysis shown in Figure 1e, equation (1), and in Section 2 indicates that this criterion is a characteristic of truncated fractal surfaces that might resemble self-arrested ruptures. These heterogeneous final slip distributions are also characterized by histograms that can be described as second order power law functions. Nevertheless, the realistic transition from initial state (no slip) to fractally distributed final slip is required. That is why the HE-B model is proposed in Section 3. The basic idea of the HE-B model came from the energy-based model proposed by Noda et al. [41] (equation (2)). Nonetheless, the HE-B model additionally considers the fact that the rupture parameters of real faults are heterogeneous. This implies that rupture parameters, such as the slip-weakening distance ( D C ), are not constant along the fault because represents fault’s rugosity [47]. As the fracture energy depends on D C , the residual energy E res also should be heterogeneous. That is, it could change from point to point within the fault. The available energy could be regarded as the initial energy that depends on initial shear stress. This implies that Δ W 0 should incorporate physical conditions of faults as the seismogenic zone S 0 [68] or the coupling function C as shown in equation (3). The incorporation of the latter implies that C could be considered as continuous and smooth patches where the stress is larger than the surrounding stress in faults [69]. That is why C is considered as a random distribution of bimodal Gaussian functions in Sections 3 and 4. Note that continuous means smooth and unified patch area. This feature modulates the residual energy which determines the rupture area [41]. As the residual energy determines where the slip occurs, it was reasonable to set the rupture nucleation point where the maximum value of E res is located. Note that E res could change in time for the inter-seismic stage. That is, E res increases due to the increase in C . This implies that there could be areas where E res is positive and no rupture could exist before E res reaches a critical value. As the point of rupture nucleation has been defined as where the maximum E res is located, this location could be interpreted as the first place where the critical value of E res has reached.

5.2 Physically consistent kinematic parameters

The kinematic parameters used here are physically consistent. For instance, the rupture velocity is proportional to E res (equation (5)) which depends on the available energy that has been regarded as proportional to the initial stress states. That is, E ref τ 0 . As τ 0 is Δ τ + τ f , where Δ τ and τ f are the stress drop and final stress, respectively [70], the rupture velocity is directly related to the stress drop ( v r Δ τ ). This relation has been used for pseudo dynamic rupture simulations [70]. This result also allows us to compare energy-based kinematic relations (equations (2) and (5)–(7)) to the relation Δ τ u / τ R 2 , where u is the slip and τ R the rise time [71]. From equation (7), the rupture velocity is v r u / G C E res , while equation (2) ( E res G C ) implies that v r u / G C 2 . As the rise time τ R used here is proportional to the fracture energy, we have v r u / τ R 2 . Thus, the relation used by Kieling et al. [71] is obtained by considering v r Δ τ . Then, the consideration of v r in equation (2), its dependence on E res (equation (5)), and the linearity of equation (7) show that the definitions of the rupture parameters used here are physically consistent. Note that this heterogeneous kinematic model incorporates non-constant rupture velocity, which differs from the Kostrov model that assumes a constant rupture velocity [72]. This new kinematic model is more realistic for considering non-constant rupture velocity, which has been shown to have a significant impact on earthquake dynamics. By accounting for variations in rupture velocity, the new model can better capture the complexities of real earthquakes and provide more accurate predictions of seismic hazard.

Another rupture parameter is the rupture time which depends on the abovementioned physics. Rupture time is shown in equation (6) and defines the causality of the ruptured front. That is, the radial dependency of t 0 impedes that other patches in the fault begins to rupture before the main rupture front reaches that place. In addition, t 0 is inverse to v r which depends on E res . The inverse relation between t 0 and E res described in equations (5) and (6) indicates that those places where the residual energy is larger (smaller) the rupture front accelerates (arrest) which implies that the rupture time is reduced (increases). The incorporation of the factor R L / ( 0.8 v s ) to the t 0 helps to constrain the rupture front below the super shear regime.

Equation (7) shows that the final slip is proportional to E res , v r , and G C because they constrain the slip to reduce its values at the rupture boundaries and to have fractal properties, respectively. This can be interpreted as u f is constrained by the energy, fault roughness, and kinematic evolution. By considering these heterogeneities and causalities, the kinematic transition from no-slip initial state up to the final slip could be done by considering the truncated Yoffe’s function [10]. This process generates the final slip that is presented here and agrees with real final slip and slip rate at each point in the fault. This allows to use the slip rate in order to determine the radiated seismic moment rate [5] and the displacement field [73]. Then, it can be said that the consideration of the fractally distributed energy-based model resembles the kinematic of real earthquakes. Additionally, it can also be claimed that the asperity criterion defined by Somerville et al. [36] is a consequence of fractal surfaces when the HE-B model is applied.

5.3 Statistically realistic ruptures and correlations in the HE-B model

Earthquakes of different magnitude were obtained by changing the coupling functions (Tables 2 and 3). Note that the areas of these C functions followed the scaling relations of Leonard [29]. That is, a prescribed M W ref generates a random distribution of C which allows us to obtain statistically realistic ruptures of similar magnitude. The examples from M W 5 to 7 are shown in Figures 46 and Tables 2 and 3. They reveal that the HE-B model can reproduce the main properties of seismic sources regardless of the considered magnitude, that is, properties such as the asperity criteria, the double corner frequency, the frequency decay, and the power law of their histograms. A special emphasis can be done regarding the physical meaning of the double corner frequency of the spectrum of M ̇ 0 . Figure 4c, and d shows that the feature of double corner frequency is related to the total rupture time and the time spent rupturing a large coupled area. Thus, the double corner frequency is a feature of the rupture entering into a large coupled zone. Here the maximum frequency f max is defined as 1 / G C . From the experiments in rock samples, the maximum frequency of the source is 1 / T C , where T C is the time required to break the breakdown zone which is proportional to the slip weakening distance D C [47]. As D C is proportional to the fracture energy, it implies that the maximum frequency depends on the inverse of G C ( f max 1 / G C ). Similar findings were also noted by Guatteri et al. [8] and Guatteri and Spudich [74]. They observed that the high frequency depends on D C . Specifically, small D C values were related to the source of high frequency, which is the same, small values of G C are related to the source of high frequency content. Thus, the definition of f max used here is physically reliable because it agrees with those found by other authors.

The HE-B model could also be used in order to obtain correlations among different rupture parameters (Figure 7). However, the change in the G C distribution (Figure 8 and Tables 4 and 5) for the same considered coupling function breaks the spatial correlations among rupture parameters (Figure 10) despite the main properties that characterize the real earthquakes hold (Figure 9). In the view of Ohnaka [47], each fault is characterized by a specific fracture energy distribution because it represents the topography or geometrical irregularities of each fault. These rock irregularities can be described as power laws because it corresponds to the geometrical manifestation of the tendency to increases the universe’s entropy [24,7579]. From this fundamental physical view, the distribution of u f inherits these spatial properties that generates the asperity criterion. Furthermore, the fact that the rupture parameters are correlated with each other to different magnitudes for a specific G C distribution (Figure 7) means that these correlations could exist only if the earthquakes are located within the same fault. The change in the distribution of G C implies that we are considering earthquakes of the same size but located in different faults. This change breaks almost all the correlations (Figure 8 and 10). In other words, the correlations among rupture parameters are not equivalent when different faults are being considered.

Finally, it is important to acknowledge that the present model focuses on heterogeneous self-arrested ruptures. However, future research should explore additional realistic approaches. These may include investigating ruptures that extend to fault boundaries [80], considering the dependence of fracture energy on depth (pressure) or temperature [81], accounting for the presence of fluids [82], and examining changes in the fault’s strength over time [83]. Furthermore, it is crucial to explore methods for decomposing ground motion into source and site effects [8487]. These avenues of investigation will enhance our understanding of seismic phenomena and improve our ability to accurately model and predict ground motion during earthquakes.

6 Conclusion

Based on the discussions presented in the previous sections, the utilization of fractal distributions to describe the distribution of final slip and rupture parameters constrained by energy conditions leads to kinematic simulations of earthquakes that can reproduce properties observed in real earthquakes. Additionally, being heterogeneous kinematic simulations, they can contribute to the study of seismic sources and, potentially, improve predictions of strong ground motion with reduced computational resources and increased efficiency. The main conclusions regarding the development of the HE-B model are listed below:

  • The final slip distribution of self-arrested ruptures resembles truncated fractal surfaces.

  • Somerville et al. [36] asperity criterion is a property of fractal surfaces.

  • The energy-based model presented by Noda et al. [41] can be spatially constrained by heterogeneous fractal fracture energy and coupling distributions.

  • The work done by Leonard [29] can constrain the coupling function in terms of a required earthquake magnitude.

  • The heterogeneous coupling function determines the residual energy which defines the total ruptured area.

  • The fractal final slip can be obtained by considering the residual energy, the fracture energy and rupture velocity in order to generate realistic self-arrested ruptures.

  • The HE-B method leads to the same spatial statistics in the final slip as those found in real self-arrested ruptures.

    As the HE-B model enables the generation of realistic simulations, it also facilitates the analysis of the rupture process, such as spatial correlations among rupture parameters and the association of corner frequencies with source properties, such as the duration of asperity rupture. These findings can be summarized as follows:

  • Double corner frequency is naturally obtained by using the HE-B model.

  • The HE-B model also leads to the decay in the frequency content characterized by f 1 and f 2 .

  • The double corner frequency could be a feature of the rupture of a large coupled area.

  • The maximum fracture energy is related to global parameters such as magnitude and peak seismic moment rate for a given fault.

  • The average fracture energy is related to the maximum source frequency for a given fault.

  • No significant correlation is found among maximum and average fracture energy, maximum source frequency, peak seismic moment rate, and magnitude for different faults even though they share the same magnitude and available energy distribution.

Future works should explore other realistic approaches such as fracture energy dependence on depth or temperature, presence of fluids, or changes in the fault’s strength over time.

Acknowledgement

I acknowledge the continuous academic and scientific support of the Pontificia Universidad Católica de Chile during the PhD path.

  1. Funding information: This research received no external funding.

  2. Conflict of interest: Author states no conflict of interest.

  3. Data availability statement: The data were generated following the equations and instructions presented in this work. For any questions or inquiries, please do not hesitate to contact the author.

References

[1] Seidl D, Berckhemer H. Determination of source moment and radiated seismic energy from broadband recordings. Phys Earth Planet Inter. Nov 1982;30(2–3):209–13. 10.1016/0031-9201(82)90108-X.Search in Google Scholar

[2] Ide S. Estimation of radiated energy of finite-source earthquake models. Bull Seismol Soc Am. 2002;92(8):2994–3005. 10.1785/0120020028.Search in Google Scholar

[3] Crempien JGF, Archuleta RJ. Within-event and between-events ground motion variability from earthquake rupture scenarios. Pure Appl Geophy. 2017;174:3451–65. 10.1007/s00024-017-1615-x.Search in Google Scholar

[4] Kurzon I, Lyakhovsky V, Sagy A, Ben-Zion Y. Radiated seismic energy and source damage evolution from analysis of simulated dynamic rupture and far-field seismograms. Geophys J Int. 2022;231(3):1705–26. 10.1093/gji/ggac279.Search in Google Scholar

[5] Renou J, Vallée M, Aochi H. Deciphering the origins of transient seismic moment accelerations by realistic dynamic rupture simulations. Bull Seismol Soc Am. 2022;112:1240–51. 10.1785/0120210221 Search in Google Scholar

[6] Hisada Y. A theoretical omega-square model considering the spatial variation on slip and rupture velocity. Bull Seism Soc Am. 2000;90:387–400. 10.1785/0119990083 Search in Google Scholar

[7] Hisada Y. A theoretical omega-square model considering the spatial variation on slip and rupture velocity. I I. Case for a two-dimensional source model. Bull Seism Soc Am. 2001;91:651–66. 10.1785/0120000097.Search in Google Scholar

[8] Guatteri M, Mai PM, Beroza GC, Boatwright J. Strong ground-motion prediction from stochastic–dynamic source models. Bull Seism Soc Am. 2003;93:301–13. 10.1785/0120020006.Search in Google Scholar

[9] Yoffe E. The moving Griffith crack. London, Edinburgh, Dublin Philos Mag J Sci. 1951;42:739–50. 10.1080/14786445108561302.Search in Google Scholar

[10] Tinti E, Fukuyama E, Piatanesi A, Cocco M. A kinematic source-time function compatible with earthquake dynamics. Bull Seismological Soc Am. 2005;95(4):1211–23. 10.1785/0120040177.Search in Google Scholar

[11] Frankel A, Wirth E, Marafi N, Vidale J, Stephenson W. Broadband synthetic seismograms for magnitude 9 earthquakes on the Cascadia megathrust based on 3D simulations and stochastic synthetics, Part 1: Methodology and overall results. Bull Seismol Soc Am. 2018;108(5A):2347–69. 10.1785/0120180034.Search in Google Scholar

[12] Melgar D, Sahakian VJ, Thomas AM. Deep coseismic slip in the Cascadia megathrust can be consistent with coastal subsidence. Geophys Res Lett. Feb 2022;49(3):e2021GL097404. 10.1029/2021GL097404.Search in Google Scholar

[13] Okubo PG, Aki K. Fractal geometry in the San Andreas Fault system. J Geophys Res. 1987;92(345–355):1987. 10.1029/JB092iB01p00345.Search in Google Scholar

[14] Mai PM, Beroza GC. A spatial random field model to characterize complexity in earthquake slip. J Geophys Res. 2002;107(B11):2308. 10.1029/2001JB000588.Search in Google Scholar

[15] Kaneko Y, Shearer PM. Seismic source spectra and estimated stress drop derived from cohesive-zone models of circular subshear rupture. Geophys J Int. 2014;197(2):1002–15. 10.1093/gji/ggu030.Search in Google Scholar

[16] Jiang Y, Samsonov SV, González PJ. Aseismic fault slip during a shallow normal-faulting seismic swarm constrained using a physically informed geodetic inversion method. J Geophys Res Solid Earth. July 2022;127(7):e2021JB022621. 10.1029/2021JB022621.Search in Google Scholar

[17] Abercrombie R, Rice J. Can observations of earthquake scaling constrain slip weakening? Geophys J Int. 2005;162:406–24. 10.1111/j.1365-246X.2005.02579.x.Search in Google Scholar

[18] Milliner CW, Dolan JF, Hollingsworth J, Leprince S, Ayoub F, Sammis C. Quantifying near-field and off-fault deformation patterns of the 1992 Mw 7.3 Landers earthquake. Geochem Geophy Geosys. 2015;16:1525–2027. 10.1002/2014GC005693.Search in Google Scholar

[19] Kamer Y, Ouillon G, Sornette D. Fault network reconstruction using agglomerative clustering: applications to southern Californian seismicity. Nat Hazards Earth Syst Sci. 2020;20(3611–3625):2020. 10.5194/nhess-20-3611-2020.Search in Google Scholar

[20] Candela T, Renard F, Klinger Y, Mair K, Schmittbuhl J, Brodsky EE. Roughness of fault surfaces over nine decades of length scales. J Geophys Res. 2012;117:B08409. 10.1029/2011JB009041.Search in Google Scholar

[21] Dascher-Cousineau K, Kirkpatrick JD, Cooke ML. Smoothing of fault slip surfaces by scale-invariant wear. J Geophys Res Solid Earth. 2018;123:7913–30. 10.1029/2018JB015638.Search in Google Scholar

[22] Lei Q, Gao K. Correlation between fracture network properties and stress variability in geological media. Geophys Res Lett. 16 May 2018;45(9):3994–4006. 10.1002/2018GL077548 Search in Google Scholar

[23] Konate L, Kondo D, Ponson L. Numerical fracture mechanics based prediction for the roughening of brittle cracks in 2D disordered solids. Int J Fract. 2021;230:225–40. 10.1007/s10704-021-00576-1.Search in Google Scholar

[24] Xie H. Fractals in rock mechanics. vol. 1, Rotterdam, Netherlands: Crc Press; 1993.Search in Google Scholar

[25] Tumarkin AG, Archuleta RJ, Madariaga R. Scaling relations for composite earthquake models. Bull Seismol Soc Am. 1994;84(4):1279–83. 10.1785/BSSA0840041279.Search in Google Scholar

[26] Schmittbuhl J, Schmitt F, Scholz C. Scaling invariance of crack surfaces. J Geophys Res. 10 April 1995;100(B4):5953–73. 10.1029/94JB02885.Search in Google Scholar

[27] Power WL, Durham WB. Topography of natural and artificial fractures in granitic rocks: Implications for studies of rock friction and fluid migration. Int J Rock Mech Min Sci. 1997;34:979–89. 10.1016/S1365-1609(97)80007-X.Search in Google Scholar

[28] Voisin C, Renard F, Grasso JR. Long term friction: From stick-slip to stable sliding. Geophys Res Lett. 2007;34:L13301. 10.1029/2007GL029715.Search in Google Scholar

[29] Leonard M. Earthquake fault scaling: Self-consistent relating of rupture length, width, average displacement, and moment release. Bull Seismol Soc Am. October 2010;100(5A):1971–88. 10.1785/0120090189.Search in Google Scholar

[30] Walter JI, Svetlizky I, Fineberg J, Brodsky EE, Tulaczyk S, Barcheck CG, et al. Rupture speed dependence on initial stress profiles: Insights from glacier and laboratory stick-slip. Earth Planet Sci Lett. 1 February 2015;411:112–20. 10.1016/j.epsl.2014.11.025.Search in Google Scholar

[31] Liu D, Duan B. Scenario earthquake and ground‐motion simulations in North China: Effects of heterogeneous fault stress and 3D basin structure. Bull Seismol Soc Am. 2018;108(4):2148–69. 10.1785/0120170374.Search in Google Scholar

[32] Le Priol C, Chopin J, Le Doussal P, Ponson L, Rosso A. Universal scaling of the velocity field in crack front propagation. Phys Rev Lett. 11 February 2020;124:1–5. 10.1103/PhysRevLett.124.065501.Search in Google Scholar PubMed

[33] Heimisson ER. Crack to pulse transition and magnitude statistics during earthquake cycles on a self-similar rough fault. Earth Planet Sci Lett. 1 May 2020;537:116202. 10.1016/j.epsl.2020.116202.Search in Google Scholar

[34] Dublanchet P. Shear stress and b-value fluctuations in a hierarchical rate-and-state asperity model. Pure Appl Geophys. 2022;179:2423–35. 10.1007/s00024-022-03039-3.Search in Google Scholar

[35] Dublanchet P. Inferring fault slip rates from cumulative seismic moment in a multiple asperity context. Geophys J Int. January 2019;216(1):395–413. 10.1093/gji/ggy431.Search in Google Scholar

[36] Somerville P, Irikura K, Graves R, Sawada S, Wald D, Abrahamson N, et al. Characterizing crustal earthquake slip models for the prediction of strong ground motion. Seismol Res Lett. 1999;70(1):59–80. 10.1785/gssrl.70.1.59.Search in Google Scholar

[37] Skarlatoudis AA, Somerville PG, Thio HK. Source‐scaling relations of interface subduction earthquakes for strong ground motion and tsunami simulation. Bull Seismol Soc Am. MAY 31, 2016;106(4):1652–62. 10.1785/0120150320.Search in Google Scholar

[38] Erickson BA, Birnir B, Lavallée D. Periodicity, chaos and localization in a Burridge–Knopoff model of an earthquake with rate-and-state friction. Geophys J Int. October 2011;187(1):178–98. 10.1111/j.1365-246X.2011.05123.x.Search in Google Scholar

[39] Ide S, Aochi H. Earthquakes as multiscale dynamic ruptures with heterogeneous fracture surface energy. J Geophys Res. November 2005;110:B11303. 10.1029/2004JB003591.Search in Google Scholar

[40] Chen Y, Yang H. Numerical simulation and pattern characterization of nonlinear spatiotemporal dynamics on fractal surfaces for the whole-heart modeling applications. Eur Phys J. 2016;89:1–6. 10.1140/epjb/e2016-60960-6.Search in Google Scholar

[41] Noda A, Saito T, Fukuyama E, Urata Y. Energy-based scenarios for great thrust-type earthquakes in the Nankai trough subduction zone, southwest Japan, using an interseismic slip-deficit model. J Geophys Res Solid Earth. 2021;126:e2020JB020417. 10.1029/2020JB020417.Search in Google Scholar

[42] Venkataraman A, Kanamori H. Observational constraints on the fracture energy of subduction zone earthquakes. J Geophys Res. 2004;109:B05302. 10.1029/2003jb002549.Search in Google Scholar

[43] Rosenau M, Horenko I, Corbi F, Rudolf M, Kornhuber R, Oncken O. Synchronization of Great Subduction Megathrust Earthquakes: Insights From Scale Model Analysis. J Geophys Res Solid Earth. 2019;124:3646–61. 10.1029/2018JB016597.Search in Google Scholar

[44] Wang K, Dixon T. “Coupling” Semantics and science in earthquake research. Eos, Trans Am Geophys Union. 2004;85(18):180. 10.1029/2004EO180005.Search in Google Scholar

[45] Ida Y. Cohesive force across the tip of a longitudinal-shear crack and Griffith’s specific surface energy. J Geophys Res. 1972;77(20):3796–805. 10.1029/jb077i020p03796.Search in Google Scholar

[46] Ohnaka M, Yamashita T. A cohesive zone model for dynamic shear faulting based on experimentally inferred constitutive relation and strong motion source parameters. J Geophys Res. 1989;94:4089–4104. 10.1029/JB094iB04p04089.Search in Google Scholar

[47] Ohnaka M. The physics of rock failure and earthquakes. Cambridge, UK: Cambridge University Press; April 2013. 10.1017/CBO9781139342865.Search in Google Scholar

[48] Williford RE. Multifractal fracture. Scr Metall (U S). 1988–11-01;22(11):1749–54. 10.1016/S0036-9748(88)80277-1.Search in Google Scholar

[49] Pramanik B, Tadepalli T, Mantena PR. Surface fractal analysis for estimating the fracture energy absorption of nanoparticle reinforced composites. Materials. 2012;5(5):922–36. 10.3390/ma5050922.Search in Google Scholar PubMed PubMed Central

[50] Li H, Zhang J-C, Branicio PS, Sha Z-D. Composition-dependent fracture energy in metallic glasses. Phys Rev Mater. March 2023;7(3):1–9. 10.1103/PhysRevMaterials.7.035602.Search in Google Scholar

[51] Ke C-Y, McLaskey GC, Kammer DS. Earthquake breakdown energy scaling despite constant fracture energy. Nat Commun. 2022;13:1005. 10.1038/s41467-022-28647-4.Search in Google Scholar PubMed PubMed Central

[52] Nielsen S, Spagnuolo E, Smith SAF, Violay M, Di Toro G, Bistacchi A. Scaling in natural and laboratory earthquakes. Geophys Res Lett. 2016;43:1504–10. 10.1002/2015GL067490.Search in Google Scholar

[53] Weng J, Chen X, Xu J. A Dynamic Explanation for the Ruptures of Repeating Earthquakes on the San Andreas Fault at Parkfield. Geophys Res Lett. 28 October 2018;45(20):11116–22. 10.1029/2018GL079140.Search in Google Scholar

[54] Ohnaka M, Kuwahara Y, Yamamoto K. Constitutive relations between dynamic physical parameters near a tip of the propagating slip zone during stick–slip shear failure. Tectonophysics. 15 December 1987;144(1–3):109–25. 10.1016/0040-1951(87)90011-4.Search in Google Scholar

[55] Schmedes J, Archuleta RJ, Lavallée D. Correlation of earthquake source parameters inferred from dynamic rupture simulations. J Geophys Res. 2010;115:B03304. 10.1029/2009JB006689.Search in Google Scholar

[56] Bizzarri A. Rupture speed and slip velocity: What can we learn from simulated earthquakes. Earth Planet Sci Lett. 1 February 2012;317–318:196–203. 10.1016/j.epsl.2011.11.023.Search in Google Scholar

[57] Song SG, Dalguer LA, Mai PM. Pseudo-dynamic source modelling with 1-point and 2-point statistics of earthquake source parameters. Geophys J Int. 2014;196:1770–86. 10.1093/gji/ggt479.Search in Google Scholar

[58] Renou J, Vallée M, Dublanchet P. How does seismic rupture accelerate? observational insights from earthquake source time month. J Geophys Res Solid Earth. 2019;124:8942–52. 10.1029/2019JB018045.Search in Google Scholar

[59] Oglesby DD, Day SM. Stochastic fault stress: Implications for fault dynamics and ground motion. Bull Seismol Soc Am. 2002;92(8):3006–21. 10.1785/0120010249.Search in Google Scholar

[60] Dalguer LA, Miyake H, Day SM, Irikura K. Surface rupturing and buried dynamic-rupture models calibrated with statistical observations of past earthquakes. Bull Seismological Soc Am. 2008;98(3):1147–61. 10.1785/0120070134.Search in Google Scholar

[61] Pavlenko OV. Simulation of ground motion from strong earthquakes of Kamchatka Region (1992–1993) at rock and soil sites. Pure Appl Geophys. 2013;170:571–95. 10.1007/s00024-012-0529-x.Search in Google Scholar

[62] Mavroeidis GP, Papageorgiou AS. Effect of fault rupture characteristics on near-fault strong ground motions. Bull Seismol Soc Am. 2010;100(1):37–58. 10.1785/0120090018.Search in Google Scholar

[63] Somala SN, Ampuero JP, Lapusta N. Resolution of rise time in earthquake slip inversions: effect of station spacing and rupture velocity. Bull Seismol Soc Am. December 2014;104(6):2717–34. 10.1785/0120130185.Search in Google Scholar

[64] Atkinson GM, Silva W. Stochastic modeling of California ground motions. Bull Seismol Soc Am. 2000;90(2):255–74. 10.1785/0119990064.Search in Google Scholar

[65] Ji C, Archuleta R. Two empirical double-cornerfrequency source spectra and their physical implications. Bull Seismol Soc Am. 2020;111:737–61. 10.1785/0120200238.Search in Google Scholar

[66] Delouis B, Nocquet JM, Vallée M. Slip distribution of the February 27, 2010 Mw = 8.8 Maule Earthquake, central Chile, from static and high-rate GPS, InSAR, and broadband teleseismic data. Geophys Res Lett. 2010;2010(37):L17305. 10.1029/2010GL043899.Search in Google Scholar

[67] Moreno M, Melnick D, Rosenau M, Baez J, Klotz J, Oncken O, et al. Toward understanding tectonic control on the Mw 8.8 2010 Maule Chile earthquak. Earth Planet Sci Lett. 2012;321-322:152–65. 10.1016/j.epsl.2012.01.006.Search in Google Scholar

[68] Heuret A, Lallemand S, Funiciello F, Piromallo C, Faccenna C. Physical characteristics of subduction interface type seismogenic zones revisited. Geochem Geophy Geosys. January 2011;12(1):1–26. 10.1029/2010GC003230.Search in Google Scholar

[69] Saito T, Noda A. Mechanically coupled areas on the plate interface in the Nankai Trough, Japan and a possible seismic and aseismic rupture scenario for Megathrust earthquakes. J Geophys Res Solid Earth. August 2022;127(8):e2022JB023992. 10.1029/2022JB023992.Search in Google Scholar

[70] Guatteri M, Mai PM, Beroza GC. A Pseudo-dynamic approximation to dynamic rupture models for strong ground motion prediction. Bull Seism Soc Am. December 2004;94(6):2051–63. 10.1785/0120040037.Search in Google Scholar

[71] Kieling K, Wang R, Hainzl S. Broadband ground-motion simulation using energy-constrained rise-time scaling. Bull Seismol Soc Am. December 2014;104(6):2683–97. 10.1785/0120140063.Search in Google Scholar

[72] Kostrov VV. Seismic moment and energy of earthquakes, and seismic flow of rock. Izvestiya, Earth. Physics. 1974;1:23–40.Search in Google Scholar

[73] Aki K, Richards PG. Quantitative seismology. 2nd edn. Sausalito, CA: University Science Books; 2002.Search in Google Scholar

[74] Guatteri M, Spudich P. What can strong motion data tell us about slip-weakening fault-friction laws? Bull Seism Soc Am. 2000;90:98–116. 10.1785/0119990053.Search in Google Scholar

[75] Mandelbrot BB. The fractal geometry of nature. New York, NY, USA: W. H. Freeman and Company; 1982.Search in Google Scholar

[76] Brown SR, Scholz CH. Broad bandwidth study of the topography of natural rock surfaces. J Geophys Res Solid Earth. 1985;90(B14):12575–82. 10.1029/JB090iB14p12575.Search in Google Scholar

[77] Turcotte DL. Fractal and chaos in geology and geophysics. Cambridge: Cambridge University Press; 1997. 10.1017/CBO9781139174695.Search in Google Scholar

[78] Bejan A, Lorente S. Constructal law of design and evolution: Physics, biology, technology, and society. J Appl Phys. 2013;113(151301):1–20. 10.1063/1.4798429.Search in Google Scholar

[79] Venegas-Aravena P, Cordaro E, Laroze D. Natural fractals as irreversible disorder: entropy approach from cracks in the semi brittle-ductile lithosphere and generalization. Entropy 2022. 2022;24(1337):1–18. 10.3390/e24101337.Search in Google Scholar PubMed PubMed Central

[80] Yao S, Yang H. Hypocentral dependent shallow slip distribution and rupture extents along a strike-slip fault. Earth Planet Sci Lett. January 2022;578(15):117296. 10.1016/j.epsl.2021.117296.Search in Google Scholar

[81] Hayashi A, Sekiguchi Y, Sato C. Effect of temperature and loading rate on the mode I fracture energy of structural acrylic adhesives. J Adv Join Process. June 2022;5:100079. 10.1016/j.jajp.2021.100079.Search in Google Scholar

[82] Cebry SBL, McLaskey GC. Seismic swarms produced by rapid fluid injection into a low permeability laboratory fault. Earth Planet Sci Lett. 1 March 2021;557:116726. 10.1016/j.epsl.2020.116726.Search in Google Scholar

[83] Bedford JD, Faulkner DR, Lapusta N. Fault rock heterogeneity can produce fault weakness and reduce fault stability. Nat Commun. 2022;13(326). 10.1038/s41467-022-27998-2.Search in Google Scholar PubMed PubMed Central

[84] Mendecki MJ, Duda A, Idziak A. Ground-motion prediction equation and site effect characterization for the central area of the Main Syncline, Upper Silesia Coal Basin, Poland. Open Geosci. 2018;10(1):474–83. 10.1515/geo-2018-0037.Search in Google Scholar

[85] Jiang Q, Rong M, Wei W, Zhang B, Wang J. Influence of thick soft superficial layers of seabed on ground motion and its treatment suggestions for site response analysis. Open Geosci. 2021;13(1):1273–89. 10.1515/geo-2020-0302.Search in Google Scholar

[86] Dujardin A, Courboulex F, Causse M, Traversa P. Influence of source, path, and site effects on the magnitude dependence of ground‐motion decay with distance. Seismol Res Lett. 2016;87(1):138–48. 10.1785/0220150185.Search in Google Scholar

[87] Nagao T, Fukushima Y. Source- and site-specific earthquake ground motions, application of a state-of-the-art evaluation method. Eng Technol Appl Sci Res. August 2020;10(4):5882–8. 10.48084/etasr.3612.Search in Google Scholar

Received: 2023-04-26
Revised: 2023-07-13
Accepted: 2023-07-16
Published Online: 2023-08-21

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Diagenesis and evolution of deep tight reservoirs: A case study of the fourth member of Shahejie Formation (cg: 50.4-42 Ma) in Bozhong Sag
  3. Petrography and mineralogy of the Oligocene flysch in Ionian Zone, Albania: Implications for the evolution of sediment provenance and paleoenvironment
  4. Biostratigraphy of the Late Campanian–Maastrichtian of the Duwi Basin, Red Sea, Egypt
  5. Structural deformation and its implication for hydrocarbon accumulation in the Wuxia fault belt, northwestern Junggar basin, China
  6. Carbonate texture identification using multi-layer perceptron neural network
  7. Metallogenic model of the Hongqiling Cu–Ni sulfide intrusions, Central Asian Orogenic Belt: Insight from long-period magnetotellurics
  8. Assessments of recent Global Geopotential Models based on GPS/levelling and gravity data along coastal zones of Egypt
  9. Accuracy assessment and improvement of SRTM, ASTER, FABDEM, and MERIT DEMs by polynomial and optimization algorithm: A case study (Khuzestan Province, Iran)
  10. Uncertainty assessment of 3D geological models based on spatial diffusion and merging model
  11. Evaluation of dynamic behavior of varved clays from the Warsaw ice-dammed lake, Poland
  12. Impact of AMSU-A and MHS radiances assimilation on Typhoon Megi (2016) forecasting
  13. Contribution to the building of a weather information service for solar panel cleaning operations at Diass plant (Senegal, Western Sahel)
  14. Measuring spatiotemporal accessibility to healthcare with multimodal transport modes in the dynamic traffic environment
  15. Mathematical model for conversion of groundwater flow from confined to unconfined aquifers with power law processes
  16. NSP variation on SWAT with high-resolution data: A case study
  17. Reconstruction of paleoglacial equilibrium-line altitudes during the Last Glacial Maximum in the Diancang Massif, Northwest Yunnan Province, China
  18. A prediction model for Xiangyang Neolithic sites based on a random forest algorithm
  19. Determining the long-term impact area of coastal thermal discharge based on a harmonic model of sea surface temperature
  20. Origin of block accumulations based on the near-surface geophysics
  21. Investigating the limestone quarries as geoheritage sites: Case of Mardin ancient quarry
  22. Population genetics and pedigree geography of Trionychia japonica in the four mountains of Henan Province and the Taihang Mountains
  23. Performance audit evaluation of marine development projects based on SPA and BP neural network model
  24. Study on the Early Cretaceous fluvial-desert sedimentary paleogeography in the Northwest of Ordos Basin
  25. Detecting window line using an improved stacked hourglass network based on new real-world building façade dataset
  26. Automated identification and mapping of geological folds in cross sections
  27. Silicate and carbonate mixed shelf formation and its controlling factors, a case study from the Cambrian Canglangpu formation in Sichuan basin, China
  28. Ground penetrating radar and magnetic gradient distribution approach for subsurface investigation of solution pipes in post-glacial settings
  29. Research on pore structures of fine-grained carbonate reservoirs and their influence on waterflood development
  30. Risk assessment of rain-induced debris flow in the lower reaches of Yajiang River based on GIS and CF coupling models
  31. Multifractal analysis of temporal and spatial characteristics of earthquakes in Eurasian seismic belt
  32. Surface deformation and damage of 2022 (M 6.8) Luding earthquake in China and its tectonic implications
  33. Differential analysis of landscape patterns of land cover products in tropical marine climate zones – A case study in Malaysia
  34. DEM-based analysis of tectonic geomorphologic characteristics and tectonic activity intensity of the Dabanghe River Basin in South China Karst
  35. Distribution, pollution levels, and health risk assessment of heavy metals in groundwater in the main pepper production area of China
  36. Study on soil quality effect of reconstructing by Pisha sandstone and sand soil
  37. Understanding the characteristics of loess strata and quaternary climate changes in Luochuan, Shaanxi Province, China, through core analysis
  38. Dynamic variation of groundwater level and its influencing factors in typical oasis irrigated areas in Northwest China
  39. Creating digital maps for geotechnical characteristics of soil based on GIS technology and remote sensing
  40. Changes in the course of constant loading consolidation in soil with modeled granulometric composition contaminated with petroleum substances
  41. Correlation between the deformation of mineral crystal structures and fault activity: A case study of the Yingxiu-Beichuan fault and the Milin fault
  42. Cognitive characteristics of the Qiang religious culture and its influencing factors in Southwest China
  43. Spatiotemporal variation characteristics analysis of infrastructure iron stock in China based on nighttime light data
  44. Interpretation of aeromagnetic and remote sensing data of Auchi and Idah sheets of the Benin-arm Anambra basin: Implication of mineral resources
  45. Building element recognition with MTL-AINet considering view perspectives
  46. Characteristics of the present crustal deformation in the Tibetan Plateau and its relationship with strong earthquakes
  47. Influence of fractures in tight sandstone oil reservoir on hydrocarbon accumulation: A case study of Yanchang Formation in southeastern Ordos Basin
  48. Nutrient assessment and land reclamation in the Loess hills and Gulch region in the context of gully control
  49. Handling imbalanced data in supervised machine learning for lithological mapping using remote sensing and airborne geophysical data
  50. Spatial variation of soil nutrients and evaluation of cultivated land quality based on field scale
  51. Lignin analysis of sediments from around 2,000 to 1,000 years ago (Jiulong River estuary, southeast China)
  52. Assessing OpenStreetMap roads fitness-for-use for disaster risk assessment in developing countries: The case of Burundi
  53. Transforming text into knowledge graph: Extracting and structuring information from spatial development plans
  54. A symmetrical exponential model of soil temperature in temperate steppe regions of China
  55. A landslide susceptibility assessment method based on auto-encoder improved deep belief network
  56. Numerical simulation analysis of ecological monitoring of small reservoir dam based on maximum entropy algorithm
  57. Morphometry of the cold-climate Bory Stobrawskie Dune Field (SW Poland): Evidence for multi-phase Lateglacial aeolian activity within the European Sand Belt
  58. Adopting a new approach for finding missing people using GIS techniques: A case study in Saudi Arabia’s desert area
  59. Geological earthquake simulations generated by kinematic heterogeneous energy-based method: Self-arrested ruptures and asperity criterion
  60. Semi-automated classification of layered rock slopes using digital elevation model and geological map
  61. Geochemical characteristics of arc fractionated I-type granitoids of eastern Tak Batholith, Thailand
  62. Lithology classification of igneous rocks using C-band and L-band dual-polarization SAR data
  63. Analysis of artificial intelligence approaches to predict the wall deflection induced by deep excavation
  64. Evaluation of the current in situ stress in the middle Permian Maokou Formation in the Longnüsi area of the central Sichuan Basin, China
  65. Utilizing microresistivity image logs to recognize conglomeratic channel architectural elements of Baikouquan Formation in slope of Mahu Sag
  66. Resistivity cutoff of low-resistivity and low-contrast pays in sandstone reservoirs from conventional well logs: A case of Paleogene Enping Formation in A-Oilfield, Pearl River Mouth Basin, South China Sea
  67. Examining the evacuation routes of the sister village program by using the ant colony optimization algorithm
  68. Spatial objects classification using machine learning and spatial walk algorithm
  69. Study on the stabilization mechanism of aeolian sandy soil formation by adding a natural soft rock
  70. Bump feature detection of the road surface based on the Bi-LSTM
  71. The origin and evolution of the ore-forming fluids at the Manondo-Choma gold prospect, Kirk range, southern Malawi
  72. A retrieval model of surface geochemistry composition based on remotely sensed data
  73. Exploring the spatial dynamics of cultural facilities based on multi-source data: A case study of Nanjing’s art institutions
  74. Study of pore-throat structure characteristics and fluid mobility of Chang 7 tight sandstone reservoir in Jiyuan area, Ordos Basin
  75. Study of fracturing fluid re-discharge based on percolation experiments and sampling tests – An example of Fuling shale gas Jiangdong block, China
  76. Impacts of marine cloud brightening scheme on climatic extremes in the Tibetan Plateau
  77. Ecological protection on the West Coast of Taiwan Strait under economic zone construction: A case study of land use in Yueqing
  78. The time-dependent deformation and damage constitutive model of rock based on dynamic disturbance tests
  79. Evaluation of spatial form of rural ecological landscape and vulnerability of water ecological environment based on analytic hierarchy process
  80. Fingerprint of magma mixture in the leucogranites: Spectroscopic and petrochemical approach, Kalebalta-Central Anatolia, Türkiye
  81. Principles of self-calibration and visual effects for digital camera distortion
  82. UAV-based doline mapping in Brazilian karst: A cave heritage protection reconnaissance
  83. Evaluation and low carbon ecological urban–rural planning and construction based on energy planning mechanism
  84. Modified non-local means: A novel denoising approach to process gravity field data
  85. A novel travel route planning method based on an ant colony optimization algorithm
  86. Effect of time-variant NDVI on landside susceptibility: A case study in Quang Ngai province, Vietnam
  87. Regional tectonic uplift indicated by geomorphological parameters in the Bahe River Basin, central China
  88. Computer information technology-based green excavation of tunnels in complex strata and technical decision of deformation control
  89. Spatial evolution of coastal environmental enterprises: An exploration of driving factors in Jiangsu Province
  90. A comparative assessment and geospatial simulation of three hydrological models in urban basins
  91. Aquaculture industry under the blue transformation in Jiangsu, China: Structure evolution and spatial agglomeration
  92. Quantitative and qualitative interpretation of community partitions by map overlaying and calculating the distribution of related geographical features
  93. Numerical investigation of gravity-grouted soil-nail pullout capacity in sand
  94. Analysis of heavy pollution weather in Shenyang City and numerical simulation of main pollutants
  95. Road cut slope stability analysis for static and dynamic (pseudo-static analysis) loading conditions
  96. Forest biomass assessment combining field inventorying and remote sensing data
  97. Late Jurassic Haobugao granites from the southern Great Xing’an Range, NE China: Implications for postcollision extension of the Mongol–Okhotsk Ocean
  98. Petrogenesis of the Sukadana Basalt based on petrology and whole rock geochemistry, Lampung, Indonesia: Geodynamic significances
  99. Numerical study on the group wall effect of nodular diaphragm wall foundation in high-rise buildings
  100. Water resources utilization and tourism environment assessment based on water footprint
  101. Geochemical evaluation of the carbonaceous shale associated with the Permian Mikambeni Formation of the Tuli Basin for potential gas generation, South Africa
  102. Detection and characterization of lineaments using gravity data in the south-west Cameroon zone: Hydrogeological implications
  103. Study on spatial pattern of tourism landscape resources in county cities of Yangtze River Economic Belt
  104. The effect of weathering on drillability of dolomites
  105. Noise masking of near-surface scattering (heterogeneities) on subsurface seismic reflectivity
  106. Query optimization-oriented lateral expansion method of distributed geological borehole database
  107. Petrogenesis of the Morobe Granodiorite and their shoshonitic mafic microgranular enclaves in Maramuni arc, Papua New Guinea
  108. Environmental health risk assessment of urban water sources based on fuzzy set theory
  109. Spatial distribution of urban basic education resources in Shanghai: Accessibility and supply-demand matching evaluation
  110. Spatiotemporal changes in land use and residential satisfaction in the Huai River-Gaoyou Lake Rim area
  111. Walkaway vertical seismic profiling first-arrival traveltime tomography with velocity structure constraints
  112. Study on the evaluation system and risk factor traceability of receiving water body
  113. Predicting copper-polymetallic deposits in Kalatag using the weight of evidence model and novel data sources
  114. Temporal dynamics of green urban areas in Romania. A comparison between spatial and statistical data
  115. Passenger flow forecast of tourist attraction based on MACBL in LBS big data environment
  116. Varying particle size selectivity of soil erosion along a cultivated catena
  117. Relationship between annual soil erosion and surface runoff in Wadi Hanifa sub-basins
  118. Influence of nappe structure on the Carboniferous volcanic reservoir in the middle of the Hongche Fault Zone, Junggar Basin, China
  119. Dynamic analysis of MSE wall subjected to surface vibration loading
  120. Pre-collisional architecture of the European distal margin: Inferences from the high-pressure continental units of central Corsica (France)
  121. The interrelation of natural diversity with tourism in Kosovo
  122. Assessment of geosites as a basis for geotourism development: A case study of the Toplica District, Serbia
  123. IG-YOLOv5-based underwater biological recognition and detection for marine protection
  124. Monitoring drought dynamics using remote sensing-based combined drought index in Ergene Basin, Türkiye
  125. Review Articles
  126. The actual state of the geodetic and cartographic resources and legislation in Poland
  127. Evaluation studies of the new mining projects
  128. Comparison and significance of grain size parameters of the Menyuan loess calculated using different methods
  129. Scientometric analysis of flood forecasting for Asia region and discussion on machine learning methods
  130. Rainfall-induced transportation embankment failure: A review
  131. Rapid Communication
  132. Branch fault discovered in Tangshan fault zone on the Kaiping-Guye boundary, North China
  133. Technical Note
  134. Introducing an intelligent multi-level retrieval method for mineral resource potential evaluation result data
  135. Erratum
  136. Erratum to “Forest cover assessment using remote-sensing techniques in Crete Island, Greece”
  137. Addendum
  138. The relationship between heat flow and seismicity in global tectonically active zones
  139. Commentary
  140. Improved entropy weight methods and their comparisons in evaluating the high-quality development of Qinghai, China
  141. Special Issue: Geoethics 2022 - Part II
  142. Loess and geotourism potential of the Braničevo District (NE Serbia): From overexploitation to paleoclimate interpretation
Downloaded on 11.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/geo-2022-0522/html
Scroll to top button