Next Article in Journal
Application Cluster Analysis as a Support form Modelling and Digitalizing the Logistics Processes in Warehousing
Previous Article in Journal
Cost-Effective Temperature Sensor for Monitoring the Setting Time of Concrete
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Stress State before the MS 6.8 Luding Earthquake on 5 September 2022 in Sichuan, China: A Retrospective View Based on the b-Value

1
Sichuan Earthquake Agency, Chengdu 610041, China
2
College of Earth Science, Chengdu University of Technology, Chengdu 610059, China
3
Crustal Deformation Monitoring Center of Sichuan Earthquake Agency, Ya’an 625000, China
4
Department of Earth and Space Sciences, Southern University of Science and Technology, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(11), 4345; https://doi.org/10.3390/app14114345
Submission received: 8 April 2024 / Revised: 17 May 2024 / Accepted: 17 May 2024 / Published: 21 May 2024

Abstract

:
On 5 September 2022 (BJT), Luding, located in southwestern Sichuan Province, China, experienced an MS 6.8 earthquake. This earthquake occurred within the historical rupture zone of the 1786 MS 7.75 event, part of the southern section of the Xianshui He Fault belt. Given the average 155-year recurrence interval for strong earthquakes in this area, the 236 years since the last event made this earthquake somewhat expected. However, prior to this event, we did not detect any anomalies indicating low surface b-values, which are often indicative of a high-stress state in the source area before strong earthquakes, as highlighted by numerous studies. Our research focused on the northern section of the eastern boundary of the Sichuan–Yunnan sub-block, encompassing the Xianshui He, Anning He, Zemu He, and Daliang Shan fault belts. We meticulously located earthquakes of ML ≥ 1.5 from 2009 to May 2022. The catalog was divided into two periods: 2009–2014 and 2015–May 2022. Using an AIC-constraint method, we analyzed the changes in b-values (Δb) in the latter period compared to the former. Our findings revealed a significant abnormal Δb zone (Δb < −0.3), with a radius of approximately 50 km, when ΔAIC ≥ 2 was selected. Intriguingly, the epicenter of the recent Luding MS 6.8 earthquake fell within this abnormal zone. Furthermore, we calculated the b-value cross-section for the southern section of the Xianshui He fault belt using a directory of precisely located small earthquakes. This revealed that the location, scale, and shape of the abnormally low-b-value area corresponded with the large displacement co-seismic area of the main earthquake, affirming the b-value’s effectiveness in identifying asperities. The b-value’s temporal evolution prior to the mainshock exhibited a nearly decade-long continuous decrease, signifying a long-term stress-loading process akin to that observed before many strong earthquakes. The b-value anomalies observed from different profiles before the Luding earthquake underline the necessity of a comprehensive, multi-dimensional analysis of such anomalies. Finally, our analysis indicates that nine earthquakes with MS ≥ 6.5, including the Luding MS 6.8 event, have contributed to increased Coulomb Failure Stress change (ΔCFS) in the Daofu (DF)–Kangding (KD) section of the Xianshui He fault belt and the northern section of the Anning He fault belt south of Shimian (SM), with amplitudes surpassing the 0.01 MPa threshold. This suggests the potential for strong earthquakes in these zones.

1. Introduction

The b-value is a pivotal parameter in seismology, quantifying the relationship between earthquake magnitudes and their occurrence frequency in a specific area. This metric plays a crucial role in decoding earthquake patterns and assessing future seismic risks. It transcends a mere statistical measure, serving as a vital gauge of the Earth’s crustal stress state. Generally, lower b-values signal higher stress levels in a region, while higher b-values often denote lower stress conditions. This inverse relationship has been substantiated through extensive studies on fault systems’ physical properties, creep–slip dynamics along fault lines, and the movement of pore pressures within the crust, as evidenced by seminal works [1,2,3,4,5].
The study of the b-value’s spatial and temporal evolution in areas prone to frequent, strong earthquakes is particularly fascinating. There is a rich history of successfully predicting the location and magnitude of potential strong earthquakes by identifying regions with low b-values, known as asperities. Notable examples include the San Jacinto–Elsinore fault system in Southern California [6], the Parkfield segment of the San Andreas fault [7], and the 2014 Parkfield M 6.0 earthquake analysis [5]. A consistent pattern observed before some major earthquakes is a prolonged, steady decrease in b-values, culminating in a significant seismic event, such as the 2008 Wenchuan Mw 7.9 earthquake in China [8] and the 2011 East Japan Mw 9.0 earthquake [9]. Regular monitoring of b-values’ spatial and temporal heterogeneity serves as a standard method for identifying precursors to potential large earthquakes.
In our quest to apply b-values more effectively in assessing seismic hazard in zones with a high probability of strong earthquakes, we conducted a retrospective analysis of the b-value’s temporal and spatial evolution before the Ms 6.8 Luding earthquake on 5 September 2022 in Sichuan, China. This study also explores the potential risks of strong earthquakes in the northern section of the eastern boundary of the Sichuan–Yunnan block.

2. b-Value Calculation Methods

In our estimation process of the b-value, we employed the widely recognized Gutenberg–Richter relationship [10] in statistical seismology, which is expressed as log N = a b M , where N denotes the cumulative number of earthquakes with magnitude equal to or greater than M, and a and b represent constants. Here, the a-value indicates the level of seismic activity within the region, while the b-value, representing the slope of the magnitude–frequency distribution, is typically associated with the region’s geological structure and prevailing stress state.
Given the linear nature of this relationship, the Least Squares Method (LSM) has been traditionally utilized to estimate a- and b-values. However, LSM tends to assign equal weight across all magnitude bins, which can lead to disproportionate influences from smaller sample sizes in higher-magnitude bins on the estimated b-value. Such disproportionate effects can result in unstable b-value estimates, which might lead to unreliable predictions in seismic risk assessments.
To address these limitations, we recommend and have utilized the maximum likelihood estimation (MLE) method in this study. The MLE method treats the magnitude M as a continuous random variable and assumes that it follows a power-law distribution, as outlined by Aki [11] and Utsu [12]. They provided foundational formulas for the maximum likelihood estimation of the b-value, enhancing the precision of b-value calculations by effectively accounting for different sample sizes across magnitude bins. This approach significantly reduces the biases associated with LSM, thereby providing more stable and reliable estimates of b-values. Based on this, Aki and Utsu separately provided expressions for the maximum likelihood estimation of the b-value:
b ^ = log ( e ) M ¯ M C
M ¯ represents the average magnitude, and M C is the magnitude of completeness. Utsu [13] further refined this method, improving the accuracy of the b-value estimations under varied seismic conditions:
b ^ = log ( e ) M ¯ ( M C Δ M / 2 )
The MLE method calculates the b-value based on the probability density values across different magnitude segments, avoiding the disproportionate weight issues characteristic of the LSM. This enhancement in methodological approach leads to more accurate and stable b-value estimates, which are crucial for understanding the underlying stress states in seismic regions.
Recent research emphasizes the critical role of accurate b-value estimation in understanding regional stress levels and seismic risk. Studies have demonstrated that lower b-values are often associated with high seismic coupling and geological asperities, particularly in regions like the Chilean megathrust, highlighting areas of significant stress accumulation that are prone to large earthquakes [14]. For instance, Tormann et al. (2015) observed rapid changes in b-values before and after the Tohoku-oki earthquake, providing insights into the randomness of megathrust earthquakes and the dynamics of stress recovery [15]. This study further illustrates how sudden shifts in b-values can signify underlying stress adjustments that might not be apparent from surface observations alone.
Other investigations have linked rapid changes in reservoir water levels to lower b-values, suggesting that such environmental modifications can induce seismic activity through increased pore pressure and resultant stress [16]. Moreover, a trend of decreasing b-values preceding major earthquakes has been observed, indicating rising shear stress within fault zones. These findings collectively validate the use of b-values as reliable indicators of stress variations, serving as essential tools for seismic risk assessment and the development of mitigation strategies [17].
Thus, by employing Utsu’s refined MLE formula for b-value calculation, we aim to enhance the reliability of seismic activity forecasts. The uncertainty of the b-value can becan be estimated by
σ ^ = b ^ N

3. Tectonic Setting

The clockwise rotation of the Tibetan Plateau, as delineated by Niu et al. [18], has fragmented the tectonic landscape into distinct sub-blocks [19] (illustrated in Figure 1). These sub-blocks are traversed by deep, major faults, each with unique kinematic attributes. The northward Bayan Har sub-block’s eastward trajectory is halted by the stable South China sub-block, culminating in the formation of the thrust-type Longmen Shan fault zone at their interface [20]. This zone has witnessed significant seismic events—notably, the 2008 Wenchuan MW 7.9 [21] and the 2013 Lushan MW 6.6 earthquakes [22]. After these events, Gui noted a decrease in stress levels in the Dayi area but an increase between Tianquan and Baoxin, suggesting dynamic stress redistributions, which have implications for future seismic activities [23]. South of Bayan Har, the Sichuan–Yunnan sub-block, emerging from the southeastern escape of the Qinghai–Tibet Plateau [24], is characterized predominantly be strike-slip and combined strike-slip with normal faults.
The Xianshui He–Anning He–Zemu He fault zone, marking the northern boundary of the Sichuan–Yunnan sub-block’s eastern edge, has been extensively studied due to its rich history of strong earthquakes [25]. Jin discussed the locking and slip deficit of the Xianshuihe fault, indicating that the varying locking depths across this fault have significant implications for earthquake potential, suggesting that different segments might respond differently to tectonic stresses [26]. Shao et al. examined the evolution of Coulomb stress over 200 years, noting that areas like Bamei, Selaha, and Kangding, which exhibit stress enhancement, are likely to experience strong earthquakes [27].
The Xianshui He fault belt (F1 in Figure 1b), commencing in Ganzi (GZ) and tapering off near Shimian (SM), stretches southeastward through Luhuo (LH), Daofu (DF), Kangding (KD), and Luding (LD), covering approximately 400 km. Since the late Cenozoic era, it has experienced notable sinistral strike-slip movement, amassing a total displacement of 60 km [28]. This fault belt is characterized by significant segmentation, with its slip rate diminishing from northwest to southeast, averaging around 10 mm/a [29,30,31,32,33,34,35].
To the north, the Anning He fault belt (F2 in Figure 1b) extends from Shimian (SM) and continues southward past Mianning (MN) to just south of Xichang (XC). Since the late Quaternary Period, it has displayed pronounced sinistral strike-slip and thrust movements. This 160 km long fault strikes north–south and dips either east or west, with angles ranging from 60° to 80° [33,36]. Zhu et al. developed a three-dimensional finite element model to analyze the stress state and fault segmentation within the Anninghe–Zemuhe–Xiaojiang faults region [37], finding that some locations with higher stress correspond to low-b-value areas, suggesting a complex interplay of factors that influence fault behavior [38]. The northern and southern segments exhibit slip rates of approximately 2.8–3.7 mm/a and 5–8 mm/a, respectively, with six of the eight recorded MS > 6.0 earthquakes occurring in its southern segment [39].
The Zemu He fault belt (F3 in Figure 1b) continues the sinistral strike-slip pattern of the Xianshui He and Anning He faults. Starting from Xichang (XC) and extending to Qiaojia (QJ) in the south, GPS studies indicate a slip rate of about 6 mm/a [40]. Paleoseismic research suggests that eight MS ≥ 7 earthquakes have occurred from the end of the Late Pleistocene Period to the present, with a recurrence interval estimated at approximately 1000 years [41].
Lastly, the Daliang Shan fault belt (F4 in Figure 1b), located east of the Anning He fault, is characterized by sinistral strike-slip movement. Since the Holocene Period, its slip rate has been recorded at approximately 3–4 mm/a [40,42,43,44,45,46]. Recent GPS analyses further highlight its evolving seismic potential by revealing substantial slip deficits and a strong potential for significant earthquakes [47].
The Luding earthquake, registering a magnitude of 6.8, struck in the Moxi area, located in the southeastern part of the Xianshui He fault belt. This area forms a crucial link between the northern termini of the Anning He and Daliang Shan fault belts [44,48]. The southeastern segment of the Xianshui He fault zone is notable for having one of the longest intervals without seismic activity in the history of the fault [25,49,50]. The accumulated strain, calculated from the slip rate and the elapsed time since the last significant event, is estimated to be between 2.3 and 3.2 m [49]. This led researchers, prior to the earthquake, to classify this segment as a high-risk area for a major seismic event, potentially reaching magnitudes of 7.0 to 7.37 [49,50].
The study region, outlined by polygons with blue lines in Figure 2, is situated in the northern portion of the eastern boundary of the Sichuan–Yunnan sub-block. The 2022 Luding MS 6.8 earthquake occurred centrally within this designated area. This area encompasses not only a section of the southern Longmen Shan fault belt but also the Xianshui He, Anning He, Zemu He, and Daliang Shan fault belts in its core. Collectively, these fault belts form the northern segment of the eastern boundary of the Sichuan–Yunnan sub-block, sharing a common dynamism and movement pattern—a left-lateral strike-slip motion attributed to the southeastward shift of the Qinghai–Tibet Plateau. The completion of the “post-disaster reconstruction project” following the 2009 Wenchuan MW 7.9 earthquake significantly enhanced earthquake monitoring capabilities in Sichuan [51]. To ensure data completeness and continuity, the year 2009 was selected as the starting point for our analysis.

4. Accurate Earthquake Locating

To ensure precise initial data for the subsequent b-value cross-section calculations, we need more seismic phase samples with more accurate arrival time data to form reasonable constraints on earthquake locating. We selected all ML ≥ 1.5 earthquakes, over 130,000, recorded by 230 stations across Sichuan from 2009 to the end of 2022. To streamline the computation and mitigate large-scale ellipsoid correction errors, we divided Sichuan into 2° × 2° sub-areas. An AI-based auto-picker constructed by the CNN network was applied to refine and supplement the phase arrival times [52]. To keep the phase arrival time more accurate, we double-checked the phases whose arrival time differences were equal to or greater than 0.5 s between the AI and manual phase files. This workflow was proven to be valid and effective in the MS 6.8 Luding earthquake sequence recognition [53]. We compared the results of manual and artificial intelligence recognition in a sub-area, and it can be seen that artificial intelligence corrected the errors of manual recognition and increased the amount of data by about 5% (Figure 3).
The earthquake location procedure adopted here is described in the following. Initially, Hypoinverse2000 [54] was employed for initial locating in each sub-area using a unified velocity model [55], yielding absolute location results. Subsequently, tomoDD [56] was used for high-precision relative locating, leveraging a detailed 0.2° × 0.2° three-dimensional velocity structure of the Sichuan–Yunnan area provided by Liu et al. [57]. We then amalgamated the location results from each sub-area, prioritizing the relative location results from tomoDD (marked as 2, black circle in Figure 4a). Events that were not amenable to relative locating due to factors like “air quakes” or connection loss were assigned absolute locating results from Hypoinverse2000 (marked as 1, green circle in Figure 4a). The few remaining earthquakes with fewer than three observations, which could not be numerically calculated for source location, were cataloged using their original parameters (marked as 0, purple circle in Figure 4a).
Ultimately, we identified 16,203 earthquakes with ML ≥ 1.5 in the study area. The distribution of these events by flags 0, 1, and 2 is 689 (4%), 1959 (12%), and 13555 (84%), respectively. The average horizontal error for events flagged as 2 is less than 1 km, and the vertical error ranges between 1 and 2 km, as determined by the bootstrap test [58]. Flag = 0 events predominantly occur at the boundaries of the seismic network. The average deviation between flag = 1 and flag = 2 events is 0.4 km horizontally and 1.3 km vertically. The focal depth is mostly within 10 km (Figure 4b), but it escalates to nearly 30 km in a small area south of Shimian (SM), a phenomenon previously reported [59] that is yet to be fully explained.
Given the occurrence of the Lushan MS 6.1 earthquake on 1 June 2022 along the southern section of the Longmen Shan fault belt adjacent to our study area, we designated the study period from 2009 to 31 May 2022 to preclude the influence of seismic sequences and mainshock-triggered effects. We computed the minimum magnitude of completeness (MC) using the Maximum Curvature method [60,61] with a 0.1° × 0.1° step and a 30 km search radius. The MC predominantly hovers around ML = 1.5 across the study area, coinciding with our earthquake locating work’s starting magnitude. It increases to about ML = 1.7 south of Ganzi (GZ) due to fewer stations, and, near Qiaojia (QJ) in the southern study area, the Mc ascends to ML 1.8–1.9, being at the seismic network’s edge (Figure 5a). Using the maximum likelihood method [11] and the MC-varied results from Figure 5a, we generated a b-value map, setting the sample’s lower limit in each grid to 30, with magnitudes exceeding Mc. This map reveals that most areas, including the future Luding MS 6.8 earthquake’s epicenter, fall within a normal range of 0.7~1.1. The only exception is the abnormally low-b-value area (b < 0.7) in the southern section of the Longmen Shan fault belt (Figure 5b).
Acknowledging the potential for underestimation of the minimum magnitude of completeness (MC) when using the Maximum Curvature method [62], we opted to set MC at a conservative ML = 2.0. The newly calculated b-value map (Figure 5c) revealed no significant deviation from the earlier version. To minimize the impact of arbitrary parameter selection on the outcomes, we employed the Hierarchical Space-Time Point-Process Model [63] for recalculating the b-value map, based on MC = ML 2.0 (Figure 5d). This approach yielded a smoother distribution of b-values, yet the overall trend remained consistent with the previous maps. Intriguingly, regardless of the method applied, the epicenter of the impending Luding MS 6.8 earthquake was not situated in a low-b-value zone.
This observation contradicts conventional wisdom, which typically associates strong earthquakes with low-b-value areas [6,7,64]. To unravel this anomaly, we investigated the stress distribution in the Xianshui He fault zone prior to the Luding MS 6.8 earthquake, examining the temporal and spatial characteristics of the b-value. The 2014 Kangding MS 6.3 earthquake, which occurred north of the Luding MS 6.8 event on the same fault belt, had exerted a positive Coulomb stress influence on the southern section of the Xianshui He fault belt [65]. To isolate the impact of this earthquake, we divided the study period into two phases, using 2015 as the demarcation point: the first phase spanned from 2009 to 2014, and the second from 2015 to 31 May 2022.
We applied a P-test based on the Akaike Information Criterion [66] to quantitatively evaluate the b-value trends across these two stages. As proposed by Utsu [67], the difference in AIC (i.e., ΔAIC) for the same statistical window between different stages can be defined as follows:
Δ A I C = 2 ( i = 1 2 N i ln b i ) 2
Ni and bi are the sample size and b-value in the ith window, respectively. The probability that the events in the two sample windows are from the same population can be derived from the AIC:
P = exp ( Δ A I C 2 2 )
A significant change in the b-value is considered to have occurred when ΔAIC ≥ 2 (P ≈ 0.05); when ΔAIC > 5 (P ≈ 0.01), it is judged to be highly significant.
The results show that, when we choose ΔAIC ≥ 2, the southern segment of the Xianshui He fault belt has the only large-scale anomaly area in the study area, with a radius of about 50 km and a ΔAIC value of around 5 (Figure 5a), corresponding to Δb ≤ −0.3, reflecting significant changes in b-values between the two periods of 2009–2014 and 2015–31 May 2022. Additionally, the b-value in the latter period significantly decreased compared to the former. The anomaly area was set at the intersection and bending parts of faults, where significant strains can accumulate [68,69]. The future Luding MS 6.8 earthquake (yellow star in Figure 6) is located in this anomaly area (Figure 6b). Considering that some earthquake cases indicate an increase in seismic activity before strong earthquakes [70,71,72,73] or some special spatiotemporal migration patterns [74,75], we analyzed the spatiotemporal distribution characteristics of earthquakes of ML ≥ 3.0 within a 100 km range, with the Luding MS 6.8 earthquake as the center (Figure 6c). We noticed that, since 2015, the seismicity rate of earthquakes of ML ≥ 3.0 within a range of 50 km has been higher than that of the background period, which is also the spatiotemporal domain where the b-value exhibits anomalies. Although we did not observe any obvious migration pattern before the mainshock, a small magnitude swarm near the future mainshock occurred from December 2021 to March 2022 (red rectangle in Figure 6c). More detailed research is needed to determine the relationship between these small earthquakes and the major earthquakes that will occur in the recent future that will occur six months later. The ΔAIC value of the adjacent southern segment of the Longmen Shan fault belt is as high as 15, indicating that seismic activity changes intensely there. However, the Δb-value in this area is positive. The Qiaojia (QJ) area in the southern part of the study area also has a high ΔAIC value, which may be due to weak monitoring capabilities at the boundary of the station network (Figure 6a).
Along the strike of the Xianshui He fault belt, we have drawn a seismic profile from Daofu (DF) to the south of Shimian (SM) (see Figure 3b for the location of the profile). The projected width of the profile is 30 km on each side, and the b-value is calculated using a rectangular statistical window with a horizontal width of 10 km and a vertical width of 3 km. The sliding increment in both directions is 1 km. At the same time, it is required that there be at least 30 samples with ML ≥ 2 in the statistical window.
The focal profile from 2009 to 2014 shows (Figure 7a) that the clustering characteristic is significant. The earthquakes from Daofu (DF) to Shimian (SM) are mainly distributed within a depth of 10 km (cluster1, C1), while the number of earthquakes within a depth of 8 km to the Shimian (SM) area is scarce, and the earthquakes are clustered between 8 and 14 km in depth (cluster2, C2). Another cluster is located south of Shimian (SM), which is clustered in the depth range of 17–30 km (cluster 3, C3). We noticed that the three clusters are not spatially continuous, especially since there is a seismic gap of about 20 km in length between C1 and C2 (Figure 7a dotted box), and the future Luding MS 6.8 earthquake sequence occurred near the boundary of this seismic gap. The entire profile does not show a significant low-b-value region in this period.
From 2015 to May 2022, the earthquake distribution showed significant changes compared to the previous period (Figure 7b): earthquakes in C1 extended into and broke the earthquake gap. At the same time, the scale of C2 and C3 expanded and connected. The b-value profile showed that, during this period, a 50 km low-b-value anomaly area (b < 0.7) formed in the lower boundary area of C1 near the seismic gap. The future Luding MS 6.8 earthquake is just located in this anomaly area. Combined with the co-seismic slip map (Figure 7c), we can see that the location, scale, and shape of the large co-seismic slip area are highly consistent with the low-b-value area, and the earthquake gap is filled with the Luding MS 6.8 aftershock. The low-b-value anomaly area and the seismic gap constitute this earthquake sequence’s rupture range.

5. Temporal Evolution of b-Value before the MS 6.8 Luding Earthquake

We selected the earthquake catalog within a 50 km radius centered on the Luding MS 6.8 earthquake (the abnormal region identified in Figure 6) from 2009 to 31 May 2022. Using a statistical window of 300 events and an increment of 1 event, we calculated the temporal evolution of b-values before the Luding MS 6.8 earthquake using the maximum likelihood method with MC = ML 2.0. The results showed that there had been a decline from the second half of 2012 to the end of 2021, from 1.3 in 2012 to 1.0 in 2022 (Figure 8a). Considering that the b-value error is around 0.1, we believe that this decline is significant. The main period of decrease in the b-value was from 2013 to 2018, during which more earthquakes with ML ≥ 3.0 occurred compared to other periods (Figure 8b). Since the end of 2021, there had been a slight rebound of b-values for about four months, with a magnitude of 0.05, corresponding to a small-magnitude earthquake cluster near the future mainshock (red rectangle in Figure 6c). This small-scale, short-term phenomenon has occurred multiple times in the past, but there has been no subsequent strong earthquake, so it is difficult to say that it was a precursor to the Luding MS 6.8 earthquake. The long-term decrease in b-values reflects the continuous accumulation of stress in the region; similar phenomena have occurred before many strong earthquakes [8,9].
To validate the observed phenomena, this study plotted the Gutenberg–Richter (G–R) relationship, fitting graphs for two different periods. Specifically, data from 2010 to 2013 are represented by red dots in Figure 9, while data from 2018 to 2021 are shown with blue dots. Comparison of Figure 9 clearly shows that, despite the identical duration and similar levels of seismic activity (in terms of both frequency and intensity) between the two periods, the later period recorded more earthquakes of magnitude 3.5 and above (4 compared to 3 in the earlier period), and the maximum magnitude increased from 4.0 in the earlier period to 4.9 in the later period. Consequently, the b-value for the later period dropped to 0.91, below the earlier period’s 1.18. This discrepancy is clearly reflected in the b-value temporal evolution curve, thereby supporting the reliability of the observed decrease in b-values over time.

6. Discussion

The occurrence of the Luding earthquake was inevitable, whether considering the historical seismic recurrence cycle or the high deformation-locking area. Since 1400, 23 MS ≥ 6.5 earthquakes have been recorded in the study region, and the rupture zones almost cover the entire boundary fault (Figure 10). Since the 21st century, there have been two MS ≥ 6.0 earthquakes on the Xianshui He fault belt, namely, the Kangding MS 6.3 in 2014 [76] and the Luding 6.8 in 2022 [77]. The section from Kangding (KD) to Shimian (SM) is located in the historic rupture zone of the Kangding south MS 7¾ in 1786; the epicenter of the Luding MS 6.8 earthquake is located at the southern end of the historic rupture zone [35]. The Xianshuihe Fault Zone, as one of the fault zones with the most frequent strong seismic activity, has a significant segmented seismic recurrence cycle. The Moxi section where the Luding MS 6.8 earthquake is located has a characteristic seismic recurrence cycle of 155 years [50]. Before the Luding MS 6.8 earthquake, the 236-year elapsed time had already passed far beyond the recurrence period.
Li et al. [78] explained this inevitability from another perspective. They used the DEFNODE algorithm [79] to invert the fault locking of the Xianshui He fault belt based on GNSS data from 2016 to 2020. The results showed that the Luding MS 6.8 earthquake sequence was located in the high-locking region of the southern section of the Xianshui He fault belt (PHI close to 1, Figure 11), indicating that the Luding seismic source area was in a state of stress accumulation during this period. The Luding MS 6.8 earthquake tore the southern part of the 1786 MS 7.75 rupture zone. However, considering the relatively low fault-locking degree, the strong earthquakes are not urgent in the northern part of the historic rupture zone (Figure 11).
Although numerous examples of earthquakes have demonstrated that the low-b-value area obtained from surface scanning is effective in predicting the location and magnitude of future strong earthquakes [5,6,7], our research indicates that neither the fixed-window method nor the non-fixed-window method (HIST-PPM) revealed a significant low-b-value state in the future epicenter area, resulting in false negatives, whereas Δb could somewhat solve this problem. It seems feasible to use AIC constraints to highlight the Δb-abnormal areas and delineate the location of future strong earthquakes from our example. In addition, using a precise location catalog to calculate the spatial and temporal b-value profiles can help discover the regions and periods of b-value anomalies from different perspectives.
The co-seismic stress transfer produced by large earthquakes has a profound impact on the occurrence of successive earthquakes [80], while the Changes in Coulomb Failure Stress (ΔCFS) can describe the transfer pattern quantitatively, which is defined as ΔCFS = Δτ + μΔσ, where Δτ is the change in shear stress, Δσ is normal stress change, and μ is the coefficient of friction [81]. We took 10 MS ≥ 6.5 earthquakes in the Sichuan region since 1973 as sources (including the 2022 Luding MS 6.8 earthquake, [82]). We used Coulomb 3.3 [83] to calculate the ΔCFS maps in our study region based on the Xianshui He fault belt (Figure 12a) and the Anning He fault belt (Figure 12b) as the receiving faults. Two segments with significant stress increases can be found in the two maps: the DF–KD segment in the Xianshui He fault belt and south of SM in the northern Anning He fault belt. The ΔCFS of the two segments is greater than the trigger threshold of 0.01 MPa [84,85], which could, therefore, infer that the segments carry a high risk of strong earthquakes.

7. Conclusions

In the retrospective analysis of the 2022 Luding MS 6.8 earthquake, the event appeared inevitable based on the recurrence interval of strong earthquakes in the region. However, conventional surface b-value scanning prior to the earthquake did not detect any low-b-value anomalies at the epicenter. This lack of detection resulted in a prediction failure.
To gain a deeper understanding of the b-value response before the earthquake and to assess the potential risk of strong earthquakes in the study area, we divided the earthquake catalog into two stages, from 2009 to 2014 and from 2015 to May 2022, and conducted segmented analysis. The AIC-constrained Δb (change in b-value) effectively revealed significant areas of b-value change, pointing out potential locations for future strong earthquakes. The Δb and ΔAIC maps calculated for these two periods show that, under the condition of ΔAIC ≥ 2, there is a clear abnormal Δb area with a radius of about 50 km (Δb < −0.3) where the Luding MS 6.8 earthquake occurred. This low-Δb-value anomaly area provides a spatial reference for the spatiotemporal evolution analysis of b-values.
Furthermore, the b-value profiles derived from precise relocation show that the abnormal low-b-value area corresponds to the position, size, and shape of the mainshock’s large-slip co-seismic zone, confirming that b-values can reveal areas of stress concentration on faults and that the aftershocks triggered by the mainshock filled the seismic gaps that existed before the earthquake. The continuous decline in b-values over the past decade reveals that the Luding MS 6.8 earthquake source area has undergone a long-term stress-loading process, a pattern that is consistent with the b-value evolution before other major earthquakes and the fault-locking results inverted from GNSS data. Given that b-values are dynamic functions of time and space, one-dimensional analysis may lose multidimensional information; therefore, comprehensive analysis from multiple dimensions such as surface, cross-section, and time is crucial for accurately identifying b-value anomalies.
Finally, the Coulomb failure stress change analysis shows that 10 historical significant earthquakes, including the Luding MS 6.8 earthquake, have increased the stress along the Dadukou (DF) to Kangding (KD) section of the Xianshuihe fault zone south of Ya’an (SM) and the northern section of the Anning River fault zone, with the stress magnitude exceeding the threshold of 0.01 MPa, indicating that these areas still have the potential for strong earthquakes.

Author Contributions

Conceptualization, F.L. and L.P.; methodology, F.L., L.P., X.R. and R.W.; software, F.L. and L.P.; validation, L.P., M.Z., X.R., D.W., R.W., W.W. and C.H.; formal analysis, L.P.; investigation, L.P., M.Z. and F.L.; resources, F.L.; data curation, M.Z., X.R., D.W., R.W., W.W. and C.H.; writing—original draft preparation, L.P.; writing—review and editing, M.Z. and F.L.; visualization, L.P.; supervision, F.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Key R&D Program of China (No. 2021YFC3000702-05) and the Spark Program of Earthquake Sciences (No. XH23033A).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used in the article is currently unavailable online.

Acknowledgments

We express our gratitude to Professor Jiang Changsheng for his insightful discussions on b-value analysis, and to Wang Xun for providing the inversion results of co-seismic slip. Most of the figures in this paper have been produced using the GMT software (version 4.5) [86].

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mogi, K. The fracture of a semi-infinite body caused by an inner stress origin and its relation to the earthquake phenomena (Second Paper). Bull. Earthq. Res. Inst. 1963, 41, 595–614. [Google Scholar]
  2. Scholz, C. Microfractures, aftershocks, and seismicity. Bull. Seismol. Soc. Am. 1968, 58, 1117–1130. [Google Scholar]
  3. Amelung, F.; King, G. Earthquake scaling laws for creeping and non-creeping faults. Geophys. Res. Lett. 1997, 24, 507–510. [Google Scholar] [CrossRef]
  4. Hainzl, S.; Fischer, T. Indications for a successively triggered rupture growth underlying the 2000 earthquake swarm in Vogtland/NW Bohemia. J. Geophys. Res. 2002, 107, ESE 5-1–ESE 5-9. [Google Scholar] [CrossRef]
  5. Schorlemmer, D.; Wiemer, S.; Wyss, M. Variations in earthquake-size distribution across different stress regimes. Nature 2005, 437, 539–542. [Google Scholar] [CrossRef]
  6. Wyss, M.; Schorlemmer, D.; Wiemer, S. Mapping asperities by minima of local recurrence time: San Jacinto-Elsinore fault zones. J. Geophys. Res. Atmos. 2000, 105, 7829–7844. [Google Scholar] [CrossRef]
  7. Wiemer, S.; Wyss, M. Mapping the frequency–magnitude distribution in asperities: An improved technique to calculate recurrence times. J. Geophys. Res. 1997, 102, 15115–15128. [Google Scholar] [CrossRef]
  8. Shi, H.X.; Meng, L.Y.; Zhang, X.M.; Chang, Y.; Yang, Z.T.; Xie, W.Y.; Fu, B.K.J.; Han, P. Decrease in b value prior to the Wenchuan earthquake (MS 8.0). Chin. J. Geophys. 2018, 61, 1874–1882. (In Chinese) [Google Scholar]
  9. Nanjo, K.Z.; Hirata, N.; Obara, K.; Kasahara, K. Decade-scale decrease in b value prior to the M9-class 2011 Tohoku and 2004 Sumatra quakes. Geophys. Res. Lett. 2012, 39, L20304. [Google Scholar] [CrossRef]
  10. Gutenberg, B.; Richter, C.F. Frequency of earthquakes in California. Bull. Seismol. Soc. Am. 1944, 34, 185–188. [Google Scholar] [CrossRef]
  11. Aki, K. Maximum likelihood estimate of b in the formula log N = a-bM and its confidence limits. Bull. Earthq. Res. Inst. Univ. Tokyo 1965, 43, 237–238. [Google Scholar]
  12. Utsu, T. A method for determining the value of b in a formula log n = a bM showing the magnitude-frequency relation for earthquakes. Geophys. Bull. Hokkaido Univ. 1965, 13, 99–103. [Google Scholar]
  13. Utsu, T. A statistical significance test of the difference in b-value between two earthquake groups. J. Phys. Earth 1966, 14, 37–40. [Google Scholar] [CrossRef]
  14. Tassara, A.; Sippl, C.; Riedel, M.; Castro, C.; Carcamo, F. Nature of asperities and barriers along the Chilean megathrust unveiled by an integrated analysis of seismicity, gravity, geodetic locking and wedge geometry. In Proceedings of the EGU General Assembly 2024, Vienna, Austria, 14–19 April 2024; EGU24-14050. [Google Scholar] [CrossRef]
  15. Tormann, T.; Enescu, B.; Woessner, J.; Wiemer, S. Randomness of megathrust earthquakes implied by rapid stress recovery after the Japan earthquake. Nat. Geosci. 2015, 58, 152–158. [Google Scholar] [CrossRef]
  16. Fu, Z.; Chen, F.; Deng, J.; Zhao, S.; Li, H.; Dai, S.; Shao, Y.; Fu, Y.; Zhu, J.; Cheng, W. Statistical investigation of induced seismicity associated with the impoundment of the Xiangjiaba Reservoir, Southwestern China. Bull. Eng. Geol. Environ. 2024, 83, 106. [Google Scholar] [CrossRef]
  17. Ito, R.; Kaneko, Y. Physical Mechanism for a Temporal Decrease of the Gutenberg-Richter b-Value Prior to a Large Earthquake. J. Geophys. Res. Solid Earth 2023, 128, e2023JB027413. [Google Scholar] [CrossRef]
  18. Niu, F.J.; Luo, J.; Lin, Z.J.; Fang, J.H.; Liu, M.H. Thaw-induced slope failures and stability analyses in permafrost regions of the Qinghai-Tibet Plateau, China. Landslides 2016, 13, 55–65. [Google Scholar] [CrossRef]
  19. Zhang, P.Z.; Zheng, D.W.; Yin, G.M.; Yuan, D.Y.; Zhang, G.L.; Li, C.Y.; Wang, Z.C. Discussion on the late Cenozoic extension and uplift of the northeastern margin of the Qinghai-Tibet Plateau. Quat. Res. 2006, 26, 5–13. (In Chinese) [Google Scholar]
  20. Burchfiel, B.C.; Chen, Z.; Liu, Y.; Royden, L.H. Tectonics of the Longmen Shan and adjacent regions, Central China. Int. Geol. Rev. 1995, 37, 661–735. [Google Scholar] [CrossRef]
  21. Wei, Z.; Yang, S.; Chen, X.F. Numerical simulation of strong ground motion for the Ms8.0 Wenchuan earthquake of 12 May 2008. Sci. China Ser. D Earth Sci. 2008, 51, 1673–1682. [Google Scholar] [CrossRef]
  22. Long, F.; Wen, X.; Ruan, X.; Zhao, M.; Yi, G. A more accurate relocation of the 2013 Ms7.0 Lushan, Sichuan, China, earthquake sequence, and the seismogenic structure analysis. J. Seismol. 2015, 19, 653–665. [Google Scholar] [CrossRef]
  23. Gui, Y. Stress state and major-earthquake risk on the southern segment of the Longmen Shan fault zone. Chin. J. Geophys. 2013, 56, 1112–1120. [Google Scholar]
  24. He, X.; Zhao, L.F.; Xie, X.B.; Tian, X.; Yao, Z.X. Weak crust in southeast Tibetan Plateau revealed by Lg-wave attenuation tomography: Implications for crustal material escape. J. Geophys. Res. Solid Earth 2021, 126, e2020JB020748. [Google Scholar] [CrossRef]
  25. Wen, X.Z.; Fan, J.; Yi, G.X.; Deng, Y.W.; Long, F. Seismic gaps on the Anninghe Fault in Western Sichuan. Sci. China Ser. D Earth Sci. 2008, 38, 797–807. (In Chinese) [Google Scholar] [CrossRef]
  26. Jin, Z. Study on dynamic characteristics of fault locking and fault slip deficit in the eastern boundary of the Sichuan-Yunnan rhombic block. Chin. J. Geophys. 2015, 58, 872–885. [Google Scholar]
  27. Shao, Z.; Xu, J.; Ma, H.-S.; Zhang, L.-P. Coulomb stress evolution over the past 200 years and seismic hazard along the Xianshuihe fault zone of Sichuan, China. Tectonophysics 2016, 670, 48–65. [Google Scholar] [CrossRef]
  28. Wang, E.; Burchfiel, B.C.; Royden, L.H.; Chen, L.Z.; Chen, J.S.; Li, W.X.; Chen, Z.L. Late Cenozoic Xianshuihe-Xiaojiang, Red River, and Dali fault systems of southwestern Sichuan and central Yunnan, China. Spec. Pap. Geol. Soc. Am. 1998, 327, 1–108. [Google Scholar]
  29. Qian, H.; Allen, C.R.; Luo, Z.L.; Wen, X.Z.; Zhou, H.W.; Huang, W.S. Activity characteristics of the Xianshuihe Fault since the Holocene. Earthq. Res. China 1988, 4, 9–18. (In Chinese) [Google Scholar]
  30. Wen, X.Z.; Allen, C.R.; Luo, Z.L.; Qian, H.; Zhou, H.W.; Huang, W.S. Segmentation.geometric features, and their seismotectonic implications for theHolocene Xianshuihe fault zone. Acta Seismol. Sin. 1989, 11, 362–372. (In Chinese) [Google Scholar]
  31. Wen, X.Z. Variable rupture scale seismic behavior of active faults and application of cascade rupture models. Acta Seismol. Sin. 2001, 23, 380–390. (In Chinese) [Google Scholar]
  32. Allen, C.R.; Luo, Z.L.; Qian, H.; Wen, X.Z.; Zhou, H.W.; Huang, W.S. Field study of a highly active fault zone: The Xianshuihe fault of Southwestern China. Geol. Soc. Am. Bull. 1991, 103, 1178–1199. [Google Scholar] [CrossRef]
  33. Tang, R.C.; Han, W.B. Active Faults and Earthquakes in Sichuan Province; Seismological Press: Beijing, China, 1993; pp. 1–368. (In Chinese) [Google Scholar]
  34. Zhou, R.J.; He, Y.L.; Huang, Z.Z.; Li, X.G.; Yang, T. The slip rate and strong earthquake recurrence interval on the Qianning-Kangding segment on the Xianshuihe fault zone. Acta Seismol. Sin. 2001, 23, 250–261. (In Chinese) [Google Scholar] [CrossRef]
  35. Zhou, R.J.; He, Y.L.; Yang, T.; He, Q.; Li, X.G. Slip Rate and Strong Earthquake Rupture on the Mox-i Mianning Segment Along the Xianshuihe-Anninghe Fault Zone. Earthq. Res. China 2001, 17, 25–34. (In Chinese) [Google Scholar]
  36. Luo, S.; Yao, H.; Wen, J.; Yang, H.; Tian, B.; Yan, M. Apparent low-velocity belt in the shallow Anninghe fault zone in SW China and its implications for seismotectonics and earthquake hazard assessment. J. Geophys. Res. Solid Earth 2023, 128, e2022JB025681. [Google Scholar] [CrossRef]
  37. Zhu, A.; Zhang, D.; Jiang, C. Numerical simulation of the segmentation of the stress state of the Anninghe-Zemuhe-Xiaojiang faults. Sci. China Earth Sci. 2016, 59, 384–396. [Google Scholar] [CrossRef]
  38. Liu, Y.; Zhao, G.; Wu, Z.; Jiang, Y. An analysis of b-value characteristics of earthquake on the southeastern margin of the Tibetan Plateau and its neighboring areas. Geol. Bull. China 2015, 34, 58–70. [Google Scholar]
  39. Ran, Y.K.; Chen, L.C.; Cheng, J.W.; Gong, H.L. Late Quaternary Surface Deformation and Rupture Behavior of strong Earthquake on the Segment north of Mianning of the Anninghe Fault. Sci. China Ser. D 2008, 51, 1224–1237. [Google Scholar] [CrossRef]
  40. Shen, Z.K.; Lü, J.N.; Wang, M.; Roland, B. Contemporary crustal deformation around the southeast borderland of the Tibetan Plateau. J. Geophys. Res. 2005, 110, B11409. [Google Scholar] [CrossRef]
  41. Du, P.S. Late Quaternary activity characteristics and paleoseismic study of the Zemuhe Fault Zone. Sichuan Earthq. 1994, 1, 42–52. (In Chinese) [Google Scholar]
  42. Zhou, R.J.; Li, X.G.; Huang, Z.Z.; He, Y.L.; Ge, T.Y. Late Quaternary average slip rate of the Daliangshan Fault Zone in Sichuan. J. Seismol. Res. 2003, 26, 191–199. (In Chinese) [Google Scholar]
  43. Xu, X.W.; Wen, X.Z.; Zheng, R.Z.; Ma, W.T.; Song, F.M.; Yu, G.H. Recent tectonic movement pattern and its dynamics in the Sichuan-Yunnan active block region. Sci. China Ser. D Earth Sci. 2003, 33, 151–162. (In Chinese) [Google Scholar]
  44. He, H.L.; Ikeda, Y.; He, Y.L.; Togo, M.; Chen, J.; Chen, C.Y.; Tajikara, M.; Echigo, T.; Okada, S. Newly generated Daliangshan fault zone: Shortcutting on the central section of Xianshuihe-Xiaojiang fault system. Sci. China Ser. D Earth Sci. 2008, 51, 1248–1258. [Google Scholar] [CrossRef]
  45. Wei, Z.Y.; He, H.L.; Shi, F.; Xu, Y.R.; Bi, S.L.; Sun, H.Y. Slip rate estimation of the southern segment of the Daliangshan Fault Zone. Seismol. Geol. 2012, 34, 282–293. (In Chinese) [Google Scholar]
  46. Sun, H.Y.; He, H.L.; Wei, Z.Y.; Gao, W. Late Quaternary activity of the northern section of the eastern branch of the Daliangshan Fault Zone—The Zhuma Fault. Seismol. Geol. 2015, 37, 440–454. (In Chinese) [Google Scholar]
  47. Li, Y.; Hao, M.; Song, S.; Zhu, L.; Cui, D.; Zhuang, W.; Yang, F.; Wang, Q. Interseismic fault slip deficit and coupling distributions on the Anninghe-Zemuhe-Daliangshan-Xiaojiang fault zone, southeastern Tibetan Plateau, based on GPS measurements. J. Asian Earth Sci. 2021, 219, 104899. [Google Scholar] [CrossRef]
  48. Chen, G.H.; Min, W.; Song, F.M.; Jiao, D.C.; Xu, H.T. Preservation of co-seismic surface rupture in different geomorphological settings from the study of the 1786 Moxi earthquake. Seismol. Geol. 2011, 33, 804–817. (In Chinese) [Google Scholar]
  49. Bai, M.; Chevalier, M.L.; Leloup, P.H.; Li, H.; Pan, J.; Replumaz, A.; Wang, S.G.; Li, K.Y.; Wu, Q.; Liu, F.C.; et al. Spatial slip rate distribution along the SE Xianshuihe fault, eastern Tibet, and earthquake hazard assessment. Tectonics 2021, 40, e2021TC006985. [Google Scholar] [CrossRef]
  50. Qiao, X.; Zhou, Y. Geodetic imaging of shallow creep along the Xianshuihe fault and its frictional properties. Earth Planet. Sci. Lett. 2021, 567, 117001. [Google Scholar] [CrossRef]
  51. Long, F.; Wen, X.Z.; Ni, S.D. Determination of the spatiotemporal distribution of the regional minimum completeness magnitude: A case study of the Longmenshan Fault Zone. Earthquake 2009, 29, 27–36. (In Chinese) [Google Scholar]
  52. Yu, Z.Y.; Wang, W.T.; Chen, Y.N. Benchmark on the accuracy and efficiency of several neural network based phase pickers using datasets from China Seismic Network. Earthq. Sci. 2023, 36, 113–131. [Google Scholar] [CrossRef]
  53. Liu, Z.M.; Li, J.; Su, J.R.; Yu, Z.Y.; Huang, C.M.; Qi, Y.P.; Chen, Y.N.; Wang, W.T. Auto-building of aftershock sequence and seismicity analysis for the 2022 Luding, Sichuan, MS6.8 earthquake. Chin. J. Geophys. 2023, 66, 1976–1990. (In Chinese) [Google Scholar] [CrossRef]
  54. Klein, F.W. User’s Guide to HYPOINVERSE, a Program for VAX Computers to Solve for Earthquake Locations and Magnitudes; U. S. Geological Survey Open-File Report; United States Geological Survey: Reston, VA, USA, 1989; pp. 89–314.
  55. Zhao, Z.; Zhang, R.S. The compilation of regional travel time table in Sichuan. Earthq. Res. Sichuan 1987, 3, 29–35. (In Chinese) [Google Scholar]
  56. Zhang, H.; Thurber, C. Double-difference tomography: The method and its application to the Hayward fault, California. Bull. Seismol. Soc. Am. 2003, 93, 1875–1889. [Google Scholar] [CrossRef]
  57. Liu, Y.; Yu, Z.; Zhang, Z.; Yao, H.; Wang, W.; Zhang, H.; Fang, H.; Fang, L. The high-resolution community velocity model V2.0 of southwest China, constructed by joint body and surface wave tomography of data recorded at temporary dense arrays. Sci. China Earth Sci. 2023, 66, 2368–2385. [Google Scholar] [CrossRef]
  58. Long, F.; Jiang, C.S.; Qi, Y.P.; Liu, Z.F.; Fu, Y. A joint probabilistic approach for merging earthquake catalogs of two neighboring seismic networks: An example of the 2014 Ludian sequence catalog. Chin. J. Geophys. 2018, 61, 2815–2827. (In Chinese) [Google Scholar]
  59. Fu, Y.; Long, F.; Wang, S.Y. Precise location of Earthquakes at the eastern boundaries of diamond block of Sichuan-Yunnan. Earthq. Res. China 2018, 34, 60–70. (In Chinese) [Google Scholar]
  60. Wyss, M.; Hasegawa, A.; Wiemer, S.; Umino, N. Quantitative mapping of precursory seismic quiescence before the 1989, m7.1 off-sanriku earthquake, Japan. Ann. Geophys. 1999, 42, 851–869. [Google Scholar] [CrossRef]
  61. Wiemer, S.; Wyss, M. Minimum magnitude of complete reporting in earthquake catalogs: Examples from Alaska, the western United States, and Japan. Bull. Seismol. Soc. Am. 2000, 90, 859–869. [Google Scholar] [CrossRef]
  62. Woessner, J.; Wiemer, S. Assessing the Quality of Earthquake Catalogues: Estimating the Magnitude of Completeness and Its Uncertainty. Bull. Seismol. Soc. Am. 2005, 95, 684–698. [Google Scholar] [CrossRef]
  63. Ogata, Y.; Katsura, K.; Tanemura, M.; Harte, D.; Zhuang, J. Hierarchical Space-Time Point-Process Models (HIST-PPM): Software Documentation. 2020. Available online: http://bemlar.ism.ac.jp/ogata/HIST-PPM-V3/ (accessed on 3 May 2023).
  64. Wyss, M. Locked and creeping patches along the Hayward Fault, California. Geophys. Res. Lett. 2001, 28, 3537–3540. [Google Scholar] [CrossRef]
  65. Xiong, W.; Tan, K.; Yu, P.F.; Chen, W. Triggering of Mw5.9 Kangding earthquake by Coulomb stress evolution along Xianshuihe fault zone since 1955. J. Geod. Geodyn. 2016, 36, 95–100. (In Chinese) [Google Scholar]
  66. Akaike, H. A new look at the statistical model identification. Sel. Pap. Hirotugu Akaike 1974, 19, 215–222. [Google Scholar]
  67. Utsu, T. Introduction to seismicity, Mathematical Seismology. Inst. Stat. Math. 1992, 7, 139–157. (In Japanese) [Google Scholar]
  68. Fischer, M.P.; Wilkerson, M.S. Predicting the orientation of joints from fold shape: Results of pseudothree-dimensional modeling and curvature analysis. Geology 2000, 28, 15–18. [Google Scholar] [CrossRef]
  69. Keating, D.P.; Fischer, M.P. An experimental evaluation of the curvature-strain relation in fault-related folds. AAPG Bull. 2008, 92, 869–884. [Google Scholar] [CrossRef]
  70. Sornette, A.; Sornette, D. Earthquake rupture as a critical point: Consequences for telluric precursors. Tectonophysics 1990, 179, 327–334. [Google Scholar] [CrossRef]
  71. Sykes, L.R.; Jaume, S. Seismic activity on neighboring faults as a long term precursor to large earthquakes in the San Francisco Bay area. Nature 1990, 348, 595–599. [Google Scholar] [CrossRef]
  72. Sornette, D.; Sammis, C.G. Complex critical exponents from renormalization group theory of earthquakes: Implications for earthquake predictions. J. Phys. I 1995, 5, 607–619. [Google Scholar] [CrossRef]
  73. Bowman, D.D.; King, G.C.P. Accelerating seismicity and stress accumulation before large earthquakes. Geophys. Res. Lett. 2001, 28, 4039–4042. [Google Scholar] [CrossRef]
  74. Martínez-Garzón, P.; Poli, P. Cascade and pre-slip models oversimplify the complexity of earthquake preparation in nature. Commun. Earth Environ. 2024, 5, 120. [Google Scholar] [CrossRef]
  75. Kato, A.; Obara, K.; Igarashi, T.; Tsuruoka, H.; Nakagawa, S.; Hirata, N. Propagation of Slow Slip Leading Up to the 2011 MW 9.0 Tohoku-Oki Earthquake. Science 2012, 335, 705–708. [Google Scholar] [CrossRef]
  76. Yi, G.X.; Long, F.; Wen, X.Z.; Liang, M.J.; Wang, S.W. Seismogenic structure of the M6.3 Kangding earthquake sequence on 22 Nov. 2014, Southwestern China. Chin. J. Geophys. 2015, 58, 1205–1219. (In Chinese) [Google Scholar] [CrossRef]
  77. Yi, G.X.; Long, F.; Wen, X.Z.; Liang, M.J.; Wang, S.W. Seismogenic structure of the MS 5.8 and MS 6.0 Barkam earthquake swarm in Sichuan Province on June 10, 2022. Seismol. Geomagn. Obs. Res. 2022, 43, 424–425. (In Chinese) [Google Scholar]
  78. Li, L.Y.; Wu, Y.Q.; Li, Y.J.; Zhan, W.; Liu, X.Z. Dynamic deformation and fault locking of the Xianshuihe Fault Zone, Southeastern Tibetan Plateau: Implications for seismic hazards. Earth Planets Space 2022, 74, 35. [Google Scholar] [CrossRef]
  79. McCafrey, R.; Stein, S.; Freymueller, J. Crustal block rotations and plate coupling. Plate Bound. Zones Geodyn. Ser. 2002, 30, 101–122. [Google Scholar]
  80. Stein, R.S. Earthquake conversations. Sci. Am. 2003, 288, 72–79. [Google Scholar] [CrossRef]
  81. King, G.C.P.; Stein, R.S.; Lin, J. Static stress changes and thetriggering of earthquakes. Bull. Seismol. Soc. Am. 1994, 84, 935–953. [Google Scholar]
  82. Ran, X.Y.; Zhu, H.; Long, F.; Li, L.H.; Wang, D. Analysis of Coulomb stress changes pre- and after the 2022 Luding MS 6.8 earthquake. Earthq. Res. Sichuan 2022, 185, 15–20. (In Chinese) [Google Scholar]
  83. Toda, S.; Stein, R.S.; Sevilgen, V.; Lin, J. Coulomb 3.3 Graphic-Rich Deformation and Stress-Change Software for Earthquake, Tectonic, and Volcano Research and Teaching-User Guide; U.S. Geological Survey Open-File Report 2011-1060; United States Geological Survey: Reston, VA, USA, 2011; 63p. Available online: https://pubs.usgs.gov/of/2011/1060/ (accessed on 1 September 2023).
  84. Toda, S.; Stein, R.S.; Reasenberg, P.A.; Dieterich, J.H.; Yoshida, A. Stress transferred by the 1995 Mw =6.9 Kobe, Japan, shock: Effect on aftershocks and future earthquake probabilities. J. Geophys. Res. Solid Earth 1998, 103, 24543–24565. [Google Scholar] [CrossRef]
  85. Stein, R.S. The role of stress transfer in earthquake occurrence. Nature 1999, 402, 605–609. [Google Scholar] [CrossRef]
  86. Wessel, P.; Smith, W.H.F. New, improved version of generic mapping tools released. Eos Trans. Am. Geophys. Union 1998, 79, 579. [Google Scholar] [CrossRef]
Figure 1. (a) Map of tectonic units in the Qinghai–Tibet plateau (modified from [11]). The yellow solid lines indicate the boundaries of the blocks, while the blue ones illustrate the borders of the sub-blocks. The grey circles show the historic M ≥ 6.5 events, and the red circle marks the position of the MS 6.8 Luding earthquake. The black arrows point out the movement directions of the blocks. The black arrows point out the movement directions of the blocks. B1: Lhasa sub-block. B2: Qiangtang sub-block. B3: Sichuan–Yunnan sub-block. B4: South China sub-block. B5: Bayan Har sub-block. B6: Qaidam sub-block. B7: Qilian sub-block. B8: Ordos sub-block. B9: Alashan sub-block. B10: Tarim sub-block. (b) From this perspective of the Sichuan–Yunnan sub-block and its surroundings. The blue dashed lines represent the boundary of the Sichuan–Yunnan sub-block, and the solid green lines envelop our study region in this work. The grey circles show the historic M ≥ 6.5 events, and the red star marks the MS 6.8 Luding earthquake. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt, F5: Longmen Shan fault belt.
Figure 1. (a) Map of tectonic units in the Qinghai–Tibet plateau (modified from [11]). The yellow solid lines indicate the boundaries of the blocks, while the blue ones illustrate the borders of the sub-blocks. The grey circles show the historic M ≥ 6.5 events, and the red circle marks the position of the MS 6.8 Luding earthquake. The black arrows point out the movement directions of the blocks. The black arrows point out the movement directions of the blocks. B1: Lhasa sub-block. B2: Qiangtang sub-block. B3: Sichuan–Yunnan sub-block. B4: South China sub-block. B5: Bayan Har sub-block. B6: Qaidam sub-block. B7: Qilian sub-block. B8: Ordos sub-block. B9: Alashan sub-block. B10: Tarim sub-block. (b) From this perspective of the Sichuan–Yunnan sub-block and its surroundings. The blue dashed lines represent the boundary of the Sichuan–Yunnan sub-block, and the solid green lines envelop our study region in this work. The grey circles show the historic M ≥ 6.5 events, and the red star marks the MS 6.8 Luding earthquake. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt, F5: Longmen Shan fault belt.
Applsci 14 04345 g001
Figure 2. Map of the ML ≥ 1.5 epicenters from 2009 to May 2022 in the Sichuan region (grey circles), active faults (red lines), stations used in earthquake relocating (green triangles), and the study region in this paper (blue lines). The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. The red star shows the epicenter of the MS 8.0 Wenchuan earthquake on 12 May 2008. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt, F5: Longmen Shan fault belt, F6: Jinsha Jiang fault belt, F7: Huaying Shan fault belt.
Figure 2. Map of the ML ≥ 1.5 epicenters from 2009 to May 2022 in the Sichuan region (grey circles), active faults (red lines), stations used in earthquake relocating (green triangles), and the study region in this paper (blue lines). The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. The red star shows the epicenter of the MS 8.0 Wenchuan earthquake on 12 May 2008. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt, F5: Longmen Shan fault belt, F6: Jinsha Jiang fault belt, F7: Huaying Shan fault belt.
Applsci 14 04345 g002
Figure 3. The Wadadi diagram in a sub-area: (a) manual picking; (b) refined and supplemented by the AI model. Red dots: P wave data, blue dots: S wave data.
Figure 3. The Wadadi diagram in a sub-area: (a) manual picking; (b) refined and supplemented by the AI model. Red dots: P wave data, blue dots: S wave data.
Applsci 14 04345 g003
Figure 4. The relocated epicentral map in the study region. (a) colored by the earthquake-locating method, black: flag = 2, green: flag = 1, purple: flag = 0; (b) colored by focal depth, the black solid line indicates the position of the profile line. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Figure 4. The relocated epicentral map in the study region. (a) colored by the earthquake-locating method, black: flag = 2, green: flag = 1, purple: flag = 0; (b) colored by focal depth, the black solid line indicates the position of the profile line. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Applsci 14 04345 g004
Figure 5. (a) Map of magnitude completeness; (b) map of b-value, derived from an MC-various mode in (a); (c) map of b-value, derived from a fixed MC = 2.0; (d) map of b-value, calculated by the HIST-PPM method from a fixed MC = 2.0. The black dots in the maps are the relocated ML ≥ 1.5 quakes from 2009 to May 2022. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt. The red arrows indicate the movement directions of the blocks.
Figure 5. (a) Map of magnitude completeness; (b) map of b-value, derived from an MC-various mode in (a); (c) map of b-value, derived from a fixed MC = 2.0; (d) map of b-value, calculated by the HIST-PPM method from a fixed MC = 2.0. The black dots in the maps are the relocated ML ≥ 1.5 quakes from 2009 to May 2022. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt. The red arrows indicate the movement directions of the blocks.
Applsci 14 04345 g005
Figure 6. (a) Map of ΔAIC—only the ΔAIC ≥ 2 areas are plotted; (b) map of Δb—only the corresponding areas with ΔAIC ≥ 2 are plotted. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. The solid circle indicates a 100 km-radius study region of the space-time diagram, which is portrayed in figure (c). The dashed box is corresponding to the 50 km-radius abnormal region of the b-value from 2015 to 2022. The red rectangle indicates the position of the nearest cluster of ML ≥ 2 to the Luding MS 6.8 earthquake. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Figure 6. (a) Map of ΔAIC—only the ΔAIC ≥ 2 areas are plotted; (b) map of Δb—only the corresponding areas with ΔAIC ≥ 2 are plotted. The yellow star shows the epicenter of the MS 6.8 Luding earthquake on 5 September 2022. The solid circle indicates a 100 km-radius study region of the space-time diagram, which is portrayed in figure (c). The dashed box is corresponding to the 50 km-radius abnormal region of the b-value from 2015 to 2022. The red rectangle indicates the position of the nearest cluster of ML ≥ 2 to the Luding MS 6.8 earthquake. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Applsci 14 04345 g006
Figure 7. Profiles of focal depth and b-value; the position of the profile is in Figure 3b: (a) 2009~2014, (b) 2015~31 May 2022, (c) the cross-section of the mainshock’s co-seismic slip and the earthquake sequence’s profile. The co-seismic slip results are provided by Dr. Wang Xun. The gray circles mark the earthquakes in the profile, dashed boxes indicate the seismic gap, and the red stars show the position of the MS 6.8 Luding earthquake.
Figure 7. Profiles of focal depth and b-value; the position of the profile is in Figure 3b: (a) 2009~2014, (b) 2015~31 May 2022, (c) the cross-section of the mainshock’s co-seismic slip and the earthquake sequence’s profile. The co-seismic slip results are provided by Dr. Wang Xun. The gray circles mark the earthquakes in the profile, dashed boxes indicate the seismic gap, and the red stars show the position of the MS 6.8 Luding earthquake.
Applsci 14 04345 g007
Figure 8. (a) The temporal b-value curve in the Luding area from 2009 to May 2022. The red line indicate the b-value, while the gray lines mark the errors of the b-value. The gray box marks the data accumulation period. (b) The M-t diagram with ML ≥ 2.0. The dashed box indicates a cluster (illustrated in Figure 5c, marked by a red rectangle) between December 2021 and March 2022, before the MS 6.8 Luding earthquake, and a rise of b-value at the same time.
Figure 8. (a) The temporal b-value curve in the Luding area from 2009 to May 2022. The red line indicate the b-value, while the gray lines mark the errors of the b-value. The gray box marks the data accumulation period. (b) The M-t diagram with ML ≥ 2.0. The dashed box indicates a cluster (illustrated in Figure 5c, marked by a red rectangle) between December 2021 and March 2022, before the MS 6.8 Luding earthquake, and a rise of b-value at the same time.
Applsci 14 04345 g008
Figure 9. The G–R relationship fitting diagram. The red dots indicate the cumulative frequencies of ML ≥ 2.0 events in the period of 2010~2013 with a b-value of 1.18, while the blue dots mark the same physical quantities in the period of 2018–2021 with a b-value of 0.91. The solid lines are the fit lines for different periods.
Figure 9. The G–R relationship fitting diagram. The red dots indicate the cumulative frequencies of ML ≥ 2.0 events in the period of 2010~2013 with a b-value of 1.18, while the blue dots mark the same physical quantities in the period of 2018–2021 with a b-value of 0.91. The solid lines are the fit lines for different periods.
Applsci 14 04345 g009
Figure 10. The rupture regions of M ≥ 6.5 historic events along the Xianshui He–Anning He–Zemu He fault belt (modified from [25]), (a) AD 1480–1700, (b) AD 1701–1800, (c) 1801–1900, (d) 1901–2022. The gray areas mark the rupture zone of historic events, the two red stars indicate the epicenters of MS 6.3 Kangding earthquake in 2014 and MS 6.8 Luding earthquake in 2022.
Figure 10. The rupture regions of M ≥ 6.5 historic events along the Xianshui He–Anning He–Zemu He fault belt (modified from [25]), (a) AD 1480–1700, (b) AD 1701–1800, (c) 1801–1900, (d) 1901–2022. The gray areas mark the rupture zone of historic events, the two red stars indicate the epicenters of MS 6.3 Kangding earthquake in 2014 and MS 6.8 Luding earthquake in 2022.
Applsci 14 04345 g010
Figure 11. Fault−locking degree derived from GNSS data in the Xianshui He fault belt (modified from [78]; the coupling coefficient PHI varies from 0 to 1, representing fully creeping to completely locked).
Figure 11. Fault−locking degree derived from GNSS data in the Xianshui He fault belt (modified from [78]; the coupling coefficient PHI varies from 0 to 1, representing fully creeping to completely locked).
Applsci 14 04345 g011
Figure 12. ΔCFS maps based on f1—Xianshuihe fault belt (a) and f2—Anninghe fault belt (b) as the receiving faults. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Figure 12. ΔCFS maps based on f1—Xianshuihe fault belt (a) and f2—Anninghe fault belt (b) as the receiving faults. F1: Xianshui He fault belt, F2: Anning He fault belt, F3: Zemu He fault belt, F4: Daliang Shan fault belt.
Applsci 14 04345 g012
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peng, L.; Long, F.; Zhao, M.; Ran, X.; Wang, D.; Wang, R.; Wu, W.; He, C. The Stress State before the MS 6.8 Luding Earthquake on 5 September 2022 in Sichuan, China: A Retrospective View Based on the b-Value. Appl. Sci. 2024, 14, 4345. https://doi.org/10.3390/app14114345

AMA Style

Peng L, Long F, Zhao M, Ran X, Wang D, Wang R, Wu W, He C. The Stress State before the MS 6.8 Luding Earthquake on 5 September 2022 in Sichuan, China: A Retrospective View Based on the b-Value. Applied Sciences. 2024; 14(11):4345. https://doi.org/10.3390/app14114345

Chicago/Turabian Style

Peng, Liyuan, Feng Long, Min Zhao, Xiyang Ran, Di Wang, Rui Wang, Weiwei Wu, and Chang He. 2024. "The Stress State before the MS 6.8 Luding Earthquake on 5 September 2022 in Sichuan, China: A Retrospective View Based on the b-Value" Applied Sciences 14, no. 11: 4345. https://doi.org/10.3390/app14114345

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop