
Citation: | Cai Y, Wu JP, Liu YN and Gao SJ (2024). Advances in seismological methods for characterizing fault zone structure. Earthq Sci 37(2): 122–138,. DOI: 10.1016/j.eqs.2024.01.019 |
Large earthquakes frequently occur along complex fault systems. Understanding seismic rupture and long-term fault evolution requires constraining the geometric and material properties of fault zone structures. We provide a comprehensive overview of recent advancements in seismological methods used to study fault zone structures, including seismic tomography, fault zone seismic wave analysis, and seismicity analysis. Observational conditions limit our current ability to fully characterize fault zones, for example, insufficient imaging resolution to discern small-scale anomalies, incomplete capture of crucial fault zone seismic waves, and limited precision in event location accuracy. Dense seismic arrays can overcome these limitations and enable more detailed investigations of fault zone structures. Moreover, we present new insights into the structure of the Anninghe-Xiaojiang fault zone in the southeastern margin of the Qinghai-Xizang Plateau based on data collected from a dense seismic array. We found that utilizing a dense seismic array can identify small-scale features within fault zones, aiding in the interpretation of fault zone geometry and material properties.
Movement along faults in the Earth’s crust leads to extensive fragmentation of rocks on each side of the fault, giving rise to a distinct low-velocity region known as a fault zone. The fault zone comprises two primary components: the fault core and the surrounding damage zones (Ben-Zion and Sammis, 2003). The fault core, typically a narrow zone ranging from tens of centimeters to a few meters in width, consists predominantly of cataclasite and ultracataclasite. The majority of displacement along the fault occurs within the fault core. The fault damage zones typically include localized fracture zones or subsidiary faults that can extend tens to hundreds of meters away from the fault core (Chester et al., 1993; Evans and Chester, 1995). Large earthquakes often occur within structurally complex fault zones, where the geometry, continuity, and seismic velocity structure of the fault zone are important for determining seismic rupture and slip localization. Therefore, accurate and comprehensive information about the fault zone is paramount for understanding seismic processes and long-term fault evolution (Share et al., 2023; Yang HF and Zhu LP, 2010; Yang HF et al., 2011; Yu HY et al., 2018).
Detailed characterization of fault zone structure offers crucial insights into past fault behavior and potential future earthquake characteristics (Allam et al., 2014). Irregular fault geometries, such as bends, bumps, or jumps, can act as either initiation or termination points of earthquake ruptures (King and Nábělek, 1985; Elliott et al., 2015; Yao SL and Yang HF, 2022). Relatively high and low seismic velocities may provide valuable information about areas of high moment release and rupture termination, respectively (Michael and Eberhart-Phillips, 1991). Moreover, velocity contrasts along and across the fault strike exert a significant influence on the dynamic migration and stagnation of earthquake ruptures (Ben-Zion and Andrews, 1998; Bennington et al., 2013; Abdelmeguid and Elbanna, 2022; Weng HH et al., 2016; Zhao P et al., 2010). Analysis of vP/vS ratios in the vicinity of faults and their surrounding areas offers essential insight into rock damage, fluid conditions, and mechanistic processes within different fault segments (Allam et al., 2014; Christensen, 1996; Jiang XH et al., 2021; Lin GQ and Shearer, 2009; Share et al., 2023). The presence of low-velocity zones close to faults amplifies ground vibration amplitudes and leads to a substantial increase in ground motion near fault zones (Ben-Zion and Aki, 1990; Ben-Zion and Andrews, 1998; Chen X and Yang HF, 2020; Kurzon et al., 2014; Luo S et al., 2023; Wirth et al., 2019; Wu CQ et al., 2009). Additionally, segmented low-velocity zones can result in asymmetric rupture propagation, which amplifies ground motion and triggers seismic activity in the direction of rupture. This information is important for seismic risk assessment (Somerville et al., 1997; Gomberg et al., 2001; Weng HH et al., 2016).
An accurate assessment of the location, magnitude, extent of rupture, ground motion, and hazard distribution of future major earthquakes within a fault zone is key for urban planning and protecting human life and property. Urban planners can use this information to make informed decisions regarding infrastructure development and land-use zoning while considering potential seismic hazards. Dynamic simulations based on detailed three-dimensional (3D) structural models offer reliable tools for studying seismic rupture and fault slip, as well as predicting strong ground motions such as peak ground velocity (PGV) and peak ground acceleration (PGA). These simulations project seismic intensity distribution, providing a robust reference for seismic hazard assessment within the fault zone (Cheng J et al., 2021; Luo S et al., 2023; Yao SL and Yang HF, 2022).
A variety of seismological techniques can be employed to investigate fault zone structures, including seismic imaging methods, such as travel-time tomography and ambient noise tomography, which enable visualization and characterization of subsurface structures (Luo S et al., 2023; Mordret et al., 2019; Ozer et al., 2019; Shao XH et al., 2022; Taylor et al., 2019; Wang YD et al., 2019; Yang HF et al., 2020; Yang Y et al., 2020). Additionally, fault zone trapped wave analysis can provide valuable information about the properties and behavior of waves confined within the fault zone (Ben-Zion et al., 2003; Li YG and Leary, 1990; Li YG et al., 2000, 2016; Shao XH et al., 2022; Zhang Z et al., 2022). Other techniques, such as head wave (Bennington et al., 2013; Najdahmadi et al., 2016; Share et al., 2017, 2023; Share and Ben-Zion, 2018; Zhao P and Peng ZG, 2008) and reflected wave analysis (Li HY et al., 2007; Yang HF et al., 2011, 2014) can offer insights into wave propagation and interaction with the fault zone. Furthermore, the spatial distribution of earthquakes can serve as a powerful tool for studying fault zone structures (Cheng YF et al., 2023; Fang LH et al., 2015; Feng T et al., 2022; Kim et al., 2016; Li YQ et al., 2019; Shelly and Hardebeck, 2019; White et al., 2019). The distribution of seismic events can provide valuable information about the geometry and connectivity of fault zones, and contribute to understanding the fault zone structure and the implications for seismic processes. Overall, combining these seismological techniques enables investigation and characterization of the structures within fault zones, which provides essential insights into fault zone properties and behavior, as well as the influence of fault zone structure on the seismic rupture process.
In recent years, dense-array seismic detection technology has emerged as an increasingly important tool for investigating deep fault zone structures. Wu JP et al. monitored fault zone behavior in China as part of the National Key Research & Development Program for Major Natural Disasters from 2019 to 2021. They established key monitoring infrastructure through a large-scale dense array within the Anninghe-Xiaojiang fault zone (AXFZ) in China. The project successfully obtained high-resolution 3D velocity structures and precise seismic spatial distribution images of the AXFZ, providing key information on the spatial geometry of the fault zone. The results enhanced our understanding of the seismogenic environment and deep tectonics and facilitated numerical simulations of seismogenic processes. In this study, we focused on reviewing advancements in seismological methods used to study fault zone structures and highlight notable results obtained from the dense seismic array deployed in the AXFZ.
High-resolution seismic imaging techniques can inform on the velocity structure of the Earth’s interior and are therefore used to study fault zone structures. Techniques include local earthquake travel-time tomography, ambient noise tomography, and tectonic interface imaging.
Travel-time tomography, an imaging technique that utilizes P and S travel times, was first proposed in the 1970s and has since become a common method for studying the Earth’s internal tectonics and geodynamics (Ajala et al., 2019; Aki and Lee, 1976; Huang JL and Zhao DP, 2004; Ozer et al., 2019; Thurber, 1983; Wu JP et al., 2013; Zhao DP et al., 1992). Ajala et al. (2019) utilized travel-time tomography to construct the 3D velocity structure of the Coachella Valley at 0–10 km depth. By analyzing lateral velocity variations, they identified a previously unrecorded fault. To enhance the capabilities of travel-time tomography, Zhang HJ and Thurber (2003) developed a joint inversion method called double-difference tomography (TomoDD) by combining double-difference earthquake location algorithm of Waldhauser and Ellsworth (2000) and travel-time tomography. TomoDD utilizes relative travel time differences between neighboring earthquakes and absolute travel times between earthquakes and seismic stations for source location and velocity structure determination. This approach has been used extensively to obtain high-resolution crustal velocity structures at local and regional scales (Deng WZ et al., 2014; Feng T et al., 2022; Hutchinson et al., 2019; Shelly et al., 2006). Feng T et al. (2022) employed machine learning seismic detection techniques to obtain a high-precision seismic catalog and seismic phase data. Subsequently, they utilized TomoDD to obtain a high-resolution 3D velocity structure at 0–20 km depth beneath parts of the AXFZ and characterized the fault’s dip angle. It is important to note that the application of travel-time tomography relies on high spatial and temporal resolution in the earthquake catalog, so it is less suitable for regions with low seismic activity.
Ambient noise tomography is commonly used to image the S-wave velocity structure (Jin JQ et al., 2023; Li DD et al., 2019; Luo S et al., 2023; Mordret et al., 2019; Ni HY et al., 2022; Shapiro et al., 2005; Taylor et al., 2019; Wang YD et al., 2019; Yang HF et al., 2020) and offers several advantages over travel-time tomography. Unlike travel-time tomography, ambient noise tomography is not constrained by the earthquake distribution, making it applicable to regions with low seismicity. This method can effectively study the shallow crustal structure of fault zones by utilizing dense array data and extracting surface wave signals from ambient noise. Mordret et al. (2019) used a dense seismic array to image the 3D velocity structure of the San Jacinto fault zone to a depth of approximately 500 m depth (Figure 1). Their model revealed significant lateral velocity variations across fault strike, providing crucial information about the fault zone’s internal dynamics. Yang HF et al. (2020) used teleseismic P-wave arrivals to determine a 3.4 km wide low-velocity anomaly within the Chenghai fault zone and subsequently expanded the detection of the low-velocity anomaly to a depth of 1.5 km through noise tomography. This demonstrates the capability of ambient noise tomography in characterizing subsurface structures and identifying anomalies within fault zones.
Traditional ambient noise tomography relies on base-order surface wave dispersion data. However, higher-order surface waves exhibit increased sensitivity to S-wave velocity variations, particularly in the presence of soft layers and for deeper subsurface investigations. Joint inversion of both base-order and high-order surface waves offers advantages in reducing the non-uniqueness of the inversion process (Luo YH et al., 2008).
Ambient noise interferometry is used to reconstruct the Green’s function between pairs of stations by cross-correlating the received signals. This technique can enable the use of ambient noise as a valuable signal, with one station acting as a source and another as a receiver. This approach enables accurate estimation of the Green’s function between stations, encompassing both body and surface waves. Body waves have higher frequencies compared to surface waves, and extracting body wave signals in ambient noise interferometry improves the resolution for imaging subsurface structures (Nakata et al., 2015). She YY et al. (2022) employed ambient noise interferometry to extract reflected wave information, thus identifying shallower fault structures in the Chenghai fault zone. Ambient noise interferometry also provided insight into fault structures that complemented the low-velocity anomalies obtained through noise tomography (Yang HF et al., 2020). Results obtained through noise interferometry were found to be consistent with those obtained through receiver function imaging (Jiang XH et al., 2021), validating the effectiveness and reliability of the method.
Although ambient noise tomography shows good lateral resolution, it has limited sensitivity to sharp vertical velocity boundaries. Therefore, ambient noise tomography is not suitable for accurately determining the depth of the sensitive cores within the low-velocity zone (LVZ). Receiver function imaging, however, assumes an equivalent source and a horizontal laminar earth model. Therefore, the method effectively removes the contributions of the seismic source, propagation path, and instrument response by deconvolving the vertical component from the radial component. The resulting receiver function provides valuable phases, including the Moho converted Ps wave and its multiple reflected waves. These waves are particularly sensitive to velocity discontinuities and the vP/vS ratio. The H-κ stacking method is widely used for estimating crustal thickness and the vP/vS ratio (Zhu LP and Kanamori, 2000; Wang P et al., 2010; Wang WL et al., 2017; Fadel et al., 2018).
The H-κ method typically assumes that the vP/vS ratio remains constant throughout the fault zone. However, this assumption can have implications for accurately determining the geometry and velocity variations within the LVZ. Therefore, Jiang XH et al. (2021) proposed a novel approach using primary and secondary Ps-converted waves (Pbs PbpPs) generated by the low-velocity structures within the fault zone (Figure 2). This approach allows for the estimation of depth and vP/vS ratio variations within the LVZ over time. Jiang XH et al. (2021) showed that their LVZ model could be successfully recovered even in the presence of noise. By incorporating primary and secondary Ps-converted waves and considering their interactions with the low-velocity structures, this method offers a more comprehensive understanding of the LVZ properties, including depth and vP/vS ratio variations.
Receiver function migration imaging can be used to investigate crustal and upper mantle velocity discontinuities. This method can determine the depth and morphology of these discontinuities by converting phase information contained in the receiver function from the time domain (or frequency domain) to the depth domain, thereby enhancing the spatial resolution of receiver function imaging (Chen JH et al., 2005). A common method for imaging these seismic discontinuities is Common Conversion Point (CCP) stacking (Chen L et al., 2006; Hong DQ et al., 2021; Li W et al., 2022; Liu Z et al., 2017; Tian XB et al., 2021; Zhu LP, 2000). CCP stacking sums receiver functions from different seismic stations that share a common conversion point corresponding to a specific depth within the Earth. By stacking the receiver functions at each common conversion point, the resulting image provides valuable insights into the structure and characteristics of crustal and upper mantle discontinuities. Liu Z et al. (2017) used CCP stacking to image the crustal structure across several faults in the eastern margin of the Qinghai-Xizang Plateau (Figure 3). The resulting images clearly showed the velocity discontinuities around the fault system, shedding light on the fault geometry and dynamic fault processes.
The spectral ratio method has emerged as a promising approach for determining the depth of the LVZ (Chen ZQ et al, 2023; Molnar et al, 2018; She YY et al., 2022; Song JH and Yang HF, 2022). This method extracts information about the impedance frequency response by employing the horizontal-to-vertical spectral ratio (HVSR), which compares the spectra of horizontal and vertical components of a waveform. This enables estimation of the shallow S-wave velocity structure and facilitates frequency-depth conversion to image the continuous seismic wave impedance depth (Arai and Tokimatsu, 2004). She YY et al. (2022) applied the HVSR method to investigate the geometry of the LVZ in the Chenghai fault zone. Their findings revealed an inverted trapezoidal LVZ with a depth of approximately 1.0 km and a width of approximately 3.0 km.
Song JH and Yang HF (2022) employed the standard spectral ratio (SSR) method to investigate the seismic site response characteristics within the Chenghai fault zone by analyzing amplitudes and spectra of local and teleseismic earthquakes recorded by a dense seismic array. Their analysis revealed the presence of a new, heightened site response within the pre-existing LVZ of the fault zone. This site response exhibited a width of 500 m and indicated a higher level of ground motion amplification (Figure 4). Song JH and Yang HF (2022) attributed this localized intensification to a concentrated low-velocity zone within the fault zone.
In general, seismic imaging techniques are limited in their ability to directly detect the structure of the fault core within a fault zone. However, analyzing fault zone trapped waves can provide valuable insights into the dimensions and characteristics of the fracture core (Zhang Z et al., 2022). When seismic waves encounter the interface between the low-velocity medium within the fault zone and the higher-velocity surrounding rocks, reflection and transmission occurs. If the angle of incidence exceeds the critical angle, the wave undergoes total reflection within the fault zone. This leads to the formation of a wave train with the properties of a Love surface wave, known as the fault zone trapped wave (Ben-Zion, 1998). This wave train arrives after the primary body waves, exhibiting low frequency, large amplitude, and dispersion characteristics (Ben-Zion, 1998). By analyzing the travel time, frequency, amplitude, and dispersion characteristics of the fault zone trapped wave, it is possible to invert and simulate waveforms to obtain the width and geometry of the fracture zone, Q values within the fracture zone, as well as velocity variations within and outside the fracture zone. This approach has been widely applied in fault studies worldwide and provides valuable information about the internal structure of fault zones (Ben-Zion et al., 2003; Li YG and Leary, 1990; Li SL et al., 2009; Li YG et al., 2000, 2016; Peng ZG et al., 2003; Qiu HR et al., 2017; Share et al., 2017; Yang HF et al., 2011; Zhang Z et al., 2022). For example, Zhang Z et al. (2022) found a deep (~4 km) and narrow (50–150 m) inner damage zone around the Haiyuan fault by using the waveform modeling of trapped waves (Figure 5).
Indeed, there is a trade-off between the depth resolution and other parameters when analyzing fault zone trapped waves. Li YG et al. (2000) investigated the Landers fault zone and determined that at the surface, the fault zone exhibited a width of 250 m with a 45% reduction in wave velocity compared to the surrounding rocks. However, at a depth of 10 km, the fault zone narrowed to 125 m width with a 35% reduction in wave velocity. Conversely, Peng ZG et al. (2003) found that the Landers fault zone had a width of about 200 m with wave velocity reduction of 30%–40% relative to the surrounding rocks at depth of approximately 2–4 km. The disparities between these results highlight the trade-off between depth-dependent fault zone width and wave velocity variation. Additional constraints in fault zone studies can be incorporated through analysis of higher-frequency body waves such as direct waves, head waves, and reflected waves (Bennington et al., 2013; Li HY et al., 2007; Qiu HR et al., 2017; Share et al., 2017; Zhang M et al., 2022). To effectively observe fault zone trapped waves, a dense array of seismic stations deployed perpendicular to the fault strike is typically required. It is important to consider the receiving distance, as excessively long distances may lead to wave attenuation, making it challenging to accurately identify and analyze the trapped waves.
In addition to Love-type trapped waves, another type of trapped wave involving normal modes or Rayleigh-type signals has been observed in the radial and vertical components of seismic recordings (Gulley et al., 2017; Malin et al., 2006). These waves exhibit amplitudes between those of the direct P and S waves. Such waves have been detected at various stations along the San Jacinto fault zone (Qin L et al., 2018; Qiu HR et al., 2017) and the Parkfield segment of the San Andreas fault (Ellsworth and Malin, 2011). After a large earthquake, the energy released during the rupture excites free vibrations within the Earth. These vibrations occur at specific eigenfrequencies and can be described by a set of eigenfunctions (Gilbert, 1971). The observed eigenfrequencies and eigenfunctions are highly sensitive to the internal structure of the Earth and have been extensively used for global-scale mapping of Earth’s deep structure. Similar to oscillatory waves found in any finite object (Geimer et al., 2020), seismic energy trapped in a fault zone waveguide can also generate normal (or resonant) modes in a finite (width and depth) waveguide. These resonant modes provide valuable constraints on the internal structure of the fault zone waveguide. Qiu HR et al. (2020) made notable advancements in this field by analyzing long-duration resonant wave signals following the arrival of trapped waves. They employed a dense array of seismic stations across the San Jacinto fault zone in southern California with a spacing of 10–30 m. The results showed a velocity contrast along the fault, which had not been achieved by previous trapped wave studies due to resolution. The resonant waves exhibit a lower frequency component (<3 Hz) than trapped waves, which typically peak around 5 Hz. This lower frequency component offers distinct spatial sensitivities for structural imaging of the fault zone.
The presence of different rock properties at fault interfaces leads to the generation of fault zone head waves when body waves propagate within a fracture zone (Ben-Zion and Aki, 1990). These head waves travel with the velocity of the faster block and are refracted from the fault to the slower velocity block (Bennington et al., 2013). Therefore, the head waves arrive earlier than direct body waves and exhibit opposite polarity. Stations located near slower interfaces record these head waves as the first arrivals (Ben-Zion and Malin, 1991). Since head waves propagate through the fault zone interface, they provide valuable information about the contrasting properties on each side of the fault (Ben-Zion and Malin, 1991; Bennington et al., 2013; Najdahmadi et al., 2016; Qiu HR et al., 2017; Share et al., 2017, 2023; Zhao P and Peng ZG, 2008). By analyzing fault zone head waves, Share and Ben-Zion (2018) found that rocks in the northeastern San Jacinto fault zone exhibited velocities more than 7.5% lower than those in the southwest. This implies that significant seismic events occurring in the northern fault zone could propagate towards the southeast.
The fault zone is commonly characterized by the presence of low-velocity anomalies, and the analysis of reflected waves offers an effective approach to determining the boundaries of the fault zone (Li HY et al., 2007; Yang HF et al., 2011, 2014; Yang HF, 2015). By tracing the paths of reflected waves that propagate through the fault zone and reflect multiple times at the fault zone boundaries, information is provided about the depth and width of the LVZ. Yang HF et al. (2011) used direct and reflected wave arrivals to investigate the structure of the Calico fault. Their study revealed a 1.3 km wide LVZ with a 40% reduction in vP and a 50% reduction in vS. The LVZ was observed to dip at an angle of 70° to the northeast, reaching a depth of 3 km. Notably, this depth was shallower than the >5 km depth determined by trapped waves. By using reflected waves, the trade-off between fault zone width and velocity can largely be avoided, allowing for enhanced depth resolution of the fault zone structure. However, the accuracy of this method is highly dependent on earthquake location. Even a small (<0.5 km) focal depth error can lead to several kilometers of error in estimating the reflection point location (Li HY et al., 2007). To improve the accuracy of seismic location and, consequently, the precision of structural studies in the fault zone, the use of dense seismic array observations can be employed.
Spatial earthquake distribution plays a crucial role in constraining the geometry of fault zones and provides insights into the rupture processes occurring at the focal depth (Fang LH et al., 2015; Feng T et al., 2022; Kim et al., 2016; Li YQ et al., 2019; Ross et al., 2017, 2019; Rubin et al., 1999; Shelly and Hardebeck, 2019; White et al., 2019). This information is essential for assessing seismic hazards and understanding earthquake behavior. For example, Kim et al. (2016) used seismic activity to investigate fault dip and geometry along the Parkfield fault in California. The vertical distribution of seismic events revealed a change in fault dip from NE to SW, suggesting a twisted fault geometry (Figure 6). However, the accuracy of hypocentral parameters depends on the precision of the velocity model employed. Therefore, it is challenging to resolve earthquake locations in regions with complex velocity structures, particularly if only a one-dimensional velocity model is used.
The application of non-traditional seismic signals has opened new avenues for investigating active fault behavior and subsurface structures (Figure 7). Brenguier et al. (2019) used train traffic as a noise source for monitoring active faults via seismic interferometry. Sheng Y (2023) introduced seismic stereometry as an alternative method, illustrating the potential of seismic signals generated by cars. Similarly, Zhang H et al. (2023) explored lateral variations along the southern San Andreas fault zone by analyzing traffic signals through a dense seismic array. These studies collectively offer new perspectives for subsurface characterization and fault monitoring, showcasing the adaptability of seismic methods to unconventional noise sources. The findings emphasize the transformative potential of traffic-generated seismic signals while highlighting the need for further fault zone and subsurface study.
Local vP/vS ratios play an important role in advancing understanding of fault zones and their underlying dynamics. Lin GQ and Shearer (2007) determined relative vP/vS ratios of different events within the same earthquake cluster, shedding light on consistent ratios among events within the cluster. Lin GQ et al. (2015) then extended this approach to near-source vP/vS ratios and constrained the size and depth of the shallow magma reservoir beneath Kīlauea caldera. This application of seismic parameters showcases the potential for providing insights into subsurface volcanic processes, and the same method could also be applied to better understand fault zone properties.
Dense seismic arrays are important for studying fault zones across all seismic methods, because a dense array enhances the ability to image high-resolution fault zone structures and capture seismic wave signals. For example, dense seismic arrays have been used to reveal structural intricacies of various segments within the San Andreas fault zone (Brenguier et al., 2019; Lewis et al., 2007; Li YG et al., 2004; Li YG, 2021; Zhang H et al., 2023) and the geometry of fracture zones within the San Jacinto fault zone (Mordret et al., 2019; Qiu HR et al., 2017, 2020; Share et al., 2017, 2023; Wang YD et al., 2019; Yang HF et al., 2014).
The AXFZ (Figure 8), situated on the southeastern margin of the Qinghai-Xizang Plateau, plays a pivotal role in eastward extrusion of the Qinghai-Xizang crust. This fault zone has experienced several notable earthquakes, including 11 earthquakes of M6–7 and 7 earthquakes of M7–8 (Wen XZ et al., 2008, 2011), so the AXFZ has high seismic hazard potential in the medium and long term. Increased observation of the AXFZ is required to further elucidate the processes involved in the formation and occurrence of large earthquakes and improve hazard assessment.
Luo S et al. (2023) used ambient noise tomography to obtain a comprehensive 3D model of the upper crustal vS model beneath the Anninghe fault zone. They revealed the presence of a narrow LVZ along the fault at a depth of 0–3 km. Interestingly, this LVZ shifts to greater depths of 4.5–8 km in the eastern part of the fault zone. Based on these observations, they proposed a seismotectonic model that interprets the narrow LVZ as a fracture zone constrained by the presence of water (Figure 9). According to their model, the western part of the fracture zone lays beneath the eastern part, resulting in the formation of a seismogenic zone characterized by increased microseismic activity at greater depths within the eastern section of the Anninghe fault zone. Moreover, the study conducted large earthquake simulations using the 3D velocity structure obtained with the dense seismic array. The results of these simulations indicated that incorporating the detailed velocity structure into the models resulted in an increase in simulated PGA during rupture by 25%–50% compared to using a smoother velocity model. This underscores the significance of considering detailed velocity models when assessing the hazard associated with large earthquakes.
We employed TomoDD to investigate the P-wave velocity structure around the Anninghe fault zone. Figure 10 shows a profile across the fault. The results revealed velocity variation across the fault. Notably, earthquakes predominantly occur within the high-velocity zone. The findings suggest that the fault may exhibit an eastward dip, which aligns with conclusions drawn by Luo S et al. (2023) from ambient noise tomography.
Shao XH et al. (2022) obtained three high-resolution P-wave velocity profiles within the Anninghe fault zone (between Mianning and Xichang) using travel-time tomography of methane gas sources (Figure 11). They particularly investigated shallow depths of 0–2 km and found this region was characterized by low velocities, primarily attributed to the presence of Quaternary sediments. Integrating data from methane gas sources and travel-time tomography offers a unique approach to investigating shallow layers surrounding a fault zone. Understanding the fault’s shallow structure provides important input data for modeling seismic events, thus contributing to improved hazard assessments in the region.
Chen ZQ et al. (2023) investigated the sedimentary structure of the Anninghe fault zone using continuous waveform data obtained from two linear dense arrays, which were positioned approximately 80 m apart across the fault zone. Chen ZQ et al. (2023) used the HVSR method to estimate the thickness of the sediment layer adjacent to the fault zone of 300–600 m. The sediment layer exhibited an eastward dip, indicating a correlation with the east-west extrusion and compression that occurred during the fault’s formation. These findings have implications for the seismic vulnerability of densely populated areas, because the resonance period of multistory buildings, particularly those with 6–9 stories, aligns with the peak frequency range of the sediment layer. Therefore, such buildings are more susceptible to earthquake damage due to resonance effects.
Feng T et al. (2022) combined the machine learning-based earthquake location workflow (LOC-FLOW; Zhang M et al., 2022) and TomoDD to investigate the north-central section of the Xiaojiang fault. The LOC-FLOW method accurately detected seismic events from raw continuous waveform data, allowing the construction of a high-accuracy seismic catalog. Subsequently, the TomoDD method was applied to derive a high-resolution 3D velocity structure of the region. The seismicity pattern indicated a west-dipping Xiaojiang main fault (Figure 12). Tomographic analysis of the velocity structure demonstrated significant lateral heterogeneity, with distinct velocity variations across the fault. Notably, most earthquakes occurred at the boundary between high-velocity and low-velocity regions or within the high-velocity region situated above the low-velocity zone. Two seismic gaps were identified along the main fault zone, representing regions with relatively low earthquake occurrence, which can be significant indicators for assessing future seismic hazards in the area.
This review paper discusses various seismological methods used to study fault zone structure, including seismic tomography, fault zone wave analysis, and seismic activity studies. These techniques provide valuable structural information for understanding fault zone evolution and seismic rupture processes. We also discussed method limitations that result in inadequate detection capabilities for many active faults with high seismic activity. For instance, imaging resolution may be insufficient for distinguishing small-scale velocity anomalies, and important fault zone seismic waves, such as trapped and resonant waves, may go undetected. Additionally, there is room for improvement in terms of seismic location accuracy. To address these limitations, it becomes necessary to deploy dense or ultra-dense arrays in important fault zones to enhance seismic imaging resolution. By combining dense arrays with analyses of fault zone seismic waves and high-precision earthquake distribution, fault zone structures can be better constrained.
The development of complementary joint inversion can lead to a qualitative improvement in the study of fault zone structure. Moreover, different fault zone seismic waves impose different constraints on the fault zone structure. In the future, the utilization of full waveform inversion techniques, incorporating multiple seismic waves, may provide a more accurate characterization of fault zone structures. Overall, it is an ongoing challenge for continuously advancing and integrating various seismological techniques to improve our understanding of fault zone structures and their associated seismic behavior.
This study was supported by the National Key R&D Program of China (No. 2022YFF0800601), the National Natural Science Foundation of China (No. U2039204), and the Special Fund of the Institute of Geophysics, China Earthquake Administration (No. DQJB23B22).
Prof. Jianping Wu serves as an editorial board member for Earthquake Science and was not involved in the editorial review or the decision-making process for this article. All authors declare that they have no competing interests.
Abdelmeguid M and Elbanna A (2022). Sequences of seismic and aseismic slip on bimaterial faults show dominant rupture asymmetry and potential for elevated seismic hazard. Earth Planet Sci Lett 593: 117648 . https://doi.org/10.1016/j.jpgl.2022.117648.
|
Ajala R, Persaud P, Stock JM, Fuis GS, Hole JA, Goldman M and Scheirer D (2019). Three-dimensional basin and fault structure from a detailed seismic velocity model of Coachella Valley, Southern California. J Geophys Res:Solid Earth 124(5): 4728 – 4750 . https://doi.org/10.1029/2018JB016260.
|
Aki K and Lee WHK (1976). Determination of three-dimensional velocity anomalies under a seismic array using first P arrival times from local earthquakes: 1. A homogeneous initial model. J Geophys Res 81(23): 4381 – 4399 . https://doi.org/10.1029/JB081i023p04381.
|
Allam AA, Ben-Zion Y, Kurzon I and Vernon F (2014). Seismic velocity structure in the Hot Springs and Trifurcation areas of the San Jacinto fault zone, California, from double-difference tomography. Geophys J Int 198(2): 978 – 999 . https://doi.org/10.1093/gji/ggu176.
|
Arai H and Tokimatsu K (2004). S-wave velocity profiling by inversion of microtremor H/V spectrum. Bull Seismol Soc Am 94(1): 53 – 63 . https://doi.org/10.1785/0120030028.
|
Bennington NL, Thurber C, Peng ZG, Zhang HJ and Zhao P (2013). Incorporating fault zone head wave and direct wave secondary arrival times into seismic tomography: Application at Parkfield, California. J Geophys Res:Solid Earth 118(3): 1008 – 1014 . https://doi.org/10.1002/jgrb.50072.
|
Ben-Zion Y and Aki K (1990). Seismic radiation from an SH line source in a laterally heterogeneous planar fault zone. Bull Seismol Soc Am 80(4): 971 – 994 . https://doi.org/10.1785/BSSA0800040971.
|
Ben-Zion Y and Malin P (1991). San Andreas fault zone head waves near Parkfield, California. Science 251(5001): 1592 – 1594 . https://doi.org/10.1126/science.251.5001.1592.
|
Ben-Zion Y (1998). Properties of seismic fault zone waves and their utility for imaging low-velocity structures. J Geophys Res: Solid Earth 103(B6): 12567 – 12585 https://doi.org/10.1029/98JB00768.
|
Ben-Zion Y and Sammis CG (2003). Characterization of fault zones. Pure Appl Geophys 160(3): 677 – 715 . https://doi.org/10.1007/PL00012554.
|
Ben-Zion Y, Peng ZG, Okaya D, Seeber L, Armbruster JG, Ozer N, Michael AJ, Baris S and Aktar M (2003). A shallow fault-zone structure illuminated by trapped waves in the Karadere-Duzce branch of the North Anatolian Fault, western Turkey. Geophys J Int 152(3): 699 – 717 . https://doi.org/10.1046/j.1365-246X.2003.01870.x.
|
Brenguier F, Boué P, Ben-Zion Y, Vernon F, Johnson CW, Mordret A, Coutant O, Share PE, Beaucé E, Hollis D and Lecocq T (2019). Train traffic as a powerful noise source for monitoring active faults with seismic interferometry. Geophys Res Lett 46(16): 9529 – 9536 . https://doi.org/10.1029/2019GL083438.
|
Chen JH, Liu QY, Li SC, Guo B and Lai YG (2005). Crust and upper mantle S-wave velocity structure across northeastern Tibetan Plateau and Ordos block. Chin J Geophys 48(2): 369 – 379 . https://doi.org/10.1002/cjg2.663.
|
Chen L, Zheng TY and Xu WW (2006). A thinned lithospheric image of the Tanlu Fault Zone, eastern China: constructed from wave equation based receiver function migration. J Geophys Res:Solid Earth 111(B9): B09312 . https://doi.org/10.1029/2005JB003974.
|
Chen X and Yang HF (2020). Effects of seismogenic width and low-velocity zones on estimating slip-weakening distance from near-fault ground deformation. Geophys J Int 223(3): 1497 – 1510 . https://doi.org/10.1093/gji/ggaa385.
|
Chen ZQ, Yao HJ, Shao XH, Luo S and Yang HF (2023). Detailed sedimentary structure of the Mianning segment of the Anninghe fault zone revealed by H/V spectral ratio. Earthq Res Adv 3(3): 100232 . https://doi.org/10.1016/j.eqrea.2023.100232.
|
Cheng J, Xu XW, Yao Q, Yang XD and Chen H (2021). Seismic hazard of multi-segment rupturing for the Anninghe-Zemuhe- Daliangshan fault region, southeastern Tibetan Plateau: constraints from geological and geodetic slip rates. Nat Hazards 107(2): 1501 – 1525 . https://doi.org/10.1007/s11069-021-04643-7.
|
Cheng YF, Allen RM and Taira TA (2023). A new focal mechanism calculation algorithm (REFOC) using inter-event relative radiation patterns: Application to the earthquakes in the Parkfield Area. J Geophys Res:Solid Earth 128(3): e2022JB025006 . https://doi.org/10.1029/2022JB025006.
|
Chester FM, Evans JP and Biegel RL (1993). Internal structure and weakening mechanisms of the San Andreas fault. J Geophys Res:Solid Earth 98(B1): 771 – 786 . https://doi.org/10.1029/92JB01866.
|
Christensen NI (1996). Poisson’s ratio and crustal seismology. J Geophys Res:Solid Earth 101(B2): 3139 – 3156 . https://doi.org/10.1029/95JB03446.
|
Deng WZ, Chen JH, Guo B, Liu QY, Li SC, Li Y, Yin XZ and Qi SH (2014). Fine velocity structure of the Longmenshan fault zone by double-difference tomography. Chin J Geophys 57(4): 1101 – 1110 . https://doi.org/10.6038/cjg2014 0408 (in Chinese with English abstract). doi: 10.6038/cjg20140408
|
Elliott AJ, Oskin ME, Liu-Zeng J and Shao YX (2015). Rupture termination at restraining bends: the last great earthquake on the Altyn Tagh Fault. Geophys Res Lett 42(7): 2164 – 2170 . https://doi.org/10.1002/2015GL063107.
|
Ellsworth WL and Malin PE (2011). Deep rock damage in the San Andreas fault revealed by P- and S-type fault-zone-guided waves. In: Geology of the Earthquake Source: A Volume in Honour of Rick Sibson. Geological Society London Special Publications 359(1): 39 – 53 . https://doi.org/10.1144/SP359.3.
|
Evans JP and Chester FM (1995). Fluid-rock interaction in faults of the San Andreas system: Inferences from San Gabriel fault rock geochemistry and microstructures. J Geophys Res:Solid Earth 100(B7): 13007 – 13020 . https://doi.org/10.1029/94JB02625.
|
Fadel I, van der Meijde M and Paulssen H (2018). Crustal structure and dynamics of Botswana. J Geophys Res:Solid Earth 123(12): 10659 – 10671 . https://doi.org/10.1029/2018JB01 6190. doi: 10.1029/2018JB016190
|
Fang LH, Wu JP, Wang WL, Du WK, Su JR, Wang CZ, Yang T and Cai Y (2015). Aftershock observation and analysis of the 2013 M S7.0 Lushan earthquake. Seismol Res Lett 86(4): 1135 – 1142 . https://doi.org/10.1785/0220140186.
|
Feng T, Zhang M, Xu LS, Wu JP and Fang LH (2022). Machine learning-based earthquake catalog and tomography characterize the middle-northern section of the Xiaojiang fault zone. Seismol Res Lett 93(5): 2484 – 2497 . https://doi.org/10.1785/0220220116.
|
Geimer PR, Finnegan R and Moore JR (2020). Sparse ambient resonance measurements reveal dynamic properties of freestanding rock arches. Geophys Res Lett 47(9): e2020GL087239 . https://doi.org/10.1029/2020GL087239.
|
Gilbert F (1971). Excitation of the normal modes of the earth by earthquake sources. Geophys J R Astr Soc 22(2): 223 – 226 . https://doi.org/10.1111/j.1365-246X.1971.tb03593.x.
|
Gomberg J, Reasenberg PA, Bodin P and Harris RA (2001). Earthquake triggering by seismic waves following the Landers and Hector Mine earthquakes. Nature 411(6836): 462 – 466 . https://doi.org/10.1038/35078053.
|
Gulley AK, Kaipio JP, Eccles JD and Malin PE (2017). A numerical approach for modelling fault-zone trapped waves. Geophys J Int 210(2): 919 – 930 . https://doi.org/10.1093/gji/ggx199.
|
Hong DQ, Huang XL, Yang Y, Bao ZW, Wang XL and Wang XY (2021). Lateral variation in Moho depth around the southern Tanlu fault zone and its adjacent area. Earthq Sci 34(1): 77 – 87 . https://doi.org/10.29382/eqs-2020-0063.
|
Huang JL and Zhao DP (2004). Crustal heterogeneity and seismotectonics of the region around Beijing, China. Tectonophysics 385(1-4): 159 – 180 . https://doi.org/10.1016/j.tecto.2004.04.024.
|
Hutchinson J, Kao H, Spence G, Obana K, Wang KL and Kodaira S (2019). Seismic characteristics of the Nootka Fault Zone: Results from the seafloor earthquake Array Japan- Canada Cascadia experiment (SeaJade). Bull Seismol Soc Am 109(6): 2252 – 2276 . https://doi.org/10.1785/012019 0008. doi: 10.1785/0120190008
|
Jiang XH, Hu SQ and Yang HF (2021). Depth extent and v P/ v S ratio of the Chenghai fault zone, Yunnan, China constrained from dense-array-based teleseismic receiver functions. J Geophys Res:Solid Earth 126(8): e2021JB022190 . https://doi.org/10.1029/2021JB022190.
|
Jin JQ, Luo S, Yao HJ and Tian XF (2023). Dense array ambient noise tomography reveals the shallow crustal velocity structure and deformation features in the Weifang segment of the Tanlu fault zone. Chin J Geophys 66(2): 558 – 575 . https://doi.org/10.6038/cjg2022P0934 (in Chinese with English abstract).
|
Kim W, Hong TK, Lee J and Taira TA (2016). Seismicity and fault geometry of the San Andreas fault around Parkfield, California and their implications. Tectonophysics 677-678: 34 – 44 . https://doi.org/10.1016/j.tecto.2016.03.038.
|
King G and Nábělek J (1985). Role of fault bends in the initiation and termination of earthquake rupture. Science 228(4702): 984 – 987 . https://doi.org/10.1126/science.228.4702.984.
|
Kurzon I, Vernon FL, Ben-Zion Y and Atkinson G (2014). Ground motion prediction equations in the San Jacinto fault zone: significant effects of rupture directivity and fault zone amplification. Pure Appl Geophys 171(11): 3045 – 3081 . https://doi.org/10.1007/s00024-014-0855-2.
|
Lewis MA, Ben-Zion Y and McGuire JJ (2007). Imaging the deep structure of the San Andreas Fault south of Hollister with joint analysis of fault zone head and direct P arrivals. Geophys J Int 169(3): 1028 – 1042 . https://doi.org/10.1111/j.1365-246X.2006.03319.x.
|
Li DD, Wang GC, Lin RH, Deng K and Tian XB (2019). The shallow crustal S-velocity structure of the Longmenshan fault zone using ambient noise tomography of a seismic dense array. Earthq Sci 32(5-6): 197 – 206 . https://doi.org/10.29382/eqs-2019-0197-02.
|
Li HY, Zhu LP and Yang HF (2007). High-resolution structures of the Landers fault zone inferred from aftershock waveform data. Geophys J Int 171(3): 1295 – 1307 . https://doi.org/10.1111/j.1365-246X.2007.03608.x.
|
Li SL, Lai XL, Yao ZX and Yang Q (2009). Fault zone structures of northern and southern portions of the main central fault generated by the 2008 Wenchuan earthquake using fault zone trapped waves. Earthq Sci 22(4): 417 – 424 . https://doi.org/10.1007/s11589-009-0417-3.
|
Li W, Chen Y, Yuan XH, Xiao WJ and Windley BF (2022). Intracontinental deformation of the Tianshan Orogen in response to India-Asia collision. Nat Commun 13(1): 3738 . https://doi.org/10.1038/s41467-022-30795-6.
|
Li YG and Leary PC (1990). Fault zone trapped seismic waves. Bull Seismol Soc Am 80(5): 1245 – 1271 . https://doi.org/10.1785/BSSA0800051245.
|
Li YG, Vidale JE, Aki K and Xu F (2000). Depth-dependent structure of the Landers fault zone from trapped waves generated by aftershocks. J Geophys Res:Solid Earth 105(B3): 6237 – 6254 . https://doi.org/10.1029/1999JB900449.
|
Li YG, Vidale JE and Cochran ES (2004). Low-velocity damaged structure of the San Andreas Fault at Parkfield from fault zone trapped waves. Geophys Res Lett 31(12): L12S06 . https://doi.org/10.1029/2003GL019044.
|
Li YG, Catchings RD and Goldman MR (2016). Subsurface fault damage zone of the 2014 M W6.0 South Napa, California, earthquake viewed from fault-zone trapped waves. Bull Seismol Soc Am 106(6): 2747 – 2763 . https://doi.org/10.1785/0120160039.
|
Li YG (2021). Spatial-temporal characterization of the San Andreas Fault by fault-zone trapped waves at seismic experiment site, Parkfield, California. Earthq Sci 34(3): 261 – 285 . https://doi.org/10.29382/eqs-2021-0014.
|
Li YQ, Wang D, Xu SH, Fang LH, Cheng YF, Luo G, Yan B, Bogdan E and Mori J (2019). Thrust and conjugate strike-slip faults in the 17 June 2018 MJMA 6.1 ( M W5.5) Osaka, Japan, earthquake sequence. Seismol Res Lett 90(6): 2132 – 2141 . https://doi.org/10.1785/0220190122.
|
Lin GQ and Shearer P (2007). Estimating local v P/ v S ratios within similar earthquake clusters. Bull Seismol Soc Am 97(2): 379 – 388 . https://doi.org/10.1785/0120060115.
|
Lin GQ and Shearer PM (2009). Evidence for water-filled cracks in earthquake source regions. Geophys Res Lett 36(17): L17315 . https://doi.org/10.1029/2009GL039098.
|
Lin GQ, Amelung F, Shearer PM and Okubo PG (2015). Location and size of the shallow magma reservoir beneath Kīlauea caldera, constraints from near-source v P/ v S ratios. Geophys Res Lett 42(20): 8349 – 8357 . https://doi.org/10.1002/2015GL065802.
|
Liu Z, Tian XB, Gao R, Wang GC, Wu ZB, Zhou BB, Tan P, Nie ST, Yu GP, Zhu GH and Xu X (2017). New images of the crustal structure beneath eastern Tibet from a high-density seismic array. Earth Planet Sci Lett 480: 33 – 41 . https://doi.org/10.1016/j.jpgl.2017.09.048.
|
Luo S, Yao HJ, Wen J, Yang HF, Tian BF and Yan MX (2023). 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 128(3): e2022JB025681 . https://doi.org/10.1029/2022JB025681.
|
Luo YH, Xia JH, Liu JP and Liu QS (2008). Joint inversion of fundamental and higher mode Rayleigh waves. Chin J Geophys 51(1): 242 – 249 . https://doi.org/10.3321/j.issn:0001-5733.2008.01.030 (in Chinese with English abstract).
|
Malin P, Shalev E, Balven H and Lewis-Kenedi C (2006). Structure of the San Andreas fault at SAFOD from P-wave tomography and fault-guided wave mapping. Geophys Res Lett 33(13): L13314 . https://doi.org/10.1029/2006GL02 5973. doi: 10.1029/2006GL025973
|
Michael AJ and Eberhart-Phillips D (1991). Relations among fault behavior, subsurface geology and three-dimensional velocity models. Science 253(5020): 651 – 654 . https://doi.org/10.1126/science.253.5020.651.
|
Molnar S, Cassidy JF, Castellaro S, Cornou C, Crow H, Hunter JA, Matsushima S, Sánchez-Sesma FJ and Yong A (2018). Application of microtremor horizontal-to-vertical spectral ratio (MHVSR) analysis for site characterization: State of the art. Surv Geophys 39(4): 613 – 631 . https://doi.org/10.1007/s10712-018-9464-4.
|
Mordret A, Roux P, Boué P and Ben-Zion Y (2019). Shallow three-dimensional structure of the San Jacinto fault zone revealed from ambient noise imaging with a dense seismic array. Geophys J Int 216(2): 896 – 905 . https://doi.org/10.1093/gji/ggy464.
|
Najdahmadi B, Bohnhoff M and Ben-Zion Y (2016). Bimaterial interfaces at the Karadere segment of the North Anatolian Fault, northwestern Turkey. J Geophys Res:Solid Earth 121(2): 931 – 950 . https://doi.org/10.1002/2015JB012601.
|
Nakata N, Chang JP, Lawrence JF and Boué P (2015). Body wave extraction and tomography at Long Beach, California, with ambient-noise interferometry. J Geophys Res:Solid Earth 120(2): 1159 – 1173 . https://doi.org/10.1002/2015JB01 1870. doi: 10.1002/2015JB011870
|
Ni HY, Zheng HG, Zhao N, Deng B, Miao P, Huang XL and Li JL (2022). Application of ambient noise tomography with a dense linear array in prospecting active faults in the Mingguang city. Chin J Geophys 65(7): 2518 – 2531 . https://doi.org/10.6038/cjg2022P0947 (in Chinese with English abstract).
|
Ozer C, Ozyazicioglu M, Gok E and Polat O (2019). Imaging the crustal structure throughout the east Anatolian fault zone, turkey, by local earthquake tomography. Pure Appl Geophys 176(6): 2235 – 2261 . https://doi.org/10.1007/s00024-018-2076-6.
|
Peng ZG, Ben-Zion Y, Michael AJ and Zhu LP (2003). Quantitative analysis of seismic fault zone waves in the rupture zone of the 1992 Landers, California, earthquake: evidence for a shallow trapping structure. Geophys J Int 155(3): 1021 – 1041 . https://doi.org/10.1111/j.1365-246X.2003.02109.x.
|
Qin L, Ben-Zion Y, Qiu H, Share PE, Ross ZE and Vernon FL (2018). Internal structure of the San Jacinto fault zone in the trifurcation area southeast of Anza, California, from data of dense seismic arrays. Geophys J Int 213(1): 98 – 114 . https://doi.org/10.1093/gji/ggx540.
|
Qiu HR, Ben-Zion Y, Ross ZE, Share PE and Vernon FL (2017). Internal structure of the San Jacinto fault zone at Jackass Flat from data recorded by a dense linear array. Geophys J Int 209(3): 1369 – 1388 . https://doi.org/10.1093/gji/ggx096.
|
Qiu HR, Allam AA, Lin FC and Ben-Zion Y (2020). Analysis of fault zone resonance modes recorded by a dense seismic array across the San Jacinto fault zone at Blackburn Saddle. J Geophys Res:Solid Earth 125(10): e2020JB019756 . https://doi.org/10.1029/2020JB019756.
|
Ross ZE, Hauksson E and Ben-Zion Y (2017). Abundant off-fault seismicity and orthogonal structures in the San Jacinto fault zone. Sci Adv 3(3): e1601946 . https://doi.org/10.1126/sciadv.1601946.
|
Ross ZE, Idini B, Jia Z, Stephenson OL, Zhong MY, Wang X, Zhan ZW, Simons M, Fielding EJ, Yun SH, Hauksson E, Moore AW, Liu Z and Jung J (2019). Hierarchical interlocked orthogonal faulting in the 2019 Ridgecrest earthquake sequence. Science 366(6463): 346 – 351 . https://doi.org/10.1126/science.aaz0109.
|
Rubin AM, Gillard D and Got JL (1999). Streaks of microearthquakes along creeping faults. Nature 400(6745): 635 – 641 . https://doi.org/10.1038/23196.
|
Shao XH, Yao HJ, Liu Y, Yang HF, Tian BF and Fang LH (2022). Shallow crustal velocity structures revealed by active source tomography and fault activities of the Mianning-Xichang segment of the Anninghe fault zone, Southwest China. Earth Planet Phys 6(2): 204 – 212 . https://doi.org/10.26464/epp2022010.
|
Shapiro NM, Campillo M, Stehly L and Ritzwoller MH (2005). High-resolution surface-wave tomography from ambient seismic noise. Science 307(5715): 1615 – 1618 . https://doi.org/10.1126/science.1108339.
|
Share PE, Ben-Zion Y, Ross ZE, Qiu HR and Vernon FL (2017). Internal structure of the San Jacinto fault zone at Blackburn Saddle from seismic data of a linear array. Geophys J Int 210(2): 819 – 832 . https://doi.org/10.1093/gji/ggx191.
|
Share PE and Ben-Zion Y (2018). A bimaterial interface along the northern San Jacinto fault through Cajon Pass. Geophys Res Lett 45(21): 11622 – 11631 . https://doi.org/10.1029/2018GL079834.
|
Share PE, Vernon FL and Ben-Zion Y (2023). The variable continuous bimaterial interface in the San Jacinto fault zone revealed by dense seismic array analysis of fault zone head waves. J Geophys Res:Solid Earth 128(2): e2022JB025070 . https://doi.org/10.1029/2022JB025070.
|
She YY, Yao HJ, Yang HF, Wang JJ and Feng JK (2022). Constraining the depth extent of low-velocity zone along the Chenghai Fault by dense array ambient noise interferometry and horizontal-to-vertical spectral ratio. Tectonophysics 827: 229265 . https://doi.org/10.1016/j.tecto.2022.229265.
|
Shelly DR, Beroza GC, Ide S and Nakamula S (2006). Low-frequency earthquakes in Shikoku, Japan and their relationship to episodic tremor and slip. Nature 442(7099): 188 – 191 . https://doi.org/10.1038/nature04931.
|
Shelly DR and Hardebeck JL (2019). Illuminating faulting complexity of the 2017 Yellowstone maple creek earthquake swarm. Geophys Res Lett 46(5): 2544 – 2552 . https://doi.org/10.1029/2018GL081607.
|
Sheng Y (2023). Seismic stereometry: an alternative two-station algorithm to seismic interferometry for analysing car-generated seismic signals. Geophys J Int 235(1): 853 – 861 . https://doi.org/10.1093/gji/ggad287.
|
Somerville PG, Smith NF, Graves RW and Abrahamson NA (1997). Modification of empirical strong ground motion attenuation relations to include the amplitude and duration effects of rupture directivity. Seismol Res Lett 68(1): 199 – 222 . https://doi.org/10.1785/gssrl.68.1.199.
|
Song JH and Yang HF (2022). Seismic site response inferred from records at a dense linear array across the Chenghai fault zone, Binchuan, Yunnan. J Geophys Res:Solid Earth 127(1): e2021JB022710 . https://doi.org/10.1029/2021JB022710.
|
Taylor G, Rost S, Houseman GA and Hillers G (2019). Near-surface structure of the North Anatolian Fault zone from Rayleigh and Love wave tomography using ambient seismic noise. Solid Earth 10(2): 363 – 378 . https://doi.org/10.5194/se-10-363-2019.
|
Thurber CH (1983). Earthquake locations and three-dimensional crustal structure in the Coyote Lake area, central California. J Geophys Res:Solid Earth 88(B10): 8226 – 8236 . https://doi.org/10.1029/JB088iB10p08226.
|
Tian XB, Bai ZM, Klemperer SL, Liang XF, Liu Z, Wang X, Yang XS, Wei YH and Zhu GH (2021). Crustal-scale wedge tectonics at the narrow boundary between the Tibetan Plateau and Ordos block. Earth Planet Sci Lett 554: 116700 . https://doi.org/10.1016/j.jpgl.2020.116700.
|
Waldhauser F and Ellsworth WL (2000). A double-difference earthquake location algorithm: Method and application to the northern Hayward fault, California. Bull Seismol Soc Am 90(6): 1353 – 1368 . https://doi.org/10.1785/0120000006.
|
Wang P, Wang LS, Mi N, Liu JH, Li H, Yu DY, Xu MJ, Wang XC and Guo ZW (2010). Crustal thickness and average v P/ v S ratio variations in southwest Yunnan, China, from teleseismic receiver functions. J Geophys Res:Solid Earth 115(B11): B11308 . https://doi.org/10.1029/2009JB006651.
|
Wang WL, Wu JP, Fang LH, Lai GJ and Cai Y (2017). Sedimentary and crustal thicknesses and Poisson’s ratios for the NE Tibetan Plateau and its adjacent regions based on dense seismic arrays. Earth Planet Sci Lett 462: 76 – 85 . https://doi.org/10.1016/j.jpgl.2016.12.040.
|
Wang YD, Allam A and Lin FC (2019). Imaging the fault damage zone of the San Jacinto Fault near Anza with ambient noise tomography using a dense nodal array. Geophys Res Lett 46(22): 12938 – 12948 . https://doi.org/10.1029/2019GL08 4835. doi: 10.1029/2019GL084835
|
Wen XZ, Ma SL, Xu XW and He YN (2008). Historical pattern and behavior of earthquake ruptures along the eastern boundary of the Sichuan-Yunnan faulted-block, southwestern China. Phys Earth Planet Inter 168(1-2): 16 – 36 . https://doi.org/10.1016/j.pepi.2008.04.013.
|
Wen XZ, Du F, Long F, Fan J and Zhu H (2011). Tectonic dynamics and correlation of major earthquake sequences of the Xiaojiang and Qujiang-Shiping fault systems, Yunnan, China. Sci China Earth Sci 54(10): 1563 – 1575 . https://doi.org/10.1007/s11430-011-4231-0.
|
Weng HH, Yang HF, Zhang ZG and Chen XF (2016). Earthquake rupture extents and coseismic slips promoted by damaged fault zones. J Geophys Res:Solid Earth 121(6): 4446 – 4457 . https://doi.org/10.1002/2015JB012713.
|
White MCA, Ben-Zion Y and Vernon FL (2019). A detailed earthquake catalog for the San Jacinto fault-zone region in Southern California. J Geophys Res:Solid Earth 124(7): 6908 – 6930 . https://doi.org/10.1029/2019JB017641.
|
Wirth EA, Vidale JE, Frankel AD, Pratt TL, Marafi NA, Thompson M and Stephenson WJ (2019). Source-dependent amplification of earthquake ground motions in deep sedimentary basins. Geophys Res Lett 46(12): 6443 – 6450 . https://doi.org/10.1029/2019GL082474.
|
Wu CQ, Peng ZG and Ben-Zion Y (2009). Non-linearity and temporal changes of fault zone site response associated with strong ground motion. Geophys J Int 176(1): 265 – 278 . https://doi.org/10.1111/j.1365-246X.2008.04005.x.
|
Wu JP, Yang T, Wang WL, Ming YH and Zhang TZ (2013). Three dimensional P-wave velocity structure around Xiaojiang fault system and its tectonic implications. Chin J Geophys 56(7): 2257 – 2267 . https://doi.org/10.6038/cjg20130713 (in Chinese with English abstract).
|
Yang HF and Zhu LP (2010). Shallow low-velocity zone of the San Jacinto fault from local earthquake waveform modelling. Geophys J Int 183(1): 421 – 432 . https://doi.org/10.1111/j.1365-246X.2010.04744.x.
|
Yang HF, Zhu LP and Cochran ES (2011). Seismic structures of the Calico fault zone inferred from local earthquake travel time modelling. Geophys J Int 186(2): 760 – 770 . https://doi.org/10.1111/j.1365-246X.2011.05055.x.
|
Yang HF, Li ZF, Peng ZG, Ben-Zion Y and Vernon F (2014). Low-velocity zones along the San Jacinto Fault, Southern California, from body waves recorded in dense linear arrays. J Geophys Res:Solid Earth 119(12): 8976 – 8990 . https://doi.org/10.1002/2014JB011548.
|
Yang HF (2015). Recent advances in imaging crustal fault zones: a review. Earthq Sci 28(2): 151 – 162 . https://doi.org/10.1007/s11589-015-0114-3.
|
Yang HF, Duan YH, Song JH, Jiang XH, Tian XF, Yang W, Wang WT and Yang J (2020). Fine structure of the Chenghai fault zone, Yunnan, China, constrained from teleseismic travel time and ambient noise tomography. J Geophys Res:Solid Earth 125(7): e2020JB019565 . https://doi.org/10.1029/2020JB019565.
|
Yang Y, Hu SQ, Yao HJ, Fang LH and Wu JP (2020). Crustal shear wave velocity and radial anisotropy in the Xiaojiang fault zone system (SE Tibet) revealed by ambient noise interferometry. Tectonophysics 792: 228594 . https://doi.org/10.1016/j.tecto.2020.228594.
|
Yao SL and Yang HF (2022). Hypocentral dependent shallow slip distribution and rupture extents along a strike-slip fault. Earth Planet Sci Lett 578: 117296 . https://doi.org/10.1016/j.jpgl.2021.117296.
|
Yu HY, Liu YJ, Yang HF and Ning JY (2018). Modeling earthquake sequences along the Manila subduction zone: Effects of three-dimensional fault geometry. Tectonophysics 733: 73 – 84 . https://doi.org/10.1016/j.tecto.2018.01.025.
|
Zhang H, Meng HR and Ben-Zion Y (2023). Lateral variations across the southern San Andreas Fault zone revealed from analysis of traffic signals at a dense seismic array. Geophys Res Lett 50(14): e2023GL103759 . https://doi.org/10.1029/2023GL103759.
|
Zhang HJ and Thurber CH (2003). Double-difference tomography: The method and its application to the Hayward fault, California. Bull Seismol Soc Am 93(5): 1875 – 1889 . https://doi.org/10.1785/0120020190.
|
Zhang M, Liu M, Feng T, Wang RJ and Zhu WQ (2022). LOC-FLOW: An end-to-end machine learning-based high-precision earthquake location workflow. Seismol Res Lett 93(5): 2426 – 2438 . https://doi.org/10.1785/0220220019.
|
Zhang Z, Deng YF, Qiu HR, Peng ZG and Liu-Zeng J (2022). High-resolution imaging of fault zone structure along the creeping section of the Haiyuan fault, NE Tibet, from data recorded by dense seismic arrays. J Geophys Res:Solid Earth 127(9): e2022JB024468 . https://doi.org/10.1029/2022JB02 4468. doi: 10.1029/2022JB024468
|
Zhao DP, Hasegawa A and Horiuchi S (1992). Tomographic imaging of P and S wave velocity structure beneath northeastern Japan. J Geophys Res:Solid Earth 97(B13): 19909 – 19928 . https://doi.org/10.1029/92JB00603.
|
Zhao P and Peng ZG (2008). Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by repeating earthquakes. Geophys Res Lett 35(1): L01303 . https://doi.org/10.1029/2007GL031810.
|
Zhao P, Peng ZG, Shi ZQ, Lewis MA and Ben-Zion Y (2010). Variations of the velocity contrast and rupture properties of M 6 earthquakes along the Parkfield section of the San Andreas fault. Geophys J Int 180(2): 765 – 780 . https://doi.org/10.1111/j.1365-246X.2009.04436.x.
|
Zhu LP (2000). Crustal structure across the San Andreas Fault, southern California from teleseismic converted waves. Earth Planet Sci Lett 179(1): 183 – 190 . https://doi.org/10.1016/S0012-821X(00)00101-1.
|
Zhu LP and Kanamori H (2000). Moho depth variation in southern California from teleseismic receiver functions. J Geophys Res:Solid Earth 105(B2): 2969 – 2980 . https://doi.org/10.1029/1999JB900322.
|