Next Article in Journal
Effectiveness of Lignin Biopolymer for Near-Surface Soil Stabilization and Sustainable Protection from Wind and Rainfall Erosion
Previous Article in Journal
Mineral Chemistry, Whole-Rock Characterization, and EnMap Hyperspectral Data Analysis of Granitic Rocks of the Nubian Shield: A Case Study from Suwayqat El-Arsha District, Central Eastern Desert, Egypt
Previous Article in Special Issue
The SISMIKO Monitoring Network and Insights into the 2024 Seismic Swarms on the Ionian Side of the Calabrian Arc
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Background Seismicity Highlights Tectonic Asperities

by
Alexey Ostapchuk
1,2,* and
Ilmir Nugmanov
1,3
1
Sadovsky Institute for Dynamics of Geospheres, Russian Academy of Sciences, 119334 Moscow, Russia
2
Moscow Institute of Physics and Technology, Department of Theoretical and Experimental Physics of Geosystems, 141701 Dolgoprudny, Russia
3
Scientific Research Center, Institute of Artificial Intelligence, Innopolis University, 420500 Innopolis, Russia
*
Author to whom correspondence should be addressed.
Geosciences 2026, 16(1), 38; https://doi.org/10.3390/geosciences16010038 (registering DOI)
Submission received: 12 December 2025 / Revised: 4 January 2026 / Accepted: 9 January 2026 / Published: 11 January 2026
(This article belongs to the Special Issue Editorial Board Members' Collection Series: Natural Hazards)

Abstract

The heterogeneity of a tectonic fault interface controls both the spatial features of seismicity and the locations of the foci of shallow earthquakes. Strong earthquakes are associated with ruptures of asperities. We present the Seismogenic Patches Detection (SPAD) algorithm to analyze background seismicity to reveal tectonic asperities. In the first stage, the algorithm detects background seismicity based on the nearest-neighbor method. In the second stage, fuzzy clustering of the background mode is performed. Dense clusters of background seismicity, called seismogenic patches, can be interpreted as tectonic asperities. The SPAD algorithm does not use a priori data about the localizations of strong earthquakes. Using SPAD, we analyzed the seismicity of the Kamchatka segment of the Kuril–Kamchatka subduction zone for the period from 1 January 1990 to 23 September 2025. The Kamchatka regional seismic catalog includes 39,104 events, and the magnitude of completeness Mc equals 3.0. We have identified seven seismogenic patches with a size of 170–270 km. Seismogenic patches correlate with the tectonic asperity determined by the maps of the slip distributions for the four largest earthquakes—Mw7.5 (8 June 1993), Mw7.8 (5 December 1997), Mw8.8 (30 July 2025), and Mw7.8 (13 September 2025).

1. Introduction

Spatiotemporal clustering of seismicity is the main feature of the seismic process that allows us to construct models for earthquake forecasting [1,2,3]. A statistical approach alone is not sufficient for improving earthquake forecasting [4]. A multilevel and multidisciplinary approach that considers the physical basis of the process is needed to expand our resources in the earthquake forecasting.
The most seismically active zones are subduction zones. Early investigations hypothesized that the maximum earthquake size is controlled by the fast convergence rate and the young buoyant lithosphere [5,6]. However, Bletery et al. [7] reported that two geological conditions (a fast convergence rate and young buoyant lithosphere) are not required to produce large earthquakes; large earthquakes preferentially rupture flat (low curvature) interfaces. Relatively high curvature can likely result in the absence of sources of major earthquakes at 600 km from the Kuril–Kamchatka subduction zone (KKSZ) to the south from N 49° (Figure 1). McCaffrey [8] suggested that any subduction zone can produce an earthquake with Mw ≥ 9. So, there is no full understanding of the nucleation and rupture of major earthquakes.
The asperity model is the predominant one for explaining the nature of earthquakes [9]. Tectonic asperities emerge at a fault interface because of the curvature of its sides. Curvature is an important structural feature of tectonic faults; the fault plane is characterized by different-scale roughness [10,11,12,13]. Originally, the asperity model was essentially a model of heterogeneous stress along a fault interface, and an asperity was assumed to have a high friction coefficient [14,15]. The size, configuration and distribution of asperities primarily determine the location of earthquakes.
The rheology of tectonic faults plays one of the most important roles in their slip behavior [16,17,18]. A modern model of asperity suggests that it shows not only high-friction but also velocity-weakening (VW) behavior. At the same time, relatively unloaded zones between asperities, as a rule, have low-friction and velocity-strengthening (VS) behavior [19]. The spatial heterogeneity of rheological properties can be traced at different scales [17,20]. An earthquake rupture is predetermined by the rheological and structural features of the fault interface [21,22,23,24,25]. Ordinary earthquakes usually nucleate in zones where VW prevails, whereas in zones where vs. prevails, slow-slip events and swarms of weak seismic events with low radiative efficiency usually occur [26,27,28]. The coseismic rupture starts at the boundary of asperity with VW behavior. The velocity of rupture decreases rapidly in the VS zones and increases again in the neighboring asperity. If the distance between asperities is large enough, the rupture stops [22,29,30,31].
Tectonic asperities concentrate a slip deficit that slowly increases during interseismic periods and is ultimately released through large earthquakes. Interseismic coupling, which represents the percentage of slip deficit with respect to the long-term tectonic loading of a fault, can be calculated from GPS data [32,33,34]. Near an asperity, which is a locked segment of a fault in the interseismic period, the seismic coupling is close to 1. In aseismic creeping segments, the coupling is small [35]. Unfortunately, no threshold coupling value is suggested to correspond to the spatial extent of tectonic asperity and, consequently, earthquake rupture. The geometrical complexity of fault rupture can be captured by an inversion of strong motion data [36,37] or the joint inversion of seismic and geodetic data [38,39]. The areas of maximum displacement are the tectonic asperities in the maps of coseismic displacement. As a rule, an earthquake source embraces an area of one asperity, but cases in which a rupture, starting at an asperity, can spread to neighboring areas, are known [40]. The sources of the strongest mega-earthquakes can embrace three or even more tectonic asperities [41,42,43].
Analyzing weak seismicity allows an improved understanding of the physics of the processes that take place in zones of strong earthquake nucleation and reveals the structures of fault interfaces [44,45,46,47,48]. Seismic data should be considered based on the current understanding of fault structure and mechanical processes that take place in fault zones [18,49]. We developed and tested the SPAD algorithm for the analysis of background seismicity to determine the long-term structure of the subduction boundary. By using the example of seismicity in the Kuril–Kamchatka subduction zone (KKSZ), we showed that the SPAD algorithm allows us to reveal the zones of tectonic asperity localization well in advance, not after a strong earthquake.

2. Geological Frameworks

A vast majority of global seismicity is located inside the circum-Pacific subduction zone. Earthquakes with magnitudes exceeding 8.5 have been registered in the Chile subduction zone (1960, Mw9.5), Japan (2011, Mw9.1) and Alaska (1964, Mw9.3) [50,51,52]. There is some information about the great historical earthquakes in Kamchatka (1952, Mw9.0) and Ecuador (1906, M8.6) [50,53]. The shallow megathrusts of subduction zones host the largest earthquakes. These faults are the only faults capable of Mw ≥ 9 ruptures.
In Kamchatka, all attributes of active subduction are clearly exhibited, including a well-established ocean trench, deep seismicity in the Benioff zone, and a continuous arc characterized by a chain of Holocene volcanoes [54]. KKSZ extends for approximately 2100 km from Hokkaido Island in Japan, along the Kuril Islands and the Pacific coast of the Kamchatka Peninsula, until it intersects the Aleutian arc near the Commander Islands (Figure 1) [55]. The Pacific oceanic plate subducts beneath the Okhotsk continental plate, and along the Kamchatka segment, the subduction angle remains nearly constant—approximately 55° between N50° and N55° [56]. Farther north, the angle decreases to approximately 35° [57]. Plate convergence occurs almost perpendicularly to the strike of the Kamchatka trench, resulting in minor active deformation inside the arc [58]. Nevertheless, the velocity of the downgoing Pacific plate varies from about 77 mm/yr at N55° to 83 mm/yr at N47° [59,60]. The age of subduction for the KKSZ increases from north to south in the range of the northern latitudes of N50°–N56° from 77 to 104.5 million years [57]. The seismicity of the KKSZ has been investigated for more than a hundred years. Fedotov [61] sets a 17-year effective period of repetition of the strongest earthquakes and their harmonics. Based on instrumental and historical data, the average duration of the seismic cycle between two consequent earthquakes with magnitudes of M ≥ 7.7 at a particular location is about 140 ± 60 years or 120 ± 50 years [62]. In the range of northern latitudes from N50° to N56°, the maximum depth of earthquakes decreases from 500 km to 100 km [57]. The Kamchatka segment ruptured during seven earthquakes with M > 7.5 in the 20th–21st centuries, and the largest earthquake was the Mw9.0 earthquake in 1952 [63,64].

3. Seismic Data

In this work, we use the regional seismic catalog of the Kamchatka Branch of the Geophysical Survey of the Russian Academy of Sciences, which contains information about 39,104 seismic events for the period of instrumental seismic observations from 1990 to 2025 (available online https://sdis.emsd.ru/info/earthquakes/catalogue.php (accessed on 8 January 2026)). Most of the stations of the network are installed in the Kamchatka Peninsula. The hypocenter locations for earthquakes in the shallow part of the Benioff zone are subject to significant trade-offs between the lateral position and depth, with errors in the latter reaching a few tens of kilometers [65]. The completeness Mc of the seismic catalog was calculated according to the method of maximum likelihood and equaled 3.0 [66]. The magnitude-frequency distribution is as follows [67]:
log N y e a r 1 = 7.25 0.93 · M .
Figure 2 shows a map of the seismicity of the Kamchatka region. The earthquake distribution clearly traces the architecture of the subduction interface. The distribution of seismicity is essentially heterogeneous in space. One can see that both areas have concentrated epicenters and areas with a set of diffuse points. Seven earthquakes with magnitudes Mw > 7 occurred during the period from January 1990 to July 2025 in the zone under consideration (above N48°). They all localized along the entire length of the Kamchatka trench. In the region of the Kamchatka Peninsula and the Kuril Islands, at depths of 1–70 km, thrust and oblique-thrust fault slips are predominantly observed in earthquake foci. For these events, the azimuths of the compression axes (P)—indicating the direction of maximum stress—range from 90° to 150°, with the axes dipping to the south-east relative to the Earth’s surface. This orientation is consistent with the azimuthal direction of the displacement of the Pacific Plate [68].
In the following study, we analyze earthquakes localized within a slab with MMc. We used the subduction zone geometry model Slab2 [69], and all the hypocenters confined to the slab were projected to its upper boundary. In total, 22,011 events from the original 39,104 seismic events were included.

4. SPAD Algorithm

The SPAD algorithm rests on the following physical concepts. Any fractured rock mass showing VW emits seismic vibrations during structural changes, and the scale of the process governs the magnitudes of seismic events [70,71]. The evolution of the stress state of an asperity continues in time, so the structural features of the interface should appear in the seismicity.
Fault interface heterogeneity can be traced at different scale levels [17,20,72,73], and some events can be triggered in the areas between asperities. The grouping of events should be observed regardless of the time of observation (timeless feature), since it is determined precisely by the localization of asperities. Since an asperity is assumed to have a fractal (self-similar) structure, dense sets of earthquakes are formed in the vicinity of asperity [74,75]. To analyze the structural timeless features of the fault interface, it is necessary to exclude events caused by foreshock and aftershock activities, as well as seismic swarms.
The SPAD algorithm includes three stages, which rest on the available ideas about the physical and geometrical peculiarities of the fault interface:
  • Declustering of the seismic catalog and detection of background seismicity.
  • Fuzzy clustering of background seismicity.
  • Mapping seismogenic patches.

4.1. Background Seismicity

The declustering of the catalog was performed via the calculation of the nearest-neighbor distance in the time–space–magnitude domain. For any two earthquakes i and j, we define the proximity function [76]:
η i j = t i j r i j d f 10 b M i , t i j > 0 + ,                                       t i j 0   ,
where t i j and r i j are the earthquake intercurrence time and the distance between earthquakes, respectively; Mi is the magnitude of event i; b is the scale parameter of the magnitude-frequency distribution (Equation (1)); and df is the fractal dimension of the earthquake spatial distribution. The nearest neighbor η* for a given event j is η * = min ( η i j , i < j ) . Additionally, for excluding aftershocks, we deselect earthquakes with magnitudes smaller than M, which are within a 0.02 × 100.5M km radius circle during the first 0.04 × 100.55M days after a magnitude M event [77].
The distributions of the nearest-neighbor distance η i j * = T i j R i j and its components ( T i j = τ i j 1 0 b M i / 2 , R i j = r i j d f 1 0 b M i / 2 ) are shown in Figure 3. The distributions are prominently bimodal, revealing the existence of two statistically distinct earthquake populations. The 2D distributions of the spatial and temporal components of η i j * clearly separate the two modes. The distribution of the nearest-neighbor proximity function may be considered a Gaussian mixture model with two modes, and the best boundary between modes is η0 = −1.72 [78]. The bimodality of the distribution of the nearest-neighbor proximity function allows us to formally attribute each event to either the background ( η i j >   η 0 ) or cluster ( η i j <   η 0 ) modes (Figure 3a).
The spatial localization of background seismicity shows severe inhomogeneity (Figure 4). At the same time, the populations of background and clustered modes localize in the same areas, and the distributions of the pairwise distances between events of two modes are similar. The characteristic values of the distribution peaks are 85–100 km (Figure 4c).
Localization of background seismicity is timeless and highlights structural features of the fault interface. According to geological studies, large-slip surfaces are smooth, elongated, and quasielliptical bumps, and the geometry of the fault surface predetermines the tectonic asperity configuration [9,10,79]. As clustered seismicity is localized in the vicinity of tectonic asperities, the characteristic peak in the distribution of pairwise distances of both clustered and background seismicity should correspond to the wavelength of bumps and the characteristic distance between tectonic asperities.

4.2. Fuzzy Clustering

Any fault surface is rough at wavelengths up to tens of kilometers [12], and earthquake locations may be used to probe fault surface roughness [80]. The fractal character of fault roughness must determine that background seismicity forms dense clusters of events. To reveal spatially dense clusters, we applied a Discrete Perfect Sets (DPS) algorithm related to fuzzy clustering [81]. The algorithm filters the original space by singling out its densest subset against a general background. The algorithm is described in detail in the work of [82]. The DPS algorithm considers the nonhomogeneity of the event distribution and was used to analyze seismicity [83,84].
The DSP has two free parameters: the index q < 0 to calculate the radius of localization and the parameter β 1 ,   1 , which represents the maximum level of the required density of detected DPS clusters. For a set of recognition objects (epicenters of background seismicity) W, the radius of localization is determined as
r q W = d D W d q D W 1 / q
where d is the Euclidean distance, and D W is the number of nontrivial pairwise distances in the set of recognition objects D(W):
D W = d w 1 , w 2 : w 1 , w 2 W , d w 1 , w 2 0 ,
where w is a point in W. For a specified radius of location r, the density of an arbitrary subset AW at point wW is defined as the sum of the weights of the points localized in the r-neighborhood B A w , r :
P A w = ξ B A w , r 1 d w , ξ r .
The determination of the density level rests on a fuzzy comparison. The measure of comparison of the densities P W W of the set W at all its points wW with values of the density level α is defined as follows [82]:
n P W W , α = P W W α max P W W , α .
The parameter β defines the maximality level of density P against the background of W. The maximality level β uniquely determines the level of density α:
n P W W , α = β
If two parameters q and β are defined, then the DPS algorithm allocates clusters as its own subsets to the set of recognition objects W and builds a set B 1 , . . . , B n , whose density is not less than α at its points:
D P S q , β : W α B 1 , . . . , B n
Applying the DPS algorithm to analyze seismicity implies expert assignment of parameters q and β. Considering the physical aspects of seismicity localization on the fault interface, we can avoid ambiguity in defining the spatial location and number of dense sets. Two physically based conditions must be applied to find a unique solution:
  • Since earthquake localization is predetermined by asperities with fractal roughness at the km scale, one should select a configuration, out of all the possible configurations, for which the number of clusters and the number of events in these clusters are maximized. This configuration of dense sets best represents the current state.
  • If there are several configurations with identical numbers of dense clusters and events in the clusters, the optimal configuration is determined via the analysis of pairwise distances between events in dense clusters. Dense clusters localize in zones of tectonic asperity localization; hence, the characteristic distance between events, which form dense sets, should correspond to the maximum distribution of distance between all the background events (Figure 4c).
Based on the analysis of the full seismic catalog (up to January 2025), 2233 local dense sets were detected in the Kamchatka segment; these sets represent 44% of the background mode (Figure 5a).

4.3. Mapping Seismogenic Patches

Dense sets should be spatially grouped to determine large-scale structural objects—seismogenic patches—which are areas of background seismicity concentration (Figure 5). We used a clustering algorithm based on the construction of the minimal spanning tree in the Euclid metrics [85]. The level of the dendrogram, at which the dense sets are divided into separate space clusters, corresponds to the peak of the distribution of earthquake pairwise distances (Figure 4c). The area where all the dense sets of separate clusters are localized is the seismogenic patch.The contours of the patch are defined as the least convex polygons that embrace all epicenters of the clusters [86].
The configuration of the seismogenic patches becomes apparent at time intervals beginning at 5 years. In short time intervals, the configuration and size of seismogenic patches change significantly; however, the localization of the patches remains unchanged. At time intervals of 15 years or longer, the relative change in the area of seismogenic patches does not exceed 10% of the asperity area. The localization and configuration of seismogenic patches are preserved, i.e., the seismogenic patches are timeless features of background seismicity. Seven seismogenic patches are detected along the Kamchatka segment of the KKSZ. Their characteristic sizes, L, vary from 170 km to 270 km. These sizes correspond to earthquake sources with magnitudes ranging from 7.8 to 8.2 [87].

5. Localization of the Strongest Earthquakes

An analysis of the seismicity along the Kamchatka segment of the KKSZ shows that epicenters of strong earthquakes are distributed along the entire Kamchatka segment. A total of 13 out of 16 strong earthquakes before 1990 and 7 out of 8 after 1990 were localized within the patches (Figure 6). Notably, these events were not considered when detecting seismogenic patches because they were removed during declustering via the nearest-neighbor method. It is worth mentioning that the maximum magnitude of background mode is 6.3.
The foci of strong earthquakes are large, and hypocenters point to only the starting point of tectonic asperity rupture. The configuration of tectonic asperities is revealed retrospectively by the inversion of seismic and/or geodetic data. At the same time, the configuration of tectonic asperities controls earthquake nucleation and failure processes. Scenarios of earthquake rupture evolution suggest the rupture of both a single asperity and several neighboring asperities synchronously [9,89,90].
Since seismogenic patches are a timeless feature of the fault interface, we assume that seismogenic patches are, in fact, seismic sketches of tectonic asperities. To support this supposition, we compared a map of seismogenic patches with the slip inversion models of the strongest earthquakes (Figure 7).
There are four earthquakes in the Kamchatka segment of the KKSZ, for which maps of the slip inversion models are available (Table 1). Earthquakes have a complex source structure, and several maxima in the slip distribution can be detected. The maxima of slip distribution for earthquakes 1 and 4 (Table 1) are localized in different sections of KKSZ within the separate patches. The 1997 Kronotsky earthquake is characterized by two distribution maxima, which are localized on neighboring patches. For the 2025 megaearthquake, there are three maxima that are localized within three different patches near their contours to the south of 52° N.

6. Discussion

When the physics of a process is uncertain, machine-learning algorithms demonstrate high efficiency [92,93]. The developed SPAD algorithm aimed at studying features of seismicity allows us to better define the geometry and heterogeneity of the subduction interface and to unambiguously contour the tectonic asperity of major earthquakes. The SPAD algorithm is based on unsupervised learning and does not use a priori data about the localization of strong earthquakes. To avoid ambiguity in detecting seismogenic patches, we introduce two criteria, which account for the physics of the process and the geological structure of the tectonic fault. It is thought that changes in the stress-strain state of tectonic asperity take place without the emission of seismic vibrations, which formally results in the occurrence of seismic gaps [94,95]. This statement was true when seismic arrays could not register earthquakes with magnitudes less than 4, but from the point of view of rock friction, it is false. When the background seismicity is reliably recorded (the magnitude of background seismicity is essentially lower than the magnitude of the main event), the seismic gap should be interpreted as a lack of strong earthquakes but not as a complete absence of seismicity. The main part of seismic events is concentrated in the zones of asperities. The background seismicity shows the timeless structural features of a tectonic fault and reflects the different-scale roughness of tectonic faults. The structure of asperities has been preserved for tens of years. We showed that large-scale roughness controls asperity spacing (Figure 7). At small scales, it manifests via the fractality and self-similarity of asperities [11,12,75]. An analysis of waveforms of seismic events confined to the Japan Trench in the Naka district revealed that all earthquakes with magnitudes ranging from 2 to 5, whose hypocenters are in a small neighborhood, were characterized by identical wave onset times, while the time interval between events was approximately 15 years [96].
Algorithms for detecting earthquake-prone areas are aimed at identifying zones where strong earthquakes can occur [97,98]. As a rule, algorithms are based on supervised learning, which means that they use information about previous strong earthquakes. A distinctive feature of the SPAD algorithm is the construction of a map of tectonic asperities, which can define the configuration of the earthquake source and the dynamics of fault rupture [99]. We have shown that zones where strong earthquakes nucleate can be detected via an analysis of background seismicity when there is no prior information about the localization of strong earthquakes. Seismogenic patches are sketchy displays of tectonic asperities. For the Kamchatka subduction zone, we can unambiguously say that we realize the configuration and sizes of tectonic asperities.
Revealing the map of tectonic asperities based on the physical aspects of fault evolution improves our understanding of earthquake nucleation and failure processes. The SPAD method can be used to prognosticate possible earthquake scenarios. In the future, if the accuracy of the location remains high enough, it is reasonable to perform estimations of the interface roughness [80], which controls the earthquake size and the stress drop on a tectonic asperity [13]. The Kamchatka earthquakes have shown that each tectonic asperity may independently rupture many times before it can rupture with neighboring asperities [33,100]. The synchronization of a rupture of two or more asperities is a challenge for assessing the geodynamic hazard, and thus far, it is uncertain.

7. Conclusions

We present a new technique, SPAD, to infer tectonic asperities at the subduction interface. SPAD does not use any information about strong earthquakes or any expertly specified parameters, and the identified seismogenic patches should be associated with tectonic asperities. The physical parameterization of subduction interface can be used to prognosticate earthquake scenarios.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/geosciences16010038/s1. CATALOG.csv—Analyzed earthquake catalog; Requirements.txt—List of Python package dependencies and library versions required to run the SPAD algorithm code; spad.py—Python implementation of the SPAD algorithm.

Author Contributions

Conceptualization, A.O.; methodology, A.O.; software, I.N.; formal analysis, I.N.; investigation, I.N. and A.O.; resources, A.O.; data curation, I.N.; writing—original draft preparation, A.O. and I.N.; writing—review and editing, A.O. and I.N.; visualization, I.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was performed by N.I. within the framework of the IDG RAS project no. 125012700824-4, and was partially funded by Russian Science Foundation, grant number 22-17-00204 to A.O.

Data Availability Statement

The seismicity catalog used in our analysis is available at the website of the Kamchatka Branch of the Geophysical Survey of the Russian Academy of Sciences: https://sdis.emsd.ru/info/earthquakes/catalogue.php (accessed on 8 January 2026). The analyzed catalog and SPAD algorithm code are presented in the Supplementary Materials. Any additional data presented in this study is available from the corresponding author upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Keilis-Borok, V.I.; Kossobokov, V.G. Premonitory activation of earthquake flow: Algorithm M8. Phys. Earth Planet. Inter. 1990, 61, 73–83. [Google Scholar] [CrossRef]
  2. Peresan, A.; Kossobokov, V.; Romashkova, L.; Panza, G.F. Intermediate-term middle-range earthquake predictions in Italy: A review. Earth-Sci. Rev. 2005, 69, 97–132. [Google Scholar] [CrossRef]
  3. Vijay, R.K.; Nanda, S.J.; Sharma, A. A review on clustering algorithms for spatiotemporal seismicity analysis. Artif. Intell. Rev. 2025, 58, 231. [Google Scholar] [CrossRef]
  4. Mignan, A.; Broccardo, M. Neural Network Applications in Earthquake Prediction (1994–2019): Meta-Analytic and Statistical Insights on Their Limitations. Seismol. Res. Lett. 2020, 91, 2330–2342. [Google Scholar] [CrossRef]
  5. Ruff, L.; Kanamori, H. Seismicity and the subduction process. Phys. Earth Planet. Inter. 1980, 23, 240–252. [Google Scholar] [CrossRef]
  6. Kanamori, H. Lessons from the 2004 Sumatra–Andaman earthquake. Philos. Trans. R. Soc. A 2006, 364, 1927–1945. [Google Scholar] [CrossRef] [PubMed]
  7. Bletery, Q.; Thomas, A.M.; Rempel, A.W.; Karlstrom, L.; Sladen, A.; De Barros, L. Mega-earthquakes rupture flat megathrusts. Science 2016, 354, 1027–1031. [Google Scholar] [CrossRef] [PubMed]
  8. McCaffrey, R. Global frequency of magnitude 9 earthquakes. Geology 2008, 36, 263–266. [Google Scholar] [CrossRef]
  9. Lay, T.; Kanamori, H.; Ruff, L. The Asperity Model and the Nature of Large Subduction Zone Earthquakes. Earthq. Predict. Res. 1982, 1, 3–71. [Google Scholar]
  10. Gzovsky, M.V. Foundation of Tectonophysics; Nauka: Moscow, Russia, 1975; p. 536. (In Russian) [Google Scholar]
  11. Sagy, A.; Brodsky, E.E.; Axen, G.J. Evolution of fault-surface roughness with slip. Geology 2007, 35, 283–286. [Google Scholar] [CrossRef]
  12. Candela, T.; Renard, F.; Klinger, Y.; Mair, K.; Schmittbuhl, J.; Brodsky, E.E. Roughness of Fault Surfaces over Nine Decades of Length Scales. J. Geophys. Res. 2012, 117, B08409. [Google Scholar] [CrossRef]
  13. Zielke, O.; Galis, M.; Mai, P.M. Fault roughness and strength heterogeneity control earthquake size and stress drop. Geophys. Res. Lett. 2017, 44, 777–783. [Google Scholar] [CrossRef]
  14. Byerlee, J.D. Frictional characteristics of granite under high contining pressure. J. Geophys. Res. 1967, 72, 3639–3648. [Google Scholar] [CrossRef]
  15. Engelder, J.T.; Scholz, C.H. The role of asperity indentation and ploughing in rock friction—II: Influence of relative hardness and normal load. Int. J. Rock Mech. Min. Sci. 1976, 13, 155–163. [Google Scholar] [CrossRef]
  16. Kaproth, B.M.; Marone, C. Slow Earthquakes, Preseismic Velocity Changes, and the Origin of Slow Frictional Stick-Slip. Science 2013, 341, 1229–1232. [Google Scholar] [CrossRef]
  17. Fagereng, A.; Beall, A. Is Complex Fault Zone Behaviour a Reflection of Rheological Heterogeneity? Phil. Trans. R. Soc. A 2021, 379, 20190421. [Google Scholar] [CrossRef]
  18. Collettini, C.; Barchi, M.R.; De Paola, N.; Trippetta, F.; Tinti, E. Rock and fault rheology explain differences between on fault and distributed seismicity. Nat. Commun. 2022, 13, 5627. [Google Scholar] [CrossRef] [PubMed]
  19. Collettini, C.; Tesei, T.; Scuderi, M.M.; Carpenter, B.M.; Viti, C. Beyond Byerlee friction, weak faults, and implications for slip behavior. Earth Planet. Sci. Lett. 2019, 519, 245–263. [Google Scholar] [CrossRef]
  20. Ruzhich, V.V.; Kocharyan, G.G.; Ostapchuk, A.A.; Shilko, E.V. Different-Scale Heterogeneities in Segments of Active Faults and Their Influence on Slip Modes. Phys. Mesomech. 2024, 27, 217–228. [Google Scholar] [CrossRef]
  21. Fagereng, A. Frequency-size distribution of competent lenses in a block-in-matrix mélange: Imposed length scales of brittle deformation? J. Geophys. Res. 2011, 116, B05302. [Google Scholar] [CrossRef]
  22. Dublanchet, P.; Bernard, P.; Favreau, P. Interactions and triggering in a 3-D rate-and-state asperity model. J. Geophys. Res. 2013, 118, 2225–2245. [Google Scholar] [CrossRef]
  23. Gounon, A.; Latour, S.; Letort, J.; El Arem, S. Rupture nucleation on a periodically heterogeneous interface. Geophys. Res. Lett. 2022, 49, e2021GL096816. [Google Scholar] [CrossRef]
  24. Bedford, J.D.; Faulkner, D.R.; Lapusta, N. Fault rock heterogeneity can produce fault weakness and reduce fault stability. Nat. Commun. 2022, 13, 326. [Google Scholar] [CrossRef]
  25. Gridin, G.A.; Kocharyan, G.G.; Morozova, K.G.; Novikova, E.V.; Ostapchuk, A.A.; Pavlov, D.V. Evolution of Sliding Along a Heterogenous Fault. A Large-Scale Laboratory Experiment. Izv. Phys. Solid Earth 2023, 59, 139–147. [Google Scholar] [CrossRef]
  26. Scholz, C.H.; Campos, J. The Seismic Coupling of Subduction Zones Revisited. J. Geophys. Res. 2012, 117, B05310. [Google Scholar] [CrossRef]
  27. Frank, W.B.; Shapiro, N.M.; Husker, A.L.; Kostoglodov, V.; Gusev, A.A.; Campillo, M. The evolving interaction of low-frequency earthquakes during transient slip. Sci. Adv. 2016, 2, e1501616. [Google Scholar] [CrossRef]
  28. Ostapchuk, A.A.; Morozova, K.G.; Besedina, A.N.; Gridin, G.A.; Grygoryeva, A.; Pavlov, D. Features of seismicity triggered by ripple-fired explosions at the Korobkovskoye iron ore deposit. Russ. J. Earth Sci. 2024, 24, 5013. [Google Scholar] [CrossRef]
  29. Luo, Y.; Ampuero, J.P. Stability of faults with heterogeneous friction properties and effective normal stress. Tectonophysics 2018, 733, 257–272. [Google Scholar] [CrossRef]
  30. Ide, S. Modeling fast and slow earthquakes at various scales. Proc. Jpn. Acad. Ser. B 2014, 90, 259–277. [Google Scholar] [CrossRef]
  31. Noda, H.; Lapusta, N. Stable creeping fault segments can become destructive as a result of dynamic weakening. Nature 2013, 493, 518–521. [Google Scholar] [CrossRef]
  32. Kositsky, A.P.; Avouac, J.-P. Inverting geodetic time series with a principal component analysis-based inversion method. J. Geophys. Res. 2010, 115, B03401. [Google Scholar] [CrossRef]
  33. Chlieh, M.; Mothes, P.A.; Nocquet, J.-M.; Jarrin, P.; Charvis, P.; Cisneros, D.; Font, Y.; Collot, J.-Y.; Villegas-Lanza, J.-C.; Rolandone, F.; et al. Distribution of discrete seismic asperities and aseismic slip along the Ecuadorian megathrust. Earth Planet. Sci. Lett. 2014, 400, 292–301. [Google Scholar] [CrossRef]
  34. Li, Y.; Shan, X.; Gao, Z.; Huang, X. Interseismic coupling, asperity distribution, and earthquake potential on major faults in southeastern Tibet. Geophys. Res. Lett. 2023, 50, e2022GL101209. [Google Scholar] [CrossRef]
  35. Métois, M.; Vigny, C.; Socquet, A. Interseismic Coupling, Megathrust Earthquakes and Seismic Swarms along the Chilean Subduction Zone (38°–18°S). Pure Appl. Geophys. 2016, 173, 1431–1449. [Google Scholar] [CrossRef]
  36. Hikima, K.; Koketsu, K. Rupture processes of the 2004 Chuetsu (mid-Niigata prefecture) earthquake, Japan: A series of events in a complex fault system. Geophys. Res. Lett. 2005, 32, L18303. [Google Scholar] [CrossRef]
  37. Cirella, A.; Piatanesi, A.; Tinti, E.; Chini, M.; Cocco, M. Complexity of the rupture process during the 2009 L’Aquila, Italy, earthquake. Geophys. J. Int. 2012, 190, 607–621. [Google Scholar] [CrossRef]
  38. Yoshida, S.; Koketsu, K.; Shibazaki, B.; Sagiya, T.; Kato, T.; Yoshida, Y. Joint Inversion of Near- and Far-field Waveforms and Geodetic Data for the Rupture Process of the 1995 Kobe Earthquake. J. Phys. Earth 1996, 44, 437–454. [Google Scholar] [CrossRef]
  39. Guo, R.; Tang, X.; Zhang, Y.; Qin, M.; Xu, J.; Zhou, J.; Zou, X.; Sun, H. Seismic versus aseismic slip for the 2023 Kahramanmaraş earthquake doublet. Nat. Commun. 2025, 16, 959. [Google Scholar] [CrossRef]
  40. Kocharyan, G.G.; Ostapchuk, A.A.; Gridin, G.A.; Kishkina, S.B. The Role of Sliding Surface Macrostructure in the Nucleation and Development of Dynamic Instability in the Upper Part of the Earth’s Crust. Izv. Phys. Solid Earth 2025, 61, 691–702. [Google Scholar] [CrossRef]
  41. Taymaz, T.; Tan, O.; Yolsal, S. Sumatra Earthquake (Mw~9.3) of December 26, 2004 Source Rupture Processes and Slip Distribution Modeling: Preliminary Rupture Model. Available online: http://web.itu.edu.tr/taymaz/sumatra/index.html (accessed on 11 December 2025).
  42. Lee, S.-J.; Huang, B.-S.; Ando, M.; Chiu, H.-C.; Wang, J.-H. Evidence of large scale repeating slip during the 2011 Tohoku-Oki earthquake. Geophys. Res. Lett. 2011, 38, L19306. [Google Scholar] [CrossRef]
  43. Lee, S.-J.; Yeh, T.-Y.; Lin, T.-C.; Lin, Y.-Y.; Song, T.-R.A.; Huang, B.-S. Two-stage composite megathrust rupture of the 2015 Mw8.4 Illapel, Chile, earthquake identified by spectral-element inversion of teleseismic waves. Geophys. Res. Lett. 2016, 43, 4979–4985. [Google Scholar] [CrossRef]
  44. Lin, Y.-Y.; Lapusta, N. Microseismicity simulated on asperity-like fault patches: On scaling of seismic moment with duration and seismological estimates of stress drops. Geophys. Res. Lett. 2018, 45, 8145–8155. [Google Scholar] [CrossRef]
  45. Valoroso, L.; Chiaraluce, L.; Collettini, C. Earthquakes and fault zone structure. Geology 2014, 42, 343–346. [Google Scholar] [CrossRef]
  46. Trugman, D.T.; Ross, Z.E. Pervasive foreshock activity across southern California. Geophys. Res. Lett. 2019, 46, 8772–8781. [Google Scholar] [CrossRef]
  47. Ostapchuk, A.; Polyatykin, V.; Popov, M.; Kocharyan, G. Seismogenic Patches in a Tectonic Fault Interface. Front. Earth Sci. 2022, 10, 904814. [Google Scholar] [CrossRef]
  48. Chalumeau, C.; Agurto-Detzel, H.; Rietbrock, A.; Frietsch, M.; Oncken, O.; Segovia, M.; Galve, A. Seismological evidence for a multifault network at the subduction interface. Nature 2024, 628, 558–562. [Google Scholar] [CrossRef]
  49. Ostapchuk, A.A. Displaying the Structure and Rheological Properties of a Fault Zone in the Characteristics of Background Seismicity. Dokl. Earth Sci. 2025, 523, 31. [Google Scholar] [CrossRef]
  50. Kanamori, H. The energy release in great earthquakes. J. Geophys. Res. 1977, 82, 2981–2987. [Google Scholar] [CrossRef]
  51. Walker, K.T.; Ishii, M.; Shearer, P.M. Rupture details of the 28 March 2005 Sumatra Mw 8.6 earthquake imaged with teleseismic P waves. Geophys. Res. Lett. 2005, 32, L24303. [Google Scholar] [CrossRef]
  52. Lay, T. A review of the rupture characteristics of the 2011 Tohoku-oki Mw 9.1 earthquake. Tectonophysics 2018, 733, 4–36. [Google Scholar] [CrossRef]
  53. Wirth, E.A.; Sahakian, V.J.; Wallace, L.M.; Melnick, D. The occurrence and hazards of great subduction zone earthquakes. Nat. Rev. Earth Environ. 2022, 3, 125–140. [Google Scholar] [CrossRef]
  54. Bushenkova, N.; Koulakov, I.; Bergal-Kuvikas, O.; Shapiro, N.; Gordeev, E.I.; Chebrov, D.V.; Abkadyrov, I.; Jakovlev, A.; Stupina, T.; Novgorodova, A.; et al. Connections between arc volcanoes in Central Kamchatka and the subducting slab inferred from local earthquake seismic tomography. J. Volcanol. Geotherm. Res. 2023, 435, 107768. [Google Scholar] [CrossRef]
  55. Fedotov, S.A.; Solomatin, A.V.; Chernyshev, S.D. A Long-Term Earthquake Forecast for the Kuril–Kamchatka Arc for the Period from September 2010 to August 2015 and the Reliability of Previous Forecasts, as well as their Applications. J. Volcanol. Seismol. 2011, 5, 75–99. [Google Scholar] [CrossRef]
  56. Hindle, D.; Fujita, K.; Mackey, K. Deformation of the Northwestern Okhotsk Plate: How is it happening? Stephan Mueller Spec. Publ. Ser. 2009, 4, 147–156. [Google Scholar] [CrossRef]
  57. Gordeev, E.I.; Bergal-Kuvikas, O.V. Structure of subduction zone and volcanism on Kamchatka. Dokl. Earth Sci. 2022, 502, 21–24. [Google Scholar] [CrossRef]
  58. Tibaldi, A. Recent surface faulting investigated in Kamchatka volcanic arc. Eos Trans. AGU 2004, 85, 133–141. [Google Scholar] [CrossRef]
  59. DeMets, C.; Gordon, R.G.; Argus, D.F.; Stein, S. Current plate motions. Geophys. J. Int. 1990, 101, 425–478. [Google Scholar] [CrossRef]
  60. Steblov, G.M.; Kogan, M.G.; King, R.W.; Scholz, C.H.; Bürgmann, R.; Frolov, D.I. Imprint of the North American plate in Siberia revealed by GPS. Geophys. Res. Lett. 2003, 30, 1924. [Google Scholar] [CrossRef]
  61. Fedotov, S.A. Seismicheskoje Rajonirovanie S.S.S.R.; Nauka: Moscow, Russia, 1968; pp. 121–150. [Google Scholar]
  62. Fedotov, S.A.; Solomatin, A.V. Long-Term Earthquake Prediction (LTEP) for the Kuril–Kamchatka island arc, June 2019 to May 2024; Properties of Preceding Seismicity from January 2017 to May 2019. The Development and Practical Application of the LTEP Method. J. Volcanol. Seismol. 2019, 13, 349–362. [Google Scholar] [CrossRef]
  63. Johnson, J.; Satake, K. Asperity Distribution of the 1952 Great Kamchatka Earthquake and its Relation to Future Earthquake Potential in Kamchatka. Pure Appl. Geophys. 1999, 154, 541–553. [Google Scholar] [CrossRef]
  64. MacInnes, B.T.; Weiss, R.; Bourgeois, J.; Pinegina, T.K. Slip Distribution of the 1952 Kamchatka Great Earthquake Based on Near-Field Tsunami Deposits and Historical Records. Bull. Seismol. Soc. Am. 2010, 100, 1695–1709. [Google Scholar] [CrossRef]
  65. Droznin, D.V.; Droznina, S.Y.; Senyukov, S.L.; Chebrov, D.V.; Shapiro, N.M.; Shebalin, P.N. Probabilistic Estimates of Hypocenters from the Data of Kamchatka Seismic Network Stations. Izv. Phys. Solid Earth 2019, 55, 677–687. [Google Scholar] [CrossRef]
  66. Wiemer, S.; Wyss, M. Minimum Magnitude of Completeness in Earthquake Catalogs: Examples from Alaska, the Western United States, and Japan. Bull. Seismol. Soc. Am. 2000, 90, 859–869. [Google Scholar] [CrossRef]
  67. Gutenberg, R.; Richter, C.F. Frequency of Earthquakes in California. Bull. Seismol. Soc. Am. 1944, 34, 185. [Google Scholar] [CrossRef]
  68. Bushenkova, N.; Kuchay, O.A. Focal mechanisms of earthquakes in the subduction zone of the Western Pacific plate. Geodyn. Tectonophys. 2022, 13, 0639. [Google Scholar] [CrossRef]
  69. Hayes, G.P.; Moore, G.L.; Portner, D.E.; Hearne, M.; Flamme, H.; Furtney, M.; Smoczyk, G.M. Slab2, a comprehensive subduction zone geometry model. Science 2018, 362, 58–61. [Google Scholar] [CrossRef]
  70. Brune, J.N. Tectonic stress and the spectra of seismic shear waves from earthquakes. J. Geophys. Res. 1970, 75, 4997–5009. [Google Scholar] [CrossRef]
  71. Kuksenko, V.S.; Lyashkov, A.I.; Mirzoev, K.M.; Negmatullaev, S.H.; Stanchits, S.A.; Frolov, D.I. Relation between sizes of the cracks formed under load and duration of elastic energy release. Dokl. AN SSSR 1982, 264, 846–848. [Google Scholar]
  72. King, G.C.P. Speculations on the geometry of the initiation and termination processes of earthquake rupture and its relation to morphology and geological structure. Pure Appl. Geophys. 1986, 124, 567–585. [Google Scholar] [CrossRef]
  73. Behr, W.M.; Bürgmann, R. What’s downthere? The structures, materials and environment of deep-seated slow slip and tremor. Phil. Trans. R. Soc. A 2021, 379, 20200218. [Google Scholar] [CrossRef]
  74. Seno, T. Fractal Asperities, Invasion of Barriers, and Interplate Earthquakes. Earth Planets Space 2003, 55, 649–665. [Google Scholar] [CrossRef]
  75. Kocharyan, G.G.; Ostapchuk, A.A. Mesostructure of a Tectonic Fault Slip Zone. Phys. Mesomech. 2023, 26, 82–92. [Google Scholar] [CrossRef]
  76. Baiesi, M.; Paczuski, M. Scale-free Networks of Earthquakes and Aftershocks. Phys. Rev. E 2004, 69, 066106. [Google Scholar] [CrossRef]
  77. Shebalin, P.; Narteau, C. Depth Dependent Stress Revealed by Aftershocks. Nat. Commun. 2017, 8, 1317. [Google Scholar] [CrossRef]
  78. Zaliapin, I.; Ben-Zion, Y. A global classification and characterization of earthquake clusters. Geophys. J. Int. 2016, 207, 608–634. [Google Scholar] [CrossRef]
  79. Alaei, B.; Torabi, A. Fault asperity and roughness, insight from faults in 3D reflection seismic data. Mar. Pet. Geol. 2024, 170, 107145. [Google Scholar] [CrossRef]
  80. Cochran, E.S.; Page, M.T.; van der Elst, N.J.; Ross, Z.E.; Trugman, D.T. Fault Roughness at Seismogenic Depths and Links to Earthquake Behavior. Seism. Rec. 2023, 3, 37–47. [Google Scholar] [CrossRef]
  81. Agayan, S.M.; Bogoutdinov, S.R.; Dobrovolsky, M.N. Discrete Perfect Sets and Their Application in Cluster Analysis. Cybern. Syst. Anal. 2014, 50, 176–190. [Google Scholar] [CrossRef]
  82. Agayan, S.M.; Bogoutdinov, S.R.; Dzeboev, B.A.; Dzeranov, B.V.; Kamaev, D.A.; Osipov, M.O. DPS Clustering: New Results. Appl. Sci. 2022, 12, 9335. [Google Scholar] [CrossRef]
  83. Soloviev, A.A.; Gvishiani, A.D.; Gorshkov, A.I.; Dobrovolsky, M.N.; Novikova, O.V. Recognition of earthquake-prone areas: Methodology and analysis of the results. Izv. Phys. Solid Earth 2014, 50, 151–168. [Google Scholar] [CrossRef]
  84. Gvishiani, A.D.; Dzeboev, B.A.; Dzeranov, B.V.; Kedrov, E.O.; Skorkina, A.A.; Nikitina, I.M. Strong Earthquake-Prone Areas in the Eastern Sector of the Arctic Zone of the Russian Federation. Appl. Sci. 2022, 12, 11990. [Google Scholar] [CrossRef]
  85. Jain, A.K.; Murty, M.N.; Flynn, P.J. Data clustering: A review. ACM Comput. Surv. 1999, 31, 264–323. [Google Scholar] [CrossRef]
  86. Chan, T.M. Optimal output-sensitive convex hull algorithms in two and three dimensions. Discrete Comput. Geom. 1996, 16, 361–368. [Google Scholar] [CrossRef]
  87. Kocharyan, G.G.; Kishkina, S.B. The Physical Mesomechanics of the Earthquake Source. Phys. Mesomech 2021, 24, 343–356. [Google Scholar] [CrossRef]
  88. Pacheco, J.F.; Sykes, L.R. Seismic moment catalog of karge shallow earthquakes, 1900 to 1989. Bull. Seismol. Soc. Am. 1992, 82, 1306–1349. [Google Scholar] [CrossRef]
  89. Vallée, M.; Douet, V. A new database of Source Time Functions (STFs) extracted from the SCARDEC method. Phys. Earth Planet. Int. 2016, 257, 149–157. [Google Scholar] [CrossRef]
  90. Ye, L.; Kanamori, H.; Lay, T. Global variations of large megathrust earthquake rupture characteristics. Sci. Adv. 2018, 4, eaao4915. [Google Scholar] [CrossRef] [PubMed]
  91. Xu, D.; Wang, W.; Ren, Z. Preliminary slip distribution of the July 29, 2025 Mw 8.8 Kamchatka, Russia earthquake. Earthq. Res. Adv. 2025; in press. [Google Scholar] [CrossRef]
  92. Bergen, K.J.; Johnson, P.A.; de Hoop, M.V.; Beroza, G.C. Machine learning for data-driven discovery in solid Earth geoscience. Science 2019, 363, eaau0323. [Google Scholar] [CrossRef]
  93. Ren, C.X.; Hulbert, C.; Johnson, P.A.; Rouet-Leduc, B. Machine learning and fault rupture: A review. Adv. Geophys. 2020, 61, 57–107. [Google Scholar] [CrossRef]
  94. McCann, W.R.; Nishenko, S.P.; Sykes, L.R.; Krause, J. Seismic gaps and plate tectonics: Seismic potential for major boundaries. Pure Appl. Geophys. 1979, 117, 1082–1147. [Google Scholar] [CrossRef]
  95. Lay, T.; Nishenko, S.P. Updated concepts of seismic gaps and asperities to assess great earthquake hazard along South America. Proc. Natl. Acad. Sci. USA 2022, 119, e2216843119. [Google Scholar] [CrossRef]
  96. Ide, S. Frequent observations of identical onsets of large and small earthquakes. Nature 2019, 573, 112–116. [Google Scholar] [CrossRef]
  97. Gelfand, I.M.; Guberman, S.A.; Keilis-Borok, V.I.; Knopoff, L.; Press, F.S.; Ranzman, E.Y.; Rotwain, I.M.; Sadovsky, A.M. Pattern recognition applied to earthquakes epicenters in California. Phys. Earth Planet. Inter. 1976, 11, 227–283. [Google Scholar] [CrossRef]
  98. Kossobokov, V.G.; Soloviev, A.A. Pattern recognition in problems of seismic hazard assessment. Chebyshevskii Sb. 2018, 19, 55–90. (In Russian) [Google Scholar] [CrossRef]
  99. Yamanaka, Y.; Kikuchi, M. Asperity Map along the Subduction Zone in Northeastern Japan Inferred from Regional Seismic Data. J. Geophys. Res. 2004, 109, B07307. [Google Scholar] [CrossRef]
  100. Kanamori, H.; McNally, K.C. Variable rupture mode of the subduction zone along the Ecuador-Colombia coast. Bull. Seismol. Soc. Am. 1982, 72, 1241–1253. [Google Scholar]
Figure 1. Tectonics of the Okhotsk plate and surroundings. Location and slip rates of tectonic plates (OK—Okhotsk, PA—Pacific, EU—Eurasia, NA—North America) are shown. The Kuril–Kamchatka subduction zone is denoted by a thick dashed line. Slip rates are presented in millimeters per year. The inset map at the bottom left shows the location of the study area (red box) in the larger map of the Asia–Pacific region.
Figure 1. Tectonics of the Okhotsk plate and surroundings. Location and slip rates of tectonic plates (OK—Okhotsk, PA—Pacific, EU—Eurasia, NA—North America) are shown. The Kuril–Kamchatka subduction zone is denoted by a thick dashed line. Slip rates are presented in millimeters per year. The inset map at the bottom left shows the location of the study area (red box) in the larger map of the Asia–Pacific region.
Geosciences 16 00038 g001
Figure 2. Map of seismicity. The earthquakes presented are recorded by the Kamchatka branch of Geophysical Survey of RAS from January 1990 to July 2025. The colors indicate the depth of earthquake hypocenters. Focal mechanisms for the earthquakes with Mw > 7 correspond to the GCMT ones (available at https://www.globalcmt.org accessed on 8 January 2026). The black line with teeth denotes the Kuril–Kamchatka (1) and the Aleutian (2) trenches location on the upper plate. Colored isolines indicate mean slab depths (available online https://www.usgs.gov/data/slab2-a-comprehensive-subduction-zone-geometry-model accessed on 8 January 2026).
Figure 2. Map of seismicity. The earthquakes presented are recorded by the Kamchatka branch of Geophysical Survey of RAS from January 1990 to July 2025. The colors indicate the depth of earthquake hypocenters. Focal mechanisms for the earthquakes with Mw > 7 correspond to the GCMT ones (available at https://www.globalcmt.org accessed on 8 January 2026). The black line with teeth denotes the Kuril–Kamchatka (1) and the Aleutian (2) trenches location on the upper plate. Colored isolines indicate mean slab depths (available online https://www.usgs.gov/data/slab2-a-comprehensive-subduction-zone-geometry-model accessed on 8 January 2026).
Geosciences 16 00038 g002
Figure 3. Declustering of seismic catalog. (a) Projection of the background (gray) and clustered (red) modes of seismicity on the upper boundary of the slab. The background mode includes 14,936 events, clustered—5880 ones. Isolines indicate mean slab depths with steps 25 km. (b) Histogram of proximity function η*; the values are normalized to sum up to unity. The dashed line corresponds to the threshold ( l g η 0 = 1.72 ) separating the clustered and background modes (solid lines show two modes of Gaussian mixture model); (c) A joint distribution of rescaled time and space components (Tij,Rij) of the nearest-neighbor distance η i j * = T i j R i j . The dashed line l g R i j + l g T i j = 1.72 separates the two modes.
Figure 3. Declustering of seismic catalog. (a) Projection of the background (gray) and clustered (red) modes of seismicity on the upper boundary of the slab. The background mode includes 14,936 events, clustered—5880 ones. Isolines indicate mean slab depths with steps 25 km. (b) Histogram of proximity function η*; the values are normalized to sum up to unity. The dashed line corresponds to the threshold ( l g η 0 = 1.72 ) separating the clustered and background modes (solid lines show two modes of Gaussian mixture model); (c) A joint distribution of rescaled time and space components (Tij,Rij) of the nearest-neighbor distance η i j * = T i j R i j . The dashed line l g R i j + l g T i j = 1.72 separates the two modes.
Geosciences 16 00038 g003
Figure 4. Localization of seismicity in the strike plane of the fault. (a) Distribution of background and clustered modes over the subduction interface (the size of the cell is 10 × 10 km2). (b) The Gutenberg–Richter statistics for the background (gray) and clustered (red) modes of seismicity; (c) relative pairwise distance between hypocenters of earthquakes for two modes.
Figure 4. Localization of seismicity in the strike plane of the fault. (a) Distribution of background and clustered modes over the subduction interface (the size of the cell is 10 × 10 km2). (b) The Gutenberg–Richter statistics for the background (gray) and clustered (red) modes of seismicity; (c) relative pairwise distance between hypocenters of earthquakes for two modes.
Geosciences 16 00038 g004
Figure 5. Mapping seismogenic patches and earthquake epicenters with Mw > 7 (stars). (a) Configuration of seismogenic patches over a 35-year interval (full catalog). Gray dots show the background seismicity (14,936 events), colored dots—dense sets of background seismicity (including 6593 events). Colored contours are seismogenic patches. (b) Configuration of seismogenic patches on 25-year interval (red), 15-year (yellow), 10-year (blue) and 5-year (black) from January 2000. (c) Configuration of seismogenic patches on 25-year interval (red), 15-year (yellow), 10-year (blue) and 5-year (black) from January 1990.
Figure 5. Mapping seismogenic patches and earthquake epicenters with Mw > 7 (stars). (a) Configuration of seismogenic patches over a 35-year interval (full catalog). Gray dots show the background seismicity (14,936 events), colored dots—dense sets of background seismicity (including 6593 events). Colored contours are seismogenic patches. (b) Configuration of seismogenic patches on 25-year interval (red), 15-year (yellow), 10-year (blue) and 5-year (black) from January 2000. (c) Configuration of seismogenic patches on 25-year interval (red), 15-year (yellow), 10-year (blue) and 5-year (black) from January 1990.
Geosciences 16 00038 g005
Figure 6. Linking seismogenic patches and epicenters of shallow strong (M > 7) earthquakes. Colored polygons denote seismogenic patches. Black stars—epicenters of earthquakes before 1990 [88], white stars—earthquakes after 1990.
Figure 6. Linking seismogenic patches and epicenters of shallow strong (M > 7) earthquakes. Colored polygons denote seismogenic patches. Black stars—epicenters of earthquakes before 1990 [88], white stars—earthquakes after 1990.
Geosciences 16 00038 g006
Figure 7. Comparison of the map of seismogenic patches to slip distribution on fault planes for the largest earthquakes: (a) Mw8.8 on 30 July 2025. (b) Mw7.8 on September 13, 2025. (c) Mw7.5 on July 8, 1993. (d) Mw7.8 on 5 December 1997. Focal mechanisms correspond to the GCMT catalog. Information on maps of slip distribution on the fault plane is presented in Table 1.
Figure 7. Comparison of the map of seismogenic patches to slip distribution on fault planes for the largest earthquakes: (a) Mw8.8 on 30 July 2025. (b) Mw7.8 on September 13, 2025. (c) Mw7.5 on July 8, 1993. (d) Mw7.8 on 5 December 1997. Focal mechanisms correspond to the GCMT catalog. Information on maps of slip distribution on the fault plane is presented in Table 1.
Geosciences 16 00038 g007
Table 1. List of the earthquakes Mw ≥ 7.5 of the Kamchatka segment with available slip distribution on the fault plane.
Table 1. List of the earthquakes Mw ≥ 7.5 of the Kamchatka segment with available slip distribution on the fault plane.
No.DataMagnitude, MwSource
18 July 19937.5https://earthquake.usgs.gov/earthquakes/eventpage/usp0005u7b/finite-fault (accessed on 8 January 2026)
25 December 19977.8https://earthquake.usgs.gov/earthquakes/eventpage/usp0008btk/finite-fault (accessed on 8 January 2026)
329 July 20258.8[91]
413 September 20257.5https://earthquake.usgs.gov/earthquakes/eventpage/us7000qx2g/finite-fault (accessed on 8 January 2026)
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

Ostapchuk, A.; Nugmanov, I. Background Seismicity Highlights Tectonic Asperities. Geosciences 2026, 16, 38. https://doi.org/10.3390/geosciences16010038

AMA Style

Ostapchuk A, Nugmanov I. Background Seismicity Highlights Tectonic Asperities. Geosciences. 2026; 16(1):38. https://doi.org/10.3390/geosciences16010038

Chicago/Turabian Style

Ostapchuk, Alexey, and Ilmir Nugmanov. 2026. "Background Seismicity Highlights Tectonic Asperities" Geosciences 16, no. 1: 38. https://doi.org/10.3390/geosciences16010038

APA Style

Ostapchuk, A., & Nugmanov, I. (2026). Background Seismicity Highlights Tectonic Asperities. Geosciences, 16(1), 38. https://doi.org/10.3390/geosciences16010038

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop