477
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Unsupervised machine learning and depth clusters of Euler deconvolution of magnetic data: a new approach to imaging geological structures

, , , , , & show all
Received 07 Jul 2023, Accepted 21 Dec 2023, Published online: 04 Jan 2024

Abstract

We present a novel approach that determines the location and dip of geologic structures by clustering Euler deconvolution depth solutions using Density-Based Spatial Clustering Applications with Noise (DBSCAN). This method and workflow rely on the association of changes in the location and relationships between Euler depth clusters and cluster boundaries with changes in rock susceptibility. We applied our method to global magnetic and high-resolution aeromagnetic datasets over Phanerozoic-Precambrian zone-bounding faults in west and central Victoria. The architecture of these structures at different scales from this imaging technique is comparable to interpreted 2D seismic reflection data. The results from the global magnetic data resolved the architecture of these structures below 5 km, while the aeromagnetic data used were limited to structural information of faults above 2 km depth. Therefore, this method shows the structural relationship of the west-dipping Avoca Fault that soles into the east-dipping Moyston Fault at a depth of ∼22 km in central Victoria and at a shallower depth of ∼15 km southward beneath the Quaternary basaltic rocks of the Newer Volcanic Province. In the vicinity of the Heathcote Zone, the method resolves the location, dip, and overprinting relationship between faults and extrusive rocks, such as the relationship between the Heathcote and Mount William Faults and the granitic Cobaw Batholith. We show how combining magnetic data at various scales can track faults from the near-surface to deeper roots while avoiding possible over-interpretation. We demonstrate how to optimise the DBSCAN parameters and a sensitivity analysis of how to determine clusters and cluster boundaries that are geologically relevant in the absence of geological constraints. Our technique provides an effective and rapid tool for imaging structures and can supplement complex and expensive imaging techniques to resolve the architecture of structures in complex geologic terrains.

Introduction

Mapping geological structures below the Earth’s surface has consistently posed challenges for geologists and geophysicists. Traditional methodologies, such as drilling and seismic surveys, often involve high costs and significant time investment and carry the potential for environmental disruption. Advancements in geophysical technology now allow us to survey and image these deep-seated geological formations in a much more efficient and non-invasive manner, revolutionising our approach to exploring and understanding Earth’s subsurface architectures.

The upper crust of numerous continents has been effectively imaged utilising potential field data, primarily through gravity and magnetic datasets (Blaikie et al. Citation2017; Fadel et al. Citation2015; Ganguli and Pal Citation2023; McLean et al. Citation2010; Pedreira et al. Citation2007; Sreejith et al. Citation2013; Williams et al. Citation2009). Collecting and processing potential field data is relatively affordable compared to other methods like seismic reflection transects, which offer detailed crust images but lack wide coverage. The signal in potential field datasets corresponds to the 3D arrangement of subsurface rocks, making it ideal for imaging geological structures (Aitken et al. Citation2009, Citation2013; Armit et al. Citation2014; Blaikie et al. Citation2014; Jessell et al. Citation1993; Le Pape et al. Citation2017). However, inherent ambiguity in the data often necessitates auxiliary a priori information, such as petrophysical or geological information, for refinement.

Machine learning (ML) has recently been applied to resolve ambiguities in interpreting geophysical responses. Recent advances in the Geoscientific application of ML applications include the determination of earthquake positions (DeVries et al. Citation2018; Mousavi and Langston Citation2016), earth system analysis (Reichstein et al. Citation2019), and seismic studies (Hu et al. Citation2019; Huang et al. Citation2006; Jia and Ma Citation2017; Lei et al. Citation2019; Li et al. Citation2018; Liang et al. Citation2014; Malfante et al. Citation2018; Mandelli et al. Citation2018). Although ML approaches aren’t yet broadly applied to gravity and magnetic analysis (Bergen et al. Citation2019; Gupta et al., Citation2022 ; Yu and Ma Citation2021), they hold promise for elucidating geological structures at varying scales. This potential stems from the ability to convert such data into the frequency domain so that the wavelengths of gravity and magnetic anomalies correspond to their source bodies. This capability underpins numerous spectral-based techniques that determine the locations and depths of sources linked to potential field anomalies (Chukwu et al. Citation2018; Ganguli and Pal Citation2023; Kumar et al. Citation2021; Phillips Citation2001; Spector and Grant Citation1970; Teknik and Ghods Citation2017).

The global magnetic data compilation is improved within southeast Australia with addition of air-based, ground and ocean data (Maus et al. Citation2009). These global data compilations capture longer wavelengths from the data, providing images of primary, deep crustal geological structures. Global data compilations are less effective at imaging near-surface geological structures associated with shorter wavelengths. However, when combined with an analysis of high-resolution potential field data that primarily captures shorter wavelength anomalies from the surface to shallow crustal levels allows for tracking near-surface structures to their deeper crustal sources (Guimarães et al. Citation2014). This permits the examination of the magnetic responses across a broad spectrum of wavelengths over an extensive area.

The Euler deconvolution technique is a widely adopted mathematical approach employed to delineate geological structures on multiple scales (Ravat Citation1996; Reid et al. Citation2002; Stavrev and Reid Citation2009). This technique is applied to gravity and magnetic data, to determine the position, depth, and geometry of underlying geological structures causing perturbations in the Earth’s gravitational and magnetic fields. Euler deconvolution is based on Euler’s homogeneity equation (Thompson Citation1982). It estimates source location and depth in its cartesian form, which can approximate the dip of geologic structures such as faults and folds (Amaral Mota et al. Citation2020; De Castro et al. Citation2014). This is due to its ability to determine subtle changes in magnetisation and density without prior knowledge of rock susceptibility and density (1). (1) (xx0)δTδx+(yy0)δTδy+(zz0)δTδz+SI(FR)=0(1) The Euler method uses a deconvolution window size of x − x0 to y − y0 to search the gridded potential field data with local (F), regional components (R), and order of homogeneity (SI, known as the structural index), for reliable depth solutions (z) (Reid et al. Citation2002). The value of SI depends on the geometry of the source. For point or sphere magnetic source, SI = 3, for thick sheet edge, SI = 1, and SI = 0 for contacts (Reid and Thurston Citation2014). The Euler deconvolution method was designed to estimate the location and depths of gravity and magnetic anomalies associated with simple geological structures (Fairhead and Williams Citation2006; FitzGerald et al. Citation2004; Mikhailov et al. Citation2003), homogenous magnetic field (Fedi et al. Citation2015 ) and is well-suited to high-resolution potential field data (Khalil Citation2016; Mohammadzadeh Citation2015; Robert and Cooper Citation2015). The method is less sensitive to noise compared to other techniques, making it more robust in the presence of measurement errors or interference (Reid et al. Citation2002; Stavrev and Reid Citation2009). The method has been used successfully to define geologic contacts and faults (Aziz et al. Citation2013; Cooper and Manzi Citation2015; Khalil Citation2016; Mohammadzadeh Citation2015; Oruç and Selim Citation2011; Ravi Kumar et al. Citation2020; Wang et al. Citation2017), locate igneous intrusions (Shaole et al. Citation2020; Wu et al. Citation2020), map suture zones (Bournas et al. Citation2003; Milano et al. Citation2016), active seismogenic faults (Mazabraud et al. Citation2005; Minelli et al. Citation2016), and identify structures in and beneath sedimentary basins (Anand et al. Citation2009; De Castro et al. Citation2014; Fang et al. Citation2015; Ferraccioli et al. Citation2009; Sridhar et al. Citation2017).

There are some difficulties with the Euler deconvolution technique. For example, it requires correct structural index selection, representing the field’s decay rate with distance from the source (Barbosa et al. Citation1999; de Melo and Barbosa Citation2017; Reid and Thurston Citation2014). Like all potential field analyses, depth solutions have inherent ambiguity, and the Euler deconvolution can produce multiple solutions, leading to several possible locations and depths for subsurface structures, making it difficult to pinpoint the exact position of geological features. The method also assumes that the subsurface geological structures have homogeneous petrophysical properties, which is geologically unreasonable.

In this paper, we present a novel approach that uses Euler deconvolution to determine the depth of magnetic sources, and Density-Based Clustering to identify clusters of irregular shapes and densities to determine the locations and dips of faults over large distances and depths. We examined the performance of this method in imaging crustal faults using global and high-resolution magnetic datasets at different spatial scales in southeast Australia. Results from a 2 arc minute global magnetic grid and aeromagnetic data of 40 m grid spacing were analysed and compared against published interpretations of crustal-scale 2D-seismic and geologic data of western and central Victoria, in southeast Australia. Using Euler results in plan view, we demonstrate how this method can be used alongside other potential geophysical algorithms to constrain the geological interpretation of magnetic data at various scales.

Geological setting

From the Early Cambrian to the Middle Triassic, southeast Australia experienced numerous cycles of basin formation and orogenesis in a convergent tectonic setting along the eastern margin of Gondwana (Cawood Citation2005; Glen and Cooper Citation2021; Gray et al. Citation2006; Gray and Foster Citation2004; Phillips and Offler Citation2011). Tectonism involved lithospheric extension and tectonic mode switching, leading to transient episodes of regional shortening and crustal thickening (Foster and Gray Citation2000; Gray and Foster Citation2004). These rocks are now preserved in the Delamerian and Lachlan orogens (Miller et al. Citation2002; VandenBerg Citation1978; VandenBerg et al. Citation2000).

The Delamerian and Lachlan orogens preserve prolonged sedimentation linked to Neoproterozoic to Cambrian break-up of Rodinia (Merdith et al. Citation2019, Citation2021), west-dipping subduction initiation (Cawood Citation2005; Gray and Foster Citation2004), back-arc basin development (Betts et al. Citation2002; Gray and Foster Citation2004), and transpressional deformation attributed to the ca 464–442sMa Benambran Orogeny and ca 400 Ma Bindian Orogeny (Huston et al. Citation2016; Phillips et al. Citation2012; Wilson et al. Citation2020). Foden et al. (Citation2006) suggested that the Late Cambrian to Early Ordovician Delamerian Orogeny developed in response to west-dipping subduction along the eastern margin of Gondwana.

The early Ordovician to early Carboniferous (ca 485–340 Ma) Lachlan Orogen (Cawood Citation2005; Cayley Citation2011; Kemp et al. Citation2002) is a subduction-accretionary system that includes deep-marine sedimentary rocks, MORB-type tholeiitic volcanic and oceanic boninitic and arc-related volcanic rocks, and an entrained Proterozoic microcontinent (VanDieland: Cayley Citation2011a; Cayley et al. Citation2002; Moore et al. Citation2016; Moresi et al. Citation2014). Ordovician and Silurian arc-related volcanic rocks, continental fragments, and shallower marine sedimentary rocks are mainly exposed in the east, while deep-marine sedimentary rocks are primarily mapped in the west. The western Lachlan Orogen is divided into the Cambrian metasedimentary rocks of the Stawell Zone, Cambrian – Ordovician metasedimentary and mafic volcanic rocks of the Bendigo Zone, and Ordovician to Upper Devonian metasedimentary rocks of the Melbourne Zone (Fergusson et al. Citation2013) that overlies the VanDieland Proterozoic basement (Cayley Citation2011a; Moore et al. Citation2016).

Within the Delamerian and Lachlan orogens are several prominent north-northeast to north-striking crustal-scale faults. Their geometries have been constrained by geological mapping (Edwards et al. Citation1998; Citation2001; Morand et al. Citation1995, Citation2003; VandenBerg Citation1978; VandenBerg et al. Citation2000), potential field analysis (Cayley et al. Citation2002; Gibson et al. Citation2013) and interpretation of regional seismic transects (Cayley Citation2011b; Korsch et al. Citation2002; Willman et al. Citation2010). The key crustal faults delineating the limits of significant tectonic elements within the Delamerian and Lachlan orogens are discussed in the following sections. Our focus is on those faults that partition Proterozoic and Paleozoic rocks, especially the NNE-SSW trending, Late Cambrian Heathcote Fault Zone (Figure ), where we showcase how our approach effectively delineates the structure of these faults.

Figure 1. Area covered by global magnetic data showing a geologic map of structural zones Delamerian and Lachlan Orogens with faults and major zone bounding faults within central and western Victoria (data source: Geoscience Australia and modified in ArcMap after VandenBerg et al. Citation2000). Faults around the northern segment of the Heathcote Fault Zone were modified after Edward et al. Citation2001. The insert Q is where further analysis is performed using aeromagnetic data. Location of profile BB′ across the Newer Volcanic province where the edges of the basaltic extrusives rocks are represented by light grey. The location of previously interpreted seismic lines (1, 2 and 3) and cross-sections of Euler depths (AA′ and BB′) are shown.

Figure 1. Area covered by global magnetic data showing a geologic map of structural zones Delamerian and Lachlan Orogens with faults and major zone bounding faults within central and western Victoria (data source: Geoscience Australia and modified in ArcMap after VandenBerg et al. Citation2000). Faults around the northern segment of the Heathcote Fault Zone were modified after Edward et al. Citation2001. The insert Q is where further analysis is performed using aeromagnetic data. Location of profile BB′ across the Newer Volcanic province where the edges of the basaltic extrusives rocks are represented by light grey. The location of previously interpreted seismic lines (1, 2 and 3) and cross-sections of Euler depths (AA′ and BB′) are shown.

Yarramyljup fault

The Yarramyljup Fault is a regional steeply west-dipping thrust fault separating the Glenelg Zone in the west and the Grampians-Stavely Zone in the east (Morand et al. Citation2003). Cambrian mafic and ultramafic igneous rocks of the Glenelg Zone are exposed in the Yarramyljup Fault’s hanging wall within the Glenelg River Metamorphic Complex. The Cambrian Mount Stavely Volcanic Complex is exposed in the footwall (Morand et al. Citation2003). The fault’s magnetic fabric indicates an NNW-SSE orientation (Figure ). Seismic data suggest it soles into a Proterozoic to Cambrian passive and syn-rift sequences of the Adelaide Rift Complex (Preiss Citation2000) that sit along a detachment that floors the Glenelg and Grampians-Stavely Zones (Cayley et al. Citation2011).

Magnetic data reveal the Yarramyljup Fault extends undercover to the south, where it becomes segmented and overlain by Devonian rocks of the Rocklands Volcanic Group, late Jurassic mafic to intermediate volcanic rocks of the Casterton Formation and the Quaternary basaltic rocks of the Newer Volcanic Province (VandenBerg et al. Citation2000; Welsh et al. Citation2011).

Moyston fault

The Moyston Fault is a major east-dipping crustal boundary that separates sub-greenschist facies Cambrian rocks of the Grampians-Stavely Zone (Delamerian orogen) in its footwall to the west from amphibolite grade Cambrian mafic volcanic and turbidite rocks of the Moornambool Metamorphic Complex in the Stawell Zone (Lachlan orogen) in the hanging wall to the east (Cayley and Taylor Citation2001; Squire et al. Citation2006). Seismic reflection data suggest this zone of high-strain rocks extends to depths greater than 30 km (Willman et al. Citation2010), where it bounds second-order west-dipping antithetic reverse faults that truncate ca 500–490 Ma Cambrian volcanic rocks (Miller et al. Citation2006) and metaturbidites of the Stawell Zone (Cayley et al. Citation2011; VandenBerg et al. Citation2000).

Avoca fault

The west-dipping Avoca Fault generally separates northwest-striking Cambrian metaturbidites in the Stawell Zone from north-striking Ordovician metaturbidites in the Bendigo Zone (Gray and Willman Citation1991; Offler et al. Citation1998; Wilson et al. Citation1992). The fault is defined by a two km-wide zone of poly-deformed Cambrian metaturbidites with small entrained lenses of volcanic rocks (Morand et al. Citation1995). Rocks in the hanging wall record a more complex poly-deformation evolution (Phillips et al. Citation2002) compared with a single regional deformation event in the footwall (Gray and Foster Citation1988; Gray and Foster Citation2004; Morand et al. Citation1995), which is inferred to have occurred around 455–440 Ma (Leader et al. Citation2013; Phillips et al. Citation2002); based on 40Ar-39Ar dating of metamorphic mica within cleavages and during the Benambran Orogeny (Glen et al. Citation2007). Seismic data suggest that the Stawell and Bendigo zones appear as a 160–250 km-wide V-shaped geometry and have been interpreted as imbricates of back-arc inversion and oceanic crust (Cayley and Taylor Citation2001; Crawford et al. Citation2003; Miller et al. Citation2005; Spaggiari et al. Citation2003; Willman et al. Citation2010) (Figure ).

Figure 2. Structural architecture along cross-section AA′ which covers the Stawell, Bendigo and Melbourne Zones. Modified from previous interpretations of seismic lines 1, 2 and 3 covered during 2006 seismic, shows major lithologic units and faults across these structural zones (Cayley et al. Citation2011; Korsch et al. Citation2002).

Figure 2. Structural architecture along cross-section AA′ which covers the Stawell, Bendigo and Melbourne Zones. Modified from previous interpretations of seismic lines 1, 2 and 3 covered during 2006 seismic, shows major lithologic units and faults across these structural zones (Cayley et al. Citation2011; Korsch et al. Citation2002).

Heathcote fault zone

The Heathcote Fault Zone (Figure ) is a major west-dipping crustal boundary with a distinct seismic character that extends to the Moho. It separates the Palaeozoic rocks of the Bendigo Zone in its hanging wall from the Proterozoic rocks of the VanDieland microcontinent and its overlying Ordovician to Devonian sedimentary succession of the Melbourne Zone (Cayley et al. Citation2011; Moore et al. Citation2016; Moresi et al. Citation2014; Willman et al. Citation2010). The fault zone is defined in the west by the Heathcote Fault and to the east by the Mount William Fault (Edward et al. Citation2001; Edwards et al. Citation1998). Movement along the fault resulted in the thrusting of Early Cambrian mafic igneous rocks along the Heathcote Fault Zone during the Benambran Orogeny, with another episode of reactivation during the Middle Devonian Tabberabberan orogeny (Edwards et al. Citation1998). Blueschist “knockers” along the Mount William Fault indicate proximity to a plate edge (Spaggiari et al. Citation2003).

Method

General processing workflow

Our novel approach for determining the dip and strike of geologic structures combines Euler deconvolution with unsupervised machine learning (see Figure for workflow). We initially precondition the data by removing the geomagnetic reference field in preparation for preprocessing. Subsequently, the data resolution is evaluated to determine deconvolution parameters.

Figure 3. The flowchart illustrates the stages involved in applying our method. The pathway in obtaining optimised clusters is shorter where there are geological or geophysical constraints.

Figure 3. The flowchart illustrates the stages involved in applying our method. The pathway in obtaining optimised clusters is shorter where there are geological or geophysical constraints.

Unsupervised machine learning algorithms learn to recognise patterns in data without being explicitly told what those patterns are. Unlike supervised learning, where the algorithm is trained using labelled data, unsupervised learning is used when the data is unstructured, or we do not know what we are looking for. This study uses Density-Based Clustering Application with Noise (DBSCAN), a machine-learning algorithm for clustering data points based on their spatial density. DBSCAN is an unsupervised machine learning algorithm that relies on the reachability and connectivity of data points to establish a cluster. Data points are reachable if they fall within a given minimum distance (or radius of the neighbourhood) and are connected if all possible paths connecting two points are within the minimum radius (Figure a) (Georgoulas et al. Citation2013; Marques and Orger Citation2019; Tran et al. Citation2013). DBSCAN requires the neighbourhood radius, ε, and the minimum number of points, minPts,as inputs (Ester et al. Citation1996; Tran et al. Citation2013).

Other clustering approaches such as partitioning, e.g. K-means and K-center (Gupta et al. Citation2022; MacQueen Citation1967), hierarchical, e.g. Chameleon and HDBSCAN (Campello et al. Citation2013; Johnson Citation1967) and grid-based, e.g. STING and CLIQUE (Amini et al. Citation2011; Duan et al. Citation2012; Yu et al. Citation2015) perform better in datasets with clusters of linear, spherical or convex shapes with similar data densities (Zhou et al. Citation2017). DBSCAN is a density-based clustering algorithm that can identify clusters of arbitrary shapes and variable densities and detect outliers in the dataset (Georgoulas et al. Citation2013; Marques and Orger Citation2019; Tran et al. Citation2013), making it suitable for geoscientific applications. An assumption made by applying this method to the Euler depth solution is that distinct clusters only exist around distinct magnetised lithologic units of a minimum thickness equivalent to the neighbourhood radius ε. Hence, the clusters’ shape is associated with the geometry of geologic structures, while the vertical variations or jumps in clusters are distinct fault planes or fold limbs and hinges. The neighbourhood radius ε can be determined from sources such as outcrop exposures, drill-hole data or other geophysical techniques that provide information about the thickness of a stratigraphic unit or dyke, fault planes, fault core or damaged zone.

Oasis Montaj was used to pre-process the data and execute Euler deconvolution. Then, cross-sections of Euler depth solutions were clustered using the density-based clustering algorithm with noise, DBSCAN, implemented using a bespoke Python code.

Potential field datasets

To demonstrate the effectiveness of our approach in determining the location and dip of geological structures across multiple scales and different geologic terrains, we employ two magnetic datasets of different scales in this study. The Earth Magnetic Anomaly Grid version 2 (EMAG2) (Maus et al. Citation2009), which provides global coverage. We focus on the southern Tasmanides and the onshore region of Victoria. The EMAG2 dataset consists of ship track and airborne magnetic data from various worldwide sources (Maus et al. Citation2009), including data donated by Australia and New Zealand to the National Geophysical Data Center (NGDC), significantly improving the resolution (from 2 to 3 arc minute, ∼3 km) of magnetic anomalies in our study region (Milligan and Franklin Citation2004). The resulting data is 4 km upward continued from the geoid, preserving longer wavelengths of more than 300 km. The processing of this data is detailed in Maus et al. (Citation2009). The upward continuation and grid cell size of EMAG2 effectively suppress noisy short-wavelength signals associated with near-surface anomalies and enhances long-wavelength signals linked to deep geologic structures.

We then conducted reduction-to-pole (RTP) (Figure ) to remove the anomaly asymmetry caused by the inclined Earth’s magnetic field and to locate the anomalies directly above their causative bodies. We initially calculated the RTP on the EMAG2 grid (Figure ) using an inclination of −68.73° and a declination of 11.73°. Although the resulting data is less detailed than high-resolution aeromagnetic or ground gravity data, it imaged first-order zone-bounding structures (Figure ).

Figure 4. The wavelength components of a 4 × 4° grid of the Reduced to Pole of (a) Earth Magnetic Anomaly Grid (EMAG2; Maus et al. Citation2009) and (b) high-resolution aeromagnetic data (HRAM; Geoscience Australia, 2016). (c) Power spectrum plot of (a) and (b) shows three major spectral zones with a strong correlation of both data at long to medium wavelength zones (see Zone A and B).

Figure 4. The wavelength components of a 4 × 4° grid of the Reduced to Pole of (a) Earth Magnetic Anomaly Grid (EMAG2; Maus et al. Citation2009) and (b) high-resolution aeromagnetic data (HRAM; Geoscience Australia, 2016). (c) Power spectrum plot of (a) and (b) shows three major spectral zones with a strong correlation of both data at long to medium wavelength zones (see Zone A and B).

Also, a relatively high-resolution magnetic data from the Australian National Magnetic Grid over an area covered by a major fault zone in central Victoria, the Heathcote Fault Zone (Poudjom et al. Citation2019). These data were re-levelled to a baseline wavelength based on the Australia Wide Airborne Geophysical Survey and had an approximate grid cell size of 40 m (Foss et al. Citation2021). An RTP was calculated using a geomagnetic inclination of −67° and a declination of 110 (Figure ) All data were georeferenced to MGA zone 54.

Depth resolution of EMAG2 in southeast Australia

In order to determine the acceptable Euler solutions, we compared the depth resolution of the EMAG2 data relative to the high-resolution aeromagnetic (HRAM) grid (Sreejith et al. Citation2013; Srinivasa and Radhakrishna Citation2017). By examining the spectral content of 4 × 4° RTP grids of EMAG2 and HRAM (Figure c) at different crustal levels. This can be done because the depth, d, and wavelength, ƛ, could be related by d∼2ƛ (Abraham et al. Citation2015; Chukwu et al. Citation2018; Ganguli et al. Citation2022). Figure c shows two significant crustal boundaries separating zones A, B and C. Zone A shows that EMAG2 and HRAM strongly correlate at long-wavelength magnetic anomalies (ƛ > 50 km) associated with the lower and middle crusts. In Zone B, magnetic anomalies with wavelengths of (2.5 km > ƛ > 50 km) are linked to mid- and upper crustal anomalies. However, in zone B, EMAG2 and HRAM have a strong-to-medium linear correlation. Here, EMAG2, with lower resolution, plots below HRAM. The relative lower power spectrum in EMAG2 indicates that HRAM contains more spectra content (or information).

Zone C has short wavelength (ƛ < 2.5 km) responses and a vague relationship between EMAG2 and HRAM that reflects upper crustal magnetic anomalies. Due to the scattering of EMAG2 spectra in Zone C, it is evident that its resolution is lower than that of HRAM. This lower resolution means it may not resolve geological structures with a wavelength of less than 2.5 km. Because of this limitation, we have chosen to use the upper limit of Zone C, 2.5 km, as a cut-off wavelength that EMAG2 can resolve geologic structures (see Figure ).

Figure 5. (a) Principle and optimisation of parameters used for DBSCAN required to generate clusters. The K-NN graph shows the number of nearest neighbourhood points (NN) across geologically constrained epsilon (ε) is sorted from farthest away from the ε in Group A to closest in Group D. Cross-section of raw Euler depth results from EMAG2 along profile AA′ (b) and BB′ (c) provides no structural /lithologic information. Location of the profiles is shown in Figure .

Figure 5. (a) Principle and optimisation of parameters used for DBSCAN required to generate clusters. The K-NN graph shows the number of nearest neighbourhood points (NN) across geologically constrained epsilon (ε) is sorted from farthest away from the ε in Group A to closest in Group D. Cross-section of raw Euler depth results from EMAG2 along profile AA′ (b) and BB′ (c) provides no structural /lithologic information. Location of the profiles is shown in Figure 1.

Euler deconvolution calculation

To determine the strike and depth of magnetic bodies that illustrate the subsurface structure of the region, the deconvolution window size must be more than half of the expected depth to magnetic source body and large enough to produce reliable and well-clustered depth results (Barbosa et al. Citation1999; Reid et al. Citation2014; Whitehead Citation2012). A deconvolution window size of 40 km was used for the EMAG2 to accurately determine source depth of >20 and 1 km for the aeromagnetic data to determine source depths > 0.5 km. A maximum error in depth locations (δxy) of 40% was maintained for all Euler depth solutions (Barbosa et al. Citation1999; Whitehead Citation2012).

Integer values were used for the structural index to avoid spurious depth solutions SI (Reid et al. Citation2014; Reid and Thurston Citation2014). Non-integer SI values in Euler deconvolution applications imply the degree of homogeneity of the anomalous potential field source changes and are prone to misleading depth solutions (Stavrev and Reid Citation2007). We have chosen 0 and 1 because they are associated with contact and dyke models linked mainly to magnetic contrast from geologic structures such as kilometre-scale lithologic boundaries, faults and shear zones. Only Euler deconvolution using SI = 1 was done for the aeromagnetic dataset, which was required for machine learning as it represents the geometry of geologic structures of interest in this study.

Theory and concept of the application of DBSCAN to Euler depth solutions from EMAG2

DBSCAN, or Density-Based Spatial Clustering of Applications with Noise, is a widely used unsupervised machine learning algorithm. It groups together points in a dataset based on their proximity to one another, following a density criterion. Notably, DBSCAN excels at handling datasets with noise or outliers, as it can detect and disregard these points during the clustering process.

We used DBSCAN to identify relevant clusters within zones of valid and spurious depth solutions to determine the architecture of faults and lithologic boundaries (see Figure ). We applied our method to cross-sections of Euler depth solutions derived from EMAG2 across major zone-bounding faults in southeast Australia. Furthermore, we chose an area where we could test the method, both with and without constraints, against a well-understood 2D reflection seismic survey.

Initially, we clustered Euler depth solutions across profile A-A′ along the previously interpreted deep 2D seismic survey 06GA-V1 to 06GA-V3 across central Victoria (Figures and ) (Cayley et al. Citation2011). Rock exposures and seismic interpretation suggest the Cambrian mafic volcanic units are thinnest at the western edge of the seismic profile and range from ∼2 to 3 km. Based on this constraint, we used a neighbourhood radius, ε, of 2500 ± 200 m (Figure ). This constraint limits the number of infinite clusters and allows clustering to be varied by changes in the minimum number of points (K-NN) alone.

We then used constrained clustering parameters derived from profile A-A′ on profile B-B which is across the Newer Volcanic Province (Figure ). Euler solutions from SI = 1 are used for this stage since they were likely to yield solutions related to two-dimensional structures like faults and dykes (Reid et al. Citation1990; Reid and Thurston Citation2014). Clusters are expected to change drastically with slight changes in the number of clustering points. Hence, we performed clustering with predetermined ε over a wide range of neighbourhood points (minNumPoints and maxNumPoints) with a significant jump in their midpoint when plotted in ak-nearest neighbourhood (k-NN) distance graph (Figure a).

This approach allowed us to determine first-degree clusters. First-degree clusters have cluster boundaries that persist for several changes within the minNumPoints and maxNumPoints. A step-change in the number of points may be necessary for a limited constraint or high uncertainty in the neighbourhood radius, ε. Subsurface structures interpreted as first-degree clusters have the highest degree of confidence. In contrast, we expect more cluster changes from other possible NumPoints and therefore their resulting cluster boundaries are regarded as less reliable. The range of minNumPoints and maxNumPoints that captures a jump and fits an average of our chosen neighbourhood radius for our Euler depth solutions is between 28 and 36 points (see Figure a).

Application of machine learning to Euler depth solutions from the HRAM

To test our approach at the outcrop scale, clustering using unconstrained parameters that are automatically derived from the datasets based on the average distances between data points (Figure ), for a cross-section of Euler depth results extracted from HRAM across the Heathcote Fault Zone in central Victoria. The parameters required for applying DBSCAN, such as the minimum number of points and the neighbourhood radius, ε, were not geologically constrained but obtained by searching for a noticeable jump in the K-NN graph (Figure ). A significant jump is observed in a minimum number of points within the ranges of 12 and 19 (Figure a). All clusters were analysed and compared to the geology of the fault zone.

Figure 6. Unconstrained parameters used for clustering depth results obtained from HRAM. (a) The K-NN graph shows a noticeable jump over an ε of 800 in a window between 12 and 19-the minimum number of points derived from the raw cross-section of Euler depth solutions shown in (b).

Figure 6. Unconstrained parameters used for clustering depth results obtained from HRAM. (a) The K-NN graph shows a noticeable jump over an ε of 800 in a window between 12 and 19-the minimum number of points derived from the raw cross-section of Euler depth solutions shown in (b).

Results

Interpretation of Euler deconvolution (ED) depth results: EMAG2 and HRAM

The results of the Euler depth solutions obtained from the EMAG2 and HRAM datasets provides insight into the geology of fault networks at multiple scales. By examining the depth variations across the two magnetic datasets at different scales, we identify areas where the locations of clusters of depth solutions deviate from major geologic faults.

Euler depth results from EMAG2

In map view, Euler results of structural index SI = 0 yield depth range from 4 to 25 km and cluster along strike of previously known extensive zone bounding faults in southeast Australia (see Figures & ). Similarly, Euler depth results from SI = 1 in map view cluster along strike and on major faults separating the structural zones, providing two-dimensional geometric information for these faults. Euler depth results from SI = 1 (Figure c) are more numerous than depth solutions from SI = 0 (Figure a) and can be used to determine the strike and dip of the faults, providing additional information to identify these faults at depth (see Figure c and d).

Figure 7. Plan view of Euler depth solutions derived from EMAG2 on the topographic map and their correlation along major zone bounding faults. (a) Euler depth solutions from SI = 0 and their interpretation in (b). (c) Euler depth solutions from SI = 1 and its corresponding interpretation in (d). Euler depth solutions in both cases appear to cluster more along the hangingwall of the major zone-bounding faults.

Figure 7. Plan view of Euler depth solutions derived from EMAG2 on the topographic map and their correlation along major zone bounding faults. (a) Euler depth solutions from SI = 0 and their interpretation in (b). (c) Euler depth solutions from SI = 1 and its corresponding interpretation in (d). Euler depth solutions in both cases appear to cluster more along the hangingwall of the major zone-bounding faults.

Interpretation of ED depth results from SI = 0

Euler solutions from SI = 0 provide almost no values within areas of strong magnetic contrast. However, they provide depth values correlating with major faults in areas with little interfering magnetic contrast, such as along the west-dipping Avoca Fault and the northern segment of the east-dipping Moyston Fault (see Figure a and b). To the north of the Newer Volcanics Province, linear NNW-SSE depth clusters are aligned with a short segment of the west-dipping Yarramyljup Fault but do not give depth results across the Newer Volcanic Province (Figure b and d). It can be difficult to establish the dip of the faults using depth clusters in a plan view.

Interpretation Euler deconvolution depth results from SI = 1

Shallower depth solutions (from <5100 to 17,500 m) from SI = 1 are present west of the Mount William Fault. Depths greater than 17,500 m east of the Mount William Fault correspond to the thicker and older crustal unit beneath the Melbourne Zone (Cayley et al. Citation2002; Edwards et al. Citation1998) (Figure d). The transition of deeper to shallower depths from east to west continues southward and connects with the Bambra Fault with both shallow and deeper depths (Figure d). Deeper clusters along the Selwyn Fault suggest that this fault lies within the relatively thick crust of VanDieland (Berry et al. Citation2005; Moore et al. Citation2015).

Depth clusters in the Newer Volcanic Province generally show the orientation of depth clusters from SI = 1 that trend east–west. These Euler depth solutions show the orientations of the young basaltic rocks with much higher magnetic susceptibilities than the basement rocks. Thus, within this zone, the orientation of the depth solutions does not reflect the NNW, NNE and N-S trends of the major faults.

Euler depth results from HRAM

The results of using the high-resolution aeromagnetic data for Euler deconvolution with SI = 1 show clear linear clusters of depth solutions around several faults close to the Heathcote Fault Zone (Figure ). These clusters are aligned with zones of abrupt changes in the magnetic signatures around the major faults. The depth solutions obtained through Euler deconvolution range from 100 meters to 2 kilometres and correlate with different geophysical responses associated with the various geological structures around the fault zone (Slater and Haydon Citation1999).

Figure 8. Correlation of Euler depth solutions generated from RTP of the aeromagnetic data using SI = 1. (a) shows the data with no annotation, while (b) shows the locations of the major faults in the vicinity of the Heathcote Fault Zone (locations in Figure ). Concentric clusters of relatively deep depth values correlate with mappable late Devonian granites G1, G2, G3 and G4 (see Edwards et al. Citation2001). Profile C-C′ shows where the cross-section of the Euler depth from the aeromagnetic data is located (Figure ).

Figure 8. Correlation of Euler depth solutions generated from RTP of the aeromagnetic data using SI = 1. (a) shows the data with no annotation, while (b) shows the locations of the major faults in the vicinity of the Heathcote Fault Zone (locations in Figure 1). Concentric clusters of relatively deep depth values correlate with mappable late Devonian granites G1, G2, G3 and G4 (see Edwards et al. Citation2001). Profile C-C′ shows where the cross-section of the Euler depth from the aeromagnetic data is located (Figure 9).

Figure 9. DBSCAN clustering of cross-section AA′ using ε = 2500 m and individual K-NN values and grouped according to their mirror distance from the inflexion point on the K-NN graph in Figure a. (a) End-member cluster assemblages farthest and (D) is the closest to the inflexion point or optimal cluster assemblages. The distinct cluster assemblages are identified by their colours. The n on each cluster assemblage for individual K-NN values indicates the number of clusters or cluster labels. All depth solutions are projected to GDA94/MGA 54.

Figure 9. DBSCAN clustering of cross-section AA′ using ε = 2500 m and individual K-NN values and grouped according to their mirror distance from the inflexion point on the K-NN graph in Figure 5a. (a) End-member cluster assemblages farthest and (D) is the closest to the inflexion point or optimal cluster assemblages. The distinct cluster assemblages are identified by their colours. The n on each cluster assemblage for individual K-NN values indicates the number of clusters or cluster labels. All depth solutions are projected to GDA94/MGA 54.

The NNE and NNW trending linear clusters runs along the major faults, including faults within the Heathcote Fault Zone (Figure ). In plan view, linear clustered depth solutions align within the hangingwall of the Heathcote and the Mount William Faults. Linear clusters of depth values from 200-750 m are aligned with the hanging wall of the west-dipping Mount William Fault, the eastern-most fault in the zone that separates Ordovician to Devonian sedimentary rocks of the Melbourne Zone in the east from Cambrian to Ordovician turbidite rocks of the Bendigo Zone to the west (Edwards et al. Citation1998; Citation2001). Linear depth clusters are also aligned along the west-dipping Heathcote Fault, the westernmost fault segment in the fault zone in this area (Figures and b). Other NNW and NNE linear depth clusters lie along the west-dipping Meadow Valley Fault and west-dipping Fosterville Fault to the west of the Heathcote Fault Zone (Figure b).

Concentric clusters with relatively deeper solutions at their centres lie around major granite intrusions. Depth range from 750 to 2000 m southwest and southeast of the Heathcote Fault Zone (G1 and G2 in Figure b). Another concentric cluster lies between linear along the Meadow Valley Fault and the Heathcote Fault Zone (G3 in Figure b). All these concentric depth clusters are located within the vicinity of the east–west oriented late Devonian granite known as the Cobaw Batholith, which is bounded by linear Euler depth clusters associated with major faults near the Heathcote Fault Zone (Clemens et al. Citation2016; VandenBerg et al. Citation2000).

Analysis of geologically constrained cluster assemblages

In Figure b and c, unclustered cross-sections of Euler depth solutions are shown along profiles A-A′ and B-B′. The top of the Euler deconvolution cross-section along profile A-A′ represents an undulating surface minimum depth of approximately 5 km. The lower part of the cross-section is less coherent at a depth of 15 km at about 200 km from the eastern edge of the traverse (Figure b). The unclustered transverse (Figures b and c) provides no structural or lithologic information as these key geological boundaries cannot be identified from the section.

To perform the sensitivity to the number of clusters and cluster boundaries due to step changes in the minimum number of points (minPts) and epsilon distance (ε), of 2500 ± 200 m (see Figure a). We categorized the cluster assemblages based on the proximity of their minPts to the inflection point on the K-NN graph (see Figure a), beginning with the end-members (Figure a) and proceeding to the assemblages nearest to the inflection point (Figure d).

End-member clusters across profile AA′

The end-member clusters that are the least optimised are those produced using the 28-NN and 36-NN (Figure a). The consistent cluster boundaries in these least optimised end-member clusters provide a means to determine cluster boundaries that will persist in all cluster assemblages. These consistent cluster boundaries are regarded as first-level cluster boundaries and are most likely to be associated with geological structures.

The end-member group have the highest variability in the number of clusters, as the 28-NN and 36-NN produced 6 and 12 clusters, respectively, yet share cluster boundaries with similar geometries (Figure a). Both cluster assemblages show cluster boundaries that runs from the western edge at a depth of 10 km and dips east down to 30 km at easting 675,000 m. Another cluster boundary in both cluster assemblages dips west from easting 860,000–790,000 m (Figure a). The boundaries in both end-member cluster assemblages are few and are first-level cluster boundaries. The first-level cluster boundaries are likely to persist in all cluster assemblages as we converge towards the inflexion point and so are considered the boundaries that are most likely to reflect faults or lithological boundaries.

Significant differences between the cluster assemblages are the number of clusters and the geometry of the resulting cluster boundaries. The 36-NN cluster assemblage has the highest number of clusters associated with the top, resulting in cluster boundaries with varying geometries. Within the 36-NN cluster assemblage, at least six major cluster boundaries extends to the top. Two clusters exist at easting of 830,000 m from depth of 5–15 km forming a west-dipping cluster boundary. A cluster boundary at easting of 670,000 m that dips west extends from depths of 10 km to depth of 20 km. In contrast, the cluster assemblage derived with 28-NN has fewer clusters and boundaries within the traverse. A cluster boundary exists within 28-NN cluster assemblage extending from the western fringe of profile A-A′ and terminating at the easting of 860,000 m (Figure a).

Other clusters across profile A-A′

Different cluster parameters impact properties of resulting cluster boundaries, allowing us to identify how geological constraints result in the most optimal parameter values that produce the most meaningful and useful clustering. We analyse other cluster assemblage pairs (Figure b–d) with corresponding parameters as they approach the inflexion point for clustering done with useful geological constraints. Variations in the number and geometry of other clusters reflect how these the geometry and number of cluster boundaries change their K-NN values approaches the inflexion point. Generally, the number of clusters tends to increase relative to the number of neighbourhood points at lower k-NN values before the inflexion point. The number of clusters becomes constant at higher numbers of neighbourhood points, usually after the inflexion point.

For clusters before the inflexion point, such as cluster assemblages, 29-NN and 30-NN produced the same number, but 31-NN has a higher number (12) of clusters. There is a large increase in the number of clusters produced after the inflexion point, as 33-NN and 34-NN give 13 clusters. Clusters continue to increase until a large drop at cluster assemblage 35-NN, which gives 11 clusters. This drastic fall in the number of clusters, after the inflexion point, suggests over-optimisation and the likely production of unreasonable clusters. The domain where the quantity of clusters stays the same or marginally increases as K-NN values increase past the inflexion point is deemed the domain with optimised clusters (e.g. cluster assemblages of 32-NN, 33-NN and 34-NN). It’s important to note the changes in the number of clusters directly affects the resulting geometry and number cluster boundaries formed within each cluster assemblage. Eliminating cluster assemblages based on their proximity to the inflexion point can be used to identify optimised clusters in absence of constraints.

The K-NN distance graph (Figure a) shows that the clustering pairs 30-NN and 32-NN, and 31-NN and 33-NN, are the closest to the inflexion point. This means that these pairs are most likely to produce optimised clusters. They demonstrate cluster boundaries with geometries and depths to those from previous clusters, including first-order clusters. The 31-NN and 33-NN pairs are the most optimised clusters and reflect the geological complexity across the traverse. While other clustering assemblages show only two major west-dipping clusters within eastings of 800,000 and 820,000 m, the 31-NN and 33-NN pairs have more varied cluster boundaries at different depths. Only these clustering pairs show similar west-dipping boundaries at 675,000 m.

Optimised parameters: clusters across B-B′

We apply clustering parameters of 33 points and an epsilon (ε) of 2500 m from the A-A traverse to Euler depths in cross-section B-B′ (Figure ). Section B-B′ differs from section A-A′ because the orientations of the Euler depth solutions from the highly magnetic basaltic cover rocks rather than the basement geology. Nevertheless, 10 distinct cluster assemblages were generated at varying depths.

Figure 10. Euler depth solutions across profile B-B′ clustered according to optimised parameters derived from profile A-A′. Also, note the short path to achieve geologically relevant clusters along B-B′ (see Figure ).

Figure 10. Euler depth solutions across profile B-B′ clustered according to optimised parameters derived from profile A-A′. Also, note the short path to achieve geologically relevant clusters along B-B′ (see Figure 3).

These cluster assemblages show generally similar geometries to those found in the optimised clusters for the A-A′ traverse. Notably, there is a similar east-dipping cluster boundary at the western edge starting at 5 km depth and extending to a depth of 30 km at easting 750,000 m (Figure ). This boundary demarcates some minor west-dipping cluster boundaries with the least number of clustering points.

A major west-dipping boundary at the western edge extends from a shallow iso-surface to depths of ∼30 km along the traverse at easting 775,000 m. Below this boundary, depth solutions are absent at greater depths. This major west-dipping cluster boundary also bounds most of the narrow east-dipping boundaries in the eastern half of the traverse. Another steep west-dipping cluster boundary exists at easting 725,000 m with shallow iso-surface. This boundary terminates at the major east-dipping boundary at a depth of ∼15 km.

Analysis of unconstrained cluster

Using constrained parameters allows us to achieve optimal clustering that quickly outlines the geological complexity, as there are an infinite number of potential clusters and cluster boundaries that could be formed. We assess the effectiveness of our method when the clustering parameters are unconstrained but are instead identified by a notable increase across an average epsilon within a range of K-NN values. Here, the epsilon is generated directly from the average distance of each Euler depth value from the other. A significant jump is observed in the K-NN window ranging from minPts of 12–19.

In general, an increase in the number of clusters aligns with an increase in K-NN values, except for outliers. As the number of nearest-neighbour distance assemblages increases, the number of clusters also increases. The number of minPts increases relative to clusters and the resulting cluster boundaries. However, an outlier appears at a minPts of 13 and 19, which shows an unusually high 27 clusters (Figure b and h). The outlier is a cluster assemblage that does not follow the general trend or the expected number of clusters. An outlier assemblage can also be identified when the number of clusters before the inflexion point is higher than that after the inflexion point. However, the outlier cluster assemblage should show the general characteristics of clusters and cluster boundaries identified in cluster assemblages derived from an adjacent number of neighbourhood points on the K-NN graph.

Figure 11. DBSCAN clustering of Euler depth solutions derived from HRAM across profile C-C′, the Heathcote Fault Zone, for each cluster assemblage using ε = 800 m and K-NN values, from 12 to 19, as shown in Figure . Clustering shown in (a) and (g) are end-member cluster assemblages. Cluste assemblages (b) and (h) are regarded as outliers. (d) and (e) are regarded as the optimised cluster assemblages. All depth solutions are projected to GDA94/MGA 54.

Figure 11. DBSCAN clustering of Euler depth solutions derived from HRAM across profile C-C′, the Heathcote Fault Zone, for each cluster assemblage using ε = 800 m and K-NN values, from 12 to 19, as shown in Figure 6. Clustering shown in (a) and (g) are end-member cluster assemblages. Cluste assemblages (b) and (h) are regarded as outliers. (d) and (e) are regarded as the optimised cluster assemblages. All depth solutions are projected to GDA94/MGA 54.

Most cluster boundaries in the end-member assemblages produced complex clusters with varied depths and geometries (Figure a and g). Consistent cluster boundaries across all cluster assemblages, arising from clusters with similar geometries, locations, and depths. These are considered first-level cluster boundaries and reflect the geology at depth. A notable difference between the cluster boundaries of end-member cluster assemblages is that minPts = 12 (Figure a) shows a single cluster between easting 830,000 and 850,000 m while minPts = 19 shows multiple clusters (Figure  h). As a result, no cluster boundary can be identified in minPts = 12 cluster assemblage while minPts = 19 cluster assemblage shows several cluster boundaries from easting 830,000 and 850,000 m. In contrast, similar or near-identical clusters occur using minPts of 17 and 18 despite the difference in the generated number of clusters (Figure f and g).

Cluster assemblages closest to the inflexion point typically show a maximum positive difference in the number of clusters. The two cluster assemblages below and above the inflexion point are minPts = 15, which gives 25 clusters and minPts = 16 with 27 clusters (Figure d and e). Though these clusters show differences in their number of clusters, they are the most optimised as they are closest to the inflexion point. Generally, they share similarities in the geometry of major cluster boundary are considered cluster assemblages that best reflect the complexity along the traverse.

Implications of cluster assemblages

Optimum clusters across A-A′ and B-B′

We initially compared the optimised cluster assemblage on profile A-A′ with the near-parallel previously interpreted 2D reflection seismic survey 06GA-V1 to 06GA-V3 and then used the knowledge gained to examine other optimised Euler depth cross-sections (Figures , d & ). For traverse A-A′, the optimised cluster assemblages were produced using a neighbourhood radius epsilon (ε) of 2500 ± 200 m and 33-NN (Figure ).

Figure 12. Optimised clustered Euler depth solutions of the cross-section along profile A-A′. Cluster boundaries strongly correlate with the locations of the major zone-bounding faults at depth. The first-order cluster boundaries are in red and, other cluster boundaries, in black.

Figure 12. Optimised clustered Euler depth solutions of the cross-section along profile A-A′. Cluster boundaries strongly correlate with the locations of the major zone-bounding faults at depth. The first-order cluster boundaries are in red and, other cluster boundaries, in black.

Figure 13. Crustal architecture beneath the Newer Volcanic Province derived from interpreted clustered Euler depth solutions along the B-B′ traverse. (a)Uninterpreted cluster assemblage (b) Interpreted cluster assemblages.

Figure 13. Crustal architecture beneath the Newer Volcanic Province derived from interpreted clustered Euler depth solutions along the B-B′ traverse. (a)Uninterpreted cluster assemblage (b) Interpreted cluster assemblages.

Comparing the optimised cluster assemblage with the pre-interpreted seismic line reveals that the cluster boundaries align with major boundaries along the seismic line associated with Proterozoic to Paleozoic faults (Cayley et al. Citation2011; Willman et al. Citation2010). Our interpretation of these clusters is also constrained by the depth solutions clustering along these faults in plan view (Figure ). The geometries and locations of the major faults interpreted in previous studies correspond with most of the first-level cluster boundaries in the end-member and other cluster assemblages. Furthermore, cluster boundaries retain the same cross-cutting relationships and preserve the structures’ geological history.

The steeply east-dipping cluster boundary that extends from the western edge to a depth of 30 km at easting 675,000 m correlates with the Moyston Fault (see Figures and ). Other first-level cluster boundaries, such as the west-dipping cluster boundaries at 790,000 and 860,000 mE, correlate with the Whitelaw and Mount William Faults. One significant west-dipping boundary at 700,000 mE terminates at a depth of ∼22 km on the east-dipping cluster boundary associated with the Moyston Fault. This cluster boundary aligns with the Avoca Fault. Other cluster boundaries correlate with less significant faults, such as west-dipping cluster boundaries at easting of 695,000, 710,000, and 760,000 m, which align with imbricated west-dipping faults identified in the seismic section (Cayley et al. Citation2011; Willman et al. Citation2010).

Using the optimised cluster parameters from profile A-A′ for the cluster assemblage on B-B′, we observed a strong correlation of geometry, location, and overprinting relationships of cluster boundaries and geologic structures despite the masking effect of highly magnetic basalts of the Newer Volcanic Province (Figure ). In the cluster assemblage along profile B-B′, the east-dipping boundary associated with the Moyston Fault truncates the west-dipping boundary associated with the Avoca Fault at a depth of ∼15 km, which is a shallower depth of than that observed in profile A-A′. The cluster boundary that coincides with the Mount William Fault continues to deeper levels in traverse B-B′ compared with profile A-A′. In profile B-B′ there are several minor east and west-dipping cluster boundaries that are not evident in profile A-A′, suggesting a subtly different structural pattern in the south. The distance between the Moyston and the Mount William Fault is greatly reduced in profile B-B′, suggesting they converge to the south (see Rawling et al. Citation2011).

Figure 14. Geologic interpretation of optimised clustering of Euler depth solutions generated from higher resolution airborne magnetic data along profile C-C′ (Figure ). Depth clusters denoted as G1, G2 and G4 are located beneath concentric clusters in plan view (see Figure ) associated with exposed granitic rocks.

Figure 14. Geologic interpretation of optimised clustering of Euler depth solutions generated from higher resolution airborne magnetic data along profile C-C′ (Figure 8). Depth clusters denoted as G1, G2 and G4 are located beneath concentric clusters in plan view (see Figure 8) associated with exposed granitic rocks.

Optimum cluster across the Heathcote Fault Zone

We apply our method across the Heathcote Fault Zone using unconstrained clustering parameters by allowing clustering parameters to be automatically determined based on a significant jump of K-NN values over an average epsilon distance (Figure ). For optimised cluster assemblages, we discuss the geologic significance and how cluster assemblages provide insight into the geometry of faults at depth. We correlate cluster interpretations with the location of depth clusters in plan view.

The boundaries of the optimised cluster assemblage along profile C-C′ strongly correlate with the known structures in and around the Heathcote Fault Zone (Figures and ). A major west-dipping cluster boundary correlates with the Mount William Fault (Figure ). This west-dipping boundary appears to truncate an east-dipping boundary west of Mount William Fault at a depth of 300 m. Along the Heathcote Fault (Figure ), there is a west-dipping cluster boundary that terminates on an unnamed east-dipping boundary at 1 km. The unnamed east-dipping boundary continues deeper to 1.8 km, where it terminates on a less evident west-dipping boundary at easting 830,000 m that appears to continue deeper as marked by dashed lines.

NNE trending linear clusters in the map view along the Meadow Valley Fault (Figure ), along the B-B′ traverse, correspond to a west-dipping cluster boundary extends to deeper depths (Figure and ). Less prominent west-dipping cluster boundaries at 830,000 mE suggests the presence of a west-dipping faults at depth of 2 km. At the western edge of the traverse is a west-dipping cluster boundary that lies along the Fosterville Fault (Figures and ).

Concentric clusters, evident in plan-view, align with known granitic rock exposures (Slater and Haydon Citation1999) and are bounded by major basement faults. East of the Mount William Fault, along profile C-C′, subcircular Euler depth cluster G1 (Figure ) coincides with basaltic outcrops. Similar elliptical clusters (G2, G3 and G4) occur between the Heathcote Fault and the Fosterville Fault. The cluster boundaries associated with major faults extend deeper than the concentric clusters (Figure ).

Discussion and conclusions

Effect of data resolution and method limitations

Data resolution is a critical factor in how combines unsupervised machine learning with Euler deconvolution solutions performs, both in determining the architecture of structures at depth and for other geophysical techniques. The method uses the capacity of the deconvolution process to distinguish subtle magnetic changes resulting from geological differences despite the homogeneous nature of the conventional Euler deconvolution equation used in this study. The number of clusters and resulting cluster boundaries are directly limited by the inherent trade-off that exist between data resolution and depth of investigation determined by the choice of the window size used during the deconvolution process.

For example, the global magnetic data grid with 2 arc-minute spacing and window size of 40 km can be used to image deep-crustal geologic structures, such as kilometre-wide fault damage zones and shear zones. In contrast, aeromagnetic data with a 40 m cell size gives depth results along individual faults within the Heathcote Fault Zone (Figures & ). Clustering of deconvolution depth results can determine the architecture of faults at metre scales and their dip at depths less than 100 m. However, this window size provides no structural information beyond 2 km due to Euler depth results’ limitations. The strong correlation of the DBSCAN clustering results with pre-interpreted 2D reflection seismic data (see Figure ), supports the effectiveness our technique even in presences of unaccounted inhomogeneity in the magnetic field during the Euler deconvolution process. Nevertheless, DBSCAN clusters of depth solutions from the Euler deconvolution that considers the inhomogeneity in the magnetic field will yield more clusters that represent geological complexities (Fedi et al. Citation2015).

The method works best when DBSCAN’s parameters are constrained, particularly the minimum radius of the cluster (see Figures and ). Using previously known constraints can quickly generate optimised clusters representing geological complexities along the traverse. However, if the minimum radius is smaller than the data resolution, the clustering process yields clusters and cluster boundary artefacts that do not reflect geological structures. Also, while most cluster boundaries determined using optimum parameters correlate with known geologic structures, not all cluster boundaries represent structures, and over-interpretation is possible. Cluster boundaries should first be interpreted from coarser geological features, and only after that should finer-scale cluster boundaries from parameters nearer to the end-members on the K-NN graph be used.

Potential applications

Our new technique demonstrates extensive potential in a variety of applications. As a swift and reliable method, it helps resolve the dip and strike of both deep and shallow geologic structures. The method can help constrain well-known, more complex and expensive imaging techniques such as seismic refraction and reflection, potential field modelling, and geophysical inversion. It provides critical information on the first 500 m depth from high-resolution magnetic data aiding the interpretation of other near-surface techniques like electromagnetic and electrical tomographic methods (Boaga et al. Citation2020; Romero-Ruiz et al. Citation2018) and multichannel analysis of surface waves (Loo and Leong Citation2018; Pelton Citation2005) also deeper crustal depth imaging techniques such as seismic reflection (Peron-Pinvidic et al. Citation2022; Stephenson et al. Citation2021), refraction (Mooney Citation2015; Stixrude and Jeanloz Citation2015) and ambient noise (Bem et al. Citation2020; O’Donnell et al. Citation2023; Zeng et al. Citation2021).

Imaging of crustal structures under igneous or volcanic provinces has long been a challenge (Ernst et al. Citation2005). The strong attenuation properties of the overlying volcanic rocks make these structures difficult to identify in seismic sections (Grijalva et al. Citation2018; Holt et al. Citation2013; Song et al. Citation2018; Yule and Spandler Citation2022). The high magnetic susceptibilities of many volcanic rocks also mask the magnetic signatures of underlying structures and rocks (Blaikie et al. Citation2014; Lillis et al. Citation2009; Raimi et al. Citation2014). However, our technique can accurately determine the locations and dips of faults in a basaltic province in southeast Australia. This has implications for not only better understanding the basic structures in these areas but also in their relationship to the volcanic plumbing systems and their ascent pathways (Ernst et al. Citation2005; Guardo et al. Citation2022). These structures help in shaping continental margins that are often sites of mafic volcanic or thick sedimentary sequences (Anudu et al. Citation2014; Bladon et al. Citation2015; Gibson et al. Citation2013; Samsu et al. Citation2021) and potentially, their links with orebodies (Blundell et al. Citation2005; Boyce et al. Citation2014; Diakov et al. Citation2002; Graham et al. Citation2017; Ji et al. Citation2023; Pinotti et al. Citation2016).

The potential to extend our approach to resolve the architecture of geologic structures at depth in 3D space represents another important but challenging area that requires attention in the future. Having resolved the location and dip of geologic structures at various scales along profiles in 2D which can improve sub-surface models from the same data, the current approach should be able to resolve the architecture of structures in 3D space if the parameters of the DBSCAN algorithm are optimised. Clustering and/or segmentation of points in 3D or multi-space has largely lagged behind 2D machine learning algorithms but has huge promise in finding relationships in Euler depth solutions that are geologically meaningful (Ahmed and Chew Citation2020; Liu et al. Citation2023).

Acknowledgement

The authors thank Geoscience Australia and Geological Survey Victoria for providing and maintaining publicly available high-resolution magnetic data. This study was supported by the Australian Society of Exploration Geophysics Foundation Grant (RF21P01) and Monash-IITB scholarship awarded to Chukwu. We also thank the editor, Mark Lackie, the associate editor, Sam Matthews, and three additional anonymous reviewers for their constructive comments that have helped improved the manuscript.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been corrected with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

This work was supported by Australian Society of Exploration Geophysics Foundation Grant [Grant Number RF21P01]; Monash-IITB scholarship awarded.

References

  • Abraham, E.M., E.G. Obande, M. Chukwu, C.G. Chukwu, and M.R. Onwe. 2015. Estimating depth to the bottom of magnetic sources at Wikki Warm Spring region, northeastern Nigeria, using fractal distribution of sources approach. Turkish Journal of Earth Sciences 24 no. 5: 494–512. doi:10.3906/yer-1407-12.
  • Ahmed, S.M., and C.M. Chew. 2020. Density-based clustering for 3D object detection in point clouds. Proceedings of the IEEE Computer Society Conference on Computer Vision and Pattern Recognition, 10605–10614. doi:10.1109/CVPR42600.2020.01062.
  • Aitken, A.R.A., P.G. Betts, R.F. Weinberg, and D. Gray. 2009. Constrained potential field modeling of the crustal architecture of the Musgrave Province in central Australia: Evidence for lithospheric strengthening due to crust-mantle boundary uplift. Journal of Geophysical Research: Solid Earth 114 no. 12. doi:10.1029/2008JB006194.
  • Aitken, A.R.A., M.L. Salmon, and B.L.N. Kennett. 2013. Australia’s Moho: A test of the usefulness of gravity modelling for the determination of Moho depth. Tectonophysics 609: 468–479. doi:10.1016/j.tecto.2012.06.049.
  • Amaral Mota, E.S. W.E. Medeiros, and R.G. Oliveira. 2020. Can Euler deconvolution outline three-dimensional magnetic sources? Geophysical Prospecting 68 no. 7: 2271–2291. doi:10.1111/1365-2478.12988.
  • Amini, A., T.Y. Wah, M.R. Saybani, and S.R.A.S. Yazdi. 2011. A study of density-grid based clustering algorithms on data streams. Proceedings - 2011 8th International Conference on Fuzzy Systems and Knowledge Discovery, FSKD 2011 3: 1652–1656. doi:10.1109/FSKD.2011.6019867.
  • Anand, S.P., M. Rajaram, T.J. Majumdar, and R. Bhattacharyya. 2009. Structure and tectonics of 85°E Ridge from analysis of geopotential data. Tectonophysics 478 no. 1–2: 100–110. doi:10.1016/j.tecto.2008.09.036.
  • Anudu, G.K., R.A. Stephenson, and D.I.M. Macdonald. 2014. Using high-resolution aeromagnetic data to recognise and map intra-sedimentary volcanic rocks and geological structures across the Cretaceous middle Benue Trough, Nigeria. Journal of African Earth Sciences 99: 625–636. doi:10.1016/j.jafrearsci.2014.02.017.
  • Armit, R.J., L. Ailleres, P.G. Betts, B.F. Schaefer, and T.N. Blaikie. 2014. High-heat geodynamic setting during the Palaeozoic evolution of the Mount Painter Province, SA, Australia: Evidence from combined field structural geology and potential-field inversions. Geophysical Journal International 199 no. 1: 253–275. doi:10.1093/gji/ggu263.
  • Aziz, A.M., W.A. Sauck, E.A.H. Shendi, M.A. Rashed, and M. Abd El-Maksoud. 2013. Application of analytic signal and Euler deconvolution in archaeo-magnetic prospection for buried ruins at the ancient city of Pelusium, NW Sinai, Egypt: A case study. Surveys in Geophysics 34 no. 4: 395–411. doi:10.1007/s10712-013-9229-z.
  • Barbosa, V.C.F., J.B.C. Silva, and W.E. Medeiros. 1999. Stability analysis and improvement of structural index estimation in Euler deconvolution. Geophysics 64 no. 1: 48–60. doi:10.1190/1.1444529.
  • Bem, T.S., H. Yao, S. Luo, Y. Yang, X. Wang, X. Wang, L. Li, and B. Liu. 2020. High-resolution 3-D crustal shear-wave velocity model reveals structural and seismicity segmentation of the central-southern Tanlu Fault zone, eastern China. Tectonophysics 778 no. August 2019: 228372. doi:10.1016/j.tecto.2020.228372.
  • Bergen, K. J., P. A. Johnson, M. V. De Hoop, and G. C. Beroza. 2019. Machine learning for data-driven discovery in solid earth geoscience. In Science, Vol. 363, Issue 6433. American Association for the Advancement of Science.
  • Berry, R.F., O.H. Holm, and D.A. Steele. 2005. Chemical U-Th-Pb monazite dating and the Proterozoic history of King Island, southeast Australia. Australian Journal of Earth Sciences 52 no. 3: 461–471. doi:10.1080/08120090500134514.
  • Betts, P.G., D. Giles, G.S. Lister, and L.R. Frick. 2002. Evolution of the Australian lithosphere. Australian Journal of Earth Sciences 49 no. 4: 661–695. doi:10.1046/j.1440-0952.2002.00948.x.
  • Bladon, A.J., S.M. Clarke, and S.D. Burley. 2015. Complex rift geometries resulting from inheritance of pre-existing structures: Insights and regional implications from the Barmer Basin rift. Journal of Structural Geology 71: 136–154. doi:10.1016/j.jsg.2014.09.017.
  • Blaikie, T.N., L. Ailleres, P.G. Betts, and R.A.F. Cas. 2014. Interpreting subsurface volcanic structures using geologically constrained 3-D gravity inversions: Examples of maar-diatremes, Newer Volcanics Province, southeastern Australia. Journal of Geophysical Research: Solid Earth 119 no. 4: 3857–3878. doi:10.1002/2013JB010751.
  • Blaikie, T.N., P.G. Betts, R.J. Armit, and L. Ailleres. 2017. The ca. 1740–1710 Ma Leichhardt Event: Inversion of a continental rift and revision of the tectonic evolution of the North Australian Craton. Precambrian Research 292: 75–92. doi:10.1016/j.precamres.2017.02.003.
  • Blundell, D., N. Arndt, P.R. Cobbold, and C. Heinrich. 2005. 9: processes of tectonism, magmatism and mineralization: Lessons from Europe. Ore Geology Reviews 27 no. 1–4: 333–349. doi:10.1016/j.oregeorev.2005.07.003.
  • Boaga, J., M. Phillips, J. Noetzli, A. Haberkorn, R. Kenner, and A. Bast. 2020. A comparison of frequency domain electro-magnetometry, electrical resistivity tomography and borehole temperatures to assess the presence of ice in a rock glacier. Frontiers in Earth Science 8 December: 1–11. doi:10.3389/feart.2020.586430.
  • Bournas, N., A. Galdeano, M. Hamoudi, and H. Baker. 2003. Interpretation of the aeromagnetic map of Eastern Hoggar (Algeria) using the Euler deconvolution, analytic signal and local wavenumber methods. Journal of African Earth Sciences 37 no. 3–4: 191–205. doi:10.1016/j.jafrearsci.2002.12.001.
  • Boyce, J., I. Nicholls, R. Keays, and P. Hayman. 2014. Victoria erupts: The Newer Volcanics Province of south-eastern Australia. Geology Today 30 no. 3: 105–109. doi:10.1111/gto.12054.
  • Campello, R.J.G.B., D. Moulavi, and J. Sander. 2013. Density-based clustering based on hierarchical density estimates, 160–172. doi:10.1007/978-3-642-37456-2_14
  • Cawood, P.A. 2005. Terra Australis Orogen: Rodinia breakup and development of the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Paleozoic. Earth-Science Reviews 69 no. 3–4: 249–279. doi:10.1016/j.earscirev.2004.09.001.
  • Cayley, R.A. 2011. Exotic crustal block accretion to the eastern Gondwanaland margin in the Late Cambrian-Tasmania, the Selwyn Block, and implications for the Cambrian-Silurian evolution of the Ross, Delamerian, and Lachlan orogens. Gondwana Research 19 no. 3: 628–649. doi:10.1016/j.gr.2010.11.013.
  • Cayley, R.A., R.J. Korsch, D.H. Moore, R.D. Costelloe, A. Nakamura, C.E. Willman, T.J. Rawling, V.J. Morand, P.B. Skladzien, and P.J. O’Shea. 2011. Crustal architecture of central Victoria: Results from the 2006 deep crustal reflection seismic survey. Australian Journal of Earth Sciences 58 no. 2: 113–156. doi:10.1080/08120099.2011.543151.
  • Cayley, R.A., R.J. Korsch, D.H. Moore, R.D. Costelloe, A. Nakamura, C.E. Willman, T.J. Rawling, V.J. Morand, P.B. Skladzien, and P.J. Shea. 2011. Crustal architecture of central Victoria: Results from the 2006 deep crustal reflection seismic survey. Australian Journal of Earth Sciences 58 no. 2: 113–156. doi:10.1080/08120099.2011.543151.
  • Cayley, R.A., and D.H. Taylor. 2001. Ararat 1:100 000.
  • Cayley, R.A., D.H. Taylor, A.H.M. VandenBerg, and D.H. Moore. 2002. Proterozoic - Early palaeozoic rocks and the Tyennan Orogeny in central Victoria: The Selwyn Block and its tectonic implications. Australian Journal of Earth Sciences 49 no. 2: 225–254. doi:10.1046/j.1440-0952.2002.00921.x.
  • Chukwu, C.G., E.E. Udensi, E.M. Abraham, A.C. Ekwe, and A.O. Selemo. 2018. Geothermal energy potential from analysis of aeromagnetic data of part of the Niger-delta basin, southern Nigeria. Energy 143: 846–853. doi:10.1016/j.energy.2017.11.040.
  • Clemens, J.D., R. Maas, T.E. Waight, and L.K. Kunneke. 2016. Genesis of felsic plutonic magmas and their igneous enclaves: The Cobaw batholith of southeastern Australia. The Journal of Geology 124 no. 3: 293–311. doi:10.1086/685509.
  • Cooper, G.R.J., and M.D. Manzi. 2015. Determining the depth of magnetic sources. Aseg-Pesa 9: 1–3.
  • Crawford, A.J., S. Meffre, and P.A. Symonds. 2003. 120 to 0 Ma tectonic evolution of the southwest Pacific and analogous geological evolution of the 600 to 220 Ma Tasman Fold Belt System. Special Paper of the Geological Society of America 372: 383–403. doi:10.1130/0-8137-2372-8.383.
  • De Castro, D.L., R.A. Fuck, J.D. Phillips, R.M. Vidotti, F.H.R. Bezerra, and E.L. Dantas. 2014. Crustal structure beneath the Paleozoic Parnaíba Basin revealed by airborne gravity and magnetic data, Brazil. Tectonophysics 614 no. April 2020: 128–145. doi:10.1016/j.tecto.2013.12.009.
  • de Melo, F.F., and V.C.F. Barbosa. 2017. Correct structural index defined by base level estimates in Euler deconvolution. 15th international Congress of the Brazilian Geophysical Society & EXPOGEF, Rio de Janeiro, Brazil, 31 July-3 August 2017, 2003, 1086–1091. doi:10.1190/sbgf2017-212
  • DeVries, P.M.R., F. Viégas, M. Wattenberg, and B.J. Meade. 2018. Deep learning of aftershock patterns following large earthquakes. Nature 560 no. 7720: 632–634. doi:10.1038/s41586-018-0438-y.
  • Diakov, S., R. West, D. Schissel, A. Krivtsov, V. Kochnev-Pervoukhov, and I. Migachev. 2002. Recent advances in the Noril’sk model and its application for exploration of Ni-Cu-PGE sulfide deposits. In Integrated Methods for Discovery Global Exploration in the Twenty-First Century (Issue January). Society of Economic Geologists. doi:10.5382/SP.09.10
  • Duan, D., Y. Li, R. Li, and Z. Lu. 2012. Incremental K-clique clustering in dynamic social networks. Artificial Intelligence Review 38 no. 2: 129–147. doi:10.1007/s10462-011-9250-x.
  • Edward, J., K.R. Slater, and I.W. McHaffie. 2001. Bendigo: 1:250 000 map area. In Department of Natural Resources and Environment, 1–55.
  • Edwards, J., K.E. Wohlt, K.R. Slater, A. Olshina, D.F. Hutchinson, and Victoria State Government. 1998. Heathcote and parts of Woodend and Echuca 1:100 000 map area geological report. In Report. Department of Natural Resources and Environment.
  • Ernst, R.E., K.L. Buchan, and I.H. Campbell. 2005. Frontiers in Large Igneous Province research. Lithos 79 no. 3-4 SPEC. ISS.: 271–297. doi:10.1016/j.lithos.2004.09.004.
  • Ester, M., H.-P. Kriegel, J. Sander, and X. Xu. 1996. A density-based algorithm for discovering clusters in large spatial databases with noise. www.aaai.org.
  • Fadel, I., M. van der Meijde, N. Kerle, and N. Lauritsen. 2015. 3D object-oriented image analysis in 3D geophysical modelling: Analysing the central part of the East African Rift System. International Journal of Applied Earth Observation and Geoinformation 35 no. PA: 44–53. doi:10.1016/j.jag.2013.11.004.
  • Fairhead, J., and S. Williams. 2006. Evaluating normalised magnetic derivatives for structural mapping. In SEG 2006 New Orleans Extended Abstract.
  • Fang, G., Z. Chen, Y. Chen, J. Li, B. Zhao, X. Zhou, Z. Zeng, and Y. Zhang. 2015. Geophysical investigations of the geology and structure of the Pangushan-Tieshanlong tungsten ore field, South Jiangxi, China-Evidence for site-selection of the 2000-m Nanling Scientific Drilling Project (SP-NLSD-2). Journal of Asian Earth Sciences 110: 10–18. doi:10.1016/j.jseaes.2015.04.004.
  • Fedi, M., G. Florio, and V. Paoletti. 2015. MHODE: A local-homogeneity theory for improved source-parameter estimation of potential fields. Geophysical Journal International 202 no. 2: 887–900. doi:10.1093/gji/ggv185.
  • Fergusson, C.L., A.P. Nutman, T. Kamiichi, and H. Hidaka. 2013. Evolution of a Cambrian active continental margin: The Delamerian-Lachlan connection in southeastern Australia from a zircon perspective. Gondwana Research 24 no. 3–4: 1051–1066. doi:10.1016/j.gr.2013.03.006.
  • Ferraccioli, F., E. Armadillo, T. Jordan, E. Bozzo, and H. Corr. 2009. Aeromagnetic exploration over the East Antarctic Ice Sheet: A new view of the Wilkes Subglacial Basin. Tectonophysics 478 no. 1–2: 62–77. doi:10.1016/j.tecto.2009.03.013.
  • FitzGerald, D., A. Reid, and P. McInerney. 2004. New discrimination techniques for Euler deconvolution. Computers and Geosciences 30 no. 5: 461–469. doi:10.1016/j.cageo.2004.03.006.
  • Foden, J., M.A. Elburg, J. Dougherty-Page, and A. Burtt. 2006. The timing and duration of the Delamerian orogeny: Correlation with the Ross Orogen and implications for Gondwana Assembly. The Journal of Geology 114 no. 2: 189–210. doi:10.1086/499570.
  • Foss, C., D. Fitzgerald, and M. Barlow. 2021. The Australian national differential RTP grid. Preview 2021 no. 215: 55–58. doi:10.1080/14432471.2021.2012039.
  • Foster, D.A., and D.R. Gray. 2000. Evolution and structure of the Lachlan Fold Belt (Orogen) of Eastern Australia. Annual Review of Earth and Planetary Sciences 28 no. 1: 47–80. doi:10.1146/annurev.earth.28.1.47.
  • Ganguli, S.S., S. Mondal, S.K. Pal, M. Lakshamana, and S. Mahender. 2022. Combined analysis of remote sensing, gravity and magnetic data across Moyar Bhavani Shear Zone, Southern Granulite Terrane (SGT), India: appraisals for crustal architecture and tectonics. Geocarto International 37 no. 26: 13973–14004. doi:10.1080/10106049.2022.2086627.
  • Ganguli, S.S., and S.K. Pal. 2023. Gravity-magnetic appraisal of the southern part of the Cauvery Basin, Eastern Continental Margin of India (ECMI): evidence of a volcanic rifted margin. Frontiers in Earth Science 11 no. July: 1–17. doi:10.3389/feart.2023.1190106.
  • Georgoulas, G., A. Konstantaras, E. Katsifarakis, C.D. Stylios, E. Maravelakis, and G.J. Vachtsevanos. 2013. ‘Seismic-mass’ density-based algorithm for spatio-temporal clustering. Expert Systems with Applications 40 no. 10: 4183–4189. doi:10.1016/j.eswa.2013.01.028.
  • Gibson, G.M., J.M. Totterdell, L.T. White, C.H. Mitchell, A.R. Stacey, M.P. Morse, and A. Whitaker. 2013. Pre-existing basement structure and its influence on continental rifting and fracture zone development along Australia’s southern rifted margin. Journal of the Geological Society 170 no. 2: 365–377. doi:10.1144/jgs2012-040.
  • Glen, R.A., and R.A. Cooper. 2021. Evolution of the East Gondwana convergent margin in Antarctica, southern Australia and New Zealand from the Neoproterozoic to latest Devonian. Earth-Science Reviews 220 May: 103687. doi:10.1016/j.earscirev.2021.103687.
  • Glen, R.A., S. Meffre, and R.J. Scott. 2007. Benambran Orogeny in the Eastern Lachlan Orogen, Australia. Australian Journal of Earth Sciences 54 no. 2–3: 385–415. doi:10.1080/08120090601147019.
  • Graham, S.D., D.A. Holwell, I. McDonald, G.R.T. Jenkin, N.J. Hill, A.J. Boyce, J. Smith, and C. Sangster. 2017. Magmatic Cu-Ni-PGE-Au sulfide mineralisation in alkaline igneous systems: An example from the Sron Garbh intrusion, Tyndrum, Scotland. Ore Geology Reviews 80: 961–984. doi:10.1016/j.oregeorev.2016.08.031.
  • Gray, D.R., and D.A. Foster. 1988. Character and kinematics of faults within the turbidite-dominated Lachlan Orogen: Implications for tectonic evolution of eastern Australia. Journal of Structural Geology 20 no. 12: 1691–1720. doi:10.1016/S0191-8141(98)00089-3.
  • Gray, D.R., and D.A. Foster. 2004. Tectonic evolution of the Lachlan Orogen, southeast Australia: Historical review, data synthesis and modern perspectives. Australian Journal of Earth Sciences 51 no. 6: 773–817. doi:10.1111/j.1400-0952.2004.01092.x.
  • Gray, D.R., and C.E. Willman. 1991. Thrust-related strain gradients and thrusting mechanisms in a chevron-folded sequence, southeastern Australia. Journal of Structural Geology 13 no. 6: 691–710. doi:10.1016/0191-8141(91)90031-D.
  • Gray, D.R., C.E. Willman, and D.A. Foster. 2006. Crust restoration for the western Lachlan Orogen using the strain-reversal, area-balancing technique: Implications for crustal components and original thicknesses. Australian Journal of Earth Sciences 53 no. 2: 329–341. doi:10.1080/08120090500499305.
  • Grijalva, A., A.A. Nyblade, K. Homman, N.J. Accardo, J.B. Gaherty, C.J. Ebinger, D.J. Shillington, et al. 2018. Seismic evidence for plume- and craton-influenced upper mantle structure beneath the northern Malawi Rift and the Rungwe Volcanic Province, East Africa. Geochemistry, Geophysics, Geosystems 19 no. 10: 3980–3994. doi:10.1029/2018GC007730.
  • Guardo, R., L. De Siena, J. Prudencio, and G. Ventura. 2022. Imaging the absorbing feeding and eruptive pathways of deception island, Antarctica. Geophysical Research Letters 49 no. 19. doi:10.1029/2022GL099540.
  • Guimarães, S.N.P., D. Ravat, and V.M. Hamza. 2014. Combined use of the centroid and matched filtering spectral magnetic methods in determining thermomagnetic characteristics of the crust in the structural provinces of central Brazil. Tectonophysics 624–625: 87–99. doi:10.1016/j.tecto.2014.01.025.
  • Gupta, N., S.K. Pal, and J. Das. 2022. GIS-based evolution and comparisons of landslide susceptibility mapping of the East Sikkim Himalaya. Annals of GIS 28 no. 3: 359–384. doi:10.1080/19475683.2022.2040587.
  • Holt, S.J., S.P. Holford, and J. Foden. 2013. New insights into the magmatic plumbing system of the South Australian Quaternary Basalt province from 3D seismic and geochemical data. Australian Journal of Earth Sciences 60 no. 8: 797–817. doi:10.1080/08120099.2013.865143.
  • Hu, L., X. Zheng, Y. Duan, X. Yan, Y. Hu, and X. Zhang. 2019. First-arrival picking with a U-net convolutional network. Geophysics 84 no. 6: U45–U57. doi:10.1190/geo2018-0688.1.
  • Huang, K.Y., J. De You, K.J. Chen, H.L. Lai, and A.J. Don. 2006. Neural network for parameters determination and seismic pattern detection. SEG Technical Program Expanded Abstracts 25 no. 1: 2285–2289. doi:10.1190/1.2369991.
  • Huston, D.L., T.P. Mernagh, S.G. Hagemann, M.P. Doublier, M. Fiorentini, D.C. Champion, A. Lynton Jaques, et al. 2016. Tectono-metallogenic systems — The place of mineral systems within tectonic evolution, with an emphasis on Australian examples. Ore Geology Reviews 76 September 2015: 168–210. doi:10.1016/j.oregeorev.2015.09.005.
  • Jessell, M.W., R.K. Valenta, G. Jung, J.P. Cull, and A. Geiro. 1993. Structural geophysics. Exploration Geophysics 24 no. 4: 599–602. doi:10.1071/EG993599.
  • Ji, X., M. Chen, L. Yang, and G. Chen. 2023. The relationship between Carlin-type Au mineralization and magmatism in the Youjiang Basin-a case study from the Mingshan gold deposit in northwest Guangxi, China. Ore Geology Reviews 157 no. 26: 105400. doi:10.1016/j.oregeorev.2023.105400.
  • Jia, Y., and J. Ma. 2017. What can machine learning do for seismic data processing? An interpolation application. Geophysics 82 no. 3: V163–V177. https://doi.org/10.1190/GEO2016-0300.1.
  • Johnson, S.C. 1967. Hierarchical clustering schemes. Psychometrika 32 no. 3: 241–254. doi:10.1007/BF02289588.
  • Kemp, A.I.S., C.M. Gray, D.J. Ellis, J.A.C. Anderson, and D.J. Ferguson. 2002. Delamerian Glenelg tectonic zone, Western Victoria: Characterisation and synthesis of igneous rocks. Australian Journal of Earth Sciences 49 no. 2: 201–224. doi:10.1046/j.1440-0952.2002.00920.x.
  • Khalil, M.H. 2016. Subsurface faults detection based on magnetic anomalies investigation: A field example at Taba protectorate, South Sinai. Journal of Applied Geophysics 131: 123–132. doi:10.1016/j.jappgeo.2016.06.001.
  • Korsch, R.J., T.J. Barton, D.R. Gray, A.J. Owen, and D.A. Foster. 2002. Geological interpretation of a deep seismic-reflection transect across the boundary between the Delamerian and Lachlan Orogens, in the vicinity of the Grampians, western Victoria. Australian Journal of Earth Sciences 49 no. 6: 1057–1075. doi:10.1046/j.1440-0952.2002.00963.x.
  • Kumar, R., A.R. Bansal, P.G. Betts, and D. Ravat. 2021. Re-assessment of the depth to the base of magnetic sources (DBMS) in Australia from aeromagnetic data using the defractal method. Geophysical Journal International 225 no. 1: 530–547. doi:10.1093/gji/ggaa601.
  • Leader, L.D., C.J.L. Wilson, and J.A. Robinson. 2013. Structural constraints and numerical simulation of strain localization in the Bendigo Goldfield, Victoria, Australia. Economic Geology 108 no. 2: 279–307. doi:10.2113/econgeo.108.2.279.
  • Lei, L., L. Yu, Z. Xiong, L. Han, X. Wei, and Z. Shifan. 2019. Convolutional recurrent neural networks based waveform classification in seismic facies analysis. SEG Technical Program Expanded Abstracts, 2599–2603. doi:10.1190/segam2019-3215237.1.
  • Le Pape, F., A.G. Jones, M.W. Jessell, S. Perrouty, L.A. Gallardo, L. Baratoux, C. Hogg, et al. 2017. Crustal structure of southern Burkina Faso inferred from magnetotelluric, gravity and magnetic data. Precambrian Research 300 no. January: 261–272. doi:10.1016/j.precamres.2017.08.013.
  • Li, Z., M.A. Meier, E. Hauksson, Z. Zhan, and J. Andrews. 2018. Machine learning seismic wave discrimination: application to earthquake early warning. Geophysical Research Letters 45 no. 10: 4773–4779. doi:10.1029/2018GL077870.
  • Liang, J., J. Ma, and X. Zhang. 2014. Seismic data restoration via data-driven tight frame. Geophysics 79 no. 3: V65–V74. https://doi.org/10.1190/GEO2013-0252.1.
  • Lillis, R.J., J. Dufek, J.E. Bleacher, and M. Manga. 2009. Demagnetization of crust by magmatic intrusion near the Arsia Mons volcano: Magnetic and thermal implications for the development of the Tharsis province, Mars. Journal of Volcanology and Geothermal Research 185 no. 1–2: 123–138. doi:10.1016/j.jvolgeores.2008.12.007.
  • Liu, Z., W. Zhou, and Y. Yuan. 2023. 3D DBSCAN detection and parameter sensitivity of the 2022 Yangtze river summertime heatwave and drought. Atmospheric and Oceanic Science Letters 16 no. November 2022. doi:10.1016/j.aosl.2022.100324.
  • Loo, P.H., and E.C. Leong. 2018. Discussion of “tool for analysis of multichannel analysis of surface waves (MASW) field data and evaluation of shear wave velocity profiles of soils”. Canadian Geotechnical Journal 55 no. 10: 1510–1511. doi:10.1139/cgj-2018-0078.
  • MacQueen, J. 1967. Some methods for classification and analysis of MultiVariate observations. Proceedings of the Fifth Berkeley Symposium on Mathematical Statistics and Probability 1: 281–297.
  • Malfante, M., M. Dalla Mura, J.I. Mars, J.P. Métaxian, O. Macedo, and A. Inza. 2018. Automatic classification of volcano seismic signatures. Journal of Geophysical Research: Solid Earth 123 no. 12: 10,645–10,658. doi:10.1029/2018JB015470.
  • Mandelli, S., F. Borra, V. Lipari, P. Bestagini, A. Sarti, and S. Tubaro. 2018. Seismic data interpolation through convolutional autoencoder. SEG Technical Program Expanded Abstracts, 4101–4105. doi:10.1190/segam2018-2995428.1.
  • Marques, J.C., and M.B. Orger. 2019. Clusterdv: A simple density-based clustering method that is robust, general and automatic. Bioinformatics (Oxford, England) 35 no. 12: 2125–2132. doi:10.1093/bioinformatics/bty932.
  • Maus, S., U. Barckhausen, H. Berkenbosch, N. Bournas, J. Brozena, V. Childers, F. Dostaler, … F.C. Tontini. 2009. EMAG2: a 2-arc min resolution earth magnetic anomaly grid compiled from satellite, airborne, and marine magnetic measurements. Geochemistry, Geophysics, Geosystems 10 no. 8. doi:10.1029/2009GC002471.
  • Mazabraud, Y., N. Béthoux, and S. Deroussi. 2005. Characterisation of the seismological pattern in a slowly deforming intraplate region: Central and western France. Tectonophysics 409 no. 1–4: 175–192. doi:10.1016/j.tecto.2005.08.021.
  • McLean, M.A., V.J. Morand, and R.A. Cayley. 2010. Gravity and magnetic modelling of crustal structure in central Victoria: What lies under the Melbourne zone? Australian Journal of Earth Sciences 57 no. 2: 153–173. doi:10.1080/08120090903416245.
  • Merdith, A.S., S.E. Williams, S. Brune, A.S. Collins, and R.D. Müller. 2019. Rift and plate boundary evolution across two supercontinent cycles. Global and Planetary Change 173 no. December: 1–14. doi:10.1016/j.gloplacha.2018.11.006.
  • Merdith, A. S., Williams, S. E., Collins, A. S., Tetley, M. G., Mulder, J. A., Blades, M. L., Young, A., Armistead, S. E., Cannon, J., Zahirovic, S., & Müller, R. D. (2021). Extending full-plate tectonic models into deep time: Linking the Neoproterozoic and the Phanerozoic. In Earth-science reviews, Vol. 214. Elsevier B.V. doi:10.1016/j.earscirev.2020.103477
  • Mikhailov, V., A. Galdeano, M. Diament, A. Gvishiani, S. Agayan, S. Bogoutdinov, E. Graeva, and P. Sailhac. 2003. Application of Artificial Intelligence for Euler Solutions Clustering 68 no. 1: 168–180.
  • Milano, M., M. Fedi, and J.D. Fairhead. 2016. The deep crust beneath the trans-European suture zone from a multiscale magnetic model. Journal of Geophysical Research: Solid Earth 121 no. 9: 6276–6292. doi:10.1002/2016JB012955.
  • Miller, J.McL., M.S. Norvick, and C.J.L. Wilson. 2002. Basement controls on rifting and the associated formation of ocean transform faults—Cretaceous continental extension of the southern margin of Australia. Tectonophysics 359 no. 1–2: 131–155. doi:10.1016/S0040-1951(02)00508-5.
  • Miller, J.M.L., D. Phillips, C.J.L. Wilson, and L.J. Dugdale. 2005. Evolution of a reworked orogenic zone: The boundary between the Delamerian and Lachlan Fold Belts, southeastern Australia. Australian Journal of Earth Sciences 52 no. 6: 921–940. doi:10.1080/08120090500304265.
  • Miller, J.M.L., C.J.L. Wilson, and L.J. Dugdale. 2006. Stawell gold deposit: A key to unravelling the Cambrian to Early Devonian structural evolution of the western Victorian goldfields. Australian Journal of Earth Sciences 53 no. 5: 677–695. doi:10.1080/08120090600823313.
  • Milligan, P.R., and R. Franklin. 2004. A new generation Magnetic Anomaly Grid Database of Australia (MAGDA). Preview, December, 25–29.
  • Minelli, L., A. Vecchio, F. Speranza, I. Nicolosi, F. D’Ajello Caracciolo, S. Chiappini, R. Carluccio, and M. Chiappini. 2016. Aeromagnetic investigation of southern Calabria and the Messina Straits (Italy): Tracking seismogenic sources of 1783 and 1908 earthquakes. Journal of Geophysical Research: Solid Earth 121 no. 3: 1297–1315. doi:10.1002/2015JB012305.
  • Mohammadzadeh, M. 2015. Interpretation of aeromagnetic data to locate buried faults in north of Zanjan Province, Iran. Journal of Geophysics & Remote Sensing 04 no. 02. doi:10.4172/2169-0049.1000143.
  • Mooney, W. D. 2015. Crust and lithospheric structure - global crustal structure. In Treatise on geophysics: second edition, Vol. 1. Elsevier Inc. doi:10.1016/B978-0-444-53802-4.00010-5
  • Moore, D.H., P.G. Betts, and M. Hall. 2015. Fragmented Tasmania: the transition from Rodinia to Gondwana. Australian Journal of Earth Sciences 62 no. 1: 37–41. doi:10.1080/08120099.2014.966757.
  • Moore, D.H., P.G. Betts, and M. Hall. 2016. Constraining the VanDieland microcontinent at the edge of East Gondwana, Australia. Tectonophysics 687: 158–179. doi:10.1016/j.tecto.2016.09.009.
  • Morand, V.J., W.R.H. Ramsay, M. Hughes, and J.M. Stanley. 1995. The southern avoca fault zone: site of a newly identified ‘greenstone’. Australian Journal of Earth Sciences 42 no. 2: 133–143. doi:10.1080/08120099508728186.
  • Morand, V.J., K.E. Wohlt, R.A. Cayley, D.H. Taylor, A.I.S. Kemp, B.A. Simons, and A.P.M. Magart. 2003. Glenelg special map area geological report. Geological Survey of Victoria Report 123.
  • Moresi, L., P.G. Betts, M.S. Miller, and R.A. Cayley. 2014. Dynamics of continental accretion. Nature 508 no. 7495: 245–248. doi:10.1038/nature13033.
  • Mousavi, S.M., and C.A. Langston. 2016. Hybrid seismic denoising using higher-order statistics and improved wavelet block thresholding. Bulletin of the Seismological Society of America 106 no. 4: 1380–1393. doi:10.1785/0120150345.
  • O’Donnell, J.P., S. Agrawal, C.M. Eakin, S. Thiel, K. Brand, A. Gorbatov, and B. Goleby. 2023. Mapping crustal structure across southern Australia using seismic ambient noise tomography. Gondwana Research 121: 307–324. doi:10.1016/j.gr.2023.04.013.
  • Offler, R., S. Mcknight, and V. Morand. 1998. Tectonothermal history of the western Lachlan Fold Belt, Australia: Insights from white mica studies. Journal of Metamorphic Geology 16 no. 4: 531–540. doi:10.1111/j.1525-1314.1998.00153.x.
  • Oruç, B., and H.H. Selim. 2011. Interpretation of magnetic data in the Sinop area of Mid Black Sea, Turkey, using tilt derivative, Euler deconvolution, and discrete wavelet transform. Journal of Applied Geophysics 74 no. 4: 194–204. doi:10.1016/j.jappgeo.2011.05.007.
  • Pedreira, D., J.A. Pulgar, J. Gallart, and M. Torné. 2007. Three-dimensional gravity and magnetic modeling of crustal indentation and wedging in the western Pyrenees-Cantabrian Mountains. Journal of Geophysical Research: Solid Earth 112 no. 12: 1–19. doi:10.1029/2007JB005021.
  • Pelton, J.R.. 2005. 7. Near-surface seismology: Wave propagation. In Near-surface geophysics,177–218 Society of Exploration Geophysicists. doi:10.1190/1.9781560801719.ch7.
  • Peron-Pinvidic, G., T. Åkermoen, and L.I. Leivestad. 2022. The North-East Atlantic Mid-Norwegian rifted margin: Insights from the deep imaging Geoex MCG RDI19 dataset. Tectonophysics 824: 229225. doi:10.1016/j.tecto.2022.229225.
  • Phillips, J.D. 2001. Designing matched bandpass and azimuthal filters for the separation of potential-field anomalies by source region and source type. ASEG Extended Abstracts 2001 no. 1: 1–4. https://doi.org/10.1071/aseg2001ab110.
  • Phillips, D., B. Fu, C.J.L. Wilson, M.A. Kendrick, A.M. Fairmaid, and J.M. Miller. 2012. Timing of gold mineralisation in the western lachlan orogen, se Australia: A critical overview. Australian Journal of Earth Sciences 59 no. 4: 495–525. doi:10.1080/08120099.2012.682738.
  • Phillips, G., J.McL. Miller, and C.J.L. Wilson. 2002. Structural and metamorphic evolution of the Moornambool Metamorphic Complex, western Lachlan Fold Belt, southeastern Australia. Australian Journal of Earth Sciences 49 no. 5: 891–913. doi:10.1046/j.1440-0952.2002.00958.x.
  • Phillips, G., and R. Offler. 2011. Contrasting modes of eclogite and blueschist exhumation in a retreating subduction system: The Tasmanides, Australia. Gondwana Research 19 no. 3: 800–811. doi:10.1016/j.gr.2010.07.009.
  • Pinotti, L.P., F.J. D’Eramo, R.F. Weinberg, M. Demartis, J.M. Tubía, J.E. Coniglio, S. Radice, M.N. Maffini, and E. Aragón. 2016. Contrasting magmatic structures between small plutons and batholiths emplaced at shallow crustal level (Sierras de Córdoba, Argentina). Journal of Structural Geology 92: 46–58. doi:10.1016/j.jsg.2016.09.009.
  • Poudjom, D.Y., B.R.S. Minty, M. Hutchens, and R.J.L. Lane. 2019. Total Magnetic Intensity (TMI) Grid of Australia 2019 - seventh edition - 40 m cell size. doi:10.26186/5dd5fb557fffb.
  • Preiss, W.V. 2000. The Adelaide Geosyncline of South Australia and its significance in Neoproterozoic continental reconstruction. Precambrian Research 100 no. 1–3: 21–63. doi:10.1016/S0301-9268(99)00068-6.
  • Raimi, J., B.B.M. Dewu, and P. Sule. 2014. An Interpretation of Structures from the Aeromagnetic Field over a Region in the Nigerian Younger Granite Province. 2014(March), 313–323.
  • Ravat, D. 1996. Analysis of the Euler method and its applicability in environmental magnetic investigations. Journal of Environmental and Engineering Geophysics 1 no. 3: 229–238. doi:10.4133/JEEG1.3.229.
  • Ravi Kumar, C., N. Kesiezie, B. Pathak, S. Maiti, and R. K. Tiwari. 2020. Mapping of basement structure beneath the Kohima Synclinorium, north-east India via Bouguer gravity data modelling. Journal of Earth System Science 129 no. 1. doi:10.1007/s12040-019-1326-z.
  • Rawling, T.J., C.R. Osborne, M.A. Mclean, P.B. Skladzien, R.A. Cayley. 2011. 3D Victoria Final Report. GeoScience Victoria 3D Victoria Report, November, 1–98. ISBN: 978-1-74264-977-1.
  • Reichstein, M., G. Camps-Valls, B. Stevens, M. Jung, J. Denzler, N. Carvalhais, and D. Prabhat. 2019. Deep learning and process understanding for data-driven earth system science. Nature 566 no. 7743: 195–204. doi:10.1038/s41586-019-0912-1.
  • Reid, A.B., J.M. Allsop, H. Granser, A.J. Millet, and I.W. Somerton. 2002. To: “Magnetic interpretation in three dimensions using Euler deconvolution” by A. B. Reid, J. M. Allsop, H. Granser, A. J. Millett, and I. W. Somerton (GEOPHYSICS, p. 80–91, January 1990). Geophysics 55 no. 4: 502–502. https://doi.org/10.1190/1.1442861.
  • Reid, A.B., J.M. Allsop, H. Granser, A.J. Millett, and I.W. Somerton. 1990. Magnetic interpretation in three dimensions using Euler deconvolution. Geophysics 55: 80–91.
  • Reid, A.B., J. Ebbing, and S.J. Webb. 2014. Avoidable Euler errors - The use and abuse of Euler deconvolution applied to potential fields. Geophysical Prospecting 62 no. 5: 1162–1168. doi:10.1111/1365-2478.12119.
  • Reid, A.B., and J.B. Thurston. 2014. The structural index in gravity and magnetic interpretation: Errors, uses, and abuses. GEOPHYSICS 79 no. 4: J61–J66. doi:10.1190/geo2013-0235.1.
  • Robert, G., and J. Cooper. 2015. The analytic downward continuation of the tilt angle. JANUARY, 2015.
  • Romero-Ruiz, A., N. Linde, T. Keller, and D. Or. 2018. A review of geophysical methods for soil structure characterization. Reviews of Geophysics 56 no. 4: 672–697. doi:10.1029/2018RG000611.
  • Samsu, A., A.R. Cruden, N.E. Molnar, and R.F. Weinberg. 2021. Inheritance without reactivation: Insights from crustal-scale analogue experiments. Tectonics 40 no. 5: e2020TC006596.
  • Shaole, A., Z. Zhixin, Z. Kefa, and W. Jinlin. 2020. Subsurface structures of the Xiaorequanzi deposit, NW China: New insights from gravity, magnetic and electromagnetic data. Geophysical Prospecting, 1365-2478.13049. doi:10.1111/1365-2478.13049.
  • Slater, K.R., and S.J. Haydon. 1999. Copper occurrences in the Heathcote Greenstone belt. ASEG Extended Abstracts 1999 no. 1: 99–112. https://doi.org/10.1071/asegspec11_07.
  • Song, J.H., S. Kim, J. Rhie, S.H. Lee, Y.H. Kim, and T.S. Kang. 2018. Imaging of lithospheric structure beneath Jeju volcanic island by teleseismic traveltime tomography. Journal of Geophysical Research: Solid Earth 123 no. 8: 6784–6801. doi:10.1029/2018JB015979.
  • Spaggiari, C.V., D.R. Gray, D.A. Foster, and S. McKnight. 2003. Evolution of the boundary between the western and central Lachlan Orogen: Implications for tasmanide tectonics. Australian Journal of Earth Sciences 50 no. 5: 725–749. doi:10.1111/j.1440-0952.2003.01022.x.
  • Spector, A., and F.S. Grant. 1970. Statistical models for interpreting aeromagnetic data. Geophysics 35 no. 2: 293–302. doi:10.1190/1.1440092.
  • Squire, R.J., C.J.L. Wilson, L.J. Dugdale, B.J. Jupp, and A.L. Kaufman. 2006. Cambrian backarc-basin basalt in western Victoria related to evolution of a continent-dipping subduction zone. Australian Journal of Earth Sciences 53 no. 5: 707–719. doi:10.1080/08120090600827405.
  • Sreejith, K.M., S. Rajesh, T.J. Majumdar, G. Srinivasa Rao, M. Radhakrishna, K.S. Krishna, and A.S. Rajawat. 2013. High-resolution residual geoid and gravity anomaly data of the northern Indian Ocean - An input to geological understanding. Journal of Asian Earth Sciences 62: 616–626. doi:10.1016/j.jseaes.2012.11.010.
  • Sridhar, M., R. Babu, V. Markandeyulu, A. Raju, B.V.S.N. Chaturvedi, A.K. Roy, and M. K. 2017. A reassessment of the Archean-Mesoproterozoic tectonic development of the southeastern Chhattisgarh Basin, Central India through detailed aeromagnetic analysis. Tectonophysics 712–713 no. June: 289–302. doi:10.1016/j.tecto.2017.05.022.
  • Srinivasa, G., and M. Radhakrishna. 2017. A comparative account of terrestrial and satellite based potential field data for regional tectonic structural interpretation and crustal scale modeling with reference to the Indian Region. International Journal of Earth Sciences and Engineering 10 no. 04: 903–914. doi:10.21276/ijee.2017.10.0426.
  • Stavrev, P., and A. Reid. 2007. Degrees of homogeneity of potential fields and structural indices of Euler deconvolution. Geophysics 72 no. 1: 1–12. doi:10.1190/1.2400010.
  • Stavrev, P., and A. Reid. 2009. Euler deconvolution of gravity anomalies from thick contact/fault structures with negative structural index. SEG Technical Program Expanded Abstracts 2009: 913–917. doi:10.1190/1.3255899.
  • Stephenson, R., T. Yegorova, and S. Stovba. 2021. An investigation of how intracratonic rifting is “seeded”: Case study of the Late Devonian Dniepr-Donets Basin rift within the East European Craton. Precambrian Research 362 no. February: 106305. doi:10.1016/j.precamres.2021.106305.
  • Stixrude, L., and R. Jeanloz 2015. Constraints on seismic models from other disciplines - Constraints from mineral physics on seismological models. In Treatise on geophysics: Second Edition, Vol. 1, Issue 1923. Elsevier B.V. doi:10.1016/B978-0-444-53802-4.00022-1
  • Teknik, V., and A. Ghods. 2017. Depth of magnetic basement in Iran based on fractal spectral analysis of aeromagnetic data. Geophysical Journal International 209 no. 3: 1878–1891. doi:10.1093/gji/ggx132.
  • Thompson, D.T. 1982. EULDPH: A new technique for making computer-assisted depth estimates from magnetic data. GEOPHYSICS 47 no. 1: 31–37. doi:10.1190/1.1441278.
  • Tran, T.N., K. Drab, and M. Daszykowski. 2013. Revised DBSCAN algorithm to cluster data with dense adjacent clusters. Chemometrics and Intelligent Laboratory Systems 120: 92–96. doi:10.1016/j.chemolab.2012.11.006.
  • VandenBerg, A.H.M. 1978. The tasman fold belt system in Victoria. Tectonophysics 48 no. 3–4: 267–297. doi:10.1016/0040-1951(78)90121-X.
  • VandenBerg, A.H.M., C.E. Willman, S. Maher, B.A. Simons, R.A. Cayley, D.H. Taylor, V.J. Morand, D.H. Moore, and A. Radojkovic, eds. 2000. The tasman fold belt system in Victoria: geology and mineralisation of proterozoic to carboniferous rocks. Geological Survey of Victoria Special Publication.
  • Wang, J., X. Meng, and F. Li. 2017. New improvements for lineaments study of gravity data with improved Euler inversion and phase congruency of the field data. Journal of Applied Geophysics 136: 326–334. doi:10.1016/j.jappgeo.2016.11.017.
  • Welsh, S.I., D.V. Higgins, and G.A. Callaway, eds. 2011. Surface Geology of Victoria 1: 250 000.
  • Whitehead, N. 2012. Perfoming Euler Deconvolution Analysis. Montaj Grav / Mag Interpretation (Tutorials and User Guilds), 57 and 58.
  • Williams, H.A., P.G. Betts, and L. Ailleres. 2009. Constrained 3D modeling of the Mesoproterozoic Benagerie Volcanics, Australia. Physics of the Earth and Planetary Interiors 173 no. 3–4: 233–253. doi:10.1016/j.pepi.2009.01.002.
  • Willman, C.E., R.J. Korsch, D.H. Moore, R.A. Cayley, V.A. Lisitsin, T.J. Rawling, V.J. Morand, and P.J. O’Shea. 2010. Crustal-scale fluid pathways and source rocks in the victorian gold province, Australia: Insights from deep seismic reflection profiles. Economic Geology 105 no. 5: 895–915. doi:10.2113/econgeo.105.5.895.
  • Wilson, C.J.L., D.H. Moore, S.A. Vollgger, and H.E. Madeley. 2020. Structural evolution of the orogenic gold deposits in central Victoria, Australia: The role of regional stress change and the tectonic regime. Ore Geology Reviews 120 January: 103390. doi:10.1016/j.oregeorev.2020.103390.
  • Wilson, C.J.L., T.M. Will, R.A. Cayley. 1992. Geologic framework and tectonic evolution in Western Victoria, Australia. Tectonophysics 214 no. 1-4: 93–127. doi:10.1016/0040-1951(92)90192-9.
  • Wu, G., G. Chen, D. Wang, Q. Cheng, Z. Zhang, J. Yang, and S. Xie. 2020. Identifying mineral prospectivity using seismic and potential field data in the Hongniangyu district, Inner Mongolia, China. Ore Geology Reviews 119 May 2019: 103317. doi:10.1016/j.oregeorev.2020.103317.
  • Yu, K., X. Dang, H. Bart, and Y. Chen. 2015. Robust model-based learning via spatial-EM algorithm. IEEE Transactions on Knowledge and Data Engineering 27 no. 6: 1670–1682. doi:10.1109/TKDE.2014.2373355.
  • Yu, S., and J. Ma. 2021. Deep learning for geophysics: current and future trends. Reviews of Geophysics 59 no. 3: 1–36. doi:10.1029/2021RG000742.
  • Yule, C.T.G., and C. Spandler. 2022. Geophysical and geochemical evidence for a New Mafic Magmatic Province within the northwest shelf of Australia. Geochemistry, Geophysics, Geosystems 23 no. 2: 1–23. doi:10.1029/2021GC010030.
  • Zeng, X., C.H. Thurber, H.F. Wang, D. Fratta, and K.L. Feigl. 2021. High-resolution shallow structure at brady hot springs using ambient noise tomography (ANT) on a trenched distributed acoustic sensing (DAS) array. Distributed Acoustic Sensing in Geophysics: Methods and Applications 2016 March 2016: 101–110. https://doi.org/10.1002/9781119521808.ch08.
  • Zhou, S., Z. Xu, and F. Liu. 2017. Method for determining the optimal number of clusters based on agglomerative hierarchical clustering. IEEE Transactions on Neural Networks and Learning Systems 28 no. 12: 3007–3017. doi:10.1109/TNNLS.2016.2608001.