Next Article in Journal
Geochronology and Geochemistry of Granitic Pegmatites from Tashidaban Li Deposit in the Central Altun Tagh, Northwest China
Next Article in Special Issue
Three-Dimensional Electrical Structure and Metallogenic Background of the Southeastern Hubei Ore Concentration Area
Previous Article in Journal
Controlling Factors of Organic Matter Enrichment in Marine–Continental Transitional Shale: A Case Study of the Upper Permian Longtan Formation, Northern Guizhou, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Relationship between the Metallogenic System of Metallic Mineral Deposits and Lithospheric Electrical Structure: Insight from Magnetotelluric Imaging

1
School of Geophysics and Information Technology, China University of Geosciences, Beijing 100083, China
2
Key Laboratory of Geo-Detection of Ministry of Education, Beijing 100083, China
*
Authors to whom correspondence should be addressed.
Current address: Xueyuan 29 Road, Haidian District, Beijing 100083, China.
Minerals 2024, 14(6), 541; https://doi.org/10.3390/min14060541
Submission received: 15 April 2024 / Revised: 19 May 2024 / Accepted: 22 May 2024 / Published: 24 May 2024
(This article belongs to the Special Issue Geoelectricity and Electrical Methods in Mineral Exploration)

Abstract

:
In development over 70 years, magnetotelluric (MT) sounding, a high-resolution technique for subsurface electrical resistivity imaging, has been widely applied in resource exploration in the Earth. The key factors of the metallogenic system of metallic mineral deposits can be closely correlated to the electrical anomalies of the lithosphere. In this paper, we review the relationship between the electrical resistivity model of the lithosphere and the metallogenic system. At the beginning, we indicate why the electrical parameters relate to the metallogenic system in all geophysical parameters. The advantage of MT sounding in sketching an electrical resistivity model of the lithosphere is subsequently discussed, and some methods of data processing, analysis and inversion are also introduced. Furthermore, we summarize how to bridge the relationship between the electrical resistivity model of the lithosphere and metallogenic system, and analyze the influence of the rheological variation estimated from conductivity in the lithosphere on mineralization. In the end, we list some typical cases of the application of MT sounding in mineral exploration, and also give some suggestions for future work. This study is aimed at providing guidance in discussing the metallogenic system using an electrical resistivity model.

1. Introduction

The material circulation between the Earth’s sub-systems maintains the dynamic equilibrium in the Earth’s system, which allows the abnormal enrichment of metals to form the mineral resources that human society depends on. The orogenic belts formed by oceanic subduction and continental collision are regions of the most active material circulation and the richest mineral resources in the Earth [1,2]. Statistical data show that 92.5 percent of the giant deposits use water as the driving force of mineralization, and 2/3 of giant deposits are related to magmatic–hydrothermal fluids, in which most metals are precipitated [3]. Therefore, the magmatic–hydrothermal deposits are the largest deposit family [4]. The upwelling, flow and thermal erosion of the asthenosphere with different types induce the partial melting of the underlying lithosphere, the rheological deformation of the crust and surface–crust–mantle interaction, which cause the activation and massive enrichment of metals and form the various ore deposits produced by the circulation of magmatic–hydrothermal fluids [5,6,7,8].
Previous studies stated that the components of a metallogenic system can be categorized into seven parts [9,10]; that is, (1) the metallogenic geodynamic setting, which facilitates the production of the metal-rich magmatic–hydrothermal fluids; (2) favorable lithospheric architecture; (3) sources of fluids; (4) drivers of fluid flow; (5) fluid-flow pathways; (6) physical and chemical processes of metal migration and deposition, and (7) preservation conditions of the ore deposit. The ore deposits can be formed on multiple scales, while the single metallogenic system can be extended over hundreds of thousands of square kilometers, from the entire crust to the upper mantle [11]. Although geophysical methods mainly represent the current subsurface material structure, the most recent geological event can modify the previous mineralization, and furthermore, the large-scale structural features revealed by geophysical methods (e.g., the magnetotelluric sounding and seismic imaging) could be preserved over a long historical period [12]. Therefore, it is feasible to discuss the metallogenic system based on the spatial variation of geophysical parameters.
At present, our discussion of the metallogenic system of ore deposits can concentrate on (1) an integrative model for revealing crust–mantle material structure; (2) the interaction of layers in the Earth; (3) tracing the circulation of metallogenic materials; (4) illustrating the enrichment mechanism of metals and (5) establishing the metallogenic model [9,12]. Therefore, besides the abundance of geological and geochemical research, geophysical models of the lithosphere are also needed to provide the geophysical evidence required to analyze the source–migration–deposit process of the magma–fluid–mineralization system.

2. Electrical Parameter and Electrical Prospecting

Electrical conductivity (EC; simplified as conductivity, σ ) is a material’s ability to conduct electric current, and takes the SI units of a Siemens per meter (S/m), which is the reciprocal of electrical resistivity (simplified as resistivity; SI units of ohm-meters (Ω·m)). Conductivity is related to the transmission of the electric current by free charge carriers, and can be described as the special charge carrier (Equation (1)).
σ = c q μ
where c, q and μ are the charge concentration, charge and mobility of the charge carrier, respectively.
The ionic conduction, electronic conduction and semi-conduction (proton conduction and small polaron conduction) are the primary conduction mechanisms in the minerals and rocks of the lithosphere, and the conditions of temperature and pressure have different influences on different conduction mechanisms. The conductivity of minerals or rocks in the lithosphere may be related to one or several conduction mechanisms. Note that it is common for a single mechanism of a material to dominate in a given thermal regime. The conductive phases in the lithosphere mainly include saline fluids, partial melts, graphite, sulfides and water (hydrogen) in nominally anhydrous minerals (NAMs). Previous studies have summarized the conducive characteristics of these phases [13,14] (Figure 1), and in this paper, these are not further described. Conductive mediums, such as graphite/graphite film, metallic oxide/sulfide, aqueous fluid and molten silicate/carbonate melt, have a conductivity that is several orders of magnitude higher than that of country rocks [15]. When conductive minerals form a connected net, the bulk conductivity can be controlled by a small number of highly conductive minerals.
Furthermore, the vapor phase (mainly volatile) also influences the bulk conductivity of rocks. The deposit types of different volatiles in the lithosphere depend on the solubility of different volatiles in the different fluids or melts. In comparison with the conductivity of the vapor phase, the volatiles in the fluids or melts have a positive effect on the bulk conductivity of rocks. Taking carbon dioxide (CO2), which is the second most abundant volatile in the lithosphere, as an example, (1) CO2 has much higher solubility in alkali-rich melts [16,17,18,19]; (2) the conductivity of the CO2 presented as the vapor phase is approximately 0.01 S/m (resistivity of 100 Ω·m), and 5–8 vol% of CO2 can reduce the conductivity from 0.04 S/m (resistivity of 250 Ω·m) to 0.01 S/m (resistivity of 100 Ω·m) [20]; (3) when CO2 is in the melt, the carbonate ion (CO2−3) can act as the effective charge carrier, and contribute significantly to electrical conduction [16,21].
Figure 1. A cartoon showing key features of the crust and upper mantle and their typical ranges of resistivity, with numbers denoting resistivity (Ω·m) [14,22].
Figure 1. A cartoon showing key features of the crust and upper mantle and their typical ranges of resistivity, with numbers denoting resistivity (Ω·m) [14,22].
Minerals 14 00541 g001
Because conductivity is extremely sensitive to the high-conductivity, high-temperature and high-rheology materials (e.g., metallic minerals, melts and aqueous fluids) [23], the electrical structure (the spatial variation of conductivity or resistivity) may be preferred when studying the metallogenic system (especially metallic minerals). Electric resistivity (direct current, DC) and electromagnetic (EM) methods (such as the time-domain electromagnetic (TEM) method and magnetotelluric method (MT)) can be used to sketch the variation of resistivity/conductivity in the underground medium [24,25]. Because the probing depth of electric resistivity methods is shallow, the EM methods are the main methods used to establish the lithospheric electrical structure and study the metallogenic system. Furthermore, the composition of the crust and mantle is heterogeneous, and therefore, the EM method is used in measuring volume current, which represents the average volume of the compositive material assembly. In comparison with the time-domain electromagnetic method (TEM) and controlled source electromagnetic methods (such as controlled sources audio magnetotelluric sounding (CSAMT) and wide-field electromagnetic method (WFEM)), magnetotelluric (MT) sounding is the primary geophysical method that can image the electrical properties of the lithosphere–asthenosphere system. At present, the geophysical electromagnetic methods play an important role in the exploration of copper, molybdenum, lead, zinc, bauxite, uranium and other metal mineral resources all over the world (see more details in Section 6).

3. Construction of the Lithospheric Electrical Structure

3.1. Magnetotelluric Sounding and Data Processing

Magnetotelluric (MT) sounding is a passive electromagnetic geophysical method, which can measure signals in a broad frequency of the electromagnetic field varying from 105 Hz to 10−5 Hz, and includes audio magnetotelluric (AMT) sounding, broad-band magnetotelluric (BMT) sounding and long-period magnetotelluric (LMT) sounding [23]. High-frequency signals (>1 Hz) originate from worldwide lightning storms, and low-frequency signals (<1 Hz) are derived from the interactions of solar wind with the Earth’s magnetosphere. In a uniformly half-space medium, the penetration depth of EM fields is described by skin depth δ = 503 ρ / f (m), where ρ denotes the subsurface resistivity and f is the frequency (Figure 2).
MT sounding can be used to image the subsurface electrical structure by measuring the time variation of natural electric (E) and magnetic (H) fields in the orthogonal horizontal direction (Ex, Ey, Hx, Hy; x and y are south–north and east–west directions) and the vertical direction (Hz) at the Earth’s surface (see Chave and Jones, 2012 [23]; Figure 3).
The time-domain electromagnetic signal is transformed into frequency-domain data by fast Fourier Transform (FFT). Subsequently, the frequency-domain MT data are processed with the least square method or statistically robust algorithm [27], and noise is reduced at the measurement locations by employing the remote reference method, which can be effective in improving the coherence and signal–noise ratio of the MT data [23]. The electric and magnetic field components are related by the complex impedance tensor Z (Equation (2)).
E x E y = Z x x Z x y Z y x Z y y H x H y = Z H x H y
The apparent resistivity ( ρ i j ) and phase ( ϕ i j ) can be calculated from four components of the impedance tensor matrix Z ,
ρ i j = 1 ω μ Z i j 2
ϕ i j = arctan ( Im Z i j Re Z i j )
where ω is the angular frequency and μ is the permeability; Im Zij and Re Zij are the imaginary and real parts of Zij, respectively; i and j are directions. The phase angle is commonly expressed in the first quadrant (0° ≤ ϕ ≤ 90°) such that it indicates whether apparent resistivity increases (ϕij < 45°) or decreases (ϕij > 45°) as frequency decreases.
Furthermore, under the condition of horizontal heterogeneity, the vertical component of the magnetic field is not equal to zero ( H z 0 ), and therefore, the ratio of horizontal and vertical components of magnetic field is defined as the tipper data (T),
H z = T x , T y H x H y
The tipper data include information about the spatial variation of resistivity; however, the absolute resistivity value needs to be constrained by the data related to the electric field. The tipper data can offer supplementary information to impedance data.

3.2. Magnetotelluric Data Analysis and Inversion

Following the processing of the MT data, some basic analyses of the MT data are required [27]. (1) Analysis of the phase tensor. This is mainly used to analyze the dimensionality of the MT data and the regional tectonic strike, which results from the advantage of not being affected by a three-dimensional distortion effect [28,29]. (2) Analysis of the impedance tensor. This is used to analyze and estimate the tectonic strike. Before the two-dimensional inversion of the MT data, the tectonic strike needs to be estimated in the research region, and then the MT data must be rotated according to the tectonic strike [30], as part of which the MT data can be divided into two perpendicular, independent polarization modes (transverse electric (TE) mode; transverse magnetic (TM) mode). (3) Analysis of the induction vector. This is sensitive to the horizontal variation in conductivity of the underground medium. Under Parkinon’s regulation, the arrows of induction vector point to the converging areas of current [31], which are the conductive zones, while under Wiese’s regulation [32], the arrows point to the high-resistance zones. The induction vector is mainly used to judge the positions of conductive zones, and can also evaluate the dimensionality. Under the ideal two-dimensional condition, the distribution of conductive zones must extend infinitely along the tectonic strike. (4) Analysis of the galvanic distortion. This results from the conductive inhomogeneity in the shallow area, which can give rise to frequency-independent field amplifications and rotations, and is also a key issue of the electromagnetic method [33]. Previous studies have proposed many methods to tackle this problem [33,34,35,36].
Subsequently, according to the different purposes of different research, the one-dimensional, two-dimensional and three-dimensional inversions of the MT data can be carried out to obtain the electrical resistivity model of the lithosphere. The nonlinear conjugate gradient algorithm (NLCG), limited-memory Broyden–Fletcher–Goldfarb–Shanno (L-BFGS) algorithm, OCCAM algorithm, base reduction OCCAM algorithm (REBOCC) and fast relaxation algorithm (RRI) are the main inversion algorithms [37,38,39,40]. Furthermore, although the natural electromagnetic (EM) signal can penetrate the resistive zones and detect the conductive zones, the EM signal cannot be easily used to determine the bottom boundary of the conductive zones because of the diffusion of the EM signal in the Earth’s interior [41,42]. Therefore, after the inversion, some sensitivity tests must be carried out to verify whether the MT data have good constraints on the electrical resistivity model.

4. Connecting the Electrical Resistivity Model of Lithosphere and Metallogenic System

Metallic minerals are mainly hosted in the rocks as metal sulfides (e.g., bornite, chalcocite, pyrite, sphalerite, galena, molybdenite) and metal oxides (e.g., cuprite, magnetite). It is generally known that the conductivity of metal-bearing rocks is much higher than that of the surrounding rocks. The structure of minerals in the rocks also has an influence on conductivity; when the metallic minerals have a good interconnection in the rocks (e.g., lumpy structure, vein-like structure, spotted structure), the conductivity increases sharply; when the metallic minerals have a bad interconnection in the rocks (e.g., dot structure, disseminated structure), conductivity increases a little, or the variation of conductivity may not be recognized in the electrical resistivity model. Furthermore, because the metallic minerals are sparsely distributed in the ore bodies, which results in the limited distribution of conductive anomalies, the metallic minerals can represent a possible factor affecting conductive zones in the shallow zones of crust, and struggle to produce large-scale conductive anomalies in the deep Earth [41,42]. Therefore, the conductive zones in the subsurface depicted by MT sounding in the high-frequency band can be interpreted as large-scale ore deposits with interconnected minerals [43,44].
Based on different migration mediums in the mineralization, the ore deposits can be divided into three types, that is, magmatic deposit, hydrothermal deposit and superficial deposit. The enrichment of metallic minerals in the magmatic deposits and hydrothermal deposits are formed by the filling process and metamorphism, while the superficial deposits are formed by either the reverse process of the selective removal of surface water from rocks of the Earth’s surface, or the enrichment of valuable elements in the solid residues through chemical dissolution and mechanical separation. Therefore, the fractures and fault zones in the rocks not only act as important migration channels of the medium, but are also filled with minerals as a vessel during mineralization. Generally, fractures and fault zones are indicated as zones of conduction from the subsurface to the deep-seated zone, or electrical gradient belts between the conductive zones and resistive zones, or electrical interval belts between the resistive zones [45,46].
Because magmatic and hydrothermal deposits mainly form in the background of magmatic events (e.g., the geological environment related to the intrusive bodies), the upwelling and evolution of the magmatic–hydrothermal materials play an essential role in the mineralization of magmatic–hydrothermal deposits. The multi-stage upwelling of the dense mafic magmas from the mantle wedge underplated crust forms the large-volume periodical magma chambers, and meanwhile, the MASH process (melting, assimilation, storage and homogenization) occurs at the crust–mantle boundary [47] (Figure 4). The magmatic–hydrothermal materials subsequently go through the processes of liquation and multi-stage penetration, and migrate into the shallow area to form metal-rich small-volume magma chambers, which act as a key point in the evolution of ore deposits [48]. Previous studies have reported that deep-seated magmatic–hydrothermal materials migrate upwards along the existing faults and weakened zones in the lithosphere [49], in which the gathering and evolution zones can also be identified as a high-rheological feature. With the decrease in temperature following mineralization, melts in the magma chambers and migration channels gradually cool and crystallize. In this process, fluids are released and migrate to the low-pressure zone above. Following current mineralization, the late tectonic–magmatic events may reactivate the existing weakened zones and fault zones in the lithosphere, which can facilitate the migration and gathering of magmatic–hydrothermal materials [46].
During the mineralization of magmatic–hydrothermal deposits, the volatiles (e.g., chlorine, sulfur, carbon dioxide, methane, water) in the magmatic–hydrothermal materials also play an essential role in the mineralization of different deposits [5,50,51,52,53]. For example, the concentration of Cl ions can increase the distribution ratio of metal elements in the fluid [50,54]. Furthermore, the alkali-rich, volatile-rich and/or water-rich melts (e.g., anatexis in the crust) also contribute to most types of ore deposits [6,53,54,55,56,57,58,59]. During or after mineralization, these volatile residues may form a chlorate, sulfide and/or graphite film under specific temperature and pressure conditions, which can increase the conductivity of the rock system [13,14,60]. The ore deposits form in a specific geological time and the metallogenic events are transient; however, the most important features of metallogenic systems are shaped before the mineralization or a long period after the metallogenic event, and the evolution of metallogenic systems may be preserved for hundreds of millions to billions of years [2,12]. Because the large-scale tectonic feature depicted by MT sounding may not change a lot over a long historical period, the conductive zones in the lithosphere can be interpreted as electrical traces of residues related to mineralization, even though they may overlap with the latest magmatic activity. Therefore, an electrical resistivity model of the lithosphere can provide constraints on the migration of magmatic–hydrothermal materials from the source area to the subsurface shallow area.

5. Influence of the Rheological Structure on the Metallogenic System

  • Relationship of conductivity and melt/fluid fraction
As mentioned above, the migration of lithospheric materials is controlled by faults and weakened zones, which show up as high-conductive anomalies. Furthermore, the weakened zones are closely related to partial melting in the lithosphere. A petrophysical experiment reported that when the volume fractions of melts are more than 5%, the effective viscosity of rocks can be reduced by an order of magnitude; when the volume fractions are more than 20%, the effective viscosity may be reduced by at least two orders of magnitude [61,62,63,64].
Silicate melts and carbonate melts, both of which are formed by the partial melting of different rocks, are the main melts in the lithosphere [65,66]. The partial melting of carbonate is generated at depths of more than 300 km in the mantle, and carbonated melts are stable under the condition of more than 2.5 GPa, while the silicate melts are more likely to exist at depths of less than 180 km [67,68]. When the upwelling carbonate melts react with normally anhydrous minerals, melts change from carbonate melts to carbonate-bearing (CO2-bearing) silicate melts [69]. Furthermore, free fluids (alkali-bearing/saline fluids) are generally present in the porous sedimentary rocks, brittle fractures and shear zones in the mid-upper crust, and dehydrated areas [70], and the chlorate and bicarbonate are the main alkali-bearing fluids in the lithosphere. Because the conductive feature of bicarbonate is controlled by the content of carbon dioxide in the fluid, the saline (alkali-bearing) fluids can be simply described as the chlorate of potassic and sodium [70,71,72].
Although MT sounding struggles in obtaining the specific distribution of conductive zones, it can estimate the total volume of melt, that is, the volume fraction of melt [73]. Note that a small number of fluids can also greatly increase the bulk conductivity of rocks, and therefore, both the thin layer, with high volume fractions of melts, and the thick layer, with low volume fractions of melts, have the same bulk conductivity, and produce the same observation regarding MT sounding [23].
2.
Mixing Models in the Lithosphere
Rocks in the crust and upper mantle generally consist of some minerals and mixtures, the bulk conductivity of which is determined by the conductivity, percentage by volume (vol%), percentage by weight (wt. %) and spatial distribution of every component. Therefore, the volume fractions of melts or fluids can be estimated by the specific mixed model when the information above is known. Because of the occurrence of some uncertainties when calculating bulk conductivity in the lithosphere, a variety of numerical models have been proposed to calculate the bulk conductivity of rocks in different geometric distributions. The mixing models include the parallel model, the perpendicular model, the Hashin–Shtrikman bound model, Archie’s law model, the random model, the Waff model, the modified brick model, the Lichtenecker–Rother equation model and the Bussian model [74,75,76,77,78,79,80,81], of which the Hashin–Shtrikman bound model and Archie’s law model the two most common models.
The Hashin–Shtrikman bound model (Equation (6)) [75] can be divided into the Hashin–Shtrikman upper (HS+) and lower bounds (HS−). The HS+ bound model is a model wherein the melt phase is completely connected and contains low-conductive spherical grains. In contrast, the HS− bound model is the resistive aggregate containing discontinuous conductive minerals.
σ H S ± [ i = 1 N v i σ i + 2 σ max , min ] 1 2 σ max , min
where σ H S ± are the maximum and minimum conductivity in the multi-phase system, N is the total number of minerals, v i is the volume percentage of every mineral, σ i is the conductivity of every mineral, and σmax,min are the maximum and minimum minerals.
Archie’s law (Equation (7)) [74,80,81] is a method that can be used to connect the conductivity of a clean reservoir rock with its porosity and the conductivity of its pore fluid. The generalized form can be expressed as:
σ e f f = i = 1 n σ i Φ i m i
where σ e f f is the bulk conductivity, σ i is the melt conductivity of every phase, Φ i is the percentage, and mi is the cementation exponent.
Therefore, based on the electrical resistivity model of the lithosphere, the volume fractions of melts or fluids are estimated according to different mixing models.
3.
Rheological constraints from electrical conductivity
Besides the rheological estimation of the lithosphere based on the relationship of conductivity and volume fractions of melts or fluids, the conductivity–viscosity relationship is also simplified as a ratio relationship between these two factors. Liu and Hasterok (2016) [82] put forward an empirical formula for the conductivity–viscosity relationship (Equation (8)):
η e f f η 0 = C 0 ρ ρ 0 C 1
The parameters are defined as follows: η e f f is the effective viscosity, η 0 is the reference viscosity, ρ is the measured resistivity, ρ 0 is the reference resistivity, and both C0 and C1 are controlling coefficients [82]. Here the reference viscosity is taken to be 1020 Pa·s and the reference resistivity is taken to be 100 Ω·m. The controlling coefficients are estimated over a narrow range and act in the following way: C0 is the controlling coefficient of the reference resistivity (values possibly in the range of 0.5–8.0), and C1 controls the maximum contrast of the viscosity (values possibly in the range of 0.5–1.5) [82]. With the systematic variation of C0 and C1, the observed conductivity can match the viscosity of the lithosphere. In addition, in the upper mantle, C0 and C1 can be set as 1 and 0.6667, respectively [83].
Certainly, the presence of partial melting can cause a correlating response in both resistivity and viscosity, with an increasing melt fraction leading to lower values of both. However, some materials with a conductive phase may be present in the crust that lower the resistivity but have a small (or no) impact on the viscosity. Examples include graphite films and sulfide minerals (important specifically in the shallow crust), both of which are important specifically in the upper crust.
The strength of rocks, which is related to a variation in viscosity, determines whether the “channel flow” occurs under the given boundary conditions, and at what rate the deformation occurs [84,85]. The “channel flow” in the partial melting zones can be represented as the Couette flow and/or Poiseuille flow. Couette flow is the laminar flow of a viscous fluid in the space between two parallel plates that are moving relative to each other. Poiseuille flow is pressure-induced flow (“channel flow”) in a layer, usually a pipe, and is distinct from drag-induced flow such as Couette flow.
In brief, the weakened zones characterized by low viscosity in the lithosphere can provide the channels of migration and the emplacement of the magmatic–hydrothermal materials, and can also influence the stress condition of the overlying rigid stratum. Therefore, rocks in the overlying stratum may locally deform, which has an impact on the exposure state and distribution feature of deposits following mineralization.

6. Typical Cases

Based on the temporal and spatial coupling relationships of other geophysical, geological and geochemical data, the electrical resistivity model obtained by MT sounding has been successfully applied in studying the metallogenic dynamics of ore deposits, and the metallogenic regularity and models of some typical ore deposits (e.g., multi-store metallogenic model [12]) have also been summarized. In the following section, three typical cases are shown that depict the relationships between the electrical resistivity models of the lithosphere and the metallogenic system.
In Australia [86,87], a cross-sectional electrical resistivity model to a depth of 60 km in the Olympic Dam is shown in Figure 5. Based on the depth and shape of the conductive zone C3 (~1 S/m), the brittle–ductile transition zone is approximately 15 km, and three narrow low-resistance zones above this zone branch to the surface. Heinson et al. (2018) [87] reported that lithospheric imaging helps portray a deep metallogenic system and map some of the pathways of metalliferous fluids moving from crust–mantle sources to discrete locations in the surface.
For southeastern China [88], an electrical resistivity model produced from a 3D MT data array in the Nanling–Xuancheng area is shown in Figure 6. The conductive zones C7 and C8 in the mid–lower crust can be attributed to hydrous fluids that are possibly related to partial melting. The upwelling fluids create a weak zone at the ductile–brittle transition (conductive zones C1–C5), and provide the conditions for crustal decoupling caused by the NW-directional compression of the Paleo-Pacific plate. The fluids migrate through NE–SW-directional channels and interact with the upper crustal rocks. Therefore, according to the electrical structure and rheological parameters as indicated by electrical conductivity, a new mechanism is offered, referred to as “crustal decoupling”, which explains the tectonic–magmatic history of the magmatic metallogenic system in the Nanling–Xuancheng area.
In the context of North China, a multi-scale 3D electrical structure of the lithosphere of the Caosiyao porphyry Mo ore district is obtained by combining the joint inversion of the AMT and MT data [44] with the electrical structure in the lithospheric mantle previously derived by Zhang et al. (2016) [89]. The multi-scale electrical structure indicates a giant upper-mantle basaltic magma reservoir, indicated by the conductor DTHCZ beneath the Datong volcanic area, which moves through a lithosphere-scale shear zone, the mid–lower crustal magma chamber CZ3, the crust-scale weak zone, and a conduit (CC1 and CC2) to transport ore-forming magmas and fluids (Figure 7).
Other typical results include the Ning (Nanjing)-Wu (Wuhu) basin in eastern China [90], the Shihu–Xishimen gold–iron ore district in North China [91], the Middle–lower Yangtze Metallogenic Belt (including Ningwu, Luzong, Xuancheng, Anqing, Guichi, and Tongling areas) in East China [12,92,93,94,95,96], the Jiaodong gold deposits in northern China [97], the Liaodong Qingchengzi orefield [98], the Yixingzhai gold deposits (Boqiang Cu–Mo–Au deposits and Nanling W-Sn ore district in China) [45,99,100], the northeastern Jiangxi metallogenic province in southeastern China [101], the Caosiyao porphyry Mo ore district in North China [43], the Beiya Cu ore district in southwestern China [46], the large-scale Hatu epithermal gold deposit in western Junggar, NW China [102], Zhaxikang in the Tethys–Himalaya area in southwestern China [103,104], the Baogutu porphyry copper–gold deposit, the Hongqiling Cu–Ni sulfide intrusions in the Central Asian Orogenic Belt [105], the Xiangshan volcanogenic uranium deposit [106,107], the Narusongduo Pb–Zn–Fe–Cu ore district, and the Qulong–Jima porphyry ore Cu deposit in southwestern China [41,42]. In the Olympic Dam mine and Tennant ore district in Australia [108], the Tsagaan Tsahir Uul Au deposit in southern Mongolia [109], the Norrbotten district in northern Sweden [110,111], and the orogenic gold district in the Red Lake greenstone belt, western Superior craton, Canada [51], the electrical resistivity models obtained by MT sounding provide better geophysical evidence of the metallogenic dynamics of metallic mineral deposits.

7. Summary

In recent years, the relationship between the metallogenic system and the multi-scale electrical structure obtained by MT sounding in the lithosphere has become a hot topic, and a frontier in the academic field. Based on many geological and geochemical studies, geophysical imaging, especially regarding geoelectrical structures, can be used to sketch the migration channels of materials and ancient magma chambers from a macro perspective. Meanwhile, the temporal and spatial coupling relationship of the geological, geophysical and geochemical data contribute to enriching our understanding of the metallogenic background and metallogenic mechanisms. Aside from metallic mineral deposits, MT sounding can also be used in the prospecting of nonmetallic deposits, such as gas hydrate and natural gas [112].
Furthermore, some suggestions regarding the relationship between the electrical structure and metallogenic system can also be given, as follows: (a) how to obtain a high-resolution, multi-scale electrical resistivity model; (2) how to synthetically analyze the different parameters obtained by different electrical methods; (3) how to improve the temporal and spatial coupling of geological, geophysical and geochemical data; (4) how to choose a mixed model of the lithosphere and the values of different parameters used for the estimation of conductivity and melt/fluid volume fraction; (5) how to improve the interpretation of coupling mechanisms in the deep and shallow lithosphere based on an electrical resistivity model; (6) how to perform a petrophysical experiment on specific rocks and minerals.
Finally, as Engels said in the «Dialectics of Nature»: “We can understand the issue under the conditions of The Times, and the extents we understand depend on what these conditions reach”. This paper seeks to throw out a minnow so as to catch a whale.

Author Contributions

Conceptualization, S.J. and Y.S.; validation, S.J.; resources, S.J., Y.S., J.J., L.Z., H.D., Y.Y. and C.X.; writing—original draft, S.J., Y.S. and C.L.; writing—review & editing, S.J. and Y.S.; supervision, W.W. and G.Y.; project administration, S.J., W.W. and G.Y.; funding acquisition, Y.S. and S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study is funded by the National Natural Science Foundation of China (No. 42304087), National Key R&D Program of China (2022YFF0800901, 2016YFC0600301, 2018YFC0604104), Second Tibetan Plateau Scientific Expedition and Research Program (STEP) (2019QZKK0701) and China Scholarship Council (202006400054).

Data Availability Statement

Not applicable.

Acknowledgments

Special thanks to the MT team at the China University of Geoscience, Beijing (CUGB); this study would not have been possible without their efforts.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Blewett, R.S.; Henson, P.A.; Roy, I.G.; Champion, D.C.; Cassidy, K.F. Scale-integrated architecture of a world-class gold mineral system: The Archaean eastern Yilgarn Craton, Western Australia. Precambrian Res. 2010, 183, 230–250. [Google Scholar] [CrossRef]
  2. Kirkby, A.; Czarnota, K.; Huston, D.L.; Champion, D.C.; Doublier, M.P.; Bedrosian, P.A.; Duan, J.; Heinson, G. Lithospheric conductors reveal source regions of convergent margin mineral systems. Sci. Rep. 2022, 12, 8190. [Google Scholar] [CrossRef] [PubMed]
  3. Fyfe, W.; Price, N.; Thompson, A. Fluids in the Earth’s Crust: Their Significance in Metamorphic, Tectonic and Chemical Transport Processes; Elsevier Scientific Pub. Co.: Amsterdam, The Netherlands, 1978. [Google Scholar] [CrossRef]
  4. Zhang, D.H.; Jin, X.D.; Mao, X.D.; Wang, L.L. The classification of ore-forming fluid and the efficiency of ore formation of magmatic hydrothermal solution. Earth Sci. Front. 2011, 18, 90–102. [Google Scholar]
  5. Hou, Z.Q.; Duan, L.; Lu, Y.; Zheng, Y.; Zhu, D.; Yang, Z.; Yang, Z.; Wang, B.; Pei, Y.; Zhao, Z.; et al. Lithospheric Architecture of the Lhasa Terrane and Its Control on Ore Deposits in the Himalayan-Tibetan Orogen. Econ. Geol. 2015, 110, 1541–1575. [Google Scholar] [CrossRef]
  6. Hou, Z.Q.; Yang, Z.M.; Wang, R.; Zheng, Y.C. Further discussion on porphyry Cu-Mo-Au deposit formation in Chinese mainland. Earth Sci. Front. 2020, 27, 20–44. [Google Scholar] [CrossRef]
  7. Hou, Z.; Wang, Q.; Zhang, H.; Xu, B.; Yu, N.; Wang, R.; Groves, D.; Zheng, Y.; Han, S.; Gao, L.; et al. Lithosphere architecture characterized by crust-mantle decoupling controls the formation of orogenic gold deposits. Natl. Sci. Rev. 2023, 10, nwac257. [Google Scholar] [CrossRef]
  8. Deng, J.; Wang, Q.F.; Gao, L.; He, W.Y.; Yang, Z.Y.; Zhang, S.H.; Chang, L.J.; Li, G.J.; Xun, X.; Zhou, D.Q. Differential crustal rotation and its control on giant ore clusters along the eastern margin of Tibet. Geology 2021, 49, 428–432. [Google Scholar] [CrossRef]
  9. Dentith, M.; Yuan, H.; Johnson, S.; Murdie, R.; Piña-Varas, P. Application of deep-penetrating geophysical methods to mineral exploration: Examples from Western Australia. Geophysics 2018, 83, WC29–WC41. [Google Scholar] [CrossRef]
  10. Davies, R.S.; Groves, D.; Trench, A.; Dentith, M. Towards producing mineral resource-potential maps within a mineral systems framework, with emphasis on Australian orogenic gold systems. Ore Geol. Rev. 2020, 119, 103369. [Google Scholar] [CrossRef]
  11. Hagemann, S.; Lisitsin, V.; Huston, D. Mineral system analysis: Quo vadis. Ore Geol. Rev. 2016, 76, 504–522. [Google Scholar] [CrossRef]
  12. Lü, Q.; Meng, G.; Zhang, K.; Liu, Z.; Yan, J.; Shi, D.; Han, J.; Gong, X. The lithospheric architecture of the Lower Yangtze Metallogenic Belt, East China: Insights into an extensive Fe–Cu mineral system. Ore Geol. Rev. 2021, 132, 103989. [Google Scholar] [CrossRef]
  13. Naif, S.; Selway, K.; Murphy, B.S.; Egbert, G.; Pommier, A. Electrical conductivity of the lithosphere-asthenosphere system. Phys. Earth Planet. Inter. 2021, 313, 106661. [Google Scholar] [CrossRef]
  14. Lin, W.L.; Yang, B.; Han, B.; Hu, X.Y. A Review of Subsurface Electrical Conductivity Anomalies in Magnetotelluric Imaging. Sensors 2023, 23, 1803. [Google Scholar] [CrossRef] [PubMed]
  15. Unsworth, M.; Rondenay, S. Mapping the distribution of fluids in the crust and lithospheric mantle utilizing geophysical methods. In Metasomatism and the Chemical Transformation of Rock; Springer: Berlin/Heidelberg, Germany, 2013; pp. 535–598. [Google Scholar]
  16. Guo, X.; Wang, Q.; Li, B.; Zha, X.; Gong, B.; Ni, H. Electrical Conductivity of CO2 and H2O-Bearing Nephelinitic Melt. J. Geophys. Res. Solid Earth 2021, 126, e2020JB019569. [Google Scholar] [CrossRef]
  17. Morizet, Y.; Paris, M.; Gaillard, F.; Scaillet, B. Carbon dioxide in silica-undersaturated melt. Part I: The effect of mixed alkalis (K and Na) on CO2 solubility and speciation. Geochim. Cosmochim. Acta 2014, 141, 45–61. [Google Scholar] [CrossRef]
  18. Morizet, Y.; Paris, M.; Sifre, D.; Di Carlo, I.; Ory, S.; Gaillard, F. Towards the reconciliation of viscosity change and CO2-induced polymerization in silicate melts. Chem. Geol. 2017, 458, 38–47. [Google Scholar] [CrossRef]
  19. Moussallam, Y.; Morizet, Y.; Massuyeau, M.; Laumonier, M.; Gaillard, F. CO2 solubility in kimberlite melts. Chem. Geol. 2015, 418, 198–205. [Google Scholar] [CrossRef]
  20. Sifré, D.; Gardés, E.; Massuyeau, M.; Hashim, L.; Hier-Majumder, S.; Gaillard, F. Electrical conductivity during incipient melting in the oceanic low-velocity zone. Nature 2014, 509, 81–85. [Google Scholar] [CrossRef]
  21. Samrock, F.; Grayver, A.V.; Bachmann, O.; Karakas, Z.; Saar, M.O. Integrated magnetotelluric and petrological analysis of felsic magma reservoirs: Insights from Ethiopian rift volcanoes. Earth Planet. Sci. Lett. 2021, 559, 116765. [Google Scholar] [CrossRef]
  22. Evans, R.L.; Elsenbeck, J.; Zhu, J.; Abdelsalam, M.G.; Sarafian, E.; Mutamina, D.; Chilongola, F.; Atekwana, E.A.; Jones, A.G. Structure of the Lithosphere Beneath the Barotse Basin, Western Zambia, From Magnetotelluric Data. Tectonics 2019, 38, 666–686. [Google Scholar] [CrossRef]
  23. Chave, A.D.; Jones, A.G. The Magnetotelluric Method: Theory and Practice; Cambridge University Press: Cambridge, UK, 2012. [Google Scholar]
  24. Meju, M.A. Geoelectromagnetic Exploration For Natural Resources: Models, Case Studies And Challenges. Surv. Geophys. 2002, 23, 133–206. [Google Scholar] [CrossRef]
  25. Guo, Z.; Xue, G.; Liu, J.; Wu, X. Electromagnetic methods for mineral exploration in China: A review. Ore Geol. Rev. 2020, 118, 103357. [Google Scholar] [CrossRef]
  26. Rippe, D. Magnetotelluric Constraints on the Role of Fluids in Convergent Plate Boundaries; University of Alberta: Edmonton, Canada, 2012. [Google Scholar]
  27. Egbert, G.D. Robust multiple-station magnetotelluric data processing. Geophys. J. Int. 1997, 130, 475–496. [Google Scholar] [CrossRef]
  28. Caldwell, T.G.; Bibby, H.M.; Brown, C. The magnetotelluric phase tensor. Geophys. J. Int. 2004, 158, 457–469. [Google Scholar] [CrossRef]
  29. Booker, J.R. The Magnetotelluric Phase Tensor: A Critical Review. Surv. Geophys. 2014, 35, 7–40. [Google Scholar] [CrossRef]
  30. Groom, R.W.; Bailey, R.C. Decomposition of magnetotelluric impedance tensors in the presence of local three-dimensional galvanic distortion. J. Geophys. Res. Solid Earth 1989, 94, 1913–1925. [Google Scholar] [CrossRef]
  31. Parkinson, W.D. Directions of Rapid Geomagnetic Fluctuations. Geophys. J. Int. 1959, 2, 1–14. [Google Scholar] [CrossRef]
  32. Wiese, H. Geomagnetic Deep Tellurics Part II: The Strike Direction of the Background Structures of Electrical Resistance, Derived from Geomagnetic Variations. Geofis. Pura Appl. 1962, 52, 83–103. [Google Scholar] [CrossRef]
  33. Jiracek, G.R. Near-surface and topographic distortions in electromagnetic induction. Surv. Geophys. 1990, 11, 163–203. [Google Scholar] [CrossRef]
  34. Beamish, D.; Travassos, J. A study of static shift removal from magnetotelluric data. J. Appl. Geophys. 1992, 29, 157–178. [Google Scholar] [CrossRef]
  35. Jones, A.G. Static shift of magnetotelluric data and its removal in a sedimentary basin environment. Geophysics 1988, 53, 967–978. [Google Scholar] [CrossRef]
  36. Becken, M.; Burkhardt, H. An ellipticity criterion in magnetotelluric tensor analysis. Geophys. J. R. Astron. Soc. 2010, 159, 69–82. [Google Scholar] [CrossRef]
  37. Rodi, W.; Mackie, R.L. Nonlinear conjugate gradients algorithm for 2-D magnetotelluric inversion. Geophysics 2001, 66, 174–187. [Google Scholar] [CrossRef]
  38. Egbert, G.D.; Kelbert, A. Computational recipes for electromagnetic inverse problems. Geophys. J. Int. 2012, 189, 251–267. [Google Scholar] [CrossRef]
  39. Kelbert, A.; Meqbel, N.; Egbert, G.D.; Tandon, K. ModEM: A modular system for inversion of electromagnetic geophysical data. Comput. Geosci. 2014, 66, 40–53. [Google Scholar] [CrossRef]
  40. Key, K. Mare2dem: A 2-D inversion code for controlled-source electromagnetic and magnetotelluric data. Geophys. J. Int. 2016, 207, 571–588. [Google Scholar] [CrossRef]
  41. Sheng, Y.; Jin, S.; Comeau, M.J.; Dong, H.; Zhang, L.; Lei, L.; Li, B.; Wei, W.; Ye, G.; Lu, Z. Lithospheric Structure Near the Northern Xainza-Dinggye Rift, Tibetan Plateau-Implications for Rheology and Tectonic Dynamics. J. Geophys. Res. Solid Earth 2021, 126, e2020JB021442. [Google Scholar] [CrossRef]
  42. Sheng, Y.; Jin, S.; Comeau, M.J.; Becken, M.; Zhang, L.; Dong, H.; Wei, W.; Ye, G. Controls on the metallogenesis of the Lhasa-Mozugongka district, Gangdese Belt, Tibetan Plateau: Constraints on melt distribution and viscosity from the 3-D electrical structure of the lithosphere. Ore Geol. Rev. 2022, 145, 104881. [Google Scholar] [CrossRef]
  43. Xu, L.Y.; Yin, Y.T.; Jin, S.; Wei, W.B.; Hao, Z.; Li, X.; Liu, H.; Li, H. CSAMT constraints on the metallogenic mechanism of the Zhaishang gold deposit, West Qinling, China. Explor. Geophys. 2021, 53, 359–371. [Google Scholar] [CrossRef]
  44. Xu, L.Y.; Jin, S.; Yin, Y.T.; Wei, W.B.; Ye, G.F.; Dong, H.; Zhang, L.T.; Jing, J.E.; Xie, C.L. Multiscale 3-D imaging of the crustal electrical structure beneath the Caosiyao porphyry Mo deposit, North China. Geophys. J. Int. 2022, 231, 1880–1897. [Google Scholar] [CrossRef]
  45. Yin, Y.T.; Jin, S.; Wei, W.; Santosh, M.; Dong, H.; Xie, C. Construction and destruction of the North China Craton with implications for metallogeny: Magnetotelluric evidence from the Hengshan–Wutai–Fuping region within Trans-North China Orogen. Gondwana Res. 2016, 40, 21–42. [Google Scholar] [CrossRef]
  46. Yu, N.; Wang, E.; Wang, X.; Kong, W.; Li, D.; Li, R. The Influence of the Ailaoshan-Red River Shear Zone on the Mineralization of the Beiya Deposit on the Southeastern Margin of the Tibetan Plateau Revealed by a 3-D Magnetotelluric Survey. J. Geophys. Res. Solid Earth 2022, 127, e2021JB022923. [Google Scholar] [CrossRef]
  47. Hildreth, W. Gradients in silicic magma chambers: Implications for lithospheric magmatism. J. Geophys. Res. 1981, 86, 10153–10192. [Google Scholar] [CrossRef]
  48. Tang, Z.L.; Qian, Z.Z.; Jiang, C.Y.; Yan, H.Q.; Jiao, J.G.; Liu, M.W.; Xu, Z.H.; Xu, G.; Wang, Y.L. Trends of research in exploration of magmatic sulfide deposits and small intrusions metallogenic system. J. Earth Sci. Environ. 2011, 33, 1–9. [Google Scholar] [CrossRef]
  49. Jin, S.; Sheng, Y.; Comeau, M.J.; Becken, M.; Wei, W.; Ye, G.; Dong, H.; Zhang, L. Relationship of the Crustal Structure, Rheology, and Tectonic Dynamics Beneath the Lhasa-Gangdese Terrane (Southern Tibet) Based on a 3-D Electrical Model. J. Geophys. Res. Solid Earth 2022, 127, e2022JB024318. [Google Scholar] [CrossRef]
  50. Zhao, Y.; Yu, J.; Xu, B.; Zhang, Y.; Wang, Z.; Hou, Z.; Kou, G.; Zhao, Z. Zircon Hf-O and apatite Sr-O isotopes jointly reveal the origin and mineralization of the Miocene adakitic porphyries in the Gangdese belt. Ore Geol. Rev. 2023, 159, 105536. [Google Scholar] [CrossRef]
  51. Adetunji, A.Q.; Launay, G.; Ferguson, I.J.; Simmons, J.M.; Ma, C.; Ayer, J.; Lafrance, B. Crustal conductivity footprint of the orogenic gold district in the Red Lake greenstone belt, western Superior craton, Canada. Geology 2023, 51, 377–382. [Google Scholar] [CrossRef]
  52. Xu, B.; Hou, Z.-Q.; Griffin, W.L.; O’Reilly, S.Y.; Zheng, Y.-C.; Wang, T.; Fu, B.; Xu, J.-F. In-situ mineralogical interpretation of the mantle geophysical signature of the Gangdese Cu-porphyry mineral system. Gondwana Res. 2022, 111, 53–63. [Google Scholar] [CrossRef]
  53. Deng, J.; Wang, Q.F. Gold mineralization in China: Metallogenic provinces, deposit types and tectonic framework. Gondwana Res. 2016, 36, 219–274. [Google Scholar] [CrossRef]
  54. Wang, R.; Weinberg, R.F.; Collins, W.J.; Richards, J.P.; Zhu, D.-C. Origin of postcollisional magmas and formation of porphyry Cu deposits in southern Tibet. Earth-Sci. Rev. 2018, 181, 122–143. [Google Scholar] [CrossRef]
  55. Wang, R.; Weinberg, R.F.; Zhu, D.-C.; Hou, Z.-Q.; Yang, Z.-M. The impact of a tear in the subducted Indian plate on the Miocene geology of the Himalayan-Tibetan orogen. GSA Bull. 2021, 134, 681–690. [Google Scholar] [CrossRef]
  56. Zhang, Z.-Y.; Hou, Z.-Q.; Lue, Q.-T.; Zhang, X.-W.; Pan, X.-F.; Fan, X.-K.; Zhang, Y.-Q.; Wang, C.-G.; Lue, Y.-J. Crustal architectural controls on critical metal ore systems in South China based on Hf isotopic mapping. Geology 2023, 51, 738–742. [Google Scholar] [CrossRef]
  57. Sun, W.; Ling, M.; Yang, X.; Fan, W.; Ding, X.; Liang, H. Ridge subduction and porphyry copper-gold mineralization: An overview. Sci. China Earth Sci. 2010, 53, 475–484. [Google Scholar] [CrossRef]
  58. Poty, B.; Holland, H.D.; Borcsik, M. Solution-mineral equilibria in the system MgO-SiO2-H2O-MgCl2 at 500 °C and 1 kbar. Geochim. Et Cosmochim. Acta 1972, 36, 1101–1113. [Google Scholar] [CrossRef]
  59. Hou, Z.; Wang, R.; Zhang, H.; Zheng, Y.; Jin, S.; Thybo, H.; Weinberg, R.F.; Xu, B.; Yang, Z.; Hao, A.-W.; et al. Formation of giant copper deposits in Tibet driven by tearing of the subducted Indian plate. Earth-Sci. Rev. 2023, 243, 104482. [Google Scholar] [CrossRef]
  60. Sheng, Y.; Jin, S.; Comeau, M.J.; Hou, Z.Q.; Becken, M.; Dong, H.; Zhang, L.T.; Wei, W.B.; Ye, G.F. Evidence for partial melting and alkali-rich fluids in the crust from a 3-D electrical resistivity model in the vicinity of the Coqen region, western Lhasa terrane, Tibetan Plateau. Earth Planet. Sci. Lett. 2023, 619, 118316. [Google Scholar] [CrossRef]
  61. Li, S.H.; Unsworth, M.J.; Booker, J.R.; Wei, W.; Tan, H.; Jones, A.G. Partial melt or aqueous fluid in the mid-crust of Southern Tibet? Constraints from INDEPTH magnetotelluric data. Geophys. J. Int. 2003, 153, 289–304. [Google Scholar] [CrossRef]
  62. Grotenhuis, S.M.T.; Drury, M.R.; Spiers, C.J.; Peach, C.J. Melt distribution in olivine rocks based on electrical conductivity measurements. J. Geophys. Res. 2005, 110, B12201. [Google Scholar] [CrossRef]
  63. Unsworth, M.J.; The INDEPTH-MT team; Jones, A.G.; Wei, W.; Marquis, G.; Gokarn, S.G.; Spratt, J.E. Crustal rheology of the Himalaya and Southern Tibet inferred from magnetotelluric data. Nature 2005, 438, 78–81. [Google Scholar] [CrossRef]
  64. Wei, W.B.; Jin, S.; Ye, G.F.; Deng, M.; Jing, J.E.; Unsworth MARTYN. The conductivity structure and rheology of the lithosphere in the southern Tibet—The result of the study of ultra-wide band magnetotelluric sounding. Sci. China Ser. D 2009, 39, 1591–1606. [Google Scholar] [CrossRef]
  65. Yoshino, T.; Laumonier, M.; McIsaac, E.; Katsura, T. Electrical conductivity of basaltic and carbonatite melt-bearing peridotites at high pressures: Implications for melt distribution and melt fraction in the upper mantle. Earth Planet. Sci. Lett. 2010, 295, 593–602. [Google Scholar] [CrossRef]
  66. Ni, H.W.; Keppler, H.; Behrens, H. Electrical conductivity of hydrous basaltic melts: Implications for partial melting in the upper mantle. Contrib. Miner. Pet. 2011, 162, 637–650. [Google Scholar] [CrossRef]
  67. Ni, H.; Hui, H.; Steinle-Neumann, G. Transport properties of silicate melts. Rev. Geophys. 2015, 53, 715–744. [Google Scholar] [CrossRef]
  68. Yang, X.Z. The origin of electrical anomalies in the uppermost mantle. Acta Petrol. Mineral. 2015, 32, 663–679. [Google Scholar]
  69. Dasgupta, R.; Mallik, A.; Tsuno, K.; Withers, A.C.; Hirth, G.; Hirschmann, M.M. Carbon-dioxide-rich silicate melt in the Earth’s upper mantle. Nature 2013, 493, 211–215. [Google Scholar] [CrossRef] [PubMed]
  70. Vlasov, K.; Keppler, H. Electrical Conductivity of KCl-H2O Fluids in the Crust and Lithospheric Mantle. J. Geophys. Res. Solid Earth 2022, 127, e2022JB024080. [Google Scholar] [CrossRef]
  71. Sakuma, H.; Ichiki, M. Electrical conductivity of NaCl-H2O fluid in the crust. J. Geophys. Res. Solid Earth 2016, 121, 577–594. [Google Scholar] [CrossRef]
  72. Guo, H.; Keppler, H. Electrical Conductivity of NaCl-Bearing Aqueous Fluids to 900 °C and 5 GPa. J. Geophys. Res. Solid Earth 2019, 124, 1397–1411. [Google Scholar] [CrossRef]
  73. Cordell, D.; Naif, S.; Troch, J.; Huber, C. Constraining Magma Reservoir Conditions by Integrating Thermodynamic Petrological Models and Bulk Resistivity from Magnetotellurics. Geochem. Geophys. Geosystems 2022, 23, e2022GC010455. [Google Scholar] [CrossRef]
  74. Archie, G.E. The electrical resistivity log as an aid in determining some reservoir characteristics. Trans. Am. Inst. Mech. Eng. 1942, 146, 54–67. [Google Scholar] [CrossRef]
  75. Hashin, Z.; Shtrikman, S. A Variational Approach to the Theory of the Effective Magnetic Permeability of Multiphase Materials. J. Appl. Phys. 1962, 33, 3125–3131. [Google Scholar] [CrossRef]
  76. Somerton, W.H. Thermal Properties and Temperature-Related Behavior of Rock/Fluid Systems; Elsevier: Amsterdam, The Netherlands, 1992. [Google Scholar]
  77. Luo, M.; Wood, J.R.; Cathles, L.M. Prediction of thermal conductivity in reservoir rocks using fabric theory. J. Appl. Geophys. 1994, 32, 321–334. [Google Scholar] [CrossRef]
  78. Yu, L.; Evans, R.L.; Edwards, R.N. Transient electromagnetic responses in seafloor with triaxial anisotropy. Geophys. J. Int. 1997, 129, 292–304. [Google Scholar] [CrossRef]
  79. Roberts, J.J.; Tyburczy, J.A. Partial-melt electrical conductivity: Influence of melt composition. J. Geophys. Res. 1999, 104, 7055–7065. [Google Scholar] [CrossRef]
  80. Glover, P.W.; Hole, M.J.; Pous, J. A modified Archie’s law for two conducting phases. Earth Planet. Sci. Lett. 2000, 180, 369–383. [Google Scholar] [CrossRef]
  81. Glover, P.W.J. A generalized Archie’s law for n phases. Geophysics 2010, 75, E247–E265. [Google Scholar] [CrossRef]
  82. Liu, L.; Hasterok, D. High-resolution lithosphere viscosity and dynamics revealed by magnetotelluric imaging. Science 2016, 353, 1515–1519. [Google Scholar] [CrossRef]
  83. Xu, Y.X.; Zhang, A.Q.; Yang, B.; Bao, X.W.; Wang, Q.Y.; Xia, J.H.; Yang, W.C. Bridging the connection between effective viscosity and electrical conductivity through water content in the upper mantle. Sci. Rep. 2018, 8, 1771. [Google Scholar] [CrossRef]
  84. Rippe, D.; Unsworth, M. Quantifying crustal flow in Tibet with magnetotelluric data. Phys. Earth Planet. Inter. 2010, 179, 107–121. [Google Scholar] [CrossRef]
  85. Le Pape, F.; Jones, A.G.; Unsworth, M.J.; Vozar, J.; Wei, W.; Jin, S.; Ye, G.; Jing, J.; Dong, H.; Zhang, L.; et al. Constraints on the evolution of crustal flow beneath Northern Tibet. Geochem. Geophys. Geosystems 2015, 16, 4237–4260. [Google Scholar] [CrossRef]
  86. Heinson, G.S.; Direen, N.G.; Gill, R.M. Magnetotelluric evidence for a deep-crustal mineralizing system beneath the Olympic Dam iron oxide copper-gold deposit, southern Australia. Geology 2006, 34, 573–576. [Google Scholar] [CrossRef]
  87. Heinson, G.S.; Didana, Y.; Soeffky, P.; Thiel, S.; Wise, T. The crustal geophysical signature of a world-class magmatic mineral system. Sci. Rep. 2018, 8, 10608. [Google Scholar] [CrossRef]
  88. Zhang, K.; Lü, Q.; Lan, X.; Guo, D.; Wang, Q.; Yan, J.; Zhao, J. Magnetotelluric evidence for crustal decoupling: Insights into tectonic controls on the magmatic mineral system in the Nanling–Xuancheng area, SE China. Ore Geol. Rev. 2021, 131, 104045. [Google Scholar] [CrossRef]
  89. Zhang, H.; Huang, Q.; Zhao, G.; Guo, Z.; Chen, Y.J. Three-dimensional conductivity model of crust and uppermost mantle at the northern Trans North China Orogen: Evidence for a mantle source of Datong volcanoes. Earth Planet. Sci. Lett. 2016, 453, 182–192. [Google Scholar] [CrossRef]
  90. Zhang, K.; Lü, Q.; Yan, J.; Hu, H.; Fu, G.; Luo, F. The three-dimensional electrical structure and metallogenic prospect of the Ning (Nanjing)-Wu (Wuhu) basin and the southern adjacent area in eastern China. J. Asian Earth Sci. 2019, 173, 304–313. [Google Scholar] [CrossRef]
  91. Li, Q.; Santosh, M.; Li, S.R. Stable isotopes and noble gases in the Xishimen gold deposit, central North China Craton: Metallogeny associated with lithospheric thinning and crust-mantle interaction. Int. Geol. Rev. 2013, 55, 1728–1743. [Google Scholar] [CrossRef]
  92. Chen, X.; Yan, J. 2D magnetotelluric imaging of the Anqing-Guichi ore district, Yangtze metallogenic belt, eastern China: An insight into the crustal structure and tectonic units. Phys. Earth Planet. Inter. 2016, 257, 1–11. [Google Scholar] [CrossRef]
  93. You, M.; Guo, D.; Lan, X.; Yang, X.; Wang, Q.; Wu, X. 3-D Magnetotelluric Resistivity Imaging of Nanling-Xuancheng Ore Concentration Area. IEEE Access 2021, 9, 68332–68342. [Google Scholar] [CrossRef]
  94. Tang, J.; Zhou, C.; Wang, X.; Xiao, X.; Lü, Q. Deep electrical structure and geological significance of Tongling ore district. Tectonophysics 2013, 606, 78–96. [Google Scholar] [CrossRef]
  95. Yan, J.; Zhang, K.; Jinhua, Z. Local uplift and copper metallogenic prospect in the Luzong ore district, Eastern China: Evidence from 3D magnetotelluric imaging. Explor. Geophys. 2019, 50, 22–30. [Google Scholar] [CrossRef]
  96. Wang, G.; Fang, H.; Qiu, G.; Pei, F.; He, M.; Du, B.; Bai, D.; Zhang, X.; Lv, Q.; Yuan, Y.; et al. Three-dimensional magnetotelluric imaging of the Eastern Qinhang Belt between the Yangtze Block and Cathaysia Block: Implications for lithospheric architecture and associated metallogenesis. Ore Geol. Rev. 2023, 158, 105490. [Google Scholar] [CrossRef]
  97. Zhang, K.; Lü, Q.T.; Yan, J.Z.; Hu, H.; Fu, G.; Shao, L. Crustal structure beneath the Jiaodong Peninsula, North China, revealed with a 3D inversion model of magnetotelluric data. J. Geophys. Eng. 2018, 15, 2442–2454. [Google Scholar] [CrossRef]
  98. An, Z.; Di, Q.; Qu, W.; Zhang, Y.; Han, Y. Shallow crustal electrical structure of the Qingchengzi orefield in Liaodong area revealed by three-dimensional inversion of Magnetotelluric data. J. Appl. Geophys. 2022, 202, 104650. [Google Scholar] [CrossRef]
  99. Yin, Y.T.; Jin, S.; Wei, W.B.; Lü, Q.T.; Ye, G.F.; En, J.; Zhang, L.; Dong, H.; Xie, C.L. Lithosphere structure and its implications for the metallogenesis of the Nanling Range, South China: Constraints from 3-D magnetotelluric imaging. Ore Geol. Rev. J. Compr. Stud. Ore Genes. Ore Explor. 2021, 131, 104064. [Google Scholar] [CrossRef]
  100. Zheng, H.; Yin, Y.; Jin, S.; Wei, W.; Xu, L.; Qi, P.; Wang, H.; Wang, Q. Three-Dimensional Lithospheric Electrical Structure beneath the Handan-Xingtai District, North China: Implications for Tectonic Control of Skarn-Iron Mineralization. Minerals 2022, 13, 14. [Google Scholar] [CrossRef]
  101. Hu, H.Y.; Ye, G.F.; Dong, J.E.; Wei, W.B.; Jin, S. 3-D lithospheric conductivity structures in the Cathaysia Block and the Jiangnan suture zone: Implications for origins of metallogenic belts. J. Appl. Geophys. 2020, 177, 104045. [Google Scholar] [CrossRef]
  102. Xu, Y.X.; Yang, B.; Wu, S.C.; Shi, Y.; Bao, X.W.; Liu, Y.; Xia, Q.K. Metal preconcentration for gold mineralization in arcs: Geophysical observations from Western Junggar, NW China. Ore Geol. Rev. 2021, 139, 104562. [Google Scholar] [CrossRef]
  103. Fan, W.X.; Li, G.M.; Liang, S.X. Electrical Characteristics of Ore-controlling structures and prospecting prediction of Zhaxikang Mining are. J. Jilin Univ. Earth Sci. Ed. 2023, 51, 1709–1719. [Google Scholar]
  104. Xue, S.; Lu, Z.; Li, W.; Liang, H.; Wang, G.; Wang, H.; Li, H.; Li, X. Three-dimensional electrical resistivity structure beneath the Cuonadong dome in the Northern Himalayas revealed by magnetotelluric data and its implication. Sci. China Earth Sci. 2022, 65, 1538–1553. [Google Scholar] [CrossRef]
  105. Xu, Z.-H.; Gu, G.-W.; Jiang, J.-Y.; Li, F.-D.; Niu, X.-G. Metallogenic model of the Hongqiling Cu-Ni sulfide intrusions, Central Asian Orogenic Belt: Insight from long-period magnetotellurics. Open Geosci. 2023, 15, 20220430. [Google Scholar] [CrossRef]
  106. Deng, J.; Yu, H.; Chen, H.; Du, Z.; Yang, H.; Li, H.; Xie, S.; Chen, X.; Guo, F. Ore-controlling structures of the Xiangshan volcanic Basin, SE China: Revealed from three-dimensional inversion of Magnetotelluric data. Ore Geol. Rev. 2020, 127, 103807. [Google Scholar] [CrossRef]
  107. Yu, H.; Deng, J.; Tang, B.; Egbert, G.; Chen, H. Electrical image of magmatic system beneath the Xiangshan volcanogenic uranium deposit, southeast China: Linking magmatic evolution and uranium metallogenesis. Geology 2023, 51, 870–874. [Google Scholar] [CrossRef]
  108. Jiang, W.P.; Duan, J.M.; Doublie, M.; Clark, A.; Schofield, A.; Brodoe, R.C.; Goodwin, J. Application of multiscale magnetotelluric data to mineral exploration: An example from the east Tennant region, Northern Australia. Geophys. J. Int. 2022, 229, 1628–1645. [Google Scholar] [CrossRef] [PubMed]
  109. Comeau, M.J.; Kaufl, J.S.; Becken, M.; Kuvshinov, A.; Grayver, A.; Kamm, J.; Demberel, S.; Sukhbaatar, U.; Batmagnai, E. Evidence for fluid and melt generation in response to an asthenospheric upwelling beneath the Hangai Dome, Mongolia. Earth Planet. Sci. Lett. 2018, 487, 201–209. [Google Scholar] [CrossRef]
  110. Juanatey, M.A.G.; Hübert, J.; Tryggvason, A.; Juhlin, C.; Pedersen, L.B.; Bauer, T.E.; Dehghannejad, M. 2D and 3D MT in the central Skellefte Ore District, northern Sweden. Tectonophysics 2019, 764, 124–138. [Google Scholar] [CrossRef]
  111. Vadoodi, R.; Rasmussen, T.M.; Smirnov, M.; Bauer, T. Towards an understanding of mineral systems—Contributions from magnetotelluric data from the Fennoscandian Shield in northern Sweden. Tectonophysics 2021, 808, 228816. [Google Scholar] [CrossRef]
  112. Unsworth, M. New developments in conventional hydrocarbon exploration with electromagnetic methods. CSEG Rec. 2005, 30, 34–38. [Google Scholar]
Figure 2. Cartoon of the skin depth of the electromagnetic waves propagating in the Earth’s interior [26]. Black lines refer to the skin depths of different angular frequencies ω (or frequencies f), while red lines indicate the attenuation of the electromagnetic wave energy.
Figure 2. Cartoon of the skin depth of the electromagnetic waves propagating in the Earth’s interior [26]. Black lines refer to the skin depths of different angular frequencies ω (or frequencies f), while red lines indicate the attenuation of the electromagnetic wave energy.
Minerals 14 00541 g002
Figure 3. Map of the data acquisition approach of magnetotelluric sounding.
Figure 3. Map of the data acquisition approach of magnetotelluric sounding.
Minerals 14 00541 g003
Figure 4. Sketching the MASH (melting, assimilation, storage and homogenization) process (modified by Hildreth, 1981) [47].
Figure 4. Sketching the MASH (melting, assimilation, storage and homogenization) process (modified by Hildreth, 1981) [47].
Minerals 14 00541 g004
Figure 5. 2D electrical resistivity model of the Olympic Dam mine, Australia (Modified from Heinson et al., 2018 [87]). (a) Location of the regional tectonic boundaries and regional mineral distribution in the Olympic Dam mine. Blue triangles mean broadband MT sites, black circles mean long-period MT sites, and white squares mean magnetometer-only sites. Yellow asterisks are ore deposits (names are marked in black). (b) The 2D electrical model to a depth of 60 km. (c) The electrical model to a depth of 20 km in the box of panel b. WW, OD and VC are the Wirrda Well, the Olympic Dam and Vulcan, respectively, as shown in panel a. C1–C3: conductive zones. R1 and R2: resistive zones. The very high resistive zones R1 and R2 are the Archean Gawler Craton and Proterozoic mobile belts, respectively. The conductive zone C2, which extends from conductive zone C3 to the surface, is indicated as a pathway.
Figure 5. 2D electrical resistivity model of the Olympic Dam mine, Australia (Modified from Heinson et al., 2018 [87]). (a) Location of the regional tectonic boundaries and regional mineral distribution in the Olympic Dam mine. Blue triangles mean broadband MT sites, black circles mean long-period MT sites, and white squares mean magnetometer-only sites. Yellow asterisks are ore deposits (names are marked in black). (b) The 2D electrical model to a depth of 60 km. (c) The electrical model to a depth of 20 km in the box of panel b. WW, OD and VC are the Wirrda Well, the Olympic Dam and Vulcan, respectively, as shown in panel a. C1–C3: conductive zones. R1 and R2: resistive zones. The very high resistive zones R1 and R2 are the Archean Gawler Craton and Proterozoic mobile belts, respectively. The conductive zone C2, which extends from conductive zone C3 to the surface, is indicated as a pathway.
Minerals 14 00541 g005
Figure 6. Cartoon of the strain and fluid movement in the crust based on the electrical resistivity model in the Nanling–Xuancheng area, SE, China (Modified from Zhang et al., 2021 [88]). (a) Location of the middle–lower Yangtze metallogenic belt in China. (b) Location of the regional tectonic boundaries and regional mineral distribution in the green box of panel a. (c) Geological map and major deposits in the Nanling–Xuancheng district in the box of panel b. The blue rectangles are the magnetotelluric (MT) stations. (d) Cross-sections of the electrical resistivity model. Locations of the AA’, BB’, CC’, and DD’ profiles are in panel c. C1–C8: conductive zones. R1–R5: resistive zones. The white dashed lines indicate crustal discontinuity and the boundary between R1 and R5, while the black dashed lines indicate the brittle–ductile transition at a depth of approximate 10 km (overlying C7 + C8).
Figure 6. Cartoon of the strain and fluid movement in the crust based on the electrical resistivity model in the Nanling–Xuancheng area, SE, China (Modified from Zhang et al., 2021 [88]). (a) Location of the middle–lower Yangtze metallogenic belt in China. (b) Location of the regional tectonic boundaries and regional mineral distribution in the green box of panel a. (c) Geological map and major deposits in the Nanling–Xuancheng district in the box of panel b. The blue rectangles are the magnetotelluric (MT) stations. (d) Cross-sections of the electrical resistivity model. Locations of the AA’, BB’, CC’, and DD’ profiles are in panel c. C1–C8: conductive zones. R1–R5: resistive zones. The white dashed lines indicate crustal discontinuity and the boundary between R1 and R5, while the black dashed lines indicate the brittle–ductile transition at a depth of approximate 10 km (overlying C7 + C8).
Minerals 14 00541 g006
Figure 7. Lithospheric electrical resistivity model of the Caosiyao ore concentration district (modified from Xu et al., 2022 [44]). (a) Location of the regional tectonic boundaries and regional mineral distribution in the Caosiyao porphyry Mo ore district in North China. Faults F1, F2 and F3 are the Datong–Shangyi fault, Shangdu–Xinghe–Yuxian fault and Daihai–Huangqihai fault, respectively. (b) Electrical structure in the crust (modified from Xu et al., 2022 [44]). Black dots are the MT stations in panel a. (c) Topography map showing major tectonic structures and MT station locations in the northern Trans North China Orogen, from Zhang et al. (2016) [89]. SX Rift: Shanxi Rift. THS Uplift: Taihangshan Uplift. BJ: Beijing. DT: Datong. HNB: Hannuoba. ZJK: Zhangjiakou. (d) Electrical structure in the mantle modified from Xu et al. (2022) [44] and Zhang et al. (2016) [89]. Black dots are the MT stations in panel c. The rectangular box is the location of panel b. CC1, CZ2, CZ3, DTHCZ, and MLCZ: conductive zones. RZ1, RZ2: resistive zones.
Figure 7. Lithospheric electrical resistivity model of the Caosiyao ore concentration district (modified from Xu et al., 2022 [44]). (a) Location of the regional tectonic boundaries and regional mineral distribution in the Caosiyao porphyry Mo ore district in North China. Faults F1, F2 and F3 are the Datong–Shangyi fault, Shangdu–Xinghe–Yuxian fault and Daihai–Huangqihai fault, respectively. (b) Electrical structure in the crust (modified from Xu et al., 2022 [44]). Black dots are the MT stations in panel a. (c) Topography map showing major tectonic structures and MT station locations in the northern Trans North China Orogen, from Zhang et al. (2016) [89]. SX Rift: Shanxi Rift. THS Uplift: Taihangshan Uplift. BJ: Beijing. DT: Datong. HNB: Hannuoba. ZJK: Zhangjiakou. (d) Electrical structure in the mantle modified from Xu et al. (2022) [44] and Zhang et al. (2016) [89]. Black dots are the MT stations in panel c. The rectangular box is the location of panel b. CC1, CZ2, CZ3, DTHCZ, and MLCZ: conductive zones. RZ1, RZ2: resistive zones.
Minerals 14 00541 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jin, S.; Sheng, Y.; Liu, C.; Wei, W.; Ye, G.; Jing, J.; Zhang, L.; Dong, H.; Yin, Y.; Xie, C. A Review of Relationship between the Metallogenic System of Metallic Mineral Deposits and Lithospheric Electrical Structure: Insight from Magnetotelluric Imaging. Minerals 2024, 14, 541. https://doi.org/10.3390/min14060541

AMA Style

Jin S, Sheng Y, Liu C, Wei W, Ye G, Jing J, Zhang L, Dong H, Yin Y, Xie C. A Review of Relationship between the Metallogenic System of Metallic Mineral Deposits and Lithospheric Electrical Structure: Insight from Magnetotelluric Imaging. Minerals. 2024; 14(6):541. https://doi.org/10.3390/min14060541

Chicago/Turabian Style

Jin, Sheng, Yue Sheng, Chenggong Liu, Wenbo Wei, Gaofeng Ye, Jianen Jing, Letian Zhang, Hao Dong, Yaotian Yin, and Chengliang Xie. 2024. "A Review of Relationship between the Metallogenic System of Metallic Mineral Deposits and Lithospheric Electrical Structure: Insight from Magnetotelluric Imaging" Minerals 14, no. 6: 541. https://doi.org/10.3390/min14060541

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

Article Metrics

Back to TopTop