ArticlePDF Available

Abstract and Figures

As iron‐bearing minerals—ferrimagnetic minerals in particular—are sensitive to stress, temperature, and presence of fluids in fault zones, their magnetic properties provide valuable insights into physical and chemical processes affecting fault rocks. Here, we review the advances made in magnetic studies of fault rocks in the past three decades. We provide a synthesis of the mechanisms that account for the magnetic changes in fault rocks and insights gained from magnetic research. We also integrate nonmagnetic approaches in the evaluation of the magnetic properties of fault rocks. Magnetic analysis unveils microscopic processes operating in the fault zones such as frictional heating, energy dissipation, and fluid percolation that are otherwise difficult to constrain. This makes magnetic properties suited as a “strain indicator,” a “geothermometer,” and a “fluid tracer” in fault zones. However, a full understanding of faulting‐induced magnetic changes has not been accomplished yet. Future research should focus on detailed magnetic property analysis of fault zones including magnetic microscanning and magnetic fabric analysis. To calibrate the observations on natural fault zones, laboratory experiments should be carried out that enable to extract the exact physicochemical conditions that led to a certain magnetic signature. Potential avenues could include (1) magnetic investigations on natural and synthetic fault rocks after friction experiments, (2) laboratory simulation of fault fluid percolation, (3) paleomagnetic analysis of postkinematic remanence components associated with faulting processes, and (4) synergy of interdisciplinary approaches in mineral‐magnetic studies. This would help to place our understanding of the microphysics of faulting on a much stronger footing.
Domain states of magnetite as a function of grain size and shapes, and examples for commonly used plots for high‐field magnetic measurements. (a) Illustration of the domain state categories for magnetite at room temperature and their relationship with shape and grain size of the magnetic particles (redrawn from Roberts et al., 2018). (b) Example of a hysteresis loop with definition of Mrs, Ms, Bc, and Bcr, with the inset showing a backfield demagnetization curve, with the definition of Bcr. The high‐field slope (typically, B > 0.7 T) on the hysteresis loop is defined as high‐field magnetic susceptibility (χhf). The dark and gray lines are the hysteresis loops before and after correction for the nonferrimagnetic matrix contribution, respectively. In this example the correction is a paramagnetic correction as χhf is positive. (c) Example of a Day plot of the hysteresis ratios Mrs/Ms and Bcr/Bc. Single domain (SD), pseudo single domain (PSD), and multidomain (MD) boundaries are after Day et al. (1977). Note that recent Day plots often use vortex rather than PSD following Roberts et al. (2018). The data points shown are host and fault zone sediments from the frontal prism in Japan Trench cored by the Integrated Ocean Drilling Program Expedition 343, Japan Trench Fast Drilling Project (JFAST) (reproduced from Yang et al., 2018). The host sediments lie mainly in the PSD (or vortex) field, whereas most of the fault zone samples are located in the SD region. (d) Example of a FORC diagram (first‐order reversal curve) for one of the fault zone sediments (FZ697) shown in Figure 4c. Bc is equivalent to particle coercivity, and Bu to the local interaction field. Colors in the diagram represent absolute values of FORC density. The Bc peak centered at 30–40 mT with a prominent “kidney” shape toward higher coercivity suggests the occurrence of SD pyrrhotite. FORC data are from Yang et al. (2018) and are reprocessed with the FORCinel package (Harrison & Feinberg, 2008) with the VARIFORC (Egli, 2013) option used. VARIFORC smoothing parameters: vertical ridge Sc0 = 4, horizontal smoothing factor Sc1 = 7, central ridge Sb0 = 3, vertical smoothing factor Sb1 = 7, horizontal lambda λc = 0.1, and vertical lambda λb = 0.1. (e) Example of an isothermal remanent magnetization (IRM) acquisition curve. After application of a stepwise increasing magnetic field, the remanent magnetization increases until a maximum value is reached, which is termed saturation IRM (SIRM or Mrs). (f) Example of unmixing of an IRM acquisition curve to identify magnetic coercivity distributions using the MAX UnMix package (reprinted from Maxbauer et al., 2016). The sample is an anoxic lake sediment from Baldeggersee, Switzerland, and three coercivity components are identified: detrital magnetite‐low coercivity component (Component 1), biogenic magnetite (Component 2), and oxidized magnetite (or hematite)—Component 3.
… 
This content is subject to copyright. Terms and conditions apply.
Faulting Processes Unveiled by Magnetic Properties
of Fault Rocks
Tao Yang
1,2,3
,YuMin Chou
4,5
, Eric C. Ferré
6
, Mark J. Dekkers
2
, Jianye Chen
7,8
,
EnChao Yeh
9
, and Wataru Tanikawa
10
1
Hubei Subsurface MultiScale Imaging Key Laboratory, Institute of Geophysics and Geomatics, China University of
Geosciences, Wuhan, China,
2
Paleomagnetic Laboratory Fort Hoofddijk, Department of Earth Sciences, Utrecht
University, Utrecht, The Netherlands,
3
Now at School of Geophysics and Information Technology, China University of
Geosciences, Beijing, China,
4
Shenzhen Key Laboratory of Marine Archaea GeoOmics, Department of Ocean Science and
Engineering, South University of Science and Technology, Shenzhen, China,
5
Southern Marine Science and Engineering
Guangdong Laboratory, Department of Ocean Science and Engineering, Southern University of Science and Technology,
Shenzhen, China,
6
School of Geosciences, University of Louisiana at Lafayette, Lafayette, LA, USA,
7
State Key Laboratory
of Earthquake Dynamics, Institute of Geology, China Earthquake Administration, Beijing, China,
8
HPT Laboratory,
Department of Earth Sciences, Utrecht University, Utrecht, Netherlands,
9
Department of Earth Sciences, National Taiwan
Normal University, Taipei, Taiwan,
10
Japan Agency for MarineEarth Science and Technology, Kochi Institute for Core
Sample Research, Nankoku, Japan
Abstract As ironbearing mineralsferrimagnetic minerals in particularare sensitive to stress,
temperature, and presence of uids in fault zones, their magnetic properties provide valuable insights into
physical and chemical processes affecting fault rocks. Here, we review the advances made in magnetic
studies of fault rocks in the past three decades. We provide a synthesis of the mechanisms that account for
the magnetic changes in fault rocks and insights gained from magnetic research. We also integrate
nonmagnetic approaches in the evaluation of the magnetic properties of fault rocks. Magnetic analysis
unveils microscopic processes operating in the fault zones such as frictional heating, energy dissipation, and
uid percolation that are otherwise difcult to constrain. This makes magnetic properties suited as a
strain indicator,ageothermometer,and a uid tracerin fault zones. However, a full understanding of
faultinginduced magnetic changes has not been accomplished yet. Future research should focus on
detailed magnetic property analysis of fault zones including magnetic microscanning and magnetic fabric
analysis. To calibrate the observations on natural fault zones, laboratory experiments should be carried
out that enable to extract the exact physicochemical conditions that led to a certain magnetic signature.
Potential avenues could include (1) magnetic investigations on natural and synthetic fault rocks after
friction experiments, (2) laboratory simulation of fault uid percolation, (3) paleomagnetic analysis of
postkinematic remanence components associated with faulting processes, and (4) synergy of
interdisciplinary approaches in mineralmagnetic studies. This would help to place our understanding of the
microphysics of faulting on a much stronger footing.
Plain Language Summary The Earth's surface is riddled with faults that largely contribute to
landscape evolution and human activities. Some of these faults produce earthquakes of different magnitudes
including some with catastrophic consequences. Understanding faulting mechanisms benets society
when predictions about rupture are made. Fault zones preserve an excellent record of chemical and physical
processes involved in failure. Among other analytical methods, magnetic studies prove to be an
emergent and untapped source of information on these processes. These methods, focused on prefaulting,
synfaulting, and postfaulting mineral changes, have resulted in signicant advances in our
understanding of the conditions of faulting. In this review, we present an extensive account of the state of
knowledge and highlight current challenges and future avenues of fault magnetism research.
1. Introduction
Fault zones comprise only a very small volume of the Earth's crust. However, their structure and associated
deformation processes are ultimately decisive for a wide range of crustal processes (e.g., Faulkner
et al., 2010; Townend & Zoback, 2000). This pertains in particular to the physical origin of earthquakes
©2020. The Authors.
This is an open access article under the
terms of the Creative Commons
Attribution License, which permits use,
distribution and reproduction in any
medium, provided the original work is
properly cited.
REVIEW ARTICLE
10.1029/2019RG000690
Key Points:
Ironbearing minerals and
ferrimagnetic minerals in particular
are sensitive to faultingassociated
physical and chemical processes
Laboratory faulting experiments and
comparison with nonmagnetic
approaches conrm results from
magnetic studies on natural rocks
Rock magnetic methods offer
novel tools to analyze strain, grain
ning, temperature trends, and
uidrock interaction in fault zones
Correspondence to:
T. Yang and M. J. Dekkers,
tyang@cugb.edu.cn;
t.yang@uu.nl;
m.j.dekkers@uu.nl
Citation:
Yang, T., Chou, Y.M., Ferré, E. C.,
Dekkers, M. J., Chen, J., Yeh, E.C.,
& Tanikawa, W. (2020). Faulting
processes unveiled by magnetic
properties of fault rocks. Reviews of
Geophysics,58, e2019RG000690.
https://doi.org/10.1029/2019RG000690
Received 28 FEB 2020
Accepted 21 SEP 2020
Accepted article online 3 OCT 2020
YANG ET AL. 1of60
within the seismogenic zone, which typically extends from 34to1520 km depth for faults in the con-
tinental crust (e.g., Scholz, 2019; Sibson, 1986). Faultrelated rocks(Wise et al., 1984), or simply fault
rocks(Sibson, 1977), form as a result of localized strain within a fault zone (Brodie et al., 2007). These
rocks are common along upper crustal fault zones (Woodcock & Mort, 2008). Physical and chemical
attributes, and textures of fault rocks hold valuable information for understanding both the longterm
behavior of faults, on a timescale of millions of years, as well as the shortterm behavior involving the
nucleation (see Appendix A for denition), rupture, cessation, and recurrence of (large) earthquakes
(Bradbury et al., 2015; Faulkner et al., 2010; Henderson et al., 2010; Rowe & Grifth, 2015; Schmid &
Handy, 1991; Ujiie & Kimura, 2014; Wibberley et al., 2008). Studies into fault rocks thus have
intrigued geologists for decades and have been and continue to be an important subject in the context
of large, active faults with repeated seismic activity (Chester et al., 1993; Rowe & Grifth, 2015;
Sibson, 1986).
It is realized that a full appreciation of earthquake dynamics requires an integration of macroscopic seismol-
ogy with microscopic and experimental studies of fault zones (e.g., Cowan, 1999; Kanamori & Heaton, 2000;
K.F. Ma, 2009; Niemeijer et al., 2012; Scholz, 2019; Sibson, 1989). The former relies on earthquake kine-
matic analysis based on seismic waveform data (Kanamori & Heaton, 2000). The latter two mainly focus
on the physical/chemical attributes and texture of fault rocks, including their formation mechanisms
(K.F. Ma, 2009; Scholz, 2019). Their combination is crucial for understanding earthquake energy dissipa-
tion, rupture processes, and seismic efciency. Substantial research effort has been dedicated during the
last two decades to understanding the development and physical/chemical attributes of fault rocks using
diverse approacheseither of mesoscopic or microscopic scalein natural faults, laboratory experiments,
and numerical simulations (e.g., Billi, 2005; Mair & Abe, 2008; Rowe, Fagereng, et al., 2012; Rowe,
Kirkpatrick, & Brodsky, 2012). These studies have covered aspects of fault rocks in the widest possible sense,
including terminology and classication (Choi et al., 2016; Woodcock & Mort, 2008), development and
architecture (Faulkner et al., 2010; Fossen & Rotevatn, 2016; Pei et al., 2015; Wibberley et al., 2008), micro-
structure (e.g., Boullier et al., 2009; Bradbury et al., 2015; Fossen & Cavalcante, 2017; Isaacs et al., 2007;
Rowe & Grifth, 2015), grainsize distribution (e.g., Billi, 2005; Hattori & Yamamoto, 1999; Wilson
et al., 2005), mineral assemblages (e.g., Boullier, 2011; Bradbury et al., 2015; J. Chen et al., 2013; Isaacs
et al., 2007; Matsuda et al., 2004), geochemical composition (e.g., Bradbury et al., 2015; J. Chen et al., 2013;
Isaacs et al., 2007; Tanaka et al., 2001), hydraulic properties (e.g., Bense et al., 2013; Carpenter et al., 2014;
J. Chen et al., 2016; Faulkner et al., 2010), nanocrystallization (e.g., Verberne et al., 2019; Viti, 2011),
geochronology (e.g., Oriolo et al., 2018), and frictional properties (e.g., Boulton et al., 2017; Carpenter
et al., 2015; Di Toro et al., 2011; Niemeijer & Vissers, 2014).
A magnetic property analysis is occasionally included in the description of fault rock properties. Magnetic
properties are primarily governed by the distribution and speciation of ironthe fourth most abundant ele-
ment in the Earth's crust. By far most Febearing minerals are not considered magneticat room tempera-
ture; common minerals include Febearing silicates, carbonates (siderite and ankerite), and suldes (pyrite).
What we refer to as magnetic minerals are ferromagnetic (sensu lato) minerals, which retain permanent or
remanent magnetism; they include iron oxides (i.e., (titano)magnetite, maghemite, and hematite), oxyhydr-
oxides (primarily goethite), and some iron suldes (i.e., pyrrhotite and greigite). They occur only in trace
amounts in the rock but nonetheless determine its magnetic signature. Metallic iron and other reduced
phases such as iron phosphide, iron carbide, and iron silicide are also magnetic; they are extremely rare
in the Earth's crust. All aforementioned minerals, magnetic and nonmagnetic, are not exempt from
(thermo)chemical alterations induced by the faulting and related uidrock interaction, resulting in changes
in magnetic properties of fault rocks in comparison to adjacent host rocks. Thus, magnetic properties of fault
zones are natural archives of the faultingassociated processes in tectonically active regions. It makes rock
magnetism, which is the study of the magnetic properties of rocks, sediments, soils, and even organisms,
a promising tool to unravel faulting processes. Rock magnetic techniques are capable of determining the nat-
ure, grain size, and concentration of magnetic minerals in a sample down to the ppm level, which lies well
below the detection limit of more conventional mineralogical techniques, for example, Xray diffraction.
Also, magnetic measurements are generally rapid, costeffective, and nondestructive (evidently with the
exception of hightemperature magnetic analysis) so they can be used in conjunction with other techniques.
These advantages make rock magnetic analysis of fault rocks attractive.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 2of60
Fault rocks, however, have been the subject of rock magnetic studies only recently, with a focus on faults
from seismically active zones (e.g., Almqvist et al., 2020; Cai et al., 2019; Chou, Song, Aubourg, Lee,
et al., 2012; Chou, Song, Aubourg, et al., 2014; Chou, Song, Aubourg, Song, et al., 2012; Chou, Song, Lee,
et al., 2014; Ferré et al., 2005, 2012, 2015, 2016, 2017; Ferré, Gébelin, et al., 2014; Ferré, Geissman, et al., 2014;
Fukuchi, 2003; Fukuchi et al., 2005, 2007, 2009; D. Liu et al., 2014, 2016; Mishima et al., 2006,
2009; Nakamura et al., 2002; Pei, Li, et al., 2014; Pei, Zhou, et al., 2014; Tanikawa et al., 2007, 2008; Yang
et al., 2012a, 2012b, 2018, 2019; Yang, Chen, et al., 2013; Yang, Mishima, et al., 2013; Yang, Dekkers, &
Zhang, 2016; Yang, Yang, et al., 2016; Zhang et al., 2017, 2018). By providing a trove of information on phy-
sical and chemical processes associated with faulting in seismically active zones, these pioneer works have
laid the foundation of fault magnetism, an emerging subdiscipline in the geosciences.
This review concentrates on the rock magnetic studies on fault rocks from seismically active fault zones
where recent megaearthquakes occurred, such as the 1995 Mw (moment magnitude) 6.9 Kobe (Japan),
the 1999 Mw 7.6 ChiChi (Taiwan), the 2008 Mw 7.9 Wenchuan (China), and the 2011 Mw 9.0 Tohoku
(Japan) earthquakes. Additionally, magnetic studies on pseudotachylytes that are the products of fossil seis-
mogenic fault zones are included. First, we describe concisely the processes that result in the formation of
fault rocks. Second, we outline the approaches to evaluate magnetism of fault rocks. Third, we discuss the
mechanisms that are responsible for changes in the magnetism of fault rocks. Fourth, we address the
insights gained from magnetic property analysis of fault rocks and compare the magnetic methods with
other approaches, the core of this review. Thereafter, we identify some of the current challenges in fault mag-
netism research and propose potential avenues that would contribute to a full appreciation of the magnetism
of fault rocks and would advance our understanding of faulting.
2. Formation of Fault Rocks and Processes During Faulting
A fault is a narrow zone of crushed rocks along which two blocks of rock have moved alongside each other in
response to stress imposed on the rock. This movement, that is, the very faulting, takes place either as creep:
stable, slow sliding, or in the form of a series of earthquakes: unstable, fast slip intertwined with long periods
of no motion. Faults thus accommodate strain on a momentous range of dimensions: on a spatial scale, from
millimeters to hundreds of kilometers (for major plate boundaries), while also the time intervals range
widely from seconds to minutes of earthquake slip, to years of slow, aseismic slip, and to millions of years
of intermittent activity (e.g., Nielsen, 2017; Sibson, 2003). The textures and structures that develop in fault
rocks depend on the amount and rate of shearing, and the physical conditions, that is, temperature and pres-
sure, under which the shearing occurred. Fault zones that develop within and above the seismogenic zone
are often characterized by highly localized brittle shear deformation (Figure 1a); they are termed brittle
faults. In contrast, faults formed at much deeper levels in the crust tend to deform by purely thermally acti-
vated creep mechanisms and are characterized by continuous ductile displacement; they are termed ductile
faults (e.g., Scholz, 2019). As depth increases, a fault zone is expected to encounter a gradual transition from
brittle to ductile deformation, occurring in a depth interval with temperatures spanning from a few tens to
hundreds of °C (typically 400650°C, depending on the fault rock composition). It is usually termed the
brittletoductiletransition zone (e.g., Aharonov & Scholz, 2019; Kawamoto & Shimamoto, 1997). The
brittletoductiletransition zone also denes the lower limit of seismic activity on faults since ductile
deformation is largely aseismic (Aharonov & Scholz, 2019).
The terminology and classication of fault rocks are not universally agreed upon yet (Woodcock &
Mort, 2008). Fault rocks formed above a depth of 14 km are generally incoherent, friable, and uncemented;
they include fault breccia and fault gouge (Figure 1a). At greater depth (up to 1015 km), typical fault rocks
are different types of cataclasites, sometimes with pockets of pseudotachylyte that were formed during seis-
mic slip events (e.g., Sibson & Toy, 2006). The cataclasite group is subdivided according to the relative pro-
portion of negrained matrix into protocataclasite, (meso)cataclasite, and ultracataclasite (Figure 1b; e.g.,
Brodie et al., 2007; Woodcock & Mort, 2008). A similar classication also applies to mylonite as the fault rock
type for ductile deformation at greater depth (Figure 1b). The brittletoductiletransition zone, as the
name suggests, is characterized by mixed mode of fault slip behavior and by extensive juxtaposing and over-
printing of different types of fault rocks, such as mylonite with pseudotachylyte interlayers and mylonitized
pseudotachylyte (e.g., Hayman & Lavier, 2014; Menegon et al., 2017; White, 1996).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 3of60
Figure 2 summarizes the formation stages of a typical brittle fault (e.g., Hattori & Yamamoto, 1999):
(i) increased stress produces a crossjoint system over the rock body (Figure 2a); (ii) rock fragments are
formed by shearing along joints (Figure 2b); (iii) movements of the host blocks in a bookshelfstyle mode
rotate these fragments and abrase them to sand, silt, and claysize particles, even down to micron/nan-
ometersize particles (Figure 2c); (iv) cementation of disintegrated fragments and subsequent formation of
new joints (Figure 2d) with one or more cycles of refracturing (Figure 2e); and eventually (v) a relatively nar-
row zone (millimeters to centimeters thick) composed of very ne grained particles is produced (Figure 2f).
The fault growth process commonly results in a fault zone architecture consisting of a socalled fault core
enclosed by broader damage zones (Figure 2g; Caine et al., 1996; Chester et al., 1993).
The fault core is the result of highly localized strain and intense shearing that accommodates the majority of
the displacement within the fault zone. It typically consists of a number of (recurring) slip surfaces and sev-
eral types of fault rocks, such as fault breccia, fault gouge, and/or cataclasite (Caine et al., 1996; Chester
et al., 1993; Choi et al., 2016; Sibson, 2003). The principal slip zone (PSZ) cuts the fault core (Figure 2g); it
is the foremost location where physicochemical processes take place as a result of individual earthquake
events (Boullier, 2011; Sibson, 2003). Examples of fault rocks from the rst borehole of the Wenchuan earth-
quake Fault Scientic Drilling project (WFSD1) are shown in Figure 2 (Wang et al., 2014).
The damage zone enveloping the fault core (Figure 2g) is characterized by what is referred to as distributed
deformation (i.e., where homogenous strain is distributed across a network of evenly spaced faults, cf. Nixon
et al., 2014). The damage zone differs structurally, mechanically, and petrophysically from the undeformed
host rock or protolith (Caine et al., 1996; Chester et al., 1993; Choi et al., 2016). The host rock surrounds the
damage zone and remains basically unchanged during faulting (Caine et al., 1996). As mentioned earlier on,
a fault undergoes cycles of creep and seismic slip. A seismic cycle is divided into three periods: (1) the coseis-
mic period (the time during an earthquake itself, typically seconds to minutes), (2) the postseismic period
(days, months, and sometimes years after a given earthquake), and (3) the interseismic period (the time
Figure 1. (a) Sketch illustrating the occurrence of diverse fault rocks in the shallow crust (reproduced from
A. Lin et al., 2005). Fault breccia, fault gouge, and cataclasite mostly occur in the brittle regime; in contrast, mylonites
develop within the ductile regime. Pseudotachylytes are frozen melt pockets occasionally formed by seismic frictional
heating during large earthquakes. (b) A revised classication of fault rocks as suggested by Woodcock and Mort (2008).
The brittle versus ductile behavior of rocks depends, to a large extent, on strain rate. The delineation of brittle and
ductile zones here reects expected behavior under typical geological strain rates of 10
14
s
1
.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 4of60
Figure 2. (af) An idealized schematic model showing the development of a brittle fault zone (modied after Hausegger et al., 2010, and Billi, 2005) and
(g) the typical architecture of a fault zone (after Chester et al., 1993, and Caine et al., 2010). A brittle fault zone is the result of several deformation processes:
(a) formation of cross joints, (b) fracturing of initial rocks and fragmentation, (c) fracturing and disintegration of fragments by bookshelf rotation, (d) cementation
of disintegrated fragments, (e) subsequent formation of new joints due to the refracturing, and (f) eventually a relatively narrow zone (i.e., principal slip
zone, PSZ) is formed composed of very ne grained particles. (g) Typical fault zone structure consisting of the fault core that includes fault gouge cut by the PSZ,
fault breccia, and/or cataclasite, the damage zone surrounded by undeformed host rock (protolith). The gure is not to scale. (h) Examples of different fault
rocks, which were retrieved from the rst borehole of the Wenchuan earthquake Fault Scientic Drilling project (WFSD1) (adapted from Wang et al., 2014). The
blue and red lines indicate the boundary between cataclasite and fault gouge, and the inferred PSZ of the 2008 Mw 7.9 Wenchuan earthquake, respectively.
The numbers on both sides of the cores indicate the depth intervals they were retrieved from.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 5of60
span between large earthquakes, lasting tens to thousands of years). A fault zone is the cumulative product
of multiple slip cycles (Figure 3). In the following, we summarize some of the processes active in fault zones
during the seismic cycle with bearing on the magnetic properties of fault zones.
Thermochemical reactions and pseudotachylytes. During an earthquake (Figure 3a), the mechanical energy
due to the high seismic slip rates is converted into heat; it constitutes the largest part (~80% to 90%) of the total
earthquake energy budget (Pittarello et al., 2008; Scholz, 2019). Frictional heating will quickly raise the tem-
perature in the slip zone after rupture (Rice, 2006). Considering typical seismic slip rates (1 m/s) and total slip
distance (tens of centimeters to meters) at a fault plane, the related temperature rise is <100°C near Earth's
surface but can become >1100°C at seismogenic depths (>5 km) (e.g., McKenzie & Brune, 1972). The tem-
perature rise may promote thermal decomposition or dehydration of certain mineral phases and the forma-
tion of breakdown products (e.g., Di Toro et al., 2011, Figure 3a). This also pertains to the magnetic minerals
(see section 4.3 for a detailed discussion). The temperature rise within the slip zone is sometimes sufcient to
trigger melting of the host rock minerals (typically >1000°C, e.g., Spray, 2010). Quick, quench cooling of
these local melt pockets of millimeter to decimeter scale produces glasses or partly devitried glasses, called
pseudotachylytes (Figure 1a; e.g., A. Lin, 2008; A. Lin et al., 2005; McKenzie & Brune, 1972; Philpotts, 1964;
Figure 3. A conceptual model showing the faultingrelated physical and chemical processes and the causes of potential magnetic changes in a fault zone during
the different stages of the earthquake cycle. (a) During a coseismic period, the high rate seismic slip may crush the ferrimagnetic grains (if present) in the
PSZ (indicated by the dark band) to a much ner size, due to the comminution of rock fragments. Meanwhile, frictional heating raises the temperature and can
induce a myriad of thermochemical reactions within a fault zone, which include transformations of Febearing minerals with immediate bearing on magnetic
properties. Coseismic hot uids and those from deeper down may dissolve the Febearing minerals and release mobile Fe
2+
. (b) During a postseismic period,
neoformation of ferrimagnetic minerals occurs through precipitation of the mobilized Fe
2+
as mixed ferrousferric iron oxide. Further, several thermochemical
reactions may proceed during cooling of the fault uids warmed by friction in the PSZ. (c) During an interseismic period, inltration and percolation of
meteoric water and/or deeplayer uids into the fault zone may dissolve (part of) the Febearing minerals again and induce precipitation of new magnetic
minerals. A fault zone is the cumulative product of many earthquake cycles. The fault core cartoons are modied after Gray et al. (2005).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 6of60
Sibson & Toy, 2006; Spray, 2010; Swanson, 1992). One should realize that the wording pseudotachylyte is also
in use to describe any aphanitic, dark colored fault rock or amorphous material formed during shear defor-
mation (see Rowe & Grifth, 2015). Throughout this review, however, we restrict pseudotachylytes to rocks
with a frictional melt origin. Pseudotachylytes commonly have a relatively high concentration of negrained
magnetite as a result of oxidation of meltsusceptible mac minerals (e.g., Ferré et al., 2005; Nakamura
et al., 2002; O'Hara, 2001; Pittarello et al., 2012; Zhang et al., 2018). This makes pseudotachylytes stand out
magnetically, in comparison with their host rocks, so that they are attractive rocks for magnetic characteriza-
tion (see section 4.4 for a detailed discussion).
Fluid movement in fault zones. Fault zones feature a dense network of fractures and secondary faults and thus
act as major uid conduits in the crust (Bense et al., 2013; Vermilye & Scholz, 1998). Fluids of multiple
sources, including meteoric waters, trapped formation brines, mineral dehydration, and volatiles from the
deep underlying layers (Hickman et al., 1995; Zoback et al., 2007), can inltrate into and percolate along fault
zones. Overpressurization may occur promoting earthquake nucleation (e.g., Miller, 2013; Scuderi &
Collettini, 2016; Sibson, 1992); pore uid pressurization during frictional heating ashes may induce fault
weakening lengthening fault slip time spans (e.g., J. Chen et al., 2017; Miller, 2013; Rice, 2006).
Importantly, uidrelated dissolutionprecipitation processes are anticipated to be common during all peri-
ods of the seismic cycle, that is, the coseismic, postseismic, and interseismic periods (Figure 3; e.g., Bense
et al., 2013; J. Chen et al., 2013, 2016; Gratier et al., 2013). This is discussed in section 4.5. These reactions play
a critical role not only in physical, chemical, and mechanical evolution of fault rocks (Bradbury et al., 2015;
J. P. Evans & Chester, 1995; Goddard & Evans, 1995; Isaacs et al., 2007; Niwa et al., 2015; Sutherland
et al., 2012) but also in the very earthquake rupture (Hickman et al., 1995; Sutherland et al., 2012; Zoback
et al., 2007).
Earthquake lightning. It is also reported that during large earthquakes (usually with Mw > 5), the crus-
tal deformation activates and releases gases and/or electrical charges. These subsequently generate
atmospheric electric elds and currents, which in turn affect the atmospheric electric circuit possibly
resulting in luminous phenomena (e.g., Derr, 1973; Enomoto et al., 2017; C.L. Kuo et al., 2011;
Lockner et al., 1983; StLaurent et al., 2006). This transient coseismic electric phenomenon is referred
to as earthquake lightning (EQL), which travels along the fault plane and may produce anomalous mag-
netizations in fault rocks (e.g., fault gouges and pseudotachylytes; Enomoto & Zheng, 1998; Enomoto
et al., 2001; Ferré et al., 2005). The mechanisms and magnetic effects are discussed in section 4.6.
All in all, a fault zone should be considered an extremely dynamic system, characterized by specic physico-
chemical conditions during each of the faulting stages (Figure 3; e.g., Cerchiari et al., 2020; Wibberley
et al., 2008). For example, the redox state associated with faulting and accompanying uid ow varies enor-
mously during seismic slip events (Ishikawa et al., 2008; Yamaguchi et al., 2011). Also, fault rocks are subject
to multiple deformation events, often under different physicochemical conditions. Consequently, a fault rock
is not simply a granulated product of its protolith but is among the most complex and heterogeneous geolo-
gical materials. It may be viewed as a lowto mediumgrade metamorphic rock (J. P. Evans & Chester, 1995).
Iron features three main valence states, zerovalent, ferrous, and ferric iron (along with several more exotic
intermediate forms) and is therefore a sensitive probe of the physicochemical conditions within fault zones.
Iron cycling as a consequence of faulting would cause deposition, leaching, and/or chemical alteration of the
Fe
2+
and/or Fe
3+
in Febearing minerals in fault zones. Very reducing conditions could lead to metallic iron.
In principle, this leads to measurable changes in magnetic properties, making rock magnetism a suitable
probing technique to decipher faulting processes. Unraveling the magnetic properties of fault rocks thus pro-
vides important clues to understand the evolution of fault zones. Also, aspects of earthquake physics and
chemistry may be unveiled. In the following section (section 3) we outline the merits of the rock magnetic
or mineral magnetic methodology (with emphasis on fault rocks) before detailing the drivers of magnetic
changes in fault rocks (section 4).
3. Rock and Mineral Magnetism of Fault Rocks
Magnetic mineralsor ferromagnetic minerals (sensu lato) refer to minerals that are magnetic at room tem-
perature: irontitanium oxides, some iron suldes, and metallic iron; the latter, however, rarely occurs in the
Earth's crust. Coupled electron spins over large atomic distances in the crystal structure result in collective
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 7of60
spin behavior. Each magnetic mineral has a magnetic ordering temperature above which it loses its collective
magnetism and becomes a paramagnet; on cooling through the ordering temperature the collective magnetic
behavior is restored. At low temperatures below ~1020 K, however, many more minerals order magnetically
and lowtemperature magnetic instrumentation can evaluate also those minerals.
The magnetic assemblage, concentration of each magnetic mineral, and magnetic granulometry of fault
rocks reect faulting processes. The rock's magnetic features are sensitive to both chemical and physical
changes occurring in rocks during the faulting. For example, variations in physical grain size are often
reected in the magnetic granulometry, which is the (inferred) grainsize distribution of magnetic particles
in a sample, usually expressed through the dominant magnetic domain structure (Figure 4a). A particle's
domain structure is the result of minimizing the overall particle energy of individual magnetic particles.
The domain structure or state depends on the size and shape of the magnetic particles, next to being tem-
perature dependent. Crystal defects and the internal stress distribution in the magnetic grains also have their
impact on the domain state (Özdemir & Dunlop, 1997), which may be particularly relevant for rocks and
minerals that have undergone faulting.
We now provide (approximate) size ranges for the domain state types distinguished for magnetite, by far the
most common magnetic mineral in nature. At room temperature, the smallest particles, with a diameter of
up to 25 nm are superparamagnetic (SP); they cannot retain a geologically stable natural remanent magne-
tization (NRM). Next in size, in the 3080 nm range (for equant grains), particles are single domain (SD).
Larger particles, up to a few μm in size, are featuring noncollinear spin structures and possibly even a
few domains; they are termed classically pseudo single domain (PSD, Stacey & Banerjee, 1974) but probably
more correctly should be referred to as vortex stateparticles (Schabes & Bertram, 1988; see also Almeida
et al., 2016, who show examples of particles in a vortex state) as recently proposed by Roberts et al. (2017). SD
and PSD or vortex particles are paleomagnetically most stable. Large grains, >510 μm or so, contain many
domains and are termed multidomain (MD). The reader is referred to Appendix A for further explanation;
textbooks include Dunlop and Özdemir (1997), M. E. Evans and Heller (2003), Stacey and Banerjee (1974),
and Tauxe (2010). Next to size also particle shape exerts a critical control on domain state threshold sizes. For
example, rodlike particles of 0.1 × 0.1 × 1 μm are SD, whereas equant particles of 1 μm can be MD
(Figure 4a).
Various types of magnetic measurements and their combinations are used to determine the magnetic
carrier(s), their concentration, and (magnetic) grain size. They are categorized along eld, frequency,
and temperaturedependent measurements. Measurement of various types of laboratoryinduced remanent
magnetizations, that is, anhysteretic remanent magnetization (ARM) and isothermal remanent magnetiza-
tion (IRM), also yields important information. The fundamentals and applications of rock and mineral mag-
netism are documented in several textbooks (e.g., Dunlop & Özdemir, 1997; M. E. Evans & Heller, 2003;
Maher & Thompson, 1999; Nagata, 1961; Stacey & Banerjee, 1974; Tauxe, 2010; Thompson &
Oldeld, 1986) and review papers (e.g., Dekkers, 1997; Hunt et al., 1995; Q. S. Liu et al., 2012; Peters &
Dekkers, 2003; Verosub & Roberts, 1995). Below we outline the more relevant approaches for the analysis
of fault rocks. In a number of cases also nonmagnetic methods are utilized in concert with magnetic methods
to constrain the latter's interpretation. Before describing those magnetic and nonmagnetic methods, how-
ever, some specics on sample collection and preparation for magnetic studies on fault rocks need to be con-
sidered. Magnetic property analysis requires physical samples since magnetic changes due to faulting
processes may be reected on the mm to μm scale. This implies drilling and coring rocks through fault zones;
borehole logging onlyenables tracking broader (magnetic) aspects of fault zones.
3.1. Samples for Magnetic Studies Into Fault Rocks
As mentioned above, fault zones are narrow zones with widths in the order of centimeters to meters.
Particularly, the PSZs are much thinner, in many cases <1 cm (Sibson, 2003). Many fault zone studies, there-
fore, share a common denominatoronly small amounts of sample are available. Such small amounts can
be delicate for the more classical geochemical and mineralogical analyses. In contrast, a relatively small
amount of material (< ~500 mg) with any shape (e.g., chip, powder, or drilling/cutting residue) is sufcient
for a complete set of magnetic analyses. Thus, in principle, magnetic changes at a small scale allow for resol-
ving the faultingrelated behavior. Magnetic scanning may reveal changes at the μm scale.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 8of60
Figure 4. Domain states of magnetite as a function of grain size and shapes, and examples for commonly used plots for
higheld magnetic measurements. (a) Illustration of the domain state categories for magnetite at room temperature
and their relationship with shape and grain size of the magnetic particles (redrawn from Roberts et al., 2018).
(b) Example of a hysteresis loop with denition of M
rs
,M
s
,B
c
, and B
cr
, with the inset showing a backeld
demagnetization curve, with the denition of B
cr
. The higheld slope (typically, B> 0.7 T) on the hysteresis loop is
dened as higheld magnetic susceptibility (χ
hf
). The dark and gray lines are the hysteresis loops before and after
correction for the nonferrimagnetic matrix contribution, respectively. In this example the correction is a paramagnetic
correction as χ
hf
is positive. (c) Example of a Day plot of the hysteresis ratios M
rs
/M
s
and B
cr
/B
c
. Single domain (SD),
pseudo single domain (PSD), and multidomain (MD) boundaries are after Day et al. (1977). Note that recent Day plots
often use vortex rather than PSD following Roberts et al. (2018). The data points shown are host and fault zone
sediments from the frontal prism in Japan Trench cored by the Integrated Ocean Drilling Program Expedition 343, Japan
Trench Fast Drilling Project (JFAST) (reproduced from Yang et al., 2018). The host sediments lie mainly in the PSD
(or vortex) eld, whereas most of the fault zone samples are located in the SD region. (d) Example of a FORC diagram
(rstorder reversal curve) for one of the fault zone sediments (FZ697) shown in Figure 4c. B
c
is equivalent to
particle coercivity, and B
u
to the local interaction eld. Colors in the diagram represent absolute values of FORC density.
The B
c
peak centered at 3040 mT with a prominent kidneyshape toward higher coercivity suggests the
occurrence of SD pyrrhotite. FORC data are from Yang et al. (2018) and are reprocessed with the FORCinel package
(Harrison & Feinberg, 2008) with the VARIFORC (Egli, 2013) option used. VARIFORC smoothing parameters: vertical
ridge Sc0 = 4, horizontal smoothing factor Sc1 = 7, central ridge Sb0 = 3, vertical smoothing factor Sb1 = 7,
horizontal lambda λc= 0.1, and vertical lambda λb= 0.1. (e) Example of an isothermal remanent magnetization
(IRM) acquisition curve. After application of a stepwise increasing magnetic eld, the remanent magnetization
increases until a maximum value is reached, which is termed saturation IRM (SIRM or M
rs
). (f) Example of unmixing
of an IRM acquisition curve to identify magnetic coercivity distributions using the MAX UnMix package (reprinted
from Maxbauer et al., 2016). The sample is an anoxic lake sediment from Baldeggersee, Switzerland, and three
coercivity components are identied: detrital magnetitelow coercivity component (Component 1), biogenic magnetite
(Component 2), and oxidized magnetite (or hematite)Component 3.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 9of60
Anisotropy of magnetic susceptibility (AMS) and/or paleomagnetic studies of fault rocks (see sections 3.5
and 3.6 for details) require oriented samples. Classically, such samples are cylinders of 25 mm dia-
meter × 22 mm height or 20 mm onaside cubes (known as standardsize samples), representing a volume
deemed statistically representative. Due to the limited width of a fault zone, oriented sampling in a narrow
fault zone is a challenge, and contaminationby the host rock adjacent to the target layer(s) may compli-
cate the interpretation. Also, bias may creep into AMS results due to the unconventionally small sample size
dictated by fault rock samples. Almqvist et al. (2020), for example, found that the mean magnetic susceptibil-
ity and magnetic anisotropy degree (see section 3.5 for denitions) are inversely proportional with sample
size for a set of pseudotachylyte cubes from western Jämtland (central Swedish Caledonides) ranging in
volume between ~0.2 and ~0.03 cm
3
.Minicoresof 12 mm diameter × 11 mm long, cubic miniAMS sam-
ples with 1 cm edges, or even down to a size of 3.5 mm onaside (hence ~250 times volumetrically smaller
than standardsize samples), often work ne. AMS analysis of pseudotachylyte fabric on such 3.5 mm cubes
enabled determining the focal mechanisms of ancient earthquakes (Ferré et al., 2015, 2016). Despite sample
size issues for very small samples where a cautious and careful interpretation is appropriate, AMS studies on
small samples are nonetheless a promising new avenue to unveil detailed geological features (Almqvist
et al., 2020; Ferré et al., 2015, 2016; Zhu et al., 2017).
Berndt et al. (2016) demonstrated that samples with a thermoremanent magnetic moment larger than
10
11
Am
2
irrespective of their sizes (ranging from the centimeter scale to tens of nm scale) contain enough
magnetic particles to be accurate magnetic recorders. Paleomagnetic measurements can thus also be carried
out on smaller samples than the standard size samples. Millimetersized samples (e.g., Böhnel et al., 2009;
Suttie et al., 2010) or even smaller samples of geological materials like single silicate crystals (e.g., zircon, pla-
gioclase, and olivine crystals) can yield accurate paleointensity and paleodirection estimates using ultrahigh
sensitivity moment magnetometry and advanced demagnetization techniques (e.g., Berndt et al., 2016; Fu
et al., 2017; Sato et al., 2015; Tarduno et al., 2015). These include, but are not limited to, scanning SQUID
(superconducting quantum interference device) microscopy (SSM; e.g., Weiss et al., 2007), quantum dia-
mond magnetometry (QDM; e.g., Glenn et al., 2017), and microwave demagnetization (e.g., Suttie
et al., 2010). This enables paleomagnetic analyses of the fault rocks. For such studies, oblique drilling across
a fault zone may offer a way to orient the samples by making use of the presentday NRM overprint that is
present in most samples.
3.2. FieldDependent Measurements: LowField Magnetic Susceptibility
One of the most widely used room temperature magnetic measurements is the loweld magnetic suscept-
ibility (volume specic: κ
lf
, or mass specic: χ
lf
), measured in small applied magnetic elds, typically,
~200300 A/m; that is, several times the strength of the Earth's magnetic eld (~2550 A/m). It is a measure
of the magnetizabilityof a sample and expresses in a general way the classes of magnetic materials (e.g.,
Dunlop & Özdemir, 1997): dominantly ferromagnetic or ferrimagnetic (very high χ
lf
), paramagnetic (small
positive χ
lf
), or diamagnetic (very small negative χ
lf
). A rock is an ensemble of ferromagnetic (sensu lato),
paramagnetic, and diamagnetic minerals. Therefore, to characterize the ferrimagnetic (χ
ferri
) component
(s) (i.e., magnetite, pyrrhotite, or greigite) in samples, contributions from paramagnetic (χ
para
), diamagnetic
(χ
dia
), and imperfect antiferromagnetic minerals (i.e., hematite and goethite) should be subtracted from the
bulk loweld magnetic susceptibility (χ
lf
). The imperfect antiferromagnetic minerals are often included in
the paramagnetic contribution since their susceptibility does not differ that much. Magnetic susceptibility
can also be measured at different frequencies of the applied alternating current (AC) eld (typically varying
between a few hundreds of Hz to a few kHz), yielding the frequencydependent magnetic susceptibility,
expressed either as a percentage of the loweld susceptibility (κ
fd
%) or as a mass(χ
fd
) or volumespecic
number (κ
fd
). With increasing measurement frequency, the SP/stable SD (SSD) boundary (~25 nm at room
temperature for a relaxation time of 100 s) shifts to smaller volumes. Hence, for a given grainsize distribu-
tion more grains become blocked at a higher measurement frequency and are SD with a lower susceptibility
than SP particles, which explains the difference in susceptibility between the low and high measurement fre-
quency. Frequencydependent magnetic susceptibility is thus an effective way to determine the concentra-
tion of magnetic particles over a small grain size window across the SP/SSD boundary (e.g., Q. S. Liu
et al., 2012).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 10 of 60
The diamagnetic and paramagnetic magnetic moment increases linearly with applied eld up to very high
eld values; hence, their susceptibility is constant for a given temperature. Also, magnetite's magnetization
is linear with the applied magnetic eld up to ~800 A/m, that is, the entire eld range of loweld suscept-
ometers. However, the magnetization of some other ferrimagnetic minerals, such as pyrrhotite, hematite,
and titanomagnetite, shows a marked applied eld dependence already in the low eld of interest here
(de Wall, 2000; Hrouda, 2011; Hrouda et al., 2006; Jackson et al., 1998; Worm et al., 1993). Variation of
the magnetic susceptibility as a function of applied eld can thus diagnose the presence of these magnetic
minerals in fault zones. The bridgetype susceptometers are the most versatile and widely used instruments
for susceptibility measurements. This includes eld, frequency, and temperaturedependent magnetic
susceptibility measurements, AMS measurements (see also section 3.5), and for some instruments, measure-
ment of the inphase and outofphase components of the magnetic susceptibility (e.g., Hrouda et al., 2016;
Hrouda & Ježek, 2014).
3.3. FieldDependent Measurements: HighField Magnetic Measurements
3.3.1. Hysteresis Loops and FORC Diagrams
Magnetic hysteresis loops (Figure 4b) are not only informative of the magnetic mineralogy but also indica-
tive of the dominant domain structure of magnetic minerals in the sample. Saturation magnetization (M
s
)is
the largest possible magnetization of a sample; it is thus a measure of the total amount of ferromagnetic
minerals in the sample. The saturation remanence (M
rs
) is the corresponding remanent magnetization after
removal of the applied eld. The coercivity (B
c
) and remanent coercivity (B
cr
) are measures of magnetic sta-
bility. The ratios M
rs
/M
s
and B
cr
/B
c
are commonly plotted on the socalled Day plot (Figure 4c; Day
et al., 1977; Dunlop, 2002a, 2002b; Dunlop & Özdemir, 1997; Lanci & Kent, 2003) as indicators of domain
states and, indirectly, grain size. As an alternative, the plot of M
rs
/M
s
versus B
c
, referred to as squareness
plotin the engineering literature, was proposed to avoid the need of determining B
cr
and to remove poten-
tial ambiguities associated with the B
cr
/B
c
ratio: Particles with markedly different B
c
values may have their
B
cr
/B
c
largely overlaying complicating interpretation of the domain state (Tauxe et al., 2002). The Day plot is
a classic rock magnetic tool and has been used extensively for domain state diagnosis since it was proposed
by Day et al. (1977). However, several fundamental ambiguities associated with its interpretation have been
documented recently by Roberts et al. (2018). Additional constraints, other than those offered by the Day
plot, are required to properly interpret the hysteresis parameters (i.e., dominant domain state) in terms of
physical grain size. Beyond the Day plot, the rstorder reversal curve (FORC) technique has been developed
to derive more information from hysteresis measurements (e.g., Roberts et al., 2000, 2006, 2014; Zhao
et al., 2015). FORC diagrams (an example in Figure 4d) provide important insights into not only the mag-
netic minerals but also the distribution of coercivity and magnetostatic interactions among magnetic parti-
cles (see review by Roberts et al., 2014, for more details). Given the interpretive ambiguities in the Day
diagram, unmixing of FORC diagrams (Lascu et al., 2015), and determination of remanent, transient, and
induced FORC diagrams (P. X. Hu et al., 2018; Zhao et al., 2017) have been suggested as suitable candidates
for diagnosing domain state on a componentbycomponent basis (Roberts et al., 2018). However, when
samples are dominantly paramagnetic or antiferromagnetic with only a small ferrimagnetic contribution,
the low signaltonoise ratio prevents acquisition of meaningful FORC data, that is, the ferrimagnetic signal
is weak in comparison to the total magnetization (e.g., Jackson & Solheid, 2010). For those samples hyster-
esis loop measurement with determination of the classic hysteresis parameters is still the foremost way to
characterize their magnetic behavior. In addition, hysteresis measurements also allow for assessing the rela-
tive magnetic contribution of ferrimagnetic minerals and that of the matrix minerals (i.e., paramagnetic and/
or diamagnetic) through comparison of the higheld magnetic susceptibility (χ
hf
, i.e., the slope of the linear
higheld segments of the hysteresis loop, cf. Figure 4b) and the loweld susceptibility (χ
lf
) of a sample
(Figure 4b).
3.3.2. Remanence Measurements at Room Temperature
Higheld methods are most commonly used to identify magnetic mineral phases. IRM acquisition curves
(Figure 4e) obtained at room temperature in successively increasing direct elds contain a wealth of infor-
mation about magnetic hardness. Unmixing of IRM acquisition curves along with forward basis functions
can provide information that is diagnostic of magnetic mineralogy (Figure 4f; Egli, 2004a, 2004b, 2004c;
Kruiver et al., 2001; Maxbauer et al., 2016). Collections of IRM acquisition curves can also be subjected to
end member modeling, an inverse bilinear unmixing approach (Heslop & Dillon, 2006; Weltje, 1997).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 11 of 60
The end members then serve as basis for a geological interpretation of the variability in the IRM acquisition
curves. It has been used for paleoenvironmental analysis, the detection of magnetotactic bacteria (e.g., Just
et al., 2012), and in the context of remagnetization, the (partial) resetting of the NRM in rocks at some point
during their geological history (e.g., Aben et al., 2014; Gong et al., 2009; Huang et al., 2015).
ARM is generated by placing a sample in a small steady direct current (DC) eld, the DC bias eld (of the
order of the Earth's magnetic eld), and superimposing an AC eld of which the amplitude is steadily
ramped down from a preset initial value to zero. It is a useful laboratory technique for characterizing
magnetic particles (Q. S. Liu et al., 2012). It senses SD magnetite particularly well. The different behavior
of ARM and χ
lf
with grain size make plots of ARM versus χ
lf
(i.e., the Banerjee diagram, Banerjee et al., 1981)
or χ
ARM
(χ
ARM
= ARM/DC bias eld) versus χ
lf
(i.e., the King diagram, King et al., 1982) useful for detecting
grain size changes, in particular for magnetite.
Another rock magnetic parameter, which is often employed to measure the relative contributions of low and
high coercivity material to a sample's saturation isothermal remanent magnetization (SIRM), is the Sratio.
It compares the SIRM obtained in a saturating forward eld with the IRM obtained subsequently in a suita-
ble backeld (usually a reversed eld of 0.3 T, yielding IRM
0.3T
). The classic Sratio, S
classic
,isdened as
(e.g., Thompson & Oldeld, 1986) follows:
Sclassic ¼IRM0:3T=SIRM (1)
It scales from 1 (only high coercivity material) to 1 (only low coercivity material). It can also be dened
as (Bloemendal et al., 1992) follows:
S¼1IRM0:3T=SIRMðÞ=2 (2)
or
Sforward ¼IRM0:3T=SIRM (3)
where the IRM
0.3T
is remanent magnetization acquired in a forward eld of 0.3 T. The Sratios in
Denitions (2) and (3)the latter is also called forward Sratio(Heslop, 2009)scale from 0 to 1; they
provide a measure of the relative contribution of the low coercivity material to the total remanence. It
should be realized that numeric values of Denition (1) on the one hand versus (2) and (3) on the other
are different and cannot be used interchangeably without the denitions being specied. When the Sratio
approaches unity, in general, ferrimagnetic minerals are dominant over antiferromagnetic minerals, and
vice versa. However, it is important to note that variations in the relative concentration of highand
lowcoercivity phases are nonlinear and interpretation of the Sratio is nonunique, because it is not solely
a function of mineral type but is also inuenced by factors such as grain size and substitution. The Sratio
is an example of socalled closed data (i.e., data within certain bounds like fractions or percentages); the
additivelogratio transform (e.g., Swan & Sandilands, 1995) has been suggested to obtain meaningful
descriptive statistics for the Sratio information (Heslop, 2009).
3.4. TemperatureDependent Magnetic Measurements
3.4.1. HighTemperature Magnetic Measurements
Hightemperature magnetic analyses, either ineld measurements or remanence measurements, are
among the most diagnostic measurements for assessing magnetic mineralogy because magnetization
decreases with increasing temperature and becomes 0 at and above the magnetic ordering temperature since
the magnetic mineral has become a paramagnet. The magnetic ordering temperature is named Curie tem-
perature (T
C
) for ferromagnet and ferrimagnet and Néel temperature (T
N
) for imperfect antiferromagnetic
materials (Dunlop & Özdemir, 1997). For example, magnetite and hematite have T
C
and T
N
values of
578°C (Figure 5a) and 675°C (Figure 5b), respectively (Dunlop & Özdemir, 1997). The magnetite
ulvöspinel (Fe
3
O
4
Fe
2
TiO
4
) solid solution series and the hematiteilmenite (αFe
2
O
3
FeTiO
3
) solid solution
series show a marked range in ordering temperatures respectively from 578°C to 153°C (T
N
of ulvöspinel)
and from 675°C to 233°C (T
N
of ilmenite) (Dunlop & Özdemir, 1997). With increasing Ti substitution
ordering temperatures decrease approximately linearly. Goethite has a T
N
of ~120°C (Figure 5c), which
decreases below room temperature with increasing Al substitution (Q. S. Liu et al., 2006). Goethite can
accommodate up to ~32 mol% Al substitution with a corresponding magnetic ordering temperature of
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 12 of 60
58°C (Murad & Bowen, 1987). Maghemite has a T
C
of 645°C (Figure 5d; Özdemir & Banerjee, 1984), but it
often inverts to hematite before the T
C
is reached: Depending on grain size and amount of substitution
maghemite can invert to hematite at any temperature between 250°C and 1000°C (Dunlop &
Özdemir, 1997). A typical inversion range for negrained maghemite is between 300°C and 400°C (e.g.,
Q. S. Liu et al., 2005). Monoclinic pyrrhotite has a T
C
of 325°C (Figure 5e); T
C
is slightly lower for
Figure 5. Examples of thermomagnetic curves of some common magnetic minerals, showing their hightemperature
behavior including Curie or Néel temperatures. Stepwisethermomagnetic analyses were performed in air with a
modied horizontal translation Curie balance (Mullender et al., 1993) in Paleomagnetic Laboratory Fort Hoofddijk,
Utrecht University (The Netherlands). Details on samples, heating/cooling rates, temperature cycles, and applied elds
are given in each panel. Kadan hematite (panel b) is from the Czech Republic and Robe River maghemite (panel d)
is from Western Australia. The Hopkinson peak (panel c) is a manifestation of superparamagnetism: the coercivity of the
goethite is very high (tens of teslas at room temperature) in comparison of the eld range used in the
thermomagnetic analysis. Then the magnetic moment can be considered a loweldsusceptibility which goes through
a maximum at the Néel/Curie temperature.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 13 of 60
Nisubstituted pyrrhotite (Kontny et al., 2000; Schwarz & Vaughan, 1972). T
C
remains undetermined for
greigite because greigite decomposes before the T
C
can be measured (like the situation of maghemite): It
exceeds 400°C (Roberts et al., 2011). Diagnostic for greigite is its thermochemical alteration to a nonmag-
netic phase starting at ~200°C and nishing at ~350°C when heated in air (Figure 5f; Dekkers et al., 2000;
Reynolds et al., 1994; Roberts, 1995; Torii et al., 1996). At higher temperatures, above 400°C magnetite forms
which in turn is converted to hematite above 500550°C yielding the typical greigiteinairthermomag-
netic signature. To date, greigite has not been reported in fault rocks; however, discrimination between grei-
gite and pyrrhotite is essential for fault zone magnetic studies, as the latter is commonly present as product of
thermal alteration of pyrite in fault zones. In addition, with stepwise thermal demagnetization of three
orthogonal IRM components the thermal behavior of high, intermediate, and lowcoercivity magnetic
phases can be evaluated separately (Lowrie, 1990).
3.4.2. LowTemperature Magnetic Measurements
A major disadvantage of hightemperature measurements is that the magnetic signal may be obscured by
newly formed magnetic mineral(s) during the thermal treatment. This applies to ineld measurements in
particular. During measurements at low temperature, down to 4 K, thermochemical alteration is nonexistent
because the sample is not heated. Therefore, lowtemperature measurements are increasingly utilized in
rock magnetism (Dunlop & Özdemir, 1997; Q. S. Liu et al., 2012; Moskowitz et al., 1998; Rochette et al., 1990).
They are designed to determine magnetic ordering temperatures below room temperature and whether or
not socalled lowtemperature transitions in magnetic minerals occur. For example, the wellknown
Verwey transition (T
V
) at ~120 K or 153°C (Figure 6a; Verwey, 1939) is diagnostic of magnetite, when it
undergoes a crystallographic phase transition from a cubic to a monoclinic structure. Similarly, the Morin
transition (T
M
,~250 K) at which hematite's magnetic structure undergoes a spin op, a change in orienta-
tion with respect to its crystal structure, is a key indicator of the presence of (crystalline) hematite
(Figure 6b; Morin, 1950). Monoclinic pyrrhotite exhibits the Besnus transition at 3034 K (Figure 6c;
Besnus & Meyer, 1964; Dekkers et al., 1989; Rochette et al., 1990, 2011). Goethite shows an increasing mag-
netization with decreasing temperature (Figure 6d; Guyodo et al., 2006; Rochette & Fillion, 1989). The
lowtemperature behavior with possible magnetic transitions is usually measured with a socalled magnetic
properties measurement system (MPMS); both ineld and remanent magnetizations are measured as a
function of applied eld and temperature between 1.8 and 400 K. The most commonly used measurement
sequences and their advantages have been described in detail by Bilardello and Jackson (2013).
However, one must exercise caution when interpreting lowtemperature data to avoid ambiguities as some
minerals show similar features at the same temperature range. At low temperatures other Febearing mate-
rials, for example, Fe or Mn carbonates (siderite (FeCO
3
), and rhodochrosite (MnCO
3
)) order magnetically
at 37 and 34 K, respectively (Frederichs et al., 2003; Housen, Banerjee, & Moskowitz, 1996). Vivianite Fe
3
(PO
4
)
2
·8H
2
O) orders as well below 20 K (Frederichs et al., 2003). Febearing silicates (clay minerals, chlorite,
amphiboles, and pyroxenes) may also order below 1520 K because of low thermal vibration at those low
temperatures. SP particles show distinctive lowtemperature behavior (e.g., Chang et al., 2009); spin glasses
may freeze in and iron oxide coatings around silicates may order below 50 K (e.g., Franke et al., 2007).
Therefore, lowtemperature measurements deliver very useful information, but their interpretation requires
expert knowledge. One should perform more elaborate measurement strategies (e.g., Bilardello &
Jackson, 2013; Kars et al., 2011) or combine the results with other magnetic and/or nonmagnetic measure-
ments. A nice example is the study of Kars et al. (2011). Through applying a magnetic eld of 5 μT inside the
SQUID MPMS during the cooling of a room temperature SIRM to detect the presence of a magnetic ordering
temperature, they successfully diagnosed siderite and rhodochrosite in overmatured claystones from a bore-
hole in the Netherlands and in outcrop samples from the Borneo Prism; the previous interpretation of
negrained pyrrhotite had to be abandoned.
3.5. Anisotropy of magnetic susceptibility and remanence
The magnetic anisotropy of rocks, that is, the magnetic properties of a given sample vary depending on the
sample orientation, results from the contributions of ferromagnetic (sensu lato), paramagnetic, and diamag-
netic minerals. Detailed coverage on the theoretical background, mineral sources, measurement procedures,
quantication, geological applications, and advantages and limitations of AMS and anisotropy of magnetic
remanence (AMR), hereafter collectively called magnetic fabric,can be found in several textbooks and
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 14 of 60
review papers (Borradaile & Henry, 1997; Borradaile & Jackson, 2004, 2010; Ferré, Gébelin, et al., 2014;
Hirt, 2007; Hrouda, 1982; Jackson, 1991; MartínHernández et al., 2004; MartínHernández & Ferré, 2007;
Parés, 2015; Potter, 2004; Rochette et al., 1992; Tarling & Hrouda, 1993). Below we briey outline the
rationale for AMS approaches.
AMS depicts geometrically the orientation and degree of alignment of the preferred orientations of mineral
grains, the mineral grain spatial distribution or their latticepreferred orientation, and the shape or
crystalline anisotropy of the grains in a rock (Figures 7a and 7b; e.g., Ferré, Gébelin, et al., 2014;
MartínHernández & Ferré, 2007). It is commonly expressed by a symmetric secondorder tensor (Tarling
& Hrouda, 1993) and described by a triaxial ellipsoid with principal axis' lengths equal to the eigenvalues
of the magnetic susceptibility tensor, κ
max
κ
int
κ
min
, which correspond to the maximum, intermediate,
and minimum susceptibility axes, respectively (Figure 7b). The degree of anisotropy of the ellipsoid is mea-
sured by the parameter P(Nagata, 1961) or the correctedP,P
j
(Jelínek, 1981), which are respectively
dened as follows:
κmax=κmin (4)
and
Pj¼exp 2 lnκmax lnκm
ðÞ
2þlnκint lnκm
ðÞ
2þlnκmin lnκm
ðÞ
2

1=2(5)
where κ
m
is the mean magnetic susceptibility, κ
m
=(κ
max
+κ
int
+κ
min
)/3. The shape of the ellipsoid is
characterized by the parameter T(Jelínek, 1981), which is dened as follows:
Figure 6. Lowtemperature treatment of (S)IRM for some common magnetic minerals, showing the lowtemperature
transition of magnetite (a), hematite (b), and monoclinic pyrrhotite (c). (a) Warming curve of SIRM acquired in 2.5 T at
5 K of a 37 nm sized stoichiometric magnetite showing the Verwey transition (T
V
) at ~120 K typical of magnetite
(Verwey, 1939). Data are from Özdemir et al. (1993). (b) Lowtemperature cycling of a 2.5 T SIRM at 300 K of a MD
hematite crystal (0.24 mm × 2.15 mm × 3.2 mm in size) from Mount Wright, Québec (Canada), showing its Morin
transition (T
M
) at ~250 K (Morin, 1950). Data are from Özdemir et al. (2008). (c) Lowtemperature cycling of SIRM
imparted at 300 K with a 4 T eld for monoclinic pyrrhotite illustrating the Besnus transition (T
Bes
)at3034 K
(Besnus & Meyer, 1964; Rochette & Fillion, 1988; Rochette et al., 1990). The sample with grain size of 250150 μm was
separated magnetically from a pyrite/pyrrhotite skarn collected from Temperino in Tuscany, Italy (Dekkers, 1988;
lowtemperature data are from Dekkers et al., 1989). (d) Lowtemperature cycling (3001040010 K) of an IRM acquired
at 300 K in a 2.5 T eld for a synthetic goethite sample showing its lowtemperature magnetic behavior. Data are from
Guyodo et al. (2006).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 15 of 60
T¼2ln κint=κmin
ðÞ=ln κmax=κmin
ðÞ1 (6)
Tranges over the spectrum of ellipsoid shapes (Figure 7c): 0 T1 for oblate ellipsoids (disk shaped),
1T0 for prolate ellipsoids (rod shaped), and T= 0 for neutral ellipsoids. Along with T, the magnetic
lineation L,
L¼κmax=κint (7)
and magnetic foliation F,
F¼κint=κmin (8)
are most informative to describe the AMS ellipsoid (Tarling & Hrouda, 1993). They are commonly plotted
on a socalled Flinn diagram (Figure 7c; Flinn, 1965; Jelínek, 1981).
AMS can detect incipient deformation well before other strain indicators (e.g., Almqvist & Koyi, 2018;
Burmeister et al., 2009). The loweld AMS, with assets of ease of data acquisition (minutes) and sensitivity
(down to one per mille), has been used almost exclusively in classic petrofabric analysis studies (e.g.,
Borradaile, 1988; Graham, 1966; Hrouda, 1982; MartínHernández et al., 2004). In the context of fault
rocksof special interest herewe note the extensive review by Ferré, Gébelin, et al. (2014) on magnetic
fabrics in ductile shear zones. In addition, measurement of AMS at high elds (e.g., Ferré et al., 2004;
Hrouda & Jelínek, 1990; Kelso et al., 2002; MartínHernández & Ferré, 2007; MartínHernández &
Hirt, 2001; Rochette et al., 1992), and temperaturedependent AMS (Issachar et al., 2016, 2018;
Parés & van der Pluijm, 2014; Schmidt et al., 2007) have been proposed to separate the ferrimagnetic from
the nonferrimagnetic (i.e., paramagnetic and diamagnetic) contributions to the AMS. Recently,
frequencydependent AMS was proposed to determine both the presence of SP particles and the relative ten-
sor contributions of the SP and larger size fractions to the overall magnetic fabric (Hrouda & Ježek, 2014).
Figure 7. Principle of the AMS method for revealing petrofabrics. (a) A schematic diagram of a rock fabric. (b) The AMS
ellipsoid is composed of three orthogonal principal axes: κ
max
: maximum, κ
int
: intermediate, and κ
min
: minimum.
The κ
max
and κ
min
axes, respectively, correspond to long and short axes of grains or crystals. (c) Flinn diagram of the
lineation (L) versus foliation (F) illustrating the shapes of AMS ellipsoids in terms of the corrected degree of
anisotropy (P
j
, Tarling & Hrouda, 1993). The shape parameter (T) is also shown. Figures 7a and 7b are reprinted from
Cho et al. (2015), and Figure 7c is modied from Dubey (2014).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 16 of 60
This would help to evaluate the occurrence of grain ning of preexisting magnetic particles as a consequence
of the intensive shearing during earthquake slip (see section 4.2).
The measurement of magnetic remanencerelated anisotropies is much more time consuming and therefore
less utilized. It includes measurement of anisotropy of anhysteretic remanent magnetization (AARM) (e.g.,
Issachar et al., 2018; Jackson et al., 1991), anisotropy of isothermal remanent magnetization (AIRM) (e.g.,
Kodama & Dekkers, 2004; Stephenson et al., 1986), and anisotropy of thermal remanent magnetization
(ATRM) (Borradaile & Lagroix, 2000; Hirt, 2007). Such AMR data have been suggested to be an efcient
way to characterize the individual contributions of ferrimagnetic and antiferromagnetic minerals
(MartínHernández & Ferré, 2007). Ferrimagnetic and paramagnetic/diamagnetic minerals in fault rocks
are usually formed at different faulting stages and/or through different faultingrelated process (e.g., fric-
tional heating and uidrock interactions; see section 4). Those minerals, therefore, may reect diverse
deformation paths. Examination of AMR across a fault zone might thus help diagnosing the subfabric
carried by the newly formed ferrimagnetic mineral(s) due to seismic frictional heating. Also, the thermal
remanent magnetization (TRM) imprint across fault zones may be evaluated, providing insights into the
dynamic deformation processes in a fault zone. The AMR techniques, including instrumental requirements,
measurement strategy, and pros and cons compared to standard loweld AMS measurements, have been
discussed in detail by MartínHernández and Ferré (2007).
3.6. Paleomagnetic Considerations
The NRM of rocks is constituted of one or several component(s) that as a rule of thumb represent(s)
different times in the geological history of that rock. The same pertains to fault rocks: The faulting gen-
erates frictional heat that (partially) resets the existing NRM. To properly isolate these NRM components,
stepwise alternating eld or thermal demagnetization should be carried out with a sufcient number of
demagnetization steps according to standard paleomagnetic practice (e.g., Tauxe, 2010). The NRM demag-
netization behavior is visually inspected on vector endpoint plots (also referred to as Zijderveld plots;
Zijderveld, 1967), and the paleomagnetic directions are typically determined with principal component
analysis (Kirschvink, 1980). Different NRM components in fault rocks, especially the new socalled
secondarycomponent(s) formed during the rupturing itself (see section 4.7), may hold valuable
information on faultingrelated effects (e.g., Chou, Song, Aubourg, Lee, et al., 2012; Ferré, Geissman,
et al., 2014; Leibovitz, 2016).
3.7. Nonmagnetic Measurement Techniques
Fault rocks contain mixed magnetic mineral assemblages of potentially different origin, grain size, and
relative concentrations. Several types of magnetic measurements must therefore be used in concert to
properly identify magnetic carriers and minimize ambiguities. In some cases, the magnetic results are
supplemented by data acquired with what is referred to as nonmagnetic techniquesto further unravel
their nature and causal links to faulting processes. The most frequently used nonmagnetic techniques
include, but are not limited to, the following: Mössbauer spectroscopy (e.g., Dyar et al., 2006), scanning
electron microscopy (e.g., Reed, 2005), transmission electron microscopy (e.g., McLaren, 2005), Xray
diffraction analysis (e.g., Lavina et al., 2014), electron probe microanalysis (e.g., Reed, 2005), magnetic
force microscopy (e.g., Grütter et al., 1995), electron spin resonance (e.g., Fukuchi, 2003; Pan &
Nilges, 2014), and diffuse reectance spectroscopy (e.g., Torrent & Barrón, 2008). All these measure-
ments can provide important insights into the valency of Fe in minerals, preferred grain alignment,
grain size, and geochemistry of Febearing minerals, and help constraining the magnetomineralogical
interpretation.
4. Drivers for Changes in Magnetic Properties of Fault Rocks
As mentioned above, faulting is a complex process involving various physical and chemical changes, which
complicates the interpretation of fault rocks by any type of measurement, thus including the rock magnetic
approach reviewed here. The magnetic properties of fault rocks are determined by several mechanisms. In
this section we focus on those for which a consensus has been reached. These include (i) straininduced mag-
netic changes, (ii) grainsize variations of preexisting magnetic minerals by grain ning, (iii) thermochemi-
cal reactions induced by frictional heating, (iv) coseismic frictional melting with pseudotachylyte formation,
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 17 of 60
(v) chemical alteration and neomineralization due to uid percolation, (vi) IRM acquisition due to EQL, and
(vii) remagnetization of the NRM due to physicochemical processes.
4.1. StrainInduced Magnetic Changes
Stress and related strain effects may have signicant inuence on magnetic properties of rocks. Stresses (dif-
ferential or even hydrostatic) can directly affect the behavior of ferrimagnetic mineral grains, especially
(titano)magnetite through the effect of magnetostriction (also known as piezomagnetism; e.g., Bezaeva
et al., 2015; Kapička, 1988, 1992; Kean et al., 1976). Strain may cause reorientation and/or internal deforma-
tion of ferrimagnetic grains (e.g., Borradaile, 1988; Till & Moskowitz, 2014). This would lead to changes in a
variety of magnetic properties, the most important of which are magnetic anisotropy, changes in remanence
orientation and intensity, and changes in bulk magnetic properties such as coercivity and magnetic suscept-
ibility (e.g., Bezaeva et al., 2015; Borradaile, 1996; Gattacceca et al., 2007; Gilder et al., 2004, 2006; Jackson
et al., 1993; Jiang et al., 2013; Kapička, 1988, 1992; Kapička et al., 2006; Kean et al., 1976; Louzada et al., 2010).
For example, up to 25% shortening due to axial compression in a set of synthetic magnetitebearing calcite
sandstonesamples irreversibly increases their coercivity and magnetic anisotropy but decreases their mean
magnetic susceptibility and the remanence component parallel to the shortening axis (Jackson et al., 1993).
Brittle or semibrittle fault rocks generally form in the upper ~1015 km of the Earth's crust. The vertical
effective stress at this depth interval is about 160240 MPa, calculated from an average rock density
(2,600 kg m
3
) under the assumption of hydrostatic pore uid pressure. The stress sensitivity of magnetic
susceptibility for various rocks ranges roughly from 0.8 to 1.3 × 10
3
MPa
1
(unit: one over pressure per unit
area; e.g., Kapička, 1988; Kean et al., 1976). The stress may induce decreases in a rock's magnetic susceptibil-
ity ranging from 12% to 30%, if only the vertical effective stress is taken into consideration. In this context,
the potential stress effects on magnetic properties of fault rocks cannot be ignored, at least not in seismogenic
zones. It has been documented that changes in magnetization of stressed rocks due to local accumulation of
stress may induce transient anomalies in the local geomagnetic eld prior to an earthquake, referred to as
piezomagnetism or tectonomagnetism (e.g., Sasai, 2001; Yamazaki, 2013).
However, an interpretation is complicated by many factors, such as variations in magnetic carriers (mag-
netic mineralogy, concentration, and grain size), ambient magnetic eld, stress loading pattern (e.g., hydro-
static or uniaxial, shock or static, single or cycling), and duration (e.g., Gilder et al., 2006, 2018;
Kapička, 1992; Louzada et al., 2010; Volk & Feinberg, 2019). For example, pressure loading removes part
of the remanent magnetization of a rock when the pressure is applied in a zero ambient magnetic eld, a
procedure conveniently termed pressure demagnetization (e.g., Bezaeva et al., 2010; Gattacceca et al., 2007;
Jiang et al., 2013; Louzada et al., 2010, 2011; Pozzi, 1975; Volk & Feinberg, 2019; Volk & Gilder, 2016). The
higher the pressure, the more remanence is removed. A detailed compilation of the available experimental
pressure demagnetization data on (titano)magnetite, (titano)hematite, and pyrrhotitebearing rocks is
provided in Louzada et al. (2011). In contrast, a rock acquires magnetic remanence when the (cycling)
pressure loading is taking place in a magnetic eld; such remanence is known as pressure remanent mag-
netization or piezoremanent magnetization (Nagata, 1966). The higher the ambient magnetic eld, the
more remanence is acquired for a given pressure (Nagata & Carleton, 1969). Also, with increasing Ti sub-
stitution, the magnetostriction coefcient of magnetite increases (Dunlop & Özdemir, 1997; Gilder & Le
Goff, 2008). Thus, the magnetic susceptibility and remanence of titanomagnetite are much more sensitive
to the stress regime than stoichiometric magnetite (Dunlop & Özdemir, 1997; Gilder & Le Goff, 2008).
Moreover, the effect of pressure on magnetic susceptibility generally decreases with decreasing magnetic
grain size; the acquisition of pressure remanent magnetization or demagnetization is quite efcient in
large MD grains in comparison to small SD particles (e.g., Gilder et al., 2006; Kean et al., 1976; Volk &
Feinberg, 2019).
4.2. Grain Fining of Preexisting Magnetic Particles
A fault core evolves by grain comminution consisting of early bulk fragmentation and late abrasion of
grains (e.g., Billi, 2005; Hattori & Yamamoto, 1999; Sammis & BenZion, 2008; Storti et al., 2007), reecting
a progressive reduction in grain size. So, in fault rocks the intensive shearing also may cause widespread
splitting of ferromagnetic (sensu lato) grains (if present) into (much) ner material (Figure 3a). Magnetic
susceptibility of ferrimagnetic particles depends strongly on grain size when the grains are close to the
SP/SSD threshold with SP particles having a distinctly higher magnetic susceptibility than SSD particles
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 18 of 60
(e.g., Walden et al., 1999). Thus, magnetic susceptibility may be elevated in
fault rocks, even when there is no change in the total concentration of
magnetic minerals. Measuring frequencydependent magnetic susceptibil-
ity and lowtemperature magnetic measurements are powerful means for
estimating the SP particle contribution (e.g., Banerjee et al., 1993; Q. S. Liu
et al., 2012). Effects of grain ning as a consequence of shearing can thus
be evaluated by these two approaches. The identication of comminuted
nanograined magnetic particles is also informative on fracture energetics
of fault gouge, as calculation of total grain surface area is a common
approach to estimate the fracture energy associated with fault gouge for-
mation (e.g., Aretusini et al., 2017; Chou, Song, Tsao, et al., 2014; Wilson
et al., 2005).
4.3. Thermochemical Reactions Induced by Frictional Heating
The transient high temperatures involved in ruptures induce thermoche-
mical reactions within the fault zone (Figure 3a), including decomposi-
tion and transformation of Febearing minerals, particularly in the PSZ
(e.g., Han et al., 2007; Pei, Zhou, et al., 2014; Tanikawa et al., 2007,
2008; Yang et al., 2018, 2019). Febearing minerals that are widely present
in fault zones are siderite (FeCO
3
), lepidocrocite (γFeOOH), goethite
(αFeOOH), pyrite (FeS
2
), smectite (Ca, Na, H) (Al, Mg, Fe, Zn)
2
(Si,
Al)
4
O
10
(OH)
2
xH
2
O) and chlorite ((Mg, Fe, Li)
6
AlSi
3
O
10
(OH)
8
). Most
relevant thermal reactions are described below.
1. Siderite's thermal decomposition product is magnetite (Koziol, 2004;
Pan et al., 2000) according to:
3FeCO3þ0:5O2
400580°C Fe3O4þ3CO2(9)
To illustrate this reaction, highvelocity friction experiments (to simulate seismic slip in the laboratory) on
sideritebearing fault gouge by Han et al. (2007) are insightful: they revealed that siderite was thermally
decomposed to nanocrystalline magnetite. This changed the gouge color from pale brown to black and
increased the magnetic susceptibility by nearly 3 orders of magnitude (Figure 8). It thus leaves a prominent
magnetic imprint of seismic slip in a fault zone, even when siderite is present only in minor amounts (Han
et al., 2007; Tanikawa et al., 2008).
2. Lepidocrocite is one of the four iron oxyhydroxides known in nature. It is a paramagnet at ambient
temperature and commonly occurs in fault gouges developed from granitic rocks (Fukuchi et al., 2005).
On annealing it converts to antiferromagnetic hematite (αFe
2
O
3
) with strongly ferrimagnetic maghe-
mite (γFe
2
O
3
) as an intermediate phase (Fukuchi, 2003; Gendler et al., 2005). The conversion of lepi-
docrocite to maghemite (γFe
2
O
3
) starts at ~200°C, and the thermal dehydroxilation is completed at
~300350°C, followed by inversion of maghemite to hematite (Gehring & Hofmeister, 1994; Gendler
et al., 2005). Maghemite is stabilized by its surface energy, so it is often prominently present as ne
particles (T. Chen et al., 2005). Particularly maghemite leaves a magnetic imprint, due to its ferrimag-
netic nature.
2γFeOOH
~200350°C
γFe2O3þH2OγFe2O3
> 500°C
αFe2O3(10)
3. Goethite, another iron oxyhydroxide, thermodynamically the most stable under ambient conditions,
also commonly occurs in fault zones. It dehydroxilates to hematite with traces of magnetite or
magnetite/hematite aggregates between 240°C and 400°C (Dekkers, 1990; Gualtieri & Venturelli, 1999;
Figure 8. Results of a highvelocity friction experiment on a simulated fault
gouge consisting of a 1:1:1 mixture of siderite, quartz, and calcite
under a normal stress of 1.3 MPa and at a slip rate of 2.0 m/s conducted by
Han et al. (2007). (a) The sheared gouge shows markedly lower Xray
diffraction (XRD) peak intensities in comparison to the original
sideritebearing gouge. The Xray diffractogram also shows new magnetite
peaks on the sheared gouge. The gouge color changes from (b) pale
brown before the experiment to (c) black after the experiment, with an
increase in magnetic susceptibility (χ) of a few orders magnitude,
supporting the XRD interpretation of magnetite.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 19 of 60
Özdemir & Dunlop, 2000; Yang et al., 2019). The magnetite in particular can be sensed readily by rock
magnetic techniques.
2αFeOOH
~240400°C
αFe2O3þγFe2O3=Fe3O4
ðÞ
þH2O (11)
4. Pyrite is often found as accessory mineral in faults developed within siltstones (e.g., Chou, Song,
Aubourg, Song, et al., 2012) and other sediments (e.g., Yang et al., 2018). It decomposes to pyrrhotite
(FeS
x
,1x< 1.08) as a result of heating to temperatures >500°C under reducing conditions that prevail
in the subsurface (Bhargava et al., 2009; Toulmin & Barton, 1964). Pyrrhotite is strongly ferrimagnetic
and its occurrence makes a magnetic imprint.
FeS2FeSxþ10:5xðÞS2(12)
5. Smectite and chlorite are among the most common constituents of fault gouge. On heating they desorb
iron which is precipitated as magnetite. Finegrained magnetite is produced when heating smectite over
250°C in the laboratory; nucleation and growth of the magnetite grains proceed up to ~450500°C (Hirt
et al., 1993; Yang et al., 2019). Thermochemical reactions of the iron adsorbed to chlorite (particularly
chamosite) result in the neoformation of magnetite on heating to 400700°C (Tanikawa et al., 2008).
Newly formed ferrimagnetic minerals (i.e., magnetite, maghemite, and pyrrhotite) as consequence of the
aforementioned thermochemical reactions may increase magnetic susceptibility and magnetization of fault
rocks up to a few orders of magnitude (e.g., Han et al., 2007; Mishima et al., 2006, 2009; Tanikawa et al., 2007,
2008; Yang et al., 2012a, 2018, 2019). These newly formed magnetic minerals can be readily diagnosed by the
different rock magnetic measurements (or combinations thereof) discussed in section 3. It is necessary to
exclude the possibility that the elevated magnetic susceptibility is caused by ning of ferrimagnetic particles
before attributing it to the magnetomineralogical changes. However, frictional heating is essentially con-
strained to the PSZ itself: The temperature rise induced by frictional heating decays exponentially away from
the PSZ, as widely reported in modeling results (e.g., Fulton et al., 2013; Yang et al., 2018; Yang, Chen,
et al., 2013). Therefore, frictional heating acts prominently only on a very narrow (millimeters to centi-
meters) zone enclosing the PSZ. Its effects attenuate quickly in other parts of a fault zone, for example,
the damage zone.
4.4. Coseismic Frictional Melting With Pseudotachylyte Formation
As mentioned earlier, pseudotachylytes may occur within seismic slip zones (e.g., Ferré et al., 2012, 2017;
Hirono, Ikehara, et al., 2006; A. Lin, 2008; A. Lin et al., 2005; Otsuki et al., 2009; Rowe & Grifth, 2015;
Sibson & Toy, 2006), due to frictional heating that raises the temperature for a short time above the rock's
melting point. The oxygen fugacity in the highly localized melt pockets is generally 24 log units above
the fayalitemagnetitequartz buffer (~11 log units at ~1000°C and 110160 MPa, e.g., O'Hara &
Huggins, 2005). Pseudotachylytes are often more magnetic than their parent rocks, due to the presence of
negrained PSD/SD ferrimagnetic minerals, predominantly magnetite, that form due to oxidation of mac
silicates during the melting and subsequent quenching (Ferré et al., 2005, 2012, 2017; Ferré, Geissman,
et al., 2014; Nakamura et al., 2002; O'Hara, 2001; Pittarello et al., 2012; Zhang et al., 2018). This enables a
potentially excellent recording of the NRM in the pseudotachylytes during cooling through the blocking
temperature, that is, the NRM is a TRM and should be considered a primary remanence, representing the
time of pseudotachylytes cooling, that is, shortly after the rupture itself. In such a context, the paleomagnetic
baked contact test and examination of the additivity of partial TRM (e.g., Tauxe, 2010), which reveals the
unblocking temperature spectrum and its gradient across a fault zone, would be a means to locate the zones
with the highest temperatures in the pseudotachylyte vein zones. Recently, Zhang et al. (2018) found metal-
lic iron in vacuum fusion experiments (in argon) on cataclasites from Borehole No. 2 of the WFSD project. It
was formed by the reducing action of Febearing minerals above 1300°C. They proposed that the presence of
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 20 of 60
metallic iron could be another reason for the magnetic highs of pseudotachylytes. Similarly, studies on
highpressure hightemperature effects in other disciplines, such as impact glasses (e.g., Rochette et al.,
2015, 2019), fulgurite (e.g., Essene & Fisher, 1986), shocked Martian meteorites (e.g., Van de Moortèle
et al., 2007), have also documented that metallic iron was formed from reduction of more oxidized material.
It is hard to identify the metal nanoparticles with conventional mineralogical techniques (e.g., scanning
electron microscopy [SEM], Raman spectrometry, and Mössbauer spectroscopy; cf. Van de Moortèle
et al., 2007), the presence of metallic iron in natural pseudotachylytes may thus have been overlooked in
previous studies.
4.5. Chemical Alteration Due to Fluid Percolation
Fault rocks can be susceptible to disaggregation and dissolution by uids during the coseismic, postseismic,
and interseismic periods (cf. section 2 and Figure 3). The shortduration pulses of hot uid induced by fric-
tional heating during the seismic slip (Ishikawa et al., 2008; MuirWood, 1994) can easily dissolve preexisting
Febearing minerals (e.g., pyrite, siderite, and Fesilicates) whereby Fe
2+
is liberated (Humbert et al., 2012;
Kopp & Humayun, 2003) (Figure 3a). This is followed by precipitation of (among others) ferrimagnetic
minerals during the postseismic period because of the low solubility of iron (oxy)(hydr)oxides (Figure 3b).
During interseismic periods that last for very long times, uids, both of meteoric and connate origin inltrate
the fault zone and percolate along a dense microcrack and macrocrack network (Figure 3c). Fluid circula-
tion would lead to destabilization of ironbearing (clay) minerals in fault zones, with release of Fe
2+
that dis-
solves in the uid and is subsequently transported (Grosz et al., 2006; Hashimoto & Kaji, 2012; Just &
Kontny, 2012). The Fe
2+
rich uids thus are potential sources to precipitate magnetic minerals (Pechersky
& Genshaft, 2001). Upon reaction with an Fe
2+
bearing uid at temperatures up to 250°C and the most com-
mon pH ranges (i.e., pH = ~4.5 at 250°C to ~6.5 at 50°C; Ohmoto, 2003), for example, hematite will trans-
form to magnetite. Magnetite in turn can be transformed to hematite in acidic solutions, with ferrous iron
getting into solution (i.e., a socalled nonredox reaction, cf. Ohmoto, 2003):
αFe2O3þH2O fluid
ðÞ
þFe2þFe3O4þ2Hþ(13)
Reaction of lepidocrocite with dissolved Fe
2+
can form magnetite in alkaline media (Cornell &
Schwertmann, 2003):
2γFeOOH þFe2þFe3O4þ2Hþ(14)
Similarly, the breakdown of pyrite produces goethite or hematite depending on temperature and pH. High
temperature and low pH generally favor hematite over goethite (Cornell & Schwertmann, 2003): at low tem-
perature and high pH (slightly alkaline conditions)
4FeS2þ15O2þ10H2O fluidðÞ4αFeOOH þ8H2SO4(15)
or at high temperature and low pH (moderately acidic conditions)
4FeS2þ15O2þ8H2O fluidðÞ2αFe2O3þ8H2SO4(16)
Magnetic properties of fault zones thus change drastically on (partial) removal of preexisting magnetic
minerals and/or formation of new (magnetic) minerals (e.g., Chou, Song, Aubourg, Song, et al., 2012;
Yang, Chen, et al., 2013; Yang, Yang, et al., 2016). However, the migration paths and composition of fault
uids may vary enormously over time and space, as fault zones act as uidow conduits or barriers at dif-
ferent stages in their development (Goddard & Evans, 1995). This leads to temporal and spatial changes in
redox state of a fault zone as function of episodic fault rupture and healing cycles during the lifetime of
the fault (Yamaguchi et al., 2011). Thus, precipitation products of uid circulation (e.g., iron oxides, iron
hydroxides, and/or suldes) are highly dependent not only on the physical and chemical conditions (perme-
ability, temperature, pressure, redox and pH, etc.) and chemical composition of the uids but also on the
uid/rock ratio and the lithology of the fault zones (Chou, Song, Aubourg, Song, et al., 2012; Grosz
et al., 2006; Guichet et al., 2006; Yang, Chen, et al., 2013; Yang, Yang, et al., 2016).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 21 of 60
4.6. EQLInduced IRM Acquisition
EQL currents have been reported on several occasions during large earthquakes (e.g., Fidani, 2010;
Heraud & Lira, 2011; Kamogawa et al., 2005; J. Y. Liu et al., 2015; StLaurent et al., 2006; Thériault
et al., 2014). However, the underlying formation mechanism is still controversial. A urry of mechanisms
have been put forward to explain the charge generation in the subsurface during earthquake motion: piezo-
magnetism, piezoelectricity, electrokinetic ow processes, opening of the tips of cracks, stressactivation of
positive holes, and hydromechanical rupture (Enomoto et al., 2017; Freund et al., 2007; Losseva &
Nemchinov, 2005; StLaurent et al., 2006). Regardless of its formation mechanism, the coseismic electric
current is expected to travel along the fault plane, which often acts as a good electrical conductor with a
conductivity of up to 4 orders higher than the surrounding insulating host rocks (Ferré et al., 2005).
According to the BiotSavart law, a transient strong azimuthal magnetic eld will be produced perpendicu-
lar to the fault plane (Ferré et al., 2005, 2012; Losseva & Nemchinov, 2005). In this arrangementlamellar
conductive rocks enveloped by insulating marginsfault gouge, pseudotachylite generation veins, parallel
or near parallel to the fault plane, will readily acquire an IRM. This IRM is a secondary NRM termed
earthquake lightninginduced remanent magnetization (EQLIRM)(Enomoto et al., 2001; Enomoto &
Zheng, 1998; Ferré et al., 2005, 2012). EQLIRM is distinct from lightning NRM in rocks struck by surface
lighting; in the latter the lightning IRM remagnetization pattern is circular (either clockwise or counter-
clockwise) centered on the lightningbolt impact location, coupled with a decrease in magnetization inten-
sity away from the electric bolt (e.g., Losseva & Nemchinov, 2005; Verrier & Rochette, 2002). In the context
of EQLIRM fault rocks, several paleointensity tests (e.g., Ferré, Geissman, et al., 2014; Gattacceca &
Rochette, 2004) and normalization of the NRM intensity by M
s
are helpful to discriminate high NRM
caused by a strong ambient magnetic eld from that caused by high concentrations of magnetic minerals.
If the latter can be excluded, the high NRM then may be attributed to a strong magnetic eld induced by
seismically generated electrical currents.
4.7. Secondary NRM Acquisition Due to Physicochemical Processes in Fault Zones
The EQLIRM dealt with in the previous section is a rather special case. In general, neoformed magnetic
minerals (resulting from any thermochemical reaction and/or uidrock interaction) may carry a chemical
remanent magnetization (CRM). Also, the mere cooling of a slip zone (hot coseismic uids included) would
lead to an imprint of a (partial) TRM in a fault zone (Figure 9). These newly imprinted NRM components in
and near the slip zone are oriented parallel to the contemporaneous local geomagnetic eld (Chou, Song,
Aubourg, Lee, et al., 2012). Evidently, a NRM direction will be different in a fault zone when the (latest or
most intense) rupture occurred any time later than the formation of the original rocks. The fault gouge
can retain the magnetic record during interseismic periods. However, these newly imprinted NRM compo-
nents may be (partially) overprinted again during later seismic events (Figure 9; see also Chou, Song,
Aubourg, Lee, et al., 2012).
Often, several of the aforementioned mechanisms act simultaneously; it is necessary to nd support for indi-
vidual mechanisms and/or their acting in concert to better understand the magnetic changes within a fault
zone. For instance, the removal or decrease in concentration of certain magnetic mineral(s), and grain ning
of magnetic particles that might be induced by the faulting, can readily be evaluated by concentration
dependent magnetic properties and grainsizedependent properties, respectively. This is a foremost asset
of rock magnetic techniques.
5. Magnetic Properties and Faulting Processes
5.1. Indicator of Seismic Slip Records
Proper slip zone identication is crucial to understand earthquake generation processes and paleoseismol-
ogy (e.g., Rowe & Grifth, 2015). Estimating earthquake recurrence rates requires proper identication of
slip zones that accommodated historic earthquake events. Also, the most intense slip conditions may be
revealed. These are fundamental ingredients for seismic risk assessment. However, to nd robust evidence
of seismic slip is not straightforward from a practical point of view. It is to be expected that the aforemen-
tioned faultingrelated magnetic changes should leave a trace in the rock record, thus making magnetic
property analysis a potential indicator of seismic slip.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 22 of 60
5.1.1. Magnetic Locating of Pseudotachylyte Veins
Classically, pseudotachylytes have been considered as the only denitive evidence of ancient seismic activity
(Cowan, 1999; Rowe & Grifth, 2015). They are well known as earthquake fossilsand less prone to natural
alteration on geological timescales (A. Lin, 2008). A strong NRM is often observed in pseudotachylyte veins
due to the presence of abundant PSD/SD magnetite (Figure 10; Ferré et al., 2005). This makes (paleo)mag-
netic measurements a robust way to locate pseudotachylyte veins, providing further linkage to past seismic
events, such as the kinematic solution (see section 5.4.1 for details). In addition, it is believed that magnetic
properties of pseudotachylytes are controlled by oxygen fugacity and thus vary systematically with depth of
formation (Ferré et al., 2012, 2017; Ferré, Geissman, et al., 2014; O'Hara & Huggins, 2005). Pseudotachylytes
that formed at shallower depths (<1012 km) commonly exhibit coseismically formed maghemite or hema-
tite (e.g., Fukuchi, 2003; Petrik et al., 2003), while those formed at intermediate crustal depths (~20 km) are
generally dominated by magnetite (e.g., Davidson et al., 2003; Ferré, Geissman, et al., 2014; Nakamura &
Nagahama, 2001). In contrast, pseudotachylytes formed at even greater depths tend to be ilmenitedomi-
nated (e.g., Ferré, Geissman, et al., 2014; Moecher & Steltenpohl, 2009). Therefore, rock magnetic properties
are informative on the ambient conditions of the earthquake rupture (Ferré et al., 2017; Ferré, Geissman,
et al., 2014).
5.1.2. Coseismic/Postseismic Neoformed Magnetic Minerals
Coseismic and/or postseismic neoformed magnetic minerals along with their related anomalies in magnetic
properties have also been widely reported during the recent rapid responsefault zone drilling campaigns
after major earthquakes. These include the Taiwan Chelungpu Fault Drilling Project (TCDP) following the
1999 ChiChi earthquake (e.g., Hirono, Lin, et al., 2006; Kano et al., 2006), the WFSD drillings following the
2008 Wenchuan earthquake (e.g., H. Li et al., 2015; Pei, Li, et al., 2014; Wang et al., 2014; Zhang et al., 2017),
and the Japan Trench Fast Drilling Project (JFAST) following the 2011 Tohokuoki earthquake
(e.g., Brodsky et al., 2020; Fulton et al., 2013; W. Lin et al., 2013; Yang et al., 2018). In TCDP HoleB, the
Fault Zone FZB1136, the PSZ of the 1999 ChiChi earthquake was estimated to contain up to ~200 ppmv
magnetite from the magnetic susceptibility and saturation IRM peaks (Figure 11; Chou, Song, Aubourg,
Song, et al., 2012; Chou, Song, Aubourg, et al., 2014). One should bear in mind that the parameters used here
(i.e., magnetic susceptibility and saturation IRM) are dependent on both magnetic particle size and concen-
tration of magnetic minerals (e.g., Dunlop & Özdemir, 1997; Peters & Dekkers, 2003; Walden et al., 1999;
Figure 9. A conceptual model showing the magnetic recording cycle in fault gouge of the Fault Zone FZB1136, TCDP HoleB, which accommodated the Taiwan
1999 ChiChi earthquake (reprinted from Chou, Song, Aubourg, Song, et al., 2012). (a) During an interseismic period, in the fault gouge the magnetic record
of the last earthquake is preserved through geological time. (b) During a coseismic period, the elevated temperature and chemical degradation lead toa
partialtocomplete demagnetization of the preexisting magnetic record in the slip zone and the baked contactzone. (c) During the subsequent postseismic
period, cooling of the slip zone and/or hot coseismic uids leads to an imprint of TRM, while the neoformed magnetic minerals resulting from any form of
chemical process may carry a CRM. The newly imprinted (T)CRM component(s) thus leave(s) a magnetic record of the recent earthquake event.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 23 of 60
see also section 3). Therefore, one has to take into account that estimates of magnetic mineral concentrations
based on these two parameters yield a range of possible concentrations rather than a single value (evidently
with an uncertainty envelope). In contrast, usage of saturation magnetization (M
s
) to this end leads to a
single number for the magnetic mineral concentrations in a given fault rock sample because M
s
is only
proportional to the concentration of magnetic minerals in a sample. In Hole WFSD1, magnetic
susceptibility peaks occur at the fault zone FZ590 which was inferred to have accommodated the 2008
Wenchuan earthquake. The peaks were credited to newly formed magnetite due to the thermal
decomposition of Febearing clays (e.g., smectite and chlorite) by frictional heating during seismic slip
(Pei, Li, et al., 2014; Yang et al., 2012a). In the JFAST drill hole, partial thermal alteration of pyrite to
pyrrhotite was exclusively identied in three slip zones developed in the frontal prism sediments (Yang
et al., 2018). The reactions were induced by seismic frictional heating and related coseismic hot uids
(Figure 12), thus intimately associated with earthquake events (Yang et al., 2018). Examination of the
magnetic mineral assemblage in fault zones may therefore be a prospective way for identifying the most
recent PSZ (Cai et al., 2019; Chou, Song, Aubourg, et al., 2014; Chou, Song, Aubourg, Song, et al., 2012;
Yang et al., 2012a, 2018, 2019).
5.1.3. Seismically Imprinted NRM Components
As mentioned in section 4.7, newly imprinted NRM component(s), that is, TRM and/or CRM, may also leave
a (paleo)magnetic record of recent seismic slip. For instance, in the aforementioned slip zone TCDP
Figure 10. (a) Image of a slab sample across a generation pseudotachylyte vein from the Santa Rosa Mountains,
California and (b) the spatial variation in the magnetic susceptibility of the slab. The map was produced by interpolation
over a grid of 170 susceptibility measurements performed with a Bartington MS2F probe with a sensitivity of
2×10
6
SI. Each 10 mm a data point was collected. (c) NRM intensity along the XXprole in Figure 10b. (d) Hysteresis
properties of pseudotachylyte veins plotted on a Day plot. Most specimens display hysteresis behavior consistent
with PSD grains. (e) NRM thermal decay curves of the two samples (Nos. 03 and 07) obtained from the pseudotachylyte
veins (Figure 10a), indicating lowTi magnetite as the principal remanence carrier. Figures 10a10d and Figure 10e
are reproduced from Ferré et al. (2005) and Ferré, Geissman, et al. (2014), respectively.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 24 of 60
FZB1136, a stable single NRM component with a direction very close to the presentday Earth's eld was
identied within an ~16 cm thick fault gouge zone (Figure 13; Chou, Song, Aubourg, Lee, et al., 2012). It
is therefore inferred to be the coseismic paleomagnetic record of the recent 1999 ChiChi earthquake
(Chou, Song, Aubourg, Lee, et al., 2012). In contrast, NRM records of two other slip zones (i.e., FZB1194
Figure 11. Magnetic susceptibility (a), Sratio (b), and estimated concentration proles of magnetite and goethite
(c) in the Fault Zone FZB1136, TCDP HoleB (Taiwan). The PSZ is characterized by a peak in the magnetic susceptibility
(a) and the Sratio (b), which is consistent with a large amount of magnetite (c). The concentrations of magnetite and
goethite are estimated from the magnetic susceptibility and SIRM of the core samples using a quantitative model
(for details on the modeling strategy, see Chou, Song, Aubourg, et al., 2014). The lowest value of the Sratio is located
near the center of the fault gouge, corresponding to the highest concentration in goethite (c). (d) Magnetic hysteresis
loops and (e) NRM thermal decay curves for selected samples from the Fault Zone FZB1136. Within the gouge (e.g., the
sample at 1136.34 m), the principal maximum unblocking temperature (T
b
) of ~120°C is consistent with the Néel tem-
perature of goethite, while in the PSZ (e.g., sample at 1136.38 m), the maximum T
b
is close to 580°C, indicating the
presence of magnetite as main magnetic carrier. The latter concurs with the open hysteresis loops saturated below 0.3 T
in Figure 11d. These observations suggest that the magnetic susceptibility anomaly at the Fault Zone FZB1136
(Figure 11a) is attributed to the formation of magnetite, as a result of the thermal decomposition (>400°C) of para-
magnetic minerals (e.g., pyrite, siderite, and chlorite) during the 1999 ChiChi earthquake. Figures 11a and 11d are
adapted from Hirono et al. (2007) and Mishima et al. (2009), respectively; Figure 11e is compiled from Chou, Song,
Aubourg, Song, et al. (2012); data in Figures 11b and 11c are from Chou, Song, Aubourg, Song, et al. (2012) and Chou,
Song, Aubourg, et al. (2014), respectively.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 25 of 60
and FZB1243) show directions far away from the presentday eld that were probably acquired before the
BrunhesMatuyama reversal (0.78 Ma). They thus have recorded ancient earthquake events and were not
reactivated during the 1999 ChiChi earthquake; otherwise, these slip zones would have been overprinted,
at least partially (Chou, Song, Aubourg, Lee, et al., 2012).
5.1.4. EarthquakeInduced AMS Fabric
AMS analysis of clastic dikes (see Appendix A for denition) has also successfully identied the unique mag-
netic fabrics of dikes emplaced by injection due to seismically triggered uidization. In the Dead Sea Basin
(Israel), one of the most active tectonic areas in the Middle East, for instance, host sediments appear to be
characterized by oblate AMS ellipsoids with vertical axes of the minimum magnetic susceptibility (κ
min
) that
indicate a normal sedimentary magnetic fabric, whereas the liqueed sediments are characterized by triaxial
fabrics with subvertical κ
int
axes and subhorizontal κ
max
axes parallel to the dike strike (Levi et al., 2006a).
Similar AMS fabrics have been reported by Jayangondaperumal et al. (2010) in the Himalayan frontal thrust
(in the western Himalaya), by Lakshmi et al. (2017) in the Dauki fault (Shillong Plateau, India), and by Cho
et al. (2017) in the Dadaepo Basin (SE Korea). These contrasting magnetic fabrics have provided crucial evi-
dence for seismically triggered uidization of clastic materials. Furthermore, Levi et al. (2018) recently pro-
posed a plot of the magnetic lineation versus the shape parameter of the AMS ellipsoid to distinguish
different seismites formed in soft sediments, that is, damage zones, gouges, earthquaketriggered folds,
Figure 12. A cartoon showing the thermal alteration of preexisting pyrite with neoformation of pyrrhotite as a result of
seismic frictional heating and subsequent coseismic hydrothermal uids in a pyritebearing fault zone during an
earthquake (reprinted from Yang et al., 2018). During the seismic slip, (a) pyrite in the slip zone thermally decomposes to
pyrrhotite once the temperature is over ~640°C and (b) a pyrrhotite rim surrounding the pyrite core is formed
indicating an incomplete reaction. (c) Leaching by coseismic hot uids, pyrite is altered and an intergrown rim of
pyrrhotite and barite is formed. (d) Precipitation of pyrite/pyrrhotite around the silicates and/or nucleation of pyrrhotite
on feldspar grains occurs at expense of iron and sulfur released by destabilization of pyrite and other minerals in fault
zone, during cooling of the coseismic hot uids. The reader is referred to Yang et al. (2018) for more details.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 26 of 60
and breccia layers. It enabled categorizing seismites that underwent different deformation processes into
distinct groups in the abovementioned Dead Sea fault system (cf. Figure 4 in Levi et al., 2018). This
illustrates that AMS can be also used as a potential petrofabric tool for recovering paleoseismic records in
soft sediments (Levi et al., 2006a, 2006b, 2014, 2018).
5.2. Thermal History of Seismic Slip
Estimation of the temperature rise induced by frictional heating during an earthquake provides crucial
information on faulting weakening mechanisms (e.g., frictional melting, thermal decomposition of carbo-
nates, and thermal pressurization), the prevailing dynamic shear stress, slip velocity, and displacement
(e.g., Di Toro et al., 2009; Scholz, 2019; Sibson, 1977; Yao et al., 2016). At the same time, earthquake
energy budgets as a whole can be assessed (e.g., Scholz, 2019). Here, temperature information derived
from melt thermodynamics from experimental igneous petrology data, and pyrometamorphism in prin-
ciple could be very useful. The approach is widely used to quantitatively determine melting and crystal-
lization temperatures (e.g., Berman et al., 1995; Ghiorso & Gualda, 2015; Palmer et al., 2015). However,
the minerals (e.g., olivine, clinopyroxene, biotite, and garnet) typically used as temperature proxyare
often not occurring in fault rocks, in particular in fault gouges. This unfortunately complicates a straight-
forward application to fault zone problems.
Thermal anomalies associated with seismic slip have been traced during the recent rapid postearthquake
fault drilling projects after major earthquakes, such as the TCDP, the WFSD, and the JFAST projects (respec-
tively, Fulton et al., 2013; Kano et al., 2006; H. Li et al., 2015). In contrast to expectation, only very weak bulk
temperature anomalies (<0.5°C) were identied across these fault zones. The thermal anomalies also appear
to be much lower than predicted from the fault strength deduced from the in situ stress measurements and
laboratory rock friction data (e.g., Brodsky et al., 2020; Fulton et al., 2013; Tanaka et al., 2006; Ujiie
et al., 2013). The temperature within a slip zone is locally elevated by coseismic frictional heating to quite
variable maximum temperatures, even at shallow crustal levels. This is surmised from widely reported ther-
mochemical reactions: clay mineral reactions (e.g., Brantut et al., 2008; L.W. Kuo et al., 2011), decarbona-
tion reactions (e.g., Han et al., 2007; Rowe, Fagereng, et al., 2012), amorphization (Aretusini et al., 2019;
Rowe et al., 2019), trace element partitioning features (e.g., Ishikawa et al., 2008; Tanikawa et al., 2015), gra-
phitization of carbonaceous material (e.g., Kuo et al., 2014; Oohashi et al., 2011), thermal maturation of
Figure 13. NRM record in the TCDP fault gouge that hosts the principal slip zone of the 1999 ChiChi earthquake
(Taiwan) (reproduced from Chou, Song, Aubourg, Lee, et al., 2012). (a) Equalarea stereoplot showing the
Chelungpu fault plane, the mean paleomagnetic components recorded in the three fault zones (FZB1136, FZB1194,
and FZB1243) and the wall rock of the TCDP HoleB, and the expected orientation of the earthquake lightning
(EQL) NRM, with an error of ±20° in declination. Solid and open symbols indicate the downward and upward
hemispheres, respectively. The wall rock's main component lies away from the modern magnetic eld, which is indicated
as the 1999 international geomagnetic reference eld (IGRF) model magnetic vector in central Taiwan by a bold plus
symbol (+). The FZB1136 fault gouge component is closest to the modern magnetic eld and statistically different from
the hypothetical EQL direction. In the FZB1194 and FZB1243 gouges, both normal and reversed components
are oriented southerly. (b) Orthogonal plot of AF demagnetization of the NRM in FZB1136 gouge (depth of 1,136.33 m).
Open and solid circles represent projection of the vector on the vertical and horizontal planes, respectively.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 27 of 60
organic molecules (e.g., Coffey et al., 2019; Rabinowitz et al., 2017; Savage et al., 2014), ssion track
signatures (e.g., D'Alessio et al., 2003), vitrinite reectance anomalies (e.g., Maekawa et al., 2014;
Sakaguchi et al., 2011), and occasionally the occurrence of pseudotachylytes (e.g., Otsuki et al., 2009;
Zhang et al., 2017). These diagnostic features provide important constraints on the temperatures in the
PSZ that prevailed during previous earthquake events. However, each of these approaches requires fairly
specic conditions to be applicable. For example, vitrinite reectance and biomarkers can only be applied
to sediments that contain certain types of organic matter, and ssion tracks require the presence of
uraniumrich inclusions in certain minerals. Also, most approaches only provide information on a certain
temperature threshold or a fairly narrow range in temperature. Vitrinite reectance, for instance,
indicates temperatures <400°C (Mukhopadhyay, 1992); higher temperatures cannot be diagnosed. Fission
tracks indicate ~120300°C because they are developed in that temperature window (Figure 14, also see
Yang, Dekkers, & Zhang, 2016, for more details). Thus, to achieve a complete picture of frictional heating,
yet other approaches are desired, complementary to existing methods.
Here, magnetic properties can be used at our advantage. Neoformation of ferrimagnetic minerals through
thermochemical reactions of Febearing minerals induced by frictional heating is deemed widespread in
seismic fault zones (see section 4.3). This opens up a new perspective in estimating the temperature
experienced during seismic slip (e.g., Chou, Song, Aubourg, Song, et al., 2012; Fukuchi et al., 2005,
2007; Han et al., 2007; Hirono et al., 2009; D. Liu et al., 2016; Mishima et al., 2006; Tanikawa et al., 2007,
2008; Yang et al., 2019, 2018; Yang, Dekkers, & Zhang, 2016). For example, a strong magnetic susceptibil-
ity anomaly in fault gouge was observed at the Fault Zone FZB1136, in TCDP HoleB (Figure 11a). It has
been attributed to the formation of large amounts of magnetite (Figures 11c11f): paramagnetic minerals
(e.g., pyrite, siderite, and chlorite) decomposed at high temperatures (>400°C) during the ChiChi earth-
quake (Chou, Song, Aubourg, et al., 2014; Mishima et al., 2006, 2009). Pyrrhotite was also identied
within the FZB1136 gouge as suggested by FORC diagrams (Figure 15): this provides evidence for thermal
decomposition of pyrite at temperatures >500°C (Chou, Song, Aubourg, Song, et al., 2012). The aforemen-
tioned occurrence of pyrrhotite in slip zones in the Japan Trench frontal prism sediments constrains the
peak temperature of seismic frictional heating to between 640°C and 800°C, combining magnetic infor-
mation with microscopy observations and pyritetopyrrhotite reaction kinetics (Yang et al., 2018).
These results provide crucial temperature constraints for understanding the earthquake energy dissipation
and weakening mechanisms, as frictional heating is regarded as the foremost factor controlling fault
weakening (e.g., Bizzarri, 2009; Yao et al., 2016).
Figure 14. Comparison of the proposed provisional rock magnetic geothermometer(Yang, Dekkers, & Zhang, 2016)
with other, more conventional, seismic frictional heating thermometers, including vitrinite reectance, ssiontrack,
and biomarkers. The text next to each approach refers to the conditions where the respective thermometers can be
applied. The bar widths indicate their working temperature ranges. The reader is referred to the main text and Yang,
Dekkers, and Zhang (2016) for more details.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 28 of 60
Recently, a preliminary rock magnetic fault zone geothermometerwas developed, which is utilizing mag-
netic susceptibility versus temperature measurements during cycling in the laboratory to increasingly ele-
vated temperatures (Yang, Dekkers, & Zhang, 2016). The thermochemical expression of magnetic phases
in a rock should not change unless the rock is heated in the laboratory beyond the maximum temperature
it underwent in nature (e.g., Hrouda et al., 2003; Spassov & Hus, 2006). This approach is very similar to esti-
mating temperatures of ancient res and burnt structures in archeology (Jordanova et al., 2018). In the Japan
Trench subduction plate boundary décollement, cored by the JFAST project, maximum temperatures were
determined ranging from ~300°C to over 500°C close to the multiple slip surfaces of previous earthquakes
(cf. Figure 6b in Yang, Dekkers, & Zhang, 2016). The rock magnetic geothermometer of fault rocks is less
dependent on lithology requirements. Moreover, it offers resolution in the 300700°C temperature range,
which is difcult to assess with other geothermometers (Figure 14). It thus is a promising complementary
approach for assessing temperature anomalies in slip zones.
Figure 15. (a) Backscatter SEM images of a pyrrhotitepyrite grain in the FZB1136 gouge in TCDP HoleB, Taiwan. The
framboidal pyrrhotite core has a 110 μm pyrite rim. With crossed polars under reected light microscopy the
pyrrhotite core remains bright and the isotropic pyrite rim is black (insets) (P: crosspolarized reected light image;
L: reected light image with plane polars). Pyrrhotite is inferred to have formed from hightemperature decomposition of
pyrite (>500°C) during coseismic slip of repeated earthquakes, while pyrite resulted from retrograde metamorphism
of pyrrhotite during cooling of coseismic uids. (b) FORC diagram for the FZB1136 gouge. A coercivity peak at 90 mT
reveals the presence of magnetically interacting pyrrhotite. (c and d) Modeled EhpH diagrams showing the stability
of pyrite and pyrrhotite at uid temperatures of 300°C and 100°C inferred for the fault gouge zone of FZB1136 in TCDP
HoleB. The pyrite stability eld at lower Eh values increases with decreasing temperature from 300°C to 100°C,
whereas that of pyrrhotite eld shrinks to a small range with high pH values (810) at 100°C, suggesting that pyrite
formed as a result of retrograde alteration of pyrrhotite. The presence of pyrite rims on pyrrhotite cores in the framboids
(Figure 15a) supports this retrograde interpretation. Compiled from Chou, Song, Aubourg, Song, et al. (2012).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 29 of 60
Substantial magnetic enhancement is also observed experimentally in highvelocity friction experiments on
natural fault gouges with increasing slip distance (e.g., Fukuchi et al., 2005; Tanikawa et al., 2007; Yang
et al., 2019). Magnetite, occurring as spherical and sintered irregular aggregates (Figure 16), was formed dur-
ing friction experiments on fault gouges from the YingxiuBeichuan fault. The original gouges are essentially
paramagnetic with one gouge containing minor goethite (~4 wt%); magnetite is the result of thermochemical
reactions of iron adsorbed on clay minerals (smectite and chlorite) and/or alteration/reduction of goethite.
Higher frictional heating temperatures (equivalent to longer slip distances) led to higher values of magnetic
susceptibility and magnetization and to larger grain sizes of the newly formed magnetite as indicated by
decreasing coercivity (Yang et al., 2019). Changes in magnetic parameters appear to be linear with the
(calculated) temperature rise induced by frictional heating (Figure 16f). Therefore, magnetic properties of
fault rocks can be used not only to detect seismic frictional heating but also to evaluate the amount of
frictional work associated with seismic slip (Tanikawa et al., 2007, 2008; Yang et al., 2019). Clearly, further
experimental work is required to establish a (semi)quantitative relation between magnetic parameters and
the heat production rate (and thus temperature rise) due to sliding. Only then meaningful extrapolation
to natural conditions would be allowed.
However, as discussed in section 4.5, uid inltration is widely present in a fault zone. During the postseis-
mic and longterm interseismic periods, uids of either meteoric or deep origin (or a combination) may inl-
trate into and percolate through the fault zone. Therefore, the latest coseismically imprinted magnetic
records (i.e., the aforementioned newly formed ferrimagnetic minerals, and consequent high magnetic sus-
ceptibility and magnetization) could be erased or overprinted soon after the earthquake. It is reported that
Figure 16. Rock magnetic properties of experimentally sheared natural fault gouge with a highvelocity apparatus (compiled from Yang et al., 2019).
(a) A schematic diagram of the rotary shear highvelocity friction (HVF) testing machine. (b) Sampling strategy after the HVF experiment. The sheared plane is
divided into ve annuluses with equivalent slip velocities during the experiments. The slip rate and displacement are 0 at the center and reach maximum
values at the edge. (c) Modeled temperature due to frictional heating in the different annuluses as a function of slip time. The reader is referred to Yang
et al. (2019) for more details on the equations and modeling strategy. (d) Hysteresis loops after and before (insets) correction for the paramagnetic contribution of
the starting material and run products after the experiments. The starting material is dominated by paramagnetic Febearing minerals; run products contain
an increasing amount of magnetite with slip distance. (e) SEM picture of an Fe oxide spherule with some adhered smaller particles present in the outermost
annulus of the sheared gouge. Inset is EDS spectrum for the analyzed spot marked by the red cross. (f) Magnetic susceptibility of run products linearly increases
with the estimated peak temperatures experienced during the experiment, suggesting enhancement of the newly formed magnetite with increasing
temperature due to frictional heating. GG indicates the starting material; red dots mark the run products. R
2
is the coefcient of determination of the line ts.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 30 of 60
fault properties can be measurably changed 2 years after an earthquake
(Brodsky et al., 2009), and the uid inltration would be one of the most
important factors driving these changes (Yang, Yang, et al., 2016). Thus,
there would be a rare opportunity to gain crucial information by observing
such transient magnetic changes related to a specic rupture only imme-
diately after a large earthquake (typically within a couple of years).
5.3. Fingerprint of Fluid Inltration in Fault Zones
Fluid ow, a major actor in fault zones, is intimately linked to the nuclea-
tion, propagation, arrest, and recurrence of earthquake ruptures (e.g.,
J. Chen et al., 2017; J. P. Evans & Chester, 1995; Goddard &
Evans, 1995; Hickman et al., 1995; Kerrich et al., 1980; Miller, 2013;
Rice, 2006; Terakawa et al., 2010; Williams et al., 2017; Zoback et al., 2007).
Detailed characterization of neoformed and/or altered magnetic minerals
due to uidrock interaction, as discussed in section 4.5 is thus important
for understanding the uidrelated processes.
In the 16 cm thick gouge of the Fault Zone FZB1136 in TCDP HoleB
drilled into the Chelungpu fault in Taiwan, the presence of goethite was
demonstrated by thermal demagnetization of NRM (Figure 11e), mea-
surements of the room temperature SIRM acquired in 2.5 T during cycling
between 300 and 400 K, and transmission Xray microscopy (Chou, Song,
Aubourg, Lee, et al., 2012; Chou, Song, Aubourg, Song, et al., 2012). Its
maximum concentration was estimated to be ~1% from the magnetic sus-
ceptibility and SIRM values of the gouge (Figure 11c). The goethite is
argued to have been formed during cooling of a hot coseismic uid
(>350°C) shortly after the 1999 ChiChi earthquake (Chou, Song,
Aubourg, Lee, et al., 2012; Chou, Song, Aubourg, Song, et al., 2012). The
coseismic uids were richer in iron in the center of the gouge zone than
at the edges, as demonstrated by the pattern of modeled goethite concen-
tration across the gouge (Figure 11c; see Chou, Song, Aubourg, et al., 2014, for detailed modeling strategy).
On cooling of hot uids after the earthquake, EhpH conditions are inferred to have changed (Chou, Song,
Aubourg, Song, et al., 2012, Figures 15c and 15d), which lednext to the formation of goethiteto partial
oxidation of magnetite, and some retrograde alteration of pyrrhotite to pyrite (Figure 15a).
Also in the rupture zone of the 2008 Wenchuan earthquake, a marked contrast in magnetic properties was
identied for the two components in the fault breccia, fragments, and matrix. The matrix has a relatively
higher magnetic susceptibility and magnetization but a lower coercivity than the fragments (Figure 17).
The contrast was attributed to fault uids that caused selective dissolution and precipitation of Febearing
minerals in the fault breccias during previous earthquake cycles. The matrix is more prone to uidrelated
effects than the fragments due to its ner grain size and thus larger specic surface (Yang, Chen, et al., 2013).
Magnetite concentrations were found to decrease starting with the host rock, via fault breccia, to (proto)cat-
aclasite. In addition, both lowand higheld magnetic susceptibility show close relationships with chlorite
that is a hydrothermal alteration product, as well as with the immobile elements (e.g., TiO
2
and P
2
O
5
; cf.
Figures 13 and 14 in Yang, Yang, et al., 2016). Since those elements are essentially immobile during
uidrock interaction processes, they can be used to infer the losses of mobile elements by mechanical wear
and dissolution, for example, by evaluating patterns in ratios of immobile to mobile elements. Finally, low
and higheld magnetic susceptibility are also positively correlated with the mass losses of fault rocks esti-
mated with the isocon method of geochemical mass balance (e.g., Tanaka et al., 2001). These observations
suggest that magnetite depletion and precipitation of Febearing clay minerals occurred in these fault
rocksexhumed from the shallow crustplumbed by uidassisted processes (Figure 18; see Yang, Yang,
et al., 2016, for details). Hence, magnetic properties of fault rocks represent an expression of uid inltration
within fault zones. Further work is necessary to establish a possible causal link between the specics of mag-
netic minerals and the fault uid properties (e.g., chemistry, temperature, and pressure).
Figure 17. (a) Sketch of a typical cemented fault breccia from the rupture
zone of the 2008 Wenchuan earthquake (Sichuan Province, China).
The fault breccia is characterized by the formation of host rock derived
fragments embedded within a nergrained matrix. (b) A plot of remanent
coercivity (B
cr
) versus the ratio of the saturation remanence and
magnetic susceptibility (M
rs
/χ) shows that the matrix, fragments of the fault
breccias, and bulk samples fall into three groups, suggesting the presence
of different magnetic carriers in the two components (fragments and
matrix) of fault breccias (adapted from Yang, Chen, et al., 2013).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 31 of 60
5.4. AMS and Strain Geometry in Fault Zones
Based on exchange equilibria, various mineral assemblages (e.g., phengite, sphalerite, hornblende, and clin-
opyroxene) have been used as geobarometerto dene differential stress conditions in crustal rocks (see
Reverdatto et al., 2019, for a review). However, these stress proxyminerals are usually absent in fault
zones. This is also the case for more conventional strain markers (e.g., fossils, ooids, pebbles, and nodules)
(Hayman et al., 2004). This makes it difcult to quantify the stress/strain conditions and deformation fabrics
in fault zones. Magnetic fabric analysis, validated for various rock types and in different structural settings
(e.g., Borradaile & Henry, 1997; Hrouda, 1982; MartínHernández et al., 2004; Parés, 2015; Tarling &
Hrouda, 1993), is a widely used tool to examine deformation in geologic structures characterized by large
strain gradients such as shear zones (for a review, see Ferré, Gébelin, et al., 2014). However, to date only a
few studies have targeted brittle fault rocks in the shallow crust (e.g., CasasSainz et al., 2018; Chou, Song,
Lee, et al., 2014; Louis et al., 2008; Marcén et al., 2019; Nakamura & Nagahama, 2001; RománBerdiel
et al., 2019; Solum & van der Pluijm, 2009; Yeh et al., 2007). In contrast to the ductile mylonitic rocks, brittle
fault rocks are typied by frictional sliding as dominant deformation process (e.g., Schmid & Handy, 1991;
Sibson, 1977; see section 2 for details). This involves grain rotation, fracturing, and translation of (magnetic)
minerals in fault rocks. The shear effects are reected in magnetic fabric changes and provide clues on the
strain states in fault zones, as outlined below.
5.4.1. Kinematics of Fault Slip
The orientation of the magnetic fabric provides constraints on the kinematics of fault slip (e.g., CasasSainz
et al., 2018; Elhanati et al., 2020; Ferré et al., 2015, 2016; Hayman et al., 2004; Levi et al., 2014; Marcén
et al., 2019; Solum & van der Pluijm, 2009). In the seismically active Dead Sea Basin (Israel), for example,
the original depositional fabric, characterized by a subvertical κ
min
axis and scattered κ
max
and κ
int
axes, is
well preserved ~2 m away from the fault plane. In contrast, the principal AMS axes of the
deformationdriven fabrics, which could be observed up to tens of cm from the fault plane, are coaxial with
the instantaneous strain ellipsoids calculated with the fault plane solutions (Figure 19). Such magnetic fab-
rics are interpreted to result from the local strain eld during the past earthquakes (Levi et al., 2014).
Another illustrative example is the study of fault rocks from the Chelungpu fault (Taiwan). AMS fabrics
Figure 18. A simple conceptual model illustrating the uidassisted cataclastic process (reprinted from Yang, Yang,
et al., 2016). Cataclasis transforms a granite protolith with magnetite as the dominant magnetic carrier to essentially
paramagnetic fault rocks, that is, cataclasite, fault breccia, and gouge, in the YingxiuBeichuan fault (Longmen Shan
thrust belt, China), which accommodated most of the displacement of the 2008 Mw 7.9 Wenchuan earthquake
(for details the reader is referred to Yang, Yang, et al., 2016).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 32 of 60
revealed that the compressional stress orientation is identical to the regional stress regime and that the mag-
netic foliation planes of fault gouges are consistent with thrust dipslip displacement across the gouge zones
(Yeh et al., 2007). Also recently, Ferré et al. (2015) reported for the rst time a successful determination of the
full kinematic solution (slip plane, direction, and sense) of a prehistoric (20.1 ± 0.5 Ma) seismic event
through AMS analysis of pseudotachylyte veins in the Dora Maira Massif (Italy; Figure 20). The analysis
relies on the notion that AMS of fault pseudotachylytes arises from coseismic viscous ow of the frictional
melt and therefore is tracking the direction of slip (Ferré et al., 2015). These studies demonstrate that mag-
netic fabrics can be employed not only to detect the local strain eld but also to determine fault plane solu-
tions of previous earthquakes.
5.4.2. Strain State Across Fault Zones
Stress and strain information across a fault zone are crucial ingredients for our understanding of earthquake
generation and related slip behavior. Rocks that host (future) faults, in particular unlithied and
Figure 19. Comparison of magnetic fabrics and fault plane solutions across a normal fault in the Lisan Formation,
Masada Plain (Israel), in the Dead Sea basin, where numerous M> 6 earthquake events occurred during the last
70,000 years (compiled from Levi et al., 2014). (a) Lower hemisphere equalarea projections of principal AMS axes of the
late Pleistocene soft rocks in the footwall (I), intermediate block (II), and hanging wall (III). Dashed lines indicate
the strike direction of the nearby fault strands. The two small stereograms in Figure 19b show the orientations of the fault
strands. (c) Two fault planes and momenttensor solutions calculated for the footwall and hanging wall. Dashed
arrows mark the dipslip direction. ε
max
,ε
int
, and ε
min
are the instantaneous maximum, intermediate, and least principal
strain axes, respectively. The principal AMS axes of soft rocks close to the three late Pleistocene syndepositional
normal faults resemble the fault plane solutions.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 33 of 60
unconsolidated sediments, respond to the progressive tectonic loading by the development of a preferred
orientation of the constituent (magnetic) minerals. This fabric is commonly coaxial with the axes of the
strain ellipsoid; more specically, the pole to the magnetic foliation is widely regarded as the orientation
of the maximum shortening strain (e.g., Cifelli et al., 2004; Parés, 2015). Therefore, comparison of the
magnetic fabrics in host rocks bounding the fault zone and fault rocks sheds light on how the strain state
varies across the fault zone. For example, AMS analysis revealed an abrupt change in strain state across
the plate boundary fault zone (décollement) at shallow depths near the Japan Trench (Yang, Mishima,
et al., 2013): The AMS data of the prism sediments above the décollement depict the maximum strain direc-
tion of lateral shortening due to the WNW convergence of the Pacic plate, while those of the underthrust
sediments appear to point to a classic sedimentary compaction fabric, thus a vertical, uniaxial strain
(Figure 21). The plate boundary décollement between the Pacic and North American plates is thus
decoupled over the long term, which favors the propagation of coseismic slip along it (Yang, Mishima,
et al., 2013). Such strain decoupling across a décollement zone revealed by AMS fabrics was also reported
at the Nankai Trough (Owens, 1993; Ujiie et al., 2003), the Costa Rica margin (Housen &
Kanamatsu, 2003), the Barbados accretionary margin (Hounslow, 1990; Housen, Tobin, et al., 1996), and
the Chile Triple Junction region (Collombat et al., 1995). AMS is thus an excellent indicator for incremental
strain that sensitively reects the stress state across a fault zone.
5.4.3. Deformation Mechanism of Fault Zones in the Shallow Crust
Fault breccia and cataclasite are considered to have been deformed via cataclastic ow (Hayman et al., 2004).
It is a specic deformation mechanism that involves continuous brittle fracturing and comminution
Figure 20. Focal mechanism of a prehistoric earthquake deduced from magnetic fabrics of pseudotachylyte veins
(the Dora Maira massif, western Italian Alps) (compiled from Ferré et al., 2015). (a) Macroscopic view of the
pseudotachylyte generation vein developed in the host mylonitic gneiss. (b) Comparison of magnetic fabrics of the host
gneiss (in black) and pseudotachylyte (in red). The host gneiss (with corrected degree of magnetic anisotropy P
j
of
1.38 ± 0.10) has a subhorizontal magnetic foliation and magnetic lineation. In contrast, the pseudotachylyte foliation
(with P
j
of 1.08 ± 0.01) has an angle of 31° with that of the host gneiss. Itis formed as a result of rapid cooling of the melt
and recorded viscous ow parallel to the seismic slip direction (c), and thus revealing a toptowest sense of shear
and a normal fault focal mechanism.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 34 of 60
of grains with frictional sliding and rolling of the fragments with respect to one another (Twiss &
Moores, 1992). Magnetic grain size reduction occurring during cataclastic ow is expected to decrease the
grainscale anisotropy of the AMS carriers (Ferré, Gébelin, et al., 2014; Jackson et al., 1993). This may
explain why notably weak and even nearly isotropic AMS tensors were determined in strongly fractured
granites in the Nojima Fault in Japan (Nakamura & Nagahama, 2001). In contrast, fault gouges are inferred
to be the product of granular ow (Twiss & Moores, 1992) which involves the rolling and sliding of rigid
particles (Morgan, 1999; Morgan & Boettcher, 1999). In this process, the physical rearrangement of the mag-
netic minerals primarily results in an increase in the AMS parameters (e.g., Pand F). One should bear in
mind, however, that brittle fault rocks often host secondary minerals that not necessarily relate directly to
strain. For example, in the Fault Zone FZB1136 in TCDP HoleB the magnetic foliation and anisotropy
degree are lowest in the PSZ itself, whereas the highest values appear in the fault gouge where goethite is
present (Chou, Song, Lee, et al., 2014). The goethite was formed postseismically by the action of
thermal uids (Chou, Song, Aubourg, Song, et al., 2012, see also section 5.3). Therefore, caution is appropri-
ate when interpreting the deformation mechanisms based on the AMS analysis of brittle fault rocks (e.g.,
RománBerdiel et al., 2019). Microscopic inspection of whether or not secondary minerals are present is
recommended.
Figure 21. (a) The general plate conguration of the Japanese island arc and map of the 2011 Mw 9.0 Tohokuoki earthquake portion of the Japan Trench,
showing the location of the IODP Expedition 343 drill site with a lled white circle. The red star indicates the epicenter of the earthquake. White arrow indicates
the Pacic plate (PAC) convergence vector with a rate of85 mm/year (Argus et al., 2011). (b) Downcore proles of degree of anisotropy (P) and magnetic
foliation (F). Both progressively increase with depth in the frontal prism and suddenly drop just below the décollement, mostly showing small values in the
underthrust sediments. Equalarea, lower hemisphere projections of the bootstrapped κ
max
,κ
int
, and κ
min
directions for (c) the prism sediments (n= 34) and (d)
the underthrust sediments (n= 22). The AMS fabric pattern of the prism sediments is consistent with horizontal tectonic shortening (red arrows in Figure 21c)
nearly parallel to the PAC convergence direction in the Japan Trench, as well as the maximum horizontal stress deduced from borehole breakouts
(W. Lin et al., 2013). The AMS in the underthrust sediments represents a vertical, uniaxial strain, that is, an initial sedimentary compaction fabric. These very
different AMS patterns reveal an abrupt strain change across the plate boundary décollement at shallow depths and imply that large coseismic slip occurred along
a weak décollement that is mainly decoupled over the long term. Compiled from Yang, Mishima, et al. (2013) with data from Argus et al. (2011) and W. Lin
et al. (2013).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 35 of 60
The magnetic fabric of pseudotachylytes is providing a snapshot of the ambient stress during or immediately
after seismic slip (e.g., Ferré et al., 2015, 2016; MolinaGarza et al., 2009). In the Chiapas Massif (Mexico), for
example, the magnetic fabrics of the host granitoid rock and the pseudotachylyte veins are essentially
parallel, with nearly eastwest and steep magnetic foliations. However, the AMS fabric of the host rock is
dominantly prolate, while that of the pseudotachylytes veins is markedly oblate. This suggests that the veins
underwent a pure shear deformation (i.e., viscous breaking) during cooling (MolinaGarza et al., 2009),
again illustrating that AMS analysis delivers important insights into the deformation mechanism of
fault zones.
5.5. Magnetic Records of Coseismic Electric Currents
As mentioned earlier, several faultingrelated magnetization processes contribute to the NRM acquisition of
pseudotachylytes, including coseismic TRM acquired upon cooling of the melt below the Curie temperature,
EQLinduced IRM acquisition (EQLIRM), and postseismic CRM due to both devitrication and alteration
(Ferré et al., 2012; Leibovitz, 2016). Deciphering these magnetization components is crucial to the under-
standing the pseudotachylytes' microstructures and the timing of their development.
Coseismic TRM is restricted to the local melt pockets and records the direction and intensity of the geomag-
netic eld at the time of cooling. It is straightforwardly demagnetized thermally in the laboratory. In con-
trast, EQLIRM typically displays an anomalously high NRM value in pseudotachylytes with characteristic
directions at a high angle to the fault plane. There is no geometric relationship to the past or present direc-
tions of the geomagnetic eld (Leibovitz, 2016). Thus, rock magnetic and paleomagnetic analyses of the
pseudotachylytes do provide independent constraints on the origin(s) of the NRM in pseudotachylytes.
Anomalously high NRM recorded in a pseudotachylyte, which would require a magnetic eld (much) stron-
ger than the typical Earth's magnetic eld, then points to EQLIRM as the dominant NRM acquisition
process.
Indeed, such abnormally high NRM intensities have been reported for fault rocks (fault gouge and pseudo-
tachylytes) along or near the fault plane where large earthquakes occurred (e.g., Enomoto & Zheng, 1998;
Enomoto et al., 2001; Ferré et al., 2005, 2012; Ferré, Geissman, et al., 2014). For instance, pseudotachylyte
with an abnormally high NRM (~132 A/m), up to 300 times higher than that of the granitic host rock
(tonalite), was found in the Santa Rosa Mountains near Palm Springs, California, USA (Ferré et al., 2005).
These pseudotachylyte veins typically have a moderatecoercivity, welldened, single component NRM iso-
lated with stepwise progressive alternating eld demagnetization (cf. Figure 6a in Ferré et al., 2005). For
comparison, typical NRM values of volcanic rocks that cool in the Earth's magnetic eld range from 0.1 to
10 A/m depending on their chemical composition. Mac and ultramac rocks are more magnetic than felsic
rocks. To acquire the reported high NRM, the fault rocks and pseudotachylytes must have been exposed to a
strong local magnetic eld, several orders of magnitude higher than the typical Earth's main eld (Ferré
et al., 2005). The most likely source of the required strong magnetic eld is believed to be the coseismic elec-
tric currents produced by EQL (Ferré et al., 2005). In the Nojima Fault (Japan), the coseismic electric current
was estimated to be as high as ~1 kA during the 1995 Kobe earthquake through simulation of the
EQLinduced sintering of the fault gouge using spark plasma sintering experiments (Enomoto et al., 2001).
Such a large current is expected to produce a strong pulsed magnetic eld, in which the foliated fault gouges
acquired an IRM which resulted in a very large NRM: ~430 times higher than that of the granitic fault brec-
cia, and ~70 times higher than that of the adjacent mudstone (Enomoto et al., 2001; Enomoto &
Zheng, 1998).
6. Comparison of Magnetic Methods to Other Approaches
The magnetic properties of fault rocks can provide intriguing insights into faulting processes. The number of
magnetic studies on fault rocks, however, is rather limited to date. A broadscale comparison of fault rock
magnetic data with other more classical approaches is therefore currently in its infancy. Below we assemble
the information concerning the temperature of frictional heating and the stress/strain state of the fault
retrieved from magnetic and nonmagnetic thermal and strain indicators. This comparison is only possible
in a few fault zones that accommodated large earthquakes, where data from multiple techniques are
available.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 36 of 60
6.1. The Seismic Frictional Heating Temperature
Frictional heating during seismic slip is evidently registered in all minerals in the slip zone, not only in mag-
netic minerals. Nonmagnetic minerals include hydrous, clay, and carbonate minerals. Thus, frictional heat-
ing might be traced and cross calibrated by a variety of thermal indicators (Table 1). In the Fault Zone
FZ1136, TCDP HoleB, the peak temperature during the 1999 ChiChi earthquake was >400500°C, inferred
from the thermal decomposition of paramagnetic minerals with the formation of ferrimagnetic minerals
(Table 1; Chou, Song, Aubourg, Song, et al., 2012; Mishima et al., 2006, 2009; Tanikawa et al., 2007, 2008).
This estimate concurs with others based on nonmagnetic thermal indicators: The signature of mobile trace
elements in the fault uid indicates a frictional heating peak temperature >350°C (Ishikawa et al., 2008).
That temperature is estimated at >300400°C based on the vaporization of water during thermal pressuriza-
tion (Boullier et al., 2009). These peak temperatures estimates are lower than the estimate of 626 ± 25°C
from a kinetic model of vitrinite thermal maturation (Maekawa et al., 2014), and the infrared and Raman
estimates from carbonaceous fault materials (700°C, Hirono et al., 2015). The clay mineral signature would
indicate that the peak temperature could even be up to 9001100°C (L.W. Kuo et al., 2011). Most recently,
nanometric geochemical analysis of fault gouge revealed that most of the siderite in the slip zone has
been evaporated into nanosized grains, suggesting that the peak temperature reached as high as ~1200°C
(W.H. Li et al., 2019). While the actual peak temperature is still not well dened, the existence of frictional
heating has been denitively veried by the 0.06°C residual temperature anomaly measured 6 years after the
1999 ChiChi earthquake rupture through direct borehole temperature logging (Table 1; Kano et al., 2006). A
similar situation exists in the Japan Trench subduction plate boundary fault zone which accommodated the
large slip of the 2011 Tohokuoki earthquake (Table 1) with a thermal of anomaly of 0.31°C measured
16 months after rupture (Fulton et al., 2013).
Table 1
Comparison of Peak Temperatures of Frictional Heating Estimated by Magnetic and Nonmagnetic Approaches in Three Fault Zones That Accommodated Large
Magnitude Earthquakes
Peak temperatures Indicators/methods Note
TCDP Hole B Fault Zone FZ1136; hosted the 1999 ChiChi (Taiwan) earthquake (Mw 7.6)
>400°C Thermal decomposition of paramagnetic minerals
a,b,c,d
Magnetic indicators
>500°C Pyrrhotite formation via thermal breakdown of pyrite
e
>350°C Fluidmobile trace element anomalies
f
Nonmagnetic indicators/methods
>300400°C Vaporization of water during thermal pressurization
g
626 ± 25°C Vitrinite reectance
h
~700°C Infrared and Raman spectra of carbonaceous materials
i
9001100°C Clay mineral anomalies (FZ1111 in Hole A that corresponds to FZ1136 Hole B)
j
~1200°C Thermal decomposition or breakdown of siderite into nanosized grains
k
0.06°C residual temperature anomaly
6 years after the earthquake
Direct borehole observation
l
Japan Trench subduction plate boundary fault zone; hosted the 2011 Mw 9.0 Tohokuoki earthquake
>300500°C Rock magnetic geothermometer
m
Magnetic indicators
640800°C Reaction of pyrite to pyrrhotite
n
120900°C Organic biomarker thermal maturity
o
Nonmagnetic indicators/methods
0.31°C residual temperature anomaly
16 months after the earthquake
Direct borehole observation
p
<1250°C Modeled with plausible slip durations and slip zone thicknesses, and a
friction coefcient of 0.08
p
Nankai Subduction Zone (Japan) associated with the 1944 Tonankai Mw 8.1 earthquake
<400°C No thermal decomposition of paramagnetic minerals occurred
q
Magnetic indicators
<300°C Fluidmobile trace elements, Raman spectra of carbonaceous material,
and inorganic carbon content
q
Nonmagnetic indicators/methods
390 ± 50°C Vitrinite reectance
r
<250°C Infrared spectroscopy, clay mineral anomalies, traceelement geochemistry,
and isotope geochemistry
s
a
Mishima et al. (2006).
b
Mishima et al. (2009).
c
Tanikawa et al. (2007).
d
Tanikawa et al. (2008).
e
Chou, Song, Aubourg, Song, et al. (2012).
f
Ishikawa
et al. (2008).
g
Boullier et al. (2009).
h
Maekawa et al. (2014).
i
Hirono et al. (2015).
j
L.W. Kuo et al. (2011).
k
W.H. Li et al. (2019).
l
Kano et al. (2006).
m
Yang, Dekkers, and Zhang (2016).
n
Yang et al. (2018).
o
Rabinowitz et al. (2020).
p
Fulton et al. (2013).
q
Hirono et al. (2009).
r
Sakaguchi et al. (2011).
s
Hirono et al. (2014).
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 37 of 60
The peak temperature estimates obtained by different methods thus appear to be quite variable. A possible
reason could be that different thermal proxies sense correctly certain temperature intervals only
(cf. Figure 14, see also Rowe & Grifth, 2015; Yang, Dekkers, & Zhang, 2016). Another reason could be that
the majority of the temperature estimates take equilibrium conditions as their basic premise, which may not
apply to the timescales of fault slip (e.g., Savage et al., 2018). This clearly warrants further work. However, at
present we may deduce that peak temperature estimates of fault rocks inferred from magnetic properties are
reasonably in line with the majority of other methods and provide at least a lower bound on the temperature
rise due to seismic frictional heating.
In other cases examination of magnetic assemblages can provide an upper limit for the temperature of the
frictional heating. This is the case in the Nankai subduction zone (Japan) associated with the 1944
Tonankai Mw 8.1 earthquake, where the absence of thermal decomposition of paramagnetic minerals indi-
cates a peak temperature <400°C (Table 1; Hirono et al., 2009). This peak temperature estimate is corrobo-
rated by estimates from other thermal indicators, such as the uidmobile trace element signature, the
Raman spectral features of carbonaceous material, and the inorganic carbon content (these parameters all
indicate <300°C, Hirono et al., 2009, 2014). Also, vitrinite reectance data (peak temperature 390°C,
cf. Sakaguchi et al., 2011) support the magnetic information. It further shows that magnetic property
analysis of fault rocks can provide at least a reasonably robust thermal record of earthquake slip. It should
be realized that the amount of frictional heating can vary widely for different earthquakes; the related
temperature rise is as yet rather poorly constrained by any individual method. Therefore, all approaches
are subject to ongoing testing and renement. Combined application, reconciliation and cross calibration
of different thermometers would thus yield the most robust picture. This is also where the magnetic methods
will be playing an important role.
6.2. Stress Orientations
In the Japan Trench subduction fault zone, the maximum AMS axes of sedimentary rocks in the frontal
prism align northeastsouthwest (36.6 ± 11.3°, Figure 21c, Yang, Mishima, et al., 2013). This is nearly per-
pendicular to the direction of the Pacic plate convergence (~292°, Argus et al., 2011) and the maximum hor-
izontal stress orientation (319 ± 23°) deduced from the azimuths of borehole breakouts (W. Lin et al., 2013;
see Appendix A for denition). The accretionary prism sediments thus probably have undergone layer par-
allel shortening due to the convergence of the Pacic plate. Similar AMS patterns have also been reported in
the TCDP core samples, where the magnetic ellipsoids are coaxial with the (palaeo)stress ellipsoids deduced
from fold axes, fault analysis, and borehole breakouts (cf. Figure 14b in Louis et al., 2008). Hence, these
observations provide a strong validation of the reliability of AMS as a (palaeo)stress indicator within fault
systems.
7. Remaining Challenges
The foregoing discussion illustrates that the rock magnetic approach is very promising to unveil valuable
information pertaining to faulting processes. Comparison of magnetic and nonmagnetic approaches, as
low as the number of comparative studies at this moment, yields a broad agreement among most of the
approaches. Moreover, they have improved our understanding of the magnetic signatures of fault rocks.
Nonetheless, some essential questions pertaining to the mechanisms that underlie the magnetic changes in
fault rocks remain currently unanswered. Most importantly, the reaction kinetics of Febearing (magnetic)
mineral reactions under the extreme faulting conditions (i.e., fast heating, intensive mechanical milling, and
nanocrystallization) are not well understood. On the other hand, other than the protolith lithology, magnetic
properties of fault rocks are strongly affected not only by physicochemical conditions (temperature, pres-
sure, oxygen fugacity, pH, etc.) during rupture but also by alteration and chemical differentiation within a
fault zone during the other stages of the seismic cycle, that is, the postseismic, and interseismic periods
(e.g., Bense et al., 2013; Ferré et al., 2017; O'Hara & Huggins, 2005; Yamaguchi et al., 2011). This complicates
the magnetic signature interpretation of fault rocks. The individual impact of each stage of the seismic cycle
must be untangled before a full understanding of faulting behavior is obtained. For example, uidrock
interaction plays a key role in the alteration, removal, and/or neoformation of Febearing minerals in fault
zones. Unfortunately, the associated kinetics are presently rather poorly understood (e.g., Chou, Song,
Aubourg, Song, et al., 2012; Yang, Chen, et al., 2013; Yang, Yang, et al., 2016). Also, identication of a
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 38 of 60
PSZ and the signature of frictional heating of individual earthquake events are not one to one, as a fault
zone is the cumulative product of many earthquake cycles, and thus, several PSZs may stack together in a
fault gouge. These challenges warrant future work.
Next, we identify some attractive research avenues to help addressing these concerns, including (1) extensive
magnetic analysis on fault rocks after friction experiments, (2) laboratory simulation of fault uid percola-
tion, (3) emerging technology for resolving faulting imprinted NRM components, (4) magnetic fabric analy-
sis of brittle fault rocks, and (5) synergy of interdisciplinary approaches in mineralmagnetic studies.
7.1. Extensive Magnetic Analysis on Fault Rocks After Frictional Experiments
Friction impacts the formation and/or destruction of Febearing (magnetic) minerals. Mechanical milling in
air can completely transform hematite to magnetite (Petrovský et al., 1996, 2000; Zdujićet al., 1998);
mechanical friction in argon favors the thermal decomposition of pyrite into pyrrhotite (H.P. Hu et al., 2002)
while mechanochemical decomposition of siderite due to crushing may produce magnetite (Criado
et al., 1988). Similar processes can readily occur in seismic slip zones during earthquakes (e.g., J. P. Evans
et al., 2014). Experimental calibration of the reaction kinetics on timetemperature paths similar to those
expected on fault surfaces during seismic slip may enable precise determination of the concomitant tempera-
ture rise. Therefore, laboratory highspeed friction experiments using rotary shear apparatus (e.g., Di Toro
et al., 2010; S. Ma et al., 2014) on natural and synthetic fault gouges are promising to precisely constrain
the processes that occurred in a fault zone during rupture. Such experiments could be carried out on
samples with various compositions focusing on the suite of Febearing minerals and under different condi-
tions (e.g., variable oxygen fugacity, permeable/impermeable, high vapor pressure, hydrostatic, and elevated
temperature). Realtime records of physical parameters, that is, temperature, velocity, stress, displacement,
provide highresolution information on the ashheatingeffects due to the friction. Therefore, studies
should ideally integrate mineral magnetic, mineralogical, geochemical, and microscopic aspects on fault
gouges before and after experimental shearing. Also, they should be complemented by kinetic chemical
modeling and temperature modeling. With such data, our understanding of the evolution of Febearing
(magnetic) mineral(s) within the PSZ during slip will be put on a much stronger footing. In addition, similar
experiments under lowspeed conditions (as low as the mm/year range, cf. Collettini et al., 2014, and S. Ma
et al., 2014) may shed light on the possible magnetic response during earthquake nucleation.
7.2. Laboratory Simulation of Fluid Percolation
Fluids along fault zones are generally of multiple sources (Zoback et al., 2007). Their migration paths and
compositions vary markedly over time and space, as fault zones act as uid conduits or barriers at different
stages in their development, that is, the coseismic, postseismic, and interseismic periods (Bense et al., 2013).
This leads to temporal and spatial changes of the redox state in a fault zone as a function of episodic fault
rupture and healing cycles during the lifetime of a fault (Yamaguchi et al., 2011). Laboratory simulation
of such processes with natural and/or synthetic fault rocks under controlled physicochemical conditions
(i.e., temperature, pressure, redox state, pH, and uidrock ratio) thus would be a promising way to improve
our understanding of alteration, removal, and/or neoformation of magnetic minerals as a result of uid inl-
tration. Answers may be retrieved from characterizing the Febearing and associated minerals which are
precipitated or recrystallized, such as iron oxides, iron hydroxides, iron carbonates, and/or iron suldes,
in incorporation with thermochemical equilibrium modeling under the specic experimental physicochem-
ical conditions.
7.3. Emerging Technology for Resolving Faulting Imprinted NRM Components
Isolation and characterization of faultinginduced NRM components are expected to yield valuable informa-
tion on faultingrelated effects. With the exception of a few early studies on paleomagnetic dating of fault
rocks (e.g., Grønlie & Torsvik, 1989; Hailwood et al., 1992; Torsvik et al., 1992); however, little attention
has been paid to newly imprinted secondarycoseismic/postseismic NRM components, such as related
(p)TRM or CRM (e.g., Chou, Song, Aubourg, Lee, et al., 2012, see section 5.3) and potential EQLIRM
(e.g., Ferré et al., 2012; see section 5.5). This is most likely dictated by the amount of sample material needed
for the required precise determination of paleomagnetic signatures by classic paleomagnetic measurements.
Alternative options include the aforementioned SSM and QDM. The SSM is capable of mapping magnetic
elds with high spatial resolution (better than 100 μm at room temperature and 4 μm or better at
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 39 of 60
cryogenic temperatures), with a limiting magnetic moment sensitivity of ~1 × 10
14
Am
2
at a sensortosam-
ple distance of 200 μm (e.g., Fukuzawa et al., 2017; Pastore et al., 2018). It has been successfully applied in
various paleomagnetic studies, such as on lunar glass spherules and Hawaiian basalt (Weiss et al., 2007) and
ferromanganese crust (Noguchi et al., 2017). In the fault slip context, magnetic stripes with positive and
negative polarities subparallel to the slip surface were recently detected by Fukuzawa et al. (2017) on the
Nojima Fault gouge (Japan) within an area of 1 cm × 1 cm, using SSM. Such NRM distribution patterns
in fault gouge could provide a means to distinguish prehistorical periods of earthquake activity on the slip
zone with the help of the geomagnetic polarity timescale. The QDM is a combination of superior spatial reso-
lution (5 μm), magnetic sensitivity (20 μTμm/Hz
1/2
), and wide eld of view (4 mm) (Glenn et al., 2017). It
can image both remanent and induced magnetization of geological samples at room temperature (Glenn
et al., 2017). These emerging microscopic magnetic scanning techniques enable micrometerscale paleomag-
netic studies of fault zones. They also offer spatially wellresolved measurements of magnetic signals of the
narrow fault zones. Thus, they represent a very promising way to resolve the cumulative frictional heating
effects in PSZs caused by recurrent (historical) earthquakes.
7.4. Magnetic Fabric Analysis of Brittle Fault Rocks
Fault zones are most complex study targets. The caveats put forward by Ferré, Gébelin, et al. (2014) for duc-
tile shear zones also apply to brittle fault rocks in the shallow crust. First, mineral assemblages and mechan-
ical properties of fault zones are heterogeneous between and within the different fault zone compartments
that is, damage zone, fault breccia, and fault gouge. The AMS fabrics in each zone could well have their indi-
vidual strain relationship. Meanwhile, faultingrelated process (frictional heating and uidrock interac-
tions) tends to induce alteration of mineral assemblages that would not necessarily relate to strain (e.g.,
Just et al., 2004). The observed changes in AMS thus cannot be solely attributed to deformation and evidently
complicate the relation of AMS with strain in a fault zone. As an example, the AMS fabric of fault gouges in
the Nojima Fault (Japan) is isotropic (Nakamura & Nagahama, 2001). However, it is anisotropic in the
Chelungpu Fault in Taiwan (Yeh et al., 2007). Integrated rock magnetism, mineralogy and geochemistry,
microstructural observations, and possibly the determination of other magnetic fabrics (high eld and low
temperature) and the AMR analysis are desired to fully unravel the relation between magnetic fabric and
the deformation mechanisms operating during faulting.
In addition, the neoformed magnetic minerals and paramagnetic/diamagnetic minerals in fault rocks may
well be formed at different faulting stages or events and could have distinct deformation paths. For example,
diamagnetic fabric of coccolith calcite from the plate boundary Dead Sea Fault indicates tectonic strain,
while the paramagnetic fabric of clays preserves the depositional fabric (Issachar et al., 2018). Thus, evalua-
tion of the relative contributions from the ferrimagnetic, paramagnetic, and diamagnetic components to the
bulk AMS (e.g., Ferré et al., 2004; Hrouda & Jelínek, 1990; Kelso et al., 2002; MartínHernández &
Ferré, 2007) and examination of the (weak) eld dependence of AMS, as well as the outofphase and in
phase components of AMS, could be insightful into the dynamic deformation processes in a fault zone.
7.5. Synergy of Interdisciplinary Approaches in MineralMagnetic Studies
The various structural domains of a fault zone that is, fault core, damage zone, and host rock, have a very
diverse level of faulting effects and number of rupture episodes. Attention should thus be paid not only to
bulk magnetic properties of these different fault rocks (e.g., fault breccia, cataclasite, and fault gouge) but
also the different components of fault rocks should be considered individually. Fault breccia, for instance,
generally comprises a mixture of fragments with different grain sizes, set in a nergrained, clayrich matrix
(Figure 17a). With their heterogeneous hydraulic structure and specic surface, alteration associated with
faulting processes (uid percolation in particular) acts on them to quite different extents (Yang,
Chen, et al., 2013; Yang, Yang, et al., 2016). Consequently, an assessment of the relative contributions of dif-
ferent components in fault rocks to the overall magnetization, and unraveling their individual alteration
processes are of great help for understanding the changes in magnetic minerals as a fault zone evolves.
This could be achieved through combination with mineralogical, geochemical, and microscopic analyses,
other than mere rock magnetic examination of the ironbearing minerals (e.g., Fe oxides, Fe hydroxides,
Fe suldes, Fe carbonates, or Fesilicates) in fault materials.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 40 of 60
The thickness of PSZ and microstructures in fault zones is often at the millimeter to even micrometer scale
(e.g., Boullier, 2011; Fondriest et al., 2013; Sibson, 2003; SimanTov et al., 2013). Thus, with magnetic studies
at a microscale one should be able to gather most meaningful information on temperature changes within a
PSZ. Next to the abovementioned SSM and QDM, advanced techniques for magnetic imaging at microscale/
nanoscale are recommended in future studies, including but not limited to magnetic tunnel junction for
mapping magnetic elds with high spatial resolution and eld sensitivity (e.g., Lima et al., 2014) and
synchrotron radiation analysis that enables imaging the threedimensional structures of nanosized magnetic
minerals (e.g., Chou, Song, Tsao, et al., 2014), as well as the deformation microfabrics in fault zones
(e.g., Fusseis et al., 2014). Orientation statistics from threedimensional structure of magnetic particles then
become valuable, which could be matched with, for example, AMS or other magnetic techniques (e.g., Petri
et al., 2020).
Figure 22. Aowchart for the magnetic property analysis of fault rocks. Multidisciplinary and integrated approaches
incorporating eld studies, laboratory simulations, highresolution microscopy, and numerical modeling are proposed as
a most promising way to gain a full appreciation of the magnetic response to dynamic physicochemical processes in
fault zones. With a research effort along such lines, better rock magnetic strain indicators,”“geothermometers,and
uid tracersare expected to become available for fault zone studies. XRD, Xray diffraction; XRF, Xray uorescence;
SEM, scanning electron microscopy; TEM, transmission electron microscopy; EDS, energy dispersive spectroscopy;
EPMA, electron probe microanalysis; ESR, electron spin resonance; DRS, diffuse reectance spectroscopy; MFM,
magnetic force microscopy; SSM, scanning SQUID microscopy; and QDM, quantum diamond magnetometry.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 41 of 60
8. Conclusions
We have examined the pros and cons of rock magnetic methods applied to fault zones. After identifying can-
didate mechanisms for magnetic changes in fault rocks, we summarize the recent advances of such studies
on fault rocks, highlighting the insights that were gained from these studies. It has been demonstrated that
the rock magnetic properties can be employed as strain indicator,geothermometer, and uid tracerin
fault zones. This opens a new window into investigating faulting processes.
This emerging subdiscipline is promising and challenging at the same time. Magnetic properties of fault
rocks hold many secrets that await discovery; multidisciplinary and integrated approaches are crucial to gain
a full appreciation of magnetic response to dynamic physicochemical processes in fault zones (Figure 22).
Other than magnetic investigations of fault rocks from earthquake faults, laboratory experiments (friction
experiments, milling, rapid thermal treatment, and uid percolation, etc.) on both natural and synthetic
fault rocks should be taken into consideration. Additionally, kinetic, equilibrium, and thermodynamic mod-
eling incorporating the results of eld studies and laboratory simulations should be included. All of these
efforts would serve to unraveling the mechanisms responsible for magnetic changes in a fault zone, and con-
tribute to the development and renement of rock magnetic strain indicators,geothermometers, and
uid tracersin fault zones. This will ultimately yield a better tool for illuminating faulting processes.
Appendix A: Glossary
A.1 Part I: Rock Magnetic Terms
Anisotropy of (loweld) magnetic susceptibility (AMS): The loweld susceptibility varies slightly as a
function of orientation of the sample with respect to the applied eld (with a strength of up to several
times the Earth's magnetic eld). This variation is described by a symmetric tensor, the AMS tensor.
AMS reects the statistical alignment of crystallographic directions and/or shape in platy or elongate
grains. Next to AMS also the anisotropy of various remanences imparted in the laboratory is occasionally
evaluated, mostly the AARM.
Antiferromagnetic: A state of magnetic order with two ferromagnetic sublattices of equal size aligned anti-
parallel so that (ideally) no net magnetic moment results. In practice, a small magnetic moment persists
because of small deviations from the perfect antiparallel state or defects in the crystal structure. Example:
hematite.
Chemical remanent magnetization (CRM): The NRM imparted to magnetic minerals by chemical pro-
cesses, at any temperature below their Curie temperatures, in the presence of an ambient geomagnetic
eld. Chemical processes involve grain growth beyond the SP grain size or thermochemical alteration
of existing magnetic minerals to other magnetic minerals.
Coercive force (B
c
) and coercivity of remanence (B
cr
): see magnetic hysteresis loop and parameters.
Curie temperature (T
C
): As the temperature increases in a ferromagnetic (sensu lato) material, interatomic
distances increase, and the magnetic exchange interaction that describes the collective ordering of atomic
magnetic moments becomes weaker. At T
C
, the thermal energy overcomes the exchange energy, and indi-
vidual atomic magnetic moments become independent (they lose their longrange ordering) so that the
material becomes paramagnetic. Named after the French physicist Pierre Curie (18591906).
Diamagnetic: A material is diamagnetic when all electron spins are paired. When exposed to a magnetic
eld, a diamagnetic material tends to reduce that magnetic eld: Its susceptibility is negative (the induced
magnetization is opposite to the applied eld). The value is very small and independent of temperature.
Examples: quartz (SiO
2
) and calcite (CaCO
3
).
Ferrimagnetic: A state of magnetic order in which two ferromagnetic sublattices of unequal magnitude are
aligned antiparallel to each other. Example: magnetite.
Ferromagnetic: A state of magnetic order in which all atomic magnetic moments are aligned parallel to
each other. Example: metallic iron.
Firstorder reversal curve (FORC) diagrams: FORCs are partial hysteresis loops determined with the fol-
lowing procedure: (1) a positive saturation eld is applied, (2) the eld is ramped down to a lower reversal
eld, and (3) the magnetization of the sample is measured, while the applied eld is ramped upward to
positive saturation again. A series of FORCs is measured for a set of reversal elds that are increasingly
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 42 of 60
away from positive saturation, and a FORC distribution is calculated as the mixed second derivative of
magnetization with respect to eld spacing; it is displayed in a socalled FORC diagram. FORC diagrams
provide information about the distribution of switching elds (i.e., coercivities) and interaction elds for
all magnetic particles that contribute to a hysteresis loop. They are widely used to characterize samples in
rock and mineral magnetism.
Frequencydependent magnetic susceptibility: The difference between susceptibility measurements made
at two frequencies (e.g., 0.47 and 4.7 kHz), expressed either as a percentage of the low frequency or as
an absolute susceptibility value (i.e., in m
3
kg
1
for massspecic susceptibility or in dimensionless
units for volumespecic susceptibility). This measurement denotes the presence of magnetically vis-
cousferrimagnetic (magnetite or maghemite) grains lying at the stable single domain (SSD)/SP
boundary.
Hysteresis parameters: see magnetic hysteresis loop and parameters.
Inphase and outofphase magnetic susceptibility: When measuring susceptibility in low alternating mag-
netic elds, the measured specimen is magnetized by a weak eld sinusoidally varying in time: H
(t)=H
0
cos(ωt), where H
0
is the eld amplitude, ωthe angular frequency, and tthe time. The magnetic
response is measured, most conveniently represented by magnetization M(t). In diamagnetic, paramag-
netic, and many ferromagnetic (sensu lato) materials, the magnetization also varies sinusoidally being
in phase with the applied eld. The magnetization to eld ratio then maintains a constant value that
represents the susceptibility. In some ferromagnetic materials, however, the magnetization is not in phase
with the applied eld but lags behind the eld usually because of electrical conduction. In that case M
and Mare the inphase and outofphase magnetization components, respectively. Then, the inphase
susceptibility is χ=M/H
0
and the outofphase susceptibility is χ=M/H
0
. See Hrouda et al. (2016)
for more details.
Magnetic domain: A portion of a ferromagnetic/ferrimagnetic material in which the atomic magnetic
moments are aligned. In a magnetic material, magnetic domains serve to reduce the magnetostatic energy
due to the aligned spontaneous magnetization. Fine magnetic grains are SD, and larger grains contain
several magnetic domains. The magnetic domains form a structure in which the magnetic moments of
individual magnetic domains cancel each other out as much as possible. As a result, the spontaneous mag-
netization of the overall magnetic body is reduced, leading to a lower magnetostatic energy state which
represents a most stable energy conguration.
Magnetic susceptibility: A measure of the ease with which a substance can be magnetized in small, Earth
like, applied magnetic elds. The magnetic moment divided by the eld strength in units of H(i.e., A m
1
)
is the susceptibility. Volume susceptibility (κ) is the induced magnetization per unit volume divided by the
eld strength, and is a dimensionless quantity. Massspecic susceptibility (χ) is the magnetic moment
expressed per unit of mass divided by the eld strength; it has units of m
3
kg
1
, that is the inverse of
the specic density of a material. Magnetic susceptibility depends principally on the type and concentra-
tion of magnetic minerals in a sample; grainsize dependence is marginal (with the exception of superpar-
amagnetism that occurs below ~25 nm at room temperature).
Magnetic remanence: Magnetization that persists in the absence of an applied magnetic eld. It differs
from induced magnetization that disappears on removal of an applied eld.
Magnetic hysteresis loop and parameters: A hysteresis loop results when the ineld magnetization of a
sample is measured as the inducing eld is cycled between high positive and negative values. A hysteresis
loop can be measured at any temperature; most hysteresis loops are acquired at room temperature. The
shape of the hysteresis loop yields pertinent information about the magnetic mineral(s) present in a sam-
ple and most importantly on their grain size. The maximuminduced magnetization that can be induced
in a sample is termed the saturation magnetization (M
s
). The corresponding remanence, when the applied
eld is reduced to 0, is the saturation remanence (M
rs
), and the reversed magnetic eld required to reduce
the magnetization to zero is the coercive force (B
c
). The coercivity of remanence (B
cr
): The oppositely
directed magnetic eld required to demagnetize the saturation remanent magnetization to zero. The
M
rs
/M
s
ratio is often plot against the B
cr
/B
c
ratio, named the Day plot (after Ron Day) which features mag-
netic grain size regions. Without additional information on the magnetic mineralogy and their oxidation
state interpretations in terms of grain size should be exercised with caution.
Magnetic saturation: The maximum magnetic moment a sample can achieve. Increasing the eld strength
even further does not impart a higher magnetic moment.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 43 of 60
Magnetic vortex state: The transitional state between SD and MD particles, more traditionally termed PSD.
When a SD particle becomes larger the magnetic moments tend to ower outin the corners of that par-
ticle. This state is termed ower state. In slightly larger particles the atomic spin magnetic moments curl
within the particle around one or more cores. This state is called vortex state. An ideal vortex has a low
magnetostatic energy because the particle's residual magnetic moment is low. Vortex cores may evolve
into classic magnetic domains in yet slightly larger grains.
Magnetostatic interactions: When magnetic particles are located sufciently close to each other, their mag-
netic elds will interact. The interaction eld depends on the conguration of neighboring particles and
will be stronger if more adjacent particles are magnetized in the same direction. Magnetostatic interac-
tions are an important factor to explain the magnetic properties of closely spaced aggregates or intergrown
magnetic particles in earth materials.
Magnetostriction: a property of ferromagnetic (sensu lato) materials which causes them to expand or con-
tract in response to an external magnetic eld. It is considered to be a coupling effect between the mag-
netic energy and the mechanical energy observed in ferromagnetic materials under the inuence of the
external magnetic eld.
Multidomain (MD): With increasing grain size, the magnetostatic energy of a magnetic particle increases.
To minimize this energy, a particle will begin to nucleate domain walls at a critical grain size threshold.
These walls divide the particle into two or more magnetic volumes or domains. The magnetization is uni-
form in each domain, but it differs in direction from domain to domain. The transition between small
grains with only one domain (termed SD) and MD grains is not sharp; a size range exists with a noncol-
linear spin structure that may contain a few domains.
Natural remanent magnetization (NRM): The remanent magnetization that has been acquired naturally
(i.e., not articially acquired in a laboratory). It is the remanent magnetization of a rock that is present
before any laboratory experiments are carried out. This preserves a record of the Earth's magnetic eld
at the time the mineral was laid down as sediment or crystallized in magma. In many cases, that original
NRM componenttermed primary NRMcan be retrieved from the composite NRM with paleomag-
netic techniques. Thus, with paleomagnetic data the original position of rock units can be reconstructed,
so their plate tectonic movement can be quantied.
Néel temperature (T
N
): In antiferromagnetic substances, the individual magnetic moments are aligned
antiparallel to one another below a certain critical temperature, which is known as Néel temperature
(T
N
). In practice, Curie and Néel temperature are being used interchangeably. Named after the French
physicist Louis Néel (19042000).
Paramagnetic: The magnetic state in materials with uncompensated electron spins that do not behave
collectively, that is each free electron spin is an individual atomic magnet. In an applied magnetic
eld these moments will partially align to that applied eld: A paramagnetic material will be magne-
tized in the direction of its inducing magnetic eld. When the eld is removed, the induced magnetic
moment disappears instantaneously. Paramagnetism is proportional to the inverse of the absolute
temperature.
Pseudo single domain (PSD): A magnetic structure that is intermediate between the singledomain (SD)
and MD states. A PSD particle contains more than one domain (up to 57) but exhibits many of the
stable magnetic properties typical of SD particles. See also magnetic vortex state.
Saturation magnetization (M
s
), saturation remanent magnetization (M
rs
): see hysteresis loops.
Single domain (SD): A uniformly magnetized magnetic particle with a single magnetic domain. In most
ferromagnetic and ferrimagnetic minerals, SSD grains are extremely small (the SD size range in equant
magnetite is 3080 nm at room temperature).
Superparamagnetism (SP): Class of magnetic behavior exhibited by very small particles (< ~2530 nm in
magnetite at room temperature) with relaxation times (the magnetic decay time equivalent to the halflife
of radioactive elements) smaller than laboratory timescales (typically taken at 100 s). For these particles,
atomic magnetic moments align in low, Earthlike, applied magnetic elds to produce a strong induced
magnetization that is rapidly attenuated by thermal vibration after removal of the eld (seconds to
minutes).
Thermal remanent magnetization (TRM), also thermoremanent magnetization: The remanent magnetiza-
tion that is acquired by a sample when it is cooled in the ambient geomagnetic eld, from above the Curie
temperature of the constituent magnetic minerals.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 44 of 60
A.2 Part II: Fault Rock and FaultingRelated Terms
Borehole breakout: Drillinginduced compressive failures of the borehole wall occur once the stress
exceeds the strength of the rock or sediment. Their geometry is believed to replicate reliably the orienta-
tion and magnitude of in situ principal stress. Borehole breakouts are deemed to be key indicators of the
stress state in the subsurface. See Zoback et al. (2003) for more details.
Brittle deformation: Deformation involving a throughgoing discontinuityin the rock. It includes tensile
cracking, shear fracturing, and frictional sliding. Brittle deformation occurs when the stresses on a rock
exceed the failure strength of the rock, causing a loss of cohesion.
Cataclasite:Anegrained, cohesive fault rock that generally forms at shallow depths in the crust (typi-
cally from ~5 to ~1015 km), dominantly by brittle deformation processes such as microcracking and
abrasion. It may be subdivided according to the relative proportion of nergrained matrix into protoca-
taclasite (with 1050% matrix), mesocataclasite (with 5090% matrix) and ultracataclasite (90100%
matrix).
Cataclastic ow :Distributed brittle deformation associated with motion of rock particles on a large collec-
tion of cracks and/or frictional surfaces.
Clast: Fragments of rock surrounded by a nergrained matrix.
Clastic dike: A seam of sedimentary material that lls in an open fracture and cuts across sedimentary
rock strata or layering in other rock types. Clastic dikes form rapidly by uidized injection (mobilization
of pressurized pore uids) or passively by water, wind, and gravity (sediment swept into open cracks).
Damage zone: The region encompassing a mappable network of subsidiary brittle deformation structures
surrounding the fault core. These structures may include small faults, veins, fractures, cleavage, and folds
(Caine et al., 1996).
Ductile deformation:Continuous and homogeneously distributeddeformation. It can involve numerous
mechanisms ranging from macroscopic cataclastic ow to crystalplastic processes (Snoke et al., 1998).
Ductile deformation occurs when the stresses on a rock cause permanent strain, without the loss of
cohesion.
Earthquake nucleation: The precursor stage preceding the dynamic rupture propagation of an earthquake
(with negligible inertial effects), which involves quasistatic, aseismic slippage of a fault surface that has a
maximum scale termed nucleation length.
Earthquake energy dissipation: During an earthquake, the elastic strain energy accumulated during the
interseismic period is released by rapid fault frictional sliding. The total energy released (E
tot
)is
partitioned into breakdown work (W
b
), frictional heat (E
H
), and radiated energy (E
R
), that is,
E
tot
=W
b
+E
H
+E
R
. Note that W
b
includes the fracture (surface) energy G, which is the energy required
to cause fracture near the tip of a fault during rupture, and the remainder is also transferred to frictional
heat. E
H
is the energy dissipated by frictional heating during slippage and is often considered to be the
largest component (~8090%) of the total energy budget. Increasing temperature during coseismic slip
can trigger various chemical reactions and phase transitions. These processes are mostly endothermic,
but the takenup energy (often referred to as the EC) is usually minor (<1%) compared with frictional heat
generated.
Fracture: A general term for any break in a rock mass, whether or not it causes displacement, including
cracks, joints and faults.
Fault breccia: A coarsegrained incohesive fault rock, formed by brittle deformation processes, consisting
of at least 30% visible fragments (>0.1 mm).
Fault core: A zone (often asymmetric) enclosed by damage zones that results from the highly localized
strain and intense shearing that accommodates the majority of the displacement within the fault zone.
It generally consists of a number of recurring slip surfaces and fault rocks such as gouges, cataclasites,
and breccias.
Fault gouge: An incohesive, clayrich neto ultranegrained cataclasite, which may possess a foliation
and contains <30% visible fragments. Lithic clasts may be present.
Matrix: Groundmass of negrained material, often surrounding larger fragments.
Mylonite: A fault rock which is cohesive and characterized by a welldeveloped foliation resulting from
tectonic grainsize reduction derived from crystal plastic processes. It commonly contains rounded por-
phyroclasts and lithic fragments of similar composition to minerals in the matrix. Mylonites may be
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 45 of 60
subdivided according to the relative proportion of negrained matrix into protomylonite, mesomylonite
and ultramylonite, which feature grainsize reduction in respectively <50%, 5090%, and >90% of the rock
volume.
Principal slip zone (PSZ): The fault plane along which the most slip is accommodated during a single
earthquake event. It is located within the fault core, surrounded by the damage zone. Synonyms: principal
slip surface (PSS) and principal slip plane (PSP).
Pseudotachylyte: A cohesive fault rock that occurs as solidied melts produced by frictional heating during
seismic slip. It is typically dark in color and is glassy in appearance. Generally, quick, quench cooling of
the melt along the main fault plane forms a pseudotachylyte generation vein. Sometimes, some of the melt
may intrude minor faults or coseismic fractures into the wall rock; the resulting veins are termed injection
veins.
Velocitystrengthening/weakening: The fault strength in the brittle crust is a function of two basic para-
meters: the effective normal stress and the coefcient of friction. A fault may be strong and slip stably,
or weak and fail in large earthquakes. To explain differences between stable sliding (aseismic) and
unstable stickslip (seismogenic) behavior, frictional velocity dependence is considered the most likely
mechanism. Materials that exhibit velocity strengthening (frictional resistance increases with sliding velo-
city) produce only inherently stable frictional slip, whereas those that exhibit velocityweakening
(frictional resistance decreases with sliding velocity) frictional behavior are capable of hosting unstable
rupture.
Data Availability Statement
Data of Figure 4d can be found at Yang et al. (2018); data of Figures 6a6d are available in Özdemir
et al. (1993, 2008), Dekkers et al. (1989), and Guyodo et al. (2006); and data in Figures 11b and 11c are from
Chou, Song, Aubourg, Song, et al. (2012) and Chou, Song, Aubourg, et al. (2014). The data used to create
Figure 5 can be found in the Magnetics Information Consortium (at https://doi.org/10.7288/V4/MAGIC/
16879).
References
Aben, F. M., Dekkers, M. J., Bakker, R. R., van Hinsbergen, D. J. J., Zachariasse, W. J., Tate, G., et al. (2014). Untangling inconsistent
magnetic polarity records through an integrated rock magnetic analysis: A case study on Neogene sections in East Timor. Geochemistry,
Geophysics, Geosystems,15, 25312554. https://doi.org/10.1002/2014GC005294
Aharonov, E., & Scholz, C. H. (2019). The brittleductile transition predicted by a physicsbased friction law. Journal of Geophysical
Research: Solid Earth,124, 27212737. https://doi.org/10.1029/2018JB016878
Almeida, T. P., Muxworthy, A. R., Kovács, A., Williams, W., Brown, P. D., & DuninBorkowski, R. E. (2016). Direct visualization of the
thermomagnetic behavior of pseudosingledomain magnetite particles. Science Advances,2, e1501801. https://doi.org/10.1126/
sciadv.1501801
Almqvist, B. S. G., Bender, H., Bergman, A., & Ring, U. (2020). Magnetic properties of pseudotachylytes from western Jämtland, central
Swedish Caledonides. Solid Earth,11(3), 807828. https://doi.org/10.5194/se118072020
Almqvist, B. S. G., & Koyi, H. (2018). Bulk strain in orogenic wedges based on insights from magnetic fabrics in sandbox models. Geology,
46(6), 483486. https://doi.org/10.1130/G39998.1
Aretusini, S., Mittempergher, S., Plümper, O., Spagnuolo, E., Gualtieri, A. F., & Di Toro, G. (2017). Production of nanoparticles during
experimental deformation of smectite and implications for seismic slip. Earth and Planetary Science Letters,463, 221231. https://doi.
org/10.1016/j.epsl.2017.01.048
Aretusini, S., Spagnuolo, E., Dalconi, M. C., Di Toro, G., & Rutter, E. H. (2019). Water availability and deformation processes in
smectiterich gouges during seismic slip. Journal of Geophysical Research: Solid Earth,124, 10,85510,876. https://doi.org/10.1029/
2019JB018229
Argus, D. F., Gordon, R. G., & DeMets, C. (2011). Geologically current motion of 56 plates relative to the nonetrotation reference frame.
Geochemistry, Geophysics, Geosystems,12, Q11001. https://doi.org/10.1029/2011GC003751
Banerjee, S. K., Hunt, C. P., & Liu, X. (1993). Separation of local signals from the regional paleomonsoon record of the Chinese Loess
Plateau: A rockmagnetic approach. Geophysical Research Letters,20(9), 843846. https://doi.org/10.1029/93GL00908
Banerjee, S. K., King, J., & Marvin, J. (1981). A rapid method for magnetic granulometry with applications to environmental studies.
Geophysical Research Letters,8(4), 333336. https://doi.org/10.1029/GL008i004p00333
Bense, V. F., Gleeson, T., Loveless, S. E., Bour, O., & Scibek, J. (2013). Fault zone hydrogeology. EarthScience Reviews,127, 171192.
https://doi.org/10.1016/j.earscirev.2013.09.008
Berman, R. G., Ya Aranovich, L., & Pattison, D. R. M. (1995). Reassessment of the garnetclinopyroxene FeMg exchange thermometer: II.
Thermodynamic analysis. Contributions to Mineralogy and Petrology,119(1), 3042. https://doi.org/10.1007/BF00310715
Berndt, T., Muxworthy, A. R., & Fabian, K. (2016). Does size matter? Statistical limits of paleomagnetic eld reconstruction from small rock
specimens. Journal of Geophysical Research: Solid Earth,121,1526. https://doi.org/10.1002/2015JB012441
Besnus, M. J., & Meyer, A. J. (1964). Nouvelles données expérimentales sur le magnétisme de la pyrrhotine naturelle (pp. 507511). Paper
presented at Proceedings of the International Conference on Magnetism, Nottingham, England.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 46 of 60
Acknowledgments
This work was supported by the
National Science Foundation of China
(NSFC) grant numbers 41874105,
41472177, and 41204062 to T. Y.
Y. M. C. was funded by the Shenzhen
Science and Technology Program under
grant KQTD20170810111725321,
Science and Technology Innovation
Committee of Shenzhen Municipality
under grant ZDSYS201802081843490,
and the Southern University of Science
and Technology under grants
K19313901 and Y01316111. This work
was also supported by the National
Science Council of Taiwan (grants NSC
1012116M003005 and NSC
1022116M003003 to E. C. Y. and
Y. M. C.) and National Science
Foundation (Instrumentation and
Facilities grant EAR0521558 and
Geophysics grant EAR0228818 to
E. C. F.). M. J. D. acknowledges support
from Netherlands Science Foundation
(NWO) Deep NL grant 2018.040. We
are grateful to Huan Wang and Tsafrir
Levi for providing the original copies of
Figures 2h and 19, respectively. We
thank Toshiaki Mishima for discussions
on an earlier version of the manuscript.
We also thank Silvia Mittempergher,
Bjarne S. G. Almqvist, and two anon-
ymous reviewers for their insightful
comments and suggestions that helped
to improve the manuscript signi-
cantly. The EditorinChief Fabio
Florindo is acknowledged for the ef-
cient handling of the manuscript.
Bezaeva, N. S., Demory, F., Rochette, P., Sadykov, R. A., Gattacceca, J., Gabriel, T., & Quesnel, Y. (2015). The effect of hydrostatic pressure
up to 1.61 GPa on the Morin transition of hematitebearing rocks: Implications for planetary crustal magnetization. Geophysical
Research Letters,42, 10,18810,196. https://doi.org/10.1002/2015GL066306
Bezaeva, N. S., Gattacceca, J., Rochette, P., Sadykov, R. A., & Trukhin, V. I. (2010). Dema gnetization of terrestrial and extraterrestrial rocks
under hydrostatic pressure up to 1.2 GPa. Physics of the Earth and Planetary Interiors,179(12), 720. https://doi.org/10.1016/j.
pepi.2010.01.004
Bhargava, S. K., Garg, A., & Subasinghe, N. D. (2009). In situ hightemperature phase transformation studies on pyrite. Fuel,88(6), 988993.
https://doi.org/10.1016/j.fuel.2008.12.005
Bilardello, D., & Jackson, M. (2013). What do the Mumpsies do? IRM Quarterly,23(3), 1115.
Billi, A. (2005). Grain size distribution and thickness of breccia and gouge zones from thin (<1 m) strikeslip fault cores in limestone.
Journal of Structural Geology,27(10), 18231837. https://doi.org/10.1016/j.jsg.2005.05.013
Bizzarri, A. (2009). What does control earthquake ruptures and dynamic faulting? A review of different competing mechanisms. Pure and
Applied Geophysics,166(57), 741776. https://doi.org/10.1007/s0002400904941
Bloemendal, J., King, J. W., Hall, F. R., & Doh, S. J. (1992). Rock magnetism of Late Neogene and Pleistocene deepsea sediments:
Relationship to sediment source, diagenetic processes and sediment lithology. Journal of Geophysical Research,97(B4), 43614375.
https://doi.org/10.1029/91JB03068
Böhnel, H. N., Dekkers, M. J., DelgadoArgote, L. A., & Gratton, M. N. (2009). Comparison between the microwave and multispecimen
parallel difference pTRM paleointensity methods. Geophysical Journal International,177(2), 383394. https://doi.org/10.1111/j.1365
246X.2008.04036.x
Borradaile, G. J. (1988). Magnetic susceptibility, petrofabric and strain. Tectonophysics,156(12), 120. https://doi.org/10.1016/00401951
(88)90279X
Borradaile, G. J. (1996). Experimental stress remagnetization of magnetite. Tectonophysics,261(4), 229248. https://doi.org/10.1016/0040
1951(95)001565
Borradaile, G. J., & Henry, B. (1997). Tectonic applications of magnetic susceptibility and its anisotropy. EarthScience Reviews,42(12),
4993. https://doi.org/10.1016/S00128252(96)00044X
Borradaile, G. J., & Jackson, M. (2004). Anisotropy of magnetic susceptibility (AMS): Magnetic petrofabrics of deformed rocks,Special
Publications (Vol. 238, pp. 299360). London: Geological Society.
Borradaile, G. J., & Jackson, M. (2010). Structural geology, petrofabrics and magnetic fabrics (AMS, AARM, AIRM). Journal of Structural
Geology,32(10), 15191551. https://doi.org/10.1016/j.jsg.2009.09.006
Borradaile, G. J., & Lagroix, F. (2000). Thermal enhancement of magnetic fabrics in high grade gneisses. Geophysical Research Letters,
27(16), 24132416. https://doi.org/10.1029/2000GL008522
Boullier, A.M. (2011). Faultzone geology: lessons from drilling through the Nojima and Chelungpu faults. Geological Society, London,
Special Publications,359(1), 1737. https://doi.org/10.1144/SP359.2
Boullier, A.M., Yeh, E.C., Boutareaud, S., Song, S.R., & Tsai, C.H. (2009). Microscale anatomy of the 1999 ChiChi earthquake fault zone .
Geochemistry, Geophysics, Geosystems,10, Q03016. https://doi.org/10.1029/2008GC002252
Boulton, C., Yao, L., Faulkner, D. R., Townend, J., Toy, V. G., Sutherland, R., et al. (2017). Highvelocity frict ional properties of Alpine Fault
rocks: Mechanical data, microstructural analysis, and implications for rupture propagation. Journal of Structural Geology,97,7192.
https://doi.org/10.1016/j.jsg.2017.02.003
Bradbury, K. K., Davis, C. R., Shervais, J. W., Janecke, S. U., & Evans, J. P. (2015). Composition, alteration, and texture of faultrelated rocks
from SAFOD core and surface outcrop analogs: Evidence for deformation processes and uidrock interactions. Pure and Applied
Geophysics,172(5), 10531078. https://doi.org/10.1007/s0002401408966
Brantut, N., Schubnel, A., Rouzaud, J.N., Brunet, F., & Shimamoto, T. (2008). High velocity frictional properties of a claybearing fault
gouge and implications for earthquake mechanics. Journal of Geophysical Research,113, B10401. https://doi.org/10.1029/
2007JB005551
Brodie, K., Fettes, D., & Harte, B. (2007). Structural terms including fault rock terms. In D. Fettes, & J. Desmonds (Eds.), Metamorphic rocks:
A classication and glossary of terms (pp. 2431). Cambridge, UK: Cambridge University Press.
Brodsky, E., Ma, K.F., Mori, J. J., Saffer, D. M., & the participants of the ICDP/SCEC International Workshop (2009). Rapid response fault
drilling: Past, present and future. Scientic Drilling,8,6674. https://doi.org/10.2204/iodp.sd.8.11.2009
Brodsky, E. E., Mori, J. J., Anderson, L., Chester, F. M., Conin, M., Dunham, E. M., et al. (2020). The state of stress on the fault before,
during, and after a major earthquake. Annual Reviews in Earth and Planetary Science,48(1), 4974. https://doi.org/10.1146/annurev
earth053018060507
Burmeister, K. C., Harrison, M. J., Marshak, S., Ferré, E., Bannister, R. A., & Kodama, K. P. (2009). Comparison of Fry strain ellipse and
AMS ellipsoid trends to tectonic fabric trends in very lowstrain sandstone of the Appalachian foldthrust belt. Journal of Structural
Geology,31(9), 10281038. https://doi.org/10.1016/j.jsg.2009.03.010
Cai, Y., Pei, J., Wang, H., Sheng, M., & Si, J. (2019). Paleoearthquakes revealed by rock magnetic evidence from the AnxianGuanxian
Fault, Sichuan Province, China. Tectonophysics,752,6880. https://doi.org/10.1016/j.tecto.2018.12.027
Caine, J. S., Bruhn, R. L., & Forster, C. B. (2010). Internal structure, fault rocks, and inferences regarding deformation, uid ow, and
mineralization in the seismogenic Stillwater normal fault, Dixie Valley, Nevada. Journal of Structural Geology,32(11), 15761589.
Caine, J. S., Evans, J. P., & Forster, C. B. (1996). Fault zone architecture and permeab ility structure. Geology,24(11), 10251028. https://doi.
org/10.1130/00917613(1996)024<1025:FZAAPS>2.3.CO;2
Carpenter, B. M., Kitajima, H., Sutherland, R., Townend, J., Toy, V. G., & Saffer, D. M. (2014). Hydraulic and acoustic properties of the
active Alpine Fault, New Zealand: Laboratory measurements on DFDP1 drill core. Earth and Planetary Science Letters,390,4551.
https://doi.org/10.1016/j.epsl.2013.12.023
Carpenter, B. M., Saffer, D. M., & Marone, C. (2015). Frictional properties of the active San Andreas Fault at SAFOD: Implications for fault
strength and slip behavior. Journal of Geophysical Research: Solid Earth,120, 52735289. https://doi.org/10.1002/2015JB011963
CasasSainz, A. M., GilImaz, A., Simón, J. L., IzquierdoLlavall, E., Aldega, L., RománBerdiel, T., et al. (2018). Strain indicators and
magnetic fabric in intraplate fault zones: Case study of Daroca thrust, Iberian Chain, Spain. Tectonophysics,730,2947. https://doi.org/
10.1016/j.tecto.2018.02.013
Cerchiari, A., Remitti, F., Mittempergher, S., Festa, A., Lugli, F., & Cipriani, A. (2020). Cyclical variations of uid source s and stress state in
a shallow megathrust zone mélange. Journal of the Geological Society,177(3), 647659. https://doi.org/10.1144/jgs2019072
Chang, L., Roberts, A. P., Rowan, C. J., Tang, Y., Pruner, P., Chen, Q., & Horng, C.S. (2009). Lowtemperature magnetic properties of
greigite (Fe
3
S
4
). Geochemistry, Geophysics, Geosystems,10, Q01Y04. https://doi.org/10.1029/2008GC002276
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 47 of 60
Chen, J., Niemeijer, A., Yao, L., & Ma, S. (2017). Water vaporization promotes coseismic uid pressurization and buffers temperature rise.
Geophysical Research Letters,44, 21772185. https://doi.org/10.1002/2016GL071932
Chen, J., Yang, X., Ma, S., & Spiers, C. J. (2013). Mass removal and clay mineral dehydration/rehydration in carbonaterich surface
exposures of the 2008 Wenchuan Earthquake fault: Geochemical evidence and implications for fault zone evolution and coseismic slip.
Journal of Geophysical Research: Solid Earth,118, 474496. https://doi.org/10.1002/jgrb.50089
Chen, J., Yang, X., Ma, S., Yang, T., & Niemeijer, A. (2016). Hydraulic properties of samples retrieved from the Wenchuan earthquake Fault
Scientic Drilling Project Hole1 (WFSD1) and the surface rupture zone: Implications for coseismic slip weakening and fault healing.
Geochemistry, Geophysics, Geosystems,17, 27172744. https://doi.org/10.1002/2016GC006376
Chen, T., Xu, H., Xie, Q., Chen, J., Ji, J., & Lu, H. (2005). Characteristics and genesis of maghemite in Chinese loess and paleosols:
Mechanism for magnetic susceptibility enhancement in paleosols. Earth and Planetary Science Letters,240(34), 790802. https://doi.
org/10.1016/j.epsl.2005.09.026
Chester, F. M., Evans, J. P., & Biegel, R. L. (1993). Internal structure and weakening mechanisms of the San Andreas Fault. Journal of
Geophysical Research,98(B1), 771786. https://doi.org/10.1029/92JB01866
Cho, H., Kim, J.S., Kim, K.K., Kang, M.H., Sohn, Y. K., Lee, Y. S., et al. (2015). Determination of rock cleavages using AMS (anisotrop y of
magnetic susceptibility): A case study on the Geochang granite stone, Korea [in Korean with English abstract]. The Journal of the
Petrological Society of Korea,24(3), 209231. https://doi.org/10.7854/JPSK.2015.24.3.209
Cho, H., Son, M., Sohn, Y. K., & Park, M. E. (2017). Magnetic fabric (anisotropy of magnetic susceptibility) constraints on emplacement
mechanism of clastic dikes. Journal of Geophysical Research: Solid Earth,122, 33063333. https://doi.org/10.1002/2016JB013583
Choi, J.H., Edwards, P., Ko, K., & Kim, Y.S. (2016). Denition and classication of fault damage zones: A review and a new methodo-
logical approach. EarthScience Reviews,152,7087. https://doi.org/10.1016/j.earscirev.2015.11.006
Chou, Y.M., Song, S.R., Aubourg, C., Lee, T.Q., Boullier, A.M., Song, Y.F., et al. (2012). An earthquake slip zone is a magnetic record er.
Geology,40(6), 551554. https://doi.org/10.1130/G32864.1
Chou, Y.M., Song, S.R., Aubourg, C., Lee, T.Q., Song, Y.F., & Yeh, E.C. (2014). Quantitative modeling of the newly formed magnetic
minerals in the fault gouge of 1999 ChiChi earthquake (M
w
7.6), Taiwan. Journal of Geophysical Research: Solid Earth,119, 67716781.
https://doi.org/10.1002/2014JB011098
Chou, Y.M., Song, S.R., Aubourg, C., Song, Y.F., Boullier, A.M., Lee, T.Q., et al. (2012). Pyrite alteration and neoformed magnetic
minerals in the fault zone of the ChiChi earthquake (M
w
7.6, 1999): Evidence for frictional heating and coseismic uids. Geochemistry,
Geophysics, Geosystems,13, Q08002. https://doi.org/10.1029/2012GC004120
Chou, Y.M., Song, S.R., Lee, T.Q., Aubourg, C., & Yeh, E.C. (2014). Magnetic fabric analysis of Chelungpu fault gouge in Taiwan and its
implications (Abstract T11B4552). Paper presented at American Geophysical Union Fall Meeting 2014, San Francisco, USA.
Chou, Y.M., Song, S.R., Tsao, T.M., Lin, C.S., Wang, M.K., Lee, J.J., & Chen, F.J. (2014). Identication and tectonic implications of
nanoparticle quartz (~50 nm) by synchrotron Xray diffraction in the Chelungpu fault gouge Taiwan. Tectonophysics,619620,3643.
https://doi.org/10.1016/j.tecto.2013.07.021
Cifelli, F., Mattei, M., Hirt, A. M., & Günther, A. (2004). The origin of tectonic fabrics in undeformedclays: The early stages of defor-
mation in extensional sedimentary basins. Geophysical Research Letters,31, L09604. https://doi.org/10.1029/2004GL019609
Coffey, G. L., Savage, H. M., Polissar, P. J., Rowe, C. D., & Rabinowitz, H. S. (2019). Hot on the trail: Coseismic heating on a localized
structure along the Muddy Mountain fault, Nevada. Journal of Structural Geology,120,6779. https://doi.org/10.1016/j.jsg.2018.12.012
Collettini, C., Di Stefano, G., Carpenter, B., Scarlato, P., Tesei, T., Mollo, S., et al. (2014). A novel and versatile apparatus for brittle rock
deformation. International Journal of Rock Mechanics and Mining Sciences,66, 114123. https://doi.org/10.1016/j.ijrmms.2013.12.005
Collombat, H., Rochette, P., & Vialon, P. (1995). Magnetic fabric as a strain indicator in unconsolidated sediments of the Chile Triple
Junction area. In S. D. Lewis, J. H. Behrmann, R. J. Musgrave, S. C. Cande (Eds.), Chile Triple Junction, Proceedings of the Ocean Drilling
Program,Scientic Results (Vol. 141, pp. 2949). Ocean Drilling Program: College Station, TX.
Cornell, R. M., & Schwertmann, U. (2003). The iron oxides: Structure, properties, reactions, occurrences and uses. Weinheim, Germany:
WileyVCH.
Cowan, D. S. (1999). Do faults preserve a record of seismic slip? A eld geologist's opinion. Journal of Structural Geology,21(89), 9951001.
https://doi.org/10.1016/S01918141(99)000462
Criado, J. M., Gonzalez, M., & Macias, M. (1988). Inuence of grinding on both the stability and thermal decomposition mechanism of
siderite. Thermochimica Acta,135, 219223. https://doi.org/10.1016/00406031(88)873891
D'Alessio, M. A., Blythe, A. E., & Bürgmann, R. (2003). No frictional heat along the San Gabriel fault, California: Evidence from
ssiontrack thermochronology. Geology,31(6), 541544. https://doi.org/10.1130/0091-7613(2003)031<0541:NFHATS>2.0.CO;2
Davidson, C., Davis, K. J., Bailey, C. M., Tape, C. H., Singleton, J., & Singer, B. (2003). Age, origin, and signicance of brittle faulting and
pseudotachylyte along the coast shear zone, Prince Rupert, British Columbia. Geology,31(1), 4346. https://doi.org/10.1130/0091-7613
(2003)031<0043:AOASOB>2.0.CO;2
Day, R., Fuller, M., & Schmidt, V. A. (1977). Hysteresis properties of titanomagnetites: Grainsize and compositional dependence. Physics of
the Earth and Planetary Interiors,13(4), 260267. https://doi.org/10.1016/00319201(77)90108X
de Wall, H. (2000). The elddependence of AC susceptibility in titanomagnetites: Implications for the anisotropy of magnetic suscept-
ibility. Geophysical Research Letters,27(16), 24092411. https://doi.org/10.1029/2000GL008515
Dekkers, M. J. (1988). Magnetic properties of natural pyrrhotite, Part 1: Behaviour of initial susceptibility and saturationmagnetization
related rock magnetic parameters in a grainsize dependent framework. Physics of the Earth and Planetary Interiors,52(34), 376393.
https://doi.org/10.1016/00319201(88)90129X
Dekkers, M. J. (1990). Magnetic properties of natural goethite. Part III. Magnetic behaviour and properties of minerals originating from
goethite dehydration during thermal demagnetization. Geophysical Journal International,103(1), 233250. https://doi.org/10.1111/
j.1365246X.1990.tb01765.x
Dekkers, M. J. (1997). Environmental magnetism: An introduction. Geologie en Mijnbouw,76(1/2), 163182. https://doi.org/10.1023/
A:1003122305503
Dekkers, M. J., Mattéi, J.L., Fillion, G., & Rochette, P. (1989). Grainsize dependence of the magnetic behavior of pyrrhotite during its
lowtemperature transition at 34 K. Geophysical Research Letters,16(8), 855858. https://doi.org/10.1029/GL016i008p00855
Dekkers, M. J., Passier, H. F., & Schoonen, M. A. A. (2000). Magnetic properties of hydrothermally synthesized greigite (Fe
3
S
4
)II. High
and lowtemperature characteristics. Geophysical Journal International,141(3), 809819. https://doi.org/10.1046/j.1365
246x.2000.00129.x
Derr, J. S. (1973). Earthquake lights: A review of observations and present theories. Bulletin of the Seismological Society of America,63,
21772187.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 48 of 60
Di Toro, G., Han, R., Hirose, T., De Paola, N., Nielsen, S., Mizoguchi, K., et al. (2011). Fault lubrication during earthquakes. Nature,
471(7339), 494498. https://doi.org/10.1038/nature09838
Di Toro, G., Niemeijer, A., Tripoli, A., Nielsen, S., Di Felice, F., Scarlato, P., et al. (2010). From eld geology to earthquake simulation: A
new stateofTheart tool to investigate rock friction during the seismic cycle (SHIVA). Rendiconti Lincei,21(S1), 95114. https:/ /doi.org/
10.1007/s122100100097x
Di Toro, G., Pennacchioni, G., & Nielsen, S. (2009). Pseudotachylytes and earthquake source mechanics. In E. Fukuyama (Ed.), Faultzone
properties and earthquake rupture dynamics (pp. 87134). Amsterdam: Elsevier.
Dubey, A. K. (2014). Anisotropy of magnetic susceptibility. In Understanding an orogenic belt (pp. 1734). Cham, Switzerland: Springer.
Dunlop, D. J. (2002a). Theory and application of the Day plot (M
rs
/M
s
versus H
cr
/H
c
): 1. Theoretical curves and tests using titanomagnetite
data. Journal of Geophysical Research,107(B3), 2056. https://doi.org/10.1029/2001JB000486
Dunlop, D. J. (2002b). Theory and application of the Day plot (M
rs
/M
s
versus H
cr
/H
c
): 2. Application to data for rocks, sediments, and soils.
Journal of Geophysical Research,107(B3), 2057. https://doi.org/10.1029/2001JB000487
Dunlop, D. J., & Özdemir, Ö. (1997). Rock magnetism: Fundamentals and frontiers. Cambridge, UK: Cambridge University Press.
Dyar, M. D., Agresti, D. G., Schaefer, M. W., Grant, C. A., & Sklute, E. C. (2006). Mössbauer spectroscopy of Earth and planetary materials.
Annual Review of Earth and Planetary Sciences,34(1), 83125. https://doi.org/10.1146/annurev.earth.34.031405.125049
Egli, R. (2004a). Characterization of individual rock magnetic components by analysis of remanence curves: 1. Unmixing natural sedi-
ments. Studia Geophysica et Geodaetica,48(2), 391446. https://doi.org/10.1023/B:SGEG.0000020839.45304.6d
Egli, R. (2004b). Characterization of individual rock magnetic components by analysis of remanence curves: 2. Fundamental properties of
coercivity distributions. Physics and Chemistry of the Earth,29, 851867. https://doi.org/10.1016/j.pce.2004.04.001
Egli, R. (2004c). Characterization of individual rock magnetic components by analysis of remanence curves: 3. Bacterial magnetite and
natural processes in lakes. Physics and Chemistry of the Earth,29(1314), 869884. https://doi.org/10.1016/j.pce.2004.03.010
Egli, R. (2013). VARIFORC: An optimized protocol for calculating nonregular rstorder reversal curve (FORC) diagrams. Global and
Planetary Change,110(2013), 302320. https://doi.org/10.1016/j.gloplacha.2013.08.003
Elhanati, D., Levi, T., Marco, S., & Weinberger, R. (2020). Zones of inelastic deformation around surface ruptures detected by magnetic
fabrics. Tectonophysics,788, 228502. https://doi.org/10.1016/j.tecto.2020.228502
Enomoto, Y., Asuke, F., Zheng, Z., & Ishigaki, H. (2001). Hardened foliated fault gouge from the Nojima Fault zone at Hirabayashi:
Evidence for earthquake lightning accompanying the 1995 Kobe earthquake? Island Arc,10(34), 447456. https://doi.org/10.1046/
j.14401738.2001.00343.x
Enomoto, Y., Yamabe, T., & Okumura, N. (2017). Causal mechanisms of seismoEM phenomena during the 19651967 Matsushiro
earthquake swarm. Scientic Reports,7, 44774. https://doi.org/10.1038/srep44774
Enomoto, Y., & Zheng, Z. (1998). Possible evidences of earthquake lightning accompanying the 1995 Kobe earthquake inferred from the
Nojima fault gouge. Geophysical Research Letters,25(14), 27212724. https://doi.org/10.1029/98GL02015
Essene, E. J., & Fisher, D. C. (1986). Lightning strike fusion: Extreme reduction and metalsilicate liquid immiscibility. Science,234(4773),
189193. https://doi.org/10.1126/science.234.4773.189
Evans, J. P., & Chester, F. M. (1995). Fluidrock interaction in faults of the SanAndreas systeminferences from SanGabriel fault rock
geochemistry and microstructures. Journal of Geophysical Research,100(B7), 13,00713,020. https://doi.org/10.1029/94JB02625
Evans, J. P., Prante, M. R., Janecke, S. U., Ault, A. K., & Newell, D. L. (2014). Hot faults: Iridescent slip surfaces with metallic luster
document hightemperature ancient seismicity in the Wasatch fault zone, Utah, USA. Geology,42(7), 623626. https://doi.org/10.1130/
G35617.1
Evans, M. E., & Heller, F. (2003). Environmental magnetism: Principles and applications of enviromagnetics. San Diego, California:
Academic.
Faulkner, D. R., Jackson, C. A. L., Lunn, R. J., Schlische, R. W., Shipton, Z. K., Wibberley, C. A. J., & Withjack, M. O. (2010). A review of
recent developments concerning the structure, mechanics and uid ow properties of fault zones. Journal of Structural Geology,32(11),
15571575. https://doi.org/10.1016/j.jsg.2010.06.009
Ferré, E. C., Chou, Y.M., Kuo, R.L., Yeh, E.C., Leibovitz, N. R., Meado, A. L., et al. (2016). Deciphering viscous ow of frictional melts
with the miniAMS method. Journal of Structural Geology,90,1526. https://doi.org/10.1016/j.jsg.2016.07.002
Ferré, E. C., Gébelin, A., Till, J. L., Sassier, C., & Burmeister, K. (2014). Deformation and magnetic fabrics in ductile shear zones: A review.
Tectonophysics,629, 179188. https://doi.org/10.1016/j.tecto.2014.04.008
Ferré, E. C., Geissman, J. W., Chauvet, A., Vauchez, A., & Zechmeister, M. S. (2015). Focal mechanism of prehistoric earthquakes deduced
from pseudotachylyte fabric. Geology,43(6), 531534. https://doi.org/10.1130/G36587.1
Ferré, E. C., Geissman, J. W., Demory, F., Gattacceca, J., Zechmeister, M. S., & Hill, M. J. (2014). Coseismic magnetization of fault pseu-
dotachylytes: 1. Thermal demagnetization experiments. Journal of Geophysical Research: Solid Earth,119, 61136135. https://doi.org/
10.1002/2014JB011168
Ferré, E. C., Geissman, J. W., & Zechmeister, M. S. (2012). Magnetic properties of fault pseudotachylytes in granites. Journal of Geophysical
Research,117, B01106. https://doi.org/10.1029/2011JB008762
Ferré, E. C., MartínHernández, F., Teyssier, C., & Jackson, M. (2004). Paramagnetic and ferromagnetic anisotropy of magnetic suscept-
ibility in migmatites: Measurements in high and low elds and kinematic implications. Geophysical Journal International,157(3),
11191129. https://doi.org/10.1111/j.1365246X.2004.02294.x
Ferré, E. C., Meado, A. L., Geissman, J. W., Di Toro, G., Spagnuolo, E., Ueda, T., et al. (2017). Earthquakes in the mantle? Insights from
rock magnetism of pseudotachylytes. Journal of Geophysical Research: Solid Earth,122, 87698785. https://doi.org/10.1002/
2017JB014618
Ferré, E. C., Zechmeister, M. S., Geissman, J. W., MathanaSekar an, N., & Kocak, K. (2005). The origin of high magnetic remanence in fault
pseudotachylites: Theoretical considerations and implication for coseismic electrical currents. Tectonophysics,402(14), 125139.
https://doi.org/10.1016/j.tecto.2005.01.008
Fidani, C. (2010). The earthquake lights (EQL) of the 6 April 2009 Aquila earthquake, in Central Italy. Natural Hazards and Earth System
Sciences,10(5), 967978. https://doi.org/10.5194/nhess109672010
Flinn, D. (1965). On the symmetry principle and the deformation ellipsoid. Geological Magazine,102(1), 3645. https://doi.org/10.1017/
S0016756800053851
Fondriest, M., Smith, S. A. F., Candela, T., Nielsen, S. B., Mair, K., & Di Toro, G. (2013). Mirrorlike faults and power dissipation during
earthquakes. Geology,41(11), 11751178. https://doi.org/10.1130/G34641.1
Fossen, H., & Cavalcante, G. C. G. (2017). Shear zonesA review. EarthScience Reviews,174, 434455. https://doi.org/10.1016/j.
earscirev.2017.05.002
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 49 of 60
Fossen, H., & Rotevatn, A. (2016). Fault linkage and relay structures in extensional settingsA review. EarthScience Reviews,154,1428.
https://doi.org/10.1016/j.earscirev.2015.11.014
Franke, C., Frederichs, T., & Dekkers, M. J. (2007). Efciency of heavy liquid separation to concentrate magnetic particles. Geophysical
Journal International,170(3), 10531066. https://doi.org/10.1111/j.1365246X.2007.03489.x
Frederichs, T., von Dobeneck, T., Bleil, U., & Dekkers, M. J. (2003). Towards the identication of siderite, rhodochrosite and vivianite in
sediments by their lowtemperature magnetic properties. Physics and Chemistry of the Earth,28(1619), 669679. https://doi.org/
10.1016/S14747065(03)001219
Freund, F., da Silva, M. A. S., Lau, B. W. S., Takeuchi, A., & Jone, H. H. (2007). Electric currents along earthquake faults and the mag-
netization of pseudotachylyte veins. Tectonophysics,431(14), 131141. https://doi.org/10.1016/j.tecto.2006.05.039
Fu, R. R., Weiss, B. P., Lima, E. A., Kehayias, P., Araujo, J. F. D. F., Glenn, D. R., et al. (2017). Evaluating the paleomagnetic potential
of single zircon crystals using the Bishop Tuff. Earth and Planetary Science Letters,458,113. https://doi.org/10.1016/j.
epsl.2016.09.038
Fukuchi, T. (2003). Strong ferrimagnetic resonance signal and magnetic susceptibility of the Nojima pseudotachylyte in Japan and their
implication for coseismic electromagnetic changes. Journal of Geophysical Research,108(B6), 2312. https://doi.org/10.1029/
2002JB002007
Fukuchi, T., Mizoguchi, K., & Shimamoto, T. (2005). Ferrimagnetic resonance signal produced by frictional heating: A new indicator of
paleoseismicity. Journal of Geophysical Research,110, B12404. https://doi.org/10.1029/2004JB003485
Fukuchi, T., Tanaka, D., Imai, N., Soh, W., & Song, S.R. (2009). Thermal analyses of black zones in the Taiwan Chelungpu fault
drilling project Hole B cores using ESR and VSM (Abstract J172013). Paper presented at Japan Geoscience Union Meeting 2009,
Chiba, Japan.
Fukuchi, T., Yurugi, J., & Imai, N. (2007). ESR detection of seismic frictional heating events in the Nojima fault drill core samples, Japan.
Tectonophysics,443(34), 127138. https://doi.org/10.1016/j.tecto.2007.01.020
Fukuzawa, T., Nakamura, N., Oda, H., Uehara, M., & Nagahama, H. (2017). Generation of billowlike wavy folds by uidization at high
temperature in Nojima fault gouge: Microscopic and rock magnetic perspectives. Earth, Planets and Space,69, 54. https://doi.org/
10.1186/s406230170638y
Fulton, P. M., Brodsky, E. E., Kano, Y., Mori, J., Chester, F., Ishikawa, T., et al. (2013). Low coseismic friction on the TohokuOki fault
determined from temperature measurements. Science,342(6163), 12141217. https://doi.org/10.1126/science.1243641
Fusseis, F., Xiao, X., Schrank, C., & De Carlo, F. (2014). A brief guide to synchrotron radiationbased microtomography in (structural)
geology and rock mechanics. Journal of Structural Geology,65,116. https://doi.org/10.1016/j.jsg.2014.02.005
Gattacceca, J., Lamali, A., Rochette, P., Boustie, M., & Berthe, L. (2007). The effects of explosivedriven shocks on the natural remanent
magnetization and the magnetic properties of rocks. Physics of the Earth and Planetary Interiors,162(12), 8598. https://doi.org/
10.1016/j.pepi.2007.03.006
Gattacceca, J., & Rochette, P. (2004). Toward a robust normalized magnetic paleointensity method applied to meteorites. Earth and
Planetary Science Letters,227(34), 377393. https://doi.org/10.1016/j.epsl.2004.09.013
Gehring, A. U., & Hofmeister, A. M. (1994). The transformation of lepidocrocite during heating: A magnetic and spectroscopic study. Clays
and Clay Minerals,42, 409415. https://doi.org/10.1346/CCMN.1994.0420405
Gendler, T. S., Shcherbakov, V. P., Dekkers, M. J., Gapeev, A. K., Gribov, S. K., & McClelland, E. (2005). The lepidocrocitemaghemite
haematite reaction chainI. Acquisition of chemical remanent magnetization by maghemite, its magnetic properties and thermal
stability. Geophysical Journal International,160(3), 815832. https://doi.org/10.1111/j.1365246X.2005.02550.x
Ghiorso, M. S., & Gualda, G. A. R. (2015). Chemical thermodynamics and the study of magmas. In H. Sigurdsson, B. F. Houghton, S. R.
McNutt, H. Rymer, & J. Stix (Eds.), Encyclopedia of volcanoes (2nd ed., pp. 143161). San Diego, CA: Academic Press.
Gilder, S. A., & Le Goff, M. (2008). Systematic pressure enhancement of titanomagnetite magnetization. Geophysical Research Letters,35,
L10302. https://doi.org/10.1029/2008GL033325
Gilder, S. A., Le Goff, M., & Chervin, J.C. (2006). Static stress demagnetization of single and multidomain magnetite with implications for
meteorite impacts. High Pressure Research,26(4), 539547. https://doi.org/10.1080/08957950601092085
Gilder, S. A., Le Goff, M., Chervin, J.C., & Peyronneau, J. (2004). Magnetic properties of single and multidomain magnetite under pres-
sures from 0 to 6 GPa. Geophysical Research Letters,31, L10612. https://doi.org/10.1029/2004GL019844
Gilder, S. A., Pohl, J., & Eitel, M. (2018). Magnetic signatures of terrestrial meteorite impact craters: A summary. In H. Lühr, J. Wicht, S. A.
Gilder, & M. Holschneider (Eds.), Magnetic elds in the solar system (pp. 357382). Cham, Switzerland: Springer.
Glenn, D. R., Fu, R. R., Kehayias, P., Le Sage, D., Lima, E. A., Weiss, B. P., & Walsworth, R. L. (2017). Micrometerscale magnetic imaging of
geological samples using a quantum diamond microscope. Geochemistry, Geophysics, Geosystems,18, 32543267. https://doi.org/
10.1002/2017GC006946
Goddard, J. V., & Evans, J. P. (1995). Chemical changes and uidrock interaction in faults of crystalline thrust sheets, northwestern
Wyoming, U.S.A. Journal of Structural Geology,17(4), 533547. https://doi.org/10.1016/01918141(94)00068B
Gong, Z., Van Hinsbergen, D. J. J., Vissers, R. L. M., & Dekkers, M. J. (2009). Early Cretaceous synrotational extension in the Organyà
basin: New constraints on the palinspastic position of Iberia during its rotation. Tectonophysics,473(34), 312323. https://doi.org/
10.1016/j.tecto.2009.03.003
Graham, J. W. (1966). Signicance of magnetic anisotropy in Appalachian sedimentary rocks. In J. S. Steinhart, & T. J. Smith (Eds.), The
Earth beneath the continents,Geophysical Monograph Series (Vol. 10, pp. 627648). Washington, DC: American Geophysics Union.
Gratier, J.P., Dysthe, D. K., & Renard, F. (2013). The role of pressure solution creep in the ductility of the Earth's upper crust. Advances in
Geophysics,54,47179. https://doi.org/10.1016/B9780123809407.000020
Gray, M. B., Stamatakos, J. A., Ferrill, D. A., & Evans, M. A. (2005). Faultzone deformation in welded tuffs at Yucca Mountain, Nevada,
USA. Journal of Structural Geology,27(10), 18731891. https://doi.org/10.1016/j.jsg.2005.01.018
Grønlie, A., & Torsvik, T. H. (1989). On the origin and age of hydrothermal thorium enriched breccias in the MøreTrøndelag Fault Zone,
central Norway. Norwegian Journal of Geology,69,119.
Grosz, S., Matthews, A., Ilani, S., Ayalon, A., & Garfunkel, Z. (2006). Iron mineralization and dolomitization in the Paran Fault zone, Israel:
Implications for lowtemperature basinal uid processes near the Dead Sea Transform. Geouids,6(2), 137153. https://doi.org/10.1111/
j.14688123.2006.00139.x
Grütter, P., Mamin, H. J., & Rugar, D. (1995). Magnetic force microscopy (MFM). In R. Wiesendanger & H. J. Güntherodt (Eds.), Scanning
tunneling microscopy II,Surface Sciences (Vol. 28, pp. 151207). Berlin, Heidelberg: Springer.
Gualtieri, A. F., & Venturelli, P. (1999). In situ study of the goethitehematite phase transformation by real time synchrotron powder
diffraction. American Mineralogist,84(56), 895904. https://doi.org/10.2138/am19995624
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 50 of 60
Guichet, X., Jouniaux, L., & Catel, N. (2006). Modication of streaming potential by precipitation of calcite in a sandwater system:
Laboratory measurements in the pH range from 4 to 12. Geophysical Journal International,166(1), 445460. https://doi.org/10.1111/
j.1365246X.2006.02922.x
Guyodo, Y., LaPara, T. M., Anschutz, A. J., Penn, R. L., Banerjee, S. K., Geiss, C. E., & Zanner, W. (2006). Rock magnetic, chemical and
bacterial community analysis of a modern soil from Nebraska. Earth and Planetary Science Letters,251(12), 168178. https://doi.org/
10.1016/j.epsl.2006.09.005
Hailwood, E. A., Maddock, R. H., Fung, T., & Rutter, E. H. (1992). Paleomagnetic analysis of fault gouge and dating fault movement,
Anglesey, North Wales. Journal of the Geological Society,149(2), 273284. https://doi.org/10.1144/gsjgs.149.2.0273
Han, R., Shimamoto, T., Ando, J., & Ree, J. H. (2007). Seismic slip record in carbonatebearing fault zones: An insight from highvelocity
friction experiments on siderite gouge. Geology,35(12), 11311134. https://doi.org/10.1130/G24106A.1
Harrison, R. J., & Feinberg, J. M. (2008). FORCinel: An improved algorithm for calculating rstorder reversal curve distributions
using locally weighted regression smoothing. Geochemistry, Geophysics, Geosystems,9, Q05016. https://doi.org/10.1029/
2008GC001987
Hashimoto, Y., & Kaji, U. (2012). Rockuid interaction along seismogenic faults inferred from clay minerals in Okitsu melange, the
Cretaceous Shimanto Belt, SW Japan. In S. D'Amico (Ed.), Earthquake research and analysisSeismology, seismotectonic and earthquake
geology (pp. 253270). Rijeka, Croatia: InTech.
Hattori, I., & Yamamoto, H. (1999). Rock fragmentation and particle size in crushed zones by faulting. The Journal of Geology,107(2),
209222. https://doi.org/10.1086/314343
Hausegger, S., Kurz, W., Rabitsch, R., Kiechl, E., & Brosch, F.J. (2010). Analysis of the internal structure of a carbonate damage zone:
Implications for the mechanisms of fault breccia formation and uid ow. Journal of Structural Geology,32(9), 13491362. https://doi.
org/10.1016/j.jsg.2009.04.014
Hayman, N. W., Housen, B. A., Cladouhos, T. T., & Livi, K. (2004). Magnetic and clast fabrics as measurements of grainscale processes
within the Death Valley shallow crustal detachment faults. Journal of Geophysical Research,109, B05409. https://doi.org/10.1029/
2003JB002902
Hayman, N. W., & Lavier, L. L. (2014). The geological record of deep episodic tremor and slip. Geology,42(3), 195198. https://doi.org/
10.1130/G34990.1
Henderson, I. H. C., Ganerød, G. V., & Braathen, A. (2010). The relationship between particle character istics and frictional strength in basal
fault breccias: Implications for faultrock evolution and rockslide susceptibility. Tectonophysics,486(14), 132149. https://doi.org/
10.1016/j.tecto.2010.02.002
Heraud, J. A., & Lira, J. A. (2011). Coseismic luminescence in Lima, 150 km from the epicenter of the Pisco, Peru earthquake of 15 August
2007. Natural Hazards and Earth System Sciences,11(4), 10251036. https://doi.org/10.5194/nhess1110252011
Heslop, D. (2009). On the statistical analysis of the rock magnetic Sratio. Geophysical Journal International,178(1), 159161. https://doi.
org/10.1111/j.1365246X.2009.04175.x
Heslop, D., & Dillon, M. (2006). Unmixing magnetic remanence curves without a priori knowledge. Geophysical Journal International,
170(2), 556566. https://doi.org/10.1111/j.1365246X.2007.03432.x
Hickman, S., Sibson, R., & Bruhn, R. (1995). Introduction to special section: Mechanical involvement of uids in faulting. Journal of
Geophysical Research,100(B7), 12,83112,840. https://doi.org/10.1029/95JB01121
Hirono, T., Ikehara, M., Otsuki, K., Mishima, T., Sakaguchi, M., Soh, W., et al. (2006). Evidence of frictional melting within diskshaped
black materials discovered from the Taiwan Chelungpu fault system. Geophysical Research Letters,33, L19311. https://doi.org/10.1029/
2006GL027329
Hirono, T., Ishikawa, T., Masumoto, H., Kameda, J., Yabuta, H., & Mukoyoshi, H. (2014). Reevaluation of frictional heat recorded in the
dark gouge of the shallow part of a megasplay fault at the Nankai Trough. Tectonophysics,626, 157169. https://doi.org/10.1016/j.
tecto.2014.04.020
Hirono, T., Lin, W., Yeh, E. C., Soh, W., Hashimoto, Y., Sone, H., et al. (2006). High magnetic susceptibility of fault gouge within Taiwan
Chelungpu fault: Nondestructive continuous measurements of physical and chemical properties in fault rocks recovered from Hole B,
TCDP. Geophysical Research Letters,33, L15303. https://doi.org/10.1029/2006GL026133
Hirono, T., Maekawa, Y., & Yabuta, H. (2015). Investigation of the records of earthquake slip in carbonaceous materials from the Taiwan
Chelungpu fault by means of infrared and Raman spectroscopies. Geochemistry, Geophysics, Geosystems,16, 12331253.
Hirono, T., Ujiie, K., Ishikawa, T., Mishima, T., Hamada, Y., Tanimizu, M., et al. (2009). Estimation of temperature rise in a shallow
slip zone of the megasplay fault in the Nankai Trough. Tectonophysics,478(34), 215220. https://doi.org/10.1016/j.
tecto.2009.08.001
Hirono, T., Yeh, E.C., Lin, W., Sone, H., Mishima, T., Soh, W., et al. (2007). Nondestructive continuous physical property measurements of
core samples recovered from Hole B, Taiwan Chelungpufault Drilling Project. Journal of Geophysical Research ,112, B07404. https://doi.
org/10.1029/2006JB004738
Hirt, A. M. (2007). Magnetic remanence, anisotropy. In D. Gubbins, & E. HerreroBervera (Eds.), Encyclopedia of geomagnetism and
paleomagnetism (pp. 535540). New York: Springer.
Hirt, A. M., Banin, A., & Gehring, A. U. (1993). Thermal generation of ferromagnetic minerals from ironenriched smectites. Geophysical
Journal International,115(3), 11611168. https://doi.org/10.1111/j.1365246X.1993.tb01518.x
Hounslow, M. W. (1990). Grain fabric measured using magnetic susceptibility anisotropy in deformed sediments of the Barbados accre-
tionary prism: Leg 110. In J. C. Moore et al. (Eds.), Proceedings of the Ocean Drilling Program,Scientic Results (Vol. 110, pp. 257275).
Ocean Drilling Program: College Station, TX.
Housen, B. A., Banerjee, S. K., & Moskowitz, B. M. (1996). Lowtemperature magnetic properties of siderite and magnetite in marine
sediments. Geophysical Research Letters,23(20), 28432846. https://doi.org/10.1029/96GL01197
Housen, B. A., & Kanamatsu, T. (2003). Magnetic fabrics from the Costa Rica margin: Sediment deformation during the initial dewatering
and underplating process. Earth and Planetary Science Letters,206(12), 215228. https://doi.org/10.1016/S0012821X(02)010762
Housen, B. A., Tobin, H. J., Labaume, P., Leitch, E. C., Maltman, A. J., & ODP Leg 156 Shipboard Science Party (1996). Strain decoupling
across the decollement of the Barbados accretionary prism. Geology,24(2), 127130. https://doi.org/10.1130/0091-7613(1996)024
<0127:SDATDO>2.3.CO;2
Hrouda, F. (1982). Magnetic anisotropy of rocks and its application in geology and geophysics. Geophysical Surveys,5(1), 3782. https://doi.
org/10.1007/BF01450244
Hrouda, F. (2011). Anisotropy of magnetic susceptibility in variable lowelds: A review. In E. Petrovský, D. Ivers, T. Harinarayana, E.
HerreroBervera (Eds.), The Earth's Magnetic Interior (pp. 281292). Dordrecht, Netherlands: Springer.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 51 of 60
Hrouda, F., Chadima, M., Ježek, J., & Pokorný, J. (2016). Anisotropy of outofphase magnetic susceptibility of rocks as a tool for direct
determination of magnetic subfabrics of some minerals: An introductory study. Geophysical Journal International,208(1), 385402.
https://doi.org/10.1093/gji/ggw399
Hrouda, F., Chlupáčová, M., & Mrázová, Š. (2006). Loweld variation of magnetic susceptibility as a tool for magnetic minera logy of rocks.
Physics of the Earth and Planetary Interiors,154(34), 323336. https://doi.org/10.1016/j.pepi.2005.09.013
Hrouda, F., & Jelínek, V. (1990). Resolution of ferrimagnetic and paramagnetic anisotropies in rocks, using combined loweld and
higheld measurements. Geophysical Journal International,103(1), 7584. https://doi.org/10.1111/j.1365246X.1990.tb01753.x
Hrouda, F., & Ježek, J. (2014). Frequencydependent AMS of rocks: A tool for the investigation of the fabric of ultrane magnetic particles.
Tectonophysics,629,2738. https://doi.org/10.1016/j.tecto.2014.01.040
Hrouda, F., Müller, P., & Hanák, J. (2003). Repeated progressive heating in susceptibility vs. temperature investigation: A new palaeo-
temperature indicator? Physics and Chemistry of the Earth,28(1619), 653657.
Hu, H.P., Chen, Q., Yin, Z., Zhang, P., Zou, J., & Che, H. (2002). Study on the kinetics of the thermal decomposition of mechanically
activated pyrite. Thermochimica Acta,389(12), 7983. https://doi.org/10.1016/S00406031(01)008504
Hu, P. X., Zhao, X., Roberts, A. P., Heslop, D., & Viscarra Rossel, R. A. (2018). Magnetic domain state diagnosis in soils, loess, and marine
sediments from multiple rstorder reversal curvetype diagrams. Journal of Geophysical Research: Solid Earth,123,9981017. https://
doi.org/10.1002/2017JB015195
Huang, W., DupontNivet, G., Lippert, P. C., Van Hinsbergen, D. J. J., Dekkers, M. J., Guo, Z., et al. (2015). Can a primary remanence be
retrieved from partially remagnetized Eocence volcanic rocks in the Nanmulin Basin (southern Tibet) to date the IndiaAsia collision?
Journal of Geophysical Research: Solid Earth,120,4266. https://doi.org/10.1002/2014JB011599
Humbert, F., Robion, P., Louis, L., Bartier, D., Ledésert, B., & Song, S.R. (2012). Magnetic inference of in situ open microcracks in sand-
stone samples from the Taiwan Chelungpu Fault Drilling Project (TCDP). Journal of Asian Earth Sciences,45,179189. https://doi.org/
10.1016/j.jseaes.2011.10.009
Hunt, C. P., Moskowitz, B. M., & Banerjee, S. K. (1995). Magnetic properties of rocks and minerals. In T. J. Ahrens (Ed.), Rock physics and
phase relations: A handbook of physical constants (pp. 189204). Washington, DC: American Geophysical Union.
Isaacs, A. J., Evans, J. P., Song, S.R., & Kolesar, P. T. (2007). Structural, mineralogical, and geochemical characterization of the
Chelungpu thrust fault, Taiwan. Terrestrial, Atmospheric and Oceanic Sciences,18(2), 183221. https://doi.org/10.3319/
TAO.2007.18.2.183(TCDP)
Ishikawa, T., Tanimizu, M., Nagaishi, K., Matsuoka, J., Tadai, O., Sakaguchi, M., et al. (2008). Coseismic uidrock interactions at high
temperatures in the Chelungpu fault. Nature Geoscience,1(10), 679683. https://doi.org/10.1038/ngeo308
Issachar, R., Levi, T., Lyakhovsky, V., Marco, S., & Weinberger, R. (2016). Improving the method of lowtemperature anisotropy of mag-
netic susceptibility (LTAMS) measurements in air. Geochemistry, Geophysics, Geosystems,17, 29402950. https://doi.org/10.1002/
2016GC006339
Issachar, R., Levi, T., Marco, S., & Weinberger, R. (2018). Separation of diamagnetic and paramagnetic fabrics reveals strain directions in
carbonate rocks. Journal of Geophysical Research: Solid Earth,123, 20352048. https://doi.org/10.1002/2017JB014823
Jackson, M. (1991). Anisotropy of magnetic remanence: A brief review of mineralogical sources, physical origins, and geological
applications, and comparison with susceptibility anisotropy. Pure and Applied Geophysics,136(1), 128. https://doi.org/10.1007/
BF00878885
Jackson, M., Borradaile, G., Hudleston, P., & Banerjee, S. (1993). Experimental deformation of synthetic magnetitebearing calcite sand-
stones: Effects on remanence, bulk magnetic properties, and magnetic anisotropy. Journal of Geophysical Research,98(B1), 383401.
https://doi.org/10.1029/92JB01028
Jackson, M. J., Banerjee, S. K., Marvin, J. A., Lu, R., & Gruber, W. (1991). Detrital remanence, inclination errors, and anhysteretic rema-
nence anisotropy: Quantitative model and experimental results. Geophysical Journal International,104(1), 95103.
Jackson, M., Moskowitz, B., Rosenbaum, J., & Kissel, C. (1998). Fielddependence of AC susceptibility in titanomagnetites. Earth and
Planetary Science Letters,157(34), 129139. https://doi.org/10.1016/S0012821X(98)000326
Jackson, M., & Solheid, P. (2010). On the quantitative analysis and evaluation of magnetic hysteresis data. Geochemistry, Geophysics,
Geosystems,11, Q04Z15. https://doi.org/10.1029/2009GC002932
Jayangondaperumal, R., Dubey, A. K., Senthil Kumar, B., Wesnousky, S. G., & Sangode, S. J. (2010). Magnetic fabrics indicating Late
Quaternary seismicity in the Himalayan foothills. International Journal of Earth Sciences,99(S1), 265278. https://doi.org/10.1007/
s0053100904945
Jelínek, V. (1981). Characterization of the magnetic fabric of rocks. Tectonophysics,79(34), T63T67. https://doi.org/10.1016/00401951
(81)901104
Jiang, Z., Rochette, P., Liu, Q., Gattacceca, J., Yu, Y., Barrón, V., & Torrent, J. (2013). Pressure demagnetization of synthetic Al substituted
hematite and its implications for planetary studies. Physics of the Earth and Planetary Interiors,224,110. https://doi.org/10.1016/j.
pepi.2013.09.005
Jordanova, N., Jordanova, D., KostadinovaAvramova, M., Lesigyarski, D., Nikolov, V., Katsarov, G., & Bacvarov, K. (2018). A mineral
magnetic approach to determine paleoring temperatures in the Neolithic settlement site of MursalevoDeveboaz (SW Bulgaria).
Journal of Geophysical Research: Solid Earth,123, 25222538. https://doi.org/10.1002/2017JB015190
Just, J., Dekkers, M. J., Von Dobeneck, T., Van Hoese l, A., & Bickert, T. (2012). Signatures and signicance of aeolian, uvial, bacterial and
diagenetic magnetic mineral fractions in late quaternary marine sediments off Gambia, NW Africa. Geochemistry, Geophysics,
Geosystems,13, Q0AO02. https://doi.org/10.1029/2012GC004146
Just, J., & Kontny, A. (2012). Thermally induced alterations of minerals during measurements of the temperature dependence of magnetic
susceptibility: A case study from the hydrothermally altered SoultzsousForêts granite, France. International Journal of Earth Sciences,
101(3), 819839. https://doi.org/10.1007/s0053101106689
Just, J., Kontny, A., de Wall, H., Hirt, A. M., & MartínHernández, F. (2004). Development of magn etic fabrics during hydrothermal alteration
in the SoultzsousForets granite from the EPS1 borehole, Upper Rhine Graben,Special Publications (Vol. 238, pp. 509526). London:
Geological Society. https://doi.org/10.1144/GSL.SP.2004.238.01.26
Kamogawa, M., Ofuruton, H., & Ohtsuki, Y.H. (2005). Earthquake light: 1995 Kobe earthquake in Japan. Atmospheric Research,76(14),
438444. https://doi.org/10.1016/j.atmosres.2004.11.018
Kanamori, H., & Heaton, T. H. (2000). Microscopic and macroscopic physic s of earthquakes,Geophysical Monograph (Vol. 120, pp. 147163).
Washington, DC: American Geophysical Union.
Kano, Y., Mori, J., Fujio, R., Ito, H., Yanagidani, T., Nakao, S., & Ma, K.F. (2006). Heat signature on the Chelungpu fault associated with
the 1999 ChiChi, Taiwan earthquake. Geophysical Research Letters,33, L14306. https://doi.org/10.1029/2006GL026733
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 52 of 60
Kapička, A. (1988). Anisotropy of magnetic susceptibility in a weak magnetic eld induced by stress. Physics of the Earth and Planetary
Interiors,51(4), 349354. https://doi.org/10.1016/00319201(88)900751
Kapička, A. (1992). Magnetic susceptibility under hydrostatic pressure of synthetic magnetite samples. Physics of the Earth and Planetary
Interiors,70(34), 248252. https://doi.org/10.1016/00319201(92)90191W
Kapička, A., Hrouda, F., Petrovský, E., & Poláčekc, J. (2006). Effect of plastic deformation in laboratory conditions on magnetic anisotropy
of sedimentary rocks. High Pressure Research,26(4), 549553. https://doi.org/10.1080/08957950601092390
Kars, M., Aubourg, C., & Pozzi, J.P. (2011). Low temperature magnetic behaviour near 35 K in unmetamorphosed claystones. Geophysical
Journal International,186(3), 10291035. https://doi.org/10.1111/j.1365246X.2011.05121.x
Kawamoto, E., & Shimamoto, T. (1997). Mechanical behavior of halite and calcite shear zones from brittle to fullyplastic deformation and
a revised fault model. In Y. Zheng, G. A. Davis, A. Yin (Eds.), Structural geology and geomechanics: Proceedings of the 30th International
Geological Congress (Vol. 14, pp. 89105). Utrecht, Netherlands: VSP International Science Publishers.
Kean, W., Day, R., Fuller, M., & Schmidt, V. (1976). The effect of uniaxial compression on the initial susceptibility of rocks as a function of
grain size and composition of their constituent titanomagnetites. Journal of Geophysical Research,81(5), 861872. https://doi.org/
10.1029/JB081i005p00861
Kelso, P. R., Tikoff, B., Jackson, M., & Sun, W. (2002). A new method for the separation of paramagnetic and ferromagnetic susceptibility
anisotropy using low eld and high eld methods. Geophysical Journal International,151(2), 345359. https://doi.org/10.1046/j.1365
246X.2002.01732.x
Kerrich, R., Allison, I., Barnett, R. L., Moss, S., & Starkey, J. (1980). Microstructural and chemical transformations accompanying defor-
mation of granite in a shear zone at Miéville, Switzerland; with implications for stress corrosion cracking and superplastic ow.
Contributions to Mineralogy and Petrology,73(3), 221242. https://doi.org/10.1007/BF00381442
King, J., Banerjee, S. K., & Marvin, J. (1982). A comparison of different magnetic methods for determining the relative grain size of mag-
netite in natural materials: Some results from lake sediments. Earth and Planetary Science Letters,59(2), 404419. https://doi.org/
10.1016/0012821X(82)90142X
Kirschvink, J. L. (1980). The leastsquares line and plane and the analysis of palaeomagnetic data. Geophysical Journal of the Royal
Astronomical Society,62(3), 699718. https://doi.org/10.1111/j.1365246X.1980.tb02601.x
Kodama, K. P., & Dekkers, M. J. (2004). Magnetic anisotropy as an aid to identifying CRM and DRM in red sedimentary rocks. Studia
Geophysica et Geodaetica,48(4), 747766. https://doi.org/10.1023/B:SGEG.0000045481.47203.33
Kontny, A., De Wall, H., Sharp, T. G., & Posfai, M. (2000). Mineralogy and magnetic behavior of pyrrhotite from a 260°C sect ion at the KTB
drilling site, Germany. American Mineralogist,85(10), 14161427. https://doi.org/10.2138/am20001010
Kopp, R. E., & Humayun, M. (2003). Kinetic model of carbonate dissolution in Martian meteorite ALH84001. Geochimica et Cosmochim ica
Acta,67(17), 32473256. https://doi.org/10.1016/S00167037(02)011146
Koziol, A. M. (2004). Experimental determination of siderite stability and application to Martian Meteorite ALH84001. American
Mineralogist,89(23), 294300. https://doi.org/10.2138/am20042306
Kruiver, P. P., Dekkers, M. J., & Heslop, D. (2001). Quantication of magne tic coercivity components by the analysis of acquisition curves of
isothermal remanent magnetisation. Earth and Planetary Science Letters,189(34), 269276. https://doi.org/10.1016/S0012821X(01)
003673
Kuo, C. L., Huba, J. D., Joyce, G., & Lee, L. C. (2011). Ionosphere plasma bubbles and density variations induced by pre-
earthquake rock currents and associated surface charges. Journal of Geophysical Research,116, A10317. https://doi.org/10.1029/
2011JA016628
Kuo, L.W., Li, H., Smith, S. A. F., Di Toro, G., Suppe, J., Song, S.R., et al. (2014). Gouge graphitization and dynamic fault weakening
during the 2008 Mw 7.9 Wenchuan earthquake. Geology,42(1), 4750. https://doi.org/10.1130/G34862.1
Kuo, L.W., Song, S.R., Huang, L., Yeh, E.C., & Chen, H.F. (2011). Temperature estimates of coseismic heating in clayrich fault gouges,
the Chelungpu fault zones, Taiwan. Tectonophysics,502(34), 315327. https://doi.org/10.1016/j.tecto.2011.02.001
Lakshmi, B. V., Gawali, P. B., Deenadayalan, K., & Ramesh, D. S. (2017). Rock magnetic and anisotropy of magnetic suscep tibility (AMS) of
earthquake affected soft sediments: Examples from Shillong and Latur (Deccan Trap), India. Geophysical Research Abstracts,19,
EGU201711760.
Lanci, L., & Kent, D. V. (2003). Introduction of thermal activation in forward modeling of hysteresis loops for singledomain magnetic
particles and implications for the interpretation of the Day diagram. Journal of Geophysical Research,108(B3), 2142. https://doi.org/
10.1029/2001JB000944
Lascu, I., Harrison, R. J., Li, Y. T., Muraszko, J. R., Channell, J. E. T., Piotrowski, A. M., & Hodell, D. A. (2015). Magnetic unmixing of
rstorder reversal curve diagrams using principal component analysis. Geochemistry, Geophysics, Geosystems,16, 29002915. https://
doi.org/10.1002/2015GC005909
Lavina, B., Dera, P., & Downs, R. T. (2014). Modern Xray diffraction methods in mineralogy and geosciences. Reviews in Mineralogy and
Geochemistry,78(1), 131. https://doi.org/10.2138/rmg.2014.78.1
Leibovitz, N. R. (2016). Magnetic paleointensities in fault pseudotachylytes and implications for earthquake lightnings (Master's thesis).
Carbondale, USA: Southern Illinois University. Retrieved from ProQuest.
Levi, T., Weinberger, R., Aïfa, T., Eyal, Y., & Marco, S. (2006a). Earthquakeinduced clastic dikes detected by anisotropy of magnetic sus-
ceptibility. Geology,34(2), 6972. https://doi.org/10.1130/G22001.1
Levi, T., Weinberger, R., Aïfa, T., Eyal, Y., & Marco, S. (2006b). Injection mechanism of clayrich sediments into dikes during earthquakes.
Geochemistry, Geophysics, Geosystems,7, Q12009. https://doi.org/10.1029/2006GC001410
Levi, T., Weinberger, R., Alsop, G. I., & Marco, S. (2018). Characterizing seismites with anisotropy of magnetic susceptibility. Geology,46(9),
827830. https://doi.org/10.1130/G45120.1
Levi, T., Weinberger, R., & Marco, S. (2014). Magnetic fabrics induced by dynamic faulting reveal damage zone sizes in soft rocks , Dead Sea
basin. Geophysical Journal International,199(2), 12141229. https://doi.org/10.1093/gji/ggu300
Li, H., Xue, L., Brodsky, E. E., Mori, J. J., Fulton, P. M., Wang, H., et al. (2015). Longterm temperature records following the Mw 7.9
Wenchuan (China) earthquake consistent with low friction. Geology,43(2), 163166. https://doi.org/10.1130/G35515.1
Li, W.H., Lee, C.H., Ma, M.H., Huang, P. J., & Wu, S. Y. (2019). Fault Dynamics of the 1999 ChiChi earthquake: Clues from nanometric
geochemical analysis of fault gouges. Scientic Reports,9, 5683. https://doi.org/10.1038/s4159801942028w
Lima, E. A., Bruno, A. C., Carvalho, H. R., & Weiss, B. P. (2014). Scanning magnetic tunnel junction microscope for highresolution
imaging of remanent magnetization elds. Measurement Science and Technology,25(10), 105401. https://doi.org/10.10 88/09570233/25/
10/105401
Lin, A. (2008). Fossil earthquakes: The formation and preservation of pseudotachylytes. Berlin: Springer.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 53 of 60
Lin, A., Maruyama, T., Aaron, S., Michibayashi, K., Camacho, A., & Kano, K. (2005). Propagation of seismic slip from brittle to ductile
crust: Evidence from pseudotachylyte of the Woodroffe thrust, central Australia. Tectonophysics,402(14), 2135. https://doi.org/
10.1016/j.tecto.2004.10.016
Lin, W., Conin, M., Moore, J. C., Chester, F. M., Nakamura, Y., Mori, J. J., et al. (2013). Stress state in the largest displacement area of the
2011 Tohokuoki Earthquake. Science,339(6120), 687690. https://doi.org/10.1126/science.1229379
Liu, D., Li, H., Lee, T.Q., Chou, Y.M., Song, S.R., Sun, Z., et al. (2014). Primary rock magnetism for the Wench uan earthquake fault zone
at Jiulong outcrop, Sichuan Province, China. Tectonophysics,619620,5869. https://doi.org/10.1016/j.tecto.2013.08.028
Liu, D., Li, H., Lee, T.Q., Sun, Z., Liu, J., Han, L., & Chevalier, M.L. (2016). Magnetic mineral characterization close to the
YingxiuBeichuan fault surface rupture zone of the Wenchuan earthquake (Mw 7.9, 2008) and its implication for earthquake slip pro-
cesses. Journal of Asian Earth Sciences,115, 468479. https://doi.org/10.1016/j.jseaes.2015.10.019
Liu, J. Y., Chen, Y. I., Huang, C. H., Ho, Y. Y., & Chen, C. H. (2015). A statistical study of lightning activities and M5.0 earthquakes in
Taiwan during 19932004. Surveys in Geophysics,36(6), 851859. https://doi.org/10.1007/s1071201593422
Liu, Q. S., Deng, C. L., Yu, Y., Torrent, J., Jackson, M. J., Banerjee, S. K., & Zhu, R. X. (2005). Temperature dependence of magnetic sus-
ceptibility in an argon environment: Implications for pedog enesis of Chinese loess/palaeosols. Geophysical Journal International,161(1),
102112. https://doi.org/10.1111/j.1365246X.2005.02564.x
Liu, Q. S., Roberts, A. P., Larrasoaña, J. C., Banerjee, S. K., Guyodo, Y., Tauxe, L., & Oldeld, F. (2012). Environmental magnetism:
Principles and applications. Reviews of Geophysics,50, RG4002. https://doi.org/10.1029/2012RG000393
Liu, Q. S., Yu, Y. J., Torrent, J., Roberts, A. P., Pan, Y. X., & Zhu, R. X. (2006). Characteristic lowtemperature magnetic properties of
aluminous goethite [α(Fe, Al)OOH] explained. Journal of Geophysical Research,111, B12S34. https://doi.org/10.1029/2006JB004560
Lockner, D. A., Johnston, M. J. S., & Byerlee, J. D. (1983). A mechanism to explain the generation of earthquake lights. Nature,302(5903),
2833. https://doi.org/10.1038/302028a0
Losseva, T. V., & Nemchinov, I. V. (2005). Earthquake lights and rupture processes. Natural Hazards and Earth System Sciences,5(5),
649656. https://doi.org/10.5194/nhess56492005
Louis, L., Chen, T.M. N., David, C., Robion, P., Wong, T.F., & Song, S.R. (2008). Anisotropy of magnetic susceptibility an d Pwave velocity
in core samples from the Taiwan ChelungpuFault Drilling Project (TCDP). Journal of Structural Geology,30(8), 948962. https://doi.
org/10.1016/j.jsg.2008.03.006
Louzada, K. L., Stewart, S. T., Weiss, B., Gattacceca, J., & Bezaeva, N. S. (2010). Shock and static pressure demagnetization of pyrrhotite
and implications for the Martian crust. Earth and Planetary Science Letters,290(12), 90101. https://doi.org/10.1016/j.
epsl.2009.12.006
Louzada, K. L., Stewart, S. T., Weiss, B., Gattacceca, J., Lillis, R. J., & Halekas, J. S. (2011). Impact demagnetization of the Martian crust:
Current knowledge and future directions. Earth and Planetary Science Letters,305(34), 257269. https://doi.org/10.1016/j.
epsl.2011.03.013
Lowrie, W. (1990). Identication of ferromagnetic minerals in a rock by coercivity and unblocking temperature properties. Geophysical
Research Letters,17(2), 159162. https://doi.org/10.1029/GL017i002p00159
Ma, K.F. (2009). Earthquake dynamicsFrom macroscopic to microscopic: Examples learnt from the 1999 ChiChi earthq uake in Taiwan.
Physics,38(7), 462470.
Ma, S., Shimamoto, T., Yao, L., Togo, T., & Kitajima, H. (2014). A rotaryshear low to highvelocity friction apparatus in Beijing to study
rock friction at plate to seismic slip rates. Earthquake Science,27(5), 469497. https://doi.org/10.1007/s1158901400975
Maekawa, Y., Hirono, T., Yabuta, H., Mukoyoshi, H., Kitamura, M., Ikehara, M., et al. (2014). Estimation of slip parameters associated with
frictional heating during the 1999 Taiwan ChiChi earthquake by vitrinite reectance geothermometry. Earth, Planets and Space,66(1),
28. https://doi.org/10.1186/188059816628
Maher, B. A., & Thompson, R. (1999). Quaternary climates, Environments and magnetism. Cambridge, UK: Cambridge University Press.
Mair, K., & Abe, S. (2008). 3D numerical simulations of fault gouge evolution during shear: Grain size reduction and strain localization.
Earth and Planetary Science Letters,274(12), 7281. https://doi.org/10.1016/j.epsl.2008.07.010
Marcén, M., RománBerdiel, T., CasasSainz, A. M., Soto, R., OlivaUrcia, B., & Castro, J. (2019). Strain variations in a seismogenic normal
fault (Baza Subbasin, Betic Chain): Insights from magnetic fabrics (AMS). Tectonophysics,765,6482. https://doi.org/10.1016/j.
tecto.2019.05.014
MartínHernández, F., & Ferré, E. C. (2007). Separation of paramagnetic and ferrimagnetic anisotropies: A review. Journal of Geophysical
Research,112, B03105. https://doi.org/10.1029/2006JB004340
MartínHernández, F., & Hirt, A. M. (2001). Separation of ferrimagnetic and paramagnetic anisotropies using a higheld torsion mag-
netometer. Tectonphysics,337(34), 209221. https://doi.org/10.1016/S00401951(01)001160
MartínHernández, F., Lüneburg, C. M., Aubourg, C., & Jackson, M. (2004). Magnetic fabric: Methods and applicationsAn introduction,
Special Publications (Vol. 238, pp. 17). London: Geological Society.
Matsuda, T., Omura, K., Ikeda, R., Arai, T., Kobayashi, K., Shimada, K., et al. (2004). Fracturezone conditions on a recently active fault:
Insights from mineralogical and geochemical analyses of the Hirabayashi NIED drill core on the Nojima fault, southwest Japan, which
ruptured in the 1995 Kobe earthquake. Tectonophysics,378(34), 143163. https://doi.org/10.1016/j.tecto.2003.09.005
Maxbauer, D. P., Feinberg, J. M., & Fox, D. L. (2016). MAX UnMix: A web application for unmixing magnetic coercivity distributions.
Computers & Geosciences,95, 140145. https://doi.org/10.1016/j.cageo.2016.07.009
McKenzie, D. P., & Brune, J. N. (1972). Melting on fault planes during large earthquakes. Geophysical Journal of the Royal Astronomical
Society,29(1), 6578. https://doi.org/10.1111/j.1365246X.1972.tb06152.x
McLaren, A. C. (2005). Transmission electron microscopy of minerals and rocks. Cambridge, UK: Cambridge University Press.
Menegon, L., Pennacchioni, G., Malaspina, N., Harris, K., & Wood, E. (2017). Earthquakes as precursors of ductile shear zones in the dry
and strong lower crust. Geochemistry, Geophysics, Geosystems,18,43564374. https://doi.org/10.1002/2017GC007189
Miller, S. A. (2013). The role of uids in tectonic and earthquake processes. Advances in Geophysics,54,146. https://doi.org/10.1016/B978
0123809407.000019
Mishima, T., Hirono, T., Nakamura, N., Tanikawa, W., Soh, W., & Song, S.R. (2009). Changes to magnetic minerals caused by
frictional heating during the 1999 Taiwan ChiChi earthquake. Earth, Planets and Space,61(6), 797801. https://doi.org/10.1186/
BF03353185
Mishima, T., Hirono, T., Soh, W., & Song, S.R. (2006). Thermal history estimation of the Taiwan Chelungpu fault using rockmagnetic
methods. Geophysical Research Letters,33, L23311. https://doi.org/10.1029/2006GL028088
Moecher, D. P., & Steltenpohl, M. G. (2009). Direct calculation of rupture depth for an exhumed paleoseismogenic fault from mylonitic
pseudotachylyte. Geology,37(11), 9991002. https://doi.org/10.1130/G30166A.1
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 54 of 60
MolinaGarza, R. S., Geissman, J., Wawrzyniec, T., Weber, B., Martínez, M. L., & ArandaGómez, J. (2009). An integrated magnetic and
geological study of cataclasitedominated pseudotachylytes in the Chiapas Massif, Mexico: A snapshot of stress orientation following
slip. Geophysical Journal International,177(3), 891912. https://doi.org/10.1111/j.1365246X.2009.04046.x
Morgan, J. K. (1999). Numerical simulations of granular shear zones using the distinct element method: 2. Effects of particle size distri-
bution and interparticle friction on mechanical behavior. Journal of Geophysical Research,104(B2), 27212732. https://doi.org/10.1029/
1998JB900055
Morgan, J. K., & Boettcher, M. S. (1999). Numerical simulations of granular shear zones using the distinct element method: 1. Shear zone
kinematics and the micromechanics of localization. Journal of Geophysical Research,104(B2), 27032719. https://doi.org/10.1029/
1998JB900056
Morin, F. J. (1950). Magnetic susceptibility of αFe
2
O
3
and αFe
2
O
3
with added titanium. Physical Review,78(6), 819820. https://doi.org/
10.1103/PhysRev.78.819.2
Moskowitz, B. M., Jackson, M., & Kissel, C. (1998). Low temperature magnetic behavior of titanomagnetites. Earth and Planetary Science
Letters,157(34), 141149. https://doi.org/10.1016/S0012821X(98)000338
MuirWood, R. (1994). Earthquakes, straincycling and the mobilization of uids,Special Publications (Vol. 78, pp. 8598). London:
Geological Society.
Mukhopadhyay, P. K. (1992). Maturation of organic matter as revealed by microscopicmethods: Applications and limitations of vitrinite
reectance, and continuous spectral and pulsed laser uorescence spectroscopy. In K. H. Wolf & G. V. Chilingarian (Eds.), Diagenesis,
III: Developments in sedimentology (Vol. 47, pp. 435510). Amsterdam: Elsevier.
Mullender, T. A. T., van Velzen, A. J., & Dekkers, M. J. (1993 ). Continuous drift correction and separate identication of ferrimagnetic and
paramagnetic contribution in thermomagnetic runs. Geophysical Journal International,114(3), 663672. https://doi.org/10.1111/j.1365
246X.1993.tb06995.x
Murad, E., & Bowen, L. H. (1987). Magnetic ordering in Alrich goethites: Inuence of crystallinity. American Mineralogist,72, 194200.
Nagata, T. (1961). Rock magnetism. Tokyo: Maruzen.
Nagata, T. (1966). Main characteristics of piezomagnetization and their qualitative interpretation. Journal of Geomagnetism and
Geoelectricity,18(1), 8197. https://doi.org/10.5636/jgg.18.81
Nagata, T., & Carleton, B. J. (1969). Notes on piezoremanent magnetization of igneous rocks II. Journal of Geomagnetism and
Geoelectricity,21(1), 427445. https://doi.org/10.5636/jgg.21.427
Nakamura, N., Hirose, T., & Borradaile, G. J. (2002). Laboratory verication of submicron magnetite production in pseudotachylytes:
Relevance for paleointensity studies. Earth and Planetary Science Letters,201(1), 1318. https://doi.org/10.1016/S0012821X(02)007045
Nakamura, N., & Nagahama, H. (2001). Changes in magnetic and fractal properties of fractured granites near the Nojima fault, Japan.
Island Arc,10(34), 486494. https://doi.org/10.1046/j.14401738.2001.00347.x
Nielsen, S. (2017). From slow to fast faulting: Recent challenges in earthquake fault mechanics. Philosophical Transactions of the Royal
Society A Mathematical Physical and Engineering Sciences,375, 20160016. https://doi.org/10.1098/rsta.2016.0016
Niemeijer, A. R., Di Toro, G., Grifth, W. A., Bistacchi, A., Smith, S. A. F., & Nielsen, S. (2012). Inferring earthquake physics and chemistry
using an integrated eld and laboratory approach. Journal of Structural Geology,39,236. https://doi.org/10.1016/j.jsg.2012.02.018
Niemeijer, A. R., & Vissers, R. L. M. (2014). Earthquake rupture propagation inferred from the spatial distribution of fault rock frictional
properties. Earth and Planetary Science Letters,396, 154164. https://doi.org/10.1016/j.epsl.2014.04.010
Niwa, M., Mizuochi, Y., & Tanase, A. (2015). Changes in chemical composition caused by waterrock interactions across a strikeslip fault
zone: Case study of the Atera Fault, Central Japan. Geouids,15(3), 387409. https://doi.org/10.1111/g.12096
Nixon, C. W., Bull, J. M., & Sanderson, D. J. (2014). Localized vs distributed deformation associated with the linkage history of an active
normal fault, Whakatane Graben, New Zealand. Journal of Structural Geology,69, 266280. https://doi.org/10.1016/j.jsg.2014.06.005
Noguchi, A., Oda, H., Yamamoto, Y., Usui, A., Sato, M., & Kawai, J. (2017). Scanning SQUID microscopy of a ferromanganese crust from
the northwestern Pacic: Submillimeter scale magnetostratigraphy as a new tool for age determination and mapping of environmental
magnetic parameters. Geophysical Research Letters,44, 53605367. https://doi.org/10.1002/2017GL073201
O'Hara, K. D. (2001). A pseudotachylyte geothermometer. Journal of Structural Geology,23(9), 13451357. https://doi.org/10.1016/S0191
8141(01)000086
O'Hara, K. D., & Huggins, F. E. (2005). A Mössbauer study of pseudotachylytes: Redox conditions during seismogenic faulting.
Contributions to Mineralogy and Petrology,148(5), 602614. https://doi.org/10.1007/s004100040622y
Ohmoto, H. (2003). Nonredox transformations of magnetitehematite in hydrothermal systems. Economic Geology,98(1), 157161. https://
doi.org/10.2113/gsecongeo.98.1.157
Oohashi, K., Hirose, T., & Shimamoto, T. (2011). Shearinduced graphitization of carbonaceous materials during seismic fault motion:
Experiments and possible implications fault mechanics. Journal of Structural Geology,33(6), 11221134. https://doi.org/10.1016/j.
jsg.2011.01.007
Oriolo, S., Wemmer, K., Oyhantçabal, P., Fossen, H., Schulz, B., & Siegesmund, S. (2018). Geochronology of shear zonesA review. Earth
Science Reviews,185, 665683. https://doi.org/10.1016/j.earscirev.2018.07.007
Otsuki, K., Hirono, T., Omori, M., Sagaguchi, M., Tanigawa, W., Lin, W., et al. (2009). Analyses of pseudotachylyte from HoleB of Taiwan
Chelungpu Fault Drilling Project (TCDP); Their implications for seismic slip behaviors during the 1999 ChiChi earthquake.
Tectonophysics,469(14), 1324. https://doi.org/10.1016/j.tecto.2009.01.008
Owens, W. H. (1993). Magnetic fabric studies of samples from Hole 808C, Nankai Trough. In I. A. Hill, A. Taira, J. V. Firth (Eds.),
Proceedings of the Ocean Drilling Program,Scientic Results (Vol. 131, pp. 301310). Ocean Drilling Program: College Station, TX.
Özdemir, Ö., & Banerjee, S. K. (1984). High temperature stability of maghemite (γFe
2
O
3
). Geophysical Research Letters,11(3), 161164.
https://doi.org/10.1029/GL011i003p00161
Özdemir, Ö., & Dunlop, D. J. (1997). Effect of crystal defects and internal stress on the domain structure and magnetic properties of
magnetite. Journal of Geophysical Research,102(B9), 20,21120,224. https://doi.org/10.1029/97JB01779
Özdemir, Ö., & Dunlop, D. J. (2000). Intermediate magnetite formation during dehydration of goethite . Earth and Planetary Science Letters,
177(12), 5967. https://doi.org/10.1016/S0012821X(00)000327
Özdemir, Ö., Dunlop, D. J., & Berquó, T. S. (2008). Morin transition in hematite: Size dependence and thermal hysteresis. Geochemistry,
Geophysics, Geosystems,9, Q10Z01. https://doi.org/10.1029/2008GC002110
Özdemir, Ö., Dunlop, D. J., & Moskowitz, B. M. (1993). The effect of oxidation on the Verwey transition in magnetite. Geophysi cal Research
Letters,20(16), 16711674. https://doi.org/10.1029/93GL01483
Palmer, C. S., Ohly, S. R., Smith, R. W., Neupane, G., McLing, T., & Mattson, E. (2015). Mineral selection for multicomponent equilibrium
geothermometry. TransactionsGeothermal Resources Council,38, 453459.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 55 of 60
Pan, Y., & Nilges, M. J. (2014). Electron paramagnetic resonance spectroscopy: Basic principles, experimental techniques and applications
to Earth and planetary sciences. Reviews in Mineralogy and Geochemistry,78(1), 655690. https://doi.org/10.2138/rmg.2014.78.16
Pan, Y. X., Zhu, R. X., Banerjee, S. K., Gill, J., & Williams, Q. (2000). Rock magnetic properties related to thermal treatment of siderite:
Behavior and interpretation. Journal of Geophysical Research,105(B1), 783794. https://doi.org/10.1029/1999JB900358
Parés, J. M. (2015). Sixty years of anisotropy of magnetic susceptibility in deformed sedimentary rocks. Frontiers in Earth Science,3,113.
https://doi.org/10.3389/feart.2015.00004
Parés, J. M., & van der Pluijm, B. A. (2014). Lowtemperature AMS and the quantication of subfabrics in deformed rocks. Tectonophysics,
629,5562. https://doi.org/10.1016/j.tecto.2014.03.005
Pastore, Z., Ter Maat, G., Church, N., McEnroe, S. A., Oda, H., & Fumagalli, P. (2018). Millimeter to micrometer scale mapping of ser-
pentinized dunite by high resolution magnetic microscopy. Lithos,323, 174190. https://doi.org/10.1016/j.lithos.2018.09.018
Pechersky, D. M., & Genshaft, Y. S. (2001). Petromagnetism of the continental lithosphere and the origin of regional magnetic anomal ies: A
review. Russian Journal of Earth Sciences,3(2), 97124. https://doi.org/10.2205/2001ES000059
Pei, J. L., Li, H. B., Wang, H., Si, J. L., Sun, Z. M., & Zhou, Z. Z. (2014). Magnetic properties of the Wenchuan Earthquake Fault Scientic
Drilling Project Hole1 (WFSD1), Sichuan Province, China. Earth, Planets and Space,66(1), 23. https://doi.org/10.1186/188059816623
Pei, J. L., Zhou, Z. Z., Dong, S. G., & Tang, L. (2014). Magnetic evidence revealing frictional heating from fault rocks in granites.
Tectonophysics,637, 207217. https://doi.org/10.1016/j.tecto.2014.10.008
Pei, Y., Paton, D. A., Knipe, R. J., & Wu, K. (2015). A review of fault sealing behaviour and its evaluation in siliciclastic rocks. EarthScience
Reviews,150, 121138. https://doi.org/10.1016/j.earscirev.2015.07.011
Peters, C., & Dekkers, M. J. (2003). Selected room temperature magnetic parameters as a function of mineralogy, concentration and grain
size. Physics and Chemistry of the Earth,28(1619), 659667. https://doi.org/10.1016/S14747065(03)001207
Petri, B., Almqvist, B. S. G., & Pistone, M. (2020). 3D rock fabric analysis using microtomography: An introduction to the opensource
TomoFab MATLAB code. Computers & Geosciences,138, 104444. https://doi.org/10.1016/j.cageo.2020.104444
Petrik, I., Nabelek, P. I., Janak, M., & Plasienka, D. (2003). Conditions of formation and crystallization kinetics of highly oxidized pseu-
dotachylytes from the High Tatras (Slovakia). Journal of Petrology,44(5), 901927. https://doi.org/10.1093/petrology/44.5.901
Petrovský, E., Alcalá, M. D., Criado, J. M., Grygar, T., Kapička, A., & Šubrt, J. (2000). Magnetic properties of magnetite prepared by
ballmilling of hematite with iron. Journal of Magnetism and Magnetic Materials,210(13), 257273. https://doi .org/10.1016/S03048853
(99)006241
Petrovský, E., Kropáček, V., Dekkers, M. J., de Boer, C. B., Hoffman, V., & Ambatiello, A. (1996). Transformation of hema tite to maghemite
as observed by changes in magnetic parameters: Effects of mechanical activation? Geophysical Research Letters,23(12), 14771480.
https://doi.org/10.1029/96GL01411
Philpotts, A. R. (1964). Origin of pseudotachylytes. American Journal of Science,262(8), 10081035. https://doi.org/10.2475/
ajs.262.8.1008
Pittarello, L., Di Toro, G., Bizzarri, A., Pennacchioni, G., Hadizadeh, J., & Cocco, M. (2008). Energy partitioning during seismic slip in
pseudotachylytebearing faults (Gole Larghe Fault, Adamello, Italy). Earth and Planetary Science Letters,269(12), 131139. https://doi.
org/10.1016/j.epsl.2008.01.052
Pittarello, L., Pennacchioni, G., & Di Toro, G. (2012). Amphibolitefacies pseudotachylytes in Premosello metagabbro and felsic mylonites
(Ivrea Zone, Italy). Tectonophysics,580,4357. https://doi.org/10.1016/j.tecto.2012.08.001
Potter, D. K. (2004). A comparison of anisotropy of magnetic remanence methodsA user's guide for application to palaeomagnetism and
magnetic fabric studies,Special Publications (Vol. 238, pp. 2135). London: Geological Society. https://doi.org/10.1144/GSL.
SP.2004.238.01.03
Pozzi, J. P. (1975). Magnetic properties of oceanic basaltsEffects of pressure and consequences for the interpretation of anomalies. Earth
and Planetary Science Letters,26(3), 337344. https://doi.org/10.1016/0012821X(75)900096
Rabinowitz, H. S., Polissar, P. J., & Savage, H. M. (2017). Reaction kinetics of alkenone and nalkane thermal alteration at seismic time-
scales. Geochemistry, Geophysics, Geosystems,18, 204219. https://doi.org/10.1002/2016GC006553
Rabinowitz, H. S., Savage, H. M., Polissar, P. J., Rowe, C. D., & Kirkpatrick, J. D. (2020). Earthquake slip surfaces identied by biomarker
thermal maturity within the 2011 TohokuOki earthquake fault zone. Nature Communications,11, 533. https://doi.or g/10.1038/s41467
020144471
Reed, S. J. B. (2005). Electron microprobe analysis and scanning electron microscopy in geology. Cambridge, UK: Cambridge University Press.
Reverdatto, V. V., Likhanov, I. I., Polyansky, O. P., Sheplev, V. S., & Kolobov, V. Y. (2019). Mineral geothermobarometr y. In The nature and
models of metamorphism (pp. 5582). Cham, Switzerland: Springer.
Reynolds, R. L., Tuttle, M. L., Rice, C. A., Fishman, N. S., Karachewski, J. A., & Sherman, D. W. (1994). Magnetization and geochemistry of
greigitebearing Cretaceous strata, North Slope Basin, Alaska. American Journal of Science,294(4), 485528. https://doi.org/10.2475/
ajs.294.4.485
Rice, J. R. (2006). Heating and weakening of faults during earthquake slip. Journal of Geophysical Research,111, B05311. https://doi.org/
10.1029/2005JB004006
Roberts, A. P. (1995). Magnetic properties of sedimentary greigite (Fe
3
S
4
). Earth and Planetary Science Letters,134(34), 227236. https://
doi.org/10.1016/0012821X(95)00131U
Roberts, A. P., Almeida, T. P., Church, N. S., Harrison, R. J., Heslop, D., Li, Y., et al. (2017). Resolving the origin of pseudosingle domain
magnetic behavior. Journal of Geophysical Research: Solid Earth,122, 95349558. https://doi.org/10.1002/2017JB014860
Roberts, A. P., Chang, L., Rowan, C. J., Horng, C. S., & Florindo, F. (2011). Magnetic properties of sedimentary greigite (Fe
3
S
4
): An update.
Reviews of Geophysics,49, RG1002. https://doi.org/10.1029/2010RG000336
Roberts, A. P., Heslop, D., Zhao, X., & Pike, C. R. (2014). Understanding ne magnetic particle systems through use of rstorder reversal
curve diagrams. Reviews of Geophysics,52, 557602. https://doi.org/10.1002/2014RG000462
Roberts, A. P., Liu, Q. S., Rowan, C. J., Chang, L., Carvallo, C., Torrent, J., et al. (2006). Characterization of hematite (αFe
2
O
3
), goethite
(αFeOOH), greigite (Fe
3
S
4
), and pyrrhotite (Fe
7
S
8
) using rstorder reversal curve diagrams. Journal of Geophysical Research,111,
B12S35. https://doi.org/10.1029/2006JB004715
Roberts, A. P., Pike, C. R., & Verosub, K. L. (2000). Firstorder reversal curve diagrams: A new tool for characterizing the magnetic prop-
erties of natural samples. Journal of Geophysical Research,105(B12), 28,46128,475. https://doi.org/10.1029/2000JB900326
Roberts, A. P., Tauxe, L., Heslop, D., Zhao, X., & Jiang, Z. (2018). A critical appraisal of the Daydiagram. Journal of Geophysical Research:
Solid Earth,123, 26182644. https://doi.org/10.1002/2017JB015247
Rochette, P., Bezaeva, N. S., Kosterov, A., Gattacceca, J., Masaitis, V. L., Badyukov, D. D., et al. (2019). Magnetic properties and redox state
of impact glasses: A review and new case studies from Siberia. Geosciences,9, 225. https://doi.org/10.3390/geosciences9050225
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 56 of 60
Rochette, P., & Fillion, G. (1988). Identication of multicomponent anisotropies in rocks using various eld and temperature values in a
cryogenic magnetometer. Physics of the Earth and Planetary Interiors,51(4), 379386.
Rochette, P., & Fillion, G. (1989). Field and temperature behavior of remanence in synthetic goethite: Paleomagnetic implications.
Geophysical Research Letters,16(8), 851854. https://doi.org/10.1029/GL016i008p00851
Rochette, P., Fillion, G., & Dekkers, M. J. (2011). Interpretation of lowtemperature data part 4: The lowtemperature magnetic transition of
monoclinic pyrrhotite. IRM Quarterly,21,111.
Rochette, P., Fillion, G., Mattéi, J.L., & Dekkers, M. J. (1990). Magnetic transition at 3034 Kelvin in pyrrhotite: Insight into a
widespread occurrence of this mineral in rocks. Earth and Planetary Science Letters,98(34), 319328. https://doi.org/10.1016/0012
821X(90)90034U
Rochette, P., Gattacceca, J., Devouard, B., Moustard, F., Bezaeva, N. S., Cournède, C., & Scaillet, B. (2015). Magnetic properties of tektites
and other related impact glasses. Earth and Planetary Science Letters,432,381390. https://doi.org/10.1016/j.epsl.2015.10.030
Rochette, P., Jackson, M., & Aubourg, C. (1992). Rock magnetism and the interpretation of anisotropy of magnetic susceptibility. Reviews of
Geophysics,30(3), 209226. https://doi.org/10.1029/92RG00733
RománBerdiel, T., CasasSainz, A. M., OlivaUrcia, B., Calvín, P., & Villalaín, J. J. (2019). On the inuence of magnetic mineralogy in the
tectonic interpretation of anisotropy of magnetic susceptibility in cataclastic fault zones. Geophysical Journal International,216(2),
10431061. https://doi.org/10.1093/gji/ggy481
Rowe, C. D., Fagereng, Å., Miller, J. A., & Mapani, B. (2012). Signature of coseismic decarbonation in dolomitic fault rocks of the Naukluft
Thrust, Namibia. Earth and Planetary Science Letters,333334, 200210. https://doi.org/10.1016/j.epsl.2012.04.030
Rowe, C. D., & Grifth, W. A. (2015). Do faults preserve a record of seismic slip: A second opinion. Journal of Structural Geology,78,126.
https://doi.org/10.1016/j.jsg.2015.06.006
Rowe, C. D., Kirkpatrick, J. D., & Brodsky, E. E. (2012). Fault rock injections record paleoearthquakes. Earth and Planetary Science Letters,
335336, 154166. https://doi.org/10.1016/j.epsl.2012.04.015
Rowe, C. D., Lamothe, K., Rempe, M., Andrews, M., Mitchell, T. M., Di Toro, G., et al. (2019). Earthquake lubrication and healing explained
by amorphous nanosilica. Nature Communications,10, 320. https://doi.org/10.1038/s4146701808238y
Sakaguchi, A., Chester, F., Curewitz, D., Fabbri, O., Goldsby, D., Kimura, G., et al. (2011). Seismic slip propagation to the updip end of plate
boundary subduction interface faults: Vitrinite reectance geothermometry on Integrated Ocean Drilling Program NanTro SEIZE cores.
Geology,39(4), 395398. https://doi.org/10.1130/G31642.1
Sammis, C. G., & BenZion, Y. (2008). Mechanics of grainsize reduction in fault zones. Journal of Geophysical Research,113, B02306.
https://doi.org/10.1029/2006JB004892
Sasai, Y. (2001). Tectonomagnetic modeling based on the piezomagnetism: A review. Annals of Geophysics,44(2), 361368.
Sato, M., Makio, M., Hayashi, T., & Ohno, M. (2015). Abrupt intensication of North Atlantic deep water formation at the
Nordic Seas during the late Pliocene climate transition. Geophysical Research Letters,42, 49494955. https://doi.org/10.1002/
2015GL063307
Savage, H. M., Polissar, P. J., Sheppard, R., Rowe, C. D., & Brodsky, E. E. (2014). Biomarkers heat up during earthquakes: New evidence of
seismic slip in the rock record. Geology,42(2), 99102. https://doi.org/10.1130/G34901.1
Savage, H. M., Rabinowitz, H. S., Spagnuolo, E., Aretusini, S., Polissar, P. J., & Di Toro, G. (2018). Biomarker thermal maturity experiment s
at earthquake slip rates. Earth and Planetary Science Letters,502, 253261.
Schabes, M. E., & Bertram, H. N. (1988). Magnetization processes in ferromagnetic cubes. Journal of Applied Physics,64(3), 13471357.
https://doi.org/10.1063/1.341858
Schmid, S. M., & Handy, M. R. (1991). Towards a genetic classication of fault rocks: Geological usage an d tectonophysical implications. In
D. W. Müller, J. A. McKenzie, & H. Weissert (Eds.), Controversies in modern geology: Evolution of geological theories in sedimentology,
Earth history and tectonics (pp. 339361). London: Academic.
Schmidt, V., Hirt, A. M., Rosselli, P., & MartínHernández, F. (2007). Separation of diamagnetic and paramagnetic anisotropy by higheld,
lowtemperature torque measurements. Geophysical Journal International,168(1), 4047. https://doi.org/10.1111/j.1365
246X.2006.03202.x
Scholz, C. H. (2019). The mechanisms of earthquake faulting (3rd ed.). New York: Cambridge University Press. https://doi.org/10.1017/
9781316681473
Schwarz, E. J., & Vaughan, D. J. (1972). Magnetic phase relations of pyrrhotite. Journal of Geomagnetics and Geoelectrics,24(4), 441458.
https://doi.org/10.5636/jgg.24.441
Scuderi, M., & Collettini, C. (2016). The role of uid pressure in induced vs. triggered seismicity: Insights from rock deformation experi-
ments on carbonates. Scientic Reports,6, 24852. https://doi.org/10.1038/srep24852
Sibson, R. H. (1977). Fault rocks and fault mechanisms. Journal of the Geological Society,133(3), 191213. https://doi.org/10.1144/
gsjgs.133.3.0191
Sibson, R. H. (1986). Earthquakes and rock deformation in crustal fault zones. Annual Review of Earth and Planetary Sciences,14(1),
149175. https://doi.org/10.1146/annurev.ea.14.050186.001053
Sibson, R. H. (1989). Earthquake faulting as a structural process. Journal of Structural Geology,11(12), 114. https://doi.org/10.1016/0191
8141(89)900321
Sibson, R. H. (1992). Implication of faultvalve behaviour for rupture nucleation and recurrence. Tectonophysics,211(14), 28 3293. https://
doi.org/10.1016/00401951(92)90065E
Sibson, R. H. (2003). Thickness of the seismic slip zone. Bulletin of the Seismological Society of America,93(3), 11691178. https://doi.org/
10.1785/0120020061
Sibson, R. H., & Toy, V. G. (2006). The habitat of faultgenerated pseudotachylyte: Presence vs. absence of frictionmelt. In R. Abercrombie,
A. McGarr, H. Kanamori, G. Di Toro (Eds.), Earthquakes: Radiated energy and the physics of faulting,Geophysical Monograph Series
(Vol. 170, pp. 153166). Washington, DC: American Geophysical Union.
SimanTov, S., Aharonov, E., Sagy, A., & Emmanuel, S. (2013). Nanograins form carbonate fault mirrors. Geology,41(6), 703706. https://
doi.org/10.1130/G34087.1
Snoke, A. W., Tullis, J., & Todd, V. R. (1998). Faultrelated rocks: A photographic atlas. Princeton, NJ: Princeton University Press.
Solum, J. G., & van der Pluijm, B. (2009). Quantication of fabrics in clay gouge from the Carboneras fault, Spain and implications for fault
behavior. Tectonophysics,475(34), 554562. https://doi.org/10.1016/j.tecto.2009.07.006
Spassov, S., & Hus, J. (2006). Estimating baking temperatures in a Roman pottery kiln by rock magnetic properties: Implications of ther-
mochemical alteration on archaeointensity determinations. Geophysical Journal International,167(2), 592604. https://doi.or g/10.1111/
j.1365246X.2006.03114.x
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 57 of 60
Spray, J. G. (2010). Frictional melting processes in planetary materials: From hypervelocity impact to earthquakes. Annual Review of Earth
and Planetary Sciences,38(1), 221254. https://doi.org/10.1146/annurev.earth.031208.100045
Stacey, F. D., & Banerjee, D. K. (1974). Physical principles of rock magnetism. Amsterdam: Elsevier.
Stephenson, A., Sadikun, S., & Potter, D. K. (1986). A theoretical and experimental comparison of the anisotropies of magnetic suscept-
ibility and remanence in rocks and minerals. Geophysical Journal International,84(1), 185200. https://doi.org/10.1111/j.1365
246X.1986.tb04351.x
StLaurent, F., Derr, J. S., & Freund, F. T. (2006). Earthquake lights and the stressactivation of positive hole charge carriers in rocks. Physics
and Chemistry of the Earth,31(49), 305312. https://doi.org/10.1016/j.pce.2006.02.003
Storti, F., Balsamo, F., & Salvini, F. (2007). Particle shape evolution in natural carbonate granular wear material. Terra Nova,19(5),
344352. https://doi.org/10.1111/j.13653121.2007.00758.x
Sutherland, R., Toy, V. G., Townend, J., Cox, S. C., Eccles, J. D., Faulkner, D. R., et al. (2012). Drilling reveals uid control on architecture
and rupture of the Alpine fault, New Zealand. Geology,40(12), 11431146. https://doi.org/10.1130/G33614.1
Suttie, N., Shaw, J., & Hill, M. J. (2010). Direct demonstration of microwave demagnetization of a whole rock sample with minimal heating.
Earth and Planetary Science Letters,292(34), 357362. https://doi.org/10.1016/j.epsl.2010.02.002
Swan, A. R. H., & Sandilands, M. (1995). Introduction to geological data analysis. Oxford, UK: Blackwell Science.
Swanson, M. T. (1992). Fault structure, wear mechanisms and rupture processes in pseudotachylyte generation. Tectonophysics,204(34),
223242. https://doi.org/10.1016/00401951(92)90309T
Tanaka, H., Chen, W. M., Wang, C. Y., Ma, K. F., Urata, N., Mori, J., & Ando, M. (2006). Frictional heat from faulting of the 1999 ChiChi,
Taiwan earthquake. Geophysical Research Letters,33, L16316. https://doi.org/10.1029/2006GL026673
Tanaka, H., Fujimoto, K., Ohtani, T., & Ito, H. (2001). Structural and chemical characterization of shear zones in the freshly activated
Nojima fault, Awaji Island, southwest Japan. Journal of Geophysical Research,106(B5), 87898810. https://doi.org/10.1029/
2000JB900444
Tanikawa, W., Ishikawa, T., Honda, G., Hirono, T., & Tadai, O. (2015). Trace element anomaly in fault rock induced by coseismic
hydrothermal reactions reproduced in laboratory friction experiments. Geophysical Research Letters,42, 32103217. https://doi.org/
10.1002/2015GL063195
Tanikawa, W., Mishima, T., Hirono, T., Lin, W., Shimamoto, T., Soh, W., & Song, S.R. (2007). High magnetic susceptibility produced in
highvelocity frictional tests on core samples from the Chelungpu fault in Taiwan. Geophysical Research Letters,34, L15304. https://doi.
org/10.1029/2007GL030783
Tanikawa, W., Mishima, T., Hirono, T., Soh, W., & Song, S.R. (2008). High magnetic susceptibility produced by thermal decomposition of
core samples from the Chelungpu fault in Taiwan. Earth and Planetary Science Letters,272(12), 372381. https://doi.org/10.1016/j.
epsl.2008.05.002
Tarduno, J. A., Cottrell, R. D., Davis, W. J., Nimmo, F., & Bono, R. K. (2015). A Hadean to Paleoarchean geodynamo recorded by single
zircon crystals. Science,349(6247), 521524. https://doi.org/10.1126/science.aaa9114
Tarling, D. H., & Hrouda, F. (1993). The magnetic anisotropy of rocks. London, UK: Chapman & Hall.
Tauxe, L. (2010). Essentials of paleomagnetism. Berkeley: University of California Press.
Tauxe, L., Bertram, H. N., & Seberino, C. (2002). Physical interpretation of hysteresis loops: Micromagnetic modeling of ne particle
magnetite. Geochemistry, Geophysics, Geosystems,3(10), 1055. https://doi.org/10.1029/2001GC000241
Terakawa, T., Zoporowski, A., Galvan, B., & Miller, S. A. (2010). Highpressure uid at hypocentral depths in the L'Aquila region inferred
from earthquake focal mechanisms. Geology,38(11), 995998. https://doi.org/10.1130/G31457.1
Thériault, R., StLaurent, F., Freund, F. T., & Derr, J. S. (2014). Prevalence of earthquake lights associated with rift environments.
Seismological Research Letters,85(1), 159178. https://doi.org/10.1785/0220130059
Thompson, R., & Oldeld, F. (1986). Environmental magnetism. London, UK: Allen and Unwin.
Till, J. L., & Moskowitz, B. M. (2014). Deformation microstructures and magnetite texture development in synthetic shear zones.
Tectonophysics,629, 211223. https://doi.org/10.1016/j.tecto.2014.04.026
Torii, M., Fukuma, K., Horng, C.S., & Lee, T.Q. (1996). Magnetic discrimination of pyrrhotiteand greigitebearing sediment samples.
Geophysical Research Letters,23(14), 18131816. https://doi.org/10.1029/96GL01626
Torrent, J., & Barrón, V. (2008). Diffuse reectance spectroscopy. In L. R. Dress, & A. L. Ulery (Eds.), Methods of soil analysis, Part 5.
Mineralogical methods,SSSA Book Serial (Vol. 5, pp. 367387). Wisconsin: Soil Science Society of America.
Torsvik, T. H., Sturt, B. A., Swensson, E., Andersen, T. B., & Dewey, J. F. (1992). Palaeomagnetic datin g of fault rocks: Evidence for Permian
and Mesozoic movements and brittle deformation along the extensional Dalsfjord Fault, western Norway. Geophysical Journal
International,109(3), 565580. https://doi.org/10.1111/j.1365246X.1992.tb00118.x
Toulmin, P., & Barton, P. B. (1964). A thermodynamic study of pyrite and pyrrhotite. Geochimica et Cosmochimica Acta,28(5), 641671.
https://doi.org/10.1016/00167037(64)900833
Townend, J., & Zoback, M. D. (2000). How faulting keeps the crust strong. Geology,28(5), 399402. https://doi.org/10.1130/0091-7613
(2000)28<399:HFKTCS>2.0.CO;2
Twiss, R. J., & Moores, E. M. (1992). Structural geology. New York: W. H. Freeman.
Ujiie, K., Hisamitsu, T., & Taira, A. (2003). Deformation and uid pressure variation during initiation and evolution of the plate boundary
décollement zone in the Nankai accretionary prism. Journal of Geophysical Research,108(B8), 2398. https://doi.org/10.1029/
2002JB002314
Ujiie, K., & Kimura, G. (2014). Earthquake faulting in subduction zones: Insights from fault rocks in accretionar y prisms. Progress in Earth
and Planetary Science,1(1), 730. https://doi.org/10.1186/2197428417
Ujiie, K., Tanaka, H., Saito, T., Tsutsumi, A., Mori, J. J., Kameda, J., et al. (2013). Low coseismic shear stress on the Tohoku oki megathrust
determined from laboratory experiments. Science,342(6163), 12111214. https://doi.org/10.1126/science.1243485
Van de Moortèle, B., Reynard, B., Rochette, P., Jackson, M., Beck, P., Gillet, P., et al. (2007). Shockinduced metallic iron nanopar-
ticles in olivinerich Martian meteorites. Earth and Planetary Science Letters,262(12), 3749. https://doi.org/10.1016/j.
epsl.2007.07.002
Verberne, B. A., Plümper, O., & Spiers, C. J. (2019). Nanocrystalline principal slip zones and their role in controlling crustal fault rheology.
Minerals,9, 328. https://doi.org/10.3390/min9060328
Vermilye, J. M., & Scholz, C. H. (1998). The process zone: A microstructural view of fault growth. Journal of Geophysical Resea rch,103(B6),
12,22312,237. https://doi.org/10.1029/98JB00957
Verosub, K. L., & Roberts, A. P. (1995). Environmental magnetism: Past, present, and future. Journal of Geophysical Research,100(B2),
21752192. https://doi.org/10.1029/94JB02713
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 58 of 60
Verrier, V., & Rochette, P. (2002). Estimating peak currents at ground lightning impacts using remanent magnetization. Geophysical
Research Letters,29(18), 1867. https://doi.org/10.1029/2002GL015207
Verwey, E. J. (1939). Electronic conduction of magnetite (Fe
3
O
4
) and its transition point at low temperature. Nature,144(3642), 327328.
https://doi.org/10.1038/144327b0
Viti, C. (2011). Exploring fault rocks at the nanoscale. Journal of Structural Geology,33(12), 17151727. https://doi.org/10.1016/j.
jsg.2011.10.005
Volk, M. W. R., & Feinberg, J. M. (2019). Domain state and temperature dependence of pressure remanent magnetization in synthetic
magnetite: Implications for crustal remagnetization. Geochemistry, Geophysics, Geosystems,20, 24732483. https://doi.org/10.1029/
2019GC008238
Volk, M. W. R., & Gilder, S. A. (2016). Effect of static pressure on absolute paleointensity recording with implications for meteorites.
Journal of Geophysical Research: Solid Earth,121, 55965610. https://doi.org/10.1002/2016JB013059
Walden, J., Oldeld, F., & Smith, J. (1999). Environmental magnetism: A practical guide,Technical Guide (Vol. 6). London: Quaternary
Research Association.
Wang, H., Li, H., Si, J., Sun, Z., & Huang, Y. (2014). Internal structure of the Wenchuan earthquake fault zone, revealed by surface outcrop
and WFSD1 drilling core investigation. Tectonophysics,619620, 101114. https://doi.org/10.1016/j.tecto.2013.08.029
Weiss, B., Lima, E., Fong, L., & Baudenbacher, F. (2007). Paleomagnetic analysis using SQUID microscopy. Journal of Geophysical
Research,112, B9105. https://doi.org/10.1029/2007JB004940
Weltje, G. J. (1997). Endmember modeling of compositional data: Numericalstatistical algorithm s for solving the explicit mixing problem.
Mathematical Geology,29(4), 503549. https://doi.org/10.1007/BF02775085
White, J. C. (1996). Transient discontinuities revisited: Pseudotachylyte, plastic instability and the inuence of low pore uid pressure on
deformation processes in the midcrust. Journal of Structural Geology,18(12), 14711486. https://doi.org/10.1016/S0191 8141(96)000594
Wibberley, C. A. J., Yielding, G., & Di Toro, G. (2008). Recent advances in the understandi ng of fault zone internal structure: A review,Special
Publications (Vol. 299, pp. 533). London: Geological Society.
Williams, J. N., Toy, V. G., Smith, S. A. F., & Boulton, C. (2017). Fracturing, uidrock interaction and mineralisation during the seismic
cycle along the Alpine Fault. Journal of Structural Geology,103, 151166. https://doi.org/10.1016/j.jsg.2017.09.011
Wilson, B., Dewers, T., Reches, Z., & Brune, J. (2005). Particle size and energetics of gouge from earthquake rupture zones. Nature,
434(7034), 749752. https://doi.org/10.1038/nature03433
Wise, D. U., Dunn, D. E., Engelder, J. T., Geiser, P. A., Hatcher, R. D., Kish, S. A., et al. (1984). Faultrelated rocks: Suggestions for ter-
minology. Geology,12(7), 391394. https://doi.org/10.1130/0091-7613(1984)12<391:FRSFT>2.0.CO;2
Woodcock, N. H., & Mort, K. (2008). Classication of fault breccias and related fault rocks. Geological Magazine,145(3), 435440. https://
doi.org/10.1017/S0016756808004883
Worm, H.U., Clark, D., & Dekkers, M. J. (1993). Magnetic susceptibility of pyrrhotite: Grain size, eld and frequency dependence.
Geophysical Journal International,114(1), 127137. https://doi.org/10.1111/j.1365246X.1993.tb01472.x
Yamaguchi, A., Cox, S. F., Kimura, G., & Okamoto, S. (2011). Dynamic changes in uid redox state associated with episodic fault rupture
along a megasplay fault in a subduction zone. Earth and Planetary Science Letters,302(34), 369377. https://doi.org/10.1016/j.
epsl.2010.12.029
Yamazaki, K. (2013). Improved models of the piezomagnetic eld for the 2011 Mw 9.0 Tohokuoki earthquake. Earth and Planetary Science
Letters,363,915. https://doi.org/10.1016/j.epsl.2012.12.019
Yang, T., Chen, J.Y., Wang, H.Q., & Jin, H.Q. (2012a). Rock magnetic properties of fault rocks from the rupture of the 2008 Wenchuan
Earthquake, China and their implications: Preliminary results from the Zhaojiagou outcrop, Beichuan County (Sichuan).
Tectonophysics,530531, 331341. https://doi.org/10.1016/j.tecto.2012.01.019
Yang, T., Chen, J.Y., Wang, H.Q., & Jin, H.Q. (2012b). Magnetic properties of fault rocks from the YingxiuBeichuan fault: Constra ints on
temperature rise within the shallow slip zone during the 2008 Wenchuan Earthquake and their implications. Journal of Asian Earth
Sciences,50,5260. https://doi.org/10.1016/j.jseaes.2012.01.013
Yang, T., Chen, J.Y., Xu, H.R., & Dekkers, M. J. (2019). Highvelocity friction experiments indicate magnetic enhancement and softening
of fault gouges during seismic slip. Journal of Geophysical Research: Solid Earth,124,2643. https://doi.org/10.1029/2018JB016341
Yang, T., Chen, J.Y., Yang, X.S., Wang, H.Q., & Jin, H.Q. (2013). Differences in magnetic properties of fragments and matrix of breccias
from the rupture of the 2008 Wenchuan earthquake, China: Relationship to faulting. Tectonophysics,601, 112124. https://doi.org/
10.1016/j.tecto.2013.05.002
Yang, T., Dekkers, M. J., & Chen, J.Y. (2018). Thermal alteration of pyrite to pyrrhotite during earthquakes: New evidence of seismic slip in
the rock record. Journal of Geophysical Research: Solid Earth,123, 11161131. https://doi.org/10.1002/2017JB014973
Yang, T., Dekkers, M. J., & Zhang, B. (2016). Seismic heating signatures in the Japan Trench subduction plateboundary fault zone:
Evidence from a preliminary rock magnetic geothermometer.Geophysical Journal International,205(1), 319331. https://doi.org/
10.1093/gji/ggw013
Yang, T., Mishima, T., Ujiie, K., Chester, F. M., Mori, J. J., Eguchi, N., Toczko, S., & Expedition 343 Scientists (2013). Strain decoupling
across the décollement in the region of large slip during the 2011 TohokuOki earthquake from anisotropy of magnetic susceptibility.
Earth and Planetary Science Letters,381,3138. https://doi.org/10.1016/j.epsl.2013.08.045
Yang, T., Yang, X.S., Duan, Q.B., Chen, J.Y., & Dekkers, M. J. (2016). Rock magnetic expression of uid inltration in the
YingxiuBeichuan fault (Longmen Shan thrust belt, China). Geochemistry, Geophysics, Geosystems,17,10651085. https://doi.org/
10.1002/2015GC006095
Yao, L., Ma, S., Niemeijer, A. R., Shimamoto, T., & Platt, J. D. (2016). Is frictional heating needed to cause dramatic weakening of nano-
particle gouge during seismic slip? Insights from friction experiments with variable thermal evolutions. Geophysical Research Letters,43,
68526860. https://doi.org/10.1002/2016GL069053
Yeh, E., Lee, T., Lin, Y., Chou, Y., & Lu, C. (2007). Analysis of magnetic and grain fabrics across fault gouge in the Che lungpu Fault of Taiwan
(Abstract T43B1331). Paper presented at American Geophysical Union Fall Meeting 2007, San Francisco, USA.
Zdujić, M., Jovalekić,Č., Karanović, L., Mitrić, M., Poleti, D., & Skala, D. (1998). Mechanochemical treatment of αFe
2
O
3
powder in air
atmosphere. Materials Science and Engineering A,245(1), 109117. https://doi.org/10.1016/S09215093(97)007156
Zhang, L., Li, H.B., Sun, Z.M., Chou, Y.M., Cao, Y., & Wang, H. (2018). Metallic iron formed by melting: A new mechanism for magne tic
highs in pseudotachylyte. Geology,46(9), 779782. https://doi.org/10.1130/G40153.1
Zhang, L., Sun, Z.M., Li, H.B., Zhao, L.S., Song, S.R., Chou, Y.M., et al. (2017). Rock record and magnetic response to large earthquakes
within Wenchuan Earthquake Fault Scientic Drilling cores. Geochemistry, Geophysics, Geosystems,18, 18891906. https://doi.org/
10.1002/2017GC006822
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 59 of 60
Zhao, X., Heslop, D., & Roberts, A. P. (2015). A protocol for variable resolution rstorder reversal curve measurements. Geochemistry,
Geophysics, Geosystems,16, 13641377. https://doi.org/10.1002/2014GC005680
Zhao, X., Roberts, A. P., Heslop, D., Paterson, G. A., Li, Y.L., & Li, J.H. (2017). Magnetic domain state diagnosis using hysteresis reversal
curves. Journal of Geophysical Research: Solid Earth,122, 47674789. https://doi.org/10.1002/2016JB013683
Zhu, K.Y., Li, M.Y., Shentu, L.F., Shen, Z.Y., & Yu, Y.H. (2017). Evaluation of a smalldiameter sampling method in magnetic sus-
ceptibility, AMS and Xray CT studies and its applications to mac microgranular enclaves (MMEs) in granite. Journal of Volcanology
and Geothermal Research,341, 208227. https://doi.org/10.1016/j.jvolgeores.2017.06.002
Zijderveld, J. D. A. (1967). A.C. demagnetization of rocks: Analysis of results. In D. W. Collinson, K. M. Creer, S. K. Runcorn (Eds.), Methods
in paleomagnetism (pp. 254286). New York: Elsevier.
Zoback, M. D., Barton, C. A., Brudy, M., Castillo, D. A., Finkbeiner, T., Grollimund, B. R., et al. (2003). Determination of stress orientation
and magnitude in deep wells. International Journal of Rock Mechanics and Mining Sciences,40(78), 10491076. https://doi.org/10.1016/
j.ijrmms.2003.07.001
Zoback, M. D., Hickman, S., & Ellsworth, W. (2007). The role of fault zone drilling. In H. Kanamori & G. Schubert (Eds.), Earthquake
seismology, treatise on geophysics (Vol. 4, pp. 649674). Amsterdam: Elsevier.
10.1029/2019RG000690
Reviews of Geophysics
YANG ET AL. 60 of 60
... Molina Garza et al., 2009;Parés, 2015;Yang et al., 2020). Hence, to interpret these fabrics, we need to determine the timing of magnetic fabric acquisition and the nature of deformation mechanisms.We chose the Heart Mountain Slide (HMS) in northwest Wyoming to address these questions. ...
... We investigate the origin of the Anisotropy of Magnetic Susceptibility (AMS) in the utracataclasite and surrounding rocks with rock magnetism methods and validate the AMS using independent methods, including shape preferred orientation (SPO), scanning electron microscopy (SEM), and energy dispersive spectroscopy (EDS). We chose to use magnetic methods to investigate this example of extreme deformation in carbonates because these approaches tend to be very effective in detecting synkinematic mineralogical changes in fault rocks and deformed materials (e.g., Yang et al., 2020). For the first time since Craddock et al. (2009), our study provides new information about the internal fabric of the slide. ...
... siderite ⇔ magnetite + carbon dioxide (6) Similar reactions are commonly described in carbonate fault rocks (e.g., Liu et al., 2022;Tanikawa et al., 2008;Yang et al., 2020). Since the Bighorn Dolomite contains up to 1.5 wt% Fe 2 O 3 (Blackwelder, 1913), it is conceivable that some of this iron forms impurities in dolomite grains. ...
Article
Full-text available
The Heart Mountain Slide in Wyoming is one of the largest known terrestrial gravity slides (3,500 km²) formed ∼49 Ma ago by the nearly horizontal detachment of Paleozoic‐Eocene cover sliding on top of autochthonous formations. At the White Mountain locality, exposures offer an exceptional opportunity to investigate high strain rate/high velocity processes in carbonates. Here we use the anisotropy of magnetic susceptibility (AMS) of 274 samples to shed light on ultracataclastic deformation along this detachment. Contrary to predictions, the carbonate ultracataclasite displays a consistent AMS fabric, particularly in the upper ultracataclasite. The AMS in this unit is controlled primarily by magnetite formed through the breakdown of iron sulfides caused by frictional heating. Additional thermomagnetic experiments reveal that the new magnetic fabric began forming ∼250ºC and continued up to ∼400ºC when calcination of carbonate minerals caused a major drop in friction. The main cataclastic slip direction inferred from AMS is ∼N033°, at odds with the previously accepted NNW‐SSE direction. We validate these AMS fabrics through 3D shape preferred orientation analysis and micro X‐ray scanning of the same specimens. These results, however, may only represent cataclastic flow directions at the local scale as a result of synkinematic rotation of the White Mountain block. Alternatively, these results may call for a re‐evaluation of the large scale movement of the slide. Finally, this study demonstrates the usefulness of a magnetic approach in deciphering deformation processes in carbonates, particularly in high strain rate cases such as seismic faults.
... The piezomagnetic effect describes in magnetization of ferromagnetic minerals under the applied mechanical stress (e.g. Hao et al. 1982 ;Hamano et al. 1989 ;Nakamura & Nagahama 1997 ;Yang et al. 2020 ). In tectonically active regions, changes in stress of the crust yield changes in magnetization due to the piezomagnetic effect, which in turn yields changes in the local magnetic field (LMFs, Davis 1974 ;Currenti et al. 2009 ). ...
... To establish the quantitative relationship between changes in stress and changes in magnetization, many e xperiments hav e been carried out by using rock sample under laboratory conditions (e.g. Hao et al. 1982 ;Hamano et al. 1989 ;Yang et al. 2020 ). Results obtained by laboratory studies have been summarized to a simple linear law (Sasai 1980 ). Based on the conclusions in the laboratory (Stacey 1963 ;Sasai 1991b ;Utsugi et al. 2000 ;Okubo & Oshiman 2004 ;Currenti et al. 2007 ;Yamazaki 2016Yamazaki , 2021, it successfully explains many magnetic anomalies caused by volcanoes and earthquakes and provides information on the piezomagnetic characteristics of rocks (Breiner 1964 ;Davis 1974 ;Johnston & Mueller 1987 ;Currenti et al. 2009 ;Okubo et al. 2011 ;Utada et al. 2011 ;Yamazaki 2013 ;Song et al. 2022 ). ...
Article
Full-text available
The fundamental behaviours of rock magnetism (piezomagnetic characteristics) at the field scale have not been confirmed because conventional experiments can be performed only in the laboratory. Here, the periodic extraction and injection of gas in the Hutubi ultralarge underground gas storage (UGS) system are used to simulate the stress loading and unloading processes at the field scale. We treat 26 gas wells in the UGS system as a multipoint-source Mogi model and calculate models of the piezomagnetic field generated during the operation of the UGS system. These models show that the local magnetic field (LMF) in the southern and central areas of the UGS system showed positive changes. In contrast, the northern area showed negative changes, and the amplitude of the negative changes was smaller than that of the positive changes. Changes in the Curie point depth and gas volume do not significantly alter the spatial characteristics of the LMF.
... Liu et al., 2014;Robustelli Test & Zanella, 2021;Yang et al., 2012Yang et al., , 2018Yang et al., , 2019L. Zhang et al., 2018; for a review, see Yang et al., 2020). ...
... It has been proposed that the candidate mechanisms responsible for magnetic changes/alterations in fault rocks occurred at the shallow crust mainly includes (e.g., Chou et al., 2012b;Ferré et al., 2012Ferré et al., , 2017Hirono et al., 2006;Mishima et al., 2006;Tanikawa et al., 2007Tanikawa et al., , 2008Yang et al., 2020): (a) grain-size variations of pre-existing ferrimagnetic minerals by grain fining, (b) coseismic frictional melting with pseudotachylyte formation, (c) thermochemical reactions induced by frictional heating, and (d) chemical alteration and neomineralization due to fluid percolation. ...
Article
Full-text available
Plain Language Summary As the products of faulting processes, physical and chemical properties of fault rocks convey important information for understanding the faulting processes. Iron, as the fourth most abundant element in the earth's crust, widely occurs in fault zones as iron‐oxides, ‐hydroxides, and/or ‐sulfides. Moreover, these Fe‐bearing (magnetic) minerals are sensitive to stress‐strain, temperature change, and fluid activity. Here, detailed rock magnetic analyses in combination with electron microscopic observations are conducted on the fault rocks and protoliths from the Shaba outcrop (Beichuan County) on the Yingxiu‐Beichuan Fault hosting the 2008 Wenchuan Mw 7.9 earthquake. It is found that pyrrhotite and goethite are only present and confined to the fault gouges just next to the principal slip surface. Their formation is resulted from the alteration of the pre‐existing pyrite and Fe‐bearing clay minerals driven by coseismic frictional heating and concomitant hot fluids. These results suggest the occurrence of thermal pressurization that played an essential role in coseismic weakening during the earthquake. Thus, it demonstrates that the rock magnetic properties of fault rocks could provide valuable insights into faulting‐related physical and chemical processes.
... Faults in intracontinental settings typically evolve through numerous stages of deformation, reflecting changing tectonic conditions over time (Allen et al., 1998;Cunningham and Mann, 2007). Such mature faults can accommodate strain over various spatial scales (from submillimeter to kilometer scales) and time intervals (from seconds for coseismic slip to years for slow slip or aseismic creep) (Sibson, 2003;Nielsen, 2017;Yang et al., 2020). Given the multidimensional nature of faults, microstructural analyses of fault rocks can reveal the mechanical processes and conditions governing the deformation of fault rocks, which is important for understanding fault evolution (Rutter et al., 1986;Bos and Spiers, 2001;Boullier et al., 2009;Hadizadeh et al., 2012;Bradbury et al., 2015). ...
Article
The Yangsan Fault, a long-lived intracontinental fault in SE Korea, exhibits various slip behaviors, including coseismic slip and aseismic creep. However, there is insufficient knowledge of deformation microstructures to reveal the deformation mechanisms operating within the fault. In this study, we present an analysis of the mechanical behaviors displayed by the Byeokgye section of the Yangsan Fault over seismic cycles. Our results are based on detailed microscopic observations of drillcore samples recovered from the Byeokgye section, using an electron backscattered diffraction (EBSD) technique. In injected calcite veins located close to the principal slip zone (PSZ) of < 2 cm in width, plastic deformation (including dynamic recrystallization by subgrain rotation and deformation twins) is concentrated in the blocky calcite grains. In a narrow microbrecciated slip zone (< 1 cm wide) within the granitic damage zone, we observed mechanical Dauphiné twins associated with fractures and microfaults in quartz, as well as intergranular pressure solution (IPS) in the quartz fragments. Given that dynamic recrystallization and IPS are indicative of mechanical behavior of aseismic creep, it is possible that aseismic creep occurs upon the fault during interseismic periods. Conversely, the presence of mechanical Dauphiné twins, coupled with the nature of the PSZ, gouge injections, and the blocky structure of calcite veins, suggests the exposure of the fault section to local seismic stresses during coseismic slip. In conclusion, various deformation processes have operated upon the Yangsan Fault at the studied section throughout multiple seismic cycles. Furthermore, our study demonstrates that EBSD analysis is an effective technique for elucidating the mechanical behavior of fault zones.
... ZIEBARTH ET AL. Hamada et al., 2009) and typical expectation (e.g., Lachenbruch & Sass, 1980;Yang et al., 2020) place the endothermic chemical energy contribution to the fault energy budget at an insignificant level compared to the frictional energy. We hence discard the chemical energy and leave a quantitative estimate for the SAF up for future work. ...
Article
Full-text available
We investigate the relation between frictional heating on a fault and the resulting conductive surface heat flow anomaly using the fault's long‐term energy budget. Analysis of the surface heat flow surrounding the fault trace leads to a constraint on the frictional power generated on the fault—the mechanism behind the San Andreas fault (SAF) heat flow paradox. We revisit this paradox from a new perspective using an estimate of the long‐term accumulating elastic power in the region surrounding the fault, and analyze the paradox using two parameters: the seismic efficiency and the elastic power. The results show that the constraint on frictional power from the classic interpretation is incompatible with the accumulating elastic power and the radiated power from earthquake catalogs. We then explore four mechanisms that can resolve this extended paradox. First, stochastic fluctuations of surface heat flow could mask the fault‐generated anomaly (we estimate 21% probability). Second, the elastic power accumulating in the region could be overestimated (≥550 MW required). Third, the seismic efficiency—ratio of radiated energy to elastic work—of the SAF could be higher than that of the remaining faults in the region (≥5.8% required). Fourth, the scaled energy—ratio of radiated energy to seismic moment—on the SAF could be lower than on the remaining faults in the region (a factor 5 difference required). In the last three hypotheses, we analyze the interplay of the energy budget on a single fault with the total energy budget of the region.
... These compounds play a significant role in chemical, biological, and geological processes. 1,2 The iron (II and III) and sulfate ions are one of the most common constituents of the lithosphere 3 and have been recently detected in extraterrestrial environments. 4−6 Moreover, the sulfate anion is a strong kosmotropic molecule in terms of a Hofmeister series. ...
Article
A new guanidinium-templated hydrated iron sulfate, [CN3H6][FeIIFeIII(SO4)3(H2O)3] (1), was prepared from strongly acidic aqueous solutions. Its crystal structure is comprised from FeIIIO6 and FeIIO3(H2O)3 octahedra linked by sulfate bridges forming a [FeIIFeII(SO4)3(H2O)3]− 3D framework with a layer-by-layer ordering of ferric and ferrous cations. The structural topology of the framework is related to the anhydrous rhombohedral mikasaite Fe2(SO4)3. The removal of part of the sulfate tetrahedra and the partial replacement of the Fe3+ cations in the [Fe3+2(SO4)3]0 framework by Fe2+ provide a negative charge and allow the incorporation of the protonated organic species in the voids. The compound 1 has been characterized by single-crystal X-ray diffraction, TG and DSC analyses, UV–vis–NIR spectroscopy, magnetic susceptibility, Mössbauer spectroscopy, IR and Raman spectroscopy, and density functional band-structure calculations. The magnetic behavior of 1 shows an interplay of FeII (S = 2) and FeIII (S = 5/2) sublattices that exhibit different types of antiferromagnetic couplings, one FeIII–FeIII (J1 ∼ 6.1 K) and two FeII–FeIII couplings (J2 ∼ 1 K, J3 ∼ 5.9 K) within corrugated honeycomb layers. These ferrimagnetic layers are coupled antiparallel to each other, resulting in an overall antiferromagnetic order below TN = 31 K.
... Therefore, as noted by Borradaile (1988) "magnetic fabrics should not be used for routine methods of 'strain analysis' without further study" as for example, composite fabrics of sedimentary, compaction and tectonic origin can result in distinct AMS due to lithology based differences in strain partitioning (Evans et al., 2003). All together this establishes the necessity to combine magnetic studies with microstructural, mineral chemistry and geochemical characterization as the relationship between fault or SZ kinematics and dynamics can be individual Narloch et al., 2021;Robustelli Test & Zanella, 2021;Yang et al., 2020). ...
Article
Full-text available
The majority of the strain in Earth crust and upper mantle is localized to the high strain zones developed at ductile-to-brittle condition at kilometer-to-micrometer scale. Therefore, they represent the key to understanding the deformation evolution of the lithosphere. The finite strain pattern recorded within these zones has been therefore a subject of research in geology. The methods studying rock magnetism such as the anisotropy of magnetic susceptibility (AMS) are frequently used techniques to characterize and quantify deformation and flow record in rocks. Numerous sedimentary, subsolidus and submagmatic deformation zones exhibit typical evolution of the AMS ellipsoid across the strain gradient suggesting indirect not straightforward correlation between AMS and strain ellipsoids. To document spatiotemporal and internal fabric evolution during strain localization, pure shear, simple shear, and shear zone (SZ) analog experiments were performed using shear-thinning thixotropic material of plaster of Paris. The experimental results closely resemble the record from natural SZs in sedimentary rock systems but also in subsolidus SZs and submagmatic mushy systems. The magnetic fabric evolution across deformation zones is interpreted to be associated with the intersection and transposition of preexisting primary fabric with shear fabrics and evolution of synkinematic subfabrics. Their development is attributed to localization of deformation at microscale due to the self-organized slip of anisometric particles forming microshear planes reflecting the symmetry of deformation. The experimental results when confronted with the natural examples implies that the localization and partitioning of deformation is one of the most important factors for the interpretation of AMS in deformation zones.
Article
Full-text available
This study examines fluctuations in the geothermal gradient, heat flow, and Curie point depth in southeast Nigeria. The study employs statistical correlation methods to examine the relationship between several geothermal indicators to gain a thorough understanding of the subsurface thermal dynamics and geological characteristics of the area. The aim of employing correlation approaches in statistical analysis is to develop a model that elucidates the connections among heat flow, Curie point depth, and geothermal gradient values in SouthEastern Nigeria. The research area spans from 5°001' N to 6°301' N latitude and from 7°001' E to 8°301' E longitude, covering approximately 27,225 km 2. The Curie point's depth, heat flow, and geothermal gradient were calculated by detailed analysis of high-resolution aircraft magnetic data, offering valuable information about the area's geothermal characteristics. Curie point depths have been verified to vary between 20.416 and 29.544 km, with geothermal gradients varying from 16.92 to 26.78 o C/m 2. The predicted heat flow varies between 42.30 and 66.95 mWm-2. The study found a negative correlation between heat flow and Curie Point Depth, indicating that heat flow reduces as Curie depth increases in the studied area. Heat flow, geothermal gradient, and Curie point depth are significantly correlated, with R 2 values exceeding 0.5, as determined through statistical analysis utilizing the correlation technique. The visual representation highlights the significance of the interaction between these components by elucidating their connection and shared principles. This indicates a strong correlation between heat flow, geothermal gradient, and Curie point depth, emphasizing their interdependence.
Article
In order to better constrain the Mesozoic-Cenozoic evolution of the Precambrian Hoggar shield, a paleomagnetic study, combined with detailed fieldwork, was carried out to date its detrital local cover, the Serouenout Formation. Thermal demagnetization yields, only in a few samples, the characteristic remanent magnetization carried by hematite. Post-tilting remagnetization was obtained in sites located along a fault affected by intense fluids circulation. The paleomagnetic directions recorded at seven widespread other sites are on the contrary associated with a positive fold test. It provides a relatively well-defined paleomagnetic pole (A95 = 4.1°), sufficient to estimate the age of the Serouenout Formation. The comparison of this pole with the reference curve of Africa suggests two possible age windows, Triassic and Upper Cretaceous-Lower Paleocene, while the discovery in the uppermost levels of the Serouenout Formation of a fern-rich level with Weichselia reticulata (Bathonian to Cenomanian; Blanco-Moreno et al., 2018) imposes a deposition during the Cenomanian. The presence of a detrital formation at least 350 m thick, with a basal conglomerate containing large pebbles, implies the existence, during this time period, of a tectonic event that generated differential uplift. In addition, structural observations indicate that the Serouenout Formation recorded later brittle tectonics, dominated by a network of vertical N–S dextral faults. The horizontal displacement generated by one well-developed fault has been estimated to be at least 1 km. This activity is related to the known Alpine reactivations of the N–S Pan-African mega-structures, which are still at the origin of the current intraplate seismicity.
Chapter
Rifts can be summarized as geographical regions consisting of extensional sedimentary basins of various sizes, with various tectonic and sedimentary geometries that are linked in various structural contexts. This chapter provides a list of the main types of rifts and case examples. It provides the definition of the active and passive rifting categories, as these are regularly mentioned in the literature and are the primary designations of extensional rift settings. The chapter lists the major tectonic structures and basin types encountered in rifts and rifted margins. Subsidence is the stage that leads to the progressive deepening of the basin floor and hence allows the accumulation of sediments in rift basins. It is the process by which the lithosphere regains isostatic equilibrium. Sedimentation, although always present in rift basins, is extremely variable from one system to the other. Traditionally, sedimentary sequences of rift basins are subdivided into three categories: pre‐, syn‐ and post‐rifts.
Article
Full-text available
Fault kinematics can provide information on the relationship and assembly of tectonic units in an orogen. Magnetic fabric studies of faults where pseudotachylytes form have recently been used to determine direction and sense of seismic slip in prehistoric earthquakes. Here we apply this methodology to study magnetic fabrics of pseudotachylytes in field structures of the Köli Nappe Complex (central Swedish Caledonides), with the aim to determine fault kinematics and decipher the role of seismic faulting in the assembly of the Caledonian nappe pile. Because the pseudotachylyte veins are thin, we focused on small (ca. 0.2 to 0.03 cm3) samples for measuring the anisotropy of magnetic susceptibility. The small sample size challenges conventional use of magnetic anisotropy and results acquired from such small specimens demand cautious interpretation. Importantly, we find that magnetic fabric results show inverse proportionality among specimen size, degree of magnetic anisotropy and mean magnetic susceptibility, which is most likely an analytical artifact related to instrument sensitivity and small sample dimensions. In general, however, it is shown that the principal axes of magnetic susceptibility correspond to the orientation of foliation and lineation, where the maximum susceptibility (k1) is parallel to the mineral lineation, and the minimum susceptibility (k3) is dominantly oriented normal to schistosity. Furthermore, the studied pseudotachylytes develop distinct magnetic properties. Pristine pseudotachylytes preserve a signal of ferrimagnetic magnetite that likely formed during faulting. In contrast, portions of the pseudotachylytes have altered, with a tendency of magnetite to break down to form chlorite. Despite magnetite breakdown, the altered pseudotachylyte mean magnetic susceptibility is nearly twice that of altered pseudotachylyte, likely originating from the Fe-rich chlorite, as implied by temperature-dependent susceptibility measurements and thin-section observations. Analysis of structural and magnetic fabric data indicates that seismic faulting occurred during exhumation into the upper crust, but these data yield no kinematic information on the direction and sense of seismic slip. Additionally, the combined structural field and magnetic fabric data suggest that seismic faulting was postdated by brittle E–W extensional deformation along steep normal faults. Although the objective of finding kinematic indicators for the faulting was not fully achieved, we believe that the results from this study may help guide future studies of magnetic anisotropy with small specimens (<1 cm3), as well as in the interpretation of magnetic properties of pseudotachylytes.
Article
Full-text available
Extreme slip at shallow depths on subduction zone faults is a primary contributor to tsunami generation by earthquakes. Improving earthquake and tsunami risk assessment requires understanding the material and structural conditions that favor earthquake propagation to the trench. We use new biomarker thermal maturity indicators to identify seismic faults in drill core recovered from the Japan Trench subduction zone, which hosted 50 m of shallow slip during the Mw9.1 2011 Tohoku-Oki earthquake. Our results show that multiple faults have hosted earthquakes with displacement ≥ 10 m, and each could have hosted many great earthquakes, illustrating an extensive history of great earthquake seismicity that caused large shallow slip. We find that lithologic contrasts in frictional properties do not necessarily determine the likelihood of large shallow slip or seismic hazard. In this study, the authors investigate thermal alteration of organic biomarkers to detect paleo earthquakes in the Japan Trench. The study shows that large earthquakes like the 2011 Tohoku-Oki earthquake can slip through different types of sediment rather than being restricted to the weakest layers.
Book
Rock Magnetism, first published in 1997, is a comprehensive treatment of fine particle magnetism and the magnetic properties of rocks. Starting from atomic magnetism and magnetostatic principles, the authors explain why domains and micromagnetic structures form in ferromagnetic crystals and how these lead to magnetic memory in the form of thermal, chemical and other remanent magnetizations. The phenomenal stability of these magnetizations, providing a record of plate tectonic motions over millions of years, is explained by thermal activation theory. One chapter is devoted to practical tests of domain state and paleomagnetic stability; another deals with pseudo-single-domain magnetism. The final four chapters place magnetism in the context of igneous, sedimentary, metamorphic, and extraterrestrial rocks. This book will be of great value to graduate students and researchers in geophysics and geology, particularly in paleomagnetism and rock magnetism, as well as physicists and electrical engineers interested in fine-particle magnetism and magnetic recording.
Book
Originally published in 2005, this book covers the closely related techniques of electron microprobe analysis (EMPA) and scanning electron microscopy (SEM) specifically from a geological viewpoint. Topics discussed include: principles of electron-target interactions, electron beam instrumentation, X-ray spectrometry, general principles of SEM image formation, production of X-ray 'maps' showing elemental distributions, procedures for qualitative and quantitative X-ray analysis (both energy-dispersive and wavelength-dispersive), the use of both 'true' electron microprobes and SEMs fitted with X-ray spectrometers, and practical matters such as sample preparation and treatment of results. Throughout, there is an emphasis on geological aspects not mentioned in similar books aimed at a more general readership. The book avoids unnecessary technical detail in order to be easily accessible, and forms a comprehensive text on EMPA and SEM for geological postgraduate and postdoctoral researchers, as well as those working in industrial laboratories.
Article
Characterizing the inelastic strain field around co-seismic faults is a key to understanding earthquake rupture processes. We aim to detect zones of inelastic deformation around surface ruptures, using Anisotropy of Magnetic Susceptibility (AMS) analyses. To characterize the paleo-seismic setting we measured 29 normal faults, which cut the late Pleistocene lacustrine Lisan Formation and were associated with shallow (<~12 km) earthquakes at the margins of the Dead Sea Basin. In order to define the geometry of those zones around a representative fault, 220 specimens were sampled in a 2D grid and analyzed by studying the room temperature AMS axes and parameters using spatial IDW (inverse distance weighting) interpolation on a cross section around the fault plane. To separate the total magnetic susceptibility of the Lisan sediments into diamagnetic, paramagnetic and ferromagnetic susceptibilities, we applied the low temperature AMS (LT-AMS) and Anhysteretic Remanent Magnetization (AARM) methods. The structural measurements demonstrate that the co-seismic faulting in the Dead Sea Basin, triggered mainly by E-W extension. The results show that the paramagnetic phase is dominant in both the aragonite-rich and the detritus-rich layers. The zones of inelastic deformation are distributed asymmetrically about the fault plane, with affinity to the observed displacement profile and in agreement with dynamic propagation models that suggest that the rupture propagated towards the surface. The discontinuous distribution of deformation around the fault plane is the result of both the faults single event propagation style and fault plane asperities. These results demonstrate a novel application of AMS analysis for defining the geometry of the inelastic strain field surrounding co-seismic faults, and improving fault propagation models dealing with the inelastic off-fault response.
Article
The study of rock fabric properties (orientation, planar, linear, anisotropy) is key to unravelling the geological processes that generated them. With advancements in data acquisition and treatment, X-ray micro-computed tomography (µXCT) represents a powerful method to analyse the shape preferred orientation (SPO) of rock-forming elements, including minerals, aggregates, and pores, in the three-dimensional space. After reconstruction and segmentation of µXCT images, we developed a novel protocol to construct and analyse the fabric tensor, a second-rank symmetric tensor constructed using the orientation and the length of the three characteristic axes of each grain (simplified to a best fit ellipsoid). The analysis of the fabric tensor permits calculation of mean principal directions and associated confidence ellipses, and quantifies the degree of anisotropy (P′) and the shape (T) of the fabric ellipsoid by eigenvalue and eigenvector analysis. We implement this method in the TomoFab open-source MATLAB package. The code integrates a graphical user interface (GUI) that allows the visualization of the full set of ellipsoid orientation, shape, and size. Density plots and contouring can be utilised to identify fabrics graphically, and a full set of fabric parameters can be calculated based on the analysis of the fabric tensor and/or the analysis of each principal direction orientation tensor. We demonstrate the versatility of TomoFab with synthetic datasets and a field- and laboratory-based investigation of a sample presenting a magmatic foliation and lineation, collected in the Mafic Complex within the lower crustal section of the Ivrea-Verbano Zone (North Italy). In the light of these developments, we stress that µXCT represents a pertinent tool for rock fabric analysis to characterise the SPO of rock components. This approach can be performed parallel or complementary to other rock fabric quantification methods (e.g., AMS, EBSD) and applied to various rock types. TomoFab is freely available for download at https://github.com/benpetri/tomofab
Article
Differences in REE patterns of calcite from extensional and shear veins of the Sestola Vidiciatico Tectonic Unit in the Northern Apennines suggest variations in fluid source during the seismic cycle in an ancient analogue of a shallow megathrust (Tmax c. 100–150°C). In shear veins, a positive Eu anomaly suggests an exotic fluid source, probably hotter than the fault environment. Small-scale extensional veins were derived instead from a local fluid in equilibrium with the fault rocks. Mutually crosscutting relations between two extensional vein sets, parallel and perpendicular to the megathrust, suggest repeated shifting of the σ1 and σ3 stresses during the seismic cycle. This is consistent with: (1) a seismic phase, with brittle failure along the thrust, crystallization of shear veins from an exotic fluid, stress drop and stress rotation; (2) a post-seismic phase, with fault-normal compaction and formation of fault-normal extensional veins fed by local fluids; (3) a reloading phase, where shear stress and pore pressure are gradually restored and fault-parallel extensional veins form, until the thrust fails again. The combination of geochemical and structural analyses in veins from exhumed megathrust analogues represents a promising tool to better understand the interplay between stress state and fluids in modern subduction zones.