Chang 2016
Chang 2016
Chang 2016
Liao Chang1,2, Andrew P. Roberts2, David Heslop2, Akira Hayashida3, Jinhua Li4, Xiang
1. School of Earth and Space Sciences, Peking University, Beijing 100871, P. R. China
100029, P. R. China
Abstract Magnetic mineral inclusions occur commonly within other larger mineral
phases in igneous rocks and have been demonstrated to preserve important paleomagnetic
signals. While the usefulness of magnetic inclusions in igneous rocks have been explored
extensively, their presence in sediments has only been speculated upon. The contribution of
magnetic inclusions to the magnetization of sediments, therefore, has been elusive. In this
study, we use transmission electron microscope (TEM) and magnetic methods to demonstrate
settings. TEM analysis reveals detailed information about the microstructure, chemical
composition, grain size, and spatial arrangement of nanoscale magnetic mineral inclusions
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1002/2016JB013109
hydrodynamic forces rather than by geomagnetic torques. Thus, even though these large
silicates may contain ideal single domain particles, they can not contribute meaningfully to
paleomagnetic recording. However, smaller silicate grains (e.g., silt- and clay-sized) silicates
with unidirectionally magnetized magnetic inclusions can potentially record a reliable DRM.
1. Introduction
Magnetic iron-titanium oxide mineral inclusions hosted within silicate minerals, e.g.,
widely in igneous and metamorphic rocks. The importance of magnetic inclusions for
paleomagnetic studies has been recognized since the discovery of ultrafine-grained iron
oxides in silicates in the Modipe Gabbro [Evans et al., 1968; Evans and Wayman, 1970].
More recently, there has been renewed interest in magnetic mineral inclusions due to the
al., 2001, 2006, 2010; Muxworthy and Evans, 2013; Sato et al., 2015]. Paleomagnetic signals
from magnetic mineral inclusions have been investigated extensively to expand knowledge of
past magnetic field behavior of the Earth and other bodies within the Solar System [e.g.,
Tarduno et al., 2001, 2010, 2015; Feinberg et al., 2005, 2006; Lappe et al., 2011, 2013;
Muxworthy et al., 2013; Usui et al., 2015]. Unlike magnetically unstable multi-domain (MD)
particles, magnetic mineral inclusions often occur as fine-grained stable single-domain (SSD)
or small pseudo-single-domain (PSD) particles [e.g., Harrison et al., 2002] that are capable
of carrying stable remanences over billions of years [Evans et al., 1968]. Moreover, silicate
host minerals can protect magnetic inclusions against changes in the local environment that
can give rise to chemical alteration. These characteristics make silicate hosted magnetic
mineral inclusions a promising candidate for retaining reliable paleomagnetic signals over
The presence of magnetic mineral inclusions in igneous rocks and their important
contributions to paleomagnetic records are well established. But knowledge of the presence
contribution to sedimentary magnetic signals have been elusive. In this study, we investigated
marine sediment samples to search for magnetic mineral inclusions using transmission
electron microscope (TEM) and magnetic analyses. We also modeled the depositional
outcrops were investigated in this study (Fig. 1). Core MD01-2421 (36°01.4′N; 141°46.8′E;
2,224 m water depth; 45.82 m long) was recovered from the North Pacific Ocean ~100 km
offshore of central Japan [Oba et al., 2006; Chang et al., 2016b]. Sediments in this core are
homogenous olive-gray silty clays with calcareous and siliceous microfossils, with high total
organic carbon (TOC) content (0.5-2.1 wt.%; Ueshima et al. [2006]). Sample “MD01-2421-
7-110” from a depth of 110 cm in core section 7 (at a depth of 10.06 m) was analyzed. The
studied sample was selected from bulk sediment with no evident disseminated volcanic ash.
Core CD143-55705 (22°22.4′ N, 60°08.0′ E; 2,193 m water depth; 10.63 m long) was
recovered on the continental margin of Oman, northwestern Arabian Sea [Rowan et al.,
hemipelagic clays, and are high in TOC (1-2%) [Rowan et al., 2009]. Sample “CD143-
55705-7-82” from a depth of 82 cm in core section 7 (at 7.49 m depth) was analyzed. Central
equatorial Pacific Ocean sediment core RR0603-03JC (2°33′ N, 117°55′ E; 4195 m water
depth) was recovered during the AMAT03 site survey cruise for Integrated Ocean Drilling
Program Proposal 626. The lithology of this core is mainly diatom nannofossil ooze. The
studied sample “RR0603-03JC-2-60” is from a depth of 60-62 cm in core section 2 (at 1.04
m depth). Dust may be an important component of the studied samples from the equatorial
Pacific Ocean and Arabian Sea. Marine sediment samples (magnetic separate sample
“BL37,38,39” and bulk sediment sample “BR49D” that are close to each other
stratigraphically) were collected from tectonically uplifted Upper Miocene marine sediments
exposed in Blind River, Lower Awatere Valley, northeastern South Island, New Zealand
[Roberts and Turner, 1993]. The succession contains siliciclastic marine sediments of the
Awatere Group and are probably derived from greywacke basement rocks and igneous
Table 1.
were made with a MicroMag vibrating sample magnetometer (model 3900) at the Australian
National University (ANU). FORC measurements [Roberts et al., 2000, 2014] were made
with a field step of 1.5 mT, maximum applied fields of 1 T, and averaging times of 200-400
ms. For some magnetically weak samples, we followed the protocol of Zhao et al. [2015],
in which 120-160 FORCs with irregular measurement grids were measured with averaging
times of 200-400 ms. FORC data were processed using the software package of Zhao et al.
[2015]. No data pre-treatments, i.e., removal of first-point artefact and subtraction of lower
branch [Egli, 2013], were applied. Low-temperature (LT) magnetic properties were measured
with a Quantum Design Magnetic Property Measurement System (MPMS; model XL7) at
ANU. For warming of a saturation isothermal remanent magnetization (SIRM), samples were
first cooled to 10 K in either zero field (zero-field cooled; ZFC) or in a 5 T field (field-
cooled; FC). At 10 K, a 5 T field was applied and was then switched off to impart a LT SIRM,
and the MPMS magnet was reset. ZFC and FC curves were measured during zero-field
Magnetic minerals were separated from bulk sediments following Chang et al. [2012]
using a Frantz isodynamic magnetic separator. TEM observations were carried out with a
JEOL 2100F field-emission (FE) TEM and a Philips CM300 TEM at the Centre for Advanced
Microscopy (CAM), ANU, and with a JEOL 2100 TEM at the Institute of Geology and
Geophysics, Chinese Academy of Sciences (CAS). The JEOL 2100F at CAM is equipped
with a FE gun and scanning transmission electron microscope (STEM) detectors, and is
operated at 200 kV. STEM observations were performed in the high-angle annular dark field
(HAADF) mode. Energy dispersive X-ray spectroscopy (EDS) analysis was performed using
STEM HAADF mode, with a focused electron beam of a few nm. The Philips CM300 TEM
at CAM is equipped with an EDAS Phoenix retractable X-ray detector and a Gatan CCD
camera, and is operated at 300 kV. The JEOL 2100 TEM at CAS was operated at 200 kV.
To model DRM, magnetic and hydrodynamic torques that act on a settling detrital
particle were compared to assess the ability of magnetic inclusion-bearing particles to align
with the geomagnetic field. Detrital sediment particles that contain inclusions are assumed to
be prolate ellipsoids. The aspect-ratio (ratio of the semi-major axis to the semi-minor axis) of
such particles plays a key role in controling their orientation as they settle through the water
as the aspect-ratio of a particle increases, the hydrodynamic torque also increases and tends to
rotate a settling ellipsoid so that its long axis is horizontal. The approximation provided by
Heslop [2007] was employed to find the maximum hydrodynamic torque, τH, that acts on a
settling particle with a given volume and aspect-ratio. Magnetic nano-inclusions within
detrital particles are assumed to be SSD magnetite particles with diameters of 100 nm that are
the magnetite assemblage, we assume that the particles carry a weak-field thermoremanent
magnetization (TRM). Dunlop [1990] demonstrated that weak-field TRM in magnetite varies
acquiring a TRM of ~10 kA/m in a 100-µT field. We assume that the TRM intensity is
proportional to field strength, so we scale this empirical value for a typical geomagnetic field
strength of 50 µT. Finally, we multiply the resulting TRM by a factor of 2 to remove the
partial cancelation that occurs over a collection of randomly oriented particles. This process
yields an estimated TRM of 10 kA/m for aligned particles in a 50-µT field. Assuming that
3. Results
were made on bulk sediment samples (sister samples of the studied TEM samples) and a
and to constrain results from TEM observations. The FORC diagram for sample “MD01-
(Fig. 2a). LT warming of SIRM reveals a weak double Verwey transition (Tv) signature (Fig.
2b) that is indicative of the presence of small amounts of both biogenic and inorganic
magnetite [e.g., Chang et al., 2016a], where the more pronounced Tv at ~120 K is mostly
likely to be associated with the presence of detrital magnetite. It should be noted that the
double Tv signature is not observed commonly for other sediment samples from core MD01-
2421 [Chang et al., 2016b]. The FORC diagram for sample “RR0603-03JC-2-60” from the
central equatorial Pacific Ocean contains a dominant central-ridge signature associated with
interactions (Fig. 2c). LT magnetic measurements for this sample did not reveal a clear Tv
signal (Fig. 2d). The FORC diagram for sample “CD143-55705-7-82” from the Oman
interactions (Fig. 2e). This sample has a pronounced Tv at ~120 K (Fig. 2f). Sample “BR49D”
from the Lower Awatere Valley, New Zealand, has two major FORC components: a SD
reveal a pronounced Tv at ~120 K (Fig. 2h). The FORC diagram for magnetic separate sample
“BL37,38,39”, which is from an outcrop close to that from which sample BR49D was taken,
contains two major FORC distributions: a SD component with moderate vertical spread and a
low-coercivity component (Fig. 2i). We plot the coercivity profiles from FORC diagrams
(Fig. 2j) for the studied samples and compare them with published results for some biogenic
magnetite samples [Roberts et al., 2013], and magnetic inclusion-bearing igneous rocks
[Muxworthy and Evans, 2013; Usui et al., 2015]. The Bc profile for sample “MD01-2421-7-
110” is broader and extends to larger fields compared to other samples, while profiles for
other samples containing magnetic nano-inclusions appear to be similar to those for biogenic
magnetite.
TEM analysis of magnetic extracts reveals abundant detrital magnetic particles with
variable grain sizes, which are probably from igneous lithic fragments sourced from Japan.
Large particles are too thick for electron transmission and appear dark under bright field
TEM observations. We selected thinner edge areas from large particles and small particles for
our detailed TEM and TEM-EDS analysis. This approach does not enable observations of
nanoparticles that occur deeper within large silicate grains. TEM observations reveal
abundant nano-sized magnetic mineral inclusions (Fig. 3) that were difficult to be observed
from scanning electron microscope observations. The nano-sized magnetic minerals must be
embedded within host minerals, rather than being attached to particle surfaces, which were
clearly visualized by HAADF-STEM imaging due to their different chemical contrast to host
minerals (Fig. 3a-d). The nanoparticle inclusions could also be observed by bright-field TEM
a clear contrast shift under TEM between inclusions and hosts from the edge to interior due to
variable inclusion depths within the respective host particles. Observed nanoparticle
inclusions have sizes that range from a few nm to several hundred nm with variable
clusters (Fig. 3a-c, h), dendrites (Fig. 3d-f), and crystallographically oriented nanoparticles
(Fig. 3g). The nanoparticle clusters consist of euhedral octahedral, sub-rounded, and
irregularly shaped crystals (Fig. 3a-c, h). Many of the nanoparticles are nearly isotropic or are
slightly elongated, some of which are closely packed (Fig. 3b, h). We observed less abundant
large magnetic mineral inclusions (i.e. ~1 m; Fig. 3i). The observed dendrites have complex
microstructures with variable one-, two-, and three-dimensional structures (Fig. 3d-f). Some
in Fig. 4a, e, f). It is possible that some of the oriented nanoparticles represent arrested
distributed within host crystals (Fig. 4h, i). We carried out selected area electron diffraction
(SAED) analysis with the TEM stage tilted at different angles (Fig. 4a-d) to determine the
nanoparticles (i.e., appearance (Fig. 4a-c, black) and extinction (Fig. 4d) of nanoparticles)
inclusions within the host crystal, which is further demonstrated by spot-like SAED patterns
(insets in Fig. 4a-c). In contrast, some nanoparticle clusters have ring-like diffraction patterns
(inset in Fig. 4g), which indicate a more random distribution of inclusion orientations.
3.2.2. High-resolution TEM (HRTEM) analysis of magnetic mineral inclusions and host
minerals
nanoparticle inclusions, including clusters and dendrites, have clear lattice fringes (Fig. 5d,
h, j, l, n, p) and strong diffraction patterns (inset in Fig. 5b), which indicate good crystallinity.
The observed d-spacing values and diffraction patterns for the inclusions match well the
crystal structure (Fd3m space group) of magnetite and titanomagnetite. The observed lattice
fringes from magnetic mineral inclusions do not reveal signs of crystal defects (Fig. 5d, 5h,
5j, 5l, 5n, 5p, 6c, 6d, 6h). However, we occasionally observed titanomagnetite nanoparticle
inclusions with crystal twinning (arrows in Fig. 6g). HRTEM and SAED analyses indicate
that the host minerals are also crystalline, as evidenced by HRTEM lattice images and SAED
patterns (Fig. 6e, f, i-l). However, the host minerals generally have weaker diffraction and
less clear lattice patterns compared to the inclusions (Fig. 6f, j). This may be attributed to
variable extents of destruction of crystal structures that were observed after a few seconds of
EDS mapping (Fig. 7a-l) and point analyses (Fig. 7m-u) were carried out to determine
the chemical composition of inclusions and host crystals. EDS mapping for one area (Fig.
7a) indicates that the magnetic mineral inclusions are rich in Fe (Fig. 7b) with a much smaller
Ti concentration (Fig. 7c). The host mineral is rich in O (Fig. 7d), and Si (Fig. 7e), and also
contains a small concentration of Al (Fig. 7f) and Ca (data not shown). EDS mapping of
another area (Fig. 7g) indicates similar characteristics, where the inclusions are Fe rich (Fig.
7h), and lacking in Ti (Fig. 7i), and the host contains O, Si, and Al (Fig. 7j-l), and Ca (data
not shown). EDS spectra of host minerals (Fig. 7m, p, s) indicate the presence of Si, O, Al,
and Ca peaks. EDS spectra of inclusions also contain these elements because they are
embedded in host mineral grains, but the EDS spectra of inclusions (Fig. 7n, o, q, r, t, u)
relative intensities of Fe and Ti peaks are variable (ratios are indicated in Fig. 7n, q, r, t, u),
but Ti contents are small. EDS mapping and point analyses, therefore, consistently indicate
that the magnetic mineral inclusions have chemical compositions that are consistent with
those in the magnetite-ulvöspinel solid solution series (mainly Ti-poor titanomagnetite). Most
of the analyzed host silicate minerals (containing O, Si, Al, and Ca; Fig. 7) are plagioclase
feldspar. Occasionally, we observed silicate host minerals with O, Si, Ca, Mg, and Fe peaks
sediment sample “CD143-55705-7-82” from the Oman continental margin reveal the
(Fig. 8a). This microstructure is consistent with that of exsolved magnetite in igneous rocks.
The presence of magnetic mineral inclusions in samples from core CD143-55705 from the
Arabian Sea confirms expectations from magnetic analyses (Fig. 2e, f; Chang et al. [2016a]).
magnetite crystals, as evidenced by apparently intact magnetosome chain structures and well-
defined magnetosome crystal morphologies (Fig. 8b). The TEM results are consistent with a
FORC diagram from the same bulk sediment (Fig. 2c), which has a strong central-ridge
signature [e.g., Egli et al., 2010; Roberts et al., 2012; Chang et al., 2014]. Despite the
hosted titanomagnetite nanoparticles in this sample (Fig. 8c). The magnetic separate sample
“BL37,38,39” from outcrops in New Zealand contains abundant silicate particles. Detailed
distributed within the silicate hosts (Fig. 8d), although possible dendritic titanomagnetite
structures are observed (Fig. 8e). The host silicate minerals within sample “BL37,38,39”
often have rough surfaces, with chemical compositions that are consistent with silicates with
major Si and O peaks, and minor Al, Ca, Na, Fe, or Mg peaks in the EDS spectra.
sediment particle to align with the ambient geomagnetic field, we compare τH and τM for a
range of equivalent particle diameters (the diameter of a sphere that has the same volume as
the ellipsoid under consideration). To illustrate the relationship between the competing
hydrodynamic and magnetic torques, we calculate the aspect-ratio for a sediment particle
with a given effective diameter and magnetite volume percentage at τH = τM (Fig. 9). Our
infeasibly high levels to achieve parity with the magnetic torque). Thus, sediment particles in
this size range will be dominated by magnetic torques and could contribute to a reliable
sedimentary paleomagnetic signal. In contrast, larger sediment particles must have aspect-
suppressed substantially to achieve parity with the magnetic torque), which is again
hydrodynamic torques, which will restrict their ability to record reliably the ambient
geomagnetic field. The shift from dominance of a magnetic to a hydrodynamic torque in our
4. Discussion
sediments. Our detailed TEM observations indicate that magnetic nanoparticle inclusions are
widely present in marine sediments. Magnetic inclusions can even dominate the magnetic
signal (Fig. 2a) [Chang et al., 2016b]. The observed magnetic mineral microstructures have
two main origins due to exsolution and inclusion (see Tarduno et al. [2006] for a discussion).
silicates, which form due to phase separation in an originally homogenous solid solution
during initial cooling of igneous rocks. Inclusions, such as euhedral crystals, in contrast, form
prior to the host silicate minerals and are incorporated into the host mineral during its
is not surprising. Silicate minerals that host magnetic mineral inclusions occur widely in
igneous rocks [e.g., Evans et al., 1968; Evans and Wayman, 1970; Haggerty, 1991; Feinberg
et al., 2006; Wakabayashi et al., 2006], so it is to be expected that such particles will occur as
detrital grains in sedimentary strata. However, magnetite is a mixed valence iron oxide
mineral that is unstable in both oxidizing and reducing sedimentary environments [Roberts,
2015]. In particular, magnetic iron oxide minerals will undergo dissolution during sulfate-
reducing diagenesis that results in significant depletion of these minerals and formation of
iron sulfide minerals [Roberts, 2015]. Unprotected iron oxides, such as coarse-grained
magnetic minerals and fine-grained biogenic magnetite, are prone to rapid dissolution in
sulfate-reducing marine environments [Karlin and Levi, 1983; Canfield and Berner, 1987;
the Japan and Oman margins have undergone extensive sulfidic diagenesis that has removed
much of the magnetite signal [Roberts and Turner, 1993; Rowan and Roberts, 2006; Rowan
et al., 2009; Chang et al., 2016a, b]. In contrast, silicate minerals are relatively stable against
Poulton et al., 2004; Roberts, 2015]. Protection from diagenesis by host silicate crystals will
paleomagnetic and environmental magnetic studies. Therefore, robust and efficient methods
are needed to identify their presence within sediments. However, this is not straightforward
because inclusions are fine-grained (often in the nanometer size range). Also, sediment
samples often contain mixed magnetic mineral assemblages. The most robust way to identify
magnetic inclusions is by direct TEM observations, as has been demonstrated in this study.
But this requires time-consuming sample preparation and analysis, which makes analysis of
large sample sets impossible. We, therefore, explore whether magnetic screening of bulk
sediment samples can provide useful indications about the possible presence of magnetic
mineral inclusions.
Our detailed TEM and magnetic analyses reveal important properties of magnetic
mineral inclusions in sediments that provide clues about their presence. Magnetic mineral
inclusions are fine-grained, and often have SSD-like magnetic properties (Fig. 2). But the
grain size distributions of magnetic inclusions can overlap with those of other type of fine
Some magnetic mineral crystals hosted in silicates occur in clusters or as complex dendrites
Such microstructures differ from those of intact biogenic magnetite chains in sediments,
which often produce a non-interacting uniaxial SSD signature [Egli et al., 2010]. Such
contrasting properties produce detectable rock magnetic signatures that enable discrimination
between these two important types of magnetic minerals in sediments. For example, FORC
diagrams with a SSD component and moderate vertical spread (Fig. 2a, e, g) are a useful
indication of the presence of magnetic mineral inclusions [Lappe et al., 2011; Muxworthy and
Evans, 2013] that contrast with the non-interacting central-ridge FORC signature observed
for biogenic magnetite [e.g., Egli et al., 2010; Roberts et al., 2012; Chang et al., 2014].
However, samples that contain dispersed magnetic nanoparticles in silicates can also give rise
to FORC signatures with weak magnetostatic interactions [e.g., Usui et al., 2015]. The size
distribution of magnetic inclusions is often broad, ranging from just a few nanometers to a
few microns (Fig. 3-8). Such size distributions can be detected magnetically. For example,
component with large dispersion parameter (DP) values (i.e., >~0.3) for the studied sample
from core MD01-2421. This may also be reflected in the FORC coercivity profiles (Fig. 2j).
For example, the Bc profile for samples with magnetic inclusions is broad and extends up to
higher fields, i.e., 200 mT for sample “MD01-2421-7-110” from the North Pacific Ocean.
This may be because titanomagnetite is magnetically harder than pure magnetite. Low-
sediments (Fig. 2). For example, it was demonstrated recently that biogenic and inorganic
magnetite in marine sediment samples have two distinct Tv temperatures clustered at ~100
and 120 K, respectively [Chang et al., 2016a]. Thus, combined magnetic analyses, such as a
inclusions from biogenic magnetite within sediment samples (Fig. 2). Nevertheless, definite
sediment samples often contain mixed magnetic mineral assemblages and also because
similar coercivity distributions and magnetostatic interactions can be observed for both
First, our results demonstrate that silicate-hosted magnetic nanoparticles are an important
[e.g., Roberts et al., 2012]. SSD magnetic minerals are important for paleomagnetic studies
because they are ideal magnetic recorders that can carry stable remanences over long periods
of geological time [Dunlop and Özdemir, 1997]. Second, compared to unprotected magnetic
integrated study of marine sediment core MD01-2421 from the continental margin of Japan
throughout the core, where monsoon events gave rise to an enhanced environmental magnetic
signal from magnetic inclusions that would otherwise have been destroyed by reductive
detrital particles is, therefore, likely to be important for interpreting diagenetically altered
marine sediment records in a wide range of settings. Third, igneous formation of silicate-
hosted magnetic nanoparticles is related to a range of factors, such as oxygen fugacity, cation
content, temperature, and pressure. The transportation pathway of erosional detritus from
important for paleomagnetic studies. Significant questions exist about their potential
paleomagnetic recording capability. For igneous rocks that contain such magnetic inclusions,
how do they acquire a TRM and is the anisotropy of elongated particles important for
nanoparticles affect paleomagnetic recording fidelity [e.g., Feinberg et al., 2006]? Our
these are important questions to address when subjecting such materials to paleomagnetic
simulate the magnetic properties of magnetic mineral inclusions with complex morphologies
and their paleomagnetic recording fidelity can be assessed quantitatively [e.g., Williams et al.,
2010; Muxworthy and Evans, 2013]. Moreover, single plagioclase crystals that contain
magnetic mineral inclusions have been used for absolute paleointensity determinations [e.g.,
Tarduno et al., 2001]. Some of our silicates differ from those documented in prior
paleointensity and paleomagnetic studies of single silicate crystals, particularly in the density
of inclusions; this may be partially due to selection criteria in those studies that excludes
crystals with visible inclusions (at low magnification) that are aimed at avoiding MD
inclusion density in our sediments highlights the continued need to test for the possibility of
interactions by nanoscale imaging [e.g., Feinberg et al., 2006; Bono and Tarduno, 2015],
FORC analyses [e.g., Tarduno and Cottrell, 2005], and the application of paleointensity
selection criteria (the latter can suggest the presence of interactions if natural remanent
The large size of some host silicate particles (ranging from microns to hundreds of microns)
means that hydrodynamic forces will be important during deposition and that large particles
are unlikely to be aligned by a geomagnetic torque (Fig. 9). It is, therefore, to be expected
signals. But how do such particles compare with the particle size distributions of sediments
that are subjected to paleomagnetic investigations? Sandstones are rarely used for
paleomagnetic analysis because even if magnetic particles occur in the finest possible sand
category (very fine sand), they will have sizes of at least 50 µm. Such magnetic particles will
have MD properties that will not enable recording of a stable paleomagnetic signal. Likewise,
rather than by magnetic torques and will not record a stable paleomagnetic signal (Fig. 9). In
contrast, clay-rich sediments (< 2 µm) are often considered ideal for paleomagnetic analysis
because fine particles are more likely to give rise to stable paleomagnetic recording.
However, clay minerals are products of weathering rather than being primary detrital
minerals that have been abraded to ultra-fine sizes, so that much of the clay size fraction in a
sediment will be due to clay minerals. Nevertheless, some part of the clay size fraction of
sediments could represent particles that have been abraded to ultra-fine sizes.
sediments that have been rounded extensively through abrasion in fluvial and other aquatic
systems, and provide a worthwhile end-member for considering particle aspect ratios and the
effects of hydrodynamic versus magnetic torques. Okada et al. [2001] analyzed particle
shapes for atmospherically transported mineral particles from three Chinese arid regions.
These fine silt- to clay-sized detrital mineral particles (0.1 to 6 µm) have irregular shapes as
expected, with aspect-ratios that are size independent and that range from values of 1 to ~3
(shaded region in Fig. 9), with skewed distributions and median aspect-ratios of 1.3 to 1.4.
Virtually 100% of their analyzed mineral particles have aspect ratios <5. The results of Okada
et al. [2001] place useful constraints on the region of Figure 9 that is likely to be meaningful
for paleomagnetic recording of host particles with magnetic nano-inclusions. From the results
of our simple models, it appears that host silicate particles that contain magnetite nano-
equivalent particle diameters below 12 µm for 1% magnetite concentrations (Fig. 9). The
effective diameter of particles that can be aligned by geomagnetic torques will increase for
important for paleomagnetic recording in silt- and clay-sized sediments. It is unlikely that
magnetite-rich host silicates will contribute to the magnetization of medium silts with particle
sizes above ~18-20 µm (Fig. 9). Sediments always contain a distribution of particle sizes.
Size distributions that cross the τH = τM line, which from Fig. 9 is likely to occur in silt- and
clay-sized sediments, will have some capacity for reliable paleomagnetic recording with
considerable partial cancelation due to both particle types. The resulting magnetization will
not be efficient, which is consistent with the low efficiency of sedimentary magnetizations
[e.g., Tauxe et al., 2006; Mitra and Tauxe, 2009; Heslop et al., 2014]. Regardless, the simple
recording. Some silicate inclusions will have two or more preferred crystallographic
orientations of inclusions [e.g., Feinberg et al., 2006] for which variable extents of magnetic
5. Conclusions
shapes), nanoparticle clusters, and dendrites. EDS analysis indicates that the magnetic
mineral inclusions consist of magnetite to titanomagnetite (with low but variable Ti contents),
while the hosts are silicate minerals (mostly plagioclase feldspar and clinopyroxene). Some
magnetic nanoparticles occur with crystallographically preferred orientations within the host
silicates. The inclusion density in some of the silicates isolated here differs from those
documented in rock magnetic and paleomagnetic studies [e.g., Feinberg et al., 2005; Bono
and Tarduno, 2015]; while such particles may have been excluded in prior paleomagnetic and
paleointensity studies by the selection criteria used [e.g., Tarduno et al., 2006], they may be
therefore, protect the embedded mineral inclusions from dissolution. Our results demonstrate
rather than by geomagnetic torques, so that even if large particles may contain ideal SSD
Nevertheless, deposition of smaller silicate particles with magnetic mineral inclusions could
Acknowledgement We are grateful to Felipe Kremer and Frank Brink at the Centre for
Advanced Microscopy, ANU, for helping with TEM analysis, and Penelope King and Andrew
Berry for useful discussions. Adrian Muxworthy and Yoichi Usui are thanked for providing
published FORC data of silicate crystals extracted from igneous rocks. We thank Richard
Harrison and John Tarduno for helpful review comments, and André Revil and an Associate
Editor for efficient editorial handling. The data in this paper can be obtained by contacting
the corresponding author (liao.chang@pku.edu.cn). The data can be found at the RMAG
(http://earthref.org/MAGIC/). This study was supported by the “1000 Talents Plan” program
of China, the National Natural Science Foundation of China (grant 41574060), and the
References
Bono, R. K., and J. A. Tarduno (2015), A stable Ediacaran Earth recorded by single silicate
doi:10.1130/G36247.1.
Canfield, D. E., and R. A. Berner (1987), Dissolution and pyritization of magnetite in anoxic
7037(87)90076-7.
Taphonomy: Releasing the Data Locked in the Fossil Record, Allison P. A., and D. E.
2016GC006344.
Cottrell, R. D., and J. A. Tarduno (1999), Geomagnetic paleointensity derived from single
821X(99)00068-0.
Dunlop, D. J. (1990), Developments in rock magnetism, Repts. Progr. Phys., 53, 707–792.
Dunlop, D. J., and Ö. Özdemir (1997), Rock Magnetism: Fundamentals and Frontiers,
doi:10.1016/j.gloplacha.2013.08.003.
Egli, R., A. P. Chen, M. Winklhofer, K. P. Kodama, and C.-S. Horng (2010), Detection of
Evans, M. E., and M. L. Wayman (1970), Investigation of small magnetic particles by means
Evans, M. E., M. W. McElhinny, and A. C. Gifford (1968), Single domain magnetite and high
513–516.
355–416.
magnetic interactions in minerals, Proc. Natl. Acad. Sci. U.S.A., 99, 16,556–16,561.
Heslop, D. (2007), Are hydrodynamic shape effects important when modelling the formation
doi:10.1111/j.1365-246X.2007.03588.x.
Heslop, D., and A. P. Roberts (2012), Estimation of significance levels and confidence
Q12Z40, doi:10.1029/2012GC004115.
Karlin, R., and S. Levi (1983), Diagenesis of magnetic minerals in recent haemipelagic
Q12Z35, doi:10.1029/2011GC003811.
and calibration of nonheating paleointensity methods: A case study using dusty olivine,
Mitra, R., and L. Tauxe (2009), Full vector model for magnetization in sediments, Earth
ultrafine magnetite inclusions in the Modipe Gabbro, Geochem. Geophys. Geosyst., 14,
921–928, doi:10.1029/2012GC004445.
from the late-Archaean Modipe Gabbro of Botswana, Geochem. Geophys. Geosyst., 14,
2198–2205, doi:10.1002/ggge.20142.
Paleoceanographic change off central Japan since the last 144,000 years based on high-
resolution oxygen and carbon isotope records, Global Planet. Change, 53, 5–20.
Okada, K., J. Heintzenberg, K. Kai, and Y. Qin (2001), Shape of atmospheric mineral
Poulton, S. W., M. D. Krom, and R. Raiswell (2004), A revised scheme for the reactivity of
68, 3703–3715.
Roberts, A. P., and G. M. Turner (1993), Diagenetic formation of ferrimagnetic iron sulphide
minerals in rapidly deposited marine sediments, South Island, New Zealand, Earth
Roberts, A. P., C. R. Pike, and K. L. Verosub (2000), First-order reversal curve diagrams: A
new tool for characterizing the magnetic properties of natural samples, J. Geophys.
Roberts, A. P., L. Chang, D. Heslop, F. Florindo, and J. C. Larrasoaña (2012), Searching for
doi:10.1029/2012JB009412.
Roberts, A. P., F. Florindo, L. Chang, D. Heslop, L. Jovane, and J.C. Larrasoaña (2013),
Roberts, A. P., D. Heslop, X. Zhao, and C. R. Pike (2014), Understanding fine magnetic
particle systems through use of first-order reversal curve diagrams, Rev. Geophys., 52,
557–602, doi:10.1002/2014RG000462.
magnetizations in Neogene marine sediments from New Zealand, Earth Planet. Sci.
(2015), Rock-magnetic properties of single zircon crystals sampled from the Tanzawa
tonalitic pluton, central Japan, Earth Planets Space, 67, 150, doi:10.1186/s40623-015-
0317-9.
Tarduno, J. A., and R. D. Cottrell (2005), Dipole strength and variation of the time-averaged
Tarduno, J. A., R. D. Cottrell, and A. V. Smirnov (2001), High geomagnetic field intensity
silicate crystals: Recording geomagnetic field strength during mixed polarity intervals,
doi:10.1029/2005RG000189.
Liu, D. G.Sibeck, L. P. Neukirch, and Y. Usui (2010), Geodynamo, solar wind, and
Toward an improved theoretical and experimental foundation, Earth Planet. Sci. Lett.,
244, 515–529.
North Pacific, during the last 145,000 years, Global Planet. Change, 53, 21–28.
Usui, Y., T. Shibuya, Y. Sawaki, and T. Komiya (2015), Rock magnetism of tiny exsolved
significant mineral grains with complex morphology, Geochem. Geophys. Geosyst., 11,
Q02Z14, doi:10.1029/2009GC002828.
Zhao, X., D. Heslop, and A. P. Roberts (2015), A protocol for variable resolution first-order
doi:10.1002/2014GC005680.
Figure 1 Locations of marine sediment core MD01-2421 from the north Pacific off the
east coast of Japan, core CD143-055705 on the continental margin of Oman, core
RR0603-03JC from the eastern equatorial Pacific Ocean, and marine sediment outcrop
from the Lower Awatere Valley, northeastern South Island, New Zealand.
d, f, h) for: (a, b) sample “MD01-2421-7-110” from the North Pacific Ocean, (c, d)
sample “RR0603-03JC-2-60” from the eastern equatorial Pacific Ocean, (e, f) sample
“CD143-55705-7-82” from the Oman margin, Arabian Sea, and (g, h) samples
“BR49D” and (i) “BL37,38,39” from the Lower Awatere Valley, New Zealand. (j)
Published data are shown in (j) for several marine sediment samples with biogenic
Roberts et al. [2013]), and igneous rocks containing magnetic inclusions (a handpicked
sample “B4HP” containing pure pyroxene crystals (Figure 3d; Muxworthy and Evans
[2013]) and a handpicked sample containing six plagioclase crystals (Figure 4b; Usui et
al. [2015]). “BL37,38,39” is a magnetic separate from Roberts and Turner [1993],
while all other studied samples are bulk marine sediments. Note that the studied
(see text for discussion). FORC diagrams in (c, e, g) were measured with variable field
steps following the protocol of Zhao et al. [2015]. All FORC diagrams were processed
using the algorithm of Zhao et al. [2015]. Thicker black lines correspond to the 0.05
significance level [Heslop and Roberts, 2012]. Dashed black lines in the FORC
diffraction patterns (a-g) indicate that some nanoparticle inclusions have a prefered
crystallographic orientation (double headed arrows) within silicate host crystals, while
(h, i) other nanoparticles appear to be more randomly oriented (see text for discussion).
host minerals for two areas (a-f, and g-l) for sample “MD01-2421-7-110”. For one area
minerals, respectively, for areas indicated in (a). The SAED pattern in (b) is from the
whole area in (a). For another area (g-l), images in (h) and (i-l) correspond to magnetic
mineral inclusions and host minerals, respectively, for areas indicated in (g). Clear
lattice fringes for the inclusions and host minerals are observed. Arrows in (g) indicate
images and (b-f, h-l) corresponding elemental maps of two areas within silicate crystals
with magnetic nanoparticle inclusions, and (m-u) EDS spectra for magnetic
nanoparticle inclusions and their host minerals for three analyzed areas in sample
“MD01-2421-7-110”. The nanoparticle inclusions are rich in Fe, but only have small Ti
concentrations. The Ti map in (i) is not as clear as that in (c), which appears to be due
to a low Ti content. The EDS spectra of the host mineral contain mainly Si and O, with
smaller concentrations of Al and Ca. The host mineral grains (a, g) are rich in O, and Si
also contain a small Ti peak. The (*) symbol indicates Cu peaks, which originate from
the TEM grid and are present in all spectra. Fe/Ti ratios for titanomagnetite inclusions
margin, Arabian Sea. The host mineral is rich in O and Si. The exsolved acicular
(double headed arrow). (b) A bright-field TEM image of biogenic magnetite crystals
and (c) a STEM image of titanomagnetite nanoparticle inclusions hosted in silicates for
core RR0603-03JC. (d) A bright-field TEM and (e, f) STEM images of magnetic
from the Lower Awatere Valley, New Zealand [Roberts and Turner, 1993]. Small black
holes in (f) are ablation pits left after EDS point analyses.
silicates. Estimated aspect ratios are shown at which τH = τM for host particles that
torques. In contrast, larger particles require aspect ratios close to 1, which indicates that
(see methods for details of the numerical simulations). The dashed lines denote
boundaries between very fine, fine, and medium silt. Schematic illustration of prolate
ellipsoids alongside the calculated curves, which represent modeled silicate particles,
highlights the aspect-ratio range of 1-3 in which most detrital particles are expected to
fall [Okada et al., 2001]. The dark gray area indicates silicates with a volumetric
magnetite content of 10%. According to the numerical model, particles in this region
should not able to acquire a significant DRM. Arrows indicate the trends of the