ArticlePDF Available

Quantifying the concentration of ferrimagnetic particles in sediments using rock magnetic methods

Wiley
Geochemistry, Geophysics, Geosystems
Authors:

Abstract and Figures

We have developed a quantification method that uses mainly room temperature rock magnetic measurements to calculate concentrations of ferrimagnetic particles in sediments. Our method uses saturation magnetization (Ms) as a total ferrimagnetic concentration proxy, the saturation remanence ratio (Mrs/Ms) as a magnetic grain-size proxy, the anhysteretic remanence ratio (χa/Mrs) to estimate inter-particle magnetostatic interactions, and the normalized susceptibility of the ferrimagnetic fraction (χf/Ms) to calculate the proportion of ultrafine, superparamagnetic particles. This approach eliminates the effect of dilution of the magnetic properties by weakly magnetic matter, and allows the calculation of direct concentrations (or fluxes for dated sedimentary profiles) of constituent ferrimagnetic components. We test our method on a short sediment core from an urban Minnesota lake, for which we calculate ferrimagnetic fluxes of four magnetic components, and compare their pre- and post-European settlement values. Our quantification technique can be applied for reconstructing past environmental changes in a range of sedimentary environments, and is particularly useful for large sets of samples, where detailed magnetic unmixing methods are unfeasible due to time or instrument constraints.
Content may be subject to copyright.
Article
Volume 11, Number 8
4 August 2010
Q08Z19, doi:10.1029/2010GC003182
ISSN: 15252027
Quantifying the concentration of ferrimagnetic particles
in sediments using rock magnetic methods
Ioan Lascu, Subir K. Banerjee, and Thelma S. Berquó
Institute for Rock Magnetism, Department of Geology and Geophysics, University of Minnesota,
Twin Cities, 310 Pillsbury Drive SE, Minneapolis, Minnesota 55455, USA (lascu003@umn.edu)
[1]We have developed a quantification method that uses mainly room temperature rock magnetic measure-
ments to calculate concentrations of ferrimagnetic particles in sediments. Our method uses saturation mag-
netization (M
s
) as a total ferrimagnetic concentration proxy, the saturation remanence ratio (M
rs
/M
s
)asa
magnetic grainsize proxy, the anhysteretic remanence ratio (c
a
/M
rs
)toestimateinterparticle magneto-
static interactions, and the normalized susceptibility of the ferrimagnetic fraction (c
f
/M
s
) to calculate the
proportion of ultrafine, superparamagnetic particles. This approach eliminates the effect of dilution of
the magnetic properties by weakly magnetic matter, and allows the calculation of direct concentrations
(or fluxes for dated sedimentary profiles) of constituent ferrimagnetic components. We test our method
on a short sediment core from an urban Minnesota lake, for which we calculate ferrimagnetic fluxes of four
magnetic components, and compare their preand postEuropean settlement values. Our quantification
technique can be applied for reconstructing past environmental changes in a range of sedimentary environ-
ments, and is particularly useful for large sets of samples, where detailed magnetic unmixing methods are
unfeasible due to time or instrument constraints.
Components: 13,700 words, 7 figures, 2 tables.
Keywords: quantification; rock magnetism; sediments; magnetic unmixing; magnetostatic interactions.
Index Terms: 1512 Geomagnetism and Paleomagnetism: Environmental magnetism; 1540 Geomagnetism and
Paleomagnetism: Rock and mineral magnetism; 1594 Geomagnetism and Paleomagnetism: Instruments and techniques.
Received 26 April 2010; Revised 1 June 2010; Accepted 9 June 2010; Published 4 August 2010.
Lascu, I., S. K. Banerjee, and T. S. Berquó (2010), Quantifying the concentration of ferrimagnetic particles in sediments using
rock magnetic methods, Geochem. Geophys. Geosyst.,11, Q08Z19, doi:10.1029/2010GC003182.
Theme: Magnetism From Atomic to Planetary Scales: Physical Principles and
Interdisciplinary Applications in Geoscience
Guest Editors: J. Feinberg, F. Florindo, B. Moskowitz, and A. P. Roberts
1. Introduction
[2]Ironbearing minerals with characteristic mag-
netic properties are ubiquitous in sedimentary
environments, and are used in environmental
magnetism studies as indicators of past climatic
conditions [e.g., Geiss and Banerjee, 1997, 1999;
Dearing, 1999a; Paasche et al., 2004; Zillén and
Snowball, 2009; HaltiaHovi et al., 2010], soil
and vegetation development [e.g., Oldfield et al.,
2003; Geiss et al., 2003, 2008], erosion [e.g.,
Rosenbaum et al., 1996; van der Post et al., 1997;
Copyright 2010 by the American Geophysical Union 1 of 22
Reynolds et al., 2004; Rosenbaum and Reynolds,
2004a, 2004b], water and sediment (bio)geochem-
istry [e.g., Snowball, 1994; Snowball et al., 2002;
Kim et al., 2005], and diagenetic conditions [e.g.,
Peck and King, 1996; Dearing et al., 1998; Gibbs
Eggar et al., 1999; Snowball et al., 1999; Demory
et al., 2005; Ortega et al., 2006; Larrasoaña et al.,
2007]. Natural samples contain mixtures of mag-
netic phases of different origins, mineralogical
compositions, and grain sizes. The most important
magnetic carriers are ferrimagnetic minerals, such
as magnetite (Fe
3
O
4
), maghemite (gFe
2
O
3
), and
greigite (Fe
3
S
4
). Antiferromagnetic minerals like
hematite (aFe
2
O
3
) and goethite (aFeO[OH])
have weak ferromagnetism, and contribute signifi-
cantly to the total magnetization of sediments only
if the total antiferromagnetic material represents at
least 90% (by mass) of the ordered magnetic phases
[Frank and Nowaczyk, 2008]. Bulk magnetic
properties such as mass susceptibility (c), isother-
mal remanent magnetization (IRM), anhysteretic
remanent magnetization (ARM), saturation mag-
netization (M
s
), saturation remanence (M
rs
), coer-
civity (H
c
), remanent coercivity (H
cr
), and a
number of their interparametric ratios are used to
characterize magnetic mineral assemblages in terms
of concentration, mineralogy, and grain size.
Magnetic mineral assemblage characteristics are
used in turn to make inferences about the processes
that lead to the accumulation of these minerals in
the depositional environment [Evans and Heller,
2003].
[3]Magnetic components (ensemble of particles
with common origin, biogeochemical history, and a
characteristic array of magnetic properties [Egli,
2004]) have been successfully modeled using
magnetization spectrum methods, which exploit the
subtleties of entire magnetization curves, reducing
the nonuniqueness of a bulk parameter approach
[e.g., Robertson and France, 1994; Roberts et al.,
2000; CarterStiglitz et al., 2001; Kruiver et al.,
2001; Heslop et al., 2002; Egli, 2003, 2004,
2006a; Egli et al., 2010]. However, the relatively
involved methodology inherent to spectral techni-
ques poses the disadvantage of only being able to
process a limited number of samples in a given
time span. On the other hand, bulk magnetic
measurements are fast, inexpensive, nondestructive,
and extremely sensitive to very small quantities of
magnetic material, but they sometimes lack the
ability to distinguish between different sedimentary
magnetic components. For the analysis of large
numbers of samples, such as typically encountered
in long core studies, a technique that makes use of a
limited number of spectral analyses on selected
samples to calibrate numerous bulk measurements
is particularly desirable.
[4]Several attempts have been made to quantify
the concentration of various magnetic components
using bulk parameters like c, ARM, or IRM. Lees
[1997] has quantified the concentration of magnetic
components in synthetic mixtures of up to six
sedimentary sources by using linear modeling
techniques, and has found cto be the most reliable
concentrationdependent parameter in case of
complex mixtures. Large errors were encountered
in the deconvolution of mixtures with a large
number of components, implying that linearity
does not apply for complex source mixing due to
the extreme heterogeneity of considered samples,
in addition to effects of magnetic viscosity of ma-
terials, interactions between particles, and inter
instrument calibration issues. In an attempt to
unmix bulk cmeasurements, von Dobeneck [1998]
introduced the concept of partial susceptibilities,
which states that the bulk susceptibility of a sample
is the sum of the susceptibilities of its constituent
components. The method was subsequently devel-
oped in studies by Frederichs et al. [1999], Xie et
al. [1999], Bleil and von Dobeneck [2004], Funk
et al. [2004], and Xie et al. [2009]. The contribu-
tion of individual components to the bulk suscep-
tibility is calculated via multiple regression,
whereby each partial susceptibility term is related
by a proportionality factor to the value of a mag-
netic parameter that is diagnostic for that particular
component (e.g., ARM for finegrained particles).
The partial susceptibility method assumes linear
proportionality between the susceptibility of the
individual component and the magnetic parameter
used as diagnostic, an assumption that does not
always hold true. The method offers the advantage
of reducing everything to susceptibilities, and the
potential to eliminate the diagnostic parameter mea-
surements if an empirical proportionality factor is
universally found for every component considered
in the model.
[5]The property of linear additivity of individual
components that contribute toward a remanence
parameter was used by Dunlop [2002] in devel-
oping theoretical mixing curves of simple two
component numerical mixtures, based on hysteresis
properties. Linear additivity of remanence was
applied by Geiss and Zanner [2006] to understand
pedogenic enhancement in loessic soils in North
American Great Plains. Bulk ARM and IRM
measurements were used to quantify magnetite
formed as a result of biogeochemical processes
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
2of22
occurring in the topsoil. The pedogenic fraction
obtained via this binary mixing model was found to
underestimate calculations based on hysteresis
parameter modeling, and coercivity spectra de-
convolution. One reason for this discrepancy is
their ARMIRM model ignores the effects of
magnetostatic interactions between particles in the
sediment matrix, which tend to lower ARM values
even for low ferrimagnetic concentrations [Egli,
2006b]. The decrease of ARM values as a result
of interparticle interactions was investigated by
Yamazaki [2008], who uses firstorder reversal
curve (FORC) diagrams in addition to bulk ARM
and IRM measurements, and coercivity analysis of
IRM acquisition curves to unmix sedimentary
components in North Pacific sediment cores. The
FORC analysis reveals the presence of a fine
grained noninteracting component, a finegrained
interacting component, and a coarsegrained
background component. The concentration of the
interacting particles is inversely related to the ARM
ratio (ARM normalized by saturation IRM), indi-
cating that interactions between particles are an
important ARMcontrolling factor. This is highly
significant because the ARM ratio is widely used as
a grainsize indicator in enviromagnetic studies.
Since ARM is sensitive to magnetostatic interac-
tions between proximal particles, the ratio can be
artificially lowered if small particles occur in
clumps, as is the case of collapsed bacterial mag-
netite chains, or clusters of pedogenic magnetite
washed into a basin. Therefore, smallscale varia-
tions in the ARM ratio could point to interactions
between particles rather than grain size changes,
potentially offering information about the packing
and arrangement of magnetic particles in the sedi-
ment matrix [Egli, 2006b; Kopp et al., 2006;
Yamazaki, 2008].
[6]In this paper we present a quantification method
that uses bulk rock magnetic measurements to
calculate mass fractions, mass per volume concen-
trations, and mass fluxes of ferrimagnetic particles
in sediments. Detailed rock magnetic and non
magnetic measurements are performed on represen-
tative samples to calibrate the modeling parameters.
The quantification method consists of three parts:
1) Calculation of total ferrimagnetic concentration
from infield magnetization parameters, 2) Model-
ing the remanencecarrying fractions using rema-
nent magnetizations to gain information about
magnetic grain size and interparticle interactions;
this is tested on a selection of synthetic mixtures
of stable single domain (SD), pseudo single
domain (PSD), and multidomain (MD) magnetite
[CarterStiglitz et al., 2001; Dunlop and Carter
Stiglitz, 2006], and 3) Calculation of the super-
paramagnetic (SP) fraction from ferrimagnetic
susceptibility and SP grain size distribution data.
The quantification method is applied to a short
sediment core from a Minnesota urban lake, but is
applicable to a range of environments, including
marine sediments, loess deposits, and soils.
2. Theoretical Considerations
2.1. Total Concentration of Ferrimagnetic
Material
[7]Magnetic susceptibility, remanent magnetiza-
tions (ARM and IRM), and saturation magnetiza-
tion are generally used to qualitatively assess the
amount of magnetic material present in sediments.
Saturation magnetization is an intrinsic property of
magnetic materials, so should theoretically offer
the most reliable concentration estimates. Suscep-
tibilitybased concentration estimates are wide-
spread due to low cost and ease of measurement,
but require information about magnetic mineral-
ogy, grain size, and grain shape. Remanence
measurements are accurate as concentration proxies
only when magnetic composition is uniform (with
respect to mineralogy and grain size) over the
sedimentary interval investigated, because of the
varying degree of remanence acquisition efficiency
for different categories of magnetic particles. In this
section we focus on cand M
s
, and discuss their use
as ferrimagnetic concentration proxies.
2.1.1. Magnetic Susceptibility as a
Ferrimagnetic Concentration Proxy
[8]Lowfield magnetic susceptibility (c
lf
) is the
most common magnetic property measured on
sediments, popularized by the use of multisensor
core loggers equipped with loop or point suscep-
tibility meters [Nowaczyk, 2001]. Largescale dril-
ling expeditions employ the use of core loggers to
measure unsplit sections of core immediately after
retrieval for initial characterization of the sedi-
ments. At this stage c
lf
is mainly used to strati-
graphically correlate between parallel cores from
the same basin, and to roughly estimate the con-
centration of magnetic minerals in the core sec-
tions. c
lf
is strongly controlled by the presence of
ferrimagnetic minerals, but this signature can be
diluted by contributions of diamagnetic, paramag-
netic, and antiferromagnetic materials [e.g.,
Dearing, 1999b]. Bulk c
lf
can be expressed as the
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
3of22
sum of the magnetic contributions of all compo-
nents to the total susceptibility signal:
lf ¼X
n
i¼1
ifi;ð1Þ
where c
i
is the susceptibility of component i, and f
i
is the mass fraction of component i (Sf
i
= 1). This
expression equates to the partial susceptibilities
concept of von Dobeneck [1998], but expressed in
terms of fractions of pure substance susceptibilities.
Since the ferrimagnetic component exerts the major
control on c
lf
, its fraction (f
ferri
) is of interest for
calculating mass or volume based concentration
parameters. In order to determine f
ferri
, the contri-
bution of diluting sedimentary matrix must be
subtracted from the bulk value. Diamagnetic and
paramagnetic substances have weak suscept-
ibilities, but can amount to significant portions of
bulk c
lf
if they are present in abundance, as in the
case of waterladen, organic, carbonaterich, or
siliciclastic sediments [Rochette, 1987]. Suscepti-
bility values for these substances are known from
experimental work (see compilations by Hunt et al.
[1995a] and Dearing [1999b]), and have average
values of ∼−9·10
9
m
3
/kg for water and organic
matter, ∼−8·10
9
m
3
/kg for calcite, ∼−6·10
9
m
3
/kg
for feldspar and quartz, 0.17·10
6
m
3
/kg for
paramagnetic and imperfect antiferromagnetic mi-
nerals. By comparison, ferrimagnetic mineral sus-
ceptibilities are on the order of 10
4
10
3
m
3
/kg.
Mass fractions f
i
can be obtained from geo-
chemical or mineralogical methods (e.g., loss on
ignition (LOI), Xray diffraction, etc.). After all
nonferrimagnetic contributions are determined and
subtracted from c
lf
, equation (1) is reduced to:
f¼fferriferri ;ð2Þ
where c
f
is the susceptibility corrected for non
ferrimagnetic contributions, and c
ferri
is the sus-
ceptibility of the ferrimagnetic component. Average
c
ferri
values are 0.21.2 · 10
3
m
3
/kg for magnetite,
0.0250.29 · 10
3
m
3
/kg for titanomagnetite, 0.3
0.5 · 10
3
m
3
/kg for maghemite, and 0.0260.194 ·
10
3
m
3
/kg for greigite [Hunt et al., 1995a;
Dearing, 1999b; Peters and Dekkers, 2003]. For
SP grains these values are expected to be one order
of magnitude higher [Dunlop and Özdemir, 1997],
because they have very high intrinsic suscept-
ibilities, and are thermally unstable, thus being
easily and efficiently magnetized even in small
magnetic fields. They will dominate the c
ferri
signal
even when in low concentrations. The main chal-
lenge posed by the exercise of determining f
ferri
from c
f
is finding an appropriate value for c
ferri
,
which implies some a priori knowledge about the
ferrimagnetic mineralogy, grain size, and grain
shape. Alternatively, if f
ferri
can be determined from
measurements of saturation magnetization, c
ferri
can be calculated from equation (2). This scenario
is discussed below.
2.1.2. Saturation Magnetization as a
Ferrimagnetic Concentration Proxy
[9]Saturation magnetization M
s
is an intrinsic
property of a magnetic mineral, and provides a
more straightforward measure of f
ferri
than mag-
netic susceptibility. The only a priori information
necessary for f
ferri
calculation is knowledge of the
ferrimagnetic mineralogy. Ferrimagnetic mass
fraction is calculated as:
fferri ¼Ms
ferri
;ð3Þ
where bulk M
s
is obtained from hysteresis
measurements after correcting for high field non
ferrimagnetic contributions, and m
ferri
is the ferri-
magnetic saturation magnetization, which can be
written as a linear combination of the saturation
magnetizations of the ferrimagnetic constituents:
ferri ¼X
m
j¼1
pjj;ð4Þ
where p
j
is the fraction of total ferrimagnetic mass
represented by component j (Sp
j
= 1) and m
j
is its
intrinsic massnormalized saturation magnetiza-
tion. Ferrimagnetic mineral composition can be
determined either from magnetic or nonmagnetic
methods. Magnetic methods are based mainly on
thermal orderdisorder transitions (Curie or Néel
points) or spin reorientations and are semiquanti-
tative, while nonmagnetic methods, such as
Mössbauer spectroscopy and Xray diffraction on
magnetic extracts, permit a direct calculation of the
ferrimagnetic mineral fractions. The main ferri-
magnetic minerals found in lake and marine
sediments are magnetite (m
Mt
=92Am
2
/kg),
maghemite (m
Mh
=74.3 Am
2
/kg) [Dunlop and
Özdemir, 1997], and greigite (m
Gr
=59 Am
2
/kg)
[Chang et al., 2008]. Magnetite and maghemite are
by far the most common, so in the absence of
reliable methods of ferrimagnetic mineral determi-
nation, calculating concentrations both as magne-
tite and maghemite will set the concentration
interval boundaries, assuring accuracy of results,
although at the expense of reduced precision.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
4of22
[10]The slope correction applied to raw hysteresis
loop measurements also offers an alternative way
to determine the contribution of nonferrimagnetic
components to c
lf
, bypassing the need for sediment
composition information. The high field suscepti-
bility c
hf
is calculated from the slope of a loop at
magnetic fields larger than the saturating field,
which is typically a few hundred mT for important
ferrimagnets, and is equivalent to the sum of partial
susceptibilities of all nonferrimagnetic compo-
nents contributing of c
lf
. The ferrimagnetic con-
tribution to c
lf
can thus be expressed as:
f¼lf hf :ð5Þ
Having calculated f
ferri
and c
f
(from equations 35),
c
ferri
can now be obtained from equation (2). This
is done in order to calculate the SP particle fraction
contributing to the total ferrimagnetic concentration
(see section 2.3).
2.1.3. Expression of Total Ferrimagnetic
Concentration
[11]The fraction of ferrimagnetic material is in
effect a mass concentration, i.e., the mass ferri-
magnetic component of total sediment mass. Since
ferrimagnetic concentrations are on the order of
thousandths of the total sediment mass, it is more
convenient to express f
ferri
as a part per thousand
(ppt, mg/g, g/kg, etc.) concentration:
cferri ¼103fferri:ð6Þ
The concentration of ferrimagnetic minerals can
also be expressed as mass ferrimagnetic material
per unit volume sediment:
Cferri ¼fferri wet fsed ;ð7Þ
where C
ferri
is the total ferrimagnetic concentration
as mass per volume unit, r
wet
is the wet sediment
density, and f
sed
is the mass fraction sediment (i.e.,
the ratio of dry mass to wet mass). Wet sediment
density is a standard parameter measured during
initial core logging, but can also be calculated on
samples of known volume by dividing the wet
mass of a sample by its volume. If sample volume
is unknown, r
wet
can be obtained from composi-
tional analysis (e.g., LOI) as a sum of pure sub-
stance densities (e.g., 1000 kg/m
3
for water)
weighed by their respective mass fractions. C
ferri
is
of interest for calculating ferrimagnetic particle
fluxes in sediment cores with established chronol-
ogies. The mass flux of ferrimagnetic material F
ferri
to the sediment column is obtained from multiplying
C
ferri
by the sediment accumulation rate R
sed
(expressed in length units per time):
ferri ¼CferriRsed :ð8Þ
In order to separate the fluxes of various ferri-
magnetic components, an unmixing method is
necessary.
2.2. Concentration of RemanenceCarrying
Ferrimagnetic Components
[12]In this section we present a threecomponent
mixing model that makes use of bulk remanence
measurements to quantify the remanencecarrying
fractions of f
ferri
. Unmixing models using either
hysteresis or bulk remanence parameters have been
used to analyze synthetic, natural, and numerical
mixtures of two components, and are based on the
linear additivity of remanence properties when the
mixture is monomineralic [Dunlop, 2002; Dunlop
and CarterStiglitz, 2006; Geiss and Zanner,
2006]. The mixing model presented here com-
bines the remanence aspect of the hysteresis
approach of Dunlop [2002] and Dunlop and
CarterStiglitz [2006] with the ARMIRM model-
ing of Geiss and Zanner [2006]. However, we use
the ARM ratio primarily to evaluate particle inter-
actions rather than grain size, essentially extending
twocomponent models to mixtures of three end
members, here MD (including PSD), uniaxial non
interacting SD (UNISD) [Egli et al., 2010], and
interacting SD (ISD). Our model accounts for all
the remanencecarrying contributors to f
ferri
, i.e.,
the nonSP ferrimagnetic fraction.
[13]Bulk ARM and saturation IRM (or M
rs
) values
can be written as linear combinations of the ARM
and M
rs
values of the mixture endmembers. Since
the absolute remanence values of endmember
components are not invariant, they are normalized
by M
s
(m
ferri
is constant across the mixture) in order
to obtain comparable ratios. The resulting system
of equations for the threecomponent (MD,
UNISD, and ISD) mixing model is:
Mrs
Ms

bulk
¼X
3
i¼1
gi
Mrs
Ms

i
a
Ms

bulk
¼X
3
i¼1
gi
a
Ms

i
X
3
i¼1
gi¼1
8
>
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
>
:
;ð9Þ
where c
a
is the ARM susceptibility (ARM nor-
malized by acquisition bias field), (M
rs
/M
s
)
bulk
and
(c
a
/M
s
)
bulk
are ratios obtained from measured
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
5of22
parameters, (M
rs
/M
s
)
i
and (c
a
/M
s
)
i
are the respec-
tive ratios of endmember i, and g
i
is the fraction of
component i. The system is solved for g
i
. Since
c
a
/M
s
is not a commonly used ratio, the second
equation in (9) can be divided by the first in order to
obtain the expression for the ARM ratio (c
a
/M
rs
).
The system then becomes:
Mrs
Ms

bulk
¼X
3
i¼1
gi
Mrs
Ms

i
a
Mrs

bulk
¼X
3
i¼1
gi
a
Ms

i
X
3
i¼1
gi
Mrs
Ms

i
X
3
i¼1
gi¼1
8
>
>
>
>
>
>
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
>
>
>
>
>
>
:
:ð10Þ
Note that ARM and M
rs
combine in a nonlinear
fashion. Theoretical values for M
rs
/M
s
and c
a
/M
rs
can be predicted by varying the endmember frac-
tions g
i
over the interval [0, 1] with the third con-
dition of the system always satisfied. The model
ratio values can be plotted in M
rs
/M
s
versus c
a
/M
rs
scatter diagrams at regular g
i
intervals (e.g., 0.1) as
mixing grids,and overlain with the experimental
values for comparison. The mixing grids are
equivalent to ternary diagrams in g
i
coordinates.
The choice of endmember ratios will be discussed
in section 4.
2.3. Concentration of Superparamagnetic
Particles
[14]In order to calculate the absolute concentra-
tions of the remanencecarrying components, the
SP concentration must be first subtracted from the
total ferrimagnetic concentration. Rearranging
equation (2) to solve for c
ferri
, and substituting f
ferri
with M
s
/m
ferri
(equation 3), we find that c
ferri
is
nothing but a multiple of c
f
/M
s
, where the multi-
plier is the ferrimagnetic saturation m
ferri
. The ratio
c
f
/M
s
has been used in a quantitative way to
express fluctuations in high susceptibility SP grains
relative to a nonSP baseline for Chinese loess
plateau deposits [Hunt et al., 1995b]. SP con-
centrations obtained this way were found to be in
good agreement with concentrations calculated via
the more reliable thermal demagnetization of
low temperature remanence method [Hunt and
Banerjee, 1992], and are much easier to perform.
The nonSP susceptibility (c
nonSP
), carried by SD
and MD grains, has values an order of magnitude
lower compared to the typical susceptibility of SP
particles (c
SP
). c
ferri
can be written as the linear
combination of c
SP
and c
nonSP
, an expression
which can be rearranged to solve for the SP fraction
(f
SP
):
fSP ¼ferri nonSP
SP nonSP
:ð11Þ
c
nonSP
is typically characterized by a narrow range
of values (0.40.8 · 10
3
m
3
/kg for magnetite
[Dunlop and Özdemir, 1997]), while c
SP
can be
determined using information about the SP grain
size distribution. For very small particles (<10 nm)
the relationship between c
SP
and particle volume is
linear at room temperature, according to Néel [1949]
theory of thermally activated magnetization. Larger
SP particles (1020 nm) exhibit frequency depen-
dence of susceptibility at room temperature,
because their magnetization relaxation times are
comparable to the observation times of the exper-
iment. Worm [1998] has investigated the nature of
the SPSD transition and the behavior of the
magnetic susceptibility across this boundary. The
relationship between c
SP
and particle volume is no
longer linear, and the transition to SD suscept-
ibilities is gradual, especially if the particles are
characterized by a distribution of coercivities [see
Worm, 1998, Figure 2]. In Worms [1998] approach,
frequencydependent susceptibility measurements
across a range of temperatures are necessary to
estimate the SP particle size distribution, and thus
c
SP
(which for magnetite is found to have values of
up to 8 · 10
3
m
3
/kg). The choice of c
SP
depends
from case to case and will be discussed in context
in Section 4.
3. Materials and Methods
3.1. Samples and Nonmagnetic
Treatments
[15]The synthetic mixtures used for testing our
remanence model (Section 2.2) have been prepared
by CarterStiglitz et al. [2001] by combining an
SD endmember with PSD and MD endmembers
in incremental mass fractions. The SD particles are
freezedried cells of vibroid magnetotactic bacte-
rium MV1, which produces prismatic magnetite
magnetosomes with mean volumes of 0.65 · 10
22
m
3
and aspect ratios of 1.5, aligned in linear chains
(samples from the same batches have also been
used by Moskowitz et al. [1993]). The PSD and
MD phases are Wright Company magnetites 3006
and 041183 with grain sizes of 1.06 ± 0.71 mm and
18.3 ± 12 mm respectively [Yu et al., 2002]. The
nonSD component was first dispersed in CaF
2
using a blender; the MV1 freezedried cells were
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
6of22
then added in the desired mixing proportions, and
gently mixed together using a mortar and pestle
[CarterStiglitz et al., 2001].
[16]The natural samples used in this study are
lacustrine sediments from Brownie Lake, a small
urban water body in Minneapolis, Minnesota.
Brownie Lake has a surface area of 5·10
4
m
2
and
a maximum depth of 14.1 m (5.6% relative depth).
The lake has been meromictic (permanently strati-
fied) since the lake level was artificially lowered by
3 m in 1917 [Swain, 1984; Tracey et al., 1996]. A
139 cm core was retrieved in 2007 from 13 m water
depth, using a driverod surface corer. The core
was logged and described at the University of
Minnesota Limnological Research Center, and
sampled at a resolution of 1 cm. Aliquots of 1cm
3
were used for LOI analysis, according to the pro-
cedure described by Dean [1974] and Heirietal.
[2001]. Separate subsamples were freezedried,
weighed, and packed in plastic boxes for magnetic
analyses. On 15 September 2009 a water column
profile of dissolved oxygen (DO) was recorded
using a Hydrolab multisensor probe. Water sam-
ples were collected using a Van Dorn sampler at a
resolution of 0.51 m across the oxicanoxic
interface (OAI) and 11.5 m below the OAI.
Between 150 and 200 ml of each sample were
filtered in the laboratory using 0.1 mm filters,
which were then frozen and packed in plastic
straws. A surface sediment (top 1015 cm of the
sediment column) sample was collected on the
same date from a water depth of 13.5 m using an
Ekman dredge. A subsample was disaggregated
with sodium hexametaphosphate (0.5 g to 100 ml
sediment slurry) and circulated through an inhouse
magnetic separator. The magnetic fraction was
extracted using a highgradient permanent magnet,
collected on a daily basis for a twoweek period,
and stored in isopropanol at 4°C. The extract was
then dried at 25°C for 24 h, and used for room
temperature
57
Fe Mössbauer spectroscopy. Möss-
bauer analysis was performed with a conventional
constantacceleration spectrometer in transmission
geometry with a source of
57
Co in an Rh matrix.
Hyperfine parameters such as magnetic hyperfine
field (B
HF
), isomer shift (IS) and quadrupole shift
(QS) have been determined by NORMOS program
[Brand, 1987], and aFe at room temperature was
used to calibrate isomer shifts and velocity scale.
3.2. Magnetic Measurements
[17]All magnetic measurements were performed at
the University of Minnesota Institute for Rock
Magnetism. Magnetic susceptibility was measured
on a Kappabridge KLY2 susceptometer operating
at a frequency of 920 Hz. A DTech 2000
demagnetizer was used for ARM acquisition,
which was imparted in a 0.1 mT steady field su-
perimposed on an AF field decaying from a peak
value of 200 mT, with a rate of 5 mT per half cycle.
Stepwise alternating frequency (AF) demagnetiza-
tion of the ARM was performed on the filtered
water samples using an automated 2G Enterprises
superconducting quantum interference device
(SQUID) magnetometer, with peak fields increas-
ing from 0.5 to 170 mT. The demagnetization
procedure contains one hundred measurement steps
spaced on a logarithmic scale. IRM was imparted
on the same samples by pulsing the samples in a
200 mT field using a 2G core pulse magnetizer.
This was done two times in order to remove any
viscous effects. The IRM was demagnetized fol-
lowing the same procedure as for the ARM. For
coercivity component analysis, first derivatives of
the demagnetization curves were fitted with
skewed generalized Gaussian distributions using
MAGMIX [Egli, 2003]. Hysteresis loops were
measured on a Princeton Measurements vibrating
sample magnetometer using a maximum applied
field of 1 T. The loop slopes at high fields (>0.7 T)
were used to correct raw M
s
values, and to calcu-
late the contribution of nonferrimagnetic minerals
to low field susceptibility. A Quantum Design
magnetic properties measurement system (MPMS2)
was used to perform low temperature demagneti-
zation experiments on representative lake sediment
samples. Thermal demagnetization of saturation
IRM (SIRM) acquired at 10 K in a 2.5 T field after
cooling the sample in zero magnetic field (the
socalled ZFC treatment) was measured from 10 K
to room temperature at 5 K intervals. Room tem-
perature SIRM imparted in a 2.5 T field was par-
tially decayed by cooling the sample to 10 K
and rewarming it back to 300 K. Measurements
were performed in zero field at 5 K intervals. The
MPMS was also used to measure susceptibility as a
function of temperature and frequency on the
selected lake samples. AC susceptibility was
measured in 10 K steps from 10 to 300 K at fre-
quencies of 1, 10, and 100 Hz.
4. Results and Discussion
4.1. Synthetic Mixtures
[18]We use the synthetic samples of known end
member proportions to test the validity of our
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
7of22
mixing model defined by system (9). The M
rs
/M
s
and c
a
/M
rs
values of the SDMD and SDPSD
mixtures are plotted in Figure 1a, together with
theoretical values for mixture bulk ratios deter-
mined from varying g
i
in (10). The measured values
fall on the PSDSD and MDSD mixing lines (i.e.,
twocomponent mixtures; g
3
= 0), where the SD
endmember ratio values are M
rs
/M
s
= 0.498, and
c
a
/M
rs
= 1.6 · 10
3
m/A. For comparison,
Moskowitz et al. [1993] measured ARM ratios on
freezedried MV1 cells and obtained values of
1.82.1 · 10
3
m/A, while Kopp et al. [2006] report
ARM ratios of 1.591.79 · 10
3
m/A for intact and
treated (with sodium dodecyl sulfate and/or ultra-
sonication followed by dilution in sucrose) freeze
dried MV1 cells. Moskowitz et al. [1993] have also
performed ARM acquisition experiments on wet
MV1 cell suspensions fixed with 1% glutaralde-
hyde, and obtained ARM ratio values of 3.13.5 ·
10
3
m/A. Fragile bacterial cell membranes must be
partially ruptured or destroyed in the process of
freeze drying, which involves freezing the sample
and extracting the water fraction by sublimation
in vacuum. The partial destruction of the cell
Figure 1. (a) Crossplot of M
rs
/M
s
versus c
a
/M
rs
for MDSD (diamonds) and PSDSD (circles) mixtures. Number
next to each symbol represents fraction SD component added to mixture. Mixing lines (dashed) for theoretical mix-
tures of two endmembers were calculated using a nil fraction for the third component in system (10). The symbols on
the mixing lines represent theoretical values of M
rs
/M
s
and c
a
/M
rs
calculated at fraction increments of 0.1. The
continuous lines are outer contours of mixing grids calculated for three endmembers, where the SD component was
subdivided into UNISD and ISD (see text for discussion of endmember ratios). (b) Ternary diagram of calculated
fractions for each component (UNISD, ISD, and MD or PSD) of the synthetic mixtures, according to the three end
member mixing model (g
i
are solutions to the system of equations). Shaded area is the interval 6075% interacting
component in a ISDUNISD mixture. (c) Inverted fractions of endmembers UNISD, ISD, and MD (left) or PSD
(right) versus fractions SD (top) and nonSD (bottom) added to each mixture.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
8of22
membranes results in magnetosome chain clump-
ing, which explains the lower ARM ratios of
freezedried samples as the effect of interactions
between the disturbed magnetosomes. Transmis-
sion electron microscope (TEM) analysis of the
untreated, freezedried MV1 samples of Kopp et al.
[2006] reveals that only about 10% of the magne-
tosomes are isolated from other crystals and chains,
while the bulk of the chains are either in sideto
side arrangements, or collapsed into loop config-
urations [Kopp et al., 2006, Figure 2a]. Given the
ARM ratio value of 1.6 · 10
3
m/A for the SD end
member component in our mixtures, we conclude
that the SD endmember physically added to the
mixtures is in fact a combination of noninteracting
and interacting particles and chains, which means
the synthetic samples can be modeled as mixtures
of three components.
[19]We first assign values to the remanence ratio
(M
rs
/M
s
) and ARM ratio of the noninteracting SD
endmember. For magnetite, M
rs
/M
s
has a value of
0.5 by definition of randomly oriented UNISD
particles [Stoner and Wohlfarth, 1948]. The deter-
mination of the value for the ARM ratio is not as
clearcut. Egli and Lowrie [2002] have developed
an analytical solution for the ARM of assemblages
of UNISD particles, and have found that the ARM
ratio of UNISD grains characterized by uniform
rotation depends chiefly on grain size and grain
shape, and only weakly on microcoercivity. They
calculate a maximum c
a
/M
rs
value of 3.7 · 10
3
m/A
for prismatic crystals with volumes of 1.252.2 ·
10
22
m
3
, and aspect (length to width) ratios of
1.432.5 [see Egli and Lowrie, 2002, Figure 11c].
Mean volumes of 0.65 · 10
22
m
3
and aspect ratios
of 1.5 (as is the case with our MV1 samples) yield
ARM ratios of 2·10
3
m/A according to the Egli
and Lowrie [2002] model, and yet c
a
/M
rs
values
obtained from wet suspensions of the same batches
of particles used in this study [Moskowitz et al.,
1993] are close to the maximum values predicted
by theory. This discrepancy can be explained by
the fact that Egli and Lowrie [2002] model ran-
domly oriented noninteracting isolated particles,
and not magnetosome chains. The arrangement of
particles in linear chains should have the same
effect as particle elongation, namely to increase
dipole moment and coercivity. The increase in
chain magnetic moment would lead to a higher
chain ARM ratio than that of its isolated magne-
tosomes. Therefore having a UNISD endmember
with a c
a
/M
rs
value of 3.7 · 10
3
m/A is not an
unreasonable assumption.
[20]The ISD endmember ratios (M
rs
/M
s
= 0.497
and c
a
/M
rs
= 0.47 · 10
3
m/A) were determined so
that a) M
rs
/M
s
of the 100% SD mixture would fall
on the ISDUNISD mixing line; b) the correlation
between the (non) SD fraction added to the mixture
and the inverted (non) SD fraction obtained from
the solution to system (9) approaches a 1:1 rela-
tionship (i.e., the slope of a linear fit to the data is
close to 1); and c) Pearsons correlation coefficient
R of the fits for each mixture is maximized (R =
0.999, Figure 1c). The ARM ratio of the ISD end
member is close to the value of 0.41 · 10
3
m/A
measured by Kopp et al. [2006] on ultrasonicated
MV1 cells diluted at 1 ppt in sucrose (their sample
V2a). The ultrasonic treatment destroys the bacte-
rial cell membrane but keeps the magnetosome
organelles intact, while the dilution prevents the
particles from clumping. TEM images of V2a show
collapsed chains throughout, but there are no major
magnetosome clumps, and some of the particles
remain isolated [Kopp et al., 2006, Figure 2b].
According to the calculations of Egli [2006b] this
corresponds to an average spacing of three particle
diameters. The SD component of our mixtures is
therefore a combination of intact chains and iso-
lated magnetosomes (the UNISD endmember),
and collapsed chains of particles that still preserve
their magnetosome membranes (the ISD end
member). The mixing grids between the three end
members are calculated using system (10), but we
plot only the outer contours of these grids (con-
tinuous lines Figure 1a) in order not to unneces-
sarily clutter the diagram. For reference, a similar
mixing grid, calculated at fraction increments of
0.1, is plotted in Figure 5a. We find that the ISD
component in our synthetic samples comprises 60
75% of the total SD grains in the mixtures (shaded
area in Figure 1b), compared to the 90% visual
estimate of Kopp et al. [2006] from TEM analysis
of samples with equivalent ARM ratios.
4.2. Lake Sediments
4.2.1. Sedimentary Units and Chronology
[21]European settlement in the Brownie Lake area
began shortly after 1850. In the core it is expressed
as a change in sedimentary characteristics at a
depth of 65 cm, marking the boundary between the
postsettlement sedimentary succession (Unit 1)
and the presettlement sediments (Unit 2). Unit 1
(065 cm; ca. 1850present) is a laminated clayey
silt with authigenic calcite and diatoms. The lami-
nae are discontinuous, and of variable thickness of
up to 0.5 cm. Organic matter (mostly sapropel) and
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
9of22
carbonate contents each average between 10 and
20% of the dry sedimentary matter. A tan sandy
layer that occurs at 50 cm marks an artificial low-
ering of the water level in 1917, the year a canal
connecting Brownie Lake with Cedar Lake to the
south was completed. Unit 2 (65140 cm; <1850)
is a dark brown, massive, partly humic peat com-
posed of terrestrial mosses and aquatic plants.
Organic matter fluctuates between 30 and 60%,
while carbonate content is constant at 5%. The
chronology of the Brownie Lake core is based on
the correlation with the
210
Pb dated core of Swain
[1984]. The correlation is drawn from organic
matter concentration variations, and is aided by the
recognition of marker horizons such as the 1917
sand lens.
4.2.2. Room Temperature Magnetic Properties
[22]Sedimentmagnetic properties of the Brownie
Lake core are presented in Figure 2, and the mean
(m) and standard deviation (s) of each parameter
are listed for both sedimentary units in Table 1. The
standard deviation was calculated for the detrended
magnetic parameters, in order to remove variability
introduced by longterm shifts in mean. In Table 1
is also computed the coefficient of variation (CV =
s/m) for each magnetic parameter within a sedi-
mentary unit. The concentrationdependent para-
Figure 2. Bulk magnetic properties of the Brownie Lake sediment core: ARM susceptibility c
a
, lowfield suscep-
tibility c
lf
, saturation remanence M
rs
, saturation magnetization M
s
, ARM ratio c
a
/M
rs
, remanence ratio M
rs
/M
s
, and
c
f
/M
s
. The magnetic parameters define two sedimentmagnetic units (separated by dashed line): Unit 1 (1850present)
and Unit 2 (<1850).
Table 1. Mean, Standard Deviation, and Coefficient of
Variation Values for the Magnetic Properties of Brownie
Lake Sediment Units
a
Magnetic Parameter
Unit 1 Unit 2
msCV msCV
c
lf
(10
7
m
3
/kg) 8.95 2.52 0.28 8.96 1.67 0.19
c
a
(10
5
m
3
/kg) 0.71 0.2 0.28 3.32 0.7 0.21
M
rs
(10
3
Am
2
/kg) 15.6 4.8 0.31 21.5 4.4 0.2
M
s
(10
3
Am
2
/kg) 96.7 34.4 0.36 69.7 13.9 0.2
c
a
/M
rs
(10
3
m/A) 0.48 0.15 0.31 1.71 0.18 0.1
M
rs
/M
s
0.16 0.02 0.1 0.31 0.02 0.06
c
f
/M
s
(10
5
m/A) 0.87 0.06 0.07 1.16 0.11 0.09
a
Here, m is the mean, sis the standard deviation, and CV = s/mis
the coefficient of variation. Standard deviations are calculated after
subtracting the effect of any longterm trends on the magnetic
parameters. CV is a unitless measure.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
10 of 22
meters c
lf
,c
a
,M
rs
, and M
s
show higher variability
in Unit 1 compared to Unit 2 (higher CVs), and a
general increase in values with the decrease in
depth in core over the past 150 years. In Unit 2 the
concentrationdependent parameters exhibit a
larger shift in mean to lower values, but the de-
trended variance is lower, indicating a steady
decrease in concentration with time, up to the
European settlement horizon. The concentration
independent magnetic parameters M
rs
/M
s
,c
a
/M
rs
,
and c
f
/M
s
have weak increasing trends with
decreasing depth in Unit 1. In Unit 2 M
rs
/M
s
and
c
a
/M
rs
have weak decreasing trends and lower
variability than in Unit 1 (lower CV values), while
c
f
/M
s
has a weak increasing trend and slightly
higher variability.
4.2.3. Suspended Sediments
[23]ARM and IRM profiles of the water filtrates
are shown in Figure 3a, along with the DO profile.
Oxygen concentrations are close to saturation va-
lues in the epilimnion (7.68 mg/l), but then
decrease to 0.2 mg/l at 4m, with a maximum gra-
dient between 2.5 and 3 m. At the depth of 3 m DO =
1.3 mg/l, and both ARM and IRM values are
lowest in the profile. DO drops to 0.75 mg/l between
3 and 3.5 m, with a simultaneous increase in ARM
and IRM, which reach maximum values at 4m,
immediately below the OAI. Below 4 m oxygen
levels remain low, with a slight increase from 0.2 to
0.35 mg/l in the bottom waters. ARM and IRM are
still elevated at 5 m, but decrease significantly below
5 m. The ARM ratio peaks (2.62.7 · 10
3
m/A) at
Figure 3. Magnetic properties of filtered water samples collected from Brownie Lake. (a) Profiles of dissolved oxy-
gen, ARM, IRM, and ARM ratio. (b) Component analysis of ARM (diamonds) and IRM (squares) demagnetization
curves showing three components defined by median coercivity and distribution width: D (detrital + dissimilatory),
BS (biogenic soft), and BH (biogenic hard). (c) ARM ratio of the bulk filtrates, individual components, and total
biogenic contribution (BS+BH).
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
11 of 22
the OAI (3.54 m), is lowest (<1 · 10
3
m/A)
between 7 and 9 m, and has values of 11.5 · 10
3
m/A
for the rest of the filtrates.
[24]The biogenic contribution to both ARM and
IRM was separated by coercivity component
analysis from first derivatives of ARM and IRM
demagnetization curves, using the fitting model
proposed by Egli [2003, 2004]. Figure 3b shows
three coercivity components present below the
OAI, each defined by two distribution parameters:
median coercivity and dispersion. According to the
definitions of Egli [2004] the low coercivity com-
ponent comprises large detrital grains, and fine
extracellular particles mediated by dissimilatory
microorganisms (D); the intermediate coercivity
component is identified as biogenic soft(BS),
and the high coercivity component as biogenic
hard(BH). The ARM ratio of each component
is plotted in Figure 3c. The two biogenic com-
ponents have maximum ARM ratios at the OAI
(3.4 · 10
3
m/A for BS; 3.82 · 10
3
m/A for BH).
Below the OAI BS decreases to <1 · 10
3
m/A at
8 m (where BH peaks at 3.2 · 10
3
m/A), and
increases again toward the bottom (where BH < 1 ·
10
3
m/A). Integrating BS and BH remanences over
the entire water column profile, we obtain average
ARM ratios of 2.44 · 10
3
m/A for component BS,
and for 2.13 · 10
3
m/A for component BH. The
ARM and IRM values of the two biogenic compo-
nents were respectively added for each sample in
order to calculate an average biogenic ARM ratio
(BS + BH in Figure 3c). At the OAI the ARM ratio
of BS + BH is 3.47 · 10
3
m/A, while its integrated
value over the water column is 2.36 · 10
3
m/A.
4.2.4. Total Ferrimagnetic Concentration
[25]To calculate the total mass concentration of
ferrimagnetic particles in a particular sample using
equation (3), the nature of the ferrimagnetic carrier
must first be identified. The low temperature
magnetization experiments reveal the presence of
partially oxidized magnetite in both sedimentary
units, with more pronounced maghemitization in
Unit 2. Zero field cooled (ZFC) 10 K SIRM
warming curves of typical samples for each sedi-
mentary unit are shown in Figure 4a. The Unit 1
sample demagnetization rate is highest between
10 and 50 K, where 30% of the SIRM is lost,
decreases between 50 and 100 K (10% loss), in-
creases again (20% loss) between 100 and 130 K
through the Verwey transition (T
V
), and is lowest
between 130 and 300 K, interval in which only
10% more is lost. The remanence retained at room
temperature is a little over 30%. By comparison,
the Unit 2 sample memory is almost double. The
initial remanence loss of this sample is steep but the
magnitude of loss (20%) is less than in Unit 1.
This is followed by a decrease in demagnetization
rate with a broad (10% loss over 30K) but still
distinct T
V
, shifted to lower temperatures (80
110 K). Above T
V
the demagnetization rate
remains constant at 7% per 100 K. The depressed
and broadened T
V
, shifted to lower temperature is a
trait of nonstoichiometric magnetite, while rema-
nence loss below and above T
V
have also been
attributed to magnetite oxidation in well charac-
terized starting material [Özdemir et al., 1993]. The
degree of oxidation also dictates the magnitude of
the demagnetization rate below and above T
V
.
The initial remanence loss is thus highest for
surface oxidized magnetite particles, while severely
maghemitized particles demagnetize at a higher
rate above T
V
[Özdemir and Dunlop, 2010].
However, this remanence loss can also occur due to
unblocking of magnetite SP particles during sample
warming. To avoid the effect of superposition of
the two phenomena, we examine the behavior of
room temperature remanence during low tempera-
ture cycling (Figure 4b). Typical Unit 1 curves
show a 1% increase in remanence with cooling
from 300 to 250 K, followed by a loss of 25% to
T
V
, and only a slight increase in remanence at
temperatures lower than 70 K. The cooling curve is
reversible from 10 to 60 K, temperature above
which the remanence recovery is incomplete.
Above T
V
, the warming curve peaks around 170 K,
followed by a steady decrease to room temperature.
The total remanence recovery is above 75%. The
Unit 2 cooling curve shows the same magnitude
increase in remanence between 300 and 250 K, but
is followed by a gradual decrease to 110 K
corresponding to a remanence loss <2%. Below T
V
there is a 2% increase in remanence, which is
reversible upon warming from 10 to 100 K. Above
100 K it steadily decreases to room temperature,
first gradually to 250 K and then more steeply
from 250 to 300 K. Compared to the Unit 1 sample,
only 5% of the room temperature SIRM is lost in
the low temperature cycling process. According to
Özdemir and Dunlop [2010] the humpshaped
cooling and warming curves above T
V
, together
with the high degree of room temperature rema-
nence memory are diagnostic for magnetite in the
advanced stages of maghemitization. Özdemir and
Dunlop [2010] have also proposed a semiquanti-
tative method for estimating the oxidation parameter
z[OReilly and Banerjee, 1967; Readman and
OReilly, 1971], by comparing the room tempera-
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
12 of 22
ture memory loss (DM
c
) with the difference in
magnetization between 20 K and T
V
(DM
m
).
Although this method has not been extensively
tested on a range of particle size fractions (or
mixtures thereof), it can be useful as a firstpass
estimate of the degree of maghemitization
[Özdemir and Dunlop, 2010]. In our case, Unit 2
DM
c
and DM
m
values fall in the range of PSD
magnetite characterized by zvalues of 0.80.9,
whereas in Unit 1 the small DM
m
value, deter-
mined by the predominance of a low coercivity MD
component [Muxworthy et al., 2003], does not
allow a comparison.
[26]The room temperature Mössbauer spectrum of
the surface sediment magnetic extract (Figure 4c)
gives us a quantitative measure of maghemite
concentration for Unit 1 in this case. The fitted
parameters correspond to the sextets of hematite,
maghemite, site A (tetrahedral) and site B (octa-
hedral) of magnetite, and doublets of Fe
2+
and Fe
3+
(Figure 4c and Table 2). The B
HF
of site A and B of
magnetite, maghemite and hematite are in agree-
ment with values from the literature and suggest
iron phases without presence of foreign elements
(e.g., Ti or Al substitutions) in their structure
[Murad and Cashion, 2004]. However, the mag-
netite site A/site B concentration ratio (Table 2) is
higher than 0.5, showings some deviation from a
stoichiometric form. The nonstoichiometry is in-
terpreted as the combined effect of a stoichiometric
magnetite core, a nonstoichiometric magnetite
transition zone (characterized by a distribution
of z), and an outer maghemite shell. The Fe
2+
doublet is associated with paramagnetic minerals
(e.g., iron phyllosilicates), and the Fe
3+
doublets
correspond to paramagnetic iron minerals (wide
Figure 4. (a) Thermal demagnetization of SIRM applied at 10 K after ZFC treatment of representative samples from
the two sedimentary units. (b) Behavior of room temperature SIRM with cooling to 10 K and warming back 300 K for
the same samples. (c)
57
Fe Mössbauer spectrum of magnetic extract from the surface sediment. (d) Time series of total
ferrimagnetic concentration by mass (c
ferri
) on a waterfree basis: thick line represents solution of equation (10) with
m
ferri
=88Am
2
/kg for Unit 1 and 78 Am
2
/kg for Unit 2, thin lines are solutions calculated for magnetite (lower bound)
and maghemite (upper bound) and delimit a solution space (shaded) for c
ferri
considering the entire range of z[0,1].
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
13 of 22
Fe
3+
doublet), as well as to superparamagnetic
iron oxides (narrow Fe
3+
doublet). If we consider
only the magnetically ordered ferrimagnetic par-
ticles, then 91% of the iron is in contained in non
stoichiometric magnetite, and 9% is in maghemite.
If we assume the SP fraction is entirely ferrimagnetic
(i.e., no ferrihydrite) and fully oxidized, then only
81% of the iron is in contained in nonstoichiometric
magnetite, and 19% is in maghemite. Since DM
c
of
Unit 1 samples is 0.25, the average remanence ratio
is 0.16 ± 0.02 (characteristic of generic PSD
behavior), and room temperature SIRM memory is
partially due to SD grains, an upper limit for zcan be
set at 0.5 for Unit 1, based on the model of Özdemir
and Dunlop [2010]. The Mössbauerbased calcula-
tion helps lower this upper limit to 0.3. The oxida-
tion parameter is necessary for determining m
ferri
values for Units 1 and 2 from equation (4), where the
two components considered are magnetite (m
Mt
=
92 Am
2
/kg) and maghemite (m
Mh
= 74.3 Am
2
/kg),
and the weighing factor is none other than z,
according to the quasilinear relationship between z
and either endmember fractions of the magnetite
maghemite oxidation series [Readman and OReilly,
1972]. For Unit 1 we find an average ferrimagnetic
component with m
ferri
=88Am
2
/kg (corresponding
to z= 0.3), while for Unit 2 the average m
ferri
=
78 Am
2
/kg (z= 0.8). The total mass concentration of
ferrimagnetic material (c
ferri
) is obtained by nor-
malizing bulk M
s
measurements by the respective
m
ferri
value (equation 3). Figure 4d shows c
ferri
delimited by total ferrimagnetic mass concentration
calculated as magnetite (lower limit) and maghe-
mite (upper limit).
4.2.5. Concentration of RemanenceCarrying
Particles
[27]The sedimentary MD, UNISD, and ISD com-
ponents are modeled using system (9), which has
been successfully tested on the synthetic mixtures.
The choice for endmember component ratios is
driven by the characteristics of the endmember
components in the depositional environment
examined.
[28]The UNISD component comprises all SD
particles or chains of particles that are isolated from
other magnetic grains in the sediment matrix. These
particles can have various sources, but in lake se-
diments they are overwhelmingly biogenic [Egli et
al., 2010]. Lakes generally have diverse magneto-
tactic bacteria communities that live in specific
geochemical conditions [Moskowitz et al., 2008],
so it is expected that populations of magnetosomes
originating in different environmental stress zones
have characteristic magnetic properties, specifically
the ratios used in this model. The magnetosomes
present in our samples have been identified as non
stoichiometric magnetite, based on the Mössbauer
analysis and the presence of Verwey transition in
the lowtemperature experiments. The remanence
ratio of the UNISD component is therefore set to
0.5 [Stoner and Wohlfarth, 1948]. The ARM ratio
should reflect the average value of noninteracting
magnetosomes and chains from all bacterial strains.
Since it is difficult to identify all the species that
produce magnetosomes in natural samples, and
there is little data for wild strain magnetic proper-
ties, the choice of UNISD ARM ratio should be
based on measurements of samples collected from
the horizons inhabited by magnetotactic bacteria.
The Brownie Lake water column ARM ratio of the
integrated biogenic coercivity component peaks at
the OAI (BS+BH in Figure 3c). The decrease of
the ARM ratio with depth is partially due to the
incipient breakdown during settling of some of the
chains produced at the OAI, and partially due to
possibly lower intrinsic ARM ratios of magneto-
some chains produced by some of the strains in the
lowoxygen layers [Egli, 2004]. It is also plausible
Table 2. Magnetic Hyperfine Parameters at Room Temperature for the Brownie Lake Surface Sediment Magnetic Extract
Iron Phase B
HFa
(T) QS
b
(mm/s) IS
c
(mm/s) %Fe
d
Magnetite (site Atetrahedral) 49.0(2)
e
0.04(2) 0.26(2) 28
Magnetite (site Boctahedral) 45.8(4) 0.02(2) 0.67(1) 34
Hematite 51.3(3) 0.22(1) 0.38(2) 9
Maghemite 50.2(5) 0.03(2) 0.43(5) 6
Fe
3+
narrow doublet 0.51(1) 0.49(1) 9
Fe
3+
wide doublet 1.53(3) 0.70(2) 7
Fe
2+
2.72(3) 1.03(1) 7
a
Magnetic hyperfine field.
b
Quadrupole splitting.
c
Isomer shift.
d
Relative concentration.
e
Errors quoted in parentheses refer to the last decimal.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
14 of 22
that some of low coercivity particles are nonin-
teracting pedogenic particles that also have lower
intrinsic ARM ratios [Geiss et al., 2008]. Coer-
civity spectra deconvolution does not discriminate
between interacting and noninteracting particles,
but isolating the biogenic component in its imme-
diate environment is the most important step in
the process of choosing a representative ARM ratio
for UNISD. Based on the BS+BH ARM ratio value
at the OAI, the habitat of magnetotactic bacteria,
we use (c
a
/M
rs
)
UNISD
= 3.5 · 10
3
m/A in our
model. The ISD endmember ratios were chosen to
represent strongly interacting SD particles (M
rs
/M
s
=
0.41; c
a
/M
rs
=0.12·10
3
m/A [Moskowitz et al.,
1993]), e.g., clumps of collapsed magnetosomes
from bacterial cells with totally degraded organic
membranes, or washedin pedogenic particle
aggregates. The MD component endmember ratios
(M
rs
/M
s
= 0.05; c
a
/M
rs
= 0.1 · 10
3
m/A) represent
average values for MD grains sensu lato (i.e.,
account for the presence PSD particles).
[29]Figure 5 illustrates the modeling results. Bulk
M
rs
/M
s
and c
a
/M
rs
are compared to theoretical
values in Figure 5a. A mixing grid is formed by the
lines connecting the three endmember components
(defined above), which are calculated using system
(10). The fractions of the endmembers in each
sample are plotted in a ternary diagram (Figure 5b),
which is equivalent to the mixing grid, but in g
i
coordinates. On average, Unit 1 samples contain
6080% MD, 2030% ISD, and <10% UNISD
particles. Comparatively, presettlement samples
contain only 3050% MD grains, while ISD and
UNISD endmembers are well represented with
2040% of the total ferrimagnetic mass each. In
Figure 5c we plot results from eight model runs that
used all possible combinations of the following
endmember ratio pairs (M
rs
/M
s
,c
a
/M
rs
), which
represent boundaries defined for each component:
(0.5, 3 · 10
3
m/A) and (0.5, 4 · 10
3
m/A) for
UNISD, (0.41, 0.1 · 10
3
m/A) and (0.45, 0.15 ·
10
3
m/A) for ISD, (0.01, 0.05 · 10
3
m/A) and
Figure 5. (a) Crossplot of M
rs
/M
s
versus c
a
/M
rs
for Brownie Lake Units 1 (triangles) and 2 (dots). Theoretical
values for bulk ratios were calculated from system (10) at 0.1 fraction increments of endmember components (see
text for endmember ratio values). (b) Ternary diagram of calculated fractions for each component (UNISD, ISD, and
MD or PSD) in Brownie Lake sediments, according to the three endmember mixing model (g
i
are solutions to system 9).
(c) Time series of endmember component fractions for model runs with eight different endmember combinations.
Lines represent mean values, and uncertainty bars are two standard deviations in length (see text for error calculation
method).
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
15 of 22
(0.05, 0.1 · 10
3
m/A) for MD. The curves are
mean g
i
values of the model runs, and the length of
the error bars is two standard deviations.
4.2.6. Concentration of SP Particles
[30]A broad distribution of particle sizes was
inferred for the SP grains based on frequency
dependence of susceptibility across the whole
spectrum of measurement temperatures (Figure 6a).
This implies that the c
SP
value to be used in
equation (11) for the calculation of f
SP
must be the
average of SP susceptibilities over the entire range
of SP volumes. We calculate c
SP
following Worm
[1998], which assumes even coercivity distribu-
tions between 40 and 60 mT for of each grain size
fraction. For a measurement frequency of 100 Hz,
average c
SP
across the frequencydependence
interval is 4 · 10
3
m
3
/kg. c
nonSP
was calculated
using an SD value of 0.4 · 10
3
m
3
/kg, and an MD
value of 0.6 · 10
3
m
3
/kg (average values for
nonstoichiometric magnetite [Dunlop and Özdemir,
1997]), weighed by their respective fractions as
obtained from system (9). SP fraction f
SP
is plotted
in Figure 6b (continuous line); Unit 1 SP fractions
are between 0.05 and 0.1, while Unit 2 values are
in the interval 0.10.2. For comparison, dashed
lines are f
SP
curves calculated using c
SP
values of
3·10
3
m
3
/kg (higher values), and 5 · 10
3
m
3
/kg
(lower values). Figure 6c illustrates the contribu-
tions of the SP and nonSP fractions to c
ferri
. The
SP particles account for as much as 50% of c
ferri
in
some intervals, even though their concentration is
only 10%.The nonSP baseline in Unit 1 has
slightly higher values than in Unit 2 because of the
prevalence of MD particles, but the average base-
line value is equivalent to the one used by Hunt et
al. [1995b]. Support for our method is provided by
the Mössbauer analysis of the magnetic extract.
The SP concentration given by the narrow Fe
3+
doublet in Figure 4c (representing paramagnetic
and SP iron oxides) is 12% of the ferrimagnetic
material (9% of total Fe). Given that not only fer-
rimagnetic SP material contributes to this doublet,
this value should be taken as a maximum SP con-
centration. By comparison, our ferrimagnetic SP
fraction calculation for the top 1015 cm of the
core (equivalent to the surface sample from which
the extract was prepared) yields an average con-
centration of 8%.
4.2.7. Ferrimagnetic Flux Model
[31]In the previous sections we have calculated the
mass of ferrimagnetic material relative to the total
dry sediment mass (Figure 4d), and the partition
between its constituent components: UNISD, ISD,
MD (Figure 5c), and SP (Figure 6b). Knowing the
preand postsettlement sedimentation rates from
the Brownie Lake age model [Swain, 1984], sedi-
ment density from core logging, and composition
from LOI, the calculated mass fractions of the
ferrimagnetic components can be expressed as
sediment loads or fluxes (Figure 7) according to
equations (7) and (8). The full interpretation of the
Figure 6. (a) Frequency and temperature dependence
of susceptibility for representative samples of each sed-
imentary unit. Both samples are frequency dependent in
the interval 30300 K, indicating a broad SP grain size
distribution. (b) SP fraction calculated for c
SP
=4·
10
3
m
3
/kg (thick line), bracketed by calculations for
c
SP
values of 5 · 10
3
m
3
/kg(left),and3·10
3
m
3
/kg
(right). (c) SP and nonSP contributions to c
ferri
.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
16 of 22
temporal fluctuations of the ferrimagnetic compo-
nents in terms of regional environmental and
anthropogenic forcing factors will be discussed in
detail elsewhere (I. Lascu et al., manuscript in
preparation, 2010); however, a synopsis follows.
The annual sedimentary ferrimagnetic flux before
European settlement is fairly constant at 0.01
0.02 mg/cm
2
, and is dominated by the SD and SP
fractions (70%). After 1850 the total flux steadily
increases and by 1950 is an order of magnitude
higher than in presettlement times. This tenfold
increase is observed in the MD, ISD, and SP
components. The 1917 lake level lowering is
marked by shortlived abrupt increase in sediment
delivery from the surrounding catchment, marked
in Figure 7 by peaks in the MD and ISD compo-
nents. The annual UNISD flux is on the order of
26mg/cm
2
before, and 412 mg/cm
2
after settle-
ment, with the lowest values (<2 mg/cm
2
) occurring
in the eighteenth and nineteenth centuries. By
comparison to the other components, whose fluxes
are controlled by increased erosion rates, UNISD
postsettlement values are only two times higher
than in presettlement times, and are tied to an
increase in nutrient delivery brought by anthro-
pogenically driven eutrophication of the lake in the
twentieth century [Swain, 1984]. The low UNISD
values that predate the 1917 lake level lowering
are interpreted as a sign of the reductive dissolution
front migrating downward into the sediment column
after the establishment of meromictic conditions.
4.3. Discussion of Modeling Parameters
Selection
[32]We here focus on the significance of the choice
of modeling parameters in understanding ferri-
magnetic components in their sedimentary context.
The total ferrimagnetic mass calculation is based
on M
s
, an intrinsic property of ferromagnetic
materials [Dunlop and Özdemir, 1997]. The error
associated with f
ferri
(or c
ferri
) is strictly related to
the determination of ferrimagnetic mineralogy, and
has a maximum value of 19% in the general case of
nonstoichiometric magnetite (i.e., unknown oxi-
dation parameter z). To decrease this error, low
temperature magnetization curves and Mössbauer
57
Fe spectroscopy have been employed to narrow
down possible zvalues for each of sedimentary
unit. The low temperature magnetic behavior of
Unit 1 samples and the Mössbauer analysis of the
surface sample indicate that the magnetite is only
oxidized at the particle surface. Brownie Lake has
been meromicitic since the 1920s, having anoxic
and reducing bottom waters that are rich in dis-
solved Fe
2+
and Mn
2+
[Swain, 1984; Tracey et al.,
1996], an unfavorable environment for oxidation of
magnetite particles. On the contrary, reductive
dissolution can potentially remove SP nanoparticles
and reduce the size of SD grains to SP [Anderson
and Rippey, 1988; Tarduno, 1995; Geiss et al.,
2004]. The magnetite particles must either have
been already oxidized before deposition and have
survived without being completely reduced, or
underwent oxidation during the laboratory storage
period of the core and samples, between collection
and measurement. Before settlement, the lake water
was only seasonally anoxic, complete mixing of
epilimnetic and hypolimnetic waters occurring at
spring and fall turnover events. Evidence for this
can be inferred from the lakes reduced relative
depth, the absence of annually laminated sedi-
ments, higher sedimentary Fe/Mn ratio, and the
presence of a diatom flora in a highertrophic status
[Swain, 1984; Tracey et al., 1996]. Because of the
more oxic conditions, and of the one order of
magnitude lower sedimentation rate [Swain, 1984],
Unit 2 ferrimagnetic particles are in a more
advanced oxidation state. The average magnetic
grain size is also lower than in Unit 1, which means
an increased surfacetovolume ratio, making the
grains more susceptible to oxidation. The choice
of m
ferri
for each unit was therefore determined by a
combination of sedimentmagnetic characteristics
Figure 7. Model of annual ferrimagnetic flux of
UNISD, ISD, MD, and SP particles in Brownie Lake
(note that Units 1 and 2 are plotted on separate axes).
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
17 of 22
and evidence of limnological conditions at the time
of deposition.
[33]The remanencecarrying components have
been modeled by way of a threecomponent mixing
model, using specific endmember values for
remanence and ARM ratios. According to our
model, bulk M
rs
/M
s
and c
a
/M
rs
of a sample can be
translated into mass fractions for each component
considered, here UNISD, ISD and MD. The basis
for defining these particular components is the
ability of the two ratios to discriminate between
grainsize categories (M
rs
/M
s
) and interparticle
interactions (c
a
/M
rs
). A judicious choice of end
member ratios will result in lower model errors. The
components for Brownie Lake were defined with the
knowledge that lake sediments in temperateclimate
areas are characterized by a mixture of detrital and
endogenic magnetic particles [Snowball et al., 2002;
Geiss et al., 2003; Egli, 2004], and that non-
stoichiometric magnetite is the ferrimagnetic carrier
in our sediments. The detrital fraction is generally
dominated by MD particles, but can also contain
finer SD and SP grains (e.g., of pedogenic origin).
The MD component sensu stricto (no PSD grains)
has low remanence and ARM ratio values (M
rs
/M
s
<
0.05, c
a
/M
rs
0.1 · 10
3
m/A). A PSD endmember
would have a higher remanence ratio (e.g., 0.2 for
1mm grains), but a similar ARM ratio. We use
the cutoff M
rs
/M
s
value of 0.05 [Day et al., 1977;
Dunlop, 2002] as an average in our model, in
order to incorporate the effects of both true MD
and PSD grains. This value corresponds to an
average grain size of 20 mm[Yu et al., 2002]. The
0.01 M
rs
/M
s
value used in the error estimation for
the MD component (g
3
in Figure 5c) corresponds to
much larger particles (100200 mm) [Day et al.,
1977; Dunlop, 2002], and would be a limit case.
The detrital SD grains can be transported into a
depositional basin either independently through
sorting, or in most cases as part of clay/organic
matter aggregates that also contain larger ferri-
magnetic particles. In both situations the SD par-
ticles are disturbed to a certain extent from their
original configuration (e.g., dispersion in a soil
matrix). It is then reasonable to assume that they
contribute to both ISD and UNISD endmembers.
The ISD component is characterized by SDlike
remanence ratios, and MDlike ARM ratios, and
was defined to represent the extreme case of strong
interactions between SD particles clumped in
aggregates [Moskowitz et al., 1993; Kopp et al.,
2006]. Both detrital (e.g., pedogenic) and endo-
genic (biogenic and inorganic) SD particles that
occur in such closely packed configurations are
expected to contribute to ISD. The UNISD com-
ponent is designed to include only isolated grains
or undisturbed linear chains of bacterial magneto-
somes [Yamazaki, 2008; Egli et al., 2010]. The
remanence ratio for magnetite UNISD particles is
0.5bydefinition[Stoner and Wohlfarth, 1948]. For
greigite magnetosomes the remanence ratio usually
exceeds 0.5, approaching a theoretical value of
0.83 as H
cr
/H
c
approaches unity, due to a [100]
easy axis of magnetization and the prevalence of
magnetocrystalline anisotropy over shape anisot-
ropy [e.g., Roberts, 1995; Sagnotti and Winkler,
1999; Chang et al., 2009]. The ARM ratio of
UNISD can be determined with the help of coer-
civity deconvolution of samples harvested from the
OAI, by calculating the ARM ratio of the biogenic
coercivityspectrum component (3.5 · 10
3
m/A in
our model). The biogenic component defined by its
coercivity is likely to include interacting particles
from magnetosome clusters and bundled or col-
lapsed chains [Egli et al., 2010], which can artifi-
cially lower (c
a
/M
rs
)
UNISD
. On the other hand,
the detrital coercivityspectrum component may
include noninteracting particles with lower initial
ARM ratios (e.g., 13·10
3
m/A for pedogenic
particles [Egli, 2004; Geiss et al., 2008]), which
should truly lower (c
a
/M
rs
)
UNISD
. Our error esti-
mation for the UNISD endmember (g
1
in Figure 5c)
was based on ARM ratios in the interval 3
10
3
m/A. The error is higher for Unit 2 samples,
indicating that the choice of ARM ratio for the
UNISD component is more critical for finer grained
magnetic assemblages.
[34]The SP component fraction is modeled using
the ferrimagnetic susceptibility c
ferri
(derived from
c
f
/M
s
) by subtracting the nonSP baseline suscep-
tibility [Hunt et al., 1995b]. The critical parameter
in the calculation of f
SP
is the susceptibility of the
SP ferrimagnetic grains, which can vary over an
order of magnitude with SP grain size [Worm,
1998]. Frequencydependent susceptibility mea-
surements across a range of temperatures are
therefore necessary to characterize the SP particle
size distribution, as a preliminary step for calcu-
lating a representative c
SP
. Sediments usually
comprise magnetic particles characterized by a
wide grainsize distribution in the SP realm
[Dearing et al., 1996], so an average c
SP
value
calculated over the SP size spectrum is a reasonable
assumption in most cases. This approach is pre-
ferred to the traditional room temperature suscep-
tibility measurements at two frequencies (c
fd
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
18 of 22
parameter), which captures only a narrow interval
of the grainsize distribution curve.
5. Conclusions
[35]1. We provide a quantitative model for calcu-
lating the concentrations of ferrimagnetic sedimen-
tary components using rock magnetic properties.
The quantification method is based mainly on bulk
room temperature measurements, and translates
raw magnetic parameters into ferrimagnetic mass
concentrations. This method eliminates dilution
effects on magnetic properties by weakly magnetic
substances in high concentrations, such as water,
organic matter, carbonates, clay minerals, etc., and
allows the calculation of ferrimagnetic particle
fluxes for dated sedimentary sequences. It can
be applied for reconstructing past environmental
changes in a range of sedimentary environments,
and is particularly useful for large sets of samples,
where detailed magnetic unmixing methods (espe-
cially lowtemperature techniques) are unfeasible
due to time or instrument constraints.
[36]2. Total ferrimagnetic concentration is most
accurately determined from infield magnetization
measurements. Saturation magnetization corrected
for nonferrimagnetic contributions (M
s
) is the
most reliable parameter, only ferrimagnetic miner-
alogy being needed for concentration calculations.
Susceptibilitybased concentrations are less reliable
because they require a priori knowledge about
ferrimagnetic grain size and shape in addition to
magnetic mineralogy. Remanence measurements
are useful as concentration proxies only when
magnetic composition is uniform with respect to
mineralogy and grain size, due to the varying
degree of remanence acquisition efficiency for
different categories of magnetic particles.
[37]3. Anhysteretic and saturation isothermal
remanences are used here to model the concentra-
tions of the remanencecarrying ferrimagnetic
fractions via a threecomponent mixing model,
which is tested on mixtures of SD and nonSD
magnetite grains of known concentrations. The
use of the ARM ratio (c
a
/M
rs
) as a proxy for inter
particle magnetostatic interactions allows the sepa-
ration of SD particles into two separate components
(UNISD and ISD), which may have independent
origins. The remanence ratio (M
rs
/M
s
) is a true
grainsize indicator, whereas c
a
/M
rs
should be used
in that capacity only when the proportion of inter-
acting particles is constant.
[38]4. For a robust quantification of the super-
paramagnetic fraction, we propose a technique that
corroborates ferrimagnetic susceptibility calcula-
tions with information about the SP particle size
distribution obtained from the frequency and tem-
perature dependence of magnetic susceptibility.
This approach is recommended, when possible, in
lieu of frequencydependent susceptibility mea-
surements at room temperature only.
Acknowledgments
[39]We thank the National Lacustrine Core Repository
(LacCore) at the University of Minnesota for access to the
Brownie Lake core and initial description data, Amy Chen
for help with water sampling and filtering, Bruce Moskowitz
and Brian The DudeCarter for access to the synthetic mix-
tures. Mike Jacksons comments and suggestions significantly
improved the development of our methodology. Constructive
reviews by Christoph Geiss, Mike Jackson, Leonardo Sagnotti
and an anonymous referee have helped improve this manu-
script. I.L. has benefited from the support of a University of
Minnesota Doctoral Dissertation Fellowship. The Institute for
Rock Magnetism is funded by the National Science Founda-
tion, the Keck Foundation, and the University Of Minnesota.
This is IRM contribution 1003.
References
Anderson, N. J., and B. Rippey (1988), Diagenesis of magnetic
minerals in the recent sediments of a eutrophic lake, Limnol.
Oceanogr.,33, 14761492, doi:10.4319/lo.1988.33.6_part_
2.1476.
Bleil, U., and T. von Dobeneck (2004), Late Quaternary terrig-
enous sedimentation in the western Equatorial Atlantic:
South American versus African provenance discriminated
by magnetic mineral analysis, in The South Atlantic in the
Late Quaternary: Reconstruction of Material Budgets and
Current Systems, edited by G. Wefer, S. Mulitza, and
V. Ratmeyer, pp. 213236, SpringerVerlag, Berlin.
Brand, R. A. (1987), Improving the validity of hyperfine field
distributions from metallic alloys. Part I: Unpolarized source,
Nucl. Instrum. Methods Phys. Res., Sect. B,28, 398416,
doi:10.1016/0168-583X(87)90182-0.
CarterStiglitz, B., B. Moskowitz, and M. Jackson (2001),
Unmixing magnetic assemblages and the magnetic behavior
of bimodal mixtures, J. Geophys. Res.,106(B11), 26,397
26,411, doi:10.1029/2001JB000417.
Chang, L., A. P. Roberts, Y. Tang, B. D. Rainford, A. R.
Muxworthy, and Q. W. Chen (2008), Fundamental magnetic
parameters from pure synthetic greigite (Fe
3
S
4
), J. Geophys.
Res.,113, B06104, doi:10.1029/2007JB005502.
Chang, L., A. P. Roberts, C. J. Rowan, Y. Tang, P. Pruner,
Q. Chen, and C. Horng (2009), Lowtemperature magnetic
properties of greigite (Fe
3
S
4
), Geochem. Geophys. Geosyst.,
10, Q01Y04, doi:10.1029/2008GC002276.
Day, R., M. Fuller, and V. A. Schmidt (1977), Hysteresis prop-
erties of titanomagnetites: Grainsize and compositional
dependence, Phys. Earth Planet. Inter.,13,260267,
doi:10.1016/0031-9201(77)90108-X.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
19 of 22
Dean, W. E. (1974), Determination of carbonate and organic
matter in calcareous sediments and sedimentary rocks by
loss on ignition; comparison with other methods, J. Sedi-
ment. Petrol.,44,242248, doi:10.1306/74D729D2-2B21-
11D7-8648000102C1865D.
Dearing, J. A. (1999a), Holocene environmental change from
magnetic proxies in lake sediments, in Quaternary Climates,
Environments and Magnetism, edited by B. A. Maher and
R. Thompson, pp. 231278, doi:10.1017/CBO9780511535635.
008, Cambridge Univ. Press, Cambridge, U. K.
Dearing, J. A. (1999b), Magnetic susceptibility, in Environ-
mental Magnetism: A Practical Guide, edited by J. Walden,
F. Oldfield, and J. P. Smith, pp. 3553, Quat. Res. Assoc.,
London.
Dearing, J. A., R. J. L. Dann, K. Hay, J. A. Lees, P. J. Loveland,
B. A. Maher, and K. OGrady (1996), Frequencydependent
susceptibility measurements of environmental materials,
Geophys. J. Int.,124,228240, doi:10.1111/j.1365-246X.
1996.tb06366.x.
Dearing, J. A., J. F. Boyle, P. G. Appleby, A. W. Mackay, and
R. J. Flower (1998), Magnetic properties of recent sediments
in Lake Baikal, Siberia, J. Paleolimnol.,20, 163173,
doi:10.1023/A:1008045029157.
Demory, F., H. Oberhaensli, N. R. Nowaczyk, M. Gottschalk,
R. Wirth, and R. Naumann (2005), Detrital input and early
diagenesis in sediments from Lake Baikal revealed by rock
magnetism, Global Planet. Change,46,145166,
doi:10.1016/j.gloplacha.2004.11.010.
Dunlop, D. J. (2002), Theory and application of the Day plot
(M
rs
/M
s
versus H
cr
/H
c
): 1. Theoretical curves and tests using
titanomagnetite data, J. Geophys. Res.,107(B3), 2056,
doi:10.1029/2001JB000486.
Dunlop, D. J., and B. CarterStiglitz (2006), Day plots of mix-
tures of superparamagnetic, singledomain, pseudosingle
domain, and multidomain magnetites, J. Geophys. Res.,
111, B12S09, doi:10.1029/2006JB004499.
Dunlop, D. J., and Ö. Özdemir (1997), Rock Magnetism: Fun-
damentals and Frontiers, 573 pp., Cambridge Univ. Press,
Cambridge, U. K.
Egli, R. (2003), Analysis of the field dependence of remanent
magnetization curves, J. Geophys. Res.,108(B2), 2081,
doi:10.1029/2002JB002023.
Egli, R. (2004), Characterization of individual rock magnetic
components by analysis of remanence curves: 1. Unmixing
natural sediments, Stud. Geophys. Geod.,48, 391446,
doi:10.1023/B:SGEG.0000020839.45304.6d.
Egli, R. (2006a), Theoretical aspects of dipolar interactions
and their appearance in firstorder reversal curves of ther-
mally activated singledomain particles, J. Geophys. Res.,
111, B12S17, doi:10.1029/2006JB004567.
Egli, R. (2006b), Theoretical considerations on the anhystere-
tic remanent magnetization of interacting particles with uni-
axial anisotropy, J. Geophys. Res.,111, B12S18,
doi:10.1029/2006JB004577.
Egli, R., and W. Lowrie (2002), Anhysteretic remanent mag-
netization of fine magnetic particles, J. Geophys. Res.,107
(B10), 2209, doi:10.1029/2001JB000671.
Egli, R., A. P. Chen, M. Winklhofer, K. P. Kodama, and C.S.
Horng (2010), Detection of noninteracting single domain
particles using firstorder reversal curve diagrams, Geochem.
Geophys. Geosyst.,11, Q01Z11, doi:10.1029/
2009GC002916.
Evans, M. E., and F. Heller (2003), Environmental Magne-
tism: Principles and Applications of Enviromagnetics,299
pp., Academic, San Diego, Calif.
Frank, U., and N. R. Nowaczyk (2008), Mineral magnetic
properties of artificial samples systematically mixed from
haematite and magnetite, Geophys. J. Int.,175, 449461,
doi:10.1111/j.1365-246X.2008.03821.x.
Frederichs, T., U. Bleil, K. Daeumler, T. von Dobeneck, and
A. M. Schmidt (1999), The magnetic view on the marine
paleoenvironment; parameters, techniques and potentials of
rock magnetic studies as a key to paleoclimatic and paleocea-
nographic changes, in Use of Proxies in Paleoceanography;
Examples From the South Atlantic, edited by G. Fischer and
G. Wefer, pp. 575599, Springer, Berlin.
Funk, J. A., T. von Dobeneck, T. Wagner, and S. Kasten
(2004), Late Quaternary sedimentation and early diagenesis
in the Equatorial Atlantic Ocean; patterns, trends and pro-
cesses deduced from rock magnetic and geochemical records,
in The South Atlantic in the Late Quaternary: Reconstruction
of Material Budgets and Current Systems, edited by G. Wefer,
S. Mulitza, and V. Ratmeyer, pp. 461497, SpringerVerlag,
Berlin.
Geiss, C. E., and S. K. Banerjee (1997), A multiparameter
rock magnetic record of the last glacialinterglacial paleocli-
mate from southcentral Illinois, USA, Earth Planet. Sci.
Lett.,152, 203216, doi:10.1016/S0012-821X(97)00133-7.
Geiss, C. E., and S. K. Banerjee (1999), Comparison of two
interglacial records from the midwestern U.S.A: Rock mag-
netism, palaeomagnetism and environmental magnetism,
Phys. Chem. Earth, Part A,24, 793798, doi:10.1016/
S1464-1895(99)00116-7.
Geiss, C. E., and C. W. Zanner (2006), How abundant is ped-
ogenic magnetite? Abundance and grain size estimates for
loessic soils based on rock magnetic analyses, J. Geophys.
Res.,111, B12S21, doi:10.1029/2006JB004564.
Geiss, C. E., C. E. Umbanhowar, P. Camill, and S. K. Banerjee
(2003), Sediment magnetic properties reveal Holocene cli-
mate change along the Minnesota prairieforest ecotone;
Lake basins as archives of continental tectonics and paleocli-
mate, J. Paleolimnol.,30,151166, doi:10.1023/
A:1025574100319.
Geiss, C. E., S. K. Banerjee, P. Camill, and C. E. Umbanhowar
(2004), Sedimentmagnetic signature of landuse and
drought as recorded in lake sediment from southcentral
Minnesota, USA, Quat. Res.,62,117125, doi:10.1016/j.
yqres.2004.06.009.
Geiss, C. E., R. Egli, and C. W. Zanner (2008), Direct esti-
mates of pedogenic magnetite as a tool to reconstruct past
climates from buried soils, J. Geophys. Res.,113, B11102,
doi:10.1029/2008JB005669.
GibbsEggar, Z., B. Jude, J. Dominik, J. Loizeau, and
F. Oldfield (1999), Possible evidence for dissimilatory bacte-
rial magnetite dominating the magnetic properties of Recent
lake sediments, Earth Planet. Sci. Lett.,168,16,
doi:10.1016/S0012-821X(99)00054-0.
HaltiaHovi, E., N. Nowaczyk, T. Saarinen, and B. Plessen
(2010), Magnetic properties and environmental changes
recorded in Lake Lehmilampi (Finland) during the Holocene,
J. Paleolimnol.,43,113, doi:10.1007/s10933-009-9309-8.
Heiri, O., A. F. Lotter, and G. Lemcke (2001), Loss on ignition
as a method for estimating organic and carbonate content
in sediments: Reproducibility and comparability of results,
J. Paleolimnol.,25,101110, doi:10.1023/A:1008119611481.
Heslop, D., M. J. Dekkers, P. P. Kruiver, and I. H. M.
van Oorschot (2002), Analysis of isothermal remanent
magnetization acquisition curves using the expectation
maximization algorithm, Geophys. J. Int.,148,5864,
doi:10.1046/j.0956-540x.2001.01558.x.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
20 of 22
Hunt, C. P., and S. K. Banerjee (1992), Thermal demagnetiza-
tion of lowtemperature SIRM: A new method for magnetic
granulometry, Eos Trans. AGU,73, 138.
Hunt, C. P., B. M. Moskowitz, and S. K. Banerjee (1995a),
Magnetic properties of rocks and minerals, in Rock Physics
and Phase Relations: A Handbook of Physical Constants,
editedbyT.J.Ahrens,pp.189204, AGU, Washington,
D. C.
Hunt, C. P., S. K. Banerjee, J. Han, P. A. Solheid, E. Oches,
W. Sun, and T. Liu (1995b), Rockmagnetic proxies of cli-
mate change in the loesspalaeosol sequences of the western
Loess Plateau of China, Geophys. J. Int.,123, 232244,
doi:10.1111/j.1365-246X.1995.tb06672.x.
Kim, B. Y., K. P. Kodama, and R. E. Moeller (2005), Bacterial
magnetite produced in water column dominates lake sedi-
ment mineral magnetism: Lake Ely, USA, Geophys. J. Int.,
163,2637, doi:10.1111/j.1365-246X.2005.02735.x.
Kopp, R. E., C. Z. Nash, A. Kobayashi, B. P. Weiss, D. A.
Bazylinski, and J. L. Kirschvink (2006), Ferromagnetic reso-
nance spectroscopy for assessment of magnetic anisotropy
and magnetostatic interactions: A case study of mutant mag-
netotactic bacteria, J. Geophys. Res.,111,B12S25,
doi:10.1029/2006JB004529.
Kruiver, P. P., M. J. Dekkers, and D. Heslop (2001), Quanti-
fication of magnetic coercivity components by the analysis
of acquisition curves of isothermal remanent magnetisation,
Earth Planet. Sci. Lett.,189,269276, doi:10.1016/S0012-
821X(01)00367-3.
Larrasoaña, J. C., A. P. Roberts, R. J. Musgrave, E. Gracia,
E. Pinero, M. Vega, and F. MartinezRuiz (2007), Diagenetic
formation of greigite and pyrrhotite in gas hydrate marine sed-
imentary systems, Earth Planet. Sci. Lett.,261, 350366,
doi:10.1016/j.epsl.2007.06.032.
Lees, J. A. (1997), Mineral magnetic properties of mixtures of
environmental and synthetic materials; linear additivity and
interaction effects, Geophys. J. Int.,131,335346,
doi:10.1111/j.1365-246X.1997.tb01226.x.
Moskowitz, B. M., R. B. Frankel, and D. A. Bazylinski (1993),
Rock magnetic criteria for the detection of biogenic magne-
tite, Earth Planet. Sci. Lett.,120,283300, doi:10.1016/
0012-821X(93)90245-5.
Moskowitz, B. M., D. A. Bazylinski, R. Egli, R. B. Frankel,
and K. J. Edwards (2008), Magnetic properties of marine
magnetotactic bacteria in a seasonally stratified coastal pond
(Salt Pond, MA, USA), Geophys. J. Int.,174,7592,
doi:10.1111/j.1365-246X.2008.03789.x.
Murad, E., and J. Cashion (2004), Mössbauer Spectroscopy of
Environmental Materials and Their Industrial Utilization,
Kluwer, Boston, Mass.
Muxworthy, A. R., D. J. Dunlop, and Ö. Özdemir (2003),
Lowtemperature cycling of isothermal and anhysteretic
remanence; microcoercivity and magnetic memory, Earth
Planet. Sci. Lett.,205, 173184, doi:10.1016/S0012-821X
(02)01039-7.
Néel, L. (1949), Théorie du traînage magnétique des ferromag-
nétiques en grains fins avec applications aux terres cuites,
Ann. Geophys.,5,99136.
Nowaczyk, N. R. (2001), Logging of magnetic susceptibility,
in Tracking Environmental Change Using Lake Sediments,
vol. 1, Basin Analysis, Coring, and Chronological Techni-
ques,editedbyW.M.LastandJ.P.Smol,pp.155170,
Kluwer, Dordrecht, Netherlands.
OReilly, W., and S. K. Banerjee (1967), The mechanism of
oxidation in titanomagnetites: A magnetic study, Mineral.
Mag.,36,2937, doi:10.1180/minmag.1967.036.277.04.
Oldfield, F., R. Wake, J. Boyle, R. Jones, S. Nolan, Z. Gibbs,
P. Appleby, E. Fisher, and G. Wolff (2003), The late
Holocene history of Gormire Lake (NE England) and its
catchment: A multiproxy reconstruction of past human impact,
Holocene,13,677690, doi:10.1191/0959683603hl654rp.
Ortega, B., M. Caballero, S. Lozano, G. Vilaclara, and
A. Rodriguez (2006), Rock magnetic and geochemical
proxies for iron mineral diagenesis in a tropical lake: Lago
Verde, Los Tuxtlas, eastcentral Mexico, Earth Planet. Sci.
Lett.,250, 444458, doi:10.1016/j.epsl.2006.08.020.
Özdemir, Ö., and D. J. Dunlop (2010), Hallmarks of maghemi-
tization in lowtemperature remanence cycling of partially
oxidized magnetite nanoparticles, J. Geophys. Res.,115,
B02101, doi:10.1029/2009JB006756.
Özdemir, Ö., D. J. Dunlop, and B. M. Moskowitz (1993), The
effect of oxidation on the Verwey transition in magnetite,
Geophys. Res. Lett.,20, 16711674, doi:10.1029/
93GL01483.
Paasche, O., R. Lovlie, S. O. Dahl, J. Bakke, and A. Nesje
(2004), Bacterial magnetite in lake sediments: Late glacial
to Holocene climate and sedimentary changes in northern
Norway, Earth Planet. Sci. Lett.,223, 319333, doi:10.1016/
j.epsl.2004.05.001.
Peck, J. A., and J. W. King (1996), Magnetofossils in the sed-
iment of Lake Baikal, Siberia, Earth Planet. Sci. Lett.,140,
159172, doi:10.1016/0012-821X(96)00027-1.
Peters, C., and M. J. Dekkers (2003), Selected room tempera-
ture magnetic parameters as a function of mineralogy, con-
centration and grain size, Phys. Chem. Earth,28, 659667,
doi:10.1016/S1474-7065(03)00120-7.
Readman, P. W., and W. OReilly (1971), Oxidation processes
in titanomagnetites, Z. Geophys.,37, 329338.
Readman, P. W., and W. OReilly (1972), Magnetic properties
of oxidized (cationdeficient) titanomagnetites (Fe,Ti)
3
O
4
,
J. Geomagn. Geoelectr.,24,6990.
Reynolds, R. L., J. G. Rosenbaum, J. Rapp, M. W. Kerwin,
P. J. Bradbury, S. Colman, and D. Adam (2004), Record of
Late Pleistocene glaciation and deglaciation in the southern
Cascade Range. I. Petrological evidence from lacustrine sed-
iment in upper Klamath Lake, southern Oregon, J. Paleolim-
nol.,31,217233, doi:10.1023/B:JOPL.0000019230.
42575.03.
Roberts, A. P. (1995), Magnetic properties of sedimentary
greigite (Fe
3
S
4
), Earth Planet. Sci. Lett.,134, 227236,
doi:10.1016/0012-821X(95)00131-U.
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. Res.,
105(B12), 28,46128,475, doi:10.1029/2000JB900326.
Robertson, D. J., and D. E. France (1994), Discrimination of
remanencecarrying minerals in mixtures, using isothermal
remanent magnetisation acquisition curves, Phys. Earth
Planet. Inter.,82,223234, doi:10.1016/0031-9201(94)
90074-4.
Rochette, P. (1987), Magnetic susceptibility of the rock matrix
related to magnetic fabric studies, J. Struct. Geol.,9, 1015
1020, doi:10.1016/0191-8141(87)90009-5.
Rosenbaum, J. G., and R. L. Reynolds (2004a), Record of Late
Pleistocene glaciation and deglaciation in the southern Cas-
cade Range. II. Flux of glacial flour in a sediment core from
upper Klamath Lake, Oregon, J. Paleolimnol.,31, 235252,
doi:10.1023/B:JOPL.0000019229.75336.7a.
Rosenbaum, J. G., and R. L. Reynolds (2004b), Basis for
paleoenvironmental interpretation of magnetic properties of
sediment from upper Klamath Lake (Oregon): Effects of
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
21 of 22
weathering and mineralogical sorting, J. Paleolimnol.,31,
253265, doi:10.1023/B:JOPL.0000019228.46421.f4.
Rosenbaum, J. G., R. L. Reynolds, D. P. Adam, J. Drexler,
A. M. SarnaWojcicki, and G. C. Whitney (1996), Record
of middle Pleistocene climate change from Buck Lake, Cas-
cade Range, southern OregonEvidence from sediment
magnetism, traceelement geochemistry, and pollen, Geol.
Soc. Am. Bull.,108,13281341, doi:10.1130/0016-7606
(1996)108<1328:ROMPCC>2.3.CO;2.
Sagnotti, L., and A. Winkler (1999), Rock magnetism and
palaeomagnetism of greigitebearing mudstones in the Italian
peninsula, Earth Planet. Sci. Lett.,165,6780, doi:10.1016/
S0012-821X(98)00248-9.
Snowball, I. F. (1994), Bacterial magnetite and the magnetic
properties of sediments in a Swedish lake, Earth Planet.
Sci. Lett.,126,129142, doi:10.1016/0012-821X(94)
90246-1.
Snowball, I., P. Sandgren, and G. Petterson (1999), The mineral
magnetic properties of an annually laminated Holocene lake
sediment sequence in northern Sweden, Holocene,9,353
362, doi:10.1191/095968399670520633.
Snowball, I., L. Zillen, and P. Sandgren (2002), Bacterial mag-
netite in Swedish varved lakesediments: A potential bio
marker of environmental change, Quat. Int.,88,1319,
doi:10.1016/S1040-6182(01)00069-6.
Stoner, E. C., and E. P. Wohlfarth (1948), A mechanism of
magnetic hysteresis in heterogeneous alloys, Philos. Trans.
R. Soc. London, Ser. A,240, 599642, doi:10.1098/
rsta.1948.0007.
Swain, E. B. (1984), The paucity of bluegreen algae in
meromictic Brownie Lake; iron limitation or heavy metal
toxicity?, Ph.D. thesis, Univ. of Minn., Twin Cities.
Tarduno, J. A. (1995), Superparamagnetism and reduction
diagenesis in pelagic sediments: Enhancement or depletion?,
Geophys. Res. Lett.,22, 13371340, doi:10.1029/
95GL00888.
Tracey, B., N. Lee, V. Card, and L. C. K. Shane (1996), Sed-
iment indicators of meromixis: Comparison of laminations,
diatoms, and sediment chemistry in Brownie Lake, Minnea-
polis, USA, J. Paleolimnol.,15, 129132, doi:10.1007/
BF00196776.
van der Post, K. D., F. Oldfield, E. Y. Haworth, P. R. J. Crooks,
and P. G. Appleby (1997), A record of accelerated erosion in
the Recent sediments of Blelham Tarn in the English Lake
District, J. Paleolimnol.,18, 103120, doi:10.1023/
A:1007922129794.
von Dobeneck, T. (1998), The concept of partial susceptibil-
ities, Geol. Carpath.,49, 228229.
Worm, H. (1998), On the superparamagneticstable single
domain transition for magnetite, and frequency dependence
of susceptibility, Geophys. J. Int.,133,201206,
doi:10.1046/j.1365-246X.1998.1331468.x.
Xie, Q., T. Chen, H. Xu, J. Chen, J. Ji, H. Lu, and X. Wang
(2009), Quantification of the contribution of pedogenic mag-
netic minerals to magnetic susceptibility of loess and paleo-
sols on Chinese Loess Plateau: Paleoclimatic implications,
J. Geophys. Res.,114, B09101, doi:10.1029/2008JB005968.
Xie, S., J. A. Dearing, and J. Bloemendal (1999), A partial sus-
ceptibility approach to analysing the magnetic properties of
environmental materials: A case study, Geophys. J. Int.,
138, 851856, doi:10.1046/j.1365-246x.1999.00915.x.
Yamazaki, T. (2008), Magnetostatic interactions in deepsea
sediments inferred from firstorder reversal curve diagrams:
Implications for relative paleointensity normalization, Geochem.
Geophys. Geosyst.,9, Q02005, doi:10.1029/2007GC001797.
Yu, Y., D. J. Dunlop, and Ö. Özdemir (2002), Partial anhys-
teretic remanent magnetization in magnetic: 1. Additivity,
J. Geophys. Res.,107(B10), 2244, doi:10.1029/2001JB001249.
Zillén, L., and I. Snowball (2009), Complexity of the 8 ka cli-
mate event in Sweden recorded by varved lake sediments,
Boreas,38,493503, doi:10.1111/j.1502-3885.2009.00086.x.
Geochemistry
Geophysics
Geosystems
G3
G3
LASCU ET AL.: FERRIMAGNETIC QUANTIFICATION 10.1029/2010GC003182
22 of 22
... The abundance of magnetite derived from the magneto-taxis prevalent in the oxic-anoxic interface of the seawater/lakes, and lacustrine sediments enhances the magnetic properties of the soils and sediments. 9,10 The magneto-tactic bacteriaderived ferrimagnetic NPs also detected in the fresh water and urban lakes. 10 Geiss and Zanner (2006) have shown that the upper horizons of many modern and buried soils have higher concentrations of ferrimagnetic magnetite and maghemite than their parent material. ...
... 9,10 The magneto-tactic bacteriaderived ferrimagnetic NPs also detected in the fresh water and urban lakes. 10 Geiss and Zanner (2006) have shown that the upper horizons of many modern and buried soils have higher concentrations of ferrimagnetic magnetite and maghemite than their parent material. 11 Intense weathering, forest fire, and eluvial soils derived from the basalts, abundant with detritus parent rock materials endow the natural environments with magnetic particles. ...
... 16,17 Moreover smelter emission and dust from the mining activities enrich soils with the magnetic spinel group minerals, one of the common hosts of metal-bearing contaminants including Pb. 18 Lascu et al. reported that a qualitative estimate of the magnetic properties of the sediments can be evaluated from the χ measurements but requires the knowledge of magnetic particles grain size. 10 Therefore, the extent of surface modifications of magnetic NPs, biomolecular conformations, grain size, and the degree of magnetic dipolar interactions evolved from polydispersity may facilitate the growth of magnetic field-directed selfassembled nanostructures in the environment. ...
Article
Sequential sorption of bovine serum albumin (BSA)-humic acid (HA)-BSA on weakly ferromagnetic maghemite nanoparticles has achieved the formation of amyloid scaffold-controlled self-assembly (B2-GFeNPs) at pH 4. Unlike BSA-modified maghemite (B1-GFeNPs),...
... The mean values of remanence coercivity fluctuate between 25.9 and 35.5 mT (Fig. 5c, Table S1), and this confirms the presence of low-coercivity magnetite/ maghemite grains. The mean values of the Mrs/χ ratio (Fig. 5c) The bi-plot of Mrs/Ms versus χ ARM /Mrs was applied according to Lascu et al. (2010) in order to estimate particle size variations and interparticle magnetostatic interactions. The values of χ ARM /Mrs ratio presented in Fig. 5d are in the range of 0.005-0.6 ...
... × 10 −3 m/A, and Mrs/Ms ratio reaches values from 0.05 to 0.25 (Fig. 5d). Such distribution of data points on the area of the Lascu diagram (Lascu et al., 2010) corresponds to MD + SD grains. Except for TMPs from glass and coke production (glass_1, coke_1 and some of coke_2 data points) that are shifted towards lower values of χ ARM /Mrs ratio. ...
Article
Variations in mineralogical composition, grain size internal structure and stoichiometry of technogenic magnetic particles (TMPs) deposited in topsoil may provide crucial information necessary to trace main pollution sources and recognize various technological processes. The aim of the study was to characterize, by means of magnetic parameters and Mössbauer spectra, the TMPs from non-ferrous metallurgy, cement, coke, glass production as well as long range transport (LRT) and compare the obtained data with previous results focused on iron mining and metallurgy. This research shows that only certain pollution sources (e.g. mainly iron mining, iron metallurgy, LRT and partly glass production) can be successfully distinguished by the applied parameters. The main features characteristic for TMPs produced by Fe-mining are: high values of concentration-dependent magnetic parameters, low values of coercivity, significant contribution from coarse MD (multi-domain) grains and a relatively high stoichiometry of magnetite. The most discriminative feature for TMPs generated by the glass industry is the abundance of goethite in the topsoil samples, which is confirmed by magnetic and Mössbauer techniques. The TMPs released by the Ni-Cu smelter and the Pb-Zn waste exhibit significant differences in the Mössbauer parameters, indicating different stoichiometry of magnetite for each group. Such variations are due to replacement of Fe by other elements at tetrahedral sites in the case of TMPs released from the Ni-Cu smelter. TMPs characteristic for the LRT emissions contain higher amount of finer fraction of low-stoichiometry magnetite (mostly single-domain SD particles) than those originating from other sources. The TMPs accumulated in the topsoils around the coking plants cannot be clearly discriminated by the applied methodology due to strong influence of the local pollution sources. Magnetic studies of the TMPs generated by cement production are complicated, since their properties mainly depend on individual technology (e.g. additives) used by the local cement plants.
... Petrol exhaust pipe and brake pad samples also lie within the fine, interacting magnetite region whereas the diesel exhaust pipe falls in a coarser region. Figure 7b shows our specimen data plotted on the Lascu plot (Lascu et al., 2010). Our leaf specimens and diesel exhaust pipe specimen lie on the interacting single-domain (ISD) to multi-domain (MD) mixing line, in agreement with Figure 7a. ...
... It is indicative of the complicated relationship of mean grain size, where MDF ARM increases with decrease in grain size. (b) Leaf, exhaust and non-exhaust specimens are plotted on the M rs /M s versus χ ARM /M rs for MD-SD (diamonds) and PSD-SD (circles) mixtures (Lascu et al., 2010). The numbers next to the symbols represent SD fraction in the total mixture. ...
Article
Full-text available
We report the characterization of anthropogenic magnetic particulate matter (MPM) collected on leaves from roadside Callistemon (bottlebrush) trees from Lahore, Pakistan, and on known sources of traffic-related particulates to assess the potential of first-order reversal curve (FORC) diagrams to discriminate between different sources of anthropogenic magnetic particles. Magnetic measurements on leaves indicate the presence of surface-oxidized magnetite spanning the superparamagnetic (<30 nm) to single domain (∼30– 70 nm) to vortex size range (∼70–700 nm). Fe-bearing particles are present both as discrete particles on the surface of larger mineral dust or carbonaceous particles and embedded within them, such that their aerodynamic sizes may be decoupled from their magnetic grain sizes. FORC diagrams of brake-pad residue specimens show a distinct combination of narrow central ridge, extending from 0 to 200 mT, and a low-coercivity, vertically spread signal, attributed to vortex and multi-vortex behavior of metallic Fe. This is in agreement with scanning electron microscopy results that show the presence of metallic as well as oxidized Fe. Exhaust-pipe residue samples display a more conventional “magnetite-like” signal comprising a lower coercivity central ridge (0–80 mT) and a tri-lobate signal attributed to vortex state and/or magnetostatic interactions. The FORC signatures of leaf samples combine aspects of both exhaust residue and brake-pad endmembers, suggesting that FORC fingerprints have the potential to identify and quantify the relative contributions from exhaust and non- exhaust (brake-wear) emissions. Such measurements may provide a cost-effective way to monitor the changing contribution; of future particulate emissions as the vehicle fleet is electrified over the coming years.
... Those with lower paramagnetic content have more pot-bellied shaped hysteresis curves (mostly σ hys >0.6), an effect which is not related to haematite content changes (see SI Fig 2d), but is likely related to slightly differing grain size populations of ferrimagnets (Tauxe et al., 1996;Tauxe et al., 2002). The domain state diagrams of Tauxe et al. (2002) and Lascu et al. (2010) in Figures 4c and 4d, also suggests MD behaviour dominates. Compared to trends seen in the ocean floor basalt compositions (TM60 and low Timagnetite trends in Fig. 4c), the data suggests that our samples largely contain low Ti-magnetites. ...
... The hysteresis data ofNiezabitowska et al. (2019a) from the Wenlock from N. Poland are also shown. d) the domain state plot ofLascu et al. (2010) showing the MD mixing relationships with various types of single domain (SD) particles (locations for 10, 20 and 30% SD material shown). Range of χ ARM /SIRM for magnetite inclusions in silicates from western European basement samples (excluding gneiss) fromHounslow and Morton (2004). ...
Article
Widespread marine red beds in the European Telychian (lower-Silurian) are one expression of an interval of unusually widespread oxic conditions in low palaeolatitude Silurian seas. This work examines in detail the geochemical and magnetic susceptibility record of cores from southern Poland, which also express the Telychian oxygenation event in grey-mudstones. The geochemical data provide an evaluation of redox conditions, palaeoweathering, sediment provenance, primary palaeoproductivity and upwelling. Sediment provenance is evaluated against possible sources on the East European Craton. The data suggest that the magnetic susceptibility is carried by both paramagnetic Fe-silicates and Fe-oxides. Magnetic data are supplemented by magnetic hysteresis and isothermal remanent magnetisations, and mineralogical data on selected samples. In Poland the oxygenation event is clearly expressed in larger Fe2O3 and magnetic susceptibility, caused by enhanced palaeoweathering, changes in sediment provenance and a redox influence on the preservation of Fe-oxides. A much briefer oxygenation event is detected in the upper Rhuddanian. Palaeoproductivity fluxes indicate that the Telychian oxygenation event was caused by a reduction in primary oceanic palaeoproductivity, possibly linked to a reduction in nutrient delivery to the margin of East European Craton, inferred to be caused by increased aridity. The increased aridity stimulated enhanced delivery of Fe-enriched aeolian dust from soils, generating a magnetic susceptibility and Fe2O3 expression of the Telychian oxygenation event.
... Often, pedogenic processes in modern soils lead to the production of strongly magnetic fine-grained magnetite and/or maghemite (referred to here as pedogenic magnetite; Maxbauer et al., 2016bMaxbauer et al., , 2017. Pedogenic magnetite dominates low-field magnetic properties, when present even in very low concentrations (Lascu et al., 2010). The fact that low-field magnetic susceptibility here is uncorrelated to ferrimagnetic susceptibility and ARM (both properties that are sensitive to the concentration of fine-grained magnetite/maghemite; see Fig. 11C) and shows a strong positive association with high-field susceptibility emphasizes that the overall concentration of ferrimagnetic minerals (including pedogenic magnetite) is exceptionally low or effectively zero (Fig. 11). ...
Article
Full-text available
Characteristics of floodplain strata document local, regional, and even global paleoenvironmental conditions. In this study, we examine floodplain deposition within the early Paleogene Wasatch Formation in the Piceance Creek Basin of northwestern Colorado, U.S.A., using sedimentology, environmental magnetism, and whole-rock geochemistry with the purpose to constrain the long-term influence of boundary conditions on basin-scale depositional patterns. Overbank strata represent a spectrum of floodplain environments from organic-rich swamps and poorly drained floodplains to well-drained, oxidized paleosols that are identifiable as distinct lithofacies. For each facies, there is a high degree of coherence between pedogenic features and magnetic properties that support drainage variation as a dominant control on floodplain development. We characterize the frequency of occurrence for depositional facies through the basin by integrating three new stratigraphic sections with fluvial and geochemical records from previous studies. In addition, we present a new subsidence analysis for the Piceance Creek Basin to aid in our understanding of temporal trends. Floodplain deposition shifts from poorly drained conditions to well-drained conditions from the Paleocene to early Eocene. This difference in long-term overbank deposition between the two epochs corresponds with an increase in basin subsidence rates. Moreover, the change in floodplain deposition is mimicked in other Laramide basins within the Western Interior of the United States, which suggests a coherent regional shift and driver. This shift correlates with a major phase of Laramide tectonism linked with accelerated basin subsidence rates, increased uplift rates, and higher paleoelevations in the Rocky Mountain region. We suggest these factors, potentially along with long-term global warming trends into the Early Eocene Climatic Optimum, led to greater trapping of sediment within Laramide basins during the Eocene and depressed water tables, which were reflected in floodplain characteristics. Surprisingly, the impact of the Paleocene–Eocene Thermal Maximum, a ~ 200 kyr global warming event, did not generate distinct changes in paleosol development apart from an abbreviated increase in the prevalence of crevasse splay events. This observation is coherent with coeval fluvial lithofacies indicative of peaked river discharge and enhanced river mobility, which may imply greater rainfall variability during the hyperthermal event.
... Numerous magnetic techniques have been developed to measure the bulk magnetic properties of sediments to provide information about the concentration, domain state (a measure of magnetic grain size), and mineralogy of magnetic particles in a sample (e.g., Verosub and Roberts, 1995;Evans and Heller, 2003;Lascu et al., 2010;Liu et al., 2012;Roberts et al., 2014;Zhao et al., 2017;Roberts et al., 2019). Mathematical unmixing methods have also been developed to identify these magnetic mineral components quantitatively based on their bulk magnetic properties (e.g., Heslop, 2015). ...
Article
Full-text available
Unambiguous magnetic mineral identification in sediments is a prerequisite for reconstructing paleomagnetic and paleoenvironmental information from environmental magnetic parameters. We studied a deep-sea surface sediment sample from the Clarion Fracture Zone region, central Pacific Ocean, by combining magnetic measurements and scanning and transmission electron microscopic analyses. Eight titanomagnetite and magnetite particle types are recognized based on comprehensive documentation of crystal morphology, size, spatial arrangements, and compositions, which are indicative of their corresponding origins. Type-1 particles are detrital titanomagnetites with micron- and submicron sizes and irregular and angular shapes. Type-2 and -3 particles are well-defined octahedral titanomagnetites with submicron and nanometer sizes, respectively, which are likely related to local hydrothermal and volcanic activity. Type-4 particles are nanometer-sized titanomagnetites hosted within silicates, while type-5 particles are typical dendrite-like titanomagnetites that likely resulted from exsolution within host silicates. Type-6 particles are single domain magnetite magnetofossils related to local magnetotactic bacterial activity. Type-7 particles are superparamagnetic magnetite aggregates, while Type-8 particles are defect-rich single crystals composed of many small regions. Electron microscopy and supervised magnetic unmixing reveal that type-1 to -5 titanomagnetite and magnetite particles are the dominant magnetic minerals. In contrast, the magnetic contribution of magnetite magnetofossils appears to be small. Our work demonstrates that incorporating electron microscopic data removes much of the ambiguity associated with magnetic mineralogical interpretations in traditional rock magnetic measurements.
... Paleomagnetic studies often exploit Curie temperature or thermal demagnetization of isothermal remanent magnetization (IRM) (Ando et al., 2001;Jiang et al., 2017;Lowrie, 1990;Oda & Suzuki, 2000). Other approaches make use of IRM curve analysis (e.g., Egli, 2003;Heslop & Dillon, 2007;Heslop et al., 2002;Kruiver et al., 2001;Maxbauer et al., 2016;Robertson & France, 1994;Zhao et al., 2018), alternating field (AF) demagnetization characteristics of isothermal and anhysteretic remanent magnetizations (ARMs) (Egli, 2004;Lagroix & Guyodo, 2017;Liu et al., 2002), a combination of hysteresis and ARM parameters (Lascu et al., 2010), and hysteresis loop unmixing . First-order reversal curve (FORC) unmixing has become possible (Lascu et al., 2015), and the improved algorithm of Harrison et al. (2018) enables solution of the linear mixing equation that facilitates identification and quantification of linearly additive magnetic components. ...
Article
Full-text available
Magnetite and hematite mixtures occur widely in nature. Magnetic unmixing of the signals recorded by these minerals can be important for assessing the origin of their respective paleomagnetic remanences and for extracting geological and paleoenvironmental information. However, unmixing magnetic signals from complex magnetite and hematite mixtures is difficult because of the weak magnetization and high coercivity of hematite. We assess here the relative effectiveness of first‐order reversal curve (FORC) and extended FORC‐type diagrams, FORC‐principal component analysis (PCA), isothermal remanent magnetization (IRM) curve decomposition, and PCA of remanent hysteretic curves for unmixing magnetic components in samples from the magnetically complex Inuyama red chert, Japan. We also further characterize the domain state and coercivity distributions of both magnetite and hematite with FORC‐PCA and IRM acquisition analysis in the red chert. We show that IRM curve decomposition can provide valuable component‐specific information linked to coercivity, while FORC‐PCA enables effective magnetic domain state identification. PCA of remanent hysteretic curves provides useful information about the most significant factors influencing remanence variations and subtle coercivity changes. To identify components in complex magnetite and hematite mixtures, we recommend PCA analysis of remanent hysteretic curves combined with FORC analysis of representative samples to identify domain states and coercivity distributions.
... Microbes capable of non-photosynthetically oxidizing Fe 2+ have been detected at the oxycline of ferruginous lakes by microscopic observation (Gorlenko et al., 1980). Magnetotactic bacteria, who perform nonmetabolic redox transformations of iron and are detectable through magnetic techniques, are in greatest abundance at the oxycline of Brownie Lake, and occur in the anoxic sediments (Lascu et al., 2010). Non-photosynthetic Fe 2+ -oxidizing and Fe 3+ -reducing bacteria, detected by 16S rRNA and culturing efforts, were more abundant than photosynthetic Fe 2+ -oxidizing bacteria within the chemocline of Lake Pavin (Berg et al., 2019;Lehours et al., 2009). ...
Article
Anoxic and iron-rich (ferruginous) conditions prevailed in the ocean under the low-oxygen atmosphere that occurred through most of the Archean Eon. While euxinic conditions (i.e. anoxic and hydrogen sulfide-rich waters) became more common in the Proterozoic, ferruginous conditions persisted in deep waters. Ferruginous ocean regions would have been a major biosphere and Earth surface reservoir through which elements passed through as part of their global biogeochemical cycles. Understanding key biological events, such as the rise of oxygen in the atmosphere, or even the transitions from ferruginous to euxinic or oxic conditions, requires understanding the biogeochemical processes occurring within ferruginous oceans, and their indicators in the rock record. Important analogs for transitions between ferruginous and oxic or euxinic conditions are paleoferruginous lakes; their sediments commonly host siderite and Ca-carbonates, which are important Precambrian records of the carbon cycling. Lakes that were ferruginous in the past, or euxinic lakes with cryptic iron cycling may also help understand transitions between ferruginous and euxinic conditions in shallow and mid-depth oceanic waters during the Proterozoic. Modern ferruginous meromictic lakes, which host diverse anaerobic microbial communities, are increasingly utilized as biogeochemical analogues for ancient ferruginous oceans. Such lakes are believed to be rare, but regional and geological factors indicate they may be more common than previously thought. While physical mixing processes in lakes and oceans are notably different, many chemical and biological processes are similar. The diversity of sizes, stratifications, and water chemistries in ferruginous lakes thus can be leveraged to explore biogeochemical controls in a range of marine systems: near-shore, off-shore, silled basins, or those dominated by terrestrial or hydrothermal element sources. Ferruginous systems, both extant and extinct, lacustrine and marine, host a continuum of biogeochemical processes that highlight the important role of iron in the evolution of Earth's surface environment.
Article
The Pliocene-Pleistocene transition is characterized by an abundance of Ice-Rafted Debris (IRD) in the North Atlantic basin. One of the regions affected by IRD during this period is the Gardar Drift, where the DSDP Leg 94 Hole 611A is located. This region received sediments from different sources during the glacial and interglacial intervals (e.g., Iceland and Greenland). We analyzed grain size and particle-size specific magnetic properties of sediments for their provenance characterization between ~2.64 and 2.52 Ma. Our results show that major proportion of bulk sediments during both glacial and interglacial periods were made up of basaltic-rich Icelandic sediments, whereas only during more intense glacial periods (Marine Isotope Stages 100 and 104), a small proportion of non-basaltic sand compositions were identified, possibly sourced from Greenland and other non-basaltic provenances. The non-basaltic sand fractions during the intense glacial periods were likely supplied as IRDs. In addition, a new level of coarse lithics (38 pcs. of >1 mm) composed of different rocks types (e.g., basalt, granite, granodiorite etc.) were identified in DSDP 611A Hole during the end of glacial MIS 104. The coarse lithic fragments showed distinctive magnetic properties than rest of the particle sizes and were classified as Iceberg-Rafted Debris (IBRD). Our results show that higher sand percentage was found during the intense glacial episodes, and their magnetic grain size analysis could help in distinguishing their provenance. We elaborate that particle size specific magnetic measurements of sand fractions could help in rapidly characterizing the glacial episodes in the subpolar North Atlantic.
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
-Introduction. List of Contents. Symbols and Abbreviations. Foreword. Acknowledgments. -1: Theory and Characteristics of the Mossbauer Effect. Theory of the Mossbauer Effect. Magnetism. Hyperfine Interactions. Relaxation Effects. Line Intensities. Diffusional Broadening. Mossbauer Spectroscopy in Mineralogy and Minerals Processing. Summary. -2: Mossbauer Instrumentation. General Considerations. Source and Reference Absorber. Conventional Spectrometers. Other Spectrometers. Sample Environment. Sample preparation. Summary. -3: Data Analysis and Interpretation. Fitting using Simple Lorentzians. Tests of Goodness of Fit. Convolution and Deconvolution. Thickness Effects and Recoilless Fraction. Fitting of Thickness Broadened Lines. Fitting of Lines Broadened by a Hyperfine Parameter Distribution. Magnetic Relaxation and Superparamagnetism. The Non Uniqueness Problem. Problems in Fitting and Interpretation. Presentation of Results. Summary. -4: Bulk and Clay Sized Phyllosilicates. 1:1 Minerals: Kaolinite. 2:1 Minerals: Illite. Montmorillonite. Nontronite. Summary. - 5: Iron Oxides and Oxyhydroxides. Anhydrous Oxides: Hematite. Magnetite. Maghemite. Oxyhydroxides: Goethite. Ferrihydrite. Summary. -6: Sediments. Terrestrial Sediments. Freshwater Sediments. Marine Sediments. Airborne Particles. Summary. -7: Soils and Clays. Soils. Clays. Summary. -8: Weathering. Silicate Weathering. Sulfide Weathering and Acid Mine Drainage. Summary. -9: Metastable Materials. Green Rusts. Fine Particle Magnetite. Anoxic Sediments. Fe2+ Sulfates Produced by Acid Mine Drainage. Summary. -10: Coal. Coal Characterization. Heat and Chemical Treatment. Hydroliquefaction. Summary. -11: ClayFiring. Individual Minerals: Kaolinite. Illite. Montmorillonite. Nontronite. Samples of Complex Mineralogy. Summary. -12: Mineral processing. Introduction. Iron Ores. Titanium Ores. Other Processing Operations Involving Iron. Gold. Summary. References. Index.
Chapter
This is an interdisciplinary and synoptic study of Equatorial Atlantic sediment formation in the Late Quaternary aimed at untangling the interlaced signatures of terrigenous and biogenous deposition and early diagenesis. It is based on a stratigraphic network of 16 gravity core records arranged along one meridional and three zonal transects (4°N, 0° and 4°S) crossing the Amazon and Sahara plumes as well as the Equatorial Divergence high productivity region. All newly introduced sediment sequences are collectively dated by their coherent CaCO3 content profiles and two available δ18O age models. To infer proxy records indicative of individual fluxes and processes, we analyze environmental magnetic parameters describing magnetite concentration, magnetic grain sizes and magnetic mineralogy along with CaCO3, Corg, Fe, Mn, Ba and color data. Diagenetically affected layers are identified by a newly introduced Fe/κ index. Reach and climatic variability of the major regional sedimentation systems is delimited from lithological patterns and glacial/interglacial accumulation rate averages. The most prominent regional trends are the N-S decrease in terrigenous accumulation and the Equatorial Divergence high in glacial Corg accumulation, which decays much faster south- than northwards. Glacial enrichments in Corg and proportional depletions in CaCO3, content appear to reflect sedimentary carbonate diagenesis more than lysoclinal oscillations and dominate temporal lithology changes. Suboxic iron mineral reduction is low at Ceara Rise and Sierra Leone Rise, but more intense on both flanks of the Mid-Atlantic ridge, where it occurs within organic rich layers deposited during oxygen isotope stages 6, 10 and 12, in particular at the terminations. To the equator, these zones reflect a full precessional rhythm with individual diagenesis peaks merging into broader magnetite-depleted zones. Rock magnetic and geochemical data show, that the depths of the Fe3+/Fe2+ redox boundary in the Equatorial Atlantic are not indicative of average productivity and were frequently shifted in the past. They are now located just above the topmost preserved productivity pulse. At 4°N, this organically enriched layer coincides with glacial stage 6, at 0° with glacial stage 2. Subsequent oxic and suboxic degradation of organic material entails stratigraphically coincident carbonate and magnetite losses opening new analytical perspectives.
Chapter
Magnetic mineral accumulation at the Ceará Rise has been studied with the aim to discriminate and reconstruct fluvial South American and eolian African terrigenous fluxes to the late Quaternary western Equatorial Atlantic. Seven sediment series recovered along two bathymetric transects were investigated with standard environmental magnetic techniques. Climatically controlled fluctuations in continental detrital discharge and marine biogenic carbonate fluxes strongly modulate the susceptibility records. Their coherent precessional and higher-frequent signal components could be used to establish a high-resolution age framework for these sediments. According to a partial susceptibility analysis, on average 79 % of the susceptibility signal originates from magnetite of different grain size, 13 % from hematite and 8 % from paramagnetic matrix compounds. In terms of absolute concentrations this implies that hematite is almost twenty times more abundant than magnetite, because of its orders of magnitude lower intrinsic susceptibility. The longitudinal gradients of their respective accumulation rates document a delivery from two major sources characterized by largely different magnetite to hematite ratios (about 1:12 versus 1:50). A mixing model of this scenario provided detailed insight into the past variability of the separate magnetic mineral fluxes and their most probable provenance. Overall about 56 % of hematite and 84 % of magnetite were transported in the Amazon fluvial load. Their accumulation is closely related to sea level changes, reaching highest (lowest) rates, when most South American shelf areas fell dry (were flooded) before and after Termination I and II. Hematite and magnetite of African provenance, 44 and 16 %, respectively, follow a distinctly different accumulation pattern with prominent maxima during cold intervals of glacial periods. By statistically linking these trace minerals to total lithogenic fluxes, we find that during the last 200 kyr, on average 79 % of total terrigenous material in the Ceará Rise area originates from South American sources in the Amazon River catchment, while African dust sources contributed 21 %.