Você está na página 1de 14

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Physics of the Earth and Planetary Interiors 214 (2013) 1–13

Contents lists available at SciVerse ScienceDirect

Physics of the Earth and Planetary Interiors


journal homepage: www.elsevier.com/locate/pepi

Role of spherical particles on magnetic field recording in sediments:


Experimental and numerical results
Dario Bilardello a,⇑, Josef Jezek b, Stuart A. Gilder a
a
Department of Earth and Environmental Sciences, LMU, Theresienstrasse 41, 80333 Munich, Germany
b
Institute of Applied Mathematics and Information Technologies, Faculty of Science, Charles University in Prague, Albertov 6, 128 43 Praha 2, Czech Republic

a r t i c l e i n f o a b s t r a c t

Article history: We report deposition experiments using spherical glass beads that possess remanent magnetizations
Received 15 May 2012 stemming from iron impurities. 15 g of glass beads with a well-characterized size distribution were
Received in revised form 5 September 2012 loaded in two different sets of tubes with diameters of 2.0 and 3.6 cm. Each tube contains identical col-
Accepted 31 October 2012
umn heights of de-ionized water, thereby allowing us to assess the effect of sediment concentration on
Available online 9 November 2012
Edited by Chris Jones
the results (352 versus 90 kg/m3 [g/l], respectively). The tubes were placed in magnetic fields of variable
inclination and intensity in a temperature-controlled environment. The full vector magnetization and
sediment accumulation rates were measured upon deposition times ranging from 10 min to 10 days.
Keywords:
Paleomagnetism
Experiments were run in triplicate to evaluate data reproducibility. Together with the lack of magnetic
Inclination shallowing interaction and the absence of clumping, the experiments elucidate an end-member scenario of how sed-
Deposition experiments iments acquire remanent magnetizations in the absence of flocculation. Our results show that inclination
Numerical models shallowing, in the range of 7–20° for field inclinations of 30° and 60°, is indeed possible with solely spher-
ical particles. More importantly, we observe a field dependence on the inclination error. Field dependence
on the moment acquisition and inclination error both exhibit non-linearity, which may complicate inter-
pretations of relative paleointensity data in paleomagnetic records. A newly developed numerical model,
whereby particle collision during settling combined with both rolling and slipping (translation) on the
substrate, is consistent with the experimental results.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction which is known as a post-depositional remanent magnetization


(pDRM) (see Tarling and Turner, 1999 and references therein).
Acquisition of remanent magnetization in sediments, called a Laboratory redeposition experiments reveal that the magneti-
depositional remanent magnetization (DRM), is typically described zation intensity of sediments grows in proportion to the strength
by spherical magnetic particles falling through a stagnant water of the applied field and that the net magnetization is orders of
column (Rees, 1961; Collinson, 1965; King and Rees, 1966; Stacey, magnitude lower than the saturation remanence (i.e., if all the par-
1972; Tauxe and Kent, 1984; Shive, 1985; von Dobeneck, 1996; ticle moments were parallel) (e.g., Barton et al., 1980; Tauxe and
Katari et al., 2000). Viewed in this way, the particle is subject to Kent, 1984). Several experiments demonstrate that the net effect
balanced inertial, viscous and magnetic torques, and spherical of a depositional remanent magnetization is to shallow the rema-
magnetic particles attain perfect alignment with the ambient field nent inclination in the rock (IR) with respect to the applied field
within seconds (Nagata, 1961; Collinson, 1965). The situation is inclination (IB) such that tan(IR) = f  tan(IB), where f is the flatten-
much more complicated in nature, where the sedimentation pro- ing factor (King, 1955; Løvlie and Torsvik, 1984; Tauxe and Kent,
cess spans a vast parameter space regarding particle size and shape 1984). Misalignment of declination is negligible. Two basic models
distributions, viscosity, pH and Eh of the fluid, etc. (Verosub, 1977). are used to explain inclination shallowing. In that of King (1955),
Contact forces between particles and Brownian motion also play a sediments are composed of spherical and platy particles: shallow-
role. Eventually the particles encounter the substrate, leading to ing depends on the relative contribution of the latter. Griffiths et al.
mechanical interaction and possibly experiencing shear from (1960) explained inclination shallowing by equal-sized spherical
bottom currents. Within the sediment column, bioturbation, dewa- particles rolling into depressions between grains lying on the sed-
tering, diagenesis and compaction can modify the magnetization, imentation plane. On the other hand, however, instances of natural
sediments yielding inclinations compatible with those predicted
⇑ Corresponding author. Current address: Instituto de Geociências, University of
from apparent polar wander paths and possessing the same incli-
São Paulo, Rua do Lago 562, 05508-080 São Paulo, Brazil. nations as lava flows, which are mostly immune to inclination
E-mail address: dario.bilardello@gmail.com (D. Bilardello). shallowing have also been reported (e.g., Opdyke, 1961).

0031-9201/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.pepi.2012.10.014
Author's personal copy

2 D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

Several experimental and theoretical studies on DRM acquisi- size distribution is well known. Numerical simulations are devel-
tion have focused on particle aggregation (flocculation) during set- oped to explain the results. We are particularly interested whether
tling, which affects the magnetic intensity and inclination recorded spherical particles can produce inclination shallowing by rolling on
by sediments based on the clay content, clay mineralogy and fluid the substrate when they settle. This necessitates a re-evaluation of
conductivity (Ellwood, 1979; Shcherbakov and Shcherbakova, the classical rolling spheres model of Griffiths et al. (1960). Our
1983; Lu et al., 1990; Deamer and Kodama, 1990; Sun and Kodama, experimental results and numerical models contradict the idea of
1992). For example, Sun and Kodama (1992) found that magnetic King (1955) that spherical particles accurately record the ambient
grains attach to clay minerals by either electrostatic or van der field direction,
Waals forces. The magnetic grains become incorporated into the
clay fabric of the sedimentary rock and rotate with the clay parti-
cles during post-depositional compaction (Arason and Levi, 1990; 2. Materials and methods
Katari and Bloxhamm, 2001). Van Vreumingen, 1993a,b) showed
that flocculation varies as a function of salinity of the sediment 2.1. Glass beads
suspension, and Tauxe et al. (2006) found a non-linear field depen-
dence on remanence for certain floc sizes. Mitra and Tauxe (2009) Fig. 1a shows a scanning electron microscope image of the solid
explored remanence acquisition as a function of applied field and glass spheres (Potters Europe, spheriglass 5000) used in this study.
floc size distributions. Their work helped explain discrepancies in The image attests that almost all the particles are spherical in
relative paleointensity and inclination data, highlighting the com- shape. Laser particle analysis (Coulter) was used to measure the
plex nature of DRM acquisition with respect to different sedimen- grain-size spectra of the beads (Fig. 1b). Five independent runs,
tary environments (variable salinity, mineralogy, organic matter made without using dispersing agents or ultrasound, are highly
content, etc. (see also Katari and Tauxe, 2000). Shcherbakov and reproducible and reveal no evidence for clumping or clustering of
Sycheva (2008, 2010) recognized that more than seven parameters the particles. Particle radii range between a fraction of a microme-
are needed to describe the magnetization acquisition of sediments. ter to 11.4 lm, showing a sharp peak for the smallest grain sizes,
This multi-parametric control on DRM hinders relative paleofield then decreases almost exponentially with increasing radii. Overall,
intensity estimates by redeposition methods since laboratory con- 10% of the particles have radii <0.48 lm, 25% are <1 lm, 50% are
ditions do not reproduce the natural environment. <3 lm, 75% are <6.8 lm and 90% are <7.7 lm. Company specifica-
Other workers have addressed the question of lock-in depth of tions list a median radius of 1.7–3.5 lm, with 90% of the spheres
the magnetic signal in sediments (Kent, 1973; Tucker, 1980; Bleil having radii between 0.3 and 9.7 lm. We accepted the company’s
and von Dobeneck, 1999; Roberts and Winklhofer, 2004). For reported density of 2.5 g/cm3 without independent verification.
example, Løvlie (1974, 1976) attributed the lag between changes Experiments on the beads indicate they contain impurities that
in ambient field and lock-in of the magnetization to post-deposi- carry a magnetic remanence. Hysteresis loops, backfield curves and
tional alignment of the magnetic grains, with consolidation-rate magnetization versus temperature curves were measured with a
significantly influencing a sediment’s magnetic intensity. Irving Petersen Instruments, variable field translation balance (dwell field
and Major (1964) showed that sediments first deposited in a null 30 mT, dwell time 1 s, ramp slope [heating and cooling] 40 °C/min).
field, and then subjected to an applied field, accurately recorded The magnetic moment upon heating undergoes a change in slope
the field direction. While post-depositional effects seem to play a at 580 °C followed by a major drop around 770 °C, suggesting the
role in influencing the final orientation of the magnetic vector, presence of both magnetite and iron, respectively (Fig. 2a). Mag-
other laboratory experiments and numerical models find a limited netic intensities during cooling lie systematically below those of
influence (Verosub et al., 1979; Shcherbakov and Shcherbakova, heating, signifying a net loss of magnetic moment during heating.
1987; Katari et al., 2000). Repeat heating–cooling cycles in 100 °C intervals indicate that al-
Despite significant efforts to understand the genesis of a detrital ready below 400 °C the cooling curve is lower than the heating
remanent magnetization, complete knowledge of the underlying curve. Our interpretation is that the glass contains only pure Fe
principles are still lacking. We thus initiated a series of experi- as a remanence carrier, but that Fe partly oxidizes into magnetite
ments to focus on a particular aspect of the problem—namely the during heating. Acid rinsed, non-annealed glass beads were used
contribution of solely spherical grains in the absence of floccula- in the experiments.
tion. Our experiments do not intend to simulate a natural detrital Fig. 2b shows a typical hysteresis loop of the material measured
remanent magnetization acquisition, rather to unravel one specific at room temperature. Hysteresis loops determined on four inde-
factor that contributes to it. Hence, we carried out deposition pendent samples exhibit a high degree of reproducibility, with hys-
experiments with synthetic spherical magnetic particles whose teresis parameters lying within the pseudo-single domain field on

Fig. 1. (A) SEM image of the glass beads used in the deposition experiments highlighting the size distribution and the predominantly spherical shape of the particles. (B)
Grain-size distribution of the particles measured with a laser counter (no de-flocculants or ultrasound used). The grain-size bins sum to 100%.
Author's personal copy

D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13 3

Fig. 2. Summary of magnetic measurements conducted on the glass beads. (A) Thermo-remanent curves for the heating (red) and cooling (blue) cycle between room
temperature and 800 °C measured with a Petersen Instruments, variable field translation balance. 580 and 770 °C correspond to Curie temperatures of magnetite and pure
iron, respectively. (B) A magnetic hysteresis loop measured with a Petersen Instruments, variable field translation balance. (C) Hysteresis parameters from four independent
measurements lie within the pseudo-single domain field on a Day et al. (1977) plot, although this is not strictly valid for iron metal. (D) A first order reversal curve diagram
measured with a Princeton Instruments, vibrating sample magnetometer. The diagram is representative of interacting single domain particles (Roberts et al., 2000). (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

a Day et al. (1977) plot (Fig. 2c). Such plots were specifically devel- remanence of a sample over several hours after exposure.
oped for titanomagnetite and might not be applicable for iron. The Remanence decreased by 2% in the first 19 h, after which it
average saturation magnetization (Ms) and remanent saturation remained stable. The sample was then placed in a 50 lT field,
magnetization (Mrs) are 3.46  103 A m2/kg (8.64 A/m) and orthogonal to the magnetization direction. No change in moment
5.05  104 A m2/kg (1.26 A/m), respectively, while the bulk coer- or direction was detected 17 h later, thus lending confidence that
cive force (Bc) and coercivity of remanence (Bcr) are 12 and the material is suitable for the deposition experiments.
30 mT, respectively. First order reversal curves, measured with a
Princeton Instruments vibrating sample magnetometer at the 2.2. Experimental set-up and procedure
Institute for Rock Magnetism (Minneapolis) (settings: field range
300 mT, moment range 2  106 A m2, averaging time 800  Helmholtz coils, 1 m on each side, were constructed to gener-
103 s, analyzed using the code of Michael Winklhofer [version ate a stable magnetic environment up to 100 lT in any direction
July, 2010]), indicate that the beads contain interacting single do- over the volume in which the experiments were performed. A
main particles (Fig. 2d). In assigning a magnetic moment to the fluxgate magnetometer probe fixed within the coils constantly
beads, we recall that the remanence ratio (Mrs/Ms) of a random monitored the field. Non-magnetic, cylindrical, flat-bottomed
assemblage of non-interacting single domain particles with uniax- borosilicate tubes with diameters of 2.0 and 3.6 cm (height =
ial anisotropy is 0.5 (Dunlop and Özdemir, 1997). Mrs/Ms. of the 30 cm) were made in triplicate for each tube size. Tubes of both
glass beads is 0.146, or 3.4 times less than the expected value, diameters were prepared with 15.000 ± 0.001 g of dry glass beads;
which we attribute to magnetic interaction. Assuming that each distilled water was added such that the height of the water column
bead contains a proportional amount of magnetic material by vol- was 20 cm. The sediment mass is thus identical for both tube-sizes,
ume, and that the magnetic remanence of each bead corresponds yet the 2 cm diameter tubes have 3.9 times higher sediment con-
to the saturation value, a magnetization of 8.64/3.4 = 2.5 A/m can centrations than the 3.6 cm diameter tubes. In this way, the paral-
be assigned to each individual bead. Moreover, after immersing lel experiments can help identify potential wall and concentration
the grains in water and observing them under a microscope, no effects (i.e., hitting of the particles against the tube walls or mutual
clumping or aligning of the grains was detected. Thus, magnetic hitting of the particles as they settle).
interactions revealed in Fig. 2d likely occurs within each grain The experiments were carried out by placing the tubes in an
(intra-grain) and not between grains (inter-grain), further support- ultrasound bath while shaking vigorously to achieve full separation
ing that 2.5 A/m is the most reasonable value for the magnetization and suspension of the sediment. The tubes were placed in the
value of the beads. controlled magnetic environment within the Helmholtz coils at a
We tested for magnetic viscosity by imposing a 600 mT field on specified field intensity and inclination. Temperature fluctuations
the beads to saturate their magnetizations and then measured the that could potentially change the moment or water viscosity were
Author's personal copy

4 D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

minimized by immersing the tubes in a temperature-controlled continues, albeit much slower, even after 10 days of settling time.
water tank (32 °C) within the Helmholtz coils. For example, the No systematic correlation is found between the slope of the mo-
experiments reported here were carried out for over a year’s time ment acquisition and B.
span where the temperature in the room fluctuated at least 10 °C. Initial loading of the tubes was in July 2010. The last experi-
Water viscosity decreases 20% from 18 to 28 °C (Kestin et al., ments performed were at IB = 0° under B of 50 (March, 2011), 25
1978), which would alter the falling time of the spheres in the (April, 2011) and then 100 lT (May–June, 2011). During the exper-
tubes depending on when the experiments were done. iments at 25 lT, 2.0 cm tube #3 displayed much higher sediment
After a given settling time, the tubes were removed from the thicknesses and lower moments than the other two tubes (Table 1,
Helmholtz coils and placed into a vertical, 3-axis, 2G Enterprises, data in gray). In the subsequent experiments conducted at IB = 0°,
755 cryogenic magnetometer to measure the magnetic vector. B = 100 lT, all 2.0 cm tubes began to mimic this behavior, so the
Typical measured magnetic moments range from 7  107 to experiments at 100 lT were discontinued. We think the glass
4  106 A m2 (Table 1), which are well above the noise of the beads started welding together, which would lower the magnetiza-
magnetometer over the measurement time (1–3  1011 A m2). tion and increase the pore space of the deposited material, thereby
The horizontal plane of the tubes was positioned in the magnetom- increasing the thickness of deposited sediment. This highlights the
eter such that the middle of the fully deposited sediment was at usefulness of measuring the sediment accumulation rate as an
the point of maximum signal strength on the magnetometer sen- independent control.
sors. Settling intervals under a given field inclination and intensity Seen at a given field intensity, the magnetization acquisition
were normally 10, 20, 30, 40, 50 and 60 min, then 2, 3, 4, 5, 18, 68 curves are compatible, independent of field inclination, and the
and 240 h. After each measurement step, the tubes were vigorously moments are systematically higher with greater applied fields at
shaken, and then placed back into the Helmholtz coils for a differ- a given time step. All curves show an initial steep rise in moment
ent settling time. In this manner, each time interval can be consid- before flattening out after 100 min (2.0 cm tubes) and 300 min
ered as an independent experiment, also meaning that settling (3.6 cm tubes). Initial moment accumulation is thus faster in the
steps may be performed in any order and repeated if necessary. 2.0 cm tubes, which makes sense, as the concentration is higher.
The fact that measurements for different time intervals yield rather Remanent inclination (IR) is fairly constant during the entire pro-
smooth curves attests to the uniformity of the underlying pro- gression of the experiments, with only one exception at IB = 60°,
cesses and increases the reliability of the results. B = 100 lT in the 3.6 cm tubes. This is a first-order observation
Three tubes of each diameter were run contemporaneously to since the process leading to inclination shallowing occurs immedi-
examine data reproducibility. The thickness of the sedimentary ately in the deposition process. It is readily apparent that for IB of
column was measured after each time step to determine the vol- 30° and 60°, IR varies as a function of B. One exception is for IB of
ume of deposited sediment, which could be compared to the accu- 30° in the 2.0 cm tubes where IR is 23° for B of both 50 and
mulated moment. This measurement also quantifies sediment 100 lT; however, IR is 15° when B = 25 lT. Extrapolating more spe-
packing and monitors the consistency of the experiments. cific trends out of the data becomes speculative as certain features
are not coherent at given IB and B among both tube diameters; e.g.,
the greatest variability in IR (and sediment height) occurs when IR
3. Results from the deposition experiments is 0° in the small tubes yet higher scatter is not observed in the
large tubes.
Figs. 3 and 4 show the results of all experiments conducted with
the 2.0 and 3.6 cm diameter tubes, respectively (data in Table 1—
figures and subsequent analyses exclude values in gray). Table 2 4. Discussion
summarizes the results for each experiment by calculating the
mean of the last three experimental steps for each of the three 4.1. Magnetic moment
tubes, and then taking the average and single standard deviation
of those three mean values. In some cases the moments or inclina- Fig. 5a plots the mean magnetic moment of the last three mea-
tions vary slightly in time until the last deposition step (Figs. 3 and surement steps for each tube, averaged over the three settling
4, Table 1); however, taking only the last step instead of the aver- tubes, versus applied field intensity (B). To a first-order, the field-
age of the last three steps yields virtually identical results. In gen- dependence on the acquired magnetization is evident in all exper-
eral, the data show a relatively good degree of reproducibility for iments, in conformity with virtually all experiments performed to
all experiments carried out at a given applied field strength (B) date (e.g., King, 1955; Tauxe and Kent, 1984; Quidelleur et al.,
and inclination (IB). That the curves of magnetization acquisition 1995).
over time match the sedimentation rates lends credence to the re- Tauxe and Kent (1984) performed deposition experiments from
sults (Figs. 3 and 4); this will be explored in further detail in the disaggregated red beds, which contained both magnetite and
numerical modeling section below. hematite. They placed approximately 10 gm of dry sediment in a
We calculated the slope of the average moments from the last 3.5 cm-diameter plastic tube and added water to make a 15 cm
three time steps reported in Table 2 (not shown), and found it to high slurry. This sediment to water concentration is comparable
be is positive in 10 of 11 cases when considering data where the to our 3.6 cm diameter tubes. The tube was sealed, agitated and
coefficient of determination (R2) from the least squares linear fit then placed in a controlled field for about 5 h. Both our results
exceeds 0.85. This could suggest that the DRM acquisition process and those of Tauxe and Kent (1984) show more variability at the
is not completed after 10 days of deposition. Alternatively, the highest applied field (100 lT) over the range of inclinations
change in slope between short (ca. <300 min) and long (ca. (Fig. 5a), yet we find the two data sets are relatively compatible.
>300 min) time scales could reflect a different acquisition process Tauxe and Kent (1984) inferred a linear relationship between ap-
with a DRM at short time scales and a pDRM over long time scales. plied field and magnetization. Fig. 5a offers an alternative interpre-
A similar analysis for the height data leads to a similar conclusion tation that the increase in moment is steeper from 25 to 50 lT than
as the moment data: the slope of the average heights from the last from 50 to 100 lT, which suggests the evolution can be better de-
three time steps (data from Table 2) is positive in 7 of 8 cases scribed by a non-linear (logarithmic) function. On the other hand,
where (R2) exceeds 0.85. Hence separating the data into two sepa- the small diameter tubes with higher sediment concentrations give
rate regimes over different time scales seems unlikely, e.g., DRM more linear trends than the large diameter tubes. This could sug-
Table 1
Moment, inclination and thickness data measured in this study. Data are subdivided into three measurement tubes of 2.0 and 3.6 cm diameter with field inclinations of 60°, 30° and 0° subdivided into applied field intensities of 25, 50
and 100 lT. Data in gray were not considered for analysis (see text for details); experiments in horizontal fields of 100 lT were terminated after two time steps.

Time 2.0 cm diameter tubes 3.6 cm diameter tubes


(min)
Tube 1 Tube 2 Tube 3 Tube 1 Tube 2 Tube 3
Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height
(A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm)
Laboratory applied field and inclination values: 25 lT, 60°
10 0.90 42.8 7.1 0.97 38.5 8.5 1.18 43.2 10.3 0.67 37.8 3.1 0.76 38.5 3.3 0.92 35.3 4.0
20 1.22 40.5 9.8 1.19 38.9 11.3 1.38 45.7 13.6 1.16 38.9 4.6 1.23 38.4 4.8 1.28 36.1 5.2
30 1.60 42.9 15.5 1.52 38.4 16.5 1.64 41.7 18.7 1.48 37.9 5.3 1.37 41.6 5.9 1.51 38.0 6.4
60 1.92 41.6 24.8 1.77 38.7 25.0 1.84 42.6 26.7 1.90 38.3 7.5 1.80 38.5 7.6 1.85 37.3 7.8
120 2.10 40.7 31.4 1.91 37.6 31.5 2.02 37.7 32.6 2.11 39.6 8.8 1.92 39.3 8.6 2.01 38.4 9.2
180 2.22 45.1 32.8 2.08 42.1 33.2 1.99 40.9 34.2 2.08 40.2 9.2 2.04 38.9 9.3 2.14 37.6 9.6
300 2.14 41.2 34.1 2.00 39.2 34.9 2.15 41.5 35.9 2.19 38.2 9.9 1.99 39.1 10.1 2.20 38.3 10.0
1020 2.22 42.6 35.5 2.06 39.4 36.1 2.15 38.6 37.3 2.29 43.0 9.1 2.13 41.8 9.3 2.27 41.0 10.2
4080 2.23 39.3 36.5 2.05 37.7 36.6 2.13 36.7 38.3 2.26 45.3 9.2 2.14 42.6 9.2 2.26 42.0 10.2
14400 2.26 43.0 36.3 2.11 42.8 36.3 2.24 37.3 37.6 2.39 40.0 10.2 2.32 41.7 10.5 2.49 40.6 10.2
Laboratory applied field and inclination values: 50 lT, 60°
10 1.32 50.4 7.5 1.10 49.1 8.2 1.10 50.6 8.4 1.29 44.9 3.4 1.18 43.7 3.2 1.24 43.2 3.2
20 2.20 48.4 13.0 2.09 49.9 13.3 2.07 50.5 14.3 2.18 45.5 5.4 1.92 45.9 5.7 2.13 43.8 5.3
30 2.59 50.3 17.7 2.29 48.9 18.0 2.55 51.0 18.6 2.50 45.9 6.5 2.34 44.4 6.4 2.50 44.1 6.2
60 3.00 47.5 25.6 2.95 48.3 25.4 2.81 49.2 26.3 3.00 44.2 7.7 2.71 44.1 8.5 2.98 43.3 8.3
120 3.13 48.0 30.9 2.95 48.0 31.2 2.96 48.4 31.9 3.20 45.5 8.7 2.90 45.1 9.0 3.22 44.9 9.0
180 3.20 48.2 33.1 2.98 48.3 33.4 3.03 49.3 34.1 3.26 46.5 9.0 3.01 45.0 9.4 3.27 44.9 9.2
240 3.22 49.4 34.2 3.00 47.5 34.5 3.05 49.8 35.1 3.29 46.9 8.8 3.06 46.5 9.4 3.34 46.0 9.5
1020 3.26 49.5 36.7 3.01 48.1 37.3 3.09 48.9 37.5 3.35 47.7 8.7 3.17 45.9 9.8 3.45 46.2 9.9
4080 3.30 49.9 36.8 3.08 48.0 37.4 3.15 48.8 37.4 3.40 46.6 9.4 3.17 45.9 10.3 3.47 45.9 10.3
14400 3.35 47.8 36.9 3.13 47.9 37.2 3.18 49.1 37.5 3.47 47.9 9.9 3.37 45.7 10.4 3.57 47.1 10.4
Laboratory applied field and inclination values: 100 lT, 60°
10 1.38 50.1 6.8 1.06 51.1 7.0 1.30 53.1 7.2 1.33 43.6 3.5 0.98 43.1 3.1 1.15 40.0 3.2
20 2.63 52.4 12.3 2.19 53.9 12.6 2.56 55.7 13.7 1.90 42.1 5.1 1.85 45.6 5.1 1.92 42.9 5.3
30 3.20 53.5 16.9 2.92 53.2 16.6 3.04 54.9 18.2 2.78 46.3 6.5 2.25 44.0 5.8 2.88 47.2 6.4
60 3.86 52.5 25.2 3.68 52.2 25.3 3.78 52.1 26.3 3.41 49.6 7.5 3.25 48.4 8.0 3.40 48.5 7.9
120 4.24 52.4 30.8 3.90 51.7 31.0 4.04 51.5 31.8 3.74 50.6 9.0 3.54 50.2 9.3 3.74 49.6 9.4
180 4.23 52.6 32.8 3.90 51.9 32.9 4.05 51.7 33.3 3.85 50.8 9.6 3.66 50.3 9.7 3.96 49.9 10.1
Author's personal copy

300 4.29 52.2 33.7 4.03 51.1 34.5 4.19 51.5 35.1 4.20 50.5 9.9 3.86 50.7 9.8 4.31 50.1 10.2
1020 4.58 53.0 36.1 4.08 52.4 36.3 4.17 53.5 36.5 4.45 51.1 10.1 3.96 50.5 10.1 4.47 50.1 10.5
4080 4.46 52.9 36.2 4.04 51.3 36.8 4.27 51.5 36.9 4.46 51.8 10.0 4.02 51.3 10.3 4.45 50.5 10.7
14400 4.70 49.8 36.6 4.27 51.1 36.9 4.37 52.5 37.1 4.69 51.2 10.6 4.10 50.4 10.4 4.64 49.7 10.9
D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

Laboratory applied field and inclination values: 25 lT, 30°


10 0.69 15.9 4.6 0.83 11.6 7.0 1.15 14.8 9.9 0.69 17.2 2.9 0.82 17.4 3.2 1.02 16.5 3.6
20 1.76 14.3 15.6 1.76 13.8 16.5 1.82 15.1 20.1 1.47 18.1 4.9 1.44 17.8 5.0 1.59 16.4 5.5
30 2.03 14.8 18.6 1.91 12.7 19.8 1.93 16.8 24.0 1.74 18.4 5.5 1.70 17.4 5.9 1.78 16.8 6.5
60 2.30 15.1 25.7 2.13 13.1 26.5 2.19 14.7 29.9 2.10 19.6 7.6 1.96 18.9 7.5 2.12 17.1 7.6
120 2.46 15.3 31.3 2.30 13.3 31.5 2.24 13.6 34.5 2.36 18.5 7.8 2.16 19.0 8.5 2.38 17.5 8.9
180 2.51 15.6 33.5 2.34 15.6 33.5 2.39 13.2 35.7 2.42 18.8 8.1 2.31 17.8 8.5 2.46 17.7 9.2
300 2.58 17.5 35.6 2.40 14.0 35.2 2.39 14.8 37.5 2.51 18.3 9.5 2.36 18.3 9.1 2.53 17.6 9.9
1020 2.60 15.9 36.3 2.43 13.7 36.8 2.47 13.6 38.9 2.43 19.1 10.6 2.45 18.1 9.9 2.56 17.8 10.7
4080 2.59 16.3 37.3 2.41 14.0 37.6 2.58 12.8 39.3 2.16 17.3 10.6 2.45 17.3 9.9 2.56 17.7 10.5
14400 2.61 21.0 36.5 2.46 17.9 36.9 2.67 17.6 37.4 2.44 17.2 10.5 2.34 16.5 10.6 2.62 17.7 10.9
Laboratory applied field and inclination values: 50 lT, 30°
10 1.12 21.3 5.9 1.34 22.2 7.5 1.78 22.2 9.0 1.51 18.9 3.7 1.06 17.9 3.3 0.86 17.1 2.6
20 2.35 22.4 11.8 2.25 22.2 13.2 2.60 23.1 15.6 2.17 19.5 5.6 1.79 18.9 5.1 1.71 19.1 4.9
30 3.03 24.0 18.6 2.84 22.8 21.3 3.09 22.0 22.8 2.34 19.9 6.3 2.25 18.5 6.7 2.40 19.0 6.5

(continued on next page)


5
6
Table 1 (continued)

Time 2.0 cm diameter tubes 3.6 cm diameter tubes


(min)
Tube 1 Tube 2 Tube 3 Tube 1 Tube 2 Tube 3
Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height Moment Inclination Height
(A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm) (A m2  10-6) (°) (mm)
60 3.19 22.6 23.0 2.87 21.9 23.6 3.15 21.6 25.3 2.81 19.5 7.7 2.59 18.5 7.9 2.81 19.1 8.4
120 3.46 24.7 30.2 3.14 22.2 30.7 3.39 21.3 31.4 3.06 19.8 9.0 2.78 19.7 9.2 3.03 19.7 9.1
180 3.51 23.9 32.5 3.20 22.2 32.7 3.42 20.9 33.4 3.21 20.2 9.0 2.84 19.8 9.5 3.12 19.6 9.4
240 3.51 23.1 33.9 3.20 22.3 33.9 3.46 21.6 34.0 3.05 19.6 9.7 2.86 19.0 9.9 3.19 19.4 10.1
300 3.54 24.2 34.7 3.25 22.3 34.1 3.45 22.2 35.0 3.11 19.0 9.9 2.90 18.4 10.3 3.11 18.6 10.4
1080 3.57 23.6 35.5 3.26 22.5 35.6 3.49 21.9 36.2 3.32 21.2 9.6 3.04 19.6 9.9 3.30 20.7 10.3
4080 3.66 24.2 36.2 3.34 23.9 36.4 3.61 23.8 36.8 3.49 20.6 9.7 3.23 20.3 9.8 3.57 21.0 10.7
14400 3.72 22.3 36.5 3.47 23.8 36.4 3.79 21.9 36.8 3.52 21.2 9.6 3.30 19.9 9.8 3.59 21.4 10.2
Laboratory applied field and inclination values: 100 lT, 30°
10 1.27 21.8 5.8 1.63 21.7 7.3 2.10 22.1 8.8 1.25 21.3 2.9 1.36 21.8 3.4 1.77 22.1 3.9
20 2.87 20.9 10.9 2.76 21.5 12.1 3.26 21.0 14.5 2.11 22.1 4.7 2.02 21.8 4.4 2.26 22.8 5.2
30 3.89 20.4 16.8 3.59 20.6 17.4 4.04 20.6 20.5 2.56 22.4 5.9 2.41 22.0 6.2 2.58 22.5 6.2
60 4.39 23.7 23.3 3.95 23.6 24.1 4.37 22.8 26.0 3.16 23.4 7.8 2.99 21.8 7.9 3.24 22.6 7.5
120 4.71 23.7 30.6 4.27 23.3 30.6 4.60 22.4 31.8 3.53 22.9 8.5 3.24 22.7 8.8 3.52 22.8 8.5
180 4.80 23.9 33.2 4.34 23.3 33.5 4.71 22.8 33.7 3.60 23.5 9.1 3.34 22.9 9.1 3.67 23.0 9.2
1080 4.86 23.8 35.8 4.44 23.2 35.7 4.77 22.7 36.3 3.78 23.2 9.5 3.45 23.5 10.0 3.84 23.1 10.2
4080 5.01 22.9 36.7 4.56 22.7 36.3 4.88 21.5 37.2 3.80 23.8 9.5 3.52 23.4 9.7 3.83 23.9 10.0
14400 4.98 23.8 36.2 4.53 23.0 36.5 4.81 22.4 37.1 3.88 23.5 9.6 3.54 23.1 9.7 3.92 23.5 9.9
Laboratory applied field and inclination values: 25 lT, 0°
10 0.74 2.2 5.7 0.63 0.0 8.2 0.35 1.2 89.8 0.48 2.8 2.7 0.48 1.6 2.8 0.97 1.2 3.5
20 1.72 3.0 15.0 1.11 5.7 17.9 0.44 1.1 74.1 0.97 1.6 4.2 0.93 1.0 4.3 1.48 1.0 4.9
30 1.97 1.9 18.5 1.43 5.2 21.2 0.36 0.7 62.5 1.26 0.5 5.4 1.17 0.2 5.4 1.75 0.9 5.8
60 2.22 1.9 26.9 1.85 3.2 27.9 0.41 0.2 54.5 1.43 1.4 7.3 1.56 0.0 7.2 2.13 0.9 7.7
120 2.41 2.0 31.8 1.63 5.5 33.9 0.54 0.8 48.1 2.06 2.1 8.2 1.81 1.7 7.9 2.33 0.9 9.0
180 2.36 0.7 34.0 1.23 5.1 38.3 0.37 0.1 57.2 2.30 2.1 8.3 2.14 1.6 8.3 2.41 1.3 9.1
300 2.54 0.1 36.3 1.48 7.6 39.3 0.65 0.6 50.0 2.41 1.3 8.7 2.28 0.6 8.8 2.49 0.9 9.5
1080 2.67 0.7 37.6 2.34 1.9 38.2 1.31 8.0 43.1 2.56 2.2 8.7 2.47 0.2 9.3 2.61 1.0 9.9
4080 2.70 0.7 37.7 2.45 2.7 37.6 1.98 9.7 40.9 2.55 1.0 9.5 2.40 1.5 9.7 2.70 0.1 9.9
14400 2.66 1.5 37.1 2.29 3.3 38.6 1.44 8.9 42.3 2.49 2.1 9.5 2.22 0.0 10.4 2.18 0.5 12.3
Laboratory applied field and inclination values: 50 lT, 0°
10 1.39 1.3 6.9 1.47 0.1 8.4 1.23 3.8 14.1 0.95 1.5 2.9 1.07 0.1 3.2 1.29 0.6 3.8
20 2.25 0.4 11.7 2.18 0.9 12.9 2.00 1.1 18.2 1.69 1.1 4.5 1.66 1.0 4.9 1.92 0.0 5.1
Author's personal copy

30 3.05 0.8 19.8 2.84 1.2 20.8 2.02 3.6 26.2 2.14 1.1 5.7 2.06 0.8 6.0 2.26 0.3 6.1
60 3.43 0.3 26.7 3.22 0.5 27.4 2.70 0.7 30.8 2.58 0.4 6.5 2.48 0.1 7.4 2.66 0.3 7.5
120 3.57 0.6 31.6 3.36 0.7 31.8 2.73 5.7 35.5 2.80 0.2 7.9 2.69 0.3 8.5 2.87 0.3 8.6
D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

180 3.67 0.7 34.1 3.41 1.8 33.6 2.90 3.6 36.9 2.86 0.5 8.4 2.74 0.9 8.7 2.98 0.3 9.2
300 3.71 0.1 35.5 3.45 0.8 35.6 3.05 2.0 37.9 2.96 1.5 8.7 2.84 0.4 9.5 3.07 0.1 9.9
1080 3.74 0.2 36.9 3.46 0.3 36.8 2.93 2.8 40.6 3.03 1.7 9.3 2.94 0.4 9.5 3.15 0.2 10.2
4080 3.73 0.3 37.9 3.47 0.6 38.1 2.91 5.6 41.4 3.02 0.4 9.0 2.93 0.1 9.6 3.17 0.1 10.1
14400 3.46 9.1 38.7 3.50 6.4 39.3 3.22 1.5 42.0 3.07 0.7 9.4 2.92 0.8 9.5 3.22 0.4 9.7
Laboratory applied field and inclination values: 100 lT, 0°
4080 1.96 1.3 44.2 0.84 1.8 51.2 3.30 0.6 9.5 2.93 0.2 10.1 3.51 0.1 10.1
14400 4.45 0.1 39.2 3.16 5.5 41.6 0.64 2.3 65.3 3.25 0.5 9.5 3.11 0.5 9.5 3.50 0.3 10.0
Author's personal copy

D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13 7

Fig. 3. Data from the deposition experiments conducted in 2.0 cm diameter tubes. Results are subdivided by total magnetic moment, recorded inclination and thickness
(height) of the sedimentary column. Experiments were carried out in three different applied field inclinations (IB) as indicated in each panel and in three different field
intensities as indicated in color. Circles, squares and triangles are data from the different tubes; thick lines are the average results from the three tubes.

Fig. 4. Data from the deposition experiments conducted in 3.6 cm diameter tubes. Results are subdivided by total magnetic moment, recorded inclination and thickness
(height) of the sedimentary column. Experiments were carried out in three different applied field inclinations (IB) as indicated in each panel and in three different field
intensities as indicated in color. Circles, squares and triangles are data from the different tubes; thick lines are the average results from the three tubes.
Author's personal copy

8 D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

Table 2 on inclination shallowing recognized that sediments composed of


Summary of the experimental and numerical results. From the data in Table 1, we ellipsoidal-shaped particles should experience more inclination
calculated the mean of the last three experimental steps for each of the three tubes,
and then took the average and single standard deviation (s.d.) of those three mean
shallowing than those composed of spheres (King, 1955; Griffiths
values. Abbreviations are: B, applied magnetic field intensity; M, magnetic moment in et al., 1960; King and Rees, 1966). Like Johnson et al. (1948) and
A m2; IR, remanent inclination (in °); IB, applied field inclination (in °). King (1955) we also observe a field dependence of the inclination
B (lT) M ± s.d. (106) IR ± s.d. IR/IB Thickness ± s.d.
error. One possibility is that this may be unique to spherical parti-
cles, for which rotation is proportional to field intensity, unlike for
2.0 cm diameter tubes
ellipsoids. Another explanation will be proposed in the following
IB 60° chapter.
25 2.16 ± 0.08 39.7 ± 2.1 0.7 37.9 ± 0.9
50 3.17 ± 0.12 48.7 ± 0.6 0.8 38.4 ± 0.4
100 4.33 ± 0.23 52.0 ± 0.5 0.9 37.8 ± 0.3 5. Comparison to numerical models
IB 30°
25 2.54 ± 0.09 15.9 ± 1.6 0.5 38.6 ± 1.0 We now assess the different phases of particle settling and DRM
50 3.55 ± 0.16 23.1 ± 0.5 0.8 37.5 ± 0.3
100 4.76 ± 0.23 22.9 ± 0.7 0.8 37.6 ± 0.4
acquisition numerically. We first estimate how fast particles settle,
and whether the measured sediment height is consistent with the
IB 0°
25 2.52 ± 0.22 1.1 ± 2.2 – 40.4 ± 2.5
numerical simulations of settling. Sediment height is a valuable
50 3.38 ± 0.32 0.9 ± 2.9 – 40.3 ± 2.0 independent control because, when comparing with the evolution
100 – – – – in magnetic moment, it confirms that a substantial fraction of
beads must fall individually, rather than in flocs. The settling rate
3.6 cm diameter tubes
also dictates how much time the particles have to orient in the field
IB 60° during their fall in the column or to be randomized by Brownian
25 2.28 ± 0.08 42.0 ± 0.8 0.7 11.2 ± 0.4
motion and mutual interactions. Through numerical simulation
50 3.38 ± 0.13 46.5 ± 0.8 0.8 11.3 ± 0.5
100 4.36 ± 0.29 50.7 ± 0.6 0.8 11.8 ± 0.3 of this process we show that the beads in the suspension are not
fully oriented by the field when they reach the substrate and that
IB 30°
25 2.45 ± 0.12 17.6 ± 0.3 0.6 11.9 ± 0.3 their inclinations can already be shallower than the field inclina-
50 3.37 ± 0.16 20.7 ± 0.6 0.7 11.4 ± 0.4 tion before hitting the bottom. Finally, using the model of Jezek
100 3.73 ± 0.20 23.4 ± 0.1 0.8 11.2 ± 0.2 et al. (2012), we establish that rolling and slipping on the substrate
IB 0° contributes significantly to inclination shallowing. Mutual magne-
25 2.46 ± 0.09 0.9 ± 0.7 – 11.3 ± 0.7 tostatic interactions of the spheres in our experiments are negligi-
50 3.05 ± 0.13 0.3 ± 0.5 – 11.0 ± 0.4 ble due to their low magnetizations and clumping plays no
100 3.27 ± 0.24 0.1 ± 0.5 – 11.2 ± 0.3
significant role in the shallowing process.
Experiments 3.6 diameter tubes Roll–Slip model
B (lT) IR f M Mrel IO f Mrel (100%)
5.1. Settling
50 (IB 60°) 46.5 0.61 3.38 0.22 48.1 0.64 0.48
50 (IB 30°) 20.7 0.65 3.37 0.22 20.4 0.64 0.60 At small Reynolds numbers, the falling velocity of spherical par-
50 (IB 0°) 00.3 – 3.05 0.20 00.4 – 0.66
ticles is given by the Stokes formula

2 ðq  qw Þgr 2
v ðrÞ ¼ ð1Þ
gest that the magnetization acquisition process is concentration 9 l
dependent. The fact that the experiments of Tauxe and Kent
(1984), which had concentrations comparable to our large tubes, where q and qw are densities of the particle and water, r is particle
yield more linear trends would suggest otherwise. However, one radius, l is viscosity and g is gravitational acceleration. The Stokes
must exercise caution since the hematite platelets present in their velocity formula (Eq. (1)) should not be used for submicron particles
experiments probably behave differently than spheres, and yield since they undergo Brownian motion (Stacey, 1972; Collinson,
different trends. Furthermore, because their experiments were 1965; Allen, 2003), as discussed in detail below. The time needed
conducted for shorter durations, the magnetic moments may not for a particle to fall from top to bottom of a water column of height
have been completely saturated. hT by this velocity is

tb ðrÞ ¼ hT =v ðrÞ ð2Þ


4.2. Recorded inclination
At tb, all particles of a given radius should be settled. Fig. 6 shows tb
The results from our experiments unambiguously show a field as a function of particle radius (solid line). To take into account the
dependence on the inclination error (Fig. 5b). Like the moment concentration of a sedimenting suspension, the velocity v can be
data, the change in inclination error with applied field can also corrected by a multiplicative factor f/(t) = (1  /(t))n where /(t) is
be globally described by a logarithmic-type function. Fig. 5c plots particle volume fraction and n  5 (Richardson and Zaki, 1954;
the recorded inclination (IR) against the applied field inclination Snabre et al., 2009):
(IB); the dashed lines correspond to different flattening factors (f)
v c ðr; tÞ ¼ v ðrÞf/ ðtÞ ð3Þ
ranging between 0.5 and 0.8. The range of sediment concentrations
(i.e., tube sizes) used in our study does not appear to systematically The settling times computed using the corrected velocity vc and
influence the amount of inclination shallowing, consistent with the initial concentration are shown in Fig. 6 for both 3.6 and 2.0 cm
findings of Blow and Hamilton (1978) and Barton et al. (1980). In tubes (dashed and dotted lines, respectively). Concentration de-
fact, the variability in the trends observed in Fig. 5 is unexpected, creases during settling and therefore a more realistic estimate of
yet must be considered inherent in the experiments since those the time at which all particles of a given radius are settled lies be-
trends were reproducible in three different tubes. tween the corrected curves (for t < 300 min) and Stokes’ curve. We
Fig. 5b also plots the data of Tauxe and Kent (1984), where no use the time tb defined by Eq. (2) as a reference value. The mea-
field dependence of the inclination error is observed. Early work sured sediment height becomes nearly constant for times
Author's personal copy

D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13 9

Fig. 5. Summary of the deposition results reported in Table 2. (A) Magnetic moment versus applied field intensity B (tube size and field inclination are specified). Results from
Tauxe and Kent (1984) are plotted for comparison (TK, moments scaled proportionately to this study). (B) Inclination error (IR/IB) as a function of B for experiments performed
in field inclinations (IB) of 30° and 60° (TK from Tauxe and Kent, 1984). (C) Applied field inclination (IB) versus measured field inclination (IR) for experiments carried out in 25,
50 and 100 lT. Large open and small solid symbols refer to 3.6 and 2.0 cm tubes, respectively. Dashed lines represent equal flattening factors [f = tan(IR)/tan(IB)], with f
ranging from 0.5 (more shallowing) to 0.8 (less shallowing).

V S ðtÞ
HS ðtÞ ¼ ð5Þ
Sfpack

where fpack is the packing ratio (the ratio of the volume of spheres to
the volume they occupy). Fig. 7 compares several estimates of HS(t)
obtained by numerical simulation [Eqs. (4) and (5)] with the sedi-
ment height observed experimentally. We use fpack = 0.55, which
corresponds to the porosity of random loose packing, consistent
with wet-packed porosities of synthetic sands with lognormal size
distributions (Weltje and Alberts, 2011). We assume there is negli-
gible compaction in the initial stages of sedimentation; initial pack-
ing would be lower if flocculation occurs. In Fig. 7, the curve
computed for fpack = 0.55 (Eq. (5)) almost reaches the average mea-

Fig. 6. Settling time (tb) for particles of different radii: solid line is the curve
according to Eq. (2); dashed and dotted lines correspond to the large (3.6 cm) and
small (2.0 cm) tubes, respectively, after correcting for particle concentration
(Richardson and Zaki, 1954).

>300 min, which according to Eq. (2) is approximately the time of


settling of particles having r  1.5 lm. We can thus assume that
the smallest particles never settle and remain in the column.
The total volume of settled particles increases in time as an
integral
Z t Z
V S ðtÞ ¼ S v c ðr; tÞf ðr; tÞdrdt ð4Þ
0 r Fig. 7. Theoretical settling curves compared to experimental measurements.
Experimental data are shown by solid points (blue for large and red for small
where S is the surface of the bottom and f(r,t) is the size distribution tubes), white circles are the mean values. Thick black lines show how the
of the sedimenting particles. Due to different settling velocities of sedimentation of individual beads should proceed according to Stokes’ velocity
particles of different radii, the size distribution of the particles in corrected by the instantaneous concentration and random loose packing (packing
the column changes over time leading to a fining upward of sedi- ratio fpack = 0.55). The shaded area is defined by an interval of fpack from 0.50 to 0.55.
Dashed lines were computed by allowing spheres of radii r < 4 lm to fall in flocs of
ment in the deposited material. Sedimenting spheres also create size 5 lm. Thick solid blue or red lines correspond to shortened floc-existence
pore spaces as they settle. The theoretical height of sediment in a times (see text for details). (For interpretation of the references to color in this
tube is therefore figure legend, the reader is referred to the web version of this article.)
Author's personal copy

10 D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

sured heights for the 3.6 cm tubes and is visibly lower for the 2.0 cm
tubes. The area between this curve and the curve for 10% lower
packing (fpack = 0.50) includes the mean values, except for the first
time step in the 3.6 cm tubes and the first three time steps for the
2.0 cm tubes.
We also tested the possibility that particles fall in flocs that
could be inherited from imperfect disaggregation of the sediment
from the previous experiment and/or from clumping of particles
during the fall. This may particularly occur in the 2.0 cm tubes,
which possess 3.9 times higher concentrations than the 3.6 cm
tubes. If we hypothesize, for example, that all particles of radii
r < 4 lm fall in flocs of 5 lm radii (we use fpack = 0.55 and Eq. (1)
with a corresponding change of radius and density) we obtain a
curve that fits the mean heights for the initial time steps
(620 min) but overshoots the points at higher time steps (dashed
lines in Fig. 7). A better agreement can be attained by limiting
Fig. 8. Comparison of the role of Brownian motion (characteristic time, sB), field
the existence of flocs to a short interval at the beginning of the
alignment (characteristic time sm) and settling (tb) for the conditions in this study.
experiments. By allowing flocs to exist for 10 min, and then to set- Particles with radii <2 lm are influenced by Brownian motion, above 3 lm,
tle as individual particles, we obtain a very good agreement with alignment by the magnetic field prevails.
the average measured sedimentation heights. We also obtain a rea-
sonably good fit when we let only one-half the number of particles
of radii r < 4 lm fall in flocs for 20 min and then sediment as indi- 7–16% of the magnetization of beads. The characteristic aligning
vidual spheres. However, further decreasing the number of parti- time sm for these flocs would be between 2 and 17 min, depending
cles in flocs and allowing those flocs to exist over longer times on the field strength. Before reaching the bottom, such flocs may be
(e.g., 1/5 of particles in aggregates for the first 60 min) yields preferentially oriented parallel to the field direction—especially if
poorer fits. These numerical experiments show that we cannot re- they were inherited since the moments of the individual grains
ject the possibility for the existence of flocs in our experiments, would be aligned. Due to their rugged surface, the flocs may either
especially in the 2.0 cm tubes. Nevertheless, a substantial fraction be immobilized when they hit the surface and preserve the pre-
of the beads must fall individually. This is supported by the fact ferred orientation, or they may roll as a whole. This effect would
that the accumulation of magnetic moment in the 3.6 cm tubes explain the faster moment accumulation and the larger variability
mimics well the height (and volume) accumulation curves. On in inclination observed in the 2.0 cm tubes.
the other hand, the moment accumulates faster than the height From the characteristic times of the individual beads, one ex-
in the 2.0 cm tubes, which could be caused by falling flocs. This ef- pects that they are well oriented with the field direction when
fect will be assessed below. approaching the bottom. However, the beads will interact during
their fall. Larger spheres that fall faster will collide with smaller
beads below them, and both will be deflected. A larger particle of
5.2. Magnetization in the column radius R may hit on average n number of smaller particles of radius
r per second, given as:
Rotation of spherical particles by torque from the magnetic field
is expressed as n ¼ pðR þ rÞ2 ðv ðRÞ  v ðrÞÞcðrÞ ð6Þ

d# mV B 1 Davis (1992) and Zhao and Davis (2002) studied the situation
¼ sin # ¼  sin # ð5Þ when a faster sphere falls past another in low Reynolds number
dt 6l sm
fluids. They considered two cases: one of rigid body rotation where
where # is angle between the particle’s magnetic moment and the spheres rotate together before splitting, and a second of mutual
field direction, mV is the particle’s magnetization and B is the field rolling with possible slipping after the maximum friction force is
intensity (Nagata, 1961). The characteristic aligning time (sm) lasts reached. The second, roll-slip, model provided a better fit to their
seconds to minutes (Nagata, 1961; Shcherbakov and Shcherbakova, experimental data and will therefore be used here. In this model,
1983; Jezek and Gilder, 2006). sm corresponding to our experiments the larger (L) and smaller (S) particles rotate by different angular
is 18, 36 and 72 s for B of 100, 50 and 25 lT, respectively. These velocities whose ratio is
characteristic times can be compared to the falling time through xL b1 þ c1 minð1; cot h= cot hS Þ
the tube (tb) and to the Brownian relaxation time, sB = 3 Vl/kT (De- ¼ ð7Þ
xS b2 þ c2 minð1; cot h= cot hS Þ
bye, 1929), which indicates how fast an aligned system becomes
randomized in a zero field. When sB > sm, particle rotation into where b1, b2, c1, and c2 are given by means of the two-sphere mobil-
the magnetic field direction is significantly disturbed. The transition ity functions of Jeffrey and Onishi (1984), and h is the angle between
between Brownian and magnetic field dominated rotation occurs the line of centers of the two spheres and the vertical, i.e., the
with beads of r  2 lm at 100 lT and r  3 lm at 25 lT (Fig. 8). Par- instantaneous slope. hS is a threshold angle. For angles h < hS,
ticles have enough time to be aligned by the field when tb > sm for spheres roll without slipping; slipping occurs when h P hS. Zhao
all radii, except during the initial stages when the larger particles and Davis (2002) found that realistic values of hS are 15–30°, which
close to the bottom have insufficient time to become fully oriented. is also supported by experimental and theoretical evaluation of
If we permit the presence of flocs, as indicated in the previous spheres rolling down an inclined plane (Smart et al., 1993; Galvin
discussion of settling, their orientation with the field can also be et al., 2001; Zhao et al., 2002).
roughly assessed by Eq. (5). For example, 5 lm-radii flocs with A numerical model of the magnetization in the column was
about 50% porosity will take 60 min to settle (tb). If the flocs were constructed as follows. We consider a multi-particle system com-
pre-formed due to the incomplete disaggregation of the substrate posed of particles of radii r = 3, . . . , 11 lm (particles smaller than
material from the previous experiment, their magnetizations 3 lm are not considered as they are influenced by Brownian mo-
would correspond to that observed experimentally, being about tion). The initial particle size distribution corresponds to our
Author's personal copy

D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13 11

experiments, and then the size distribution evolves in time accord- the substrate. This in turn should contribute to lower moments
ing to Eq. (4). At each time step, particles are oriented by the mag- and shallower inclinations upon final deposition. Note that, be-
netic field (Eq. (5)) and disoriented by collision. The number of hits cause the mutual rotation of colliding spheres proceeds around
is given by Eq. (6). Each interaction causes both particles to rotate horizontal rotation axes, the nature of inclination shallowing in
around a horizontal axes. The larger particle can hit the smaller one the column is similar to the Griffiths et al. (1960) model of inclina-
at any position on its upper hemisphere (determined randomly) tion shallowing.
that in turn defines the angle h, the azimuth of the rotation axis,
and the amount of rotation (Eq. (7)). We use hS = 25° and the ratio 5.3. Rolling and slipping on the substrate
of slipping to rolling 0.5. This means that for h < 25° (when a larger
particle hits a small one close to its top), particles fully rotate, and King (1955) assumed that spherical particles deposited on a
for h > 25° (when a larger particle hits a small one more to its side), substrate will correctly record the field inclination, whereas Grif-
particles half rotate and half slip. fiths et al. (1960) showed that spherical particles hitting the bot-
This model provides an estimate for the magnetization of the tom roll equally in all directions, thus producing inclination
suspension in the column before reaching the substrate. Results shallowing. However, the model of Griffiths et al. (1960) is limited
of the simulations (Fig. 9) indicate that (1) the particles are not by not-knowing a representative value of the angle of rotation –
fully oriented with the magnetic field, (2) the inclination is shal- this ambiguity persists until today; moreover, they considered only
lower than expected, and (3) there exists a significant field depen- populations of equi-dimensional spheres. To overcome these pit-
dency on the magnetization and inclination. Because equation (6) falls we constructed an algorithm that treats random sedimenta-
also accounts for concentration, there is a clear difference between tion with rolling and slipping spherical particles (Jezek et al.,
the two tube-sizes, with the 2.0 cm tubes showing lower moments. 2012). Here we briefly describe the model and its implications to
It is probable that mutual interactions in the column, especially in our experiments.
the 2.0 cm tubes with higher concentrations, leads to rolling and We generate spherical particles in random x, y and z coordinates
slipping processes away from the ideal case described by the above above a substrate and allow them to fall until they hit previously
model. But even if rolling and slipping only partially occurs, the re- deposited spheres (Fig. 10). Depending on the geometry of the
sults from the 2.0 cm tubes will still be more influenced, yielding spheres on the substrate and where the descending particle
lower net moments and shallower inclinations when arriving at encounters the substrate, several scenarios are possible concerning

Fig. 9. Results from numerical simulations of the suspension above the substrate. Thick lines are for 3.6 cm tubes, thin lines for 2.0 cm tubes. Green, blue and red correspond
to fields 25, 50 and 100 lT, respectively. Field inclination in this case was 60°. Mr is the magnetization relative to the saturation value; IO is the observed inclination.

Fig. 10. Random sedimentation of spherical particles of variable sizes.


Author's personal copy

12 D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13

how/if the particle will roll according to the geometry of the bot- The main results can be summarized as follows: (1) For two differ-
tom layer. The end result occurs when the sedimenting sphere ar- ent diameters of settling tubes, the detrital remanent magnetiza-
rives at a stable position (stops moving), supported by at least tion increases then plateaus at a nearly constant level as the
three stationary spheres, and becomes embedded into the sub- settling time increases. (2) With few exceptions, the measured sed-
strate. By repeatedly solving the equations of mutual contact of iment inclination is shallower than the ambient field inclination
two or more spheres, we track the sphere’s path and record the from the outset of the experiment, i.e. at the shortest settling
magnetic vector upon cessation of motion. This is done for tens times. No discernible change in inclination is observed as the set-
of thousands of particles. tling time increases. Inclination shallowing is similar to what is ob-
Not only do the particles roll, but they can also slide. As de- served for natural sediments, with a flattening factor of about
scribed above we define slipping by means of the threshold angle f = 0.6; e.g., sediments composed entirely of spherical particles
hS and the ratio of slipping to rolling when the angle hS is exceeded can produce inclination shallowing, on the order observed in nat-
(see Jezek et al. (2012) for a detailed description). In the simula- ure. (3) The amount of inclination shallowing generally follows
tions we used hS = 25° and the ratio of slipping to rolling equals the King (1955) tangent-tangent relationship with respect to the
0.5, as in Section 5.2. magnetic field inclination. Moreover, the degree of inclination
Repeated simulations of this model show that a characteristic shallowing depends on the field strength, with less shallowing
value of the inclination shallowing factor (f) is on the order of 0.6 for stronger magnetic fields.
(simulations with equal-sized spheres produced a standard devia- Through numerical modeling, we improved upon the model of
tion of 0.02). Interestingly, f  0.6 also holds valid for spherical Griffiths et al. (1960) by calculating the amount of rotation pro-
populations with non-equal radii, including uniform (i.e., same duced by mutual interaction of spherical particles in the column
number of spheres of given r) and exponential size-distributions. and their rolling on a surface created by randomly falling particles.
The latter can be taken as an end member for our experiments. A roll/slip model (Zhao and Davis, 2002) of particle interaction in
Note that the model is sequential in that each sphere stops rolling the column and on the sediment–water interface accounts for low-
before another one falls. There is no influence of the sphere’s er moments and inclinations than otherwise predicted in the ab-
momentum. A falling or depositing sphere on the substrate cannot sence of particle interaction. The importance of rolling is damped
displace another sphere, cause avalanches, etc. as field intensity increases, which in turn leads to less inclination
In the lower part of Table 2 we compare the average measured shallowing and heightened magnetization. It is possible that the
inclination, IR, and relative moment (Mrel) in the large tubes for flocculation process restricts rolling, hence limits the field depen-
IB = 60° with those predicted by the roll-slip model. In these simu- dency on inclination shallowing. Compaction ill explains the
lations all beads were perfectly oriented parallel to the field before relationship between volume of accumulated material and acquisi-
reaching the substrate and then subjected to rolling and slipping. tion of remanence in our experiments. On the other hand, the roll/
Therefore the results reflect the separate role of rolling and sliding slip formulation predicts a difference in moment acquisition and
on the substrate. The inclinations are well predicted, but the model inclination shallowing between the two tube diameters, as the
yields higher relative moments than the experimental values. concentration and likelihood to collide is higher in tubes with
However, it was shown above that particles are not fully aligned smaller diameters. However, the concentration in the small tubes
with the field when they reach the bottom, so the moment given may exceed the applicable limit of the roll/slip model as multiple
in Table 2 must be lowered accordingly, which will be different interactions restrict rotational freedom. If true, experiments using
for small and large tubes. Although the moment and inclination lower concentrations (e.g., 3 g of sediment in the 3.6 cm diameter
of the suspension above the substrate are influenced more in the tubes) should find higher (relative) moments and lower amounts
small tubes than in the large ones (Fig. 9), proportional to particle of inclination shallowing.
concentration, the situation is reversed on the substrate—the high- Future deposition experiments should monitor accumulation
er concentration of the small tubes restricts rolling and sliding rates and use multiple tubes to assess data reproducibility. Tem-
more than in the large tubes. Both processes are complementary peratures should be held constant during the entire span of the
(counteracting), which can partly explain small differences ob- experiments. The relative contribution of different factors or con-
served between results from the large and small tubes. tributions from the processes involved (rolling, flocculating, hit-
To summarize the numerical models, we find that one factor con- ting, etc.) could be determined experimentally by varying the
trolling the observed field dependence on both moment and inclina- concentration and the size-distribution in future experiments.
tion originates while the particles are in the water column due to Experiments employing strictly oblate particles may become feasi-
mutual interaction. An additional process is the reorientation by ble in the near future with the emergence of techniques to grow
magnetic torque of particles lying on the surface before they are bur- hematite grains with narrow size distributions and provide an
ied by newly falling particles. During slipping, the spheres can reori- alternative end-member to the inclination shallowing problem.
ent to the field direction in response to the field. Furthermore, the
proportion of particles controlled by Brownian motion is also Acknowledgements
field-dependent (Fig. 8). Finally, stable recorded moments and incli-
nations over longer time scales show that realignment of particles in Funding from the national science agencies of Germany and the
the substrate seems to be inhibited or very restricted by grain–grain Czech Republic supported this project (Grants DFG GI712/3-1,
contact; pDRM effects, if present, are hardly discernible. GACR 205/09/J028 and MSM0021620855). Jessica Till kindly
performed the FORC measurements at the Institute for Rock
Magnetism. We thank Ramon Egli and Michael Winklhofer for
6. Conclusions stimulating discussions. We also thank Mark J. Dekkers and an
anonymous reviewer for comments that improved the manuscript.
Our experimental set up was designed to examine the remanent
magnetization acquisition process during sedimentation of solely
spherical particles in the absence of flocculation. These experi- References
ments and numerical models represent an end-member of a myr-
Allen, T., 2003. Powder Sampling and Particle Size Determination. Elsevier (600pp).
iad of potential natural scenarios, yet they highlight how spherical Arason, P., Levi, S., 1990. Compaction and inclination shallowing in deep sea
particles, in particular, should behave in the natural environment. sediments from the Pacific Ocean. J. Geophys. Res. 95, 4501–4510.
Author's personal copy

D. Bilardello et al. / Physics of the Earth and Planetary Interiors 214 (2013) 1–13 13

Barton, C.E., McElhinny, M.W., Edwards, D.J., 1980. Laboratory studies of Nagata, T., 1961. Rock Magnetism. Maruzen, Tokyo (348pp).
depositional DRM. Geophys. J. R. Astron. Soc. 61, 355–377. Opdyke, N.D., 1961. The paleomagnetism of the New Jersey Triassic: A field study of
Bleil, U., von Dobeneck, T., 1999. Geomagnetic events and relative paleointensity the inclination error in red sediments. J. Geophys. Res. 66, 1941–1949.
records – clues to high resolution paleomagnetic chronostratigraphies of late Quidelleur, X., Valet, J.P., LeGoff, M., Bouldoire, X., 1995. Field dependence on
quaternary marine sediments. In: Fischer, G., Wefer, G. (Eds.), Use of Proxies in magnetization of laboratory-redeposited deep-sea sediments: first results.
Paleoceanography: Examples from the South Atlantic. Springer, Berlin Earth Planet. Sci. Lett. 133, 311–325.
Heidelberg, pp. 635–654. Rees, A.I., 1961. The effect of water currents on the magnetic remanence and
Blow, R.A., Hamilton, N., 1978. Effect of compaction on the acquisition of a detrital anisotropy of susceptibility of some sediments. Geophys. J. R. Astron. Soc. 5,
remanent magnetization in fine-grained sediments. Geophys. J. R. Astron. Soc. 235–251.
52, 13–23. Richardson, J.F., Zaki, W.N., 1954. Sedimentation and fluidisation: part 1. Trans. Inst.
Collinson, D.W., 1965. DRM in sediments. J. Geophys. Res. 70, 4663–4668. Chem. Eng. 32, 35–53.
Davis, R.H., 1992. Effects of surface roughness on a sphere sedimenting through a Roberts, A., Winklhofer, M., 2004. Why are geomagnetic excursions not always
dilute suspension neutrally buoyant spheres. Phys. Fluids A 4, 2607–2619. recorded in sediments? Constraints from post-depositional remanent
Day, R., Fuller, M., Schmidt, V.A., 1977. Hysteresis properties of titanomagnetites: magnetization lock-in modeling. Earth Planet. Sci. Lett. 227, 345–359.
Grain-size and compositional dependence. Phys. Earth Planet. Int. 13, 260–267. Roberts, A.P., Pike, C.R., Verosub, K.L., 2000. First-order reversal curve diagrams: a
Deamer, G., Kodama, K.P., 1990. Compaction-induced inclination shallowing in new tool for characterizing the magnetic properties of natural samples. J.
synthetic and natural clay-rich sediments. J. Geophys. Res. B 95, 4511–4529. Geophys. Res. 105 (29), 461–475.
Debye, P., 1929. Polar Molecules. Dover, 172pp. Shcherbakov, V., Shcherbakova, V., 1983. On the theory of depositional remanent
Dunlop, D.J., Özdemir, Ö., 1997. Rock Magnetism: Fundamentals and Frontiers. magnetization in sedimentary rocks. Geophys. Surv. 5, 369–380.
Cambridge Studies in Magnetism, Cambridge University Press, Cambridge, 573 Shcherbakov, V., Shcherbakova, V., 1987. On the physics of acquisition of post-
pp. depositional remanent magnetization. Phys. Earth Planet. Inter. 46, 64–70.
Ellwood, B., 1979. Particle Flocculation: One possible control on the magnetization Shcherbakov, V., Sycheva, N.K., 2008. Flocculation mechanism of the acquisition of
of deep-sea sediments. Geophys. Res. Lett. 6, 237–240. remanent magnetization by sedimentary rocks. Phys. Solid Earth 44, 804–815.
Galvin, K.P., Zhao, Y., Davis, R.H., 2001. Time-averaged hydrodynamic roughness of a Shcherbakov, V., Sycheva, N.K., 2010. On the mechanism of formation of
noncolloidal sphere in low Reynolds number motion down an inclined plane. depositional remanent magnetization. Geochem. Geophys. Geosyst. 11, 2.
Phys. Fluids 13, 3108–3120. http://dx.doi.org/10.1029/2009GC002830.
Griffiths, D.H., King, R.F., Rees, A.I., Wright, A.E., 1960. Remanent magnetism of some Shive, P.N., 1985. Alignment of magnetic grains in fluids. Earth Planet. Sci. Lett. 72,
recent varved sediments. Proc. R. Soc. A 256, 359–383. 117–124.
Irving, E., Major, A., 1964. Post-depositional remanent magnetization in a synthetic Smart, J.R., Beimfohr, S., Leighton, D.T., 1993. Measurement of the translational and
sediment. Sedimentology 3, 135–143. rotational velocities of a noncolloidal sphere rolling down a smooth inclined
Jeffrey, D.J., Onishi, Y., 1984. Calculation of the resistance and mobility functions for plane at low Reynolds number. Phys. Fluids A5, 13–24.
two unequal rigid spheres in low-Reynolds-number flow. J. Fluid Mech. 139, Snabre, P., Pouligny, B., Metayer, C., Nadal, F., 2009. Size segregation and particle
261–290. velocity fluctuations in settling concentrated suspensions. Rheol. Acta 48, 855–
Jezek, J., Gilder, S.A., 2006. Competition of magnetic and hydrodynamic forces on 870.
ellipsoidal particles under shear: influence of the Earth’s magnetic field on Stacey, F.D., 1972. On the role of Brownian motion in the control of detrital
particle alignment in viscous media. J. Geophys. Res. 111, B12S123. http:// remanent magnetization in sediments. Pure Appl. Geophys. 98, 139–145.
dx.doi.org/10.1029/2006JB004541. Sun, W.W., Kodama, K.P., 1992. Magnetic anisotropy, scanning electron microscopy,
Jezek, J., Gilder, S.A., Bilardello, D., 2012. Numerical simulation of inclination and X-ray pole figure goniometry study of inclination shallowing in a
shallowing by rolling and slipping of spherical particles. Comput. Geosci.. compacting clay-rich sediment. J. Geophys. Res. 97, 9599–9615.
http://dx.doi.org/10.1016/j.cageo.2012.06.013. Tarling, D.H., Turner, P., 1999. Palaeomagnetism and Diagnesis in Sediments, vol.
Johnson, E.A., Murphy, T., Torreson, O.W., 1948. Pre-history of the Earth’s magnetic 151. Geol. Soc. London, Spec. Publ., 306pp.
field. Terr. Magn. Atm. Elect. 53, 349–372. Tauxe, L., Kent, D.V., 1984. Properties of a detrital remanence carried by hematite
Katari, K., Bloxhamm, J., 2001. Effects of sediment aggregate size on DRM intensity: from study of modern river deposits and laboratory redeposition experiments.
a new theory. Earth Planet. Sci. Lett. 186, 113–122. Geophys. J. R. Astron. Soc. 77, 543–561.
Katari, K., Tauxe, L., 2000. Effects of pH and salinity on the intensity of Tauxe, L., Steindorf, J.L., Harris, A., 2006. Depositional remanent magnetization:
magnetization in redeposited sediments. Earth Planet. Sci. Lett. 181, 489–496. toward an improved theoretical and experimental foundation. Earth Planet. Sci.
Katari, K., Tauxe, L., King, J., 2000. A reassessment of post-depositional remanent Lett. 244, 515–529.
magnetism: preliminary experiments with natural sediments. Earth Planet. Sci. Tucker, P., 1980. A grain mobility model of post-depositional realignment. Geophys.
Lett. 183, 147–160. J. R. Astron. Soc. 63, 149–163.
Kent, D.V., 1973. Post-depositional remanent magnetisation in deep-sea sediment. Van Vreumingen, M.J., 1993a. The magnetization intensity of some artificial
Nature 246, 32–34. suspensions while flocculating in a magnetic field. Geophys. J. Int. 114, 601–
Kestin, J., Sokolov, M., Wakeham, W.A., 1978. Viscosity of liquid water in range 8 606.
to 150 °C. J. Phys. Chem. Ref. Data 7 (3), 941–948. Van Vreumingen, M.J., 1993b. The influence of salinity and flocculation upon the
King, R.F., 1955. The remanent magnetism of artificially deposited sediments. Mon. acquisition of remanent magnetization in some artificial sediments. Geophys. J.
Notic. R. Astron. Soc. Geophys. Suppl. 7, 115–134. Int. 114, 607–614.
King, R.F., Rees, A.I., 1966. Detrital magnetizationin sediments: an examination of Verosub, K.L., 1977. Depositional and postdepositional processes in the
some theoretical models. J. Geophys. Res. 71, 561–571. magnetization of sediments. Rev. Geophys. Space Phys. 15 (2), 129–143.
Løvlie, R., 1974. Post-depositional remanent magnetization in a re-deposited deep- Verosub, K.L., Ensley, R.A., Ulrick, J.S., 1979. The role of water content in the
sea sediment. Earth Planet. Sci. Lett. 21, 315–320. magnetization of sediments. Geophys. Res. Lett. 6, 226–228.
Løvlie, R., 1976. The intensity pattern of post-depositional remanence acquired in Von Dobeneck, T., 1996. A systematic analysis of natural magnetic mineral
some marine sediments deposited during a reversal of the external magnetic assemblages based on modelling hysteresis loops with coercivity-related
field. Earth Planet. Sci. Lett. 30 (2), 209–214. hyperbolic basis functions. Geophys. J. Int. 124, 675–694.
Løvlie, R., Torsvik, T., 1984. Magnetic remanence and fabric properties of laboratory- Weltje, G.J., Alberts, L.J.H., 2011. Packing states of ideal reservoir sands: Insights
deposited hematite-bearing red sandstone. Geophys. Res. Lett. 11, 221–224. from simulation of porosity reduction by grain rearrangement. Sed. Geol. 242,
Lu, R., Banerjee, S., Marvin, J., 1990. Effects of clay mineralogy and the electrical 52–64.
conductivity of water in the acquisition of depositional remanent magnetism in Zhao, Y., Davis, R.H., 2002. Interaction of two touching spheres in a viscous fluid.
sediments. J. Geophys. Res. B 95, 4531–4538. Chem. Eng. Sci. 57, 1997–2006.
Mitra, R., Tauxe, L., 2009. Full vector model for magnetization in sediments. Earth Zhao, Y., Galvin, K.P., Davis, R.H., 2002. Motion of a sphere down a rough plane in a
Planet. Sci. Lett. 286, 535–545. viscous fluid. Int. J. Multiph. Flow 28, 1787–1800.

Você também pode gostar