Você está na página 1de 12

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. B6, 10.

1029/2001JB000706, 2002

A short, reverse polarity interval within the Jaramillo subchron:


Evidence from the Jingbian section, northern Chinese Loess Plateau
Bin Guo1 and Rixiang Zhu
Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing, China

Fabio Florindo
Istituto Nazionale di Geofisica e Vulcanologia, Rome, Italy

Zhongli Ding and Jimin Sun


Institute of Geology and Geophysics, Chinese Academy of Sciences, Beijing, China
Received 18 June 2001; revised 25 November 2001; accepted 30 November 2001; published 26 June 2002.

[1] A high-resolution paleomagnetic and rock magnetic investigation supported by


scanning electron microscopy observations, X-ray diffraction determinations, and
anisotropy of magnetic susceptibility analyses has been carried out on a loess-paleosol
sequence from Jingbian (37.4N, 108.8E), at the northern extremity of the Chinese
Loess Plateau. The results indicate that the magnetic assemblage is dominated by
pseudo-single-domain magnetite with associated maghemite and traces of a
higher-coercivity phase, probably hematite. The Brunhes-Matuyama and upper Jaramillo
transitions have been carefully documented and a record of a short reverse polarity interval
(SRI), entirely within the Jaramillo normal polarity subchron, has been obtained. The
duration of the SRI, recorded over a stratigraphic interval of 48 cm, is 10.1 kyr (29 kyr if
the polarity transitions are included). These results seem to exclude that rock magnetic
variations are adversely affecting the recording process and compel us to reconsider the
Jaramillo subchron as characterized by a pair of normal polarity intervals interrupted by a
INDEX TERMS: 1520 Geomagnetism and Paleomagnetism:
short reverse polarity interval.
Magnetostratigraphy; 1519 Geomagnetism and Paleomagnetism: Magnetic mineralogy and petrology; 1530
Geomagnetism and Paleomagnetism: Rapid time variations; 1518 Geomagnetism and Paleomagnetism:
Magnetic fabrics and anisotropy; KEYWORDS: Chinese loess, Jingbian section, magnetostratigraphy, Jaramillo

1. Introduction
[2] The Jaramillo event was first discovered from a
series of volcanic rocks from the Valles Caldera (Santa Fe,
New Mexico) by Doell and Dalrymple [1966], who dated it
at 0.9 Ma. The discovery of the Jaramillo event (named
after Jaramillo Creek) added further detail to the previous
geomagnetic reversal polarity timescale of Doell et al.
[1966]. In the last three decades this event has been
observed in many different environments, from both volcanic and sedimentary sequences (including lacustrine,
marine, and aeolian sequences), confirming its global nature
[Jacobs, 1994]. Recently, new 40Ar/39Ar ages from the
Punaruu Valley in Tahiti indicate that the Jaramillo normal
polarity subchron (JNS) lasted 67 kyr, starting at 1.053
0.006 Ma and ending at 0.986 0.005 Ma [Singer et al.,
1999], in agreement with astronomical estimates [Berggren
et al., 1995; Shackleton et al., 1990].

Also at Istituto Nazionale di Geofisica e Vulcanologia, Rome, Italy.

Copyright 2002 by the American Geophysical Union.


0148-0227/02/2001JB000706$09.00

EPM

[3] In recent years, there has been increasing enthusiasm


to investigate the Jaramillo subchron after several authors
identified the presence of directional excursions, with amplitudes above that of background secular variations, within this
subchron in Pleistocene sedimentary sequences with high
sedimentation rates [Pillans et al., 1994; Zhu et al., 1994;
McIntosh et al., 1996; Richter et al., 1998; Biswas et al.,
1999; Channell and Mazaud, 2000]. The evidence comes
from different lithologies and localities (New Zealand,
China, Japan Sea, Mediterranean Sea, Iceland), suggesting
a worldwide occurrence. On the other hand, an examination
of the ocean floor magnetic anomaly patterns [Cande and
Kent, 1992, 1995], does not support the existence of any
short-term feature. This is probably related to the difficulties
in resolving such short-period features on ocean magnetic
profiles [Jacobs, 1994; Florindo, 1996]. In this paper, we
present a detailed paleomagnetic record of the JNS termination from the Jingbian loess section at the northern
extremity of the Chinese Loess Plateau. This site appears
to be an ideal recorder of such short-lived geomagnetic
features because of high dust input rates and relatively weak
pedogenesis compared to the southern and central Chinese
Loess Plateau, such as the Duanjiapo and Luochuan areas
[Sun, 1994; Zheng et al., 1995; Florindo et al., 1999].

2-1

EPM

2-2

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

2. Geological Setting and Paleomagnetic


Sampling
[4] Loess deposits and intercalated paleosols in the Chinese Loess Plateau consist of aeolian dust coming from the
dry deserts of the north and northwest, which have been
deposited over an area of 440,000 km2 [Liu et al., 1985].
Loess is accumulated under dry and cold conditions during
ice ages, whereas paleosols originated on the top of the
loess horizons during warmer and more humid periods
(interglacials and interstadials).
[5] The initiation of the Chinese aeolian dust deposition
has been recently extended back to more than 7 Ma,
marking the onset of the present-day East Asian monsoon
system [Ding et al., 1998; Sun et al., 1998; Porter, 2001].
According to the direction of the aeolian transport by means
of the cold winter monsoon winds, the thickness of loess
outcrops varies throughout the Chinese Loess Plateau, with
a maximum in the northwest (over 300 m in thickness) and
a decreasing thickness toward the east and southeast [e.g.,
Heller and Evans, 1995; Evans and Heller, 2001]. Along
the same direction, from the northwest to south and east,
there is also a decrease in the sediment grain size that
reflects the direction of the wind transport and suggests a
primary source of dust in the desert basins northwest of the
Chinese Loess Plateau. These variations reflect a contemporary climatic gradient, from relatively dry in the northwest to increasingly humid in the south and east.
[6] The Jingbian loess section (108.8E, 37.4N), located
in Jingbian County, Shannxi Province, China, lies in the
northern part of the Chinese Loess Plateau, east of the
Liupan Mountains (Figure 1). It stands on a 23-m-thick
upper Tertiary eolian sedimentary sequence, the Red Clay
Formation, and consists of a sequence of 33 loess-paleosol
levels with a total thickness of 237 m. In this area,
characterized by an arid/semiarid climate, the paleosols
are weakly developed and visually very indistinct in the
field.
[7] In previous magnetostratigraphic investigations of
loess sequences in China, the upper and lower boundaries
of the JNS were identified in loess units L10 (between S9
and S10) and L12 (between S11 and S12), respectively,
though some conflict exists [Heller and Liu, 1982; Liu et
al., 1987; Kukla and An, 1989; Rutter et al., 1991; Zheng et
al., 1992; Heller and Evans, 1995; Heslop et al., 2000;
Evans and Heller, 2001]. These differences have been
related either to low sampling resolution or to previous,
and now refined, stratigraphic subdivision of the complete
loess-soil sequence [e.g., Hus and Han, 1992; Heslop et al.,
2000].
[8] In the investigated area we first identified the two
thick and coarse-grained loess beds L9 and L15, which
are typically observed in the upper portions of the loess
sequences and used as stratigraphic markers in field work
[Ding et al., 1993]. After, in order to identify the other
loess-soil units, we measured the low-field magnetic
susceptibility from the top of the section to unit L15
using a portable Bartington magnetic susceptibility meter
equipped with a MS2F probe (operating frequency of
0.465 kHz).
[9] At first, a preliminary magnetostratigraphy was
developed from 253 oriented pilot samples, collected at

Figure 1. Distribution of loess in central China. The


Jingbian section is marked with a star.

20-cm intervals, from the bottom of S7 to the top of the


S13 (Figure 2). On the basis of thermal demagnetization
of these samples, we recognized (1) the BrunhesMatuyama boundary in loess layer L8, (2) the upper
boundary of the JNS close to the stratigraphic boundary
L10/S9, and (3) the lower boundary of the JNS in the
lower part of loess layer L12. A thin reversed polarity
interval was found near the stratigraphic boundary L10/
S10, and aiming to better characterize this short feature,
contiguous oriented block samples were taken from the
lower part of S9 to the lower part of L11 through a
stratigraphic interval of 6 m.
[10] According to the refined Chinese loess-paleosol
chronology [Heslop et al., 2000], the L10 layer (278 cm
in thickness) and the S10 paleosol unit (104 cm in thickness), which embrace this thin reversed polarity interval,
span time intervals of 28 and 22 kyr, with average accumulation rates of 9.9 and 4.7 cm kyr 1, respectively. The
low estimated sedimentation rate observed in the S10 unit is
probably related to reduced wind activity during the
emplacement of parent atmospheric dust from which the
paleosol formed.
[11] The sampling procedure for the overlapped oriented
loess blocks was as follows. First, in a well-exposed natural
outcrop, after cleaning the surface to a depth of >50 cm to
eliminate surface weathering effects, oriented and overlapped loess blocks with a 5  5 cm base and up to 30
cm high were sampled. A magnetic compass was used to
orient samples in the field, with magnetic north direction
marked on the top surface. In the laboratory, using an
electric saw with a very thin steel cutting wire, the blocks
were cut into four parallel columns (sets A, B, C and D)
with a 2  2 cm section and 30 cm long; 15 cubic samples

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

EPM

2-3

Figure 2. Preliminary magnetostratigraphy of Jingbian loess section from the bottom of S7 to the top of
S13. The nomenclature of the loess sequence follows Kukla and An [1989].

of 2  2  2 cm size were finally taken from each of these


loess columns.

3. Laboratory Procedures
[12] All samples from sets A, B, and C were progressively thermally demagnetized from room temperature up to
585C, using a magnetically shielded oven. Natural magnetic remanences were measured using a cryogenic magnetometer (2G Enterprises model 750R or 755R) installed in a
field free space.
[13] Magnetic coercivity and thermal unblocking spectra
were used to investigate the magnetic mineralogy and its
homogeneity throughout the investigated loess blocks.
For all the samples from the four sets (A, B, C, D) the
volume-specific low-field (and low-frequency) magnetic
susceptibility (k) was determined using a Bartington
MS2B magnetic susceptibility meter. The frequency
dependence of magnetic susceptibility (kfd) was measured
on samples from set A. For all samples from set D an
anhysteretic remanent magnetization (ARM) was imparted
in an 80-mT alternating field with a superimposed 0.05-mT
dc bias field using a modified GSD-1 AC Geophysical
Specimen Demagnetizer. For the same samples from set D
an isothermal remanent magnetization (IRM) was imparted
in a dc field of 2.7 T using a 660 2G-pulse magnetizer.

[14] On selected samples we investigated (1) the anisotropy of magnetic susceptibility, measured using a Kappabridge KLY 3 (AGICO), (2) the stepwise acquisition of
IRM up to 2.7 T (these high-intensity artificial remanences
were measured using a Digital Spinner Magnetometer
(DSM-1)), (3) the coercivity of remanence (Bcr) determined
by back-field application to the maximum IRM, (4) hysteresis loops using a Molspin vibrating sample magnetometer
(VSM), and (5) the temperature dependence of susceptibility, from room temperature up to 700C, measured in an
argon atmosphere to minimize oxidation, by a furnaceequipped Kappabridge KLY 3 [Hrouda, 1994]. The rock
magnetic analyses were performed at the paleomagnetism
laboratory of the Institute of Geology and Geophysics,
Beijing (China) and at the Istituto Nazionale di Geofisica
e Vulcanologia, Rome (Italy).

4. Results
4.1. Rock Magnetic Analyses
[15] Acquisition curves of IRM in progressively increasing fields up to 2.7 T reveal that 90% of the saturation
IRM (SIRM) is acquired below an inducing field of 300
350 mT for both loess and paleosol samples (Figure 3a).
The IRM acquisition curves are typical of sediments where
low-coercivity phases (magnetite/maghemite) dominate. At

EPM

2-4

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

Figure 3. (a) Stepwise acquisition of isothermal remanent magnetization (IRM) in progressively


increasing fields up to 2.7 T and (b) back coercivity curve (IRM back-field curve) for two samples from
unit S10 and four samples for unit L10.

fields 300 350 mT, IRM acquisition is due to traces of an


imperfect antiferromagnetic substance (i.e., haematite). The
progressive removal of the SIRM, obtained by applying
back fields, shows that the coercivity of remanence (Bcr) is
62 mT for loess unit L10 and 42 mT for paleosol unit
S10 (Figure 3b). Both values are consistent with a remanence mainly carried by magnetite/maghemite, but the
presence of a higher-coercivity phase in loess unit L10 is
suggested by the higher value of Bcr. These higher coercivities might also be due to the presence of low-temperature oxidation of stoichiometric magnetite grains [Van
Velzen and Dekkers, 1999].
[16] Similar values were obtained from the analysis of the
hysteresis loops measured on small chips of samples (0.7
g) coming from L10 (two samples at 27.64 and 27.92 m
from the top), S10 (two samples at 28.60 and 28.95 m from
the top), and L11 (one sample at 30.05 m from the top)
(Figure 4). Saturation magnetization (Ms), saturation remanence (Mrs), coercive force (Bc), and coercivity of remanence (Bcr) were determined. The range of hysteresis ratios
found are 0.19 < Mrs/Ms < 0.21 and 3.36 < Bcr/Bc < 3.53.
These values indicate a pseudo-single-domain state for the
magnetite grains in a narrow grain-size range [Day et al.,
1977; Vlag et al., 1996]. No evidence of wasp-waisted
characteristics [Roberts et al., 1995] was detected in the
hysteresis loops.
[17] The temperature dependence of magnetic susceptibility curves shows a clear decrease in susceptibility near
580C for both loess and paleosol samples (Figure 5),
which indicates that magnetite is the major contributor to
the susceptibility. The heating curves for the three loess
samples from different depths are quite similar, displaying a
broad maximum from 150C to 350C followed by a
slight decrease in magnetic susceptibility. For the paleosol
sample (J14) the heating curve also shows a similar bump
from 150C to 350C, but it is followed by a relatively
rapid decrease in magnetic susceptibility from 300C to
500C. These results may be interpreted in terms of the
thermally induced oxidation of some metastable cubic
maghemite (g-Fe2O3) converted to weakly magnetic rhombohedral hematite (a-Fe2O3) [Stacey and Banerjee, 1974].
Moreover, the absence of marked Hopkinson peaks could
indicate the absence of magnetite grains with a narrow

zdemir, 1997]. All samples


grain-size range [Dunlop and O
from both L10 and S10 display magnetite cooling curves
(not shown here) that are far above the heating curves. This
increase may be due to the thermally induced growth of new
magnetite from iron-containings silicates/clays.

Figure 4. Typical hysteresis loops for two representative


samples, before and after correction for the paramagnetic
contribution: (a) sample from the L10 layer at 27.92 m and
(b) sample from the S10 layer at 28.6 m.

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

EPM

2-5

These values are comparable to those observed in the sandy


loess units L9 (mean value of the susceptibility, 79  10 5
SI) and L15 (mean value of the susceptibility, 100.5  10 5
SI) in the southern loess plateau of China [Zhu et al., 2000],
which gives further evidence of the relatively weak level of
pedogenesis for the Jingbian section.
[19] The frequency dependence of magnetic susceptibility is indicative of magnetically viscous grains near the
superparamagnetic (SP)/stable single-domain (SD) boundary and is insensitive to multidomain grains. The parameter kfd is expressed as the difference between the magnetic
susceptibility measured at 0.465 kHz (kl) and 4.65 kHz (kh),
expressed as a percentage of the kl measurement (kfd% = (kl
kh)/kl)100) [e.g., Bloemendal et al., 1985]. The kfd values
obtained here are relatively low (mean of 4.3 with standard
deviation 0.43 for the L10 loess unit and 6.9 with standard
deviation 0.75 for the S10 paleosol unit), suggesting a
relatively weak pedogenesis in the investigated interval
(Figure 6b).
[20] Finally, considering that the magnetic mineralogy is
dominated by magnetite, the stratigraphic variations of the
ratios SIRM/k give indications of changes in magnetic grain
size (Figure 6e). The minima in the paleosol horizons S9,
S10, and S11 reflect a relatively higher abundance of SP
and/or smaller SD magnetite/maghemite grains compared to
the loess units L10, L11, and L12.
[21] Summarizing, magnetic coercivity and thermal
unblocking characteristics suggest that the magnetic mineralogy at Jingbian is dominated by pseudo-single-domain
magnetite with associated maghemite and traces of a higher
coercivity phase, probably hematite.

Figure 5. Temperature dependence of magnetic susceptibility (TDS) for selected samples from unit L10 and S10.

[18] The stratigraphic variation of concentration-dependent parameters such as k, SIRM, and ARM are shown in
Figures 6a, 6c, and 6d. The profiles of these magnetic
parameters provide boundaries that agree with the lithologic
ones observed in the field. All of these magnetic parameters
exhibit the lowest values in the loess layers and significant
magnetic increases in the paleosol horizons S9, S10, and
S11. From the variation of these bulk parameters we can
estimate that the concentration of magnetic grains does not
vary by more than 1 order of magnitude. It is noteworthy
that relatively low values of k are observed in the soil units.

4.2. Anisotropy Magnetic Susceptibility Analyses


[22] In this study, aiming to demonstrate that the magnetic record is not distorted by sedimentological phenomena
and/or by errors in the sampling procedure, the magnetic
fabric of samples coming from the investigated interval
(both inside and outside the short reverse polarity interval)
was analyzed by means of anisotropy magnetic susceptibility (AMS). The maximum (kmax), intermediate (kint), and
minimum (kmin) axes of AMS are shown in equal-area
stereogram in Figure 7. The kmin directions are vertical (D
= 5.8, I = 89.8, a95 = 2.7), perpendicular to the bedding
plane, whereas the inclinations of orientation kmax and kint
are very shallow (D = 208, I = 4.6, a95 = 7.5 and D =
3.5, I = 6.6, a95 = 7, respectively). These features are
typical of a primary sedimentary fabric without evidences
for an anomalous, distorted fabric [Marino and Ellwood,
1978]. We believe that this fabric is related to the direction
of transport [Ellwood and Howard, 1981] and that the eastwest orientations of kmax directions reflect a north-south
direction of strong dust-bearing monsoon winds.
4.3. Scanning Electron Microscopy Observations and
X-Ray Diffraction of Magnetic Extracts
[23] Three samples extracted from loess unit L10 (2) and
paleosol unit S10 (1) were gently disaggregated in a small
volume of water and thoroughly dispersed ultrasonically.
From the dispersed sediments, magnetic extracts were
obtained using a magnetic finger and examined with a
scanning electron microscope (SEM) model Amray KA1000B with an energy dispersive spectrometer.

EPM

2-6

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

Figure 6. Mineral magnetic parameters as a function of stratigraphic position. (a) Low-field magnetic
susceptibility, k; (b) frequency dependence of magnetic susceptibility, kfd; (c) saturation remanent
magnetization (at 2.7 T), SIRM; (d) anhysteretic remanent magnetization, ARM; and (e) ratio of the
SIRM/k. The paleosols are indicated by the light shaded regions in each plot.

[24] Scanning electron micrographs of these magnetic


extracts show that most of the grains comprise a mixture
of intact and chipped geometric crystals, subrounded to
rounded in shape. The size of these grains range from 1 to
15 mm, although some larger grains (30 mm) are visible. The
size distribution, and especially the morphology, of the
extracted grains suggests a detrital, windblown origin
[e.g., Tsoar and Pye, 1987] and excludes an in situ precipitation during, and as a response to, weathering and soil
formation [Maher and Taylor, 1988]. These evidences are
consistent with the presence of fresh biotite that is a mineral
highly sensitive to weathering processes [Sun, 1994].
[25] Analysis of the magnetic extracts by X-ray diffraction were performed using a DMAX 2400 X-ray diffractometer, with a Cu-Ka radiation and a scanning step of
0.02. The definitely matched minerals are magnetite,
maghemite, and hematite (Figure 8). Quartz is also present
in the extracts, possibly due to the presence of magnetic
grains as inclusions. These analyses were conducted at the
X-ray diffraction (XRD) analyses laboratory of the Institute
of Geology and Geophysics, Chinese Academy of Sciences,
in Beijing.
4.4. Paleomagnetic Analyses
[26] Initial natural remanence magnetization directions of
most of the samples exhibit normal polarity with directions
close to the present-day geomagnetic field value (D = 3,

Figure 7. Plot of results from anisotropy magnetic


susceptibility (AMS) measurements. Directions of kmax
(squares), kint (triangles) and kmin (circles) are plotted on an
equal-area stereographic projection.

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

EPM

Figure 8. X-ray diffraction spectra of magnetic extracts from samples from loess unit L10 and paleosol
unit S10 samples. M, magnetite; Mh, maghemite; H, hematite; Q, quartz.

Figure 9. Stratigraphic plot of (a) declination and (b) inclination of the NRM20C (dashed line indicates
the inclination of the geocentric axial dipolar field for the present latitude of the sampling site), and (c)
NRM20C intensities. The paleosols are indicated by the shaded regions in each plot.

2-7

EPM

2-8

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

Figure 10. Normalized intensity decay plots, vector component diagrams (solid and open squares
represent projections of vector endpoints onto the vertical and horizontal plane, respectively), and equalarea stereographic projections (lower hemisphere) of representative reverse, intermediate, and normal
polarity samples. Numbers indicate the heating temperature (in C).

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

EPM

2-9

Figure 11. Stratigraphic plot of (a) declination and (b) inclination of the ChRM (dashed line indicates
the inclination of the geocentric axial dipole field for the present sampling site latitude), and (c) virtual
geomagnetic poles (VGPs) latitude. The ChRM directions were determined by linear regression fits to
multiple demagnetization steps. Solid symbols, open circles, and triangles represent records from sets A,
B, and C. Stars represent the average value of three samples at the same stratigraphic level. (d) Magnetic
polarity zonation. Black (white) represents normal (reverse) polarity intervals. Intermediate directions are
indicated by the striped intervals. The paleosols are indicated by the light shaded regions in each
stratigraphic plot.
I = 56.8) and a few samples with reverse polarity (Figure
9). As also observed at Jiuzhoutai, northwestern Chinese
Loess Plateau [McIntosh et al., 1996], secondary components of magnetization did not fully obscure the reverse
polarity directions recorded between 27.22 and 27.24 m and
between 27.58 and 27.78 m from the top. Samples from set
A were stepwise thermally demagnetized in steps of 30
50C from room temperature to the limit of reproducible
results, which was generally higher than 500C. After the
removal of a low-temperature component of magnetization
in the interval 250 300C, the demagnetization diagrams
reveal a well-defined characteristic component of magnetization (ChRM) (Figure 10). We attribute the low-temperature component of magnetization to a viscous remanent
magnetization component [Pan et al., 2001] or to varying
degrees of low-temperature oxidation of stoichiometric
magnetite during the first stages of loess deposition [Heller
and Liu, 1984; Van Velzen and Dekkers, 1999]. This
evidence is supported by the rock magnetic investigation
which indicates the absence of magnetic phases with
unblocking temperatures close to 250 300C (i.e., Ti-rich
titanomagnetite).
[27] The ChRM directions were determined using orthogonal vector component diagrams [Zijderveld, 1967], stereographic projections, and intensity decay curves [Dunlop,
1979]; best fit lines for the progressive demagnetization data

were evaluated by principal component analysis [Kirschvink, 1980]. Generally, demagnetization at higher temperature steps (T > 500C) yielded rather scattered directions
due to the thermally induced growth of new magnetite
during the heating treatment. Virtually identical ChRM
directions were obtained from samples on the same stratigraphic level from sets B and C, which were only stepwise
demagnetized at temperatures of 300, 350, 400, 450, and
500C (Figure 11). The ChRM directions of multiple specimens, taken from the same stratigraphic horizon, have been
analyzed using Fisher [1953] statistics. Mean directions
with a95  20 were considered as well defined and 84% of
the analyzed samples satisfied this criterion. The normal and
reverse ChRM directions are from the stable polarity
intervals group (excluding the pilot samples) in two nearly
antipodal clusters defined by D = 1.2, I = 51 (a95 =
3.0) for 103 normal polarity levels and D =178.4, I =
59 (a95 = 3.6) for 45 reverse polarity levels.
[28] After the evaluation of the ChRMs, the corresponding north VGPs were calculated. In the interval between 31
m and the top, four stable polarity stages and three polarity
transition intervals were obtained (Figure 11). The recorded
polarity transitions are not characterized by a smooth
progressive change from reverse (normal) to normal
(reverse). Rather, the transition zones display some shortlived directional fluctuations in declination and inclination.

EPM

2 - 10

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

The presence of small-amplitude fluctuations of the declination and inclination record (Figure 11), within loess unit
L10, may result from difficulties in acquiring a substantial
DRM in the coarse-grained sediments that dominate this
unit (Figure 12).
[29] Starting from the middle of L11, from 31 to 29.09 m,
the VGPs exhibit a stable normal polarity. Transitional
VGPs are recorded from 29.09 to 28.84 m and are followed
by a stable reverse polarity interval, from 28.84 to 28.36 m.
This is followed by a complex transitional pattern of VGPs
over an interval of 1.15 m, from 28.36 to 27.21 m, which
contains a thin reverse polarity interval (from 27.85 to 27.70
m). The VGPs exhibit stable normal polarity from 27.21 to
25.84 m and then a normal-reverse transition from 25.84 to
25.17 m. Finally, stable reverse polarities are recorded in the
uppermost 0.17 m of the section.
[30 ] Statistical analysis [Fisher, 1953] of 64 mean
ChRMs from loess unit L10 and 34 mean ChRMs from
paleosol S10 yields angular standard deviation (dispersion)
of 24.0 and 16.7 respectively, about a mean direction of
loess with D = 1.2, I = 51.7, and a95 = 4.2; and paleosol
with D = 179.2, I = 57.1, and a95 = 5.1. The angular
standard deviation of the corresponding VGPs is 21.6 and
20.0, respectively (Table 1). These values are greater than
those predicted for secular variation at the latitude of
Jingbian (37.4N) as estimated from globally distributed
lava flows [McFadden et al., 1991]. This provides evidence
that the paleomagnetic record was not severely altered by
smoothing effects.

5. Discussion
[31] The most striking feature in this record is a short
reverse polarity interval, which is bounded by transitional
intervals, recovered between the middle lower part of S10
and the middle part of L10 over a stratigraphic thickness of
1.88 m (from 29.09 to 27.21 m). Rock magnetic investigations supported by SEM observation, X-ray diffraction,
and AMS analyses indicate that the magnetization is not
anomalous and support the conclusion that the short reverse
polarity interval recorded within the Jaramillo subchron
reflects a real geomagnetic phenomenon. Assuming that
this feature represents part of the upper Jaramillo (UJ)
polarity transition, spanning the stratigraphic interval
between 29.09 and 25.17 m and assuming the same sedimention rate for this interval as for soil S10, we obtain a
duration of 51.9 kyr for the transition. This conclusion is in
serious conflict with most published estimates for the
duration of polarity transitions [Jacobs, 1994].
[32] We also considered the possibility that the polarity
transition recorded between 29.09 and 28.84 m represents
the UJ polarity transition and that the stable normal polarity
interval between 25.84 and 27.21 m could be either the
Kamikatsura or the Santa Rosa events, which have been
dated at  0.88 Ma and  0.92 Ma, respectively [Singer et
al., 1999; Guyodo et al., 1999; Dinare`s-Turell et al., 2002].
Under this hypothesis the UJ termination (0.986 0.005 Ma
[Singer et al., 1999]) is interpreted to occur in the lower part
of paleosol S10. This conclusion is (1) inconsistent with the
acknowledged result that the UJ transition is located in L10
[e.g., Ding et al., 1998] and (2) in contrast with the new
Chinese loess-paleosol chronology [Heslop et al., 2000]

Figure 12. Grain-size data as function of stratigraphic


depth at Jingbian. Grain sizes were determined at 5-cm
interval with a computer operated PRO-700 SK laser
Micron Sizer. Ultrasonic pretreatment with addition of 20%
(NaPO3)6 solution has been used to disperse the samples for
particle determination. The median grain size (Md) in the
loess units L10 and L11 is larger than in the soils S9 and
S10, and, in particular, the Md of loess unit L10 (26.2 mm 
Md  43.1 mm) is coarser than that of loess unit L10 in the
southern loess plateau [Ding et al., 1999]. This variation
reflects the direction of the transport by means of the cold
winter monsoon winds, from the northwest to south and
east, across the Chinese Loess Plateau [Liu et al., 1985;
Ding et al., 1999].

EPM

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

2 - 11

Table 1. Mean Directions and Statistical Parameters From Units L10 and S10a
L10
S10

a95

sd

Longitude

Latitude

A95

SD

64
34

1.2
179.2

51.7
57.1

4.2
5.1

18.7
23.9

24.0
19.2

277.7
224.2

84.9
89.3

6.2
6.2

9.1
16.9

21.6
20.0

a
Samples with VGP latitudes <50 were excluded. N, number of samples; D and I, mean declination and inclination, respectively, of characteristic
remanent magnetization (ChRM) directions; Longitude and Latitude, mean longitude and latitude of ChRM poles, respectively; a95 and A95, radii of 95%
confidence circles about mean ChRM direction and pole, respectively; k and K, precision parameters for ChRM directions and poles, respectively; sd and
SD, angular standard deviation equal to dispersion of ChRM directions and poles, respectively.

which assigns numerical ages of 1.012 Ma to the L10-S10


boundary and 1.034 Ma to the S10-L11 boundary, respectively. On the other hand, if the geomagnetic episodes
recorded between 29.09 and 27.21 m and between 25.84
and 25.17 m (including stable and transitional states)
represent a short reverse polarity interval and the UJ
transition, respectively, the duration of the UJ polarity
transition is 9.3 kyr, which is consistent with most reported
estimates [Zhu et al., 1994; Verosub et al., 1996; Singer et
al., 1999]. The short reverse polarity interval can be broken
down into three parts: a normal (N) to reverse (R) polarity
transition recorded between 29.09 and 28.84 m and lasting
5.3 kyr; a stable state, recorded between 28.84 and 28.36
m, spanning some 10.1 kyr; and a R to N polarity transition
recorded between 28.36 m and 27.21 m which spans the
boundary between S10 and L10 and which lasted about 13.6
kyr.
[33] In summary, our detailed paleomagnetic and rock
magnetic investigations support the conclusion that a short
reverse polarity interval is recorded within the JNS. This
geomagnetic feature lasted 10.1 kyr (29 kyr if the polarity
transitions are included). Its short duration implies that it
might be difficult to resolve in paleomagnetic investigations
of deep-sea and most continental sequences, as well as in
ocean floor magnetic anomaly patterns. Several authors
have suggested that delayed remanence acquisition might
be important in the Chinese loess sequences [Heller et al.,
1987; Tauxe et al., 1996; Zhou and Shackleton, 1999], and
it is possible that this short polarity interval represents an
interval with delayed remanence acquisition (acquired during the late Matuyama Chron). The best test for the reality
of the short reverse polarity interval that we have reported
would be to document it in other types of geological
archives (e.g., marine or continental sediments or lavas) in
other locations around the world.
[34] More than 30 years after the discovery of the JNS,
the evidence presented above compels us to reconsider it as
characterized by a pair of normal polarity intervals interrupted by a short reverse polarity interval. This conclusion should renew ongoing debate regarding the influence
of missing reversals on the analysis of the geomagnetic
polarity timescale [Marzocchi, 1997]. We are currently
undertaking further studies to improve the resolution and
to verify the presence of this short geomagnetic feature at
other loess sections.
[35] Acknowledgments. This paper benefited greatly from critical
comments and suggestions by S. Brachfeld, B. Clement, F. Heller, M. E.
Evans, J. Pares, and L. Tauxe. We gratefully acknowledge the assistance of
G. L. Yang during the sampling. Financial assistance was supported by the
National Science Foundation of China (grant 49834001). G.B. gratefully
acknowledges E. Boschi and support from the Istituto Nazionale di Geo-

fisica e Vulcanologia in Rome in the form of a fellowship during his visit to


the paleomagnetism laboratory.

References
Berggren, W. A., F. J. Hilgen, C. G. Langereis, D. V. Kent, J. D. Obradovich, I. Raffi, M. E. Raymo, and N. J. Shackleton, Late Neogene chronology: New perspectives in high-resolution stratigraphy, Geol. Soc. Am.
Bull., 107, 1272 1287, 1995.
Biswas, D. K., M. Hyodo, Y. Taniguchi, M. Kaneko, S. Katoh, H. Sato, Y.
Kinugasa, and K. Mizuno, Magnetostratigraphy of Plio-Pleistocene sediments in a 1700-m core from Osaka Bay, southern Japan and short
geomagnetic events in the middle Matuyama and early Brunhes chrons,
Palaeogeogr. Palaeoclimatol. Palaeoecol., 148, 233 248, 1999.
Bloemendal, J., C. E. Barton, and C. Radhakrishnamurthy, Correlation
between Rayleigh loops and frequency-dependent and quadrature susceptibility: Application to magnetic granulometry of rocks, J. Geophys.
Res., 90, 8789 8792, 1985.
Cande, S. C., and D. V. Kent, A new geomagnetic polarity timescale for the
late Cretaceous and Cenozoic, J. Geophys. Res., 97, 13,951 13,971,
1992.
Cande, S. C., and D. V. Kent, Revised Calibration of the geomagnetic
polarity timescale for the Late Cretaceous and Cenozoic, J. Geophys.
Res., 100, 6093 6095, 1995.
Channell, J. E., and A. Mazaud, The Matuyama Chron at Sites 983 and 984
(Iceland Basin), Eos Trans. AGU, 81(48), Fall Meet. Suppl., Abstract
GP62A-04, 2000.
Day, R., M. D. Fuller, and V. A. Schmidt, Hysteresis properties of titanomagnetites: Grain-size and compositional dependence, Phys. Earth Planet. Inter., 13, 260 267, 1977.
Dinare`s-Turell, J., L. Sagnotti, and A. P. Roberts, Relative geomagnetic
paleointensity from the Jaramillo subchron to the Matuyama/Brunhes
boundary as recorded in a Mediterranean piston core, Earth Planet.
Sci. Lett., 194, 327 341, 2002.
Ding, Z. L., N. W. Rutter, and T. S. Liu, Pedostratigraphy of Chinese loess
deposits and climatic cycles in the last 2.5 Ma, Catena, 20, 73 91, 1993.
Ding, Z. L., J. M. Sun, S. L. Yang, and T. S. Liu, Preliminary magnetostratigraphy of a thick eolian red clay-loess sequence at Lingtai, the Chinese Loess Plateau, Geophys. Res. Lett., 25, 1225 1228, 1998.
Ding, Z. L., J. M. Sun, N. W. Rutter, D. Rokosh, and T. S. Liu, Changes in
sand content of loess deposits along a north-south transect of the Chinese
Loess Plateau and the implications for desert variations, Quat. Res., 52,
56 62, 1999.
Doell, R. R., and G. B. Dalrymple, Geomagnetic polarity epochsA new
polarity event and the age of the Brunhes-Matuyama boundary, Science,
152, 1060 1061, 1966.
Doell, R. R., G. B. Dalrymple, and A. Cox, Geomagnetic polarity epochsSierra Nevada data, 3, J. Geophys. Res., 71, 531 541, 1966.
Dunlop, D. J., On the use of Zijderveld vector diagrams in multicomponent
paleomagnetic studies, Phys. Earth Planet. Inter., 20, 12 24, 1979.
. O
zdemir, Rock Magnetism: Fundamentals and
Dunlop, D. J., and O
Frontiers, 573 pp., Cambridge Univ. Press, New York, 1997.
Ellwood, B. B., and J. H. Howard, Magnetic fabric development in an
experimentally produced barchan dune, J. Sediment. Petrol., 51, 97
100, 1981.
Evans, M. E., and F. Heller, Magnetism of loess/palaeosol sequences: Recent developments, Earth Sci. Rev., 54, 129 144, 2001.
Fisher, R. A., Dispersion on a sphere, Proc. R. Soc. London, Ser. A, 217,
295 305, 1953.
Florindo, F., Record of a previously unidentified short geomagnetic event
from an upper Miocene sedimentary sequence, and preferred path of the
transitional VGPs, Geophys. J. Int., 126, F1 F5, 1996.
Florindo, F., R. X. Zhu, B. Guo, L. P. Yue, Y. X. Pan, and F. Speranza,
Magnetic proxy climate results from the Duanjiapo Loess section, southermost extremity of the Chinese loess plateau, J. Geophys. Res., 104,
645 659, 1999.

EPM

2 - 12

GUO ET AL.: INTRA-JARAMILLO REVERSE POLARITY INTERVAL

Guyodo, Y., C. Richter, and J.-P. Valet, Paleointensity record from Pleistocene sediments (1.4 0 Myr) off the California margin, J. Geophys. Res.,
104, 22,953 22,964, 1999.
Heller, F., and M. E. Evans, Loess magnetism, Rev. Geophys., 33, 211
240, 1995.
Heller, F., and T. S. Liu, Magnetostratigraphical dating of loess deposits in
China, Nature, 300, 431 433, 1982.
Heller, F., and T. S. Liu, Magnetism of Chinese loess deposits, Geophys. J.
R. Astron. Soc., 77, 125 141, 1984.
Heller, F., B. Meili, J. D. Wang, H. M. Li, and T. S. Liu, Magnetization and
sedimentation history of loess in the central loess plateau of China, in
Aspects of Loess Research, edited by T. S. Liu, pp. 147 163, China
Ocean Press, Beijing, 1987.
Heslop, D., C. G. Langereis, and M. J. Dekkers, A new astronomical timescale for the loess deposits of northern China, Earth Planet. Sci. Lett.,
184, 125 139, 2000.
Hrouda, F., A technique for the measurement of thermal changes of magnetic susceptibility of weakly magnetic rocks by the CS-2 apparatus and
KLY-2 Kappabridge, Geophys. J. Int., 118, 604 612, 1994.
Hus, J. J., and M. Han, The contribution of loess magnetism in China to the
retrival of the past global changes-Some problems, Phys. Earth Planet.
Inter., 70, 154 168, 1992.
Jacobs, J. A., Reversals of the Earths Magnetic Field, 346 pp., Cambridge
Univ. Press, New York, 1994.
Kirschvink, J. L., The least-squares line and plane and the analysis of
paleomagnetic data, Geophys. J. R. Astron. Soc., 62, 699 718, 1980.
Kukla, G., and Z. S. An, Loess stratigraphy in central China, Paleogeogr.
Paleoclimatol. Paleoecol., 72, 203 225, 1989.
Liu, T. S., et al., Loess and the Environment, 251 pp., China Ocean Press,
Beijing, 1985.
Liu, X. M., T. S. Liu, T. C. Xu, C. Liu, and M. Y. Cheng, A preliminary
study on magnetostratigraphy of a loess profile in Xifeng area, Gansu
Province, in Aspects of Loess Research, edited by T. S. Liu, pp. 164 174,
China Ocean Press, Beijing, 1987.
Maher, B. A., and R. M. Taylor, Formation of ultrafine-grained magnetite in
soils, Nature, 336, 368 370, 1988.
Marino, R. J., and B. B. Ellwood, Anomalous magnetic fabric in sediments
which record an apparent geomagnetic field excursion, Nature, 274, 581,
1978.
Marzocchi, W., Missing reversals in the geomagnetic polarity timescale:
Their influence on the analysis and in constraining the process that generates geomagnetic reversals, J. Geophys. Res., 102, 5157 5171, 1997.
McFadden, P. L., R. T. Merrill, M. W. McElhinny, and S. Lee, Reversals of
the Earths magnetic field and temporal variations of the dynamo families, J. Geophys. Res., 96, 3923 3933, 1991.
McIntosh, G., T. C. Rolph, J. Shaw, and P. Dagley, A detailed record of
normal-reversed-polarity transition obtained from a thick loess sequence
at Jiuzhoutai, near Lanzhou, China, Geophys. J. Int., 127, 651 664,
1996.
Pan, Y. X., R. X. Zhu, J. Shaw, Q. S. Liu, and B. Guo, Can relative
paleointensities be determined from the normalized magnetization of
the wind-blown loess of China?, J. Geophys. Res., 106, 19,221
19,232, 2001.
Pillans, B. J., A. P. Roberts, G. S. Wilson, S. T. Abbott, and B. V. Alloway,
Magnetostratigraphic, lithostratigraphic and tephrostratigraphic constraints on Lower and Middle Pleistocene sea-level changes, Wanganui
Basin, New Zealand, Earth Planet. Sci. Lett., 121, 81 98, 1994.
Porter, S. C., Chinese loess record of monsoon climate during the last
glacial-interglacial cycle, Earth Sci. Rev., 54, 115 128, 2001.
Richter, C., A. P. Roberts, J. S. Stoner, L. D. Benning, and C. T. Chi,
Magnetostratigraphy of Pliocene-Pleistocene sediments from the eastern
Mediterranean Sea, Proc. Ocean Drill. Program Sci. Results, 160, 61
73, 1998.
Roberts, A. P., Y. Cui, and L. Verosub, Wasp-waisted hysteresis loops:

Mineral magnetic characteristics and discrimination of components in


mixed magnetic systems, J. Geophys. Res., 100, 17,909 17,924,
1995.
Rutter, N. W., Z. L. Ding, M. E. Evans, and T. S. Liu, Magnetostratigraphy
of the Baoji loess-paleosol section in the north-central China Loess Plateau, Quat. Int., 7, 97 102, 1991.
Shackleton, N. J., A. Berger, and W. R. Peltier, An alternative astronomical
calibration of the lower Pleistocene timescale based on ODP Site 677,
Trans. R. Soc. Edinburgh Earth Sci., 81, 251 261, 1990.
Singer, B. S., K. Hoffman, A. Chauvin, R. S. Coe, and M. S. Pringle,
Dating transitionally magnetized lavas of the late Matuyama Chron: Toward a new 40Ar/39Ar timescale of reversals and events, J. Geophys. Res.,
104, 679 693, 1999.
Stacey, F. D., and S. K. Banerjee, The Physical Principles of Rock Magnetism, Elsevier Sci., New York, 1974.
Sun, J. M., Climatic change and environmental evolution in the desert-loess
transitional Zone of north China, Ph.D. thesis, 145 pp., Inst. of Geol.,
Chin. Acad. Sci., Beijing, 1994.
Sun, D. H., J. Shaw, Z. An, M. Cheng, and L. Yue, Magnetostratigraphy
and paleoclimatic interpretation of a continuous 7.2 Ma Late Cenozoic
eolian sediments from the Chinese Loess Plateau, Geophys. Res. Lett., 25,
85 88, 1998.
Tauxe, L., T. Herbert, N. J. Shackleton, and Y. S. Kok, Astronomical
calibration of the Matuyama-Brunhes boundary: Consequences for magnetic remanence acquisition in marine carbonates and the Asian loess
sequences, Earth Planet. Sci. Lett., 140, 133 146, 1996.
Tsoar, K., and K. Pye, Dust transport and the question of desert loess
formation, Sedimentology, 34, 139 153, 1987.
Van Velzen, A. J., and M. J. Dekkers, Low-temperature oxidation of magnetite in loess-paleosol sequences: A correction of rock magnetic parameters, Stud. Geophys. Geod., 43, 357 375, 1999.
Verosub, K. L., E. Herrero-Bervera, and A. P. Roberts, Relative paleointensity across the Jaramillo subchron and the Matuyama/Brunhes boundary, Geophys. Res. Lett., 23, 467 470, 1996.
Vlag, P., P. Rochette, and M. J. Dekkers, Some additional hysteresis parameters for a natural (titano)magnetite with known grain size, Geophys.
Res. Lett., 23, 2803 2806, 1996.
Zheng, H. B., Z. S. An, and J. Shaw, New contributions to Chinese PlioPleistocene magnetostratigraphy, Phys. Earth Planet. Inter., 70, 146
153, 1992.
Zheng, H. B., T. C. Rolph, J. Shaw, and Z. S. An, A detailed palaeomagnetic record for the last interglacial period, Earth Planet. Sci. Lett., 133,
339 351, 1995.
Zhou, L. P., and N. J. Shackleton, Misleading positions of geomagnetic
reversal boundaries in Eurasian loess and implications for correlation
between continental and marine sedimentary sequences, Earth Planet.
Sci. Lett., 168, 117 130, 1999.
Zhu, R. X., C. Laj, and A. Mazaud, The Matuyama-Brunhes and Upper
Jaramillo transitions recorded in a loess section at Weinan, north-central
China, Earth Planet. Sci. Lett., 125, 143 158, 1994.
Zhu, R. X., B. Guo, Z. L. Ding, Z. T. Guo, A. Kazansky, and G. Matasova,
Gauss-Matuyama polarity transition obtained from a loess section at
Weinan, north-central China, Chin. J. Geophys, 43, 621 634, 2000.
Zijderveld, J. D. A., A.C. demagnetization of rocks: Analysis of results, in
Methods in Paleomagnetism, edited by D. W. Collinson, pp. 254 286,
Elsevier Sci., New York, 1967.

Z. Ding, J. Sun, and R. Zhu, Institute of Geology and Geophysics,


Chinese Academy of Sciences, Beijing 100101, China.
F. Florindo and B. Guo, Istituto Nazionale di Geofisica e Vulcanologia,
Via di Vigna Murata 605, I-00143, Rome, Italy. (florindo@ingv.it)

Você também pode gostar