Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Goudge et al.

A 40,000 yr record of clay mineralogy at Lake Towuti,


Indonesia: Paleoclimate reconstruction from reflectance
spectroscopy and perspectives on paleolakes on Mars

Timothy A. Goudge1,2,†, James M. Russell1, John F. Mustard1, James W. Head1, and Satria Bijaksana3
1
Department of Earth, Environmental and Planetary Sciences, Brown University, 324 Brook Street, Box 1846, Providence,
Rhode Island 02912, USA
2
Jackson School of Geosciences, The University of Texas at Austin, Austin, Texas 78712-1722, USA
3
Faculty of Mining and Petroleum Engineering, Institut Teknologi Bandung, Jalan Ganesa 10, Bandung 40132, Indonesia

ABSTRACT method for developing paleoenvironmental A common remote sensing technique used to
records from sedimentary phyllosilicate min- study the surface mineralogy of Mars is visible
Sediment deposited within lake basins eralogy. Exposed paleolake deposits on Mars to near-infrared (VNIR) reflectance spectroscopy
can preserve detailed records of past envi- should preserve similar paleoenvironmental (e.g., Bibring et al., 2005, 2006; Mustard et al.,
ronmental conditions on planetary surfaces, information that can be accessed through de- 2005, 2008; Murchie et al., 2009; Ehlmann et
including both Earth and Mars. Establishing tailed remote sensing observations of stratig- al., 2011; Carter et al., 2013). VNIR reflectance
how to best characterize these paleoclimate raphy and VNIR reflectance spectroscopy in spectroscopy studies how radiation in the VNIR
records is thus critical for understanding the a source-to-sink framework. portion of the electromagnetic spectrum interacts
evolution of past planetary climates. Here, we with a surface and is reflected back to a detector
present an ~40 k.y. lake sediment record from INTRODUCTION (i.e., a spectrometer). Minerals with infrared ac-
Lake Towuti, Indonesia, developed using tive components can have unique VNIR absorp-
visible to near-infrared (VNIR) reflectance Lake basins on planetary surfaces are effec- tion features, and analysis of the precise position
spectroscopy. Source sediment from the main tive traps for sediment and can provide detailed and shape of absorptions in an unknown measured
river input to Lake Towuti, the Mahalona records of the evolution of past environmental spectrum can be used to confidently identify many
River, is spectrally dominated by Mg-rich conditions throughout their lifetime of activ- of the minerals present (Farmer, 1974; Clark et al.,
serpentine; however, we also identify a dis- ity. Paleolimnological records are well known 1990; Gaffey et al., 1993; Clark, 1999).
tinct Al-phyllosilicate component, which we from Earth (e.g., Sly, 1978; Cohen, 2003), VNIR reflectance spectroscopy is par-
interpret as kaolinite, that increases in rela- and paleolake basins have also been identi- ticularly sensitive to vibrational absorptions
tive proportion to serpentine with decreasing fied on Mars, where they provide some of the caused by molecular components (e.g., H2O,
grain size. Sink sediment from two cores col- most convincing geologic evidence for ancient OH–, or CO32–) in the structure of minerals
lected at the distal margins of the Mahalona fluvial activity (e.g., Goldspiel and Squyres, such as phyllosilicates (i.e., clay minerals) or
River delta has similar spectral signatures to 1991; Cabrol and Grin, 1999, 2001, 2010; Ir- carbonates. These molecular components have
the input source sediment. The cores capture win et al., 2002, 2005; Malin and Edgett, 2003; their fundamental vibrational absorptions in
systematic variations in the proportion of Al- Fassett and Head, 2005, 2008). While paleo- the mid-infrared (MIR) portion of the electro-
phyllosilicate to serpentine over time, which lake activity on Mars has been investigated magnetic spectrum, which manifest as strong
is also expressed in changes in bulk elemen- with in situ rover analyses in a handful of lo- overtone and combination tone absorptions in
tal chemistry of the sediment. We show that cations (e.g., Christensen et al., 2004; Squyres the VNIR (Farmer, 1974; Clark et al., 1990;
the abundance of serpentine relative to Al-­ et al., 2004a, 2004b; McGlynn et al., 2012; Gaffey et al., 1993). The sensitivity of infra-
phyllosilicate increases dramatically during Grotzinger et al., 2014, 2015; McLennan et red spectroscopy to H2O and OH– makes it an
the globally cooler, regionally drier climate al., 2014; Mangold et al., 2016), the study of ideal tool for studying clay mineralogy. Pre-
of the Last Glacial Maximum. This change martian paleolake basins and associated sedi- vious workers have investigated the detailed
records the grain size–dependent mineral- mentary deposits is primarily restricted to the structure of phyllosilicates using both MIR
ogy of deltaic sediment, which is ultimately analysis of orbital remote sensing data (e.g., (e.g., Cuadros and Dudek, 2006) and VNIR
driven by forced delta progradation and river Goldspiel and Squyres, 1991; Cabrol and Grin, (e.g., Petit et al., 1999) spectroscopy; however,
incision during lake lowstands. Our analyses 1999, 2001, 2010; Irwin et al., 2002, 2005; VNIR reflectance spectroscopy has proved
show that VNIR reflectance spectroscopy Malin and Edgett, 2003; Fassett and Head, most useful for studying clay mineralogy on
offers a rapid, nondestructive, and effective 2005, 2008; Ehlmann et al., 2008; Milliken Mars (e.g., Bibring et al., 2005, 2006; Mustard
and Bish, 2010; Ansan et al., 2011; Wray et al., et al., 2008; Murchie et al., 2009; Ehlmann et
2011; Goudge et al., 2015). al., 2011; Carter et al., 2013).
tgoudge@jsg.utexas.edu

GSA Bulletin; July/August 2017; v. 129; no. 7/8; p. 806–819; doi: 10.1130/B31569.1; 9 figures; 1 table; Data Repository item 2017076;
published online 3 March 2017.

806 For permission


Geological Societytoof
copy, contact editing@geosociety.org
America Bulletin, v. 129, no. 7/8
© 2017 Geological Society of America

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

VNIR reflectance spectroscopy has also been


used to study the mineralogy and biogeochemical
characteristics of lacustrine sediment from terres-
trial lake basins, and has the potential for use in
paleolimnologic studies (e.g., Bishop et al., 1996,
2003; Vogel et al., 2008; Rosén et al., 2010; Lynch
et al., 2015; Weber et al., 2015). However, the sen-
sitivity of VNIR reflectance spectroscopy to the
presence of clay minerals in terrestrial lake sedi-
ment has yet to be widely exploited (e.g., Lynch et
al., 2015; Weber et al., 2015).
Here, we use VNIR reflectance spectroscopy,
complemented with more traditional laboratory
analysis techniques, to study the mineralogy
and chemistry of sediment from Lake Towuti, a
large tectonic lake basin in Indonesia. We char-
acterize the source-to-sink mineralogy of Lake
Towuti and demonstrate the utility of VNIR
reflectance spectroscopy for assessing paleoen-
vironmental signals preserved within lake sedi-
ment. We observe large changes in Lake Towuti
sediment mineralogy and composition during
the past ~40,000  yr that improve our under-
standing of lake sedimentation and regional cli-
mate variability for this basin. Our work also in-
dicates that paleoenvironmental records may be
developed through remote sensing analyses of
paleolake deposits on Mars, which we suggest
will be best achieved through careful analysis of
both mineralogy and stratigraphy in the frame-
work of a source-to-sink analysis (e.g., Milliken
and Bish, 2010; Goudge et al., 2015).

METHODS

Lake Towuti, Indonesia—Geologic Context


and Sampling

Lake Towuti (2.5°S, 121.5°E) is a hydrologi-


cally open lake located on the island of Sulawesi,
Indonesia (Fig.  1). Lake Towuti has a surface

Figure  1. Lake Towuti, Indonesia. (A) Re-


gional context for Lake Towuti on the is-
land of Sulawesi in Indonesia. Red box
indicates approximate location of part B.
Background is the Ocean base map from
ESRI et al. (2015). (B) Generalized geologic
map of Lake Towuti and the surrounding
area showing the dominance of ultramafic
material. Map is modified from Costa et
al. (2015). (C) Bathymetry of Lake Towuti
showing the location of the two analyzed
sediment cores at the distal margins of the
Mahalona River delta (white dots labeled 4
for TOW4 and 5 for TOW5). River inputs
are shown in thin blue lines, with the Maha-
lona River shown as a thick blue line.

Geological Society of America Bulletin, v. 129, no. 7/8 807

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

area of ~560 km2, a maximum depth of 203 m, cal Spectral Devices (ASD) FieldSpec 3 spec- tion source with a Ni filter that operates with
and a catchment area of ~1500  km2 (Russell et trometer, which covers the wavelength range 10 mA and 30 kV. Measurements were made
al., 2014). The catchment area primarily consists 350–2500 nm at a spectral resolution of ~10 nm with a 0.6  mm divergence, a 3  mm antiscatter
of laterites/Oxisols (Golightly and Arancibia, and a spectral sampling of 1 nm. VNIR spectra shield, and a 2.5° Soller slit. Our samples were
1979; Soil Survey Staff, 1999) derived from, and of the input sediment samples were measured run over the 2θ range 2°–70° with a step size of
overlain on, the East Sulawesi ophiolite, which with an external light source and the bare ASD ~0.02°. We used a dwell time of 4 s per step for
is largely composed of harzburgite, dunite, and fiber optic cable. Spectra of the two sediment the >1 mm, 0.25–1 mm, and 63–250 µm input
serpentinized lherzolite (Monnier et al., 1995; cores were measured at 1 cm intervals along the sediment grain-size separates, and 12 s per step
Kadarusman et al., 2004; Fig. 1B). Lake Towuti entire core length using the ASD contact probe for all other samples.
is fed by a number of river networks, although the attached to the instrument stage of a Geotek™
primary input is the Mahalona River, which feeds multisensor core logger (MSCL). The scanned RESULTS
the lake from the north (Fig. 1C) and supplies the VNIR reflectance data for the two sediment
majority of water and sediment flux to the basin cores were then converted from a data set rela- VNIR Spectroscopy
(Costa et al., 2015). tive to core depth (i.e., a spectrum each 1  cm
For our source-to-sink analysis, we have fo- along the entire length of core) to one relative Input Sediment Samples
cused on source sediment input to the lake from to absolute age using linear interpolated age The VNIR reflectance spectra of the input
the Mahalona River, and sink sediment depos- models (see the GSA Data Repository1 for ad- sediment samples are characterized by a series
ited at the distal margins of a delta deposit build- ditional details). of distinct vibrational absorptions related to
ing out from the mouth of this river (Fig. 1C). the presence of phyllosilicate minerals (Fig. 2;
We analyzed three sets of input sediment sam- Major Element Chemistry Table 1). All of the grain-size separates are spec-
ples from the Mahalona River collected during trally dominated by absorption features charac-
the 2013 field season: (1) suspended load and Major element chemistry (Al, Ca, Cr, Fe, K, teristic of Mg-rich serpentine, including narrow,
(2) bed load sediment collected ~500  m up- Mg, Mn, Na, Ni, Si, and Ti) of the input sedi- vibrational absorptions centered at ~1390, 2150,
stream from the river mouth, and (3) river mouth ment samples and lake core subsamples was and 2320 nm (Fig.  2). The absorption feature
sediment collected within the lake in water ~1 m measured using inductively coupled plasma– centered at ~1390 nm is caused by the first over-
deep. We also analyzed sink sediment from two atomic emission spectrometry (ICP-AES). tone of OH stretch, and the absorption feature
cores ~10 m in length collected from both the Prior to analysis by ICP-AES, aliquots of the centered at ~2320 nm is caused by a combina-
eastern and southern distal margins of the Ma- samples were prepared via flux fusion using Li- tion tone of Mg-OH bend and OH stretch (King
halona River delta (Fig. 1C). The two sediment metaborate flux following the methods of Mur- and Clark, 1989; Clark et al., 1990). The broader
cores, IDLE-TOW10-4B (herein referred to as ray et al. (2000). Aliquots of the bulk, >1 mm, absorption feature centered at ~2150 nm is char-
TOW4) and IDLE-TOW10–5B (herein referred 0.25–1  mm, and 63–250 µm input sediment acteristic of serpentine (King and Clark, 1989;
to as TOW5), were both collected from Lake grain-size separates were ground to <100 µm Clark et al., 1990; Fig. 2D); however, the under-
Towuti during the 2010 field season using a prior to flux fusion. All data are reported as wt% lying structural cause of this absorption is not
modified Kullenberg piston corer. elemental oxide, with all Fe reported as FeO fully understood, although it is likely to be re-
(see the GSA Data Repository for additional lated to fundamental vibrational modes of Mg-
Initial Sample Preparation details on ICP-AES methods [see footnote 1]). OH (Clark et al., 1990).
Additionally, all of the input sediment spectra
The Mahalona River sediment samples were X-Ray Diffraction have a strong absorption centered near ~1900 nm
freeze-dried, and bulk samples were sieved to (Fig.  2), which is caused by a combination
grain sizes: >1  mm, 0.25–1  mm, 63–250 µm, Powder X-ray diffraction (XRD) data were tone of OH stretch and H-O-H bend from
32–63 µm, and <32 µm. The bed load and river collected for all of the grain-size separates of the H2O (Gaffey et al., 1993; Clark et al., 1990).
mouth samples were dry sieved, while the sus- input sediment samples. Prior to analysis, ali- While the ideal formula of Mg-rich serpentine,
pended load sample was wet sieved and subse- quots of the >1 mm, 0.25–1 mm, and 63–250 µm Mg3Si2O5(OH)4, contains no structural water,
quently freeze-dried. Bulk sample aliquots were input sediment grain-size separates were ground laboratory spectra of pure serpentine com-
also kept for analysis. to <45 µm. Randomly oriented samples were monly show a broad ~1900 nm absorption due
We extracted continuous 1–1.5-m-long subsec- measured using a Bruker D2 PHASER instru- to either adsorbed water or small amounts of
tions (U-channels, ~2.5 × 2.5 cm in cross section) ment, and the data were analyzed using the structural water (King and Clark, 1989; Clark
from the two sediment cores. The U-channels Bruker DIFFRAC.EVA (v3.1) program, with et al., 1990; Fig. 2D).
were frozen and freeze-dried using a custom-built the International Centre for Diffraction Data The input sediment samples also exhibit sev-
polyvinyl chloride (PVC) chamber. Based on our (ICDD) PDF-2 mineral database used for ma- eral spectral differences associated with varia-
analysis of the freeze-dried lake cores, we selected jor mineral phase identifications. The Bruker tions in grain size. The ~2150 nm absorption
subsamples (11 from TOW4 and 10 from TOW5) D2 PHASER instrument uses a CuKα radia- is greatly subdued with decreasing grain size
from each of the two cores for further laboratory (Fig. 2). Additionally, spectra of the finest grain-
analysis. The subsamples were extracted from the size separates (i.e., 32–63 µm and <32 µm) show
original core sections and freeze-dried. a subtle absorption feature centered at ~2200 nm
(Fig.  2), which is caused by a combination
VNIR Spectroscopy
1
GSA Data Repository item 2017076, ad- tone of Al-OH bend and OH stretch within the
ditional detail on methods, Figures  DR1–DR3,
and Table  DR1, is available at http://www structure of an Al-bearing phyllosilicate, such
We measured the VNIR reflectance spectra .­geosociety.org/datarepository/2017 or by request to as kaolinite or montmorillonite (Clark et al.,
of the input and core sediment using an Analyti- editing@geosociety.org. 1990; Bishop et al., 2008). Finally, the shape

808 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

of the ~1900 nm absorption feature exhibits


changes associated with grain-size variations.
The coarsest grain-size separates have a broad,
“boxy” ~1900 nm absorption (Fig.  2), similar
to that seen in laboratory serpentine (King and
Clark, 1989; Clark et al., 1990; Fig.  2D). As
grain size decreases, the width of the ~1900 nm
absorption decreases, and it becomes more
asymmetric (Fig.  2), similar to the ~1900 nm
absorption feature seen for many other phyllo-
silicate minerals (Fig. 2D).
Differences in the spectra of the input sedi-
ment associated with variations in grain size
may be due to either changes in the mineral-
ogy of the samples or spectral effects associated
with variations in grain size of a homogeneous
material (e.g., Pieters, 1983; King and Clark,
1989; Mustard and Hays, 1997). The change in
strength of the ~2150 nm absorption feature is
likely to be due to the latter, as the same trend
was observed by King and Clark (1989) in
spectra of grain-size separates of pure serpen-
tine. However, the cause(s) of the ~2200 nm
absorption and the change in the shape of the
~1900 nm absorption is not as readily apparent.
To investigate this issue further, we took the
coarsest grain-size separate (>1 mm) of the river
mouth sediment and ground it to the same size
fractions as the natural size separates. We then
measured these powdered grain-size separates
with the ASD FieldSpec and compared the spec-
tra of the powdered and natural grain-size sepa-
rates (Fig.  DR1 [see footnote 1]). The change

Figure  2. (A–C) Visible to near-infrared


(VNIR) reflectance spectra of input sedi-
ment from the Mahalona River: (A) river
mouth, (B) bed load, (C) and suspended
load sediment. Top plots show VNIR reflec-
tance for grain-size separates. Dashed lines
are located at ~1400, 1900, 2100, 2200, and
2320 nm. Bottom plots show continuum-
removed reflectance in the ~2200 nm region
showing the subtle Al-OH absorption fea-
ture. Continuum-removed reflectance was
calculated by dividing the original spectrum
by a continuum spectrum calculated using a
convex hull method, which accentuates local
absorption features as opposed to broader
spectral shapes. (D) VNIR reflectance spec-
tra of phyllosilicate minerals from the U.S.
Geological Survey spectral library (Clark
et al., 2007). Dashed lines are located at
~1400, 1900, 2100, 2200, and 2320 nm. Spec-
tra are of samples SapCa-1 (saponite), NG-
1.a (nontronite), NMNH17958 (antigorite
serpentine), KGa-2 (kaolinite), and SWy-1
(montmorillonite).

Geological Society of America Bulletin, v. 129, no. 7/8 809

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

in the strength of the ~2150 nm absorption


occurs in both the natural and powdered grain-
TABLE 1. MAJOR IDENTIFIED PHYLLOSILICATE ABSORPTION FEATURES size separates; however, both the presence of an
Approximate Structural cause Reference(s) ~2200 nm absorption and the narrowing of
absorption center the ~1900 nm absorption at finer grain sizes
(nm)
1390 First overtone of OH stretch King and Clark (1989), occur only in the natural grain-size separates
(primarily serpentine in our samples). Clark et al. (1990) (Fig. DR1 [see footnote 1]).
1900 Combination tone of OH stretch and Clark et al. (1990), Therefore, we interpret these spectral
H-O-H bend in structural/adsorbed H2O. Gaffey et al. (1993)
changes to indicate mineralogic differences in
2150 Characteristic of serpentine, but unassigned; Clark et al. (1990)
likely related to vibrational modes of Mg-OH. the grain-size separates. The ~2200 nm absorp-
2200 Combination tone of Al-OH bend and Clark et al. (1990), tion indicates a proportionally larger component
OH stretch in Al-phyllosilicate. Bishop et al. (2008) of Al-rich phyllosilicate compared to Mg-rich
2320 Combination tone of Mg-OH bend and King and Clark (1989),
OH stretch in serpentine. Clark et al. (1990)
serpentine in the finest grain-size separates. The
narrowing of the ~1900 nm absorption indicates
a change in the location of bound or adsorbed
water. We interpret the cause of this change as a
decrease in the relative amount of Mg-rich ser-
pentine, which has a broad ~1900 nm absorp-
tion (King and Clark, 1989; Clark et al., 1990;
Fig. 2D), in the finest grain-size separates.
The observed qualitative spectral changes
(Fig.  2) can be readily quantified using spec-
tral parameters, such as band depth (Clark and
Roush, 1984; Clark, 1999; Pelkey et al., 2007).
Absorption strength measured as a band depth
(BD) is typically correlated with mineral abun-
dance, and combinations of BD measurements
provide discrimination among closely spaced
absorption features (Clark and Roush, 1984;
Clark, 1999; Pelkey et al., 2007). The primary
parameter we make use of, BD1975:BD1915,
quantifies the shape of the ~1900 nm absorption
feature by comparing the relative absorption
strength at 1975 and 1915 nm (Fig.  3; see the
GSA Data Repository for additional details [see

Figure  3. (A) Schematic explanation of the


BD1975:BD1915 parameter, which quanti-
fies the shape of the ~1900 nm absorption
feature by comparing the relative absorp-
tion strength at 1975 and 1915 nm. Spec-
tra with broader ~1900 nm absorptions
will have larger BD1975:BD1915 values.
(B) BD1975:BD1915 values vs. grain size
for the input river mouth (diamonds), bed
load (circles), and suspended load (squares)
sediment from the Mahalona River. Right
plot shows a histogram of BD1975:BD1915
values for cores TOW4 and TOW5. Note
the similarity in absolute parameter values
for the input and lake core sediment. (C–D)
BD1975:BD1915 values vs. time for (C)
core TOW4 and (D) core TOW5. Thin line
indicates raw data, thick line is a 21 point
moving average, and thickest line is a 101
point moving average. Triangles at bottom
of both plots indicate locations of collected
subsamples.

810 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

footnote 1]). The BD1975:BD1915 parameter is


also strongly negatively correlated to the band
depth at 2200 nm (BD2200; Fig. DR2 [see foot-
note 1]), indicating this parameter accurately
records the spectral differences we observe in
association with variations in grain size.
Application of the BD1975:BD1915 param-
eter to the input sediment spectra shows, as ex-
pected, smaller BD1975:BD1915 values (i.e.,
narrower, more asymmetric ~1900 nm absorp-
tions) for finer grain-size samples (Fig.  3B).
An exception to this is the >1  mm grain-size
separates, which have BD1975:BD1915 values
that are lower than expected (Fig. 3B). We sug-
gest that this is due to the very low albedo and
low overall ~1900 nm band strength for these
samples (Figs. 2A and 2B).

Lake Cores
VNIR reflectance spectra of the two sediment
cores show the same primary vibrational ab-
sorption features as the input sediment, and both
cores are spectrally dominated by serpentine
along their entire length. However, qualitative
analysis of the spectra from each of the cores
reveals systematic changes in the down-core
spectral signature. This is best observed in varia-
tions in BD1975:BD1915 values for each of the
>1000 spectra from the two cores, which show
coherent variability (Figs. 3C and 3D). The ab-
solute BD1975:BD1915 values of the sediment
cores are also very similar to those of the input
sediment (Fig.  3B), emphasizing the spectral
similarity of the source and sink sediment.
The spectral parameter records from both
cores show the same trends over the past
~20  k.y. (Figs.  3C and 3D), with a period of
high BD1975:BD1915 values prior to ca.  17–
15  ka, followed by a rapid transition to low

Figure 4. Variations in major element chem-


istry from inductively coupled plasma–
atomic emission spectrometry (ICP-AES)
data. (A) Al2O3 (wt%) vs. grain size and
(B) MgO (wt%) vs. grain size for input
river mouth (diamonds), bed load (circles),
and suspended load (squares) sediment
from the Mahalona River. (C, E) Al2O3
(wt%) vs. time and (D, F) MgO (wt%) vs.
time for core TOW4 (C, D) and TOW5 (E,
F) subsamples. The 21 point moving aver-
age of BD1975:BD1915 values versus time
(Figs.  3C and 3D) is shown in the back-
ground for qualitative comparison. The
y-axis scale for BD1975:BD1915 is inverted
for the Al2O3 plot. Note that changes in
major element chemistry track changes in
spectral signature.

Geological Society of America Bulletin, v. 129, no. 7/8 811

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

BD1975:BD1915 values in the Holocene, tionally, all but the >1 mm grain-size separates ment from the Mahalona River and the sink
which persist from ca.  12  ka to the present. contain quartz, identified by 101 and 100 reflec- sediment from the Lake Towuti basin are domi-
Core TOW4 ends at ca. 20 ka, during the phase tion peaks, which both increase in strength with nated by Mg-rich serpentine. This conclusion
of high BD1975:BD1915 values; core TOW5 decreasing grain size. The primary systematic is confirmed by XRD analyses of sample min-
shows low BD1975:BD1915 values (similar to variations in the XRD patterns of the input sedi- eralogy (Fig. 5) and ICP-AES analyses of ma-
Holocene levels) prior to ca. 35 ka that transi- ment grain-size separates are subtle changes to jor element chemistry (Fig. 4; Table DR1 [see
tion to the period of high values by ca.  27  ka the 001 and 100 serpentine peaks (Fig.  5). As footnote 1]). This conclusion is also consistent
(Fig. 3D). grain size decreases, the primary serpentine 001 with the catchment geology of the East Sulawesi
peak develops a notable shoulder at slightly ophiolite, which contains large regions of ser-
Major Element Chemistry (ICP-AES) larger 2θ values. Additionally, the position of pentinized lherzolite (Kadarusman et al., 2004;
the 100 serpentine peak near ~20° 2θ shifts Fig. 1B).
Both the input sediment and lake core sub- to slightly larger 2θ values with decreasing While serpentine is the dominant mineral
samples have distinct variations in their major grain size. phase in all of our samples, we also identify a
element chemistry (Table  DR1 [see footnote distinct, systematically varying Al-rich phyllo-
1]). Al2O3, K2O, NiO, Na2O, and TiO2 concen- DISCUSSION silicate phase in both the input and lake sediment
trations vary coherently and inversely to varia- (Figs.  2–5). We interpret this Al-­phyllosilicate
tions in MgO concentrations. CaO, Cr2O3, FeO, Overview of Lake Towuti Clay Mineralogy to be kaolinite based on three primary obser-
MnO, and SiO2 concentrations show either very vations from both our data and previous work
weak coherence with other major elements or VNIR spectra (Figs.  2 and 3) of river and on Lake Towuti sediment. First, Weber et al.
no systematic variation. lake sediment show that both the source sedi- (2015) identified kaolinite in sediment samples
Systematic trends in major element chem-
istry of both the input sediment samples and
lake core subsamples are best observed through
relative variations in Al2O3 and MgO concen-
tration (Fig.  4). Input sediment samples have
increasing Al2O3 concentration and decreasing
MgO concentration with decreasing grain size
(Figs.  4A and 4B). Thus, MgO concentration
is positively correlated with BD1975:BD1915
values across varying sediment grain sizes,
while the Al2O3 concentration trend is inverse
to that of BD1975:BD1915 (Figs. 3B, 4A, and
4B). The lake core subsamples show a simi-
lar agreement with BD1975:BD1915 values,
with Al2O3 concentration negatively correlated
with BD1975:BD1915 and MgO concentration
positively correlated with BD1975:BD1915
(Figs. 4C–4F).
Additionally, in both the input sediment
samples and the lake core subsamples, the MgO
concentration is generally higher than the Al2O3
concentration, by up to an order of magnitude in
some cases (Fig. 4). This is consistent with our
VNIR spectral interpretation that the samples
are dominated by Mg-rich serpentine with a mi-
nor component of Al-rich phyllosilicate (Fig. 2).

X-Ray Diffraction

Results from XRD analysis of the input


sediment samples indicate that the major min-
eral phase in all of these samples is serpentine
(Fig. 5), consistent with our interpretation from
VNIR reflectance spectroscopy (Fig.  2). Ser- Figure  5. X-ray diffraction patterns for (A) the river mouth, (B) bed load, and
pentine is identified from strong 001 and 002 (C) suspended load sediment from the Mahalona River. Left plots show the diffrac-
reflection peaks, as well as a prominent 100 re- tion pattern over the 2θ range 5°–40°. The major peaks used for mineral interpreta-
flection. All of the samples also have a minor tions are labeled, with the diffraction plane shown in parentheses, with Sm—smectite,
component of smectite, indicated by a weak S—serpentine, K—kaolinite, and Q—quartz. Right plots show the diffraction pattern
001 peak; however, the peak position and height over the 2θ range 18°–22°, showing the subtle shift in the serpentine 100 peak to
do not vary significantly with grain size. Addi- larger 2θ values, consistent with the kaolinite 020 peak.

812 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

from Lake Towuti to the south of our study area In the input sediment, the proportion of ka- nal formulation of the CIA was developed for
based on VNIR spectra that show an ~2200 nm olinite increases relative to serpentine with upper-crustal materials such as granite or ba-
absorption with a distinct shoulder near ~2160 decreasing grain size (Figs.  2–5). In the lake salt (Nesbitt and Young, 1982, 1989; Nesbitt
nm, which is diagnostic of kaolinite (Clark et sediment, the proportion of kaolinite relative and Wilson, 1992); however, our sediment
al., 1990; Bishop et al., 2008). Those sedi- to serpentine is high prior to ca. 35 ka, during samples—derived from the partially serpen-
­
ments were finer grained and more Al-rich than marine isotope stage 3 (MIS 3), lowest from tinized East Sulawesi ophiolite (Kadarusman
our samples, which led to a more diagnostic ca. 27–17 ka, during the Last Glacial Maximum et al., 2004)—have very low concentrations of
~2200 nm absorption. (LGM), and again high from ca.  12  ka to the CaO, Na2O, and K2O (<2 wt%; Table DR1 [see
The variations in the XRD patterns of our present, during the Holocene. There are also footnote 1]), the labile oxides traditionally used
grain-size separates (Fig.  5) further suggest higher-frequency fluctuations in the proportion to calculate CIA. This makes calculation of CIA
the presence of kaolinite. The position of the of kaolinite to serpentine superimposed on these values less robust and more prone to influence
shoulder at larger 2θ values on the serpentine orbital-scale trends (Figs. 3C and 3D). from sample variability and systematic propaga-
001 peak is consistent with the position of the Based on the similarity of both the spectral tion of measurement error. Therefore, we used
kaolinite 001 peak. Additionally, the shift of signature (Fig.  3) and elemental chemistry a modified CIA value to incorporate MgO as
the serpentine 100 peak near ~20° 2θ to larger (Fig. 4; Table DR1 [see footnote 1]) of the in- a labile oxide, since MgO is expected to leach
2θ values with decreasing grain size is consis- put source compared to the lake sink sediment, out of basaltic and mafic materials before Al2O3
tent with an increasing abundance of kaolinite, we conclude that the core sediment is primarily (Nesbitt and Wilson, 1992). The modified CIA
as the kaolinite 020 peak is at slightly larger detrital in origin and that there is no evidence is calculated as:
2θ values than the serpentine 100 peak. for major in situ alteration or authigenesis of
Finally, there are no systematic variations phyllosilicate minerals. This conclusion is con- modified CIA = [Al2O3/(Al2O3 + CaO +
in the weak smectite 001 peak associated with sistent with the relatively low salinity of Lake Na2O + K2O + MgO)] × 100, (1)
changes in grain size (Fig.  5), which would Towuti (Costa et al., 2015), as in situ altera-
be expected if the Al-phyllosilicate were an tion and authigenesis of phyllosilicate minerals where all the oxide values are in molecular pro-
Al-rich smectite, such as montmorillonite. is most common in highly alkaline and saline portions, and higher values indicate increased
Therefore, we conclude that kaolinite is the lakes (e.g., Jones and Weir, 1983; Deocampo et degrees of chemical weathering (Nesbitt and
most probable interpretation for the identified al., 2009). The provenance of this detrital sedi- Young, 1982).
Al-phyllosilicate, and we refer to the identified ment represents a mix of materials from Lake The river input sediment samples show a
Al-phyllosilicate as kaolinite in the rest of the Towuti’s catchment, including primary phases clear trend of increasing modified CIA values
discussion. However, we note that our interpre- from the serpentinized East Sulawesi ophiolite with decreasing grain size (Fig. 6A). The same
tation of kaolinite is not a definitive identifica- (Kadarusman et al., 2004) and pedogenic altera- general trend is observed with traditional CIA
tion and that our data are also potentially con- tion phases from laterites/Oxisols (Golightly values, although there is increased scatter due
sistent with a different Al-rich phyllosilicate. and Arancibia, 1979; Golightly, 1981; Soil Sur- to the low overall concentrations of CaO, Na2O,
In both the input and lake sediment, we vey Staff, 1999). and K2O (Table DR1 [see footnote 1]). The river
tracked the variation in kaolinite content rela- input sediment samples also show a distinct
tive to serpentine using our developed VNIR Variation in Input Sediment Mineralogy trend of increasing TiO2/Al2O3 ratios with de-
parameter BD1975:BD1915. While, strictly creasing grain size (Fig. 6B). Ti-rich phases are
speaking, this parameter measures the shape The systematic variation in the proportion typically more resistant to chemical weather-
of the ~1900 nm absorption due to structural/ of kaolinite to serpentine with changes in grain ing than Al-rich phases, and in heavily leached
adsorbed water, it is a reliable proxy for the size of the input sediment (Figs. 2–5) indicates soils, TiO2/Al2O3 ratios are expected to increase
grain size–dependent mineralogy of our stud- that there is mineralogic sorting operating in in the most weathered portions of the soil profile
ied samples (Fig.  3). Calculated values of the catchment of the Mahalona River, or per- (e.g., Young and Nesbitt, 1998).
BD2200, which quantify the strength of the haps within the river itself, whereby kaolin- Therefore, these data indicate that the fin-
Al-OH absorption centered at ~2200 nm and ite is preferentially accumulated in finer grain est grain-size input sediment has experienced
more accurately capture the variation in ka- sizes compared to serpentine. Al-rich alteration the highest degree of chemical alteration and
olinite abundance, show the same trends as phases, such as kaolinite, are expected only with leaching, including loss of MgO. In light of
those observed in the BD1975:BD1915 pa- progressive alteration of crustal materials due to the chemical gradients that exist in regional
rameter, and the two parameters are strongly the relative stability of Al to leaching (e.g., Nes- soil profiles (Golightly and Arancibia, 1979;
negatively correlated (Fig.  DR2 [see foot- bitt and Young, 1982, 1989; Nesbitt and Wilson, Golightly, 1981), we conclude that the increase
note 1]). However, many of our measured 1992). Lake Towuti’s catchment is characterized in the proportion of kaolinite to serpentine in
samples have no clear absorption at ~2200 by highly weathered laterites/Oxisols that are the finest grain-size separates of the Mahalona
nm (Fig. 2), which results in a BD2200 value enriched in Al, Fe, Cr, and other recalcitrant ele- River input sediment is primarily controlled by
of 0 (Fig. DR2 [see footnote 1]). In contrast, ments relative to the underlying parent bedrock, varying sources of the detrital sediment from
measured BD1975:BD1915 values vary con- which is enriched in Mg, among other leachable the Lake Towuti catchment. The fine, more
tinuously from ~0.5 to 0.9 (Fig.  2), making elements (Golightly and Arancibia, 1979; Go- chemically weathered, kaolinite-enriched
for easier data analysis and interpretation. lightly, 1981; Soil Survey Staff, 1999). sediment is derived from fluvial erosion and
Furthermore, the general subtlety of the The relationship between grain size and transport of the heavily leached soil (Golightly
~2200 nm absorption (Fig.  2) leads to in- alteration can be tested using the chemical in- and Arancibia, 1979; Golightly, 1981). In con-
creased scatter in the trends for BD2200 as dex of alteration (CIA), which is a measure of trast, the coarse, less chemically weathered,
compared to BD1975:BD1915, which again the overall degree of chemical weathering of a ­serpentine-rich sediment is derived from flu-
hinders data interpretation. sample (Nesbitt and Young, 1982). The origi- vial erosion and transport of bedrock, i.e., the

Geological Society of America Bulletin, v. 129, no. 7/8 813

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

serpentinized East Sulawesi ophiolite (Kada- ophiolite. In theory, decreased chemical weath- in the mineralogy and spectral signature of the
rusman et al., 2004). ering intensity would lead to a decreased pro- sediment cores (Figs. 3C and 3D) are primarily
portion of kaolinite in the weathered soil (Go- indicative of grain size–controlled variations in
Variation in Lake Core Mineralogy lightly and Arancibia, 1979; Nesbitt and Young, the relative proportion of (fine) kaolinite com-
1982, 1989; Nesbitt and Wilson, 1992). Assum- pared to (coarse) serpentine accumulated at the
Our ~40  k.y. VNIR spectroscopy record of ing no change in physical transport processes, core sites over the past ~40 k.y. This conclusion
clay mineralogy from the Mahalona River delta this should result in a decrease in the amount of is also consistent with the sedimentologic analy-
sediment shows coherent changes in the relative kaolinite input to the basin relative to serpentine. sis of Vogel et al. (2015).
abundances of kaolinite and serpentine through Changes in the dominant grain size accu-
time (Figs.  3C and 3D). We hypothesize that mulated at the core sites should also result in VNIR Spectroscopy as a
this variation in mineralogy is controlled by sys- a change in mineralogy given the strong grain Paleoenvironmental Record
tematic changes over time in: (1) the chemical size dependence we observe in input sediment
weathering intensity on the landscape and/or composition (Figs. 2–5). Decreased input (e.g., Modern Holocene sediment in both cores
(2) the relative proportion of fine versus coarse by changes in the grain-size distribution of input is composed of fine-grained material that is
sediment accumulated at the core sites. sediment) or accumulation (e.g., by changes in enriched in kaolinite (Figs.  3C and 3D). Dur-
Changes in the chemical weathering intensity hydrodynamic sorting processes during trans- ing MIS 3, prior to ca. 35 ka, sediment deposi-
on the landscape would change the proportion port and deposition) of fine-grained sediment at tion at the TOW5 site was similar to Holocene
of kaolinite produced from chemical weather- the distal margins of the Mahalona River delta conditions, with deposition of fine-grained,
ing of the partially serpentinized East Sulawesi would result in a decrease in the proportion of kaolinite-enriched sediment. This was followed
kaolinite compared to serpentine in the cores. by a gradual transition to an interval composed
To test between these two hypotheses, we ex- of serpentine-rich, coarse-grained sediment,
amined lacustrine grain-size separates of three from ca. 27 to 17 ka, i.e., during the LGM. We
sediment core subsamples collected from pe- have no record of MIS 3 in core TOW4, but
lagic clay beds spanning a wide range of ages both cores record a relatively abrupt ~3–5 k.y.
(ca.  11.6  ka to ca.  32.0  ka). While the fine- transition during the last glacial termination to
grained nature of the samples precludes analyses the regime of kaolinite-enriched, fine-grained
of coarse sediment fractions, VNIR reflectance sediment deposition occurring from ca. 12 ka to
data from these subsamples show that the finest the present.
grain-size separates (<32 µm) have a stronger What was the cause for the period of de-
~2200 nm absorption and a weaker ~2320 nm position of serpentine-rich, coarse-grained
absorption than the coarser, 32–63 µm fraction sediment during the LGM? One potential ex-
(Fig. 7). This observation is consistent with an planation is an increase in the discharge of the
increasing component of kaolinite relative to Mahalona River, which might be expected to
serpentine in the finest grain-size separates, as lead to increased fluvial transport of coarse
is also seen for modern input sediment (Figs. 2 sediment. However, increased surface runoff
and 3). Similarly, ICP-AES analyses show that should also deliver more fine-grained, kaolin-
the lake core subsamples have an MgO/Al2O3 ite-enriched sediment to the basin, which may
ratio that is very similar to values for the same lead to a nonlinear net change in the proportion
grain sizes of modern input sediment (Fig. 8). of transported serpentine to kaolinite with in-
Taken together, these data suggest that there creasing discharge. Furthermore, recent work
are similar proportions of kaolinite to serpen- on geochemical and sedimentological proxy
tine in the silt and clay grain-size fractions of records from Lake Towuti has shown that
both the modern input sediment and the lake during the last glacial period, from ca.  33 to
core sediment. Furthermore, the lake core sub- 16 ka, this part of Indonesia experienced rela-
sample grain-size separates have similar modi- tively drier conditions compared with the Ho-
fied CIA values (Fig. 6A) and TiO2/Al2O3 ratios locene and MIS 3 (Russell et al., 2014; Costa
(Fig. 6B) as modern input sediment, indicating et al., 2015; Tamuntuan et al., 2015; Vogel et
similar degrees of chemical weathering for the al., 2015).
Figure  6. (A) Modified chemical index of same grain-size fractions through time. These Increased hydrodynamic sorting processes,
alteration (CIA) values and (B) TiO2/Al2O3 data argue against dramatic changes in chemi- namely, winnowing of fine-grained sediment
ratios vs. grain size for the input river cal weathering intensity at this site over the from strong bottom currents, could also lead to a
mouth (diamonds), bed load (circles), and past ~40 k.y. decrease in the amount of kaolinite accumulated
suspended load (squares) sediment from While this result may seem surprising in light at the core sites during the LGM. Changes in
the Mahalona River, as well as the grain- of paleoclimate records that document a drier hydrodynamic sorting processes within the ba-
size separates of lake core subsamples C4- climate from ca. 33 to 16 ka in this region dur- sin would be expected in association with major
7, C5-5, and C5-8 (triangles). Symbol for ing the LGM (e.g., Russell et al., 2014), we sug- changes in lake level and circulation, and indeed
the lake core subsamples is the mean value, gest that the Lake Towuti basin remained warm previous workers have shown that water levels
and error bars denote the range of values. and moist enough to maintain strong chemical in Lake Towuti dropped during the drier LGM
TiO2/Al2O3 ratios were calculated in molar weathering throughout this time period. There- interval (Russell et al., 2014; Costa et al., 2015;
proportions. fore, we conclude that the observed changes Tamuntuan et al., 2015; Vogel et al., 2015).

814 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

Figure 7. Visible to near-infrared (VNIR) reflectance spectra of the grain-size separates of lake core subsamples (A) C4-7, (B) C5-5,
and (C) C5-8. Top plots show full VNIR reflectance spectra. Dashed lines are located at ~1400, 1900, 2100, 2200, and 2320 nm. Bottom
plots show continuum-removed reflectance in the ~2200–2400 nm region, showing a stronger ~2200 nm absorption and a weaker ~2320
nm absorption for the <32 μm grain-size fraction. Dashed lines are located at ~2205 and 2325 nm. Continuum-removed reflectance was
calculated by dividing the original spectrum by a continuum spectrum calculated using a convex hull method.

However, Vogel et al. (2015) demonstrated that Instead, we interpret the increase in coarse- Tamuntuan et al., 2015; Vogel et al., 2015),
water level changes in the Lake Towuti basin grained, serpentine-rich sediment in the two should cause progradation of the Mahalona
during the LGM were on the order of tens of cores during the LGM as a signal of multiple River delta, exposure of coarse-grained topset
meters, which would be insufficient to drive sig- sedimentologic processes internal to the Lake sediment, and increased river incision due to the
nificant changes in lake circulation and hydro- Towuti basin driven by the Mahalona River fall in local base level.
dynamic sorting processes affecting the distal and delta responding to the external forcing Delta progradation would cause increased
portions of the Mahalona River delta. of a drier climate. Lower water levels in Lake coarse sediment deposition further into the ba-
Towuti, such as those experienced during the sin, including at the core sites, which would
LGM (Russell et al., 2014; Costa et al., 2015; have been enhanced by the readily available
coarse-grained topset sediment exposed from
lowered lake levels. As suggested by sedimen-
tologic analyses (Vogel et al., 2015), additional
Figure 8. MgO/Al2O3 (in molar proportions) coarse sediment was delivered to the core sites
vs. grain size for the input river mouth (dia- through discrete slumping and turbidity current
monds), bed load (circles), and suspended events along the delta front. The drop in local
load (squares) sediment from the Mahalona base level from lower Lake Towuti water lev-
River, as well as the grain-size separates els would also have driven increased bedrock
of lake core subsamples C4-7, C5-5, and incision by the Mahalona River. The resultant
C5-8 (triangles). Symbol for the lake core increase in fluvially transported coarse-grained,
subsamples is the mean value, and error bedrock-derived sediment relative to fine-
bars denote the range of values (range for grained, soil-derived sediment would cause an
<32 μm samples is smaller than the symbol). increase in the amount of serpentine delivered

Geological Society of America Bulletin, v. 129, no. 7/8 815

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

to the basin relative to kaolinite. All of these our VNIR spectroscopy record captures a cyclic oscillate on a 21,000 yr cycle due to Earth’s pre-
processes would result in increased deposition variation in the Mahalona River delta miner- cession (Short et al., 1991; Fig. 9).
of coarse-grained, serpentine-rich sediment alogy with a period of ~11  k.y. superimposed In the present day, seasonal precipitation at
relative to fine-grained, kaolinite-enriched sedi- upon the MIS 3/2/1 trends (Fig. 9). Lake Towuti does peak during March and April
ment at the distal deltaic core sites during the Interestingly, this periodicity is suggestive of as the tropical rain belt migrates over the equa-
drier LGM climate, a signal that is accurately a half-precession rainfall cycle. At the equato- tor (Russell et al., 2014). Speleothem-based
captured by our VNIR spectroscopy record rial latitude of Lake Towuti, insolation is stron- rainfall reconstructions from nearby Borneo
(Figs. 3C and 3D). gest during both the spring and fall equinoxes also suggest that changes in rainfall and convec-
A comparison of our VNIR spectroscopy re- (Berger, 1978; Fig.  9), which coincide with tion responded to variations in boreal fall (Sep-
cord to previous paleoclimate reconstructions the twice-annual passage of the tropical rain tember to November) insolation intensity during
from Lake Towuti (Russell et al., 2014) shows belt over the equator. Equatorial rainfall might the last glacial cycle (Carolin et al., 2013). How-
highly consistent trends (Fig. 9); however, there therefore exhibit a half-precession cycle if the ever, Lake Towuti experiences a dry season dur-
are notable systematic differences between our intensity of rainfall within the tropical rain belt ing boreal fall due to local air-sea interactions
VNIR spectroscopy record and the record pre- responds to the strength of local insolation dur- (Hendon, 2003). More importantly, during the
sented by Russell et al. (2014). In particular, ing the spring and fall equinoxes, which each past 40,000 yr, boreal spring and fall insolation
at the equator peaked at ~6, 17, 28, and 39 k.y.
B.P., which correspond to drier intervals in our
record (Fig. 9).
Instead, wetter time periods in our data align
most closely with insolation maxima during
austral and boreal summer (December and June,
respectively; Fig.  9). Thus, our data may indi-
cate that the half-precession cycle observed at
Lake Towuti reflects the intensification of the
austral and boreal summer Asian monsoons.
Similar dynamics have been observed in a re-
cord from equatorial East Africa (Verschuren
et al., 2009); however, a longer paleoclimate
record from Lake Towuti is needed to robustly
document the relationship between insolation
and rainfall through the Pleistocene.
Regardless, the general coherence between
our VNIR spectroscopy record and a more tradi-
tional geochemical proxy record (Russell et al.,
2014; Fig. 9) clearly indicates that our record ac-
curately captures the major paleoclimatic forc-
ings affecting this basin over the past ~40 k.y.
Our approach of core scanning VNIR reflec-
tance spectroscopy provides an effective method
for rapidly and nondestructively characterizing
the phyllosilicate mineralogy of terrestrial lake
sediment. Our results show that this technique
can be used to develop high-­resolution records
of paleoenvironmental change, similar to re-
cords from more traditionally used geochemical
proxies (e.g., Hughen et al., 2004; Castaneda et
al., 2009; Russell et al., 2014).

Figure 9. (A) Previously published paleoclimate reconstruction from Lake Towuti sediment, Implications for Paleolakes on Mars
d13Cwax (Russell et al., 2014), compared to (B–C) BD1975:BD1915 values vs. time for core
TOW4 (B) and core TOW5 (C) and (D–G) equatorial insolation curves. Note the coher- The ultramafic nature of the East Sulawesi
ence of the overall structure for the top three records (A–C). Interpretations for driving ophiolite (Monnier et al., 1995; Kadarusman
mechanisms of variation in both d13Cwax and BD1975:BD1915 values are shown on right. et al., 2004) makes the Lake Towuti catchment
Vertical gray bars highlight periods with insolation maxima at the winter and summer sol- area compositionally comparable to the mafic,
stices, which correspond to wetter time intervals in our visible to near-infrared (VNIR) basaltic martian crust (e.g., Mustard et al., 2005;
spectroscopy record. The Russell et al. (2014) paleoclimate reconstruction was developed us- McSween et al., 2009); however, the specific
ing d13Cwax, a proxy for regional precipitation based on carbon isotopic values of terrestrial alteration mineral phases that dominate the in-
leaf waxes buried in lacustrine sediment. Insolation curves were calculated for the fall (D) put and lake sediment in the Lake Towuti sys-
and spring (E) equinoxes and the winter (F) and summer (G) solstices at 0° latitude using tem (serpentine and kaolinite) are not as com-
the Analyseries software package (Paillard et al., 1996). monly identified on the martian surface as other­

816 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

basaltic weathering products, primarily Fe- and eralogy can also offer detailed paleoenviron- lated to variations in the dominant grain size de-
Mg-rich smectites (e.g., Poulet et al., 2005; mental information (Fig. 9). We thus emphasize posited at the core sites over time. Fine-grained,
Bibring et al., 2006; Mustard et al., 2008; Ehl- both the importance and utility of considering Al-phyllosilicate–enriched sediment was depos-
mann et al., 2011; Carter et al., 2013, 2015). hydrodynamic sorting as a mechanism for driv- ited at the core locations during the Holocene
Although this is the case, our spectral record of ing mineralogic variation within sedimentary (ca.  12–0  ka) and MIS 3 (prior to ca.  35  ka).
Lake Towuti sediment demonstrates that VNIR deposits on Mars. Furthermore, our results show During the LGM, from ca. 27 to 17 ka, the core
reflectance spectroscopy can provide a detailed that processes of hydrodynamic sorting and the sites received coarse-grained, serpentine-rich
record of paleoenvironmental changes over a resultant effects on sedimentary mineralogy sediment. We conclude that this change in grain
period of tens of thousands of years (Fig.  9). can be accurately characterized using VNIR size and associated mineralogy was driven by
The sedimentary rock record on Mars should spectroscopy in a source-to-sink framework, an a drier regional climate and a drop in the Lake
also preserve distinct paleoenvironmental sig- approach that can be readily applied to orbital Towuti water level (Russell et al., 2014; Costa et
nals that can be studied in a similar fashion with remote sensing studies of the sedimentary rock al., 2015; Tamuntuan et al., 2015; Vogel et al.,
combined analyses of high-resolution, remote record of Mars (e.g., Milliken and Bish, 2010; 2015), which forced delta progradation, expo-
sensing observations of stratigraphy and VNIR Goudge et al., 2015). sure and offshore transport of coarse-grained
spectroscopy (e.g., Wray et al., 2008, 2011; Martian paleolake deposits, in particular, are topset sediment, and increased river incision
Loizeau et al., 2010; Milliken and Bish, 2010; highly suitable for analysis of mineralogy using into serpentine-rich bedrock.
Milliken et al., 2010; Ansan et al., 2011). a source-to-sink approach, because they often Our VNIR spectroscopy record is highly
Major variations in the orbital spectral signa- have a clear regional geologic context; however, consistent with geochemical proxy records for
ture of exposed stratigraphic sections on Mars there have been relatively few source-to-sink the Lake Towuti basin over the past ~40  k.y.
have previously been interpreted to indicate mineralogy studies of paleolake systems on (Russell et al., 2014; Fig.  9), and it captures
changes in environmental weathering condi- Mars (e.g., Milliken and Bish, 2010; Goudge et changes in delta dynamics, lake level, and re-
tions (e.g., Wray et al., 2008; Loizeau et al., al., 2015). Our results from Lake Towuti sug- gional precipitation over this time period. VNIR
2010; Milliken et al., 2010) or protolith mate- gest that future studies of paleolake deposits on reflectance spectroscopy offers a relatively
rial (e.g., Loizeau et al., 2010). However, our re- Mars using VNIR reflectance spectroscopy may rapid, nondestructive method for assessing the
sults from Lake Towuti demonstrate that major be able to extract unique information on fluvial phyllosilicate mineralogy of modern lake sedi-
changes in sedimentary spectral signatures can hydrodynamic sorting processes, in addition to ments (e.g., Bishop et al., 1996, 2003; Vogel et
also be driven by grain size–controlled miner- records of paleoenvironmental change. Such in- al., 2008; Rosén et al., 2010; Lynch et al., 2015;
alogy and the internal response of sedimentary formation will be crucial for testing competing Weber et al., 2015). Our results demonstrate
systems to external environmental forcing fac- hypotheses of the evolution of ancient aqueous the utility of VNIR reflectance spectroscopy, in
tors, and these factors should also play a role in surface environments on Mars and, in particular, combination with a core scanning approach, for
controlling mineralogic variations in sedimen- two of the end-member hypotheses that invoke providing detailed, high-resolution records of
tary environments on Mars. significantly different climatic and geochemi- past environmental change preserved in sedi-
Indeed, in situ analyses from the Mars Ex- cal conditions: a warm and wet early Mars with ment from terrestrial lake basins.
ploration Rovers and Mars Science Laboratory widespread surface weathering (e.g., Craddock Paleolake deposits on Mars are also likely to
have revealed the potential effects of hydrody- and Howard, 2002; Carter et al., 2013, 2015) or contain similar paleoenvironmental informa-
namic sorting and grain size–dependent miner- a cold and icy early Mars with limited surface tion that can be accessed with remote VNIR re-
alogy on the composition of both modern sur- weathering and more widespread subsurface al- flectance spectroscopy. Interpretation of VNIR
face soils (e.g., McGlynn et al., 2012; Meslin et teration (e.g., Ehlmann et al., 2011; Wordsworth reflectance spectroscopy data requires careful
al., 2013) and exposed sedimentary rocks (e.g., et al., 2013; Head and Marchant, 2014). analysis, and we emphasize the importance of
Christensen et al., 2004; Grotzinger et al., 2014; mineralogic grain-size sorting (e.g., Christensen
McLennan et al., 2014; Mangold et al., 2016). CONCLUSIONS et al., 2004; McGlynn et al., 2012; Meslin et
Building upon these results, Fedo et al. (2015) al., 2013; Grotzinger et al., 2014; McLennan et
used laboratory studies to show that hydrody- We used VNIR reflectance spectroscopy to al., 2014; Mangold et al., 2016) as a potential
namic sorting in an eolian environment has a assess the source-to-sink mineralogy of Lake control on observed spectral signatures. Future
significant effect on the composition and min- Towuti, a hydrologically open lake in Indonesia. studies of martian paleolake deposits aimed at
eralogy of sediment derived from an unaltered Our results show that the input sediment from paleoenvironmental reconstructions will greatly
basaltic protolith. Therefore, it is entirely pos- the Mahalona River is spectrally dominated by benefit from combined stratigraphic and spectral
sible that stratigraphic variations in VNIR spec- Mg-rich serpentine, with an increasing compo- analyses, as well as mineralogic analysis in the
tral signatures of exposed martian sedimentary nent of Al-phyllosilicate, which we interpret as framework of a source-to-sink approach, which
deposits are due to grain size–controlled min- kaolinite, with decreasing grain size (Figs. 2 and has been applied to only a select few martian pa-
eralogy, as opposed to weathering or protolith 3B). Two sediment cores collected from the dis- leolake systems to date (e.g., Milliken and Bish,
changes (e.g., Wray et al., 2008; Loizeau et al., tal margins of the Mahalona River delta show 2010; Goudge et al., 2015).
2010; Milliken et al., 2010). similar overall spectral signatures to the input
ACKNOWLEDGMENTS
Although mineralogic variations of martian sediment (Fig.  3), suggesting a limited impact
sedimentary deposits due to weathering or pro- of clay authigenesis on the phyllosilicate miner- We express our gratitude to Associate Editor An-
tolith changes are enticing for inferring paleoen- alogy in the lake. drew Cohen and two anonymous reviewers for their
vironmental changes on Mars (e.g., Wray et The VNIR spectral records from both cores insightful and constructive comments, and to David
al., 2008; Loizeau et al., 2010; Milliken et al., show coherent, systematic trends over the past Schofield for editorial handling. We thank Rebecca
Greenberger, Io Wicaksono, Danielle Cares, Dave
2010), our results from Lake Towuti demon- ~40 k.y. (Figs. 3C and 3D), which we conclude Murray, Joe Orchardo, Tom Kiefer, Taki Hiroi, and
strate that records of grain size–dependent min- were controlled by changes in mineralogy re- especially Kevin Robertson for their help with various

Geological Society of America Bulletin, v. 129, no. 7/8 817

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
Goudge et al.

components of sample preparation, laboratory tech- of America, v.  106, p.  20,159–20,163, doi:10.1073/ als, in Pieters, C.M., and Englert, P.A.J., eds., Remote
niques, and data analysis. We also thank Ralph Mil- pnas.0905771106. Geochemical Analysis: Elemental and Mineralogical
liken for use of his X-ray diffraction instrument and Christensen, P.R., Wyatt, M.B., Glotch, T.D., et al., 2004, Composition: Cambridge, UK, Cambridge University
Mineralogy at Meridiani Planum from the Mini-TES Press, p. 43–77.
helpful discussions, and Hendrik Vogel for sharing a
experiment on the Opportunity rover: Science, v. 306, Goldspiel, J.M., and Squyres, S.W., 1991, Ancient aqueous
preprint of his work and helpful discussions. Goudge p. 1733–1739, doi:10.1126/science.1104909. sedimentation on Mars: Icarus, v. 89, p. 392–410, doi:
gratefully acknowledges support for this work from Clark, R.N., 1999, Spectroscopy of rocks and minerals, and 10.1016/0019-1035(91)90186-W.
the Natural Sciences and Engineering Research Coun- principles of spectroscopy, in Rencz, A.N., ed., Remote Golightly, J.P., 1981, Nickeliferous laterite deposits, in
cil of Canada (NSERC) Postgraduate Scholarships Sensing for the Earth Sciences: Manual of Remote Skinner, B.J., ed., Economic Geology 75th Anniver-
Program (PGSD3-421594-2012). Head acknowledges Sensing, Volume 3 (3rd ed.): New York, John Wiley & sary Volume: Lancaster, Pennsylvania, Lancaster Press
support from National Aeronautics and Space Admin- Sons, Inc., p. 3–58. Inc., p. 710–735.
istration (NASA) Mars Data Analysis Program grant Clark, R.N., and Roush, T.L., 1984, Reflectance spectro­ Golightly, J.P., and Arancibia, O.N., 1979, The chemical
scopy: Quantitative analysis techniques for remote composition and infrared spectrum of nickel- and iron-
NNX11AI81G. Russell acknowledges support from
sensing applications: Journal of Geophysical Research, substituted serpentine from a nickeliferous laterite pro-
National Science Foundation grant EAR-1401448. v. 89, p. 6329–6340, doi:10.1029/JB089iB07p06329. file, Soroako, Indonesia: Canadian Mineralogist, v. 17,
Clark, R.N., King, T.V.V., Klejwa, M., Swayze, G.A., and p. 719–728.
REFERENCES CITED Vergo, N., 1990, High spectral resolution reflectance Goudge, T.A., Mustard, J.F., Head, J.W., Fassett, C.I., and
spectroscopy of minerals: Journal of Geophysi- Wiseman, S.M., 2015, Assessing the mineralogy of the
Ansan, V., Loizeau, D., Mangold, N., et al., 2011, Stratig- cal Research, v.  95, p.  12,653–12,680, doi:10.1029/ watershed and fan deposits of the Jezero crater paleo-
raphy, mineralogy, and origin of layered deposits in- JB095iB08p12653. lake system, Mars: Journal of Geophysical Research,
side Terby crater, Mars: Icarus, v.  211, p.  273–304, Clark, R.N., Swayze, G.A., Wise, R., Livo, E., Hoefen, T., v. 120, p. 775–808, doi:10.1002/2014JE004782.
doi:10.1016/j.icarus.2010.09.011. Kokaly, R., and Sutley, S.J., 2007, USGS Digital Spec- Grotzinger, J.P., Sumner, D.Y., Kah, L.C., et al., 2014, A
Berger, A., 1978, Long-term variations of daily insola- tral Library splib06a: U.S. Geological Survey Digital habitable fluvio-lacustrine environment at Yellowknife
tion and Quaternary climatic changes: Journal of Data Series 231 (available at http://speclab.cr.usgs.gov/ Bay, Gale crater, Mars: Science, v. 343, p. 1242777-1–
the Atmospheric Sciences, v.  35, p.  2362–2367, spectral.lib06/; last accessed February 2015). 1242777-14, doi:10.1126/science.1242777.
doi:10.1175/1520-0469(1978)035<2362:LTVODI>2 Cohen, A.S., 2003, Paleolimnology: The History and Evolu- Grotzinger, J.P., Gupta, S., Malin, M.C., et al., 2015, Deposition,
.0.CO;2. tion of Lake Systems: Oxford, UK, Oxford University exhumation, and paleoclimate of an ancient lake deposit,
Bibring, J.-P., Langevin, Y., Gendrin, A., et al., 2005, Mars Press, 500 p. Gale crater, Mars: Science, v. 350, p. aac7575-1–aac7575-
surface diversity as revealed by the OMEGA/Mars Costa, K.M., Russell, J.M., Vogel, H., and Bijaksana, 12, doi:10.1126/science.aac7575.
Express observations: Science, v. 307, p. 1576–1581, S., 2015, Hydrological connectivity and mixing of Head, J.W., and Marchant, D.R., 2014, The climate history
doi:10.1126/science.1108806. Lake Towuti, Indonesia, in response to paleoclimatic of early Mars: Insights from the Antarctic McMurdo
Bibring, J.-P., Langevin, Y., Mustard, J.F., et al., 2006, changes over the last 60,000 years: Palaeogeography, Dry Valleys hydrologic system: Antarctic Science,
Global mineralogical and aqueous Mars history de- Palaeoclimatology, Palaeoecology, v. 417, p. 467–475, v. 26, p. 774–800, doi:10.1017/S0954102014000686.
rived from OMEGA/Mars Express data: Science, doi:10.1016/j.palaeo.2014.10.009. Hendon, H.H., 2003, Indonesian rainfall variabil-
v. 312, p. 400–404, doi:10.1126/science.1122659. Craddock, R.A., and Howard, A.D., 2002, The case ity: Impacts of ENSO and local air-sea interac-
Bishop, J.L., Koeberl, C., Kralik, C., Froschl, H., Englert, for rainfall on a warm, wet early Mars: Journal tion: Journal of Climate, v.  16, p.  1775–1790,
P.A.J., Andersen, D.W., Pieters, C.M., and Wharton, of Geophysical Research, v.  107, p.  21-1–21-36, doi:10.1175/1520-0442(2003)016<1775:IRVIOE>2
R.A., 1996, Reflectance spectroscopy and geochemi- doi:10.1029/2001JE001505. .0.CO;2.
cal analyses of Lake Hoare sediments, Antarctica: Im- Cuadros, J., and Dudek, T., 2006, FTIR investigation of the Hughen, K.A., Eglinton, T.I., Xu, L., and Makou, M., 2004,
plications for remote sensing of the Earth and Mars: evolution of the octahedral sheet of kaolinite-smectite Abrupt tropical vegetation response to rapid climate
Geochimica et Cosmochimica Acta, v. 60, p. 765–785, with progressive kaolinization: Clays and Clay Miner- changes: Science, v. 304, p. 1955–1959, doi:10.1126/
doi:10.1016/0016-7037(95)00432-7. als, v. 54, p. 1–11, doi:10.1346/CCMN.2006.0540101. science.1092995.
Bishop, J.L., Anglen, B.L., Pratt, L.M., Edwards, H.G.M., Deocampo, D.M., Cuadros, J., Wing-Dudek, T., Olives, Irwin, R.P., Maxwell, T.A., Howard, A.D., Craddock, R.A.,
Des Marais, D.J., and Doran, P.T., 2003, A spectro­ J., and Amouric, M., 2009, Saline lake diagenesis as and Leverington, D.W., 2002, A large paleolake basin
scopy and isotope study of sediments from the Antarc- revealed by coupled mineralogy and geochemistry of at the head of Ma’adim Vallis, Mars: Science, v. 296,
tic Dry Valleys as analogues for potential paleolakes multiple ultrafine clay phases: Pliocene Olduvai Gorge, p. 2209–2212, doi:10.1126/science.1071143.
on Mars: International Journal of Astrobiology, v.  2, Tanzania: American Journal of Science, v. 309, p. 834– Irwin, R.P., Howard, A.D., Craddock, R.A., and Moore,
p. 273–287, doi:10.1017/S1473550403001654. 868, doi:10.2475/09.2009.03. J.M., 2005, An intense terminal epoch of widespread
Bishop, J.L., Lane, M.D., Dyar, M.D., and Brown, A.J., Ehlmann, B.L., Mustard, J.F., Fassett, C.I., Schon, S.C., fluvial activity on early Mars: 2. Increased runoff and
2008, Reflectance and emission spectroscopy study Head, J.W., Des Marais, D.J., Grant, J.A., and Mur- paleolake development: Journal of Geophysical Re-
of four groups of phyllosilicates: Smectites, kaolinite- chie, S.L., 2008, Clay minerals in delta deposits and search, v. 110, E12S15, doi:10.1029/2005JE002460.
serpentines, chlorites, and micas: Clay Minerals, v. 43, organic preservation potential on Mars: Nature Geosci- Jones, B.F., and Weir, A.H., 1983, Clay minerals of Lake Ab-
p. 35–54, doi:10.1180/claymin.2008.043.1.03. ence, v. 1, p. 355–358, doi:10.1038/ngeo207. ert, an alkaline, saline lake: Clays and Clay Minerals,
Cabrol, N.A., and Grin, E.A., 1999, Distribution, classifica- Ehlmann, B.L., Mustard, J.F., Murchie, S.L., Bibring, J.-P., v. 31, p. 161–172, doi:10.1346/CCMN.1983.0310301.
tion, and ages of martian impact crater lakes: Icarus, Meunier, A., Fraeman, A.A., and Langevin, Y., 2011, Kadarusman, A., Miyashita, S., Maruyama, S., Parkinson,
v. 142, p. 160–172, doi:10.1006/icar.1999.6191. Subsurface water and clay mineral formation during C.D., and Ishikawa, A., 2004, Petrology, geochemis-
Cabrol, N.A., and Grin, E.A., 2001, The evolution of la- the early history of Mars: Nature, v.  479, p.  53–60, try and paleogeographic reconstruction of the East
custrine environments on Mars: Is Mars only hydro- doi:10.1038/nature10582. Sulawesi ophiolite, Indonesia: Tectonophysics, v. 392,
logically dormant?: Icarus, v. 149, p. 291–328, doi:10 ESRI, G.E.B.C.O., NOAA, National Geographic, DeLorme, p. 55–83, doi:10.1016/j.tecto.2004.04.008.
.1006/icar.2000.6530. HERE, Geonames.org, and other contributors, 2015, King, T.V.V., and Clark, R.N., 1989, Spectral charac-
Cabrol, N.A., and Grin, E.A., 2010, Searching for lakes on Ocean Basemap: http://services.arcgisonline.com/ teristics of chlorites and Mg-serpentines using
Mars: Four decades of exploration, in Cabrol, N.A., and ArcGIS/rest/services/Ocean_Basemap/MapServer high-resolution reflectance spectroscopy: Journal
Grin, E.A., eds., Lakes on Mars: Oxford, UK, Elsevier, (last accessed February 2015). of Geophysical Research, v.  94, p.  13,997–14,008,
p. 1–29, doi:10.1016/B978-0-444-52854-4.00001-5. Farmer, V.C., 1974, The Infrared Spectra of Minerals: Lon- doi:10.1029/JB094iB10p13997.
Carolin, S.A., Cobb, K.M., Adkins, J.F., Clark, B., Conroy, don, UK, Mineralogical Society, 539 p., doi:10.1180/ Loizeau, D., Mangold, N., Poulet, F., Ansan, V., Hauber,
J.L., Lejau, S., Malang, J., and Tuen, A.A., 2013, Var- mono-4. E., Bibring, J.-P., Gondet, B., Langevin, Y., Masson,
ied response of western Pacific hydrology to climate Fassett, C.I., and Head, J.W., 2005, Fluvial sedimentary de- P., and Neukum, G., 2010, Stratigraphy in the Mawrth
forcings over the last glacial period: Science, v.  340, posits on Mars: Ancient deltas in a crater lake in the Vallis region through OMEGA, HRSC color imagery
p. 1564–1566, doi:10.1126/science.1233797. Nili Fossae region: Geophysical Research Letters, and DTM: Icarus, v.  205, p.  396–418, doi:10.1016/j
Carter, J., Poulet, F., Bibring, J.-P., Mangold, N., and Mur- v. 32, L14201, doi:10.1029/2005GL023456. .icarus.2009.04.018.
chie, S., 2013, Hydrous minerals on Mars as seen by Fassett, C.I., and Head, J.W., 2008, Valley network-fed, open- Lynch, K.L., Horgan, B.M., Munkata-Marr, J., Hanley,
the CRISM and OMEGA imaging spectrometers: Up- basin lakes on Mars: Distribution and implications for H., Schneider, R.J., Rey, K.A., Spear, J.R., Jackson,
dated global view: Journal of Geophysical Research, Noachian surface and subsurface hydrology: Icarus, W.A., and Ritter, S.M., 2015, Near-infrared spectro­
v. 118, p. 831–858, doi:10.1029/2012JE004145. v. 198, p. 37–56, doi:10.1016/j.icarus.2008.06.016. scopy of lacustrine sediments in the Great Salt Lake
Carter, J., Loizeau, D., Mangold, N., Poulet, F., and Bibring, Fedo, C.M., McGlynn, I.O., and McSween, H.Y., 2015, Desert: An analog study for martian paleolake basins:
J.-P., 2015, Widespread surface weathering on Mars: A Grain size and hydrodynamic sorting controls on Journal of Geophysical Research, v. 120, p. 599–623,
case for a warmer and wetter climate: Icarus, v.  248, the composition of basaltic sediments: Implica- doi:10.1002/2014JE004707.
p. 373–382, doi:10.1016/j.icarus.2014.11.011. tions for interpreting martian soils: Earth and Plan- Malin, M.C., and Edgett, K.S., 2003, Evidence for per-
Castaneda, I.S., Mulitza, S., Schefuss, E., dos Santos, etary Science Letters, v. 423, p. 67–77, doi:10.1016/j sistent flow and aqueous sedimentation on early
R.A.L., Sinninghe Damste, J.S., and Schouten, S., .epsl.2015.03.052. Mars: Science, v.  302, p.  1931–1934, doi:10.1126/
2009, Wet phases in the Sahara/Sahel region and hu- Gaffey, S.J., McFadden, L.A., Nash, D., and Pieters, C.M., science.1090544.
man migration patterns in North Africa: Proceedings of 1993, Ultraviolet, visible, and near-infrared reflectance Mangold, N., Thompson, L.M., Forni, O., et al., 2016, Com-
the National Academy of Sciences of the United States spectroscopy: Laboratory spectra of geologic materi- position of conglomerates analyzed by the Curiosity

818 Geological Society of America Bulletin, v. 129, no. 7/8

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user
A 40,000 yr record of clay mineralogy at Lake Towuti, Indonesia

rover: Implications for Gale crater crust and sediment ment chemistry of lutites: Nature, v. 299, p. 715–717, ronment at Meridiani Planum, Mars: Science, v. 306,
sources: Journal of Geophysical Research, v.  121, doi:10.1038/299715a0. p. 1709–1714, doi:10.1126/science.1104559.
p. 353–387, doi:10.1002/2015JE004977. Nesbitt, H.W., and Young, G.M., 1989, Formation and dia- Tamuntuan, G., Bijaksana, S., King, J., Russell, J., Fauzi,
McGlynn, I.O., Fedo, C.M., and McSween, H.Y., 2012, genesis of weathering profiles: The Journal of Geology, U., Maryunani, K., Aufa, N., and Safiuddin, L.O.,
Soil mineralogy at the Mars Exploration Rover land- v. 97, p. 129–147, doi:10.1086/629290. 2015, Variation of magnetic properties in sediments
ing sites: An assessment of the competing roles of Paillard, D., Labeyrie, L., and Yiou, P., 1996, Macintosh pro- from Lake Towuti, Indonesia, and its paleoclimatic
physical sorting and chemical weathering: Journal gram performs time-series analysis: Eos (Washington, significance: Palaeogeography, Palaeoclimatology,
of Geophysical Research, v.  117, E01006, doi:10 D.C.), v. 77, p. 379, doi:10.1029/96EO00259. Palaeoecology, v.  420, p.  163–172, doi:10.1016/j
.1029/2011JE003861. Pelkey, S.M., Mustard, J.F., Murchie, S., et al., 2007, .palaeo.2014.12.008.
McLennan, S.M., Anderson, R.B., Bell, J.F., et al., 2014, CRISM multispectral summary products: Param- Verschuren, D., Sinninghe Damsté, J.S., Moernaut, J., et al.,
Elemental geochemistry of sedimentary rocks eterizing mineral diversity on Mars from reflectance: 2009, Half-precessional dynamics of monsoon rainfall
at Yellowknife Bay, Gale crater, Mars: Science, Journal of Geophysical Research, v.  112, E08S14, near the East African equator: Nature, v. 462, p. 637–
v.  343, p.  1244734-1–1244734-10, doi:10.1126/ doi:10.1029/2006JE002831. 641, doi:10.1038/nature08520.
science.1244734. Petit, S., Madejova, J., Decarreau, A., and Martin, F., Vogel, H., Rosén, P., Wagner, B., Melles, M., and Persson,
McSween, H.Y., Taylor, G.J., and Wyatt, M.B., 2009, El- 1999, Characterization of octahedral substitutions P., 2008, Fourier transform infrared spectroscopy, a
emental composition of the martian crust: Science, in kaolinites using near infrared spectroscopy: Clays new cost-effective tool for quantitative analysis of bio-
v. 324, p. 736–739, doi:10.1126/science.1165871. and Clay Minerals, v.  47, p.  103–108, doi:10.1346/ geochemical properties in long sediment records: Jour-
Meslin, P.-Y., Gasnault, O., Forni, O., et al., 2013, Soil diver- CCMN.1999.0470111. nal of Paleolimnology, v. 40, p. 689–702, doi:10.1007/
sity and hydration as observed by ChemCam at Gale Pieters, C.M., 1983, Strength of mineral absorption features s10933-008-9193-7.
crater, Mars: Science, v. 341, p. 1238670-1–1238670- in the transmitted component of near-infrared reflected Vogel, H., Russell, J.M., Cahyarini, S.Y., Bijaksana, S.,
10, doi:10.1126/science.1238670. light: First results from RELAB: Journal of Geo- Wattrus, N., Rethemeyer, J., and Melles, M., 2015,
Milliken, R.E., and Bish, D.L., 2010, Sources and sinks of physical Research, v. 88, p. 9534–9544, doi:10.1029/ Depositional modes and lake-level variability at Lake
clay minerals on Mars: Philosophical Magazine, v. 90, JB088iB11p09534. Towuti, Indonesia, during the past ~29 kyr BP: Jour-
p. 2293–2308, doi:10.1080/14786430903575132. Poulet, F., Bibring, J.-P., Mustard, J.F., et al., 2005, Phyl- nal of Paleolimnology, v. 54, p. 359–377, doi:10.1007/
Milliken, R.E., Grotzinger, J.P., and Thomson, B.J., 2010, losilicates on Mars and implications for early martian s10933-015-9857-z.
Paleoclimate of Mars as captured by the stratigraphic climate: Nature, v.  438, p.  623–627, doi:10.1038/ Weber, A.K., Russell, J.M., Goudge, T.A., Salvatore, M.R.,
record in Gale crater: Geophysical Research Letters, nature04274. Mustard, J.F., and Bijaksana, S., 2015, Tracing clay
v. 37, L04201, doi:10.1029/2009GL041870. Rosén, P., Vogel, H., Cunningham, L., Reuss, N., Conley, mineralogy in Lake Towuti, Indonesia, with reflec-
Monnier, C., Girardeau, J., Maury, R.C., and Cotton, J., D.J., and Persson, P., 2010, Fourier transform in- tance spectroscopy: Journal of Paleolimnology, v. 54,
1995, Back-arc basin origin for the East Sulawesi ophi- frared spectroscopy, a new method for rapid deter- p. 253–261, doi:10.1007/s10933-015-9844-4.
olite (eastern Indonesia): Geology, v. 23, p. 851–854, mination of total organic and inorganic carbon and Wordsworth, R., Forget, F., Millour, E., Head, J.W., Mad-
doi:10.1130/0091-7613(1995)023<0851:BABOFT>2 biogenic silica concentration in lake sediments: Jour- eleine, J.-B., and Charnay, B., 2013, Global modelling
.3.CO;2. nal of Paleolimnology, v. 43, p. 247–259, doi:10.1007/ of the early martian climate under a denser CO2 atmo-
Murchie, S.L., Mustard, J.F., Ehlmann, B.L., et al., 2009, s10933-009-9329-4. sphere: Water cycle and ice evolution: Icarus, v. 222,
A synthesis of martian aqueous mineralogy after 1 Russell, J.M., Vogel, H., Konecky, B.L., Bijaksana, S., p. 1–19, doi:10.1016/j.icarus.2012.09.036.
Mars year of observations from the Mars Recon- Huang, Y., Melles, M., Wattrus, N., Costa, K., and Wray, J.J., Ehlmann, B.L., Squyres, S.W., Mustard, J.F.,
naissance Orbiter: Journal of Geophysical Research, King, J.W., 2014, Glacial forcing of central Indone- and Kirk, R.L., 2008, Compositional stratigraphy
v. 114, E00D06, doi:10.1029/2009JE003342. sian hydroclimate since 60,000 y B.P.: Proceedings of clay-bearing layered deposits at Mawrth Vallis,
Murray, R.W., Miller, D.J., and Kryc, K.A., 2000, Analysis of the National Academy of Sciences of the United Mars: Geophysical Research Letters, v.  35, L12202,
of major and trace elements in rocks, sediments, and States of America, v. 111, p. 5100–5105, doi:10.1073/ doi:10.1029/2008GL034385.
interstitial waters by inductively coupled plasma– pnas.1402373111. Wray, J.J., Milliken, R.E., Dundas, C.M., et al., 2011, Colum-
atomic emission spectrometry (ICP-AES): ODP Tech- Short, D.A., Mengel, J.G., Crowley, T.J., Hyde, W.T., and bus crater and other possible groundwater-fed paleo­
nical Note, v. 29, p. 1–27. North, G.R., 1991, Filtering of Milankovitch cycles lakes of Terra Sirenum, Mars: Journal of Geophysical
Mustard, J.F., and Hays, J.E., 1997, Effects of hyperfine par- by Earth’s geography: Quaternary Research, v.  35, Research, v. 116, E01001, doi:10.1029/2010JE003694.
ticles on reflectance spectra from 0.3 to 25 µm: Icarus, p. 157–173, doi:10.1016/0033-5894(91)90064-C. Young, G.M., and Nesbitt, W.N., 1998, Processes control-
v. 125, p. 145–163, doi:10.1006/icar.1996.5583. Sly, P.G., 1978, Sedimentary processes in lakes, in Ler- ling the distribution of Ti and Al in weathering pro-
Mustard, J.F., Poulet, F., Gendrin, A., Bibring, J.-P., Lan- man, A., ed., Lakes: Chemistry, Geology, Phys- files, siliciclastic sediments and sedimentary rocks:
gevin, Y., Gondet, B., Mangold, N., Bellucci, G., and ics: New York, Springer-Verlag, p.  65–89, Journal of Sedimentary Research, v.  68, p.  448–455,
Altieri, F., 2005, Olivine and pyroxene diversity in doi:10.1007/978-1-4757-1152-3_3. doi:10.2110/jsr.68.448.
the crust of Mars: Science, v.  307, p.  1594–1597, Soil Survey Staff, 1999, Soil Taxonomy: A Basic System
doi:10.1126/science.1109098. of Soil Classification for Making and Interpreting Soil
Mustard, J.F., Murchie, S.L., Pelkey, S.M., et al., 2008, Hy- Surveys (2nd ed.): U.S. Department of Agriculture
drated silicate minerals on Mars observed by the Mars Handbook 436, Natural Resources Conservation Ser-
Reconnaissance Orbiter CRISM instrument: Nature, vice, 886 p. Science Editor: David I. Schofield
v. 454, p. 305–309, doi:10.1038/nature07097. Associate Editor: Andrew Cohen
Squyres, S.W., Arvidson, R.E., Bell, J.F., et al., 2004a, The
Nesbitt, H.W., and Wilson, R.E., 1992, Recent chemical Spirit rover’s Athena science investigation at Gusev Manuscript Received 10 May 2016
weathering of basalts: American Journal of Science, crater, Mars: Science, v. 305, p. 794–799, doi:10.1126/ Revised Manuscript Received 2 November 2016
v. 292, p. 740–777, doi:10.2475/ajs.292.10.740. science.3050794. Manuscript Accepted 30 December 2016
Nesbitt, H.W., and Young, G.M., 1982, Early Proterozoic Squyres, S.W., Grotzinger, J.P., Arvidson, R.E., et al.,
climates and plate motions inferred from major ele- 2004b, In situ evidence for an ancient aqueous envi- Printed in the USA

Geological Society of America Bulletin, v. 129, no. 7/8 819

Downloaded from https://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/129/7-8/806/1002176/806.pdf


by Universiteit Ghent user

You might also like