Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Catena 212 (2022) 106113

Contents lists available at ScienceDirect

Catena
journal homepage: www.elsevier.com/locate/catena

Paleosols record dry and humid paleoenvironments during the Upper


Pleistocene in the Brazilian Pantanal
Francisco Sérgio Bernardes Ladeira a, *, Patricia Colombo Mescolotti b,
Fabiano do Nascimento Pupim c, Laura Milani Dias Mathias de Faria a, Mario Luis Assine b
a
Laboratory of Pedology (LabPed), Department of Geography, Institute of Geosciences, UNICAMP, R. Carlos Gomes, 250, Campinas, São Paulo 13083-855, Brazil
b
Departament of Geology, Institute of Geosciences and Exact Sciences-UNESP, Avenida 24A, 1515, Rio Claro 13506-900, Brazil
c
Department of Environmental Sciences, Federal University of São Paulo, Diadema, Rua São Nicolau, 210, São Paulo, SP, 09913-030, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Information on Quaternary landscapes and paleoenvironmental changes in the interior of Brazil are scarce. One
Paleoclimate of such areas, particularly sensitive to these variations, is the modern tropical Pantanal wetland, whose alluvial
Pedocomplex deposits record paleoenvironmental and paleoclimatic changes during the late Quaternary. Based on detailed
Optically stimulated luminescence (OSL)
pedological and sedimentological descriptions, stimulated optical luminescence dating and analytical data we
described a well-exposed sedimentary succession that crops out in slope terraces of the Aquidauana River,
located in the southern area of the Pantanal. We interpret environmental changes that have occurred since the
beginning of the Upper Pleistocene, and which affected sedimentation and biogeochemical processes. Basal
deposits comprise cross-stratified sands, which record active fluvial sedimentation from ~106 to ~70 ka, a
period of probable increased river discharge. Fluvial channel facies are overlaid by pedogeneized floodplain
deposits, in which we recognized a pedocomplex composed of a sequence of three late Pleistocene truncated
paleosol profiles and a Holocene soil profile. The two lower paleosols are Vertisols with mottling, slickensides,
rhizolites and septarian glaebules of carbonate. These characteristics indicate prolonged drought periods under
semi-arid to arid climates ca. ~70 and ~30 ka. The upper paleosol profile is a Planosol (Ultisol) with block
structure, waxiness, iron glaebules, and manganese films, pointing to prevailing humid climates during the
terminal Pleistocene (since ~30 ka). Covering this truncated Planosol is a layer of fine to medium-grained
Holocene sands, with development of a horizon A of an Arenosol (Entisol). These paleoenvironmental changes
correlate with regional changes in precipitation deduced from speleothems in nearby plateaus and in south­
eastern Brazilian sites.

1. Introduction pedocomplex (Morrison, 1967; Stremme, 1998; Karlstrom et al., 2008;


Feng and Wang, 2006; Jacobs et al., 2009), and their analysis is a central
Vertical and horizontal relationships between pedogenesis and issue in studies that relate processes of sedimentation and pedogenesis.
sedimentation are complex in areas of active sedimentation, where the Although widely used in pre-Quaternary paleosols (Alekseeva et al.,
development of soils records non-deposition events. The interaction of 2016), Quaternary deposits are an ideal field of observation for these
processes can obliterate or even erase sediments deposited, or soils relationships (Zhao, 2003; Karlstrom et al., 2008), as the processes that
formed. The non-preservation of soils and sediments is a consequence of form soils/sediments are active or have striking relict characteristics in
soil surface truncation by erosive processes, obliteration of sedimentary the landscape, thus, are models for studies of pre-Quaternary units.
structures, and granulometric modifications by pedogenetic processes The Brazilian Pantanal is a large tropical wetland of the world
(Kraus, 1999; Retallack, 2019). Different soil profiles may present (Assine, 2015a), an interesting area to study the relationships between
superimposed pedogenetic processes, e.g. where the deposit that over­ soils and sediments. It encompasses a wide variety of modern alluvial
laps the profile is thin or when the pedogenetic processes of the upper environments and past sedimentary deposits (Assine and Soares, 2004;
deposit reach the lower profile. These relationships are named Assine et al., 2015a, b) that significantly affect pedogenetic processes

* Corresponding author.
E-mail address: ladeira@unicamp.br (F.S.B. Ladeira).

https://doi.org/10.1016/j.catena.2022.106113
Received 28 April 2021; Received in revised form 24 January 2022; Accepted 1 February 2022
Available online 17 February 2022
0341-8162/© 2022 Elsevier B.V. All rights reserved.
F.S.B. Ladeira et al. Catena 212 (2022) 106113

and soil formation (Oliveira Junior, et al. 2017). The Pantanal is char­ The association between soils/sediments/terrain morphology was
acterized by the presence of fluvial megafans, systems in which rivers pointed out by Nascimento et al. (2013) and Oliveira Junior et al.
frequently change their course due fluvial avulsion (Assine, 2005; Assine (2019), who identified deposits from the Upper Pleistocene in the
et al, 2014; Assine et al., 2016). Several proxies are being used for northern area of the Pantanal. They highlighted that Quartzipsammets
paleoenvironmental reconstructions, with data from lake sediment cores are associated with paleochannels; Natrustalfs and Albaqualfs are
(Whitney et al., 2011; Metcalfe et al., 2014; Fornace et al., 2016; McGlue associated with the old marginal dikes and small elevations; and
et al. 2012, 2017), geomorphological and geochronological data (Assine Aquents, Aquepts, Plinthaquults, and Fluvaquentic Inceptisols are
et al, 2014; Pupim et al., 2017), and O18 delta isotope data obtained developed in old floodplains.
from speleothems inside caves in the high areas surrounding the basin The Aquidauana River megafan is in the southern part of the modern
(Novello et al. 2016, 2017). alluvial depositional tract (Fig. 1). In the catchment basin located to the
In fluvial megafans, small topographic variations produce distinct east, on the Maracaju plateau - Campo Grande, the river carves Paleo­
soil types, as exemplified in the northern Pantanal where topographi­ zoic and Mesozoic sedimentary rocks of the Paraná Basin, whose sedi­
cally low-amplitude sand ridges present conditions of satisfactory ments are the main source of sediments for the megafan (Facincani and
drainage with associated Arenosols, and with clayey soils between the Assine, 2010; Assine et al., 2015a). A secondary source is the sediment
elevations, through where water flows during wet periods (Zeilhofer and from small rivers present on the left bank of the Aquidauana River,
Schessl, 2000; Pupim et al., 2017). Soils associated with active megafans whose catchment area is in crystalline Precambrian terrains located to
were studied by several authors, as well as paleosols from the Pleisto­ the south. The Aquidauana River flows towards NNW, at the top of the
cene and in older stratigraphic units (Mohindra et al. 1992; Mohindra, megafan, deflecting west at the bottom where it forms the current
1995; Srivastava and Parkash, 2002; Srivastava et al., 2007; Fontana depositional lobe. In the upper part of the megafan, the Aquidauana
et al., 2008; Latrubesse et al., 2012; Trendell et al., 2013). River flows in a confined valley, where it forms a modern meander belt.
In the Pantanal, studies relating soils with distinct geomorphological The belt is limited by marginal terraces that reach six meters in height in
features of fluvial megafans are relatively recent (Sousa and Souza, the uppermost part of the megafan, but that decrease downstream until
2013, Nascimento et al., 2013, Cardoso et al., 2016, Oliveira Junior, disappearing at the entrance of the current depositional lobe. In the
2017). They are especially focused on soils developed in the Holocene, depositional lobe, the river acquires a distributional pattern with
seeking to associate the development of soils with recent depositional deflection and several channel-levee complexes with higher altitudes
processes. than the surrounding floodplains, standing out in the landscape
Soil/sediment relationships are complex in ancient sedimentary (Facincani and Assine 2010; Assine et al., 2015a). The climate is tropical
units (Mack and James, 1994; Tabor and Myers, 2015; Zhou et al., 2015) with a dry winter season (Aw), according to the Köppen-Geiger system
including paleosols in old fluvial megafans deposits (Trendell et al., (1928). The rainy season is between November and April, followed by a
2013). Interpretation of past environmental changes using pedological dry season between May and October. July is the driest month. Its
analysis are rare for the Pantanal wetland due to the scarcity of outcrops annual averages for temperature, precipitation, and evapotranspiration
and the lack of geochronological data needed to accurately determine are 23.3 ◦ C, 1323 mm, and 1171 mm, respectively (INMET, 2017).
the age of sedimentary deposits and rates of pedogenetic processes that Prance and Schaller (1982) point out the diversity of the Pantanal
have acted over time. vegetation. Its richness of plant species may be due to a mosaic of
Despite these difficulties, the discovery of an excellent exposure in permanently flooded areas, others that are seasonally flooded, and non-
slope terraces of the Aquidauana River megafan (Anhumas vertical flooded lands, besides its distinct plant domains. According to Pott et al.
section), allowed us to measure a vertical stratigraphic section and (2011), the vegetation cover of the Pantanal sedimentary plain is
recognize events of fluvial sedimentation and pedogenesis that took characterized by a mosaic of aquatic plants, flooded fields, riparian
place from the Upper Pleistocene to the Early Holocene. Three sequen­ forests, cerrado (Brazilian Savanna), cerrado woodland, and deciduous
tial Upper Pleistocene paleosol profiles and a Holocene soil profile were forests. The literature presents few studies on Pantanal paleosols and
recognized and several questions were addressed: a) the time intervals in even fewer on paleosols in the Aquidauana fan region. A particular study
which pedogenesis developed; b) the importance of deposits in defining stands out (Facincani et al., 2009), describing calcretes and Calcic
the morphological characteristics of pedological horizons, as well as the Vertisols in the research area.
identification of diagnostic horizons; c) the association of paleosols with The studied area is located at the coordinates 20◦ 20′ 48′′ S and
paleoclimates; d) the comprehension of pedocomplexes in this type of 55 53′ 25′′ W, with an altitude of 140 m (Figs. 1 and 2), in the megafan

system and their importance in soil morphology; and e) the correlation upper portion. The exposure is significant as the river erodes the mar­
with regional paleoenvironmental events. ginal terrace that is constituted by Pleistocene deposits.

2. Physical settings 3. Materials and methods

The Pantanal is an active sedimentary basin with a sedimentary 3.1. Description and sampling
succession over 500-m-thick (Assine and Soares, 2004; Assine et al.
2015a, b), housing one of the largest tropical wetlands in the world (Por, A vertical section of 7.3 m exposed on the slope of the marginal
1995; Assine, 2015; Assine et al. 2015), and an area of approximately terrace of the Aquidauana River was selected for analysis (Fig. 2 A, B).
150,000 km2. The modern depositional systems tract in Pantanal is The description of sedimentary features and interpretation of deposi­
characterized by fluvial and lacustrine systems mainly related to fluvial tional processes were based on sedimentary textures and structures and
megafans (Assine et al., 2015a, b). Megafans present different charac­ lithological contacts (Fig. 3). The criteria for collecting samples for
teristics of alluvial fans and tributary river systems in terms of di­ dating by Optically Stimulated Luminescence (OSL) prioritized sampling
mensions, gradient, textural characteristics, and depositional processes in different sedimentary features and pedological horizons.
(Leier et al., 2005; Trendell et al., 2013). They also present specific Soil description followed the Manual for Describing and Collecting
characteristics in terms of soil and soil/sediments (Hartley et al., 2013). Soil in the Field (Santos et al., 2013) and the Field Book for Describing
Soil diversity and complexity are great in the Pantanal, and ten soil and Sampling Soils (Schoenerberger et al., 2012). The description
orders according to the Brazilian soil classification system have already included thickness of the horizons, transition between them, soil color
been identified (Brasil, 1982a, b; Fernandes et al. 2007). The diversity of (according to the Munsell Soil Color Chart), structure, texture, consis­
soils is due to small variations in local characteristics of soil formation tency, waxiness, glaebules (nodules and concretions), slickenside, and
factors and is associated with the dynamics of deposits in the megafan. presence of rhizolites. Deformed samples were collected from all

2
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Fig. 1. Location of the study area (modified from Assine et al., 2015).

horizons of all profiles described, and granulometric, chemical, miner­ recommended because of bioturbation that may result in remobilization,
alogical, and dating analyses were performed. We use the WRB/FAO and underestimated ages (Bateman et al., 2003).
classification system, in its most general classification (Reference Soil Sample preparation for equivalent dose (De) measurements followed
Groups) (IUSS Working Group WRB, 2015). standard procedures (Aitken, 1985). Samples were wet-sieved to isolate
the 180–250 µm grain-size fractions. The target fraction was treated
3.2. Optically stimulated luminescence dating with hydrogen peroxide (H2O2, 27%), and hydrochloric acid (HCl, 10%)
to remove organic matter and carbonate minerals. Heavy mineral
OSL dating was performed on quartz sand grains from six samples. (>2.75 g/cm3) and feldspar (<2.62 g/cm3) grains were removed by
Samples were collected along the described vertical section (Fig. 3), heavy liquid separation with lithium metatungstate solution. Then,
except for the sample at the top that was collected a few meters laterally quartz concentrates (2.62 and 2.75 g/cm3) were etched with hydro­
where the horizon is thicker. Sampling close to the surface is not fluoridric acid (HF, 40%) for 40 min to remove remnant feldspar grains

3
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Fig. 2. A. Location of the section studied on the Aquidauana River; B. Sediment exposure due incision of the Aquidauana River channel with delimitation of
pedological profiles. C. Section studied, with delimitation of pedogenetic horizons (white dashed lines).

and etch a ~10 µm outer layer of quartz grains, thus removing the alpha tube (Thorn EMI 9635QB).
particles’ contribution from the radiation dose rate. Aliquots of quartz The single-aliquot regenerative dose (SAR) protocol (Murray and
grains (100–200 grains) were mounted on stainless steel cups for Wintle, 2000) was used to estimate the De (Table 1). A dose recovery test
luminescence measurements on an automated Risø TL/OSL DA-20 was carried out using 6 aliquots with a pre-heat temperature of 240 ◦ C
reader system equipped with a 90Sr/90Y beta source that delivered a that showed a calculated-to-given dose ratio of 0.98 ± 0.02 for a given
dose rate of 0.109 Gy/s, blue LEDs (470 ± 20 nm) operated at 90% dose of 21.8 Gy. The De calculations used aliquots with recycling ratio
power (~40 mW/cm2) for stimulation, and a Hoya U-340 filter for light values between 0.90 and 1.10, recuperation of less than 5%, and
detection in the ultraviolet band (290 and 340 nm) with a bialkali PM negligible feldspar signal tested by infrared stimulation, as suggested by

4
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Fig. 3. Stratigraphic section showing pedological and sedimentological data.

Table 1
Equivalent doses, radiation dose rates and OSL ages from quartz grains of the Aquidauana River, Pantanal - Brazil.
Lab Depth N U (ppm) Th (ppm) K (%) Cosmic dose rate Water sat. Total dose rate Equivalent Dose OD Age (ka)
Code (m) (Gy/ka) (%) (Gy/ka) (Gy) (%)

L0446 0.85 19 0.50 ± 1.33 ± 0.27 ± 0.18 ± 0.01 3.3 0.65 ± 0.04 5.4 ± 0.3 23 8.3 ± 0.7
0.02 0.05 0.01
L0222 0.8 12 0.53 ± 2.52 ± 0.24 ± 0.19 ± 0.01 19.1 0.64 ± 0.03 19.1 ± 1.1 15 29.8 ± 2.3
0.01 0.01 0.01
L0434 2 18 0.56 ± 2.62 ± 0.62 ± 0.15 ± 0.01 44.6 0.77 ± 0.04 44.2 ± 2.1 20 57.1 ± 4.0
0.02 0.08 0.02
L0435 5 23 0.27 ± 0.48 ± 0.22 ± 0.11 ± 0.01 0.7 0.44 ± 0.01 31.5 ± 0.6 8 72.2 ± 5.5
0.02 0.04 0.01
L0436 6.2 24 0.25 ± 0.52 ± 0.15 ± 0.09 ± 0.01 3.0 0.34 ± 0.02 23.4 ± 0.5 10 69.1 ± 4.9
0.02 0.04 0.01
L0437 7.2 22 0.35 ± 0.77 ± 0.17 ± 0.08 ± 0.01 58.6 0.27 ± 0.01 28.8 ± 0.8 13 106.7 ±
0.02 0.04 0.01 5.6

N = number of aliquots; OD = overdispersion.

Murray and Wintle (2000, 2003). Equivalent doses for age dating of each using factors proposed by Guérin et al. (2011). Water saturation was
sample were calculated through the Central Age Model (Galbraith et al. determined by the ratio between water weight and dry sample weight.
1999). The cosmic dose rate contribution was calculated using the sample
For radiation dose-rate calculation, natural radionuclides concen­ depth, elevation, latitude, and longitude, as described by Prescott and
trations were measured through gamma-ray spectrometry in a high- Hutton (1994).
purity germanium detector (HPGe; relative efficiency of 55%; 2.1 keV
energy resolution) encased in an ultralow background shield (Canberra
Industries). U, Th, and K concentrations were converted to dose rates

5
F.S.B. Ladeira et al. Catena 212 (2022) 106113

3.3. Chemical analysis dispersed in water), silt, very coarse, coarse, medium, fine and very fine
sand, degree of flocculation, and textural class (Camargo et al., 2009).
The chemical analysis for pedological analysis routine followed
Camargo et al. (2009), defining the pH, exchangeable aluminum (Al+3), 3.5. Mineralogy
saturation by aluminum, exchangeable sodium (Na+), exchangeable
sodium, calcium, and magnesium saturation index (Ca+2, Mg+2), levels The clay fraction of the subsurface diagnostic horizons was deposited
of P, K, Ca +2, Mg+2, H + Al (potential acidity), carbon, sum of bases on oriented thin films. All preparations were analyzed using a PAN­
(SB), base saturation (V value), total cation exchange capacity (T), and alytical diffractometer (Empyrean) with measurements made by using
phosphorus resin (P-res). We defined the elements through X-Ray CuKα1 radiation (WL = 1.54056 Å), and Ni filter. The 2θ start angle was
Fluorescence Spectrometry analyzes, and sprayed samples with particle 3◦ , and the end angle 65◦ , with 3.8′′ step, and 0.008◦ step size (0.27◦ /s
size smaller than 200 mesh and pressed at 30 T/cm2 in 25% wax. The 8- scan speed, 30 mA current, and 40 kV voltage). The diffractograms were
g-aliquot of samples were dried at 110 ◦ C before pressing. The concen­ interpreted using the X’Pert Highscore Plus software, by the IMA min­
trations of major elements, expressed as a percentage of weight, were eral database (Anthony et al., 2017) supported by the ICDD PDF-2
calculated according to Franzini et al. (1976). The error of major ele­ database. The samples were analyzed in their natural form after being
ments was less than 5%. The loss on ignition (LOI) was obtained by subjected to an atmosphere of ethylene glycol for about 48 h (glyco­
heating the samples at 1000 ◦ C for about 3 h. lagen), and after burning at 300 ◦ C and 550◦ C to confirm the presence of
certain mineral phases.
3.4. Granulometric analysis
4. Results and interpretation
The samples underwent mechanical disaggregation, dispersion, and
evaluation of the relative proportion of primary particles by sedimen­ 4.1. Vertical stratigraphic section
tation in an aqueous medium. The granulometry was carried out in the
air-dried fine soil fraction (TFSA) with a sieve smaller than 2 mm to The vertical section described comprises 7.2 m of siliciclastic de­
determine the percentage distribution of the primary soil particles. The posits (Fig. 3). The basal 7.2–6.0 m interval is a finning-upward channel
pipette method was applied to determine the fractions of clay (total and facies sequence made up of coarse-grained sand, poorly sorted, with

Fig. 4. Base of profile. A. Coarse pebbly sandstone, poorly-sorted, with cross-stratification; B. Intercalation of laminae of fine-grained sand and mud; C. Fine-grained
sandy deposits, well sorted, with cross stratification.

6
F.S.B. Ladeira et al. Catena 212 (2022) 106113

granules and rounded pebbles of quartz and crystalline basement rocks, 2–4, from top to bottom) and a soil horizon at the top (profile 1),
with cross-stratification and cemented by iron oxide (Fig. 4A). The facies comprising a total thickness of 4.2 m (Fig. 2, Table 2). No major
sequence ends in fine-grained facies, with an erosional surface on the top morphological changes have affected the paleosols since their incorpo­
(Fig. 4B). ration into the stratigraphic archive, except the truncation of the top
The 6.0–3.0 m interval records a younger fluvial cycle beginning profile horizons by erosion. At the top, the Profile 1 presents a sandy
with beds of fine/medium cross-stratified sands, which are successively horizon A with dark colors, iron glaebules at its base, overlapping the
overlaid by pebbly cross-stratified sands (Fig. 4C), and a thin sand layer horizon 2Bt1 abruptly (Profile 2), featuring high textural contrast.
of fines that records low energy clay settling. Finally, the succession is The horizon 2Bt1 presents a strong blocky structure and very waxy.
overlaid by trough-stratified sands that grade to pedogeneized silt layers The 2Bt1 and 2Bt2 presents abundant mottled spots, iron glaebules at
towards the top. their intermediate level, and manganese films covering the peds in 2Bt2.
Three paleosol and one soil profiles were characterized at the upper The base of the profile is characterized by a thin, dark horizon, associ­
part of the vertical section. No primary sedimentary structures are ated with a high presence of manganese films (mangans) covering
preserved, but the contact between paleosols profiles are erosional fissure planes. The contact with the profile below is an abrupt and flat
surfaces evidenced by truncation of carbonate rhizolites. The topmost transition, which separates different pedogenetic conditions, indicating
soil profile developed in well-sorted, medium-grained aeolian sands, truncation of the profile positioned below.
with sharp basal contact and laterally up to 1.2 m-thick, that compose Profile 3 (intermediate) presents two Btss horizons (the “t” was used
the uppermost sedimentary unit. to indicate the presence of illuvial clay, not a textural change) with a
parallelepiped structure and without the presence of horizon A. Later­
4.2. Morphology of soil and paleosols ally in the outcrop, a great development (in dimensions and quantity) of
slickensides (usually over 1 m in size) typically associated with Vertisols
The lower 3 m of the stratigraphic section corresponds to a non- (Fig. 5 A, B). The horizons present gray color with abundant mottling,
pedogenized fluvial channel sands (Horizon R, Fig. 3). The pedocom­ and several rhizolites truncated in the contact with the upper profile.
plex sequence comprises three successive paleosol profiles (numbered Manganese films are common on the surface of the peds. Slickensides are

Table 2
Macromorphology.
Horizon Horizon Color Soil Structure Waxy Consistence Glaebules Transition
Depth Texture
(cm)

A 0–24 10YR 6/2 (d), 10YR 4/2 (m) Sandy Moderate fine and – Slightly hard, friable, Many, fine, Very
loam medium granular; slightly sticky, spherical of Fe, at Abrupt
moderate fine slightly plastic the base of the Smooth
subangular blocky horizon
2Bt1 24–33 10YR 5,5/2 (d), 10YR 5/1 Sand Strong medium and Prominent Hard, firm, – Clear
(m);38 mottled 10R 5/8 (d), clay coarse angular blocky continuous many moderately sticky, Smooth
10R 4/8 (m); black mottled loam moderately plastic
2Bt2 33–80 10YR 6/2 (d), 10YR 5,5/2 Clay Strong fine and Prominent Hard, firm, Many, medium Abrupt
(m); glebules cortex 10YR 7/ loam medium angular continuous many moderately sticky, and coarse, Smooth
8 (d), 10YR 6/8 (m); glebules blocky moderately plastic spherical of Fe
core 10YR 2/1 (d), 10YR 2/1
(m)
2C 80–92 10YR 2/1 (d m), abundant Sand Massive – Hard, very firm, Many, medium, Very
mottled 10YR 5/2 (d), clay moderately sticky, spherical of Mn Abrupt
10YR5/1(m) loam moderately plastic Smooth
3Btss1 92–170 Mottled horizon 10YR 6,5/ Sand Strong medium Prominent Extremely hard, Few coarse and Clear
2–7,5YR 6/8 − 10R 3/2 (d), clay prismatic; Strong continuous many slightly rigid, very very coarse Smooth
10YR 5/1–7,5Y 5/8 − 10R 3/ loam medium and coarse sticky, very plastic spherical
2 (m) angular blocky septarian of
CaCO3
3Btss2 170–251 Mottled horizon 10YR 6/ Clay Strong medium and Distinct Extremely hard, Few coarse and Very
2–7,5YR 6/8 − 10YR 2/1 (d), loam coarse angular blocky discontinuous extremely firm, very very coarse Abrupt
10YR 5/2–7,5YR 5/8 − 10YR common sticky, very plastic spherical Smooth
2/1 (m) septarian of
CaCO3
4Bss1 251–299 Mottled horizon 7,5YR 6/8 − Sand Strong medium Distinct Extremely hard, Few coarse and Clear
10YR 6/2 − 5YR 6/8 − 10YR clay angular blocky discontinuous slightly rigid, very very coarse Smooth
2/1 (d); 7,5YR 6/8 − 10YR loam common sticky, very plastic spherical
5/2 − 5YR 5/8 − 10YR 2/1 septarian of
(m) CaCO3
4Bss2 299–352 10YR 7/2 (d), 10YR 6/2 (m); clay Moderate fine and Distinct Extremely hard, Few coarse and Clear
mottled 10YR 6/2 (d), 7,5YR medium angular discontinuous extremely firm, very very coarse Smooth
5/8 (m); black mottled blocky common sticky, very plastic spherical
septarian of
CaCO3
4C1 352–373 10YR 8/1 (d), 10YR 7/1 (m); Sandy Massive – Extremely hard, – Abrupt
rare mottled 7,5YR 6/8 (d), loam extremely firm, Smooth
7,5YR 5/8 (m) moderately sticky,
moderately plastic
4C2 373–426 Mottled horizon 2,5YR 6/ Sandy Massive – Extremely hard, – Clear
6–10 YR 7/2 (d); 2,5YR 5/8 loam extremely firm, Smooth
− 10YR 6/2 (m) moderately sticky,
moderately plastic

(d) – dry, (m) – moist;

7
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Fig. 5. Pedological characteristics of the profile. A) slickenside in profile 2; B) slickenside, observed in the dry season; C, D) aggregate defined by slickenside planes;
E) CaCO3 rhizolite in profile 4; F) septarian carbonate nodule; G, H) sand injection in the slickenside planes.

abundant and define the structures (Fig. 5C, D). It presents abundant H). The 4C2 horizons are massive, sandy, and present carbonate con­
oriented clay films not associated with slickenside. The transition with centrations and mottling. Below the horizons, the deposits show sedi­
the lower profile is abrupt, truncating the profile and eliminating hori­ mentary structures without evidence of pedogenesis, however, a
zon A. transition occurs where the sedimentary structures were partially
Profile 4 (bottom) presents two Bss horizons and two C horizons, obliterated.
with colors similar to the horizons of profile 3, few manganese films and
less wax of illuvial origin, despite the abundant presence of slickenside 4.3. Granulometric and chemical data
(usually over 1 m in size). Mottling is less frequent than in the upper
profile. The structure is also parallelepiped, however with smaller blocks Granulometric results are presented in Table 3. Horizons B present
than the previous profile. Carbonate rhizolites are frequent (Fig. 5E) and values equal to or greater than 30% of clay (lower limit to classify the
extend to the base of the profile (horizon 4C2). The rhizolites are trun­ profile with a horizon Bss) (USDA, 1999; IUSS Working Group WRB,
cated at the top, indicating an erosive process that eliminated upper 2015). The clay values exceed 44% in the 4Bss2 horizon, classifying
horizons. The profile presents septarian carbonate glaebules (Fig. 5F). them as clay, clay loam, and eventually sand clay loam. Clay values are
The base of horizon 4Bss2 shows slickenside planes filled with sand from significantly lower in horizons A, 4C1 and 4C2. Silt values are high in all
the lower horizon, indicating that the wetting/drying cycles injected profiles, but they are significantly lower than those found by Srivastava
sand into the slickenside planes, starting from the horizon 4C2 (Fig. 5G, (2001) in Gangetic Plains in India in similar alluvial conditions. The

8
F.S.B. Ladeira et al. Catena 212 (2022) 106113

sand values are high, according to the characteristics already observed

g/dm3
for the Pantanal (Nascimento et al., 2013; Cardoso et al., 2016). Between
profiles 1 and 2 there is a high textural gradient, reaching 3.4, what

Hor. = horizon; CAS = gravel; AMG = very coarse sand; AG = coarse sand; AM = medium sand; AF = fine sand; AMF = very fine sand; SB = sum of bases; CTC = cation exchange capacity; V = base saturation
C

5
7
9
8
6
5
5
7
3
4
would characterize a Planosol if they were a unique profile.
X-ray fluorescence spectrometry data indicate widespread presence

24,21
11,78
36,49
66,58
71,56
82,47
91,99
93,67
88,89
91,07
of SiO2 (values above 68.81%; Table 4). The Al2O3, MnO, and CaO
V variations indicate significant pedogenetic processes. Al2O3 values are
significantly high in clayey horizons. The MnO presents low values in

17,91
15,43
11,37
13,36
16,54
19,98
25,29
4,75

11,2
the section, except in horizon 2C, where MnO values reach 0.18%,
CTC

9
coincident with the occurrence of black films on the peds. The values
indicate reducing conditions (Kämpf et al, 2009; Schulze et al., 1995)

13,64
18,38
23,69
and may correspond to a drainage impediment linked to the lower ho­
1,15
2,11
5,63
7,57
9,56

10,2
SB

8
rizon, generating a suspended water table, and indicating past drainage
conditions (Kämpf et al, 2009). The sorptive complex results indicate
0,25
0,11
0,13
0,17
0,26
0,34
0,38
0,49
that the characteristics are repeated (Table 4 and Fig. 6). Exchangeable
0,2
0,2
K

aluminum amounts are much larger in the profile 2 and decrease


significantly with depth. It directly reflects base saturation, low at the
Mg

0,3
0,4

1,8
2,5
3,5
4,5
5,2
2,3
2,8
profiles 1 and 2, and high in the profiles 3 and 4, with the profile 1 and 2
1

presenting dystrophic characteristics, and the others eutrophic. CaO


presents two relevant distinct conditions from the upper profile (profiles
13,5
0,6
1,6
4,5
5,6
6,8
9,8

5,5
7,2
cmol/Kg %
Ca

18

1 and 2), showing low values that significantly increase in the under­
lying Bss horizons, especially in the profile 4. Exchangeable calcium is
3,8
3,2
2,6
0,7
0,6
0,2

noticeably high in the lower profiles. The values reach 20 meq, being the
Al




main responsible for the base saturation of the profiles. However, CaO
values are relatively low for carbonate profiles, indicating that the cal­
pH (CaCl2)

cium in its ionic form suffered intense mobility, solubilizing, and then
precipitating, as the paleosols were humid and dried. In addition, the
3,8

4,2
4,2
4,6
5,6
5,8
5,7
5,8
Chemical Analysis

carbonate concentrations are associated with bioturbation, and not so


4

significantly with the soil matrix, indicating the association of calcium


values with vegetation growth. Thus, it indicates that the carbonate
pH (H20)

concentrations occurred when the profiles were on the surface, not


4,6
4,2
4,5
4,8
4,8
5,3
6,3
6,5
6,5
6,5

representing concentrations after burial. The carbonate concentration


associated with roots indicates dry conditions during soil evolution
(Nash and McLaren, 2003, Oliveira, 2013).
Clay

30,4
35,9
33,8
29,9
30,9
28,4
44,7
17,3
17,3
8,7

4.4. Mineralogical data


14,4

17,2
14,5

40,3
20,9
29,7
14,2
17,3
Silt

16

16

X-ray diffractograms are shown on Fig. 7. The 2Bt1 and 2Bt2 sub­
surface horizons present kaolinite. The main peak presented is 7.20 Å,
76,9
53,6
46,9
51,7
54,1
28,8
50,7
25,6
68,5
65,4

which after heating, collapses and disappears completely, confirming


AT

the presence of kaolinite. A secondary peak at 14.53 Å corresponds to


expansive clay minerals (montmorillonite). After glycolate and burning,
AMF

24,1
17,4
14,9
15,6
23,6
11,7
22,3
13,3
20,9

the peak presented a slight displacement of the basal spacing, due to the
28

dehydration of the expansive mineral layer containing water/hydroxyl.


However, the little accentuated change indicates a small contribution of
AF

30
20
18
17
20
11
19

32
25
9

the group of clay minerals in the sample, which are predominantly


kaolinite. Profiles 3 and 4 (3Btss1, 3Btss2, 4Bss1, and 4Bss2) show a
AM

3,4
9,1
5,5
7,7
3,1

8,2

predominance of clay minerals from the smectite group, especially


17
14
12

12

montmorillonite. In the profiles, the main peak is at 15.67 Å, corre­


sponding to minerals such as chlorite, montmorillonite, and vermiculite,
3,6
1,9
2,5
6,9

0,7
1,1
0,2
2,6
2,9
AG

identified after glycolate and burning. A secondary peak occurs at 7.23


Å, which spacing corresponds to the presence of kaolinite. In the gly­
AMG

colate sample, the main peak was moved to 18.15 Å, and corresponds to
Granulometry

1,9
0,4
0,3
8,8
0,2
0,2
0,5

1,4
1,8
0

the expansion of the basal spacing of 15.67 Å presented in the natural


fraction (approximately 2.5 Å expansion), confirming the presence of
Granulometry and Chemical Analysis.

10,8
CAS

montmorillonite. The expansion is significantly high, corroborating the


%

0
0
0

0
0
0
0
0
0

presence of extremely plastic clay minerals, and a great capacity for


absorbing cations and water. After burning, the peak collapsed to 15.67
Hor. Depth (cm)

Å, a position close to 10 Å, which corresponds to the dehydration of the


montmorillonite layer containing water/hydroxyl. The 7.23 Å peak, like
170–251
251–299
299–352
352–373
373–426
92–170

other secondary peaks, collapses and disappears completely, confirming


24–33
33–80
80–92
0–24

the presence of kaolinite (Fig. 7). The diffractograms show two distinct
situations for the occurrence of clay minerals, indicating different con­
ditions for pedogenetic formation. The occurrence of kaolinite repre­
Table 3

3Btss1
3Btss2
4Bss1
4Bss2

sents perfect conformation with hot and humid environments with


2Bt1
2Bt2
Hor.

4C1
4C2
2C
A

prolonged periods of rain, providing intense leaching, thus, the effective

9
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Table 4
Major elements concentration in Anhumas profile obtained from XRF measurements.
Horizons SiO2 TiO2 Al2O3 Fe2O3 MnO MgO CaO Na2O K2O P2O5

(%)

A 91,73 1,316 2,49 3,12 0,084 0,09 0,02 0,03 0,28 0,038
2Bt1 79,46 1,484 9,11 4,94 0,024 0,24 0,05 0,03 0,32 0,047
2Bt2 74,65 1,649 10,98 6,15 0,039 0,3 0,13 0,08 0,42 0,045
2C 71,36 1,5 10,73 10,3 0,439 0,36 0,17 0,09 0,49 0,071
3Btss1 80,16 1,5 6,9 5,9 0,05 0,38 0,24 0,12 0,74 0,056
3Btss2 77,58 1,46 7,5 6,66 0,08 0,5 0,3 0,17 0,94 0,077
4Bss1 75,26 1,769 8,48 7,28 0,149 0,65 0,39 0,27 1,18 0,062
4Bss2 68,81 1,91 11,06 8,5 0,183 0,92 0,58 0,32 1,3 0,056
4C1 87,26 1,332 4,4 3 0,146 0,34 0,23 0,19 0,85 0,028
4C2 83,42 1,613 4,75 5,17 0,179 0,41 0,31 0,21 0,95 0,039

Fig. 6. Results of FRX analyzes and routine chemical analysis.

Fig. 7. Diffractograms. A. Standard for Bt horizons comparing natural sample (Bt_n), glycolate (Bt_gl), heated to 300 ◦ C (Bt_300) and 550 ◦ C (Bt1_550); B. Standard
for Bss horizons comparing natural sample (Bss_n), glycolate (Bss_gl), heated to 300 ◦ C (Bss_300) and 550 ◦ C (Bss_550).

removal of cations and SiO2 (Resende et al., 1995; Alleoni et al., 2009, presence of silica, Ca, and Mg, bases that would commonly be leached in
Melo and Wypych, 2009). However, montmorillonite tends to form and other environments. In these areas, soils with vertic character are
remain stable in poorly drained or non-leaching environments, such as common (e.g. profiles 3 and 4 of Anhumas section), as well as the natural
those with long dry seasons and shallow groundwater areas (Kämpf and processes of expansion and contraction of clays that play an important
Curi, 2003, Kämpf et al., 2009). Montmorillonite indicates an abundant role in their genesis during dry and wet seasonal cycles.

10
F.S.B. Ladeira et al. Catena 212 (2022) 106113

4.5. OSL dating deposits were completely obliterated. The development of soil profiles 3
and 4 was in drought areas associated with semi-arid or arid climates.
Six OSL dates were obtained (Fig. 3) to establish the chronology of The paleosol profile 2 show very different characteristics that indi­
the pedo-sedimentary evolution of the Anhumas section, three in hori­ cate the change to a more humid climatic condition. Hydromorphic
zon R, and three in the succession of paleosols and soil. The sediments features at the base of profile 2 (2C horizons) are associated with the low
used for OSL dating present bright quartz sand grains with signal permeability of the underlying Vertisol. The occurrence of a perched
dominated by fast component (0.8 s corresponding to ~25% of light water table is indicated by the presence of a horizon abundant in Mn
emission), indicating quartz with high luminescence sensitivity (Sawa­ films at the base of the profile 2. Horizons Bt are common in the Pantanal
kuchi et al., 2011). All OSL signals grow with dose, and their dos­ wetland, specially associated with Planosols, cover 35.06% of its area
e–response curves are well described by fitting a saturating exponential (Couto et al., 2017), and is characterized by a high textural gradient
function. The calculated equivalent doses for the six samples range from between A and B.
44.2 ± 2.1 to 5.4 ± 0.3 Gy, with De distributions showing overdispersion Despite the gradient between the horizon A and the underlying 2Bt1,
from 8% to 23%, which is consistent with the reported range for well- high gradient may be associated with processes of ferrolysis, gleization,
bleached sediments, and absence of sediment mixing after deposition illuviation/eluviation, elutriation, bioturbation, granulometric/miner­
(Arnold and Roberts, 2009). Equivalent dose distributions with low to alogical differences of the source material, or simply coarse sedimentary
moderate overdispersion are reported for several Brazilian rivers (e.g., contribution at the top of the profile (Philips, 2004). In studied section,
Pupim et al., 2017; Oliveira et al., 2019), indicating low feldspar content the textural gradient does not imply a pedogenetic origin because there
and well-bleached quartz grains (Sawakuchi et al., 2016). The radiation is no genetic relation between these horizons because the horizon A
dose rates range from 0.77 to 0.27 Gy/ka. developed in a sand unit deposited in a younger depositional event.
The estimated OSL ages indicate Holocene to late Pleistocene de­ There is a truncation surface with a gap of Σ22 ka between the A and 2Bt
posits, ranging from 8.3 ± 0.7 to 106.7 ± 5.6 ka (Table 1). The dating horizons. Therefore, the abrupt transition surface between horizons A
carried out in the deposits of the lower sequence of Horizon R shows the and Bt is a lithological discontinuity.
oldest age (106.7 ± 5.63 ka.). In the upper deposits of Horizon R, two In the northern Pantanal, particle size differences were interpreted as
OSL dates were obtained, one at the base and one at the top, yielding non-pedogenetic and attributed to seasonal flood flows or associated
ages of 69.1 ± 4.9 ka and 72.2 ± 5.5 ka, respectively (Fig. 3). Consid­ with avulsions (Nascimento et al., 2013). These may be responsible for
ering the error margin for the ages, the two dates indicate the same deposits of different particle sizes, creating textural discontinuities.
deposition age for the upper sequence of horizon R. The deposition of Regarding profiles 1 and 2, we cannot state whether the textural dif­
both sequences of R features refers to the Upper Pleistocene, and the ference is associated or not with the dynamics of the Aquidauana River
record of the degrading Pleistocene paleochannel. Although all lith­ channel. Contrary to previous studies, the present work identifies an
ofacies belong to the paleochannel, due to the sedimentological differ­ Arenosol at the top, over the Bt horizon (probably from a truncated
ences and the dates that showed a time gap of about Σ20 ka, a non- Planosol), and not only does not present genetic affiliation between
conformity between the two feature sequences is interpreted. horizons A and Bt, but that it also has a significant gap between the
The baseline age obtained for the paleosols profile was 57.1 ± 4.0 ka, deposits that generated it. It does not present a pedogenesis that
on the 3Btss2 horizon. As the top of Horizon R was dated 72.2 ± 5.5 ka, inherited a lithological discontinuity as indicated in the literature and
profile 1 of the paleosol presented a maximum gap for its development identified by Nascimento et al. (2013) in the North Pantanal, or by
of Σ15 ka, and pedogenesis occurred in the Upper Pleistocene. The Esfandiarpour-Borujeni, et al. (2018) in Iran. On the contrary, it is a
second age in paleosols was performed at the base of horizon 2Bt2, aged pedogenesis that generated changes in very sandy materials (which are
29.8 ± 2.3 ka. Thus, profile 3 of the paleosol developed in the Upper identified as horizon A) over a previous pedogenesis that had generated
Pleistocene, in an interval of approximately 28 ka. The younger age a profile 2, characterized by Bt horizons, whose original horizon A has
obtained corresponds to horizon A, at the top of the profile 1. been eliminated by erosive processes. Thus, the subsurface horizons of
the Planosol indicate that they are paleosol horizons, like the profiles of
5. Discussion Vertisols below them.

5.1. Pedocomplex 5.2. Pedoenvironmental changes

The sedimentary section described allows us to define two distinct Two evolutionary sets in distinct environmental and climatic terms
stratigraphic intervals regarding soil/sediment relationships. The lower were recognized in the Anhumas paleosol profiles. One is associated
part essentially presents sedimentary deposits without evidence of with the Vertisols and the other with the Planosol/Arenosol. According
pedogenesis, indicating a period with no development or conservation to Buol et al. (2011), Vertisols develop associated with three basic
of soil profiles. The absence of soil profiles may be explained by envi­ conditions that may or may not be associated: (a) in an arid or semi-arid
ronmental conditions that did not allow formation of soils (such as climate; (b) associated with a low permeability layer or horizon below
climate and short depositional gaps), continuous fluvial accumulation, the vertic horizons, and (c) in low permeability settings produced by the
with no pause in in-channel sedimentation or by erosional processes that smectite clays that characterize the horizons. The second condition is
eliminated the soils eventually developed before the occurrence of a new absent, since the Vertisol positioned at the base (profile 4) is directly in
deposit (Marriott and Wright, 1993; Kraus and Aslan, 1999; Kraus, contact with sandy deposits with high porosity, indicating that the cli­
1999). Large temporal gaps occur if soils are partially or completely matic and mineralogical types are what justifies the presence of Verti­
eroded before the next depositional event. sols. The origin of expansive clay may be associated with processes of
The three paleosol profiles without any preserved sedimentary local weathering and/or inherited by deposition. According to Pal et al.
structures may be associated with events of 103 to 104 years, which (2009), Vertisols can occur in different climate types, as already indi­
resulted in long pauses in the deposition, as well as erosive processes cated by different authors (Coulombe et al. 1996; Coulombe et al. 2000,
responsible for the decapitation of superficial horizons associated with Nordt and Driese, 2010). However, Vertisols that develop in different
Btss and Bt, affecting the soil development (Kraus, 1999). The profiles 3 climates have different characteristics.
and 4 are Vertisols with significant clay values (30–40%). The erosional Morphologically, the cracks are deeper in drier climates and exceed
surface separating these two paleosol profiles indicates an erosive pro­ 1 m. Although the profile is truncated, the cracks reach up to 1 m in
cess that truncated the previous profile and generated juxtaposed pro­ depth (Fig. 5A, B) and are often filled with sand (Fig. 5G,H). Carbonate
files (e.g. Fedoroff et al. 2010). The sedimentary structures of these concentrations occur in the form of rhizoconcretions (Fig. 5E) and

11
F.S.B. Ladeira et al. Catena 212 (2022) 106113

carbonate glaebules (Fig. 5F), indicating environmental conditions with 72.2 ± 5.5 ka, marked by high variability of precipitation and sparse
significant water deficit (Pal et al., 2001). The white color, due to car­ vegetation cover in the source areas (Fig. 8). The high variability in
bonate concentrations, highlights the environmental conditions. Ac­ precipitation during the period has been associated with weakening and
cording to Pal et al. (2001) and Srivastava et al. (1994, 1998), white strengthening of the South American Summer Monsoon (SASM - Cruz
colors in alluvial deposits point to arid and semi-arid climates, mean­ et al., 2005; Novello et al., 2017), whereas the colder temperatures and
while the formation of Fe-Mn films occurs in slightly humid conditions, decrease of CO2 in the atmosphere suggest a sparse vegetation cover
generating red and black colors associated with carbonate features. (Burbridge et al., 2004). The environmental conditions indicate high
Profile 2 and 1, which overlaps the Vertisol (profile 3), presents low variability in river discharge and sediment yield at the catchment area in
saturation of bases, iron glaebules, manganese films, and kaolinite in the the surrounding plateaus, leading to a dynamic fluvial system in the
clay fraction. These characteristics, completely different from Vertisols, lowland areas (Hansford and Plink-Björklund, 2020). The reported
indicate a distinct pedoclimate, though without significant water deficit, fluvial deposits with absolute ages older than 50 ka are scarce in the
indicating notable presence of water in the system. They also indicate literature of South American rivers. Deposition in dynamic fluvial sys­
partially impeded drainage. tems (braided channels and/or distributary systems) with high season­
The profiles 2 and 1 are more complex, in terms of paleoenvir­ ality in river discharge has been interpreted for chronocorrelatable
onmental evolution, maybe as two overlapped soil profiles that generate deposits in the Paraná River, Araguaia River, Pantanal rivers, and
another profile, which due to organization of the horizons, is classified Amazon River and its tributaries (Stevaux, 2000; Valente and Latru­
as a single profile. Horizons 2Bt and 2C show deposition in 30 ka, and besse, 2012; Assine et al., 2014; Rossetti et al., 2015; Pupim et al., 2017;
the A horizon presents deposition age of 8 ka. Independently of whether Oliveira et al., 2019; Pupim et al., 2019).
there are two deposits and two pedogenetic events or two deposits and The Anhumas profile paleosols show a sequence of sedimentary de­
one pedogenesis, these indicate, in pedogenetic terms, relatively more posits in which the OSL ages corroborate relative drier events and
humid conditions. In summary, the presence of two overlapping Verti­ marked environmental transitions during the Upper Pleistocene. The
sols profiles indicates that their respective developments were associ­ paleosols originated in slightly humid conditions (profiles 3 and 4) and
ated with climatic and/or hydrological variations. However, between with higher water availability (profiles 1 and 2), particularly in the 2Bt
the Vertisol and Planosol profiles, at the top a significant change in horizons (Fig. 8). Our interpretation matches the Knox model (Knox,
environmental conditions occurs, especially climatic, featuring differ­ 1972), where most geomorphic processes of erosion and deposition
ences of allogenic character. occur during periods of environmental transition. Therefore, the
development of soil is associated with periods of relative landscape
stability under an increase in regional humidity, in conditions equally or
5.3. Paleosols and regional climate variations wetter than today.
The sequence of well-developed Btss and Bss horizons (Vertisols with
The paleoenvironmental changes recognized in the Anhumas section carbonate septarian glaebules and carbonate rhizolites) highlights a
(Aquidauna River) in southern Pantanal may be related to regional long period of arid or semi-arid conditions between ≈70 and ≈30 ka.
changes in precipitation (Cruz et al., 2005; Novello et al., 2017) and The drier period during the marine isotope stage 3 (MIS3) was also
vegetation (Whitney et al., 2011; Fornace et al., 2016). The lower part of interpreted from geomorphological and sedimentological records of
the studied Anhumas section consists of medium to coarse-grained sands fluvial systems in northern Pantanal (Assine et al., 2014; Pupim et al.,
with trough cross-bedding strata without paleosol horizons preserved, 2017). Moreover, oxygen isotope records of speleothems from Paraiso
suggesting a dynamic fluvial system with high competence to transport Cave (eastern Amazon; Wang et al., 2017) and Botuverá Cave (southern
sediment. The deposition of sediments occurred from 106.7 ± 5.6 to

Fig. 8. OSL dating and climate proxy. A. OSL dating of sediments in the Anhumas section of the Aquidauna River with error interval; B. Isotopic data from the
Jaraguá Cave (Novello et al., 2017); C. Isotopic data from the Botuverá Cave (Cruz et al., 2005); D. Isotopic data from the Paraíso Cave, east of the Amazon State
(Wang et al., 2017); Marine isotope stages (MIS) are indicated at the top of the figure (Lisiecki and Raymo, 2005). LGM – Last Glacial Maximum. Top of curves - drier;
lower - wetter.

12
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Brazil, Cruz et al., 2005) suggest a weakening of the SASM towards dating facilities; FSBL (#307951/2018-9), FNP (#302411/2018-6) and
central and northern Brazil during the MIS3, limiting the wetter con­ MLA (#304925/2017-9) are research fellows of Brazilian National
ditions to the south. The well-developed Bt horizons suggest relatively Council for Scientific and Technological Development - CNPq/Brazil.“;
wetter conditions and a stable landscape from 29.8 ± 2.3 ka to 8.3 ± 0.7 the authors are grateful to Isabel Leli and José Cândido Stevaux for
ka. Thus, an increase of the tropical summer precipitation can be fieldwork assistance; we thank an anonymous reviewer and Dra María
interpreted, due to the intense SASM associated with an increase in the Sol Raigemborn for their review and suggestions that improved the
atmospheric CO2 concentration, during the last deglaciation (18–11.7 manuscript.
ka) and early Holocene (McGlue et al., 2012; Fornace et al., 2016).
Despite the limitation of our data to precisely constrain the humid References
period, oxygen isotopes from Jaraguá Cave (Bonito, MS) suggest that it
extended through the Last Glacial Maximum (LGM − 26 to 18 ka) Aitken, M.J., 1985. Thermoluminescence dating. Academic Press, p. 359.
Alekseeva, T., Kabanov, P., Alekseev, A., Kalinin, P., Alekseeva, V., 2016. Characteristics
(Novello et al., 2017). The OSL age of 8.3 ± 0.7 ka points to sediment of early Earth’s Critical Zone based on Middle-Late Devonian paleosol properties
reworking and deposition from early to mid-Holocene. It corroborates (Voronezh High, Russia). Clays Clay Miner. 64 (5), 677–694.
relatively drier conditions reported from lakes (McGlue et al., 2012; Alleoni, L.R.F.; Camargo, O.A.; Casagrande, J.C., Soares, M.R., 2009. Química dos solos
altamente intemperizados. In: Melo, V.F., Alleoni, L.R.F (Eds.). Química e
McGlue et al., 2017) and speleothems records (Novello et al., 2017) in mineralogia do solo. Viçosa, MG, Sociedade Brasileira de Ciência do Solo, v.2. p.
the Pantanal region. The development of horizon A may be related to 381-447.
modern environmental conditions that have been established since 4–2 Anthony, J.W., Bideaux, R.A., Bladh, K.W., Nichols, M.C., 2017. Handbook of
Mineralogy. http://www.handbookofmineralogy.org/ (accessed in july 2017).
ka with an overall increase of rainfall, due to the current monsoon Arnold, L.J., Roberts, R.G., 2009. Stochastic modeling of multi-grain equivalent dose
system (McGlue et al., 2012; McGlue et al., 2017). (De) distributions: implications for OSL dating of sediment mixtures. Quat.
Geochronol. 4, 204e230.
Assine, M.L., Soares, P.C., 2004. Quaternary of the Pantanal, west-central Brazil. Quat.
6. Conclusions
Int. 114, 23–34.
Assine, M.L., 2005. River avulsions on the Taquari megafan, Pantanal Wetland, Brazil.
The Anhumas outcrop in the Aquidauana River terrace records sig­ Geomorphology 70 (3-4), 357–371.
nificant diversity of paleoenvironmental information. By identifying the Assine, M.L., Corradini, F.A.L., Pupim, F.N., Mcglue, M.M., 2014. Channel arrangements
and depositional styles in the São Lourenço fluvial megafan, Brazilian Pantanal
processes associated with sedimentation and pedogenesis, remarkable wetland. Sediment. Geol. 301, 172–184.
environmental changes from the late Pleistocene to Holocene could be Assine, M.L., 2015. Brazilian Pantanal: A Large Pristine Tropical Wetland. In: Vieira, B.
discerned. The succession of events was interpreted based on the dis­ C., Salgado, A.A.R., Santos, L.J.C. (Eds.) Landscapes and Landforms of Brazil.
Springer Netherlands, (World Geomorphological Landscapes), 135–146.
continuities observed in the field (vertical succession of facies, horizons Assine, M.L., Merino, E., Pupim, F.N., Macedo, H.A., Santos, M.G.M., 2015a. The
and pedological features, physical discontinuities) that defined the Quaternary alluvial systems tract of the Pantanal Basin, Brazil. Brazil. J. Geol. 45 (3),
sample collection. The OSL dates proved the gaps defined in the field. 475–489.
Assine, M.L., Merino, E., Pupim, F.N., Warren, L., Guerreiro, R., Mcglue, M., 2015b.
The sedimentary and pedogenetic processes recorded in the Anhu­ Geology and Geomorphology of the Pantanal Basin. In: Bergier, I., Assine, M.L.
mas sequence show that the SASM plays a key role in the landscape (Eds.), Dynamics of the Pantanal Wetland in South America. Springer International
evolution in southern Pantanal. Data shows that most of the aggrada­ Publishing, (The Handbook of Environmental Chemistry, volume 37), p. 23–50.
Assine, M.L., Macedo, H., Stevaux, J., Bergier, I., Padovani, C., Silva, A., 2016. Avulsive
tional periods occurred over drier and transitional conditions, whereas Rivers in the Hydrology of the Pantanal Wetland. In: Bergier, I., Assine, M.L. (Eds.)
soil development and preservation were related to wetter and more Dynamics of the Pantanal Wetland in South America. Springer International
stable conditions. The development of Vertisols with an abundance of Publishing, (The Handbook of Environmental Chemistry, volume 37), p. 83–110.
Bateman, M.D., Frederick, C.D., Jaiswal, M.K., Singhvi, A.K., 2003. Investigations into
carbonate glaebules occurred between Σ≈70 and ≈30 ka. It presents
the potential effects of pedoturbation on luminescence dating. Quat. Sci. Rev. 22
well-marked, arid or semi-arid climatic conditions for the region during (10–13), 1169–1176. https://doi.org/10.1016/S0277-3791(03)00019-2.
MIS 4 and MIS 3 with increased aridity in the transition from MIS 4 to 3, Brasil, 1982a. Projeto RadamBrasil – Folha SE.21 – Corumbá e parte da Folha SE.20, Rio
and from MIS 3 to 2. de Janeiro. vol. 27, p. 448.
Brasil, 1982b. Projeto RadamBrasil – Folha SF.21 – Campo Grande, Rio de Janeiro. vol.
Pedogenetic processes that form the Bt horizon suggest relatively 28, p. 412.
wetter conditions during the MIS 2 and early Holocene. Sedimentary Burbridge, R.E., Mayle, F.E., Killeen, T.J., 2004. Fifty-thousand-year vegetation and
reworking during the middle Holocene points to drier conditions, which climate history of Noel Kempff Mercado National Park, Bolivian Amazon. Quat. Res.
61 (2), 215–230.
were followed by rainfall increase during the late Holocene. Buol, S.W., Southard, R.J., Graham, R.C., Mcdaniel, P.A., 2011. Soil Genesis and
The described profiles are commonly identified in the Pantanal Classification, sixth ed. John Wiley & Sons Inc, p. 543.
pedological mappings. However, many of them can be exhumed pale­ Camargo, O.A., Moniz, A.C., Jorge, J.A., Valadares, J.M.A.S., 2009. Métodos de análises
química, mineralógica e física de solos do Instituto Agronômico de Campinas.
osols as the case here documented. This helps explain the complexity Campinas, Instituto Agronômico de Campinas, p. 77 (Boletim Técnico, 106).
and pedological diversity that characterizes the Pantanal. The soils that Cardoso, E.L., Santos, S.A., Urbanetz, C., Carvalho Filho, A.d., Naime, U.J., Silva, M.L.N.,
currently occur on the surface, usually associated with different envi­ Curi, N., 2016. Relação entre solos e unidades da paisagem no ecossistema Pantanal.
Pesquisa Agropecuária Brasileira 51 (9), 1231–1240.
ronmental conditions, can coexist on the same geomorphic surface. Couto, E.G., Oliveira, V.A., Beirigo, R.M., Oliveira Junior, J.C., Nascimento, A.F., Vidal-
Finally, we consider the Pantanal a model for paleopedology in­ Torrado, P., 2017. Solos do Pantanal Mato-Grossense. In. Curi, N., Ker, J.C., Novais,
vestigations of fluvial distributive deposits. R.F., Vidal-Torrado, P., Schaefer, C.E.G.R. (Eds.) Pedologia – Solos dos Biomas
Brasileiros. SBCS. pp. 303–352.
Coulombe, C., Wilding, L., Dixon, J., 1996. Overview of Vertisols: characteristics and
Declaration of Competing Interest impacts on society. In: Sparks, D. (Ed.), Advances in Agronomy, vol. 57. Academic
Press, San Diego, pp. 289–375.
Coulombe, C., Wilding, L., Dixon, J., 2000. Vertisols. In: Sumner, M.E. (Ed.), Handbook
The authors declare that they have no known competing financial
of Soil Science. CRC Press, pp. E269–E286.
interests or personal relationships that could have appeared to influence Cruz, F.W., Burns, S.J., Karmann, I., Sharp, W.D., Vuille, M., Cardoso, A.O., Ferrari, J.A.,
the work reported in this paper. Silva Dias, P.L., Viana, O., 2005. Insolation-driven changes in atmospheric
circulation over the past 116,000 years in subtropical Brazil. Nature 434 (7029),
63–66.
Acknowledgments Esfandiarpour-Borujeni, I., Mosleh, Z., Farpoor, M.H., 2018. Comparing the ability of Soil
Taxonomy (2014) and WRB (2015) to distinguish lithologic discontinuity and an
The authors thank the FAPESP (grant #2014/06889-2, São Paulo abrupt textural change in major soils of Iran. Catena 165, 63–71.
Facincani, E.M., Souza, E.P., Bacani, V.M., Ladeira, F.S.B., 2009. Calcretes: indicação de
Research Foundation-Brazil) and the CNPq (grant # 432985/2018-2 paleoambientes no leque do Aquidauana na borda sudeste da bacia do Pantanal
National Council for Scientific and Technological Development-Brazil) Matogrossense. Anais 2◦ Simpósio de Geotecnologias no Pantanal, Corumbá, 7–11
and CAPES for financial support; the Luminescence and Gamma Spec­ novembro 2009, Embrapa Informática Agropecuária/INPE, p. 140–149.
trometry Laboratory (LEGaL) at the University of São Paulo for OSL

13
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Facincani, E.M., Assine, M.L., 2010. Geomorfologia fluvial do rio Aquidauana, borda Mohindra, R., Parkash, B., Prasad, J., 1992. Historical gebmorphology and pedology of
sudeste do Pantanal Mato-Grossense. In: Martins Junior, C., Oliveira Neto, A.F. the Gandak megafan, middle Gangetic plains, India. Earth Surf. Proc. Land. 17,
(Eds.) Revelando Aquidauana. Campo Grande, Editora da UFMS, (Serie Fronteiras n◦ 643–662.
3), 267–284. Mohindra, R., 1995. Holocene soil chronoassociation in part of the Middle Gangetic
Fedoroff, N., Courty, M.A., Guo, Z., 2010. Palaeosoils and relict soils. In: Stoops, G., Plain: morphological and micromorphological characteristics. Terra Nova 7 (3),
Marcelino, V., Mees, F. (Eds.), Interpretation of Micromorphological Features of Soils 305–314.
and Regoliths. Elsevier, pp. 623–662. Morrison, R.B., 1967 Principles of quaternary stratigraphy. In: Morrison, R.B., Wright, H.
Feng, Z.-D., Wang, H.B., 2006. Geographic variations in particle size distribution of the E. (Eds.). Quaternary soils. Congress of the International Association of Quaternary
last interglacial pedocomplex S1 across the Chinese Loess Plateau: Their Research, 7, Reno, 1967. Proceedings. Reno, v.9. p. 1–69.
chronological and pedogenic implications. Catena 65 (3), 315–328. Murray, A.S., Wintle, A.G., 2000. Luminescence dating of quartz using an improved
Fernandes, F.A., Fernandes, A.H.B.M., Soares, M.T.S, Pellegrin, L.A., Lima, I.B.T., 2007. single-aliquot regenerative-dose protocol. Radiat. Meas. 32 (1), 57–73.
Atualização do Mapa de Solos da Planície Pantaneira para o Sistema Brasileiro de Murray, A.S., Wintle, A.G., 2003. The single aliquot regenerative dose protocol: potential
Classificação de Solos. Comunicado Técnico 61. Embrapa. Corumbá. p. 6. for improvements in reliability. Radiat. Meas. 37 (4-5), 377–381.
Fontana, A., Mozzi, P., Bondesan, A., 2008. Alluvial megafans in the Venetian-Friulian Nascimento, A.F.d., Furquim, S.A.C., Couto, E.G., Beirigo, R.M., Oliveira Júnior, J.C.d.,
Plain (north-eastern Italy): evidence of sedimentary and erosive phases during Late Camargo, P.B.d., Vidal-Torrado, P., 2013. Genesis of textural contrasts in subsurface
Pleistocene and Holocene. Quat. Int. 189 (1), 71–90. soil horizons in the Northern Pantanal-Brazil. R. Bras. Ci. Solo 37 (5), 1113–1127.
Fornace, K.L., Whitney, B.S., Galy, V., Hughen, K.A., Mayle, F.E., 2016. Late Quaternary Nash, D.J., McLaren, S.J., 2003. Kalahari valley calcretes: their nature, origins, and
environmental change in the interior South American tropics: New insight from leaf environmental significance. Quat. Int. 111 (1), 3–22.
wax stable isotopes. Earth Planet. Sci. Lett. 438, 75–85. Nordt, L.C., Driese, S.G., 2010. A modern soil characterization approach to
Franzini, M., Leoni, L., Saitta, M., 1976. Enhancement effects in X-Ray fluorescence reconstructing physical and chemical properties of paleo-vertisols. Am. J. Sci. 310
analysis of rocks. X-Ray Spectrom. 5 (4), 208–211. (1), 37–64.
Galbraith, R.F., Roberts, R.G., Laslett, G.M., Yoshida, H., Olley, J.M., 1999. Optical Novello, V.F., Vuille, M., Cruz, F.W., Strikis, N.M., Paula, M.S., Edwards, R.L., Cheng, H.,
dating of single and multiple grains of quartz from Jinmium rock shelter, northern Karmann, I., Jaqueto, P.F., Trindade, R.I.F., Hartmann, G.A., Moquet, J.S., 2016.
Australia: Part I, experimental design and statistical models. Archaeometry 41 (2), Centenial-scale solar forcing of the South American Monsoon System recorded in
339–364. stalagmites. Sci. Rep. 6, 24762.
Guérin, G., Mercier, N., Adamiec, G., 2011. Dose-rate conversion factors: update. Anc. TL Novello, V.F., Cruz, F.W., Vuille, M., Strikis, N.M., Edwards, R.L., Cheng, H., Emerick, S.,
29, 5–8. Paula, M.S., Li, X., Barreto, E.S., Karmann, I., Santos, R.V., 2017. A high-resolution
Hansford, M.R., Plink-Björklund, P., 2020. River discharge variability as the link between history of the South American Monsoon from Last Glacial Maximal to the Holocene.
climate and fluvial fan formation. Geology 48 (10), 952–956. https://doi.org/ Sci. Rep. 7, 44267.
10.1130/G47471.1. Oliveira, E.C., 2013. Rochas Carbonáticas Continentais (Quaternário) do Pantanal
Hartley, A.J., Weissmann, G.S., Bhattacharayya, P.R.O.M.A., Nichols, G.J., Scuderi, L.A., Matogrossense e Adjacências. São Paulo, Tese (Doutorado em Geociências). Instituto
Davidson, S.K., Driese, S., 2013. Soil development on modern distributive fluvial de Geociências, Universidade de São Paulo, p. 133.
systems: preliminary observations with implications for interpretation of paleosols in Oliveira, S.C., Pupim, F.N., Stevaux, J.C., Assine, M.L., 2019. Luminescence chronology
the rock record. New Front. Paleopedol. Terrestrial Paleoclimatol.: SEPM, Special of terrace development in the Upper Paraná River, southeast Brazil. Front. Earth Sci.
Publication 104, 149–158. 7, 200.
INSTITUTO NACIONAL DE METEOROLOGIA (INMET), 2017. http://www.inmet.gov.br/ Oliveira Junior, J.C., Beirigo, R.M., Chiapini, M., Nascimento, J.F., Couto, E.G., Vidal-
portal/index.php?r=estacoes/estacoesAutomaticas. Accessed in 03 de sept. 2017. Torrado, P., 2017. Origin of mounds in the Pantanal wetlands: an integrated
IUSS Working Group WRB. 2015. World Reference Base for Soil Resources 2014, update approach between geomorphology, pedogenesis, ecology and soil micromorphology.
2015. International soil classification system for naming soils and creating legends Plos one 12 (7), e50179197.
for soil maps. World Soil Resources Reports No. 106. FAO, Rome. Oliveira Junior, J.C., Furquim, S.A.C., Nascimento, A.F., Beirigo, R.M., Barbiero, L.,
Jacobs, P.M., Konen, M.E., Curry, B.B., 2009. Pedogenesis of a catena of the Farmdale- Valles, V., Couto, E.G., Vidal-Torrado, P., 2019. Salt-affected soils on elevated
Sangamon Geosol complex in the north central United States. Palaeogeogr. landforms of an alluvial megafan, northern Pantanal, Brazil. Catena 172, 819–830.
Palaeoclimatol. Palaeoecol. 282 (1-4), 119–132. Pal, D.K., Balpande, S.S., Srivastava, P., 2001. Polygenetic Vertisols of the Purna Valley
Kämpf, N., Curi, N., 2003. Argilominerais em solos brasileiros. In: Curi, N., Marques, J.J., of Central India. Catena 43 (3), 231–249.
Guilherme, L.R.G., Lima, J.M., Lopes, A.S., Alvarez V. (Eds.). Tópicos em ciência do Pal, D.K., Bhattacharyya, T., Chandran, P., Ray, S.K., Satyavathi, P.L.A., Durge, S.L.,
solo. Viçosa: Sociedade Brasileira de Ciência do Solo, p.1-54. Raja, P., Maurya, U.K., 2009. Vertisols (cracking clay soils) in a climosequence of
Kämpf, N., Curi, N., Marques, J.J., 2009. Óxidos de alumínio, silício, manganês e titânio Peninsular India: Evidence for Holocene climate changes. Quat. Int. 209 (1-2), 6–21.
In. Melo, V.F., Alleoni, L.R.F. (Eds.) Química e Mineralogia do Solo Parte I – Phillips, J.D., 2004. Geogenesis, pedogenesis, and multiple causality in the formation of
conceitos básicos. SBCS pp. 573–610. texture-contrast soils. Catena. 58 (3), 275–295.
Karlstrom, E.T., Oviatt, C.G., Ransom, M.D., 2008. Paleoenvironmental interpretation of Por, F.D., 1995. The Pantanal of Mato Grosso (Brazil) - World’s Largest Wetlands. Kluwer
multiple soil-loess sequence at Milford Reservoir, northeastern Kansas. Catena 72 Academic Publishers, Dordrecht, The Netherlands, p. 12.
(1), 113–128. Pott, A., Oliveira, A.K.M., Damasceno-Junior, G.A., Silva, J.S.V., 2011. Plant diversity of
Knox, J.C., 1972. Valley alluviation in southwestern Wisconsin. Ann. Assoc. Am. Geogr. the Pantanal wetland. Braz. J. Biol. 71 (1 suppl 1), 265–273.
62 (3), 401–410. Prance, G.T., Schaller, G.B., 1982. Preliminary study of some vegetation types of the
Köppen, W., Geiger, R., 1928. Klimate der Erde. Gotha: Verlag Justus Perthes. Wall-map Pantanal, Mato Grosso, Brazil. Brittonia 34, 228–251.
150cmx200cm. Prescott, J.R., Hutton, J.T., 1994. Cosmic ray contributions to dose rates for
Kraus, M.J., 1999. Paleosols in clastic sedimentary rocks: their geologic applications. luminescence and ESR dating: large depths and long-term time variations. Radiat.
Earth Sci. Rev. 47 (1-2), 41–70. Meas. 23 (2-3), 497–500.
Kraus, M.J., Aslan, A., 1999. Paleosol sequences in floodplain environments: a Pupim, F.d.N., Assine, M.L., Sawakuchi, A.O., 2017. Late Quaternary Cuiabá megafan,
hierarchical approach. In: Thiry, M. (Ed.), Palaeoweathering, Palaeosurfaces and Brazilian Pantanal: Channel patterns and paleoenvironmental changes. Quat. Int.
Related Continental Deposits. Int. Assoc. Sedimentol., Spec. Publ. 27, 303–321. 438, 108–125.
Latrubesse, E.M., Stevaux, J.C., Cremon, E.H., May, J.-H., Tatumi, S.H., Hurtado, M.A., Pupim, F.N., Sawakuchi, A.O., Almeida, R.P., Ribas, C.C., Kern, A.K., Hartmann, G.A.,
Bezada, M., Argollo, J.B., 2012. Late Quaternary megafans, fans and fluvio-aeolian Chiessi, C.M., Tamura, L.N., Mineli, T.D., Savian, J.F., Grohmann, C.H., Bertassoli, D.
interactions in the Bolivian Chaco, Tropical South America. Palaeogeogr. J., Stern, A.G., Cruz, F.W., Cracraft, J., 2019. Chronology of Terra Firme formation in
Palaeoclimatol. Palaeoecol. 356-357, 75–88. Amazonian lowlands reveals a dynamic Quaternary landscape. Quat. Sci. Rev. 210,
Leier, A.L., DeCelles, P.G., Pelletier, J.D., 2005. Mountains, monsoons, and megafans. 154–163.
Geology 33 (4), 289. https://doi.org/10.1130/G21228.110.1130/2005049. Resende, M., Curi, N., de Rezende, S.B., Corrêa, G.F., 1995. Pedologia: base para
Mack, G.H., James, W.C., 1994. Paleoclimate and global distribution of paleosols. J Geol distinção de ambientes, first ed. NEPUT- Viçosa, Viçosa, MG, Brazil, p. 304.
102, 360–366. Retallack, G.J., 1990. Soils of the Past: an introdution to paleopedology. In: Retallack, G.
Marriott, S.B., Wright, V.P., 1993. Paleosols as indicators of geomorphic stability in two J. (Ed.), Soils of the Past. Springer Netherlands, Dordrecht, pp. 3–8. https://doi.org/
Old Red Sandstone alluvial suites, South Wales. J. Geol. Soc. London 150, 10.1007/978-94-011-7902-7_1.
1109–1120. Rossetti, D.F., Cohen, M.C.L., Tatumi, S.H., Sawakuchi, A.O., Cremon, É.H., Mittani, J.C.
McGlue, M.M., Silva, A., Zani, H., Corradini, F.A., Parolin, M., Abel, E.J., Cohen, A.S., R., Bertani, T.C., Munita, C.J.A.S., Tudela, D.R.G., Yee, M., Moya, G., 2015. Mid-late
Assine, M.L., Ellis, G.S., Trees, M.A., Kuerten, S., Gradella, F.D.S., Rasbold, G.G., Pleistocene OSL chronology in western Amazonia and implications for the
2012. Lacustrine records of Holocene flood pulse dynamics in the Upper Paraguay transcontinental Amazon pathway. Sed. Geol. 330, 1–15.
River watershed (Pantanal wetlands, Brazil). Quat. Res. 78 (2), 285–294. Santos, R.D., Lemos, R.C., Santos, H.G., Ker, J.C., Anjos, L.H.C., Shimizu, S.H., 2013.
McGlue, M.M., Guerreiro, R.L., Bergier, I., Silva, A., Pupim, F.N., Oberc, V., Assine, M.L., Manual de descrição e coleta de solo no campo, 6ª ed. Sociedade Brasileira de
2017. Holocene stratigraphic evolution of saline lakes in Nhecolândia, southern Ciência do Solo, Viçosa, p. 100.
Pantanal wetlands (Brazil). Quat. Res. 88 (3), 472–490. Sawakuchi, A.O., Blair, M.W., DeWitt, R., Faleiros, F.M., Hyppolito, T., Guedes, C.C.F.,
Melo, V. De F., Wypych, F., 2009. Caulinita e Haloisita. In: Melo, V. F., Alleoni, L.R.F. 2011. Thermal history versus sedimentary history: OSL sensitivity of quartz grains
(Eds.). Química e Mineralogia do Solo. Parte II: Aplicações. Viçosa, Sociedade extracted from rocks and sediments. Quat. Geochronol. 6 (2), 261–272.
Brasileira de Ciência do Solo, p. 427–504. Sawakuchi, A.O., Mendes, V.R., Pupim, F.d.N., Mineli, T.D., Ribeiro, L.M.A.L., Zular, A.,
Metcalfe, S.E., Whitney, B.S., Fitzpatrick, K.A., Mayle, F.E., Loader, N.J., Street- Guedes, C.C.F., Giannini, P.C.F., Nogueira, L., Sallun Filho, W., Assine, M.L., 2016.
perrott, F.A., Mann, D.G., 2014. Hydrology and climatology at Laguna La Gaiba, Optically stimulated luminescence and isothermal thermoluminescence dating of
lowland Bolivia: complex responses to climatic forcings over the last 25,000 years. high sensitivity and well bleached quartz from Brazilian sediments: from Late
J. Quat. Sci. 29 (3), 289–300. Holocene to beyond the Quaternary? Braz. J. Geol. 46 (suppl 1), 209–226.

14
F.S.B. Ladeira et al. Catena 212 (2022) 106113

Schoenerberger, P.J., Wysocki, D.A., Bernham, E.C., Soil Survey Staff, 2012. Field Book Stremme, H.E., 1998. Correlation of Quaternary pedostratigraphy from western to
for Describing and Sampling Soils. Version 3.0. NRCS/NSSC. Lincoln, USA. eastern Europe. Catena 34 (1-2), 105–112.
Schulze, D.G., Sutton, S.R., Bajt, S., 1995. Determining manganese oxidation state in soils Tabor, N.J., Myers, T.S., 2015. Paleosols as Indicators of Paleoenvironment and
using x-ray adsorption near-edge structure (XANES) spectroscopy. Soil Sci. Soc. Am. Paleoclimate. Annu. Rev. Earth Planet. Sci. 43 (1), 333–361.
J. 59, 1540–1548. Trendell, A.M., Atchley, S.C., Nordt, L.C., 2013. Facies analysis of a probable large-
Sousa, J.B., Souza, C.A., 2013. Caracterização morfológica e mineralógica de solos em fluvial-fan depositional system: the upper Triassic Chinle Formation at Petrified
ambientes de cordilheira e campo de inundação no Pantanal de Poconé, Mato Forest National Park, Arizona, USA. J. Sediment. Res. 83 (10), 873–895.
Grosso. Boletim de Geografia 31, 53–66. USDA, 1999. Soil Taxonomy - A Basic System of Soil Classification for Making and
Srivastava, P., Parkash, B., Sehgal, J.L., Kumar, S., 1994. Role of neotectonics and Interpreting Soil Surveys. NRCS, p. 871.
climate in development of the Holocene geomorphology and soils of the Gangetic Valente, C.R., Latrubesse, E.M., 2012. Fluvial archive of peculiar avulsive fluvial patterns
Plains between Ramganga and Rapti rivers. Sediment. Geol. 94, 119–151. in the largest Quaternary intracratonic basin of tropical South America: The Bananal
Srivastava, P., Parkash, B., Pal, D.K., 1998. Clay minerals in soils as evidence for Basin, Central-Brazil. Palaeogeogr. Palaeoclimatol. Palaeoecol. 356-357, 62–74.
Holocene climatic changes in the central Indo-Gangetic Plains, north-central India. Wang, X., Edwards, R.L., Auler, A.S., Cheng, H., Kong, X., Wang, Y., Cruz, F.W., Dorale, J.
Quat. Res. 50, 230–239. A., Chiang, H.-W., 2017. Hydroclimate changes across the Amazon lowlands over the
Srivastava, P., 2001. Paleoclimatic implications of pedogenic carbonates in Holocene past 45,000 years. Nature 541 (7636), 204–207.
soils of the Gangetic Plains, India. Palaeogeogr. Palaeoclimatol. Palaeoecol. 172 (3- Whitney, B.S., Mayle, F.E., Punyasena, S.W., Fitzpatrick, K.A., Burn, M.J., Guillen, R.,
4), 207–222. Chavez, E., Mann, D., Pennington, R.T., Metcalfe, S.E., 2011. A 45 kyr palaeoclimate
Srivastava, P., Parkash, B., 2002. Polygenetic soils of the north-central part of the record from the lowland interior of tropical South America. Palaeogeogr. Palaeoclim.
Gangetic Plains: a micromorphological approach. Catena 46 (4), 243–259. Palaeoecol. 307 (1-4), 177–192.
Srivastava, P., Singh, A.K., Parkash, B., Singh, A.K., Rajak, M.K., 2007. Paleoclimatic Zeilhofer, P., Schessl, M., 2000. Relationship between vegetation and environmental
implications of micromorphic features of Quaternary paleosols of NW Himalayas and conditions in the Northern Pantanal of Mato Grosso. Brazil. J. Biogeogr. 27 (1),
polygenetic soils of the Gangetic Plains—a comparative study. Catena 70 (2), 159–168.
169–184. Zhao, J., 2003. Paleoenvironmental significance of a paleosol complex in Chinese Loess.
Stevaux, J.C., 2000. Climatic Events during the Pleistocene and Holocene in the upper Soil Sci. 168 (1), 63–72.
Paraná River: correlation with NE Argentina and South-Central Brazil. Quat. Int. 77, Zhou, Y.u., Retallack, G.J., Huang, C., 2015. Early Eocene paleosol developed from basalt
87–94. in southeastern Australia: implications for paleoclimate. Arab J Geosci. 8 (3),
1281–1290.

15

You might also like