You are on page 1of 11

Geoderma 154 (2010) 518–528

Contents lists available at ScienceDirect

Geoderma
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / g e o d e r m a

Soil mineral genesis and distribution in a saline lake landscape of the Pantanal
Wetland, Brazil
S.A.C. Furquim a,⁎, R.C. Graham b, L. Barbiero c, J.P. Queiroz Neto a, P. Vidal-Torrado d
a
Departamento de Geografia, Universidade de São Paulo, São Paulo, Brazil
b
Soil & Water Sciences Program, Department of Environmental Sciences, University of California, Riverside, USA
c
Laboratoire de Mécanisms et Transferts en Géologie, UMR 5563 CNRS-IRD-UPS-OMP, Toulouse, France
d
Departamento de Ciência do Solo, Universidade de São Paulo, Piracicaba, Brazil

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

Article history: The origin of the saline lakes in the Pantanal wetland has been classically attributed to processes occurring in
Received 20 August 2008 past periods. However, recent studies have suggested that saline water is currently forming from evaporative
Received in revised form 6 January 2009 concentration of fresh water, which is provided annually by seasonal floods. Major elements (Ca, Mg, K) and
Accepted 20 March 2009
alkalinity appear to be geochemically controlled during the concentration of waters and may be involved in
Available online 23 April 2009
the formation of carbonates and clay minerals around the saline lakes. The mineralogy of soils associated
with a representative saline lake was investigated using XRD, TEM-EDS, and ICP-MS in order to identify the
Keywords:
Nhecolândia
composition and genesis of the secondary minerals suspected to be involved in the control of major
Salinity elements. The results showed that Ca, Mg, and K effectively undergo oversaturation and precipitation as the
Mineral neoformation waters become more saline. These elements are incorporated in the authigenically formed carbonates,
Carbonates smectites, and micas surrounding the saline lake. The control of Ca occurs by precipitation of calcite and
Mg-smectite dolomite in nodules while Mg and K are mainly involved in the neoformation of Mg-smectites (stevensitic
Fe-mica and saponitic minerals) and, probably, iron-enriched micas (ferric–illite) in surface and subsurface horizons.
Therefore, our study confirms that the salinity of Pantanal, historically attributed to inheritance from former
regimes, has a contribution of current processes.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction The Nhecolândia sub-region (27,000 km2), located in the central-


southern portion of the Pantanal, is distinguished by the existence of
The Pantanal wetland, located in the central area of South America about 7000 fresh water lakes (“baías”) and 1500 alkaline–saline lakes
(16°–20°S and 58°–50°W), is the largest complex of continental (“salinas”), generally associated with temporary channels (“vazantes”)
wetlands in the world, with an area of about 200,000 km2 (Por, 1995) and sand hills (“cordilheiras”) (Fig. 1) (Cunha, 1943; Brum and Sousa,
(Fig. 1A). Characterized by low altitudes (b200 m) and extremely low 1985; Mourão et al., 1988; Fernandes, 2007). The fresh water lakes are
gradients (0.02 to 0.03°), the wetland is a still active alluvial plain, temporary or permanent water bodies with different forms and
partially covered by seasonal flooding mainly in the summer dimensions and up to 2 m deep. Coalescence of fresh water lakes during
(November to March) (Silva, 1986; Scott, 1991). The Quaternary sedi- the flooding periods forms the temporary channels, which are several
ments underlying the plain are mainly sandy alluvial deposits of the kilometers long and up to 30 m wide. The alkaline–saline lakes are round
Paraguai River and tributaries. Most of the flooding water and the depressions with permanent salty water, generally 500 to 1000 m in
deposited sediments are carried from surrounding highlands (200 to diameter and 2 to 3 m in depth. They are located inside the sand hills,
900 m), which are formed by Precambrian crystalline and Mesozoic which are narrow (200–300 m wide) and elongated stripes of dryland, 2
sedimentary rocks (Del'Arco et al., 1982; Alvarenga et al., 1984; Godoi to 3 m higher than other landscape features and covered by savanna
Filho, 1986; Por, 1995). In spite of the recurrent presence of water, the vegetation.
wetland presents an annual hydrological deficit of 300 mm, resulting Different hypotheses have been addressed to explain the origin of
from a mean annual rainfall of 1100 mm and a mean annual these Nhecolândia landforms and the subject is still a matter of
evapotranspiration around 1400 mm (Alvarenga et al., 1984; Por, debate. A set of explanations invokes eolian processes occurring in the
1995). last arid phase (Late Pleistocene) as being the responsible for the
formation of sand hills and lakes. In this view, sand hills are seen as
paleo-longitudinal dunes (Klammer, 1982) and fresh water and/or
⁎ Corresponding author. Tel./fax: +55 11 30913794. saline lakes are wind deflation depressions (Almeida and Lima, 1956;
E-mail address: sfurquim@usp.br (S.A.C. Furquim). Tricart, 1982; Soares et al., 2003; Assine and Soares, 2004). Saline lakes

0016-7061/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.geoderma.2009.03.014
S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528 519

were specifically defined by Tricart (1982) as old sebkhas, where the during water concentration should provide insight into the role of
eolian deflation was facilitated due to aggregation of fine particles by current processes in the origin of salt waters in Nhecolândia.
salts accumulated on the soil surface. Another hypothesis attributes A detailed mineralogical investigation of the soils surrounding a
the origin of sand hills and lakes to the shifting of river courses, which representative alkaline–saline lake (Salina do Meio, Fig. 1B) was carried
occurred during and after the more humid climate prevailing during out by our research team, the results of which are beginning to be
deglaciation (Early Holocene). In this context, lakes would be inun- disseminated (e.g. Furquim et al., 2008). In spite of the extensive number
dated depressions located inside cutoff meander banks or isolated by of saline lakes in the Nhecolândia, our studies represent a pioneering
embankments while sand hills would develop through transformation effort in understanding the silt and clay mineralogy in these landscapes.
of shortcut meanders banks into elongated sand ribbons. The distinc- In this paper, we further elucidate the linkages between soil mineralogy,
tion between fresh water and saline lakes would depend on the geomorphology, and geochemistry of these lakes by providing a com-
recurrent input of fresh water from seasonal floodings (Wilhelmy, prehensive view of the composition and genesis of the fine-grained
1958; Ab'Saber, 1988). Eiten (1983) provided a different interpreta- minerals in the Salina do Meio landscape.
tion, attributing lake genesis to relict karstic or pseudo-karstic sys-
tems. Although these studies referred to different processes, they 2. Study area
all agree with a past origin of the saline lakes and other typical
Nhecolândia landforms. Soils were sampled along a toposequence (T1) extending from
However, recent geochemical studies involving surface and the border of the local sand hill to the surface level of Salina do Meio
subsurface waters of Nhecolândia suggested that saline water has (Figs. 1B and 2B). The surface water level of the saline lake varies
currently arisen from evaporative concentration of subsurface fresh throughout the year according to the watertable dynamics (Fig. 2B):
water (Barbiero et al., 2002, 2007). The water concentration seems to from November to March, the level is high and the water is close to the
be triggered by the presence of a subsurface impervious green horizon border of the sand hill; from April to October, the level goes down,
surrounding the saline lakes, which prevents the groundwater exposing most of the lake shore. Thus, the field work was done at the
outflow through the sand hill during the dry season, thereby retaining peak of the dry season, when the soils were at maximum exposure.
the water so it is concentrated by evaporation. The authors show that The morphological and analytical properties of the soil cover along
fresh and saline waters belong to the same chemical family and T1 (Fig. 2B) are summarized below and more fully discussed by
suggest that most of the chemical compositional changes in the Furquim et al. (2008). The soil consists of a pale brown (10YR 6/3),
system are related to precipitation of some elements (mainly Ca, Mg, single grain, sand surface material (horizon 1), which is laterally
K) as solutions become more saline. The chemical elements controlled replaced toward the lake by a dark gray (10YR 4/1), mostly prismatic,
in this way are suspected to be involved in present-day authigenic loamy sand, surface material, rich in decomposed algae (horizon 2).
formation of carbonates (calcite or Mg-calcite) and silicates (steven- Beneath this, there is an organic-enriched, grayish brown (10YR 4/2),
site, sepiolite, Mg-montmorillonite, or amorphous) in the soils massive horizon with a sand texture (horizon 3). This, in turn, overlies a
surrounding the saline lakes. The identification of the minerals light brownish gray (2.5Y 6/2), single grain horizon, with a sand texture
supposed to be responsible for the control of chemical elements and abundant vertical organic-rich volumes (horizon 4). The deepest

Fig. 1. A — Location of Pantanal wetland and Nhecolândia sub-region; B — Aerial picture showing the main landscape features of Nhecolândia, the studied toposequence (T1), and the
water survey transect.
520 S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528

Fig. 2. A — Location of piezometers (Pz) and watertable samplers (G) in the water survey transect (Barbiero et al., 2007); B — Soil horizons and pH values along toposequence 1 (T1).

horizons consist of a gray (2.5Y 5/1), massive, loamy sand material, important to highlight that the green and indurated horizon typically
with common presence of millimeter- to centimeter-diameter associated with saline lakes is the feature pointed out by Barbiero et al.
nodules (horizon 5), and a greenish (5Y 5/2), massive, sandy loam (2007) as being responsible for the isolation of groundwater around
material, which has rare to common presence of nodules and is saline lakes during the dry season, allowing the evaporative
indurated by amorphous silica (horizon 6). The amounts of clay in concentration of waters.
the soil range from 3 to 20% and are mainly concentrated in horizons The geochemistry study of surface and subsurface waters of a
2 (8.5%) and 6 (10 to 18%). Soil pH varies from 5.4 to 11 along the transect involving T1, the adjacent sand hill, and a neighboring fresh
toposequence, but 70% of the measurements are higher than 10. Gray water lake (Figs. 1B and 2A) was carried out by Barbiero et al. (2007).
and greenish horizons (5 and 6) are generally the most alkaline, with The results showed that fresh waters (piezometers Pz5 to Pz15 and
pH values up to 11 (Fig. 2B). Electrical conductivity (EC) values samplers G7 to G15, Fig. 2A) were of Na–K–HCO3 + CO3 type, whereas
display a wide range (0.3 to 43 dS m− 1), but are higher toward the saline waters (Pz2, Pz3, G0 to G6, and lake, Fig. 2A) were of Na–HCO3 +
saline lake. Na is the main cation on exchange sites in all horizons, CO3 type. Similar to the regional water conditions measured by
ranging from 3 to 190 cmolc/kg. Barbiero et al. (2002), the geochemical control of Ca, Mg, and K in the
Preliminary soil morphology surveying in different places of more saline waters of the transect suggested the involvement of
Nhecolândia showed that soil organization of T1 is representative of these elements in the formation of minerals around Salina do Meio
the saline lakes. The sand surface horizon (1) and the subsurface gray (Fig. 3; A, B, and C). Equilibrium diagrams show that saline waters
(5) and green (6) horizons were identified around all the surveyed of the transect are at equilibrium or slightly oversaturated with
saline lakes (Sakamoto, 1997; Barbiero et al., 2000; Fernandes, 2000; respect to Mg-silicates (saponite, stevensite, and sepiolite) (Fig. 3; D
Silva and Sakamoto, 2003; Silva et al., 2004) while the dark surface and E), attesting that those minerals could originate from these
horizon (2) and the light sand subsurface horizon with vertical waters. The chemical characteristics of water samples collected
organic volumes (4) were found in most of the studied saline lakes specifically along T1 (Pz2, G0, and G1) are listed in Table 1. These
(Sakamoto, 1997; Silva and Sakamoto, 2003; Silva et al., 2004). It is results include alkaline pH, high EC (mainly G0 and lake), high K
S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528 521

Fig. 3. A, B, and C— Concentration diagrams showing the evolution of Ca, Mg, and K during the concentration of waters in the transect involving Salina do Meio. Na amounts represent
the concentration factor. The solid line denotes the simulation of evaporation with precipitation of Mg-calcite and a Mg-silicate (sepiolite). D and E — Diagrams showing equilibrium
conditions in solution for different minerals (Barbiero et al., 2007).

content (mainly G0 and lake), and low contents of Al. Silica method (Nelson, 1982) and total C was determined by dry combustion
concentrations are lower than 100 mg/L, but increase toward the (Nelson and Sommers, 2001). Organic C was calculated by the
saline lake. difference between total C and inorganic C, the last derived from
calcium carbonate equivalent values.
3. Mineralogical methods Fine silt (2–20 μm), coarse clay (0.2–2 μm), and fine clay (b0.2 μm)
fractions of the 20 collected soil samples were separated by
Twenty whole soil samples and 17 nodules were collected along T1 centrifugation and sedimentation following destruction of organic
from horizons of 5 pedons (P1 to P5) exposed by trenches (Fig. 2B). matter with NaOCl (pH 9.5) (Anderson, 1963). These fractionated
Soil samples and nodules were crushed with a mortar and pestle samples were used in the following analyses.
and passed through a 100-mesh sieve. These powdered materials XRD was carried out for the 3 fractions of all the 20 samples. Silt
were randomly packed in an aluminum sample holder and submitted and clays underwent 5 treatments: Mg saturation (52% relative
to X-ray diffraction (XRD) for identification of carbonate minerals and humidity), ethylene glycol (EG) solvation of Mg-saturated clays, air
measurement of d060 values of phyllosilicates. Calcium carbonate dry K saturation, and heating of the K-saturated samples at 350 °C and
equivalent of the ground samples was determined by a manometric 550 °C (Jackson, 1979). Glass slides of oriented specimens were

Table 1
Characteristics of waters from Salina do Meio and surrounding watertable (cations with 2 repetitions).

Pz2 G1 G0 Salina do Meio


pH – 7.66 8.66 9.25 9.64
EC dS/m 6.19 8.76 19.18 12.74
Eh – − 25.2 − 217.0 − 364.5 − 125.0
Temp. °C 31.1 31.2 34.2 40.2
HCO3 + CO3 mEq/L 29.73 26.49 194.08 67.62
Cl− mmol/L 21.58 15.78 76.17 66.64
Si(OH)4 3.2098 3.5155 36.9463 38.6839 51.4686 52.3721 97.6426 89.5579
Al3+ 0.0007 0.0007 0.0196 0.0594 0.0039 0.0040 0.0514 0.0272
Mg2+ mg/L 1.0813 1.1843 0.0356 0.0372 0.2202 0.2241 2.2365 2.0513
Ca2+ 6.2313 6.8249 1.4463 1.4720 4.2313 4.3056 8.5382 7.8313
Fe 0.0826 0.4466 0.0319 0.0439 0.0120 0.0342 0.0553 0.0880
Na+ – – 982.2820 1028.479 5827.8413 5930.1401 – –
K+ 34.6911 37.9959 230.1342 240.9576 1094.8175 1114.0353 945.9906 867.66

EC: electrical conductivity; Temp.: temperature; Pz.: piezometer; G: watertable samplers.


522 S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528

prepared by smearing a paste of clay or dropping a water–silt slurry on patterns of these carbonate-enriched nodules show 3 carbonate
the slide (Theisen and Harward, 1962). All the XRD analyses were minerals: calcite (CaCO3), dolomite (CaMg(CO3)2) and/or nahcolite
performed in a Siemens D-500 diffractometer (CuK radiation with (NaHCO3) (Fig. 4).
graphite crystal monochromator), employing a step size of a 0.02° 2θ Carbonate precipitation has been observed in soils that are fre-
and a count time of 1.0 s per step. The XRD peak intensities from the quently subjected to evaporative concentration of solutions (Fehren-
EG- and Mg-treated clays were used to calculate the semi-quantitative bacher et al., 1963; Mahjoory, 1979; Kohut and Dudas, 1995). Alkaline
amount (%) of each mineral in the assemblages (Biscaye, 1965), and to earth carbonates, such as calcite and dolomite, are generally the first
estimate the relative content of iron in the mica structure, using the minerals to precipitate in the concentration process. Calcite generally
formula I(001) / I(002), where I(001) and I(002) represent, respec- contains less than 5%mol of Mg(CO3) because it usually precipitates
tively, the intensity of the d001 and d002 mica peaks (Brown and from more diluted solutions, which typically have low Mg/Ca ratios
Brindley, 1980; Huggett et al., 2001). Results higher than 2 indicate a (b1). However, this ratio increases as calcite formation continues,
relatively large Fe-content in the octahedral sheet (Brown and allowing the subsequent precipitation of Mg-carbonates such as mag-
Brindley, 1980; Deconinck et al., 1988). nesite and dolomite (Eugster and Hardie, 1978; Boettinger and
Quantitative analyses of major elements (Si, Al, Fe, Mg, Ca, Na, K) Richardson, 2001).
and rare earth elements-REE (La, Ce, Pr, Nd, Sm, Eu, Gd, Tb, Dy, Ho, Er, As already stated, the concentration diagrams displayed in Fig. 3
Tm, Yb, Lu, and Y) were carried out on the fine-clay fraction of 18 (A and B) show a trend of decreasing concentration of Ca and Mg in
samples by inductively coupled plasma-mass spectrometry (ICP-MS), more saline waters of the transect involving Salina do Meio (Barbiero
using a Perkin-Elmer equipment, model Elan 6000. et al., 2007), which is in agreement with the regional results obtained
The fine-clay fraction of 5 selected samples was analyzed using a FEI- by Barbiero et al. (2002). The geochemical control of these elements in
CM300 transmission electron microscope (TEM) linked with a Phoenix more saline waters suggests their involvement in the precipitation of
energy-dispersive X-ray microanalyser (EDS). Six crystals of kaolinite Ca- and Mg-carbonates in the soils associated with Salina do Meio,
(P1—horizon 6, P5—horizon 6), 6 crystals of smectite (P1—horizon 6; inasmuch as the solubility limits of these minerals are the first to be
P4—horizon 2; P5—horizon 2), and 1 crystal of mica (P3—horizon 6) exceeded in the concentration process. The presence of calcite in the
were examined. For analysis, a dilute suspension containing the fine surface horizons within the area of seasonal lake-level variation (P3
clay was dropped on a standard Cu grid with carbon film. The to P5) and the occurrence of calcite- and dolomite-enriched nodules
chemical composition of individual crystals was determined by EDS confirm the prediction made by water analyses and support the
for each sample. The collected data were used to calculate the neoformation of these carbonate minerals around the saline lake.
chemical formulas of the individual crystals.
5. Clay minerals
4. Carbonates
X-ray diffraction patterns indicate an overwhelming dominance of
The studied soil samples generally have very low contents of calcium quartz and kaolinite in the coarse clay, and the presence of smectite,
carbonate equivalent, with values ranging between 0.2 and 3.2%. The mica, kaolinite, and quartz in the fine clay. The composition and
greatest values (1.2 and 3.2%) occur in the surface horizon 2 of P4 and P5. genesis of the 3 detected clay minerals are presented below.
Consistent with this, XRD analysis revealed the presence of calcite
(CaCO3) in the fine silt of the surface horizons of P3, P4 and P5 (Fig. 4). 5.1. Kaolinite
Most of the analyzed nodules have much higher calcium carbonate
equivalents than the whole soil samples. Values vary from 0.2 to 44%, Kaolinite occurs in all the analyzed horizons of the toposequence
but 65% of the nodules have CaCO3eq amounts greater than 6%. XRD in both coarse and fine-clay fractions. Semi-quantitative calculations

Fig. 4. XDR patterns showing the presence of carbonates in fine silt samples (air dry K-saturated peaks) and nodules (b 100-mesh powder). Identified minerals: Q — quartz; M —
microcline; C — calcite; D — dolomite; N — nahcolite.
S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528
Fig. 5. Percentages of clay minerals in coarse and/or fine-clay assemblages along T1.

523
524 S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528

estimate 3 to 77% of kaolinite in coarse clay assemblages, with no Maracaju”, where soils are acid, leached Psamments (Orioli et al.,
trend of vertical or lateral distribution along the toposequence (Fig. 5). 1982), ideal for kaolinite formation. Iron-rich kaolinites are very
Fine-clay assemblages contain 4 to 53% of kaolinite, with a general commonly formed in tropical soils (Herbillon, 1976; Newman and
decrease toward the saline lake (P5) (Fig. 5). Brown, 1987) and this may be the case in “Planalto de Maracaju”.
TEM/EDS results were used to calculate structural formulas of 6 Experiments of kaolinite dissolution under room temperature and
individual kaolinite crystals. One representative formula is listed in Table 2 pressure show very high concentrations of dissolved Si(OH)4 and Al in
(number 1). In all the analyzed specimens, tetrahedral sheet is formed by strongly acid (≤ 3) and strongly alkaline conditions (≥10) (Huertas
Si, with no Al substitution, and octahedral sheet is formed by Al and Fe3+. et al., 1999). Because most of the soil and water pH values are strongly
Aluminum ranges from 1.43 to 1.81 atoms per unit formula on the basis of alkaline in the Salina do Meio toposequence, kaolinite is probably
14 negative charges (apuf/14) and Fe3+ varies from 0.08 to 0.25 apuf/14, undergoing dissolution and releasing Si, Al, and Fe into solution. The
but the values are ≥0.13 in 5 of the 6 crystals. Iron substitution of pseudo-hexagonal crystals of kaolinite with subangular vertices and
octahedral Al in kaolinites is commonly less than 0.04 apuf/14 (Newman rough surfaces are probably evidences of this weathering process.
and Brown, 1987), which qualifies the studied specimens as iron-rich
kaolinites. The octahedral sum is slightly lower than 2.00 apuf/14 in all 5.2. Smectites
formulas probably because of the absence of Na in the results, which is
due to an interference of Cu from the Cu grid of the TEM samples holder Smectites in the Salina do Meio toposequence have been thor-
(Hover et al., 1999). TEM images show that the analyzed kaolinite crystals oughly studied by Furquim et al. (2008) and the most relevant aspects
are hexagonal or pseudo-hexagonal shaped, with angular or subangular are presented here. Based on semi-quantitative estimates for the fine-
vertices and smooth to rough surfaces (Fig. 6; A and B). clay fraction, substantial percentages of smectite are present in two
Some evidence suggests that kaolinite is allochthonous in the zones of the toposequence (Fig. 5): an upper zone, formed by all
study area. First, stability diagrams available in the literature showed horizons of P1 and P2; and a lower zone, formed by surface horizons 1
that kaolinites are stable under low values of pH–pK and pSi(OH)4 at and 2 of P3, P4, and P5. The results indicate the occurrence of different
room pressure and temperature, typical conditions of acid, well- smectite minerals and mechanisms of genesis in these zones.
drained, and leached soils (Tardy et al., 1973; Bohn et al., 1985; Velde,
1995). Soils associated with Salina do Meio are strongly alkaline, 5.2.1. Smectites from the upper zone
poorly drained, and have high base saturations, the opposite In this zone, XRD 00l patterns show a random interstratification of
conditions for the ideal formation of this mineral. Second, neoformed smectite with mica and vermiculite (Fig. 7) while 060 XRD peaks (0.149–
clays tend to concentrate in the fine-clay size and kaolinites of the 0.150 nm) indicate a dioctahedral domain in the fine-clay samples. Si, Al,
study area are abundant in both coarse and fine-clay fractions. and Fe3+ are the most abundant major elements of these samples, as
Therefore, the studied kaolinites were probably transported by the indicated by ICP-MS analysis, with amounts varying from 49 to 63% for
tributaries of the Paraguai River and deposited in the alluvial plain Si, 15 to 23% for Al (Fig. 8), and 5 to 10% for Fe3+. Results from TEM-EDS
along with the dominant sand particles. The source area of analysis of a single flake-shaped smectite crystal from P1—horizon 6
Nhecolândia sediments is a tropical highland known as “Planalto de (Table 2, number 2) agree with the ICP-MS analysis in that they yielded a
structural formula with high tetrahedral Al (0.8 apuf/22) and dominance
of Al (1.7 apuf/22) and Fe3+ (0.5 apuf/22) in the octahedral sheet. These
Table 2
characteristics qualify the analyzed specimen as a ferribeidellite (Güven,
Oxides and structural formulas of clay minerals and amorphous specimen.
1988; Reid-Soukup and Ulery, 2002). This crystal probably has some
1 2 3 4 5 6 hydroxyl-Al, -Fe in the interlayer position, raising the octahedral sum
Upper zonea Lower zonea above 2.00 apuf/22 (Weaver and Pollard, 1973).
P5-hor 6 P1-hor 6 P4-hor.2 P5-hor. 2 P3-hor.6 P3-hor. 6 The following characteristics suggest that these dioctahedral
SiO2 53.8 52.3 62.0 60.7 54.3 73.8 smectites originated from transformation of micas: 1) interstratifica-
Al2O3 39.7 35.0 7.0 10.7 27.7 10.2 tion with mica, as showed by 00l XRD patterns (Fig. 7); 2) high Al
Fe2O3 5.1 10.0 6.3 4.2 8.2 9.1
tetrahedral substitution in the structural formula (Table 2, number 2),
MgO 0.0 1.3 17.7 17.1 2.9 2.8
CaO 0.0 0.8 1.2 1.3 0.0 0.0 as expected for smectites transformed from micas (Borchardt, 1989);
MnO 0.0 0.0 0.0 0.0 0.0 0.9 3) dominance of iron-rich micas in the studied toposequence, as
K2O 1.5 0.6 1.6 0.9 4.8 4.1 discussed in the section on micas (Section 5.3); and 4) interstratifica-
Cl− 0.0 0.0 1.3 2.4 0.0 0.0 tion with vermiculite (Fig. 7), which is commonly an intermediate
Si4+ 2.04 3.18 3.96 3.87 3.43 –
phase in the transformation of mica toward smectite in soils
Al3+ 0.00 0.82 0.04 0.13 0.57 – (Crawford et al., 1983; Badraoui et al., 1987; Ransom et al., 1988).
∑Tetrahedral 2.04 4.00 4.00 4.00 4.00 –
5.2.2. Smectites from the lower zone
Al3+ 1.78 1.69 0.49 0.67 1.49 –
In this zone, all 060 XRD patterns display a strong peak of
Fe3+ 0.15 0.46 0.30 0.20 0.39 –
Mg2+ 0.00 0.04 1.69 1.63 0.12 – 0.152 nm, revealing a trioctahedral domain in the fine-clay samples.
∑Octahedral 1.93 2.19 2.48 2.50 2.00 – ICP-MS results show dominance of Si (44–51%) and Mg (13–15%)
among major elements (Fig. 8) and indicate much lower amounts
K+ 0.07 0.05 0.13 0.07 0.39 – of REE than samples from the upper zone (Fig. 9). Major element
Ca2+ 0.00 0.05 0.08 0.09 0.00 –
contents are corroborated by the calculated structural formulas of
Mg2+ 0.00 0.08 0.00 0.00 0.15 –
5 flake-shaped smectite crystals. Two of these formulas are listed
Anions O5 O10 O10 O10 O10 – in Table 2 (numbers 3 and 4) and a representative picture of one
OH− 4.00 2.00 1.86 1.74 2.00 – crystal is displayed in Fig. 6 (C). In all formulas, tetrahedral Al is low
Cl− 0.00 0.00 0.14 0.26 0.00 –
(0–0.14 apuf/22) and Mg is the main cation in the octahedral sheet
Tetrahedral charge 0.16 − 0.82 − 0.04 − 0.13 − 0.57 –
Octahedral charge − 0.24 0.51 − 0.25 − 0.12 − 0.12 – (1.49 to 2.30 apuf/22). Substitution of Al for Si in the tetrahedral
Total charge − 0.08 − 0.31 − 0.29 − 0.25 − 0.69 – sheet is zero or near zero in two crystals (e.g. Table 2, number 3) and
Interlayer charge 0.07 0.31 0.29 0.25 0.69 – between 0.12 and 0.14 apuf/22 in the other 3 crystals (e.g. Table 2,
Hor.: horizon. number 4). These are typical characteristics of the Mg-rich end-
a
Zones of smectite concentration. members stevensite, which has near-zero Al substitution, and
S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528 525

Fig. 6. TEM pictures of: A— Kaolinite crystal with angular vertices and smooth surfaces; B— Kaolinite crystals with subangular vertices and smooth to rough surfaces; C: Smectite
crystal with flake shape; D: Mica crystal with lath shape; E: Amorphous silica-rich phase with mineral crystallization.

saponite, which presents slightly higher Al substitution (Weaver and potentially as Mg-smectites; 3) equilibrium diagrams reveal that
Pollard, 1973). However, because of the presence of certain proper- saline waters of the transect at Salina do Meio are at equilibrium or
ties not found in stevensite and saponite, such as very low octahedral oversaturated with respect to Mg-silicates (saponite, stevensite, and
sum and Cl substitution for OH, the studied specimens are designated sepiolite) (Fig. 3; D and E), attesting that these minerals can
as stevensitic and saponitic minerals. The considerably lower precipitate from the saline waters; and 4) horizons having Mg-
octahedral sum is mainly attributed to the presence of a dioctahedral smectites occur only on the surface and exactly within the area of
component in the crystals and, to a lesser degree, to the absence of Na seasonal lake-level variation (Fig. 5), indicating that these minerals
in the analysis due to an interference of Cu grids of TEM samples are directly associated with the lake dynamics.
holder (Hover et al., 1999).
Some lines of evidences show that stevensitic and saponitic
minerals originate by direct precipitation from the saline lake: 1) 5.3. Micas
minerals formed by chemical precipitation from waters have much
lower REE values than inherited or transformed minerals (Torrez Ruiz XRD patterns from oriented clays show that mica is present in all
et al., 1994; Jamoussi et al., 2003), and this is true of the Mg-smectites horizons analyzed for the toposequence. The higher percentages (67
of the studied area (Fig. 9); 2) the concentration diagram displayed in to 84%) clearly occur in the gray (5) and green (6) horizons of the
Fig. 3(B) shows that Mg precipitates during the concentration process, pedons nearest to the saline lake (P3, P4, and P5) (Fig. 5).
526 S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528

Fig. 9. Sum of REEs in upper and lower zone samples.

cations of the octahedral sheet. This amount of octahedral Fe3+,


intermediate between illite and glauconite minerals, qualifies the
studied specimen as a ferric–illite (Fe-illite) (Porrenga, 1968; Baker,
1997; Hugget et al., 2001). Abundant crystallites associated with Si-
rich amorphous materials were also detected in the analyzed sample
(Fig. 6E; Table 2, number 6).
Although research about this mica is still underway, the following
water chemistry (Barbiero et al., 2007) and mineralogy data suggest
an authigenic formation: 1) geochemical control of K in the more
saline waters of the transect involving Salina do Meio (Fig. 3C) shows
that the involvement of this element in the precipitation of micas is
possible; 2) synthesis experiments under room conditions carried out
Fig. 7. Example of smectite peaks behavior under different treatments. Mixed-layer mica
smectite is identified in the upper zone (P1, horizons 1, 3, 5, and 6; P2, horizons 1, 5, and by Harder (1974) verified neoformation of mixed-layer illite–smectite
6) by the partial coalescence of the d001 mica and smectite peaks in K (22 °C) and/or Mg from waters that had similar chemical characteristics to those of the
patterns and a clear separation of these peaks in the EG patterns. Mixed-layer study area, such as alkaline pH, Si(OH)4 concentration lower than
vermiculite–smectite is recognized in this zone by total or partial collapse of smectite
100 mg/L, and high K concentration; 3) the mechanism of neoforma-
peaks to 10.0 Å in K-saturated samples (22 °C) (P1 and P2, horizons 1 and 6).
tion observed by Harder (1974) involved initial precipitation of
amorphous hydroxides from solutions, followed by mineral crystal-
XRD 001 diffraction peaks suggest an interstratification of mica lization after aging. The presence of amorphous materials associated
with smectite layers in all samples. This interpretation is based on with mica crystals in the sample investigated by TEM-EDS might be
very broad and asymmetrical mica peaks from Mg- and/or K-saturated evidence of the same mechanism of mica neoformation in the study
clays and a decrease of width and asymmetry of mica peaks area. Further investigation will probably give more consistent
accompanied by a better definition of smectite peaks upon EG- information about genesis of mica around Salina do Meio.
solvation of the Mg-clays (Fig. 10). Results of the I(001) / I(002) index
shows values higher than 2, suggesting that all fine-clay samples 6. Conclusions
contain Fe-rich micas.
TEM-EDS analysis of one lath-shaped mica crystal from P3— The current evaporative concentration of waters is promoting
horizon 6 (Fig. 6D; Table 2, number 5) confirms the trend pointed out precipitation of Ca, Mg, and K from the more saline solutions and their
by the I(001) / I(002) index. The calculated structural formula of this consequent participation in authigenic mineral phases (carbonates,
crystal defines Al (1.5 apuf/22) and Fe3+ (0.4 apuf/22) as the main smectites, and probably micas) around Salina do Meio. The

Fig. 8. Relationship between Si%–Al% and Si%–Mg% in the upper and lower zone samples.
S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528 527

historically attributed to inheritance from former regimes, has a con-


tribution of current processes.

Acknowledgments

The authors wish to thank Capes for the doctoral fellowship to


S.A.C.F. and for the financial support to the research through a Capes-
Cofecub cooperation (412/03); Dr. Arnaldo Sakamoto, Dr. Rosely
Pacheco Dias Ferreira, Dr. Sônia Furian, Dr. Nádia Nascimento, MS Ary
Tavares, undergraduate and graduate students of the University of
Mato Grosso do Sul (UFMS), and employees of the Nhumirim Farm
for their assistance in the field; Paul Sternberg, Marcos Pinheiro, and
Luís Silva for general help with the laboratory work; and Dr.
Krassimir Bozhilov for TEM assistance.

References
Ab'Saber, A.N., 1988. O Pantanal Mato-Grossense e a Teoria dos Refúgios. Revista Brasileira
de Geografia 50, 9–57.
Almeida, F.F.M., Lima, M.A., 1956. Excursion guidebook. 18th International Geography
Congress, vol. 1. Rio de Janeiro.
Alvarenga, S.M., Brasil, A.E., Pinheiro, R., Kux, H.J.H., 1984. Estudo Geomorfológico Aplicado
à Bacia do Alto Rio Paraguai e Pantanais Matogrossenses. Boletim Técnico Projeto
Radambrasil 1, 1–187.
Anderson, J.U., 1963. An improved pretreatment for mineralogical analysis of samples
containing organic matter. Clays and Clay Minerals 10, 380–388.
Assine, M.L., Soares, P.C., 2004. Quaternary of the Pantanal, west-central Brazil. Quaternary
International 114, 23–34.
Badraoui, M., Bloom, P.R., Rust, R.H., 1987. Occurrence of high-charge beidellite in a
vertic Haplaquoll of northwestern Minnesota. Soil Science Society of America
Journal 51, 813–818.
Baker, J.C., 1997. Green ferric clay in non-marine sandstones of the Rewan Group,
southern Bowen Basin, Eastern Australia. Clay Minerals 32, 499–506.
Barbiero, L., Furquim, S.A.C., Vallès, V., Furian, S., Sakamoto, A., Rezende Filho, A., Fort,
M., 2007. Natural arsenic in groundwater and alkaline lakes at the Upper Paraguay
Basin, Pantanal, Brazil. In: Bhattacharya, P., Mukherjee, A.B., Bundschuh, J.,
Zevenhoven, R., Loeppert, R.H. (Eds.), Arsenic in Soil and Groundwater Environ-
ment: Biogeochemical Interactions, Health Effects and Remediation. Trace Metals
and Other Contaminants in the Environment series, vol. 9. Elsevier, pp. 101–126.
Barbiero, L., Furian, S., Queiroz Neto, J.P., Ciornei, G., Sakamoto, A.Y., Capellari, B., Fernandes,
E., Vallès, V., 2002. Geochemistry of water and ground water in the Nhecolândia,
Pantanal of Mato Grosso, Brazil: variability and associated processes. Wetlands 22,
528–540.
Barbiero, L., Queiroz Neto, J.P., Sakamoto, A.Y., 2000. Características Geoquímicas dos Solos
Relacionadas à Organização Pedológica e à Circulação da Água (Fazenda Nhumirim:
Embrapa CPAP, Nhecolândia, MS). Anais do 3o Simpósio sobre Recursos Naturais e
Sócio-Econômicos do Pantanal, Corumbá (MS), pp. 90–100.
Biscaye, P., 1965. Mineralogy and sedimentation of recent deep-seaclay in the Atlantic
Ocean and adjacent seas and oceans. Geological Society of America Bulletin 76,
803–832.
Boettinger, J.L., Richardson, J.L., 2001. Saline and wet soils of wetlands in dry climates.
In: Richardson, J.L., Vepraskas, M.J. (Eds.), Wetland soils. Genesis, hydrology, land-
scapes, and classification. Lewis Publishers, New York, pp. 383–390.
Bohn, H.L., McNeal, B.L., O'Connor, G.A., 1985. Soil Chemistry. A Wiley-Interscience
Fig. 10. Example of mica peaks behavior under different treatments: broad and Publication. John Wiley and Sons, New York.
asymmetrical mica peaks on Mg- and/or K-saturated patterns, decrease of width and Borchardt, G., 1989. Smectites. In: Dixon, J.B., Weed, S.B. (Eds.), Minerals in Soil Environ-
asymmetry of mica peaks and better definition of smectite peaks on EG patterns. ments. Soil Science Society of America Book Series, vol. 1. Soil Science Society of
America, Inc., Madison, pp. 675–728.
Brown, G., Brindley, G.W., 1980. X-ray diffraction procedures for clay mineral identi-
involvement of Ca and part of the Mg in the precipitation of fication. In: Brindley, G.W., Brown, G. (Eds.), Crystal Structures of Clay Minerals and
carbonates is confirmed by nodules containing up to 44% of CaCO3eq, their X-ray identification. Mineralogical Society, London, pp. 305–359.
Brum, P.A.R., Sousa, J.C., 1985. Níveis de Nutrientes minerais para gado, em lagoas (“baías” e
consisting of calcite (CaCO3) and dolomite (CaMg(CO3)2). The major “salinas”) no Pantanal Sul-Matogrossense. Pesquisa Agropecuária Brasileira 20,
control of Mg, however, occurs by precipitation of Mg-smectites 1451–1454.
(stevensitic and saponitic minerals) within surface horizons (1 and 2) Crawford, T.W., Whittig, L.D., Begg, E.L., Huntington, G.L., 1983. Eolian influence on develop-
ment and weathering of some soils of Point Reyes Peninsula, California. Soil Science
of the area of seasonal lake-level variation (P3 to P5). Potassium Society of America Journal 47, 1179–1185.
probably participates in the neoformation of Fe-micas in the deepest Cunha, J., 1943. Cobre do Jauru e lagoas alcalinas do Pantanal (Mato Grosso). Boletim do
horizons (5 and 6). The dissolution of kaolinites, which are Laboratório de Produção Mineral 6, 1–43.
Deconinck, J.F., Strasser, A., Debrabant, P., 1988. Formation of illitic minerals at surface
allochthonous in the study area, probably increases the supply of Si,
temperatures in the Purbeckian sediments (lower Berrisian, Swiss and French Jura).
Al, and Fe of the solutions, contributing to the neoformation of the Clay Minerals 23, 91–103.
other clays. Del'Arco, J.O., Silva, R.H., Tarapanoff, I., Freire, F.A., Pereira, L.G.M., Souza, S.L., Luz, D.S.,
Palmeira, R.C.B., Tassinari, C.C.G., 1982. Geologia da Folha SE.21- Corumbá e Parte da
The identification of these mineralogical processes within the soils
Folha SE.20. RADAMBRASIL-Levantamento dos Recursos Naturais. Rio de Janeiro,
surrounding the studied alkaline–saline lake corroborates the hypoth- pp. 25–160.
esis formulated by Barbiero et al. (2002, 2007) that chemical variation Eiten, G., 1983. Classificação da vegetação do Brasil. CNPq/Coordenação Editorial, Brasília.
between fresh and saline waters of Nhecolândia is related to mecha- Eugster, H.P., Hardie, L.A., 1978. Saline lakes. In: Lerman, A. (Ed.), Lakes: chemistry, geology
and physics. Springer-Verlag, Berlin, pp. 237–294.
nisms triggered by present-day evaporative concentration of fresh Fehrenbacher, J.B., Wilding, L.P., Odell, R.T., Melsted, S.W.,1963. Characteristics of solonetzic
waters. Therefore, our study confirms that the salinity of Pantanal, soils in Illinois. Soil Science Society of America Proceedings 27, 421–431.
528 S.A.C. Furquim et al. / Geoderma 154 (2010) 518–528

Fernandes, E., 2000. Caracterização dos Elementos do Meio Físico e da Dinâmica da Orioli, A.L., Amaral Filho, Z.P., Oliveira, A.B., 1982. Pedologia: levantamento exploratório
Nhecolândia (Pantanal Sulmatogrossense). Dissertação de Mestrado. Departa- de solos da Folha SE.21 Corumbá e Parte da Folha SE.20: As regiões fitoecológicas,
mento de Geografia, Universidade de São Paulo (USP). sua natureza e seus recursos econômicos. In: RADAMBRASIL-Levantamento dos
Fernandes, E., 2007. Organização espacial dos componentes da paisagem da baixa Recursos Naturais. Rio de Janeiro, pp. 225–328
Nhecolândia. Tese de Doutorado. Departamento de Geografia. Universidade de São Por, F.D., 1995. The Pantanal of Mato Grosso (Brazil) — World's Largest Wetland. Kluwer
Paulo (USP). Academic Publishers, Dordrecht.
Furquim, S.A.C., Graham, R.C., Barbiero, L., Queiroz Neto, J.P., Vallès, V., 2008. Mineralogy Porrenga, D.H., 1968. Non-marine glauconitic illite in the Lower Oligocene of Aardebrug,
and genesis of smectites in an alkaline–saline environment of Pantanal wetland, Belgium. Clay minerals 7, 421–430.
Brazil. Clays and Clay Minerals 56, 580–596. Ransom, M.D., Bigham, J.M., Smeck, M.E., Jaynes, W.F., 1988. Transitional vermiculite–
Godoi Filho, J.D., 1986. Aspectos Geológicos do Pantanal Mato-grossense e de sua Área smectite phases in Aqualfs in southwestern Ohio. Soil Science Society of America
de Influência. Anais do 1o Simpósio sobre Recursos Naturais e Sócio-Econômicos do Journal 52, 873–880.
Pantanal. Corumbá, pp. 63–76. Reid-Soukup, D.A., Ulery, A., 2002. Smectites. In: Dixon, J.B., Schulze, D.G. (Eds.), Soil
Güven, N.,1988. In: Bailey, S.W. (Ed.), Smectites. Reviews in Mineralogy, vol.19, pp. 497–559. Mineralogy with Environmental Application. Soil Science Society of America Book
Harder, H., 1974. Illite mineral synthesis at surface temperatures. Chemical Geology 14, Series, vol. 7. Soil Science Society of America, Inc., Madison, pp. 467–499.
241–253. Sakamoto, A.Y., 1997. Dinâmica hídrica em uma lagoa salina e seu entorno no Pantanal
Herbillon, A.J., 1976. Iron in kaolinite with special reference to kaolinite from tropical da Nhecolândia: contribuição ao estudo das relações entre o meio físico e a
soils. Clay Minerals 11, 201–220. ocupação, Fazenda São Miguel Firme, MS. Tese de Doutorado. Departamento de
Hover, V.C., Walter, L.M., Peacor, D.R., Martini, A., 1999. Mg-smectite authigenesis in a Geografia, Universidade de São Paulo (USP).
marine evaporative environment, salina Ometepec, Baja California. Clays and Clay Scott, D.A., 1991. Latin America and Caribbean. In: C. M., M.E., Moser (Eds.), Wetlands: A
Minerals 47, 252–268. Global Perspective. Facts on File, New York, pp. 85–114.
Huertas, F.J., Chou, L., Wollast, R., 1999. Mechanism of kaolinite dissolution at room tem- Silva, M.H.S., Sakamoto, A., 2003. Perfis pedomorfológicos do Pantanal da Nhecolândia-
perature and pressure. Part II — kinetic study. Geochimica et Cosmochimica Acta 63, MS: Um estudo comparativo. XII Encontro Sul-Matogrossense de Geografia, Três
3261–3275. Lagoas, MS, pp. 544–552.
Huggett, J.M., Gale, A.S., Clauer, N., 2001. The nature and origin of non-marine 10A clay Silva, M.H.S., Bacani, V.M., Sakamoto, A., 2004. Caracterização do solo de uma lagoa salina
from the Late Eocene and Early Oligocene of the Isle of Wight (Hampshire Basin), na área da fazenda Santo Inácio, Pantanal da Nhecolândia, MS. XIII Encontro Estadual
UK. Clay Minerals 36, 447–464. de Geografia. Aquidauana, MS.
Jackson, M.L., 1979. Soil Chemical Analysis—Advanced Course. By author, Madison. Silva, T.C., 1986. Contribuição da Geomorfologia para o Conhecimento e Valorização do
Jamoussi, F., Ben Aboud, A., López Galindo, A., 2003. Palygorskite genesis through silicate Pantanal. Anais do 1o Simpósio sobre Recursos Naturais e Sócio-Econômicos do
transformation in Tunisian continental Eocene deposits. Clay Minerals 38, 187–199. Pantanal. Corumbá, pp. 77–90.
Klammer, G., 1982. Die palaeowuste des Pantanal von Mato Grosso und die pleistozane Soares, A.P., Soares, P.C., Assine, M.L., 2003. Areiais e lagoas do Pantanal, Brasil: herança
Klimageschchte der brasilianischen Randtropen. Zeitschrift fur Geomorphologie paleoclimática? Revista Brasileira de Geociências 33, 211–224.
26, 393–416. Tardy, Y., Bocquier, G., Paquet, H., Millot, G., 1973. Formation of clay from granite and its
Kohut, C.K., Dudas, M.J., 1995. Evaporite mineralogy and trace-element content of salt- distribution in relation to climate and topography. Geoderma 10, 271–284.
affected soil in Alberta. Canadian Journal of Soil Science 73, 399–409. Theisen, A.A., Harward, M.E., 1962. A paste method for preparation of slides for clay
Mahjoory, R.A., 1979. The nature and genesis of some salt-affected soils in Iran. Soil mineral identification by x-ray diffraction. Soil Science Society of America
Science Society of America Journal 43, 1019–1024. Proceedings 26, 90–91.
Mourão, G.M., Ishii, I.H., Campos, Z.M.S., 1988. Alguns fatores limnológicos relacionados Torrez-Ruiz, J., López-Galindo, A., Gonzalez-López, J.M., Delgado, A., 1994. Geochemistry
com a ictiofauna de baías e salinas do Pantanal da Nhecolândia, MS, Brasil. Acta of Spanish sepiolite–palygorskite deposits: genetic considerations based on trace
Limnológica Brasileira 2, 181–198. elements and isotopes. Chemical Geology 112, 221–245.
Nelson, D.W., Sommers, L.E., 2001. Total carbon, organic carbon, and organic matter. In: Tricart, J., 1982. El Pantanal: Un ejemplo del impacto de la Geomorfología sobre el medio
Sparks, D.L. (Ed.), Methods of Soil Analysis — Chemical Analysis 3th edition. Soil ambiente. Geografia 7, 37–50.
Science Society of America, Madison, pp. 961–1010. Velde, B., 1995. Origin and Mineralogy of Clays. Clays and the Environment. Springer,
Nelson, R.E., 1982. Carbonate and gypsum, In: Page, A.L. (Ed.), Methods of Soil Analysis — New York.
Chemical and Microbiological Properties, part 2, 2nd edition. SSSA Inc., Madison, Weaver, C.E., Pollard, L.D., 1973. The Chemistry of Clay Minerals. Elsevier, Amsterdam.
pp. 181–198. Wilhelmy, H., 1958. Umlaufseen und Dammuferseen tropischer Tieflandflusse. Zeitschrift
Newman, A.C.D., Brown, G., 1987. The chemical constitution of clays. In: Newman, A.C.D. fur Geomorphologie NF 2, 27–54.
(Ed.), Chemistry of clays and clay minerals. Longman Scientific & Technical.
Mineralogical Society, New York, pp. 1–128.

You might also like