J Paleolimnol (2011) 46:273–289 DOI 10.1007/s10933-011-9538-5 ORIGINAL PAPER Limnogeology in Brazil’s ‘‘forgotten wilderness’’: a synthesis from the large floodplain lakes of the Pantanal Michael M. McGlue • Aguinaldo Silva • Fabrı́cio A. Corradini • Hiran Zani Mark A. Trees • Geoffrey S. Ellis • Mauro Parolin • Peter W. Swarzenski • Andrew S. Cohen • Mario L. Assine • Received: 17 March 2011 / Accepted: 27 June 2011 / Published online: 13 July 2011 ! Springer Science+Business Media B.V. 2011 Abstract Sediment records from floodplain lakes have a large and commonly untapped potential for inferring wetland response to global change. The Brazilian Pantanal is a vast, seasonally inundated Electronic supplementary material The online version of this article (doi:10.1007/s10933-011-9538-5) contains supplementary material, which is available to authorized users. M. M. McGlue (&) ! M. A. Trees ! A. S. Cohen Department of Geosciences, The University of Arizona, 1040 East 4th Street, Tucson, AZ 85721, USA e-mail: mmmcglue@email.arizona.edu M. A. Trees e-mail: treesma@email.arizona.edu A. S. Cohen e-mail: cohen@email.arizona.edu A. Silva Departamento de Ciências do Ambiente, Universidade Federal de Mato Grosso do Sul—UFMS-CPAN, Av. Rio Branco, 1270, Corumbá, MS 79304-902, Brazil e-mail: aguinald_silva@yahoo.com.br F. A. Corradini Faculdade de Geografia, Universidade Federal do ParáUFPA, Folha 31, Quadra 7, Lote Especial S/N, Marabá, PA 68501-970, Brazil e-mail: f_coradini@yahoo.com.br savanna floodplain system controlled by the flood pulse of the Upper Paraguay River. Little is known, however, about how floodplain lakes within the Pantanal act as sedimentary basins, or what influence hydroclimatic variables exert on limnogeological processes. This knowledge gap was addressed through an actualistic analysis of three large, shallow (\5 m) floodplain lakes in the western Pantanal: Lagoa Gaı́va, G. S. Ellis Energy Resources Program, U.S. Geological Survey, Denver, CO, USA e-mail: gsellis@usgs.gov M. Parolin Faculdade Estadual de Ciências e Letras de Campo Mourão, Av. Comendador Norberto Marcondes, 733, Campo Mourão, PR 87303-100, Brazil e-mail: mauroparolin@gmail.com P. W. Swarzenski U.S. Geological Survey, Santa Cruz, CA, USA e-mail: pswarzen@usgs.gov M. L. Assine Departamento de Geologia Aplicada—IGCE, Universidade Estadual Paulista—UNESP/Campus Rio Claro, Av. 24-A, 1515, Rio Claro, SP 13506-900, Brazil e-mail: assine@rc.unesp.br H. Zani Divisão de Sensoriamento Remoto, Instituto Nacional de Pesquisas Espaciais—INPE, Av. dos Astronautas, 1758, São José dos Campos, SP 12201-970, Brazil e-mail: hzani@dsr.inpe.br 123 274 Lagoa Mandioré and Baia Vermelha. The lakes are dilute (CO32- [ Si4? [ Ca2?), mildly alkaline, freshwater systems, the chemistries and morphometrics of which evolve with seasonal flooding. Lake sills are bathymetric shoals marked by siliciclastic fans and marsh vegetation. Flows at the sills likely undergo seasonal reversals with the changing stage of the Upper Paraguay River. Deposition in deeper waters, typically encountered in proximity to margin-coincident topography, is dominated by reduced silty-clays with abundant siliceous microfossils and organic matter. Stable isotopes of carbon and nitrogen, plus hydrogen index measured on bulk organic matter, suggest that contributions from algae (including cyanobacteria) and other C3-vegetation dominate in these environments. The presence of lotic sponge spicules, together with patterns of terrigenous sand deposition and geochemical indicators of productivity, points to the importance of the flood pulse for sediment and nutrient delivery to the lakes. Flood-pulse plumes, waves and bioturbation likewise affect the continuity of sedimentation. Short-lived radioisotopes indicate rates of 0.11–0.24 cm year-1 at sites of uninterrupted deposition. A conceptual facies model, developed from insights gained from modern seasonal processes, can be used to predict limnogeological change when the lakes become isolated on the floodplain or during intervals associated with a strengthened flood pulse. Amplification of the seasonal cycle over longer time scales suggests carbonate, sandy lowstand fan and terrestrial organic matter deposition during arid periods, whereas deposition of lotic sponges, mixed aquatic organic matter, and highstand deltas characterizes wet intervals. The results hold substantial value for interpreting paleolimnological records from floodplain lakes linked to large tropical rivers with annual flooding cycles. Keywords Pantanal ! Limnogeology ! Floodplain lakes ! Tropical wetlands ! Sedimentary organic matter ! Freshwater sponges Introduction Global estimates indicate that the total land surface area occupied by wetlands in the humid tropics and subtropics exceeds 3.3 9 106 km2 (Maltby and Turner 1983). Prior to the twentieth century, attitudes 123 J Paleolimnol (2011) 46:273–289 towards such wetlands were largely negative, fueled by the desire to reclaim perennially-inundated lands for agriculture or industry (Mitsch and Gosselink 2000). More recently, international efforts have demonstrated the value of wetlands conservation, especially for the protection of numerous aquatic species and for maintaining hydrologic resources critical to downstream human populations. Importantly, vast tropical wetlands are now recognized for the unique role they play in global biogeochemical cycles, particularly in the production and sequestration of greenhouse gases (Bartlett and Harriss 1993; Cao et al. 1998; Mitsch et al. 2009). Global changes in air temperature or increased variability in the water cycle, such as those predicted by the IPCC AR4, are thought to place wetland ecotones at substantial risk for terrestrialisation (Chauhan and Gopal 2001; Bates et al. 2008). Few in-depth paleo-records from unaltered, tropical wetlands are available, however, to assess the response of these sensitive systems to wellknown climatic perturbations of the Quaternary (Donders et al. 2005). Stratigraphic sequences from large lakes in the Pantanal may provide a unique lens through which the response of tropical wetlands to environmental change can be viewed (Fig. 1). The Pantanal is the world’s largest freshwater wetland and the hydrologic basin of the Upper Paraguay River (PR; Heckman 1998). Scientific inquiry into the region was delayed until the twentieth century, as early attempts at exploration were deterred by harsh conditions along the Brazilian frontier (Por 1995). Sixteenth-century maps depicted the Pantanal as a single large lake, purportedly because of reports given to Spanish conquistadors by lakeshore-dwelling Indians (Por 1995). Surveying conducted centuries later revealed that thousands of lakes dot the Pantanal landscape (Ab’Saber 1988; Assine and Soares 2004). Limnogeological datasets from the Pantanal are, however, generally absent, constraining an understanding of the lakes and their linkages to the PR floodplain. The purpose of this study is to provide the first synthesis of modern limnogeological knowledge on three lakes situated along the floodplain of the PR: Lagoa Gaı́va, Lagoa Mandioré and Baia Vermelha (Fig. 1B). Several research groups noted the potentially valuable archive of paleoclimate information stored within deposits from these and nearby floodplain lakes (Mayle et al. 2004; de Oliveira Bezerra J Paleolimnol (2011) 46:273–289 275 Fig. 1 A = Regional map of the Brazilian Pantanal, the world’s largest neotropical wetland and the hydrological basin of the Upper Paraguay River (adapted from Assine and Silva 2009). Ten large perennial lakes mark the western floodplain of this river, straddling the border of Brazil and Bolivia. The rectangle outlines the area of B. B = LANDSAT (2000) image showing the position of LG, LM, and BV with respect to the main channel of the PR and the highlands of the Serras do Amolar. C = bathymetric map and sample grid for LG; sediment core locations are marked with stars. D = bathymetric map and sample grid for LM. Heavy red lines indicate the location of high-relief lake margins and rocky shorelines; orange lines mark sand bars. E = bathymetric map and sample grid for BV and Mozeto 2008). Inferences from lake sediment may likewise be vital for understanding more recent environmental dynamics in the greater Pantanal. As tropical wetlands are ecologically sensitive, such information holds considerable value for conservation and sustainability planning. Before longer sediment core-based reconstructions can be fully realized, a number of key questions need to be addressed, including: How do these lakes act as depositional basins? What are the dominant limnological processes that influence patterns of sedimentation? What baseline variability exists in the geochemistry of organic matter in modern lake sediments? What relationships exist between the lakes and the greater Pantanal wetland system? The focus of this study is to address these questions using observations and data acquired from the lakes from 2007 to 2009. et al. 1988; Por 1995). Estimates from remote sensing indicate that the surface area of the Pantanal that is inundated at maximum flood conditions exceeds 130,000 km2 (Hamilton et al. 2002). The region is particularly renowned for the biodiversity of its flora and for its role as a natural wildlife sanctuary, especially for numerous species of fish and aquatic birds (Heckman 1998; Junk et al. 2006; Lopes et al. 2007). Vegetation is a mosaic of cerrado (tropical savanna), Amazon-derived semi-deciduous forest, Chaco-derived seasonal dry forest and aquatic plants (Cole 1960; Prance and Schaller 1982). The spatial distribution of plant communities is heterogeneous and controlled dominantly by patterns of seasonal flooding, topography and soil type (Pinder and Rosso 1998). Lagoa Gaı́va (LG), Lagoa Mandioré (LM) and Baia Vermelha (BV) are located along the western margin of the PR (Fig. 1). The lakes are amongst the largest perennial standing bodies of water in the Pantanal and its Paraguai-Corixo Grande sub-region (Electronic Supplementary Material Table 1; de Magalhães 1992). The lakes are interspersed within the Serra do Amolar, a mountain range with[800 m of local relief. Site description The Pantanal (16–20"S, 55–58"W) straddles the borders of Brazil, Bolivia and Paraguay (Fig. 1; Alho 123 276 Geology The Pantanal sits in a low-altitude (mean elevation \200 m above sea level), elliptically-shaped, seismically-active basin (Fig. 1; Assine and Soares 2004). Andean tectonics are generally invoked to explain the presence of the Pantanal basin, although its position within the foreland remains equivocal. Horton and DeCelles (1997) argued that the Pantanal forms the backbulge depozone of the Modern Andean foreland basin system. More recent research by Chase et al. (2010) supports this hypothesis, as geoid anomalies indicate the presence of a flexural forebulge [200 km west of the Pantanal. In contrast, Ussami et al. (1999) used geophysical data to suggest that the Pantanal depression formed through extension, set in motion by the interaction of the Andean forebulge with the Neoproterozoic Paraguai foldthrust belt. Regardless of its tectonic origin, at least 500 m of Quaternary sediments fill the Pantanal basin (Ussami et al. 1999). Although a number of fault trends are apparent in the basin, the NE-SW-trending Transbrasiliano Lineament is most significant, as the course of the PR is strongly altered by its presence in a number of locations (Assine and Soares 2004). Several large rivers draining basin-margin highlands enter the Pantanal, forming expansive, low-gradient fluvial megafans north and east of the PR (Assine and Silva 2009). A wide variety of lacustrine environments are present in association with distal megafan surfaces, including thousands of saline and fresh ponds in the Nhecolândia region of the southern Pantanal (Tricart 1982; Barbiéro et al. 2002). J Paleolimnol (2011) 46:273–289 net annual loss approaching 300 mm (Alfonsi and Paes de Camargo 1986; Por 1995). A shift in wind direction accompanies the change in seasons, as austral summer winds are from the northwest, whereas winds in the austral winter are dominantly from the east-northeast. Over the past 40 years, dry season wind velocities average 5–6 m/s near the study lakes, whereas wind speeds during the rainy season are comparatively slower. Hydrology and limnology The hydrology of the Pantanal is controlled by the flooding cycle of the PR, as well as the combination of precipitation and evapo-transpiration. A particularly notable aspect of PR’s annual flooding cycle is the delayed passage of its flood pulse moving from north to south, altering stage in the main channel by more than 5 m over the course of several months (Hamilton et al. 1997). Small tributary channels connect each of the study lakes to the PR (Fig. 1A, C, D, E). Exchange between the lakes and the PR is strongest following the passage of the flood wave, which commonly occurs several months after peak austral summer precipitation. Seasonal retention of the flood waters in the northern Pantanal results from the different mechanisms of floodplain inundation, including: (1) overbank flow of the PR and smaller tributaries, commonly in association with heavy seasonal rainfall; (2) a backwater effect, where overland flow on floodplains becomes impeded by the elevated stage of the PR; and (3) local ponding of rainfall on heavilyvegetated floodplain soils (Hamilton 1999). Climate Tropical latitude and seasonal migration of the Intertropical Convergence Zone (ITCZ) controls climate patterns in the Pantanal. Mean annual air temperature in the region is *25"C. Precipitation is spatially variable and strongly seasonal. Near the study area, annual precipitation typically exceeds 1,000 mm, whereas precipitation to the north is greater, but more seasonally variable (Electronic Supplementary Material Fig. 1). The majority of precipitation ([70%) falls during a prolonged wet season that lasts from late October to early April. Evaporation exceeds precipitation during most of the year in the Pantanal, with some areas experiencing a 123 Materials and methods Approach Tropical wetlands typically house a wide variety of depositional environments that are linked by prevailing climate and the hydrology of large rivers. Inundation and passage of the PR flood pulse are key mechanisms controlling floodplain ecological interactions and productivity in the Pantanal (Junk et al. 1989; de Oliveira and Calheiros 2000). However, the implications of flood-pulse dynamics for limnogeological processes represent a key knowledge J Paleolimnol (2011) 46:273–289 gap for the region. In order to redress this gap, we undertook an assessment of modern floodplain lakewater chemistry, bathymetry, morphometry, physical sedimentology and sedimentary biogeochemistry. Integrated analyses of limnological processes and their influence on geological products are vital, as accurate interpretations of paleoenvironments derived from sediment cores hinges on robust calibration datasets that validate the meaning and fidelity of sedimentary proxy data. We conducted reconnaissance bathymetric and water sampling surveys in 2007, 2008, and 2009 using a hand-held fathometer. Water chemistry was determined in situ using a YSI model 85 multi-meter for dissolved oxygen, conductivity and temperature, and a Hach Sension1 meter was used to measure pH (Electronic Supplementary Material Table 2). Major ions were determined using a Perkin-Elmer Optima 5300 ICP-OES at the University of Arizona. The upper 2–3 cm of the modern lake floors (n = 40, 69 and 45 for LG, LM and BV, respectively) were collected using a Ponar grab-sampler in order to assess the physical and geochemical attributes of sediments that have recently accumulated in these lakes. Sample grids were constructed to provide maximum coverage of the different environments encountered in the field. Spacing between individual sampling stations varied, but in all cases was \2.5 km. Short (\1.5 m) sediment cores were collected and dated to evaluate rates of sedimentation. Cores were collected from BV and LM in 2008 and LG in 2009 using a Livingstone-style square rod corer or an aluminum-barrel vibracorer system. Sediment geochronologies and linear sedimentation rates (cm year-1) were derived for LG cores using the decay of excess 210Pb (t1/2 = 22.3 year) at the USGS in Santa Cruz, CA, following the calculations presented in Swarzenski et al. (2006). Precision in the activity of excess 210Pb was typically better than 10%. Smear slides were inspected under a petrographic microscope to estimate sediment components. Concentrations of biogenic silica (BiSi), total organic carbon (TOC), total nitrogen (TN) and stable isotopes of C and N were measured to assess productivity and the composition of organic matter (OM) in the lake sediments. Biogenic silica analyses were completed at the University of Minnesota utilizing multiple extractions of hot alkaline digestions following a modified protocol of DeMaster (1979). Reported 277 values have an analytical precision of *1.0%. Elemental and stable isotopic analyses of sediment OM were conducted at the University of Arizona. Total organic carbon, TN, d13COM, and d15NOM were measured on a continuous-flow gas-ratio mass spectrometer (Finnigan Delta PlusXL) coupled to a Costech elemental analyzer. Samples were decarbonated using 1 M HCl at room temperature for 1 h, washed four times with deionized water, and dried at 40"C prior to combustion in the elemental analyzer. Standardization is based on acetanilide for elemental concentration and NBS-22 and USGS-24 for d13Com. Precision was better than ±0.09 and 0.20 for d13Com and d15Nom, respectively, based on repeated standard analyses. d13COM values are given with respect to the PDB standard and d15NOM values are reported with respect to air. A select number of decarbonated samples were analyzed by Rock–Eval pyrolysis at the University of Houston to further discriminate the provenance of sediment OM (Espitalie et al. 1977). Particle size analyses were conducted at the University of Arizona using a Malvern laser-diffraction analyzer coupled to a Hydro 2000S sample dispersion bench. Particle size provides key insights into environmental energy and sedimentation processes (traction flow vs. suspension settling). Samples were pre-treated according to Johnson and McCave (2008) to remove BiSi and OM, whereas authigenic minerals were removed by excess hot 1 M HCl digestion for 24 h. Following completion of the digestions, samples were disaggregated using a solution of sodium hexametaphosphate and a mechanical shaker. Prior to analysis, sediments were examined optically to ensure the removal of non-terrigenous particles. A subset of samples from each lake was processed for sponge microfossils at the Faculdade Estadual de Ciências e Letras de Campo Mourão, following the procedure outlined in Volkmer-Ribeiro and Turcq (1996). Sample aliquots were suspended in water, dried on microscopic slides, Entellan-mounted and permanently sealed with cover slips. Species identification and nomenclature followed Volkmer-Ribeiro and Pauls (2000) and included: (1) megascleres, which are spicules that integrate the sponge skeletal network; (2) microscleres, a smaller spicule within the sponge skeleton; and (3) gemmoscleres, spicules that cover the gemmules and which ultimately define families, genera and species in freshwater sponges. 123 278 J Paleolimnol (2011) 46:273–289 Results Sedimentology and geochemistry Morphometrics, bathymetry, and water chemistry Mean siliciclastic particle size trends and the concentrations of BiSi, TOC, and TN vary among LG, LM and BV (Fig. 2). In LG and LM, dark green silts and clays are commonly associated with the deepest areas of the basins, whereas similar sediments in BV were deposited only west of the headland, where water depths exceeded 2.5 m. Photomicrographs of sediment components in these environments are presented in Electronic Supplementary Material Fig. 2. Terrigenous grains in the sand fraction ([63 lm) are common in the lakes, especially proximal to basin sill points. In LG, a broad, sandy fan emanating from the basin sill occupies the northern sub-basin (Fig. 2). Likewise, lake floor substrates at the northern end of the southern subbasin, where the lake receives weak overflow from Lagoa Uberaba, are dominantly sandy. In LM, axial margins and cuspate bays are commonly sand-rich environments. Inflow from the PR at the basin sill forms a sandy deltaic fan with a surface area [40 km2, whereas a much smaller sand-rich delta exists along the northern axis of LM (Fig. 2). Approximately 3 km offshore from the northern delta, a 1.5-km-long, partially-vegetated sand bar creates a pronounced bathymetric shoal; a similar shoreline-parallel bar exists off a cuspate bay on the southwestern margin of the lake. Whereas coarsegrained sediment west of the headland in BV is confined near the shoreline, lake floor sediment east of the headland is commonly sandy. A northeastoriented, *30-km2, sand-rich fan occupies much of the center of the basin east of the headland. Along the eastern lake margin, sandy sediments mark floodpulse connection points with the PR. Values of BiSi in LG range between 0.0 and 3.5 wt%, with a mean of 1.5 wt% (Electronic Supplementary Material Table 3). Biogenic silica concentrations are highest on the eastern side of southern LG, where samples exceed 2.5 wt% (Fig. 2). Elsewhere in the basin, lake floor sediment has less BiSi, with samples generally exhibiting values below 2.0 wt%. Total nitrogen ranges from 0.0 to 0.4 wt%, with a mean of 0.2 wt%, whereas values of TOC range from 0.0 to 3.3 wt%, with a mean of 1.5 wt%. Contours of TN and TOC are simple and concentric in the southern sub-basin. Geochemical analyses of OM from LG indicate relatively narrow ranges of Figure 1 and Electronic Supplementary Material Tables 1 and 2 present morphometrics, bathymetry and water chemistry data for LG, LM and BV. The lakes share a number of limnological characteristics in common after the passage of the PR flood pulse, including: (1) dominant CO32- [ Si4? [ Ca2? hydrochemistries; (2) well-oxygenated water columns; (3) low (freshwater) conductivity values; (4) weaklybasic pH values; (5) elongate basin shapes with fetch distances \25 km; (6) shallow mean water depths (\3.5 m); and (7) maximum water depths proximal to high relief lake margins. Complete thermal stratification was not detected in any of the lakes, but confirmation of mixing state awaits future studies. Recent satellite imagery indicates that LG has a surface area of *100 km2 during passage of the flood. The basin sill is located along the eastern margin of the northern sub-basin of the lake, at an altitude of *95 m a.s.l. Proximal to the sill, LG is very shallow, with water depths B2.5 m at a distance of *2 km from shore (Fig. 1C). A bathymetric low exists \3 km from the eastern high-relief margin of the southern sub-basin; water depths were routinely [4.5 m in this area. Lagoa Mandioré has a surface area [150 km2 during the passage of the flood pulse. The basin sill is on the southeastern lake margin at an altitude of *93 m a.s.l. Water depths generally increase with distance from the shoreline in LM. An exception to this trend occurs\1 km from the basin’s northeastern high-relief margin, where measurements of water depth were C4.0 m (Fig. 1D). Baia Vermelha exhibits a highly variable surface area (B150 km2), resulting from an expansive, low-gradient plain on the northern lake margin (Fig. 1E). Several intermittent channels (*91–92 m a.s.l.) connect BV with the PR on the eastern side of the basin. The bathymetry of BV is complicated by the presence of a circular headland *3.5 km offshore from the northwestern axial lake margin. West of the headland, bathymetric contours are approximately concentric and reach a maximum depth of *3.1 m (Fig. 1E). Baia Vermelha is deepest (*3.5 m) east of the headland, *2 km from its southern, high-relief lake margin. 123 J Paleolimnol (2011) 46:273–289 279 Fig. 2 Concentration maps of mean particle size, biogenic silica, total nitrogen and total organic carbon for LG, LM, and BV. Top row = LG. Middle row = LM. Bottom row = BV. The contour interval is the same for each of the lakes, illustrating the elevated productivity and burial of OM in LM versus the other lakes C/N, d13COM, and d15NOM values (Fig. 3). Values of C/N and d13COM typically plot within known fields for lacustrine algae mixed with C3-pathway terrestrial plant matter (Meyers and Ishiwatari 1993). Values of d15NOM range from ?1.8 to ?3.3%, with a mean of ?2.5%. Hydrogen index values for LG sediments are \200 mg HC/gm TOC and plot near the Type II– Type III kerogen boundary (Fig. 3). Lake floor sediments in LM exhibit BiSi values between 0.0 and 12.9 wt%, with a mean of 4.3 wt% 123 280 J Paleolimnol (2011) 46:273–289 Fig. 3 Cross plots of key biogeochemical indicators in the surface sediments of LG, LM, and BV. A = TN versus TOC. Sediments in LM are particularly rich in carbon and nitrogen, suggesting elevated rates of primary productivity. B = d13C versus C/N. Note the elevated C/N values exhibited by BV sediments, probably due to an increased role of vascular plant material in the basin’s carbon cycle. Sediments from LG fall within well-defined ranges for mixed algal and C3-terrestrial vegetation. C = d15N versus d13C. Low d15N values in some LM sediments likely results from the presence of nitrogenfixing cyanobacteria in the basin. D = HI versus OI. Rock– Eval pyrolysis data support a mixed organic provenance interpretation for lakes in the Pantanal (Electronic Supplementary Material Table 3). Axial concentrations of BiSi are high ([7.0 wt%) in LM, especially \2.5 km from the western lake shore (Fig. 2). Sediments deposited near the basin sill have less BiSi (B2.0 wt%). Total nitrogen ranges from 0.0 to 1.0 wt%, with a mean of 0.5 wt%, whereas values of TOC are particularly high, ranging from 0.0 to 8.5 wt%, with a mean of 4.9 wt%. Contours of TN and TOC indicate elevated concentrations on the western side of the basin axis, away from fluvial point sources (Fig. 2). Total nitrogen and TOC demonstrate statistically significant correlations with water depth in LM (r2 = 0.51 and 0.46, respectively). Sediment OM in LM exhibits the broadest range of d13C and d15N values, whereas the range of C/N values (10.0–13.6) is comparable to the other lakes (Fig. 3). The mean d13COM isotopic composition (-24.0%) is at least 2.5% more positive than the means of LG and BV. In contrast, values of d15NOM are the smallest in the study area, with a range of -0.1 to ?3.2% and a mean of ?1.4% (Electronic Supplementary Material Table 3). Hydrogen index values for LM surface sediments range from 206 to 357 mg HC/gm TOC and fall within the Type II kerogen field (Fig. 3). Lake floor sediments in BV exhibit BiSi values between 0.0 and 4.3 wt%, with a mean of 1.6 wt% (Electronic Supplementary Material Table 3). The highest concentrations of BiSi are located west of the headland and near the southeastern lake margin; most of the lake floor substrates between the axial margins are \2.0 wt% BiSi (Fig. 2). Total nitrogen ranges from 0.0 to 0.4 wt%, with a mean of 0.2 wt%, whereas values of TOC range from 0.0 to 4.4 wt%, with a mean of 2.0 wt%. The patterns of TN and TOC contours are similar to BiSi, with elevated concentrations along the axial margins (Fig. 2). C/N values on BV samples are slightly higher (mean = 12.5) than those measured in the other lakes, whereas d13COM sample means are comparable to LG (-26.5 and -27.0%, respectively). Sediment OM in BV exhibits d15NOM values that range from ?0.3 to ?4.4%, with a mean of ?3.1%. Hydrogen index 123 J Paleolimnol (2011) 46:273–289 values for BV sediments range from 123 to 287 mg HC/gm TOC and plot near the Type II–Type III kerogen boundary (Fig. 3). Sedimentation rates were developed on two cores from LG using 210Pb (Fig. 4). The upper 20–30 cm of both cores contain massive, dark-green, silty clay with OM, diatoms and sponge spicules commonly present. Linear sedimentation rates vary spatially within the southern sub-basin. Recent sedimentation has been relatively slow (0.11 cm year-1) at core site GVA09-3A, which is situated near the northern end of the southern sub-basin (Fig. 1). At the southern core site (GVA09-2A) recent sedimentation has progressed more rapidly, at a rate of 0.24 cm year-1. A 137Cs-derived geochronology for GVA09-3A of 0.1 cm year-1 supports the sedimentation rate obtained from the excess 210Pb distribution. Sponge microfossils Figure 5 and Electronic Supplementary Material Table 4 illustrate the variety of sponge microfossils in the modern sediments of LG, LM and BV. Megascleres and gemmoscleres of two lentic species are common to all three lakes: Radiospongilla amazonensis Volkmer-Ribeiro and Maciel 1983 and Trochospongilla variabilis Bonetto and Ezcurra de Drago 1973. In BV, spicules of the lentic sponge Corvoheteromeyenia sp. range from common to abundant in lake floor sediments west of the headland, but are much less common proximal to connection points with the PR. Modern sediments in the southern sub-basin of LG contained spicules of Metania spinata Carter 1881 and Heteromeyenia sp. sponges; both are commonly associated with lakes and wetlands in the cerrado biome (Volkmer-Ribeiro 1992). Notably, all three lakes contain the robust megascleres and gemmules of Corvospongilla secktii Bonetto and Ezcurra de Drago 1966, a sponge known to colonize rocky substrates and channel-margin vegetation in large neotropical rivers (Fig. 5; Batista et al. 2003). These spicules are found in sediments across LG and BV, but are more localized in LM. Lagoa Mandioré also contains abundant spicules of the lotic sponge Oncosclera navicella Carter 1881 in sediments deposited near the basin sill; such spicules were not detected in the center and northern axis of the basin. In LG, spicules of lotic sponges are present 281 across the basin, and include gemmoscleres of Uruguaya corallioides Bowerbank 1863. Discussion Any discussion of limnogeological processes in the floodplains of the Pantanal must consider the downstream variability exhibited in the passage of the PR flood pulse. A key characteristic distinguishing LG, LM and BV from floodplain lakes in many other tropical watersheds is the near anti-phasing of the dominant rainy season with the arrival of riverine flood waters (Veloso 1972; Heckman 1998). The 2- to 4-month time lag associated with the passage of the PR flood pulse from the north to the central Pantanal is important to lake hydrochemistry, morphometrics, and sedimentation processes. The arrival of the PR flood pulse following the austral summer causes lake surface areas, fetch distances and bathymetric gradients to reach their maxima as wind speeds begin to reach peak velocity (Electronic Supplementary Material Fig. 1). This combination has implications for wave development, sediment re-suspension and the temporal continuity of stratigraphic records. Because of its tapered shape and short effective fetch, the deepest regions of LG likely escape the influence of waves (critical depth = 4.1 m), whereas the potential for wave reworking and sediment re-suspension is much higher in LM and BV due to their morphometries. This difference provides a plausible explanation for the ‘‘well-behaved’’ decay of excess 210Pb in the uppermost sediments in LG (Fig. 4). Shortterm stratigraphic records from the deepest parts of LG are continuous, whereas those from LM and BV appear to be more punctuated. Unsupported 210Pb in LM is confined to the upper 2–3 cm of sediment cores, and is interpreted to result from recent wavedriven re-suspension of sediments at the core site. The impact of waves on LM is further illustrated by shoreline-parallel sand bars and cuspate sandy beaches. Such features are well known from lakes with high-energy coastlines (Adams and Wesnousky 1998). In BV, erratic downcore concentrations of excess 210Pb suggest that bioturbation may impact the continuity of the stratigraphy of this basin (Sharma et al. 1987). Feral cows and gregarious water birds in the central Pantanal make the potential for shallowwater bioturbation high (Por 1995). Highly variable 123 282 J Paleolimnol (2011) 46:273–289 Fig. 4 210Pb profiles for sediment cores from LG; location of cores shown in Fig. 1. Continuity of sedimentation among the study lakes is highly variable, probably due to differences in morphometries and bathymetry. Fairly linear decline of log excess 210Pb with depth in LG due to the lake’s water depth, which at many locales exceeds the critical wave depth. Dpm/ g = disintegrations per minute per gram. Dark symbols = mixed layer samples, grey symbols = samples used in age model sedimentation rates could produce a similar downcore 210 Pb decay pattern, but massive lithologic units bounded by irregular bedding planes in the BV sediment core suggest this alternative is less likely. 123 Another key difference between the Pantanal lakes and other floodplain lakes is the mechanism of riverine inundation. The water cycles of many tropical floodplain lakes rely on groundwater seepage or overland flow when bankfull conditions are exceeded (Cohen 2003). In contrast, LG, LM and BV are silled basins linked to the PR by channels. Tributaries connecting the lakes to the PR not only influence water cycling, but also impact lake sedimentology. Channels impart a first-order control on flow velocity and the routing of bed-load into the lake basins, whereas unconfined overland flows are more diffuse, allowing vegetated levees and floodplain soils to act as sediment traps. Terrigenous particlesize contours show the impact of fluvial tractionflows on the lakes (Fig. 2). Fan-shaped sand bodies occupy broad areas of the lake floors and are key components of the modern facies architecture that are genetically linked to the passage of the flood pulse. The orientation of fan apices confirms that flows originating at basin sills are the source of these deposits. Another hallmark of ‘‘fill-phase’’ sedimentology is the presence of lotic sponge microfossils, which reflect at least seasonal connectivity between the lakes and the PR. The ecological characteristics of sponge remains are well known, and the spicules of lotic species are typically well preserved, albeit with some taphonomic overprint from transport (Batista et al. 2003; Volkmer-Ribeiro and Pauls 2000). Prior research in Brazil has demonstrated the utility of spicule analysis to discriminate between fluvial and lacustrine environments in sediment cores (Parolin et al. 2007). Results from the present study show that down-core variability in lotic versus lentic assemblages could provide an important ecological perspective on fluvial-lacustrine connectivity for the floodplain lakes of the Pantanal. In our study lakes, mean particle sizes are generally in the silt range. Coarser (sandy) sediments are most prominent near sill points, although this J Paleolimnol (2011) 46:273–289 Fig. 5 Common fluvial and lacustrine sponge spicules encountered in lake-bottom sediments from the study lakes. A = gemmosclere of the lotic sponge Corvospongilla secktii. B = gemmosclere of the lotic sponge Oncosclera navicella. C = gemmosclere of the lentic sponge Radiospongilla amazonensis. 283 D = gemmosclere of the lentic sponge Trochospongilla variabilis. E = gemmosclere of the lotic sponge Uruguaya corallioides. F = megasclere of the lentic sponge Heteromeyenia sp. G = megasclere of the lentic sponge Corvoheteromeyenia sp. I = megasclere of the lentic sponge Metania spinata 123 284 relation is complicated in BV given its low-relief eastern margin, which likely receives inflow from the PR at several points. Particle sizes tend to be finest in deep-water regions, which in most cases are located proximal to margins with coincident topography; suspension settling and organic sedimentation are common in these locales. Although sub-surface datasets from the lakes are unavailable, the correlation of bathymetric lows and high-relief rocky margins suggests the potential for neotectonic controls on accommodation space in the basins (Fig. 1). In the Pantanal, structural influences on fluvial channel patterns are not uncommon and shallow earthquakes have been recorded in the region (Assine and Soares 2004). Por (1995) suggested that karst plays an intimate role in large lake formation in the Pantanal, but our data suggest that this is not likely for LG, LM and BV. Hydrochemical datasets show that the lakes are extremely dilute, whereas karst basin waters are typically rich in divalent metals due to dissolution of carbonate aquifers (Electronic Supplementary Material Table 2). During sampling, the lake waters broadly reflected the chemical composition of the PR, especially for conservative cations; these waters are undersaturated with respect to calcite (Hamilton et al. 1997). We interpret a three-phase seasonal evolution that links lacustrine hydrochemistry, morphometrics and sedimentology to stage of the PR, which is strongly modulated by regional climate (Electronic Supplementary Material Fig. 3). Seasonal evolution of the large lakes of the Pantanal shares a number of similarities with lakes on the floodplain of the Orinoco River in Venezuela described by Hamilton and Lewis (1987). Channelized linkages to the PR may serve as conduits for reversing flows following the passage of the flood pulse, during a transient ‘‘draining-phase’’ leading to full isolation of the lakes from the river (Electronic Supplementary Material Fig. 3). In the central Pantanal, the stage height of the PR drops by C5 m in the austral summer, when the flood pulse is retained in the north and evaporation increases due to elevated air temperatures. Theoretical estimates of lake level decline due to evaporation alone indicate lake surfaces drop by at least 0.6 m annually. At peak inundation, lake-surface altitudes for LG, LM and BV are typically 5, 4, and 6 m below their sill heights, respectively. As a result, short-lived reversed flows are probable for LG and LM, and 123 J Paleolimnol (2011) 46:273–289 possible, but likely rare, for BV. Complete isolation of the lakes generally occurs in the austral summer, suggesting direct rainfall and groundwater inputs are less important to morphometrics and bathymetry than the PR flood pulse. Both fluvial and lacustrine processes control organic facies development in the large lakes of the Pantanal. River-borne solutes probably provide a key source of nutrients that help control algal productivity in LG, LM and BV. Junk et al. (1989) were among the first to recognize the importance of flood pulse dynamics for nutrient delivery and productivity in the tropics, whereas Hamilton et al. (1997) noted that the floodplains of the central Pantanal are sinks for flood pulse-derived OM and sites of inorganic nutrient transformation. Concentrations of N, P and suspended solids, however, are lower in the PR than in other rivers draining tropical savannas (Hamilton et al. 1997). As a consequence, local allochthonous sources of nutrients may likewise be important. Macrophyte decomposition and internal recycling are two possible local sources of phytoplanktonlimiting nutrients. Floating macrophytes and marsh vegetation are commonly encountered along the margins of each of the lakes, and dry-season decomposition of such vegetation is a known source of dissolved nutrients in other floodplain lakes (Furch et al. 1983). Likewise, mixing-driven re-suspension of lake bottom sediments can lead to the release of nutrients, but this process also prolongs the exposure of labile organic compounds to degradation by aerobic respiration (Forsberg et al. 1988; Henrichs 1992). Ultimately, OM-rich facies development across broad regions of the lakes is mediated by dilution during ‘‘fill-phase’’ sedimentation. Biogeochemical signals in LM samples are complex and appear to reflect elevated productivity and subsequent degradation (Fig. 2). Organic sediments from LM are relatively rich in 13C (Fig. 3). The factors that can influence d13COM values include: (1) the isotopic composition of the carbon source (commonly lake water DIC); (2) phytoplankton assemblage and metabolic rates; (3) the proportion of C3 and C4 terrestrial vegetation within particulate organic matter; and (4) diagenesis (Meyers and Teranes 2001; Talbot and Johannessen 1992). At present, d13CDIC measurements on lake or river water samples are unavailable, but we assume that the concentration of aqueous CO2 is relatively low due to LM’s pH (8.3). J Paleolimnol (2011) 46:273–289 Therefore, utilization of HCO3- during photosynthesis may partially contribute to the enriched carbon isotopic signature observed in these samples (Fig. 3). Hamilton et al. (1997) noted that C/N values for particulate organic matter (POM) in the PR averages 8.5, but that algae comprised only a small portion of this OM. Wantzen et al. (2002) reported a range of -26.0 to -32.0% for d13C values in POM in the northern Pantanal. These values are consistent with geochemical characteristics of both C3 terrestrial plants and algae. Smear slides show that LM sediments contain abundant green algae and diatoms, with cyanobacteria, including Anabaena sp., common in some samples (Electronic Supplementary Material Fig. 2). Algae blooms were observed during our sampling and the relatively high d13C values can at least be partially accounted for by the preferential uptake of 12C during photosynthesis, thus depleting the residual DIC reservoir in the lake (McKenzie 1985; Amorim et al. 2009). Given its long fetch (*22 km), wind-driven internal nutrient fertilization provides a compelling explanation for algal blooms and elevated carbon burial in LM. Intriguingly, samples with anomalously high d13C values generally exhibit nitrogen isotope ratios below 2.0% (Fig. 3). The presence of nitrogen-fixing cyanobacteria in these samples provides one viable explanation for these d15N values, as well as the high TN concentrations (Talbot 2001). However, HI values from LM are below 400 mg HC/gm TOC, suggesting that cyanobacteria do not dominate the phytoplankton assemblage and local allochthonous sources may also contribute to the lake’s carbon cycle (Talbot and Livingstone 1989). Importantly, oxygen index values exceeding 200 mg CO2/gm TOC indicates that diagenesis also affects the geochemistry of OM in LM. Re-suspension and oxidation of buried sediments can alter primary geochemical signals, due to selective loss of N and 12C (in the latter, due to oxidative fractionation that may increase d13COM values by 1–2%). Therefore, elevated C/N and d13COM may in part reflect the influence of oxidation in some samples. Biogeochemical data from LG and BV lack the variability of samples from LM and suggest lower primary productivity. Sediment OM samples from LG and BV have low d13C values and C/N values \20, commonly thought to reflect admixtures of lacustrine algae and vascular plants (Meyers and Ishiwatari 1993). Values of lake-water pH suggest that dissolved 285 CO2 in isotopic equilibrium with the atmosphere is likely the DIC source for primary producers in these lakes. Hydrogen index values below 300 mg HC/gm TOC for both LG and BV support the interpretation of a mixed organic provenance. Examination of smear slides confirms the presence of algae and terrestrial plant fragments in both lakes (Electronic Supplementary Material Fig. 2). In general, the nitrogen isotope composition of OM from LG and BV is consistent with d15N values associated with lake sediments composed of both algae, which commonly preserve the isotopic signature of lake-water dissolved inorganic nitrogen, and terrestrial vegetation, which utilize atmospheric N2. The data are also consistent with common aquatic macrophytes and flood-pulsederived POM in the Pantanal. Several BV samples, collected west of the headland, exhibit high d13C values and low d15N values (\2.6%); lake-margin marshes, dominated by C4-pathway Paspalum repens likely contribute to the isotopic composition of these samples (Fellerhoff et al. 2003). Rock–Eval data suggest that post-depositional alteration of OM is less pronounced in LG and BV when compared with LM, likely the result of differences in wave development and sediment re-suspension. Figure 6 presents a conceptual model of sedimentation derived from the observed modern seasonal sedimentation cycle and its linkages to the PR flood pulse. Data described in this study provide constraints on the processes that influence modern lithofacies patterns and geochemical baselines. With caution, these inferences can be used to predict facies migrations during intervals of major environmental change, such as those that might be preserved in sediment core records. Evidence of meaningful alteration of climate is already well established by relict landforms (dunes, lunettes, and deflation pans) in the southern Pantanal (Soares et al. 2003). Thus, over short intervals of time, climate change in the PR headwaters is clearly an important mechanism that could lead to significant change in sedimentation in LG, LM and BV. Climatic conditions that promote a stronger flood pulse and deeper lakes, such as sustained southerly advance of the ITCZ, are likely to: (1) create additional fluvial entry points into the lakes, leading to ephemeral marginal fans; (2) backflood existing incipient river channels, creating highstand deltas; (3) increase primary productivity and algal OM deposition due to enhanced riverine 123 286 J Paleolimnol (2011) 46:273–289 Fig. 6 Conceptual model of sedimentation in the floodplain lakes of Pantanal, based on modern inferences. Solid line around lake perimeter in B indicates modern shoreline, and broken lines in A and C denote position of modern shoreline relative to hypothetical lake highstands and lowstands. See text for details nutrient supply; (4) enhance the accumulation of sponge spicules from lotic species assemblages; and (5) develop fine-grained organic facies where suspension fall-out processes dominate, possibly with higher temporal continuity due to increased water depths. In contrast, sustained lake level lowstands associated with reduced effective precipitation in the PR headwaters will serve to magnify ‘‘isolationphase’’ processes and sedimentation, including: (1) incision of incipient river channels, with the potential for lowstand fan deposition; (2) sporadic deposition of lotic sponges; (3) possible precipitation of reduced metals and carbonate as isolated lake waters concentrate and mix; (4) reworking of sub-aerially exposed lake-bottom deposits, through deflation or bioturbation; (5) greater deposition of terrestrial OM from the local watershed; and (6) oxidation of labile biogenic sediments and diagenetic evolution of elemental and stable isotopic values on OM (Fig. 6). A space-fortime substitution with large lakes in the southern Pantanal, e.g. Lagoa Cáceres (Fig. 1), provides insights into chemical sedimentation during lowstands. This region experiences lower yearly precipitation and delayed passage of the flood pulse; minor calcite precipitation occurs in these lakes today (Electronic Supplementary Material Table 5). Clearly, changes in sediment accumulation rates can be expected as major hydrologic thresholds are crossed, especially during lowstands when sedimentation is more discontinuous. The processes described herein are likely common to many floodplain lake systems, often resulting in low-to-intermediate-resolution records with highly variable sedimentation rates (Cohen 2003). Importantly, neotectonic alterations of drainage patterns are an alternative mechanism that could substantially alter the PR and its connection to floodplain lakes in the Pantanal. For example, faulting of the shallow crust has been implicated in the capture and redirection of large rivers in tropical Africa, which in some cases have influenced sedimentation processes in wetlands and lakes (Gumbricht et al. 2001; McGlue et al. 2006). Similar tectonic processes could affect the position of the PR, cutting the lakes off from the flood pulse and leading to ‘‘isolation-phase’’ sedimentation. Likewise, crevasse splay development and channel migrations have been documented in the central Pantanal over the past several decades (Assine 2005). Avulsion of the PR could spur meaningful changes in the magnitude of the flood pulse on LG, LM and BV without coeval climate change. Independent constraints on fluvial geomorphic evolution are therefore necessary for interpreting paleoenvironments in the Pantanal and similar wetlands. 123 Conclusions Owing to their role in biogeochemical cycles, biodiversity and as a water resource, understanding the J Paleolimnol (2011) 46:273–289 response of large tropical wetlands to global change is critically important. A potentially rich paleolimnological archive of Quaternary environmental dynamics for tropical wetlands exists in the sediments of floodplain lakes. Interpreting these records effectively, however, requires a highly nuanced understanding of how these lakes act as depositional basins. The results of this actualistic assessment yield new insights into these limnogeological processes for three large floodplain lakes in the Brazilian Pantanal. 1. 2. 3. Situated along the western margin of the PR and strongly influenced by its flood pulse, LG, LM and BV are freshwater, well-oxygenated, weakly basic, dilute limnological systems with elongate shorelines and shallow maximum water depths (\5 m). Correlation of bathymetric maxima and coincident margins suggests a neotectonic control on lake formation, likely modulated by channel migrations of the PR. The flood pulse of the PR impacts the limnogeology of each of the study lakes through delivery of terrigenous grains, lotic sponge spicules, POM, nutrients and through modulation of lake shape, depth, and water chemistry. Sediment biogeochemistry (BiSi, TN and TOC) suggests that production and burial of biomass is highest in LM and somewhat lower in LG and BV. Stable isotopes of carbon and nitrogen and HI values indicate a mixed (lacustrine algae, macrophytes and vascular plants) organic matter provenance for the lakes. Post-depositional resuspension of lake-bottom sediments by waves and bioturbation influence geochemical signals and depositional continuity in these shallow basins. Sites of uninterrupted sedimentation were only encountered in the deepest regions of LG, where the decay of short-lived radioisotopes indicates that sedimentation rates ranged between 0.11 and 0.24 cm year-1. Sensitivity of the PR flood pulse to climate change, neotectonics and fluvial autocyclicity has important implications for floodplain limnogeology. Our conceptual depositional model suggests that intensification of the PR flood pulse is likely to be represented by higher lake levels and a concomitant decrease in grain size, an increase in the burial of aquatic OM due to elevated productivity, depositional continuity, and lotic 287 sponge assemblages. Weakening of the flood pulse promotes chemical and sandy lowstand fan sedimentation, terrestrial OM deposition, oxidation and bioturbation as the lakes become isolated and contract on the floodplain. Actualistic datasets and space-for-time considerations developed herein provide a predictive framework for facies migrations that can be captured and dated in sediment cores from these or similar lakes with linkages to large tropical rivers with sustained seasonal flooding. Acknowledgments The title for this contribution was adapted from: The Pantanal—Brazil’s Forgotten Wilderness by Vic Banks (1991). The research presented in this paper was supported by the American Chemical Society (PRF program grant 45910-AC8), the São Paulo Research Foundation (FAPESP grant 2007/55987-3) and the UA-Exxon Mobil COSA project. Generous grants to MM from Laccore, the Chevron Corporation, Kartchner Caverns, and PAGES assisted the completion of this project. Research in Brazil would not have been possible without logistical and scientific support from ECOA-Brazil, EMBRAPA-Brazil, UFMS—Campus do Pantanal, the Fazenda Santa Teresa and the citizens of Amolar (MS—Brazil). We are extremely grateful to K. Wendt, F. dos Santos Gradella, S. Kuerten, B. Lima de Paula, D. Calheiros, R. Lins, A. Lins, and T. Matsushima for their assistance. C. Gans and E. Guerra de Lima provided vital support at UA during our 2009 field season. L. Helfrich, C. Landowski, X. Zhang, C. Eastoe, J. Ash and the staff of the Limnological Research Center at the University of Minnesota provided support in the laboratory. Critical reviews by S. Harris, C. Turner, S. Ivory, M. Blome, M. Brenner and two anonymous reviewers substantially improved the quality of the text. References Ab’Saber AN (1988) O Pantanal Mato-Grossense e a teoria dos refúgios. Revista Brasileira de Geografia 50 especial, pp 9–57 Adams KD, Wesnousky SG (1998) Shoreline processes and age of the Lake Lahontan high-stand in the Jessup embayment, Nevada. Geol Soc Am Bull 110:1318–1332 Alfonsi RR, Paes de Camargo MB (1986) Condições Climáticas para a Região do Pantanal Mato-Grossense. In: Boock A (ed) Anais do 1 Simpósio sobre Recursos Naturais e Sócio-Econômico do Pantanal. EMBRAPA-CPAP, vol 5, pp 29–42 Alho CJR, Lacher TE, Goncalves HC (1988) Environmental degradation in the Pantanal ecosystem. Bioscience 38:164–171 Amorim MA, Turcq PFM, Turcq BJ, Cordeiro RC (2009) Origem e dinâmica da deposição dos sedimentos superficiais na Várzea do Lago Grande de Curuai, Pará, Brasil. Acta Amazônica 39:165–172 123 288 Assine ML (2005) River avulsions on the Taquari megafan, Pantanal wetland, Brazil. Geomorphology 70:357–371 Assine ML, Silva A (2009) Contrasting fluvial styles of the Paraguay River in the northwestern border of the Pantanal wetland, Brazil. Geomorphology 113:189–199 Assine ML, Soares PC (2004) Quaternary of the Pantanal, west-central Brazil. Quat Int 114:23–34 Banks V (1991) The Pantanal: Brazil’s forgotten wilderness. Sierra Club Books, San Francisco Barbiéro L, de Querioz Neto JP, Ciomai G, Sakamoto AY, Capellari B, Fernandes E, Valles V (2002) Geochemistry of water and ground water in the Nhecolandia, Pantanal of Mato Grosso, Brazil: variability and associated processes. Wetlands 22:528–540 Bartlett KB, Harriss RC (1993) Review and assessment of methane emissions from wetlands. Chemosphere 26: 261–320 Bates BC, Kundzewicz ZW, Wu S, Palutikof JP (eds) (2008) Climate change and water. Technical paper of the intergovernmental panel on climate change, IPCC Secretariat, Geneva, 210 pp Batista TCA, Volkmer-Ribeiro C, Darwich C, Alves LF (2003) Freshwater sponges as indicators of floodplain lake environments and of river rocky bottom in Central Amazônia. Amazoniana 18:525–549 Cao M, Gregson K, Marshall S (1998) Global methane emission from wetlands and its sensitivity to climate change. Atmos Environ 32:3293–3299 Chase CG, Sussman AJ, Coblentz DD (2010) Curved Andes: Geoid, forebulge, and flexure. Lithosphere 1:358–363 Chauhan M, Gopal B (2001) Biodiversity and management of Keoladeo National Park (India): a wetland of international importance. Biodiversity in wetlands: assessment, function and conservation, vol 2. Backhuys Publishers, Leiden Cohen AS (2003) Paleolimnology: the history and evolution of lake systems. Oxford University Press, Oxford Cole MM (1960) Cerrado, caatinga and Pantanal: the distribution and origin of the savanna vegetation of Brazil. Geogr J 126:168–179 de Magalhães NW (1992) Conheca o Pantanal. Terragraph Artes e Informática São Paulo de Oliveira MD, Calheiros DF (2000) Flood pulse influence on phytoplankton communities of the southern Pantanal floodplain, Brazil. Hydrobiologia 427:101–112 de Oliveira Bezerra MA, Mozeto AA (2008) Deposição de carbono orgânico na planı́cie de inundação do Rio Paraguai durante o Holoceno médio. Oecol Bras 12:155–171 DeMaster DJ (1979) The marine budgets of silica and Si32. PhD Thesis, Yale University, New Haven, 308 pp Donders TH, Wagner F, Dilcher DL, Visscher H (2005) Midto late-Holocene El Niño-Southern Oscillation dynamics reflected in the subtropical terrestrial realm. Proc Natl Acad Sci USA 102:10904–10908 Espitalie J, Laporte JL, Madec M, Marquis F, Leplat G, Paulet J, Boutefeu A (1977) Methode rapide de caracterisation des roches meres, de leur potential petrolier et de leur degre d’evolution. Rev Inst Fr Pétrol 32:23–42 Fellerhoff C, Voss M, Wantzen KM (2003) Stable carbon and nitrogen isotope signatures of decomposing tropical macrophytes. Aquat Ecol 37:361–375 123 J Paleolimnol (2011) 46:273–289 Forsberg BR, Devol AH, Richey JE, Martinelli LA, Dos Santos H (1988) Factors controlling nutrient concentrations in Amazon floodplain lakes. Limnol Oceanogr 33:41–56 Furch K, Junk WJ, Dieterich J, Kochert N (1983) Seasonal variation in the major cation (Na, K, Mg, and Ca) content of the water of Lago Camaleao, an Amazonian floodplain lake near Manaus, Brazil. Amazoniana 8:75–89 Gumbricht T, McCarthy TS, Merry CL (2001) The topography of the Okavango Delta, Botswana, and its tectonic and sedimentological implications. S Afr J Geol 104:243–264 Hamilton SK (1999) Potential effects of a major navigation project (Paraguay-Parana Hidrovia) on inundation in the Pantanal floodplains. Regul River 15:289–299 Hamilton SK, Lewis WM (1987) Causes of seasonality in the chemistry of a lake on the Orinoco River floodplain, Venezuela. Limnol Oceanogr 32:1277–1290 Hamilton SK, Sippel SJ, Calheiros DF, Melack JM (1997) An anoxic event and other biogeochemical effects of the Pantanal wetland on the Paraguay River. Limnol Oceanogr 42:257–272 Hamilton SK, Sippel SJ, Melack JM (2002) Comparison of inundation patterns among major South American floodplains. J Geophys Res 107(D20):8038 Heckman CW (1998) The Pantanal of Poconé. Kluwer, Den Haag Henrichs SM (1992) Early diagenesis of organic matter in marine sediments: progress and perplexity. Mar Chem 39:119–149 Horton BK, DeCelles PG (1997) The modern foreland basin system adjacent to the central Andes. Geology 25: 895–898 Johnson TC, McCave IN (2008) Transport mechanism and paleoclimatic significance of terrigenous silt deposited in varved sediments of an African rift lake. Limnol Oceanogr 53:1622–1632 Junk WJ, Bayley PB, Sparks RE (1989) The flood pulse concept in river-floodplain systems. Can Spec Publ Fish Aquat Sci 106:110–127 Junk WJ, Nunes Da Cunha K, Wantzen KM, Petermann P, Strussmann C, Marques MI, Adis J (2006) Biodiversity and its conservation in the Pantanal of Mato Grosso, Brazil. Aquat Sci 68:278–309 Lopes I, Minõ C, Del Lama S (2007) Genetic diversity and evidence of recent demographic expansion in waterbird populations from the Brazilian Pantanal. J Biol 67: 849–857 Maltby E, Turner RE (1983) Wetlands of the world. Geog Mag 55:12–17 Mayle FE, Beerling DJ, Gosling WD, Bush MB (2004) Responses of Amazonian ecosystems to climatic and atmospheric carbon dioxide changes since the last glacial maximum. Philos Trans R Soc Lond B Biol Sci 359: 499–514 McGlue MM, Scholz CA, Karp T, Ongodia B, Lezzar K-E (2006) Facies architecture of flexural margin lowstand delta deposits in Lake Edward, East African rift: constraints from seismic reflection imaging. J Sed Res 76: 942–958 McKenzie JA (1985) Carbon isotopes and productivity in the lacustrine and marine environment. In: Stumm W (ed) J Paleolimnol (2011) 46:273–289 Chemical processes in lakes. Wiley, New York, pp 99–118 Meyers PA, Ishiwatari R (1993) Lacustrine organic geochemistry: an overview of indicators of organic matter sources and diagenesis in lake sediments. Org Geochem 20: 867–900 Meyers PA, Teranes JL (2001) Sediment organic matter. In: Last WM, Smol JP (eds) Tracking environmental change using lake sediments volume 2: physical and geochemical methods, vol 2. Springer, New York, pp 239–269 Mitsch WJ, Gosselink JG (2000) Wetlands, 3rd edn. Wiley, New York Mitsch WJ, Nahlik A, Wolski P, Bernal B, Zhang L, Ramberg L (2009) Tropical wetlands: seasonal hydrologic pulsing, carbon sequestration, and methane emissions. Wetl Ecol Manag 18:573–586. doi:10.1007/s11273-009-9164-4 Parolin M, Volkmer-Ribeiro C, Stevaux JC (2007) Sponge spicules in peaty sediments as paleoenvironmental indicators of the Holocene in the upper Parana River, Brazil. Rev Bras Paleo 10:17–26 Pinder L, Rosso S (1998) Classification and ordination of plant formations in the Pantanal of Brazil. Plant Ecol 136:151–165 Por FD (1995) The Pantanal of Mato Grosso (Brazil). Kluwer, Dordrecht Prance GT, Schaller GB (1982) Preliminary study of some vegetation types of the Pantanal, Mato Grosso, Brazil. Brittonia 34:228–251 Sharma P, Gardner LR, Moore WS, Bollinger MS (1987) Sedimentation and bioturbation in a salt marsh estuary as revealed by 210Pb, 137Cs, and 7Be studies. Limnol Oceanogr 32:313–326 Soares AP, Soares PC, Assine ML (2003) Areiais e lagoas do Pantanal, Brasil: herança paleoclimática? Rev Bras Geo 33:211–224 Swarzenski PW, Baskaran M, Orem WH, Rosenbauer RJ (2006) Historical reconstruction of contaminant inputs in 289 Mississippi River delta sediments. Estuar Coast Mar Sci 29:1094–1107 Talbot MR (2001) Nitrogen isotopes in palaeolimnology. In: Last WM, Smol JP (eds) Tracking environmental change using lake sediments: physical and geochemical methods, vol 2. Springer, New York, pp 401–439 Talbot MR, Johannessen T (1992) A high resolution palaeoclimatic record for the last 27, 500 years in tropical West Africa from the carbon and nitrogen isotopic composition of lacustrine organic matter. Earth Planet Sci Lett 110: 23–37 Talbot MR, Livingstone DA (1989) Hydrogen index and carbon isotopes of lacustrine organic-matter as lake level indicators. Palaeogeogr Palaeoclimatol Palaeoecol 70: 121–137 Tricart J (1982) El Pantanal: Un ejemplo del impacto de la geomorfologia sobre el medio ambiente. Geografia 7: 37–50 Ussami N, Shiraiwa S, Dominguez JML (1999) Basement reactivation in a sub-Andean foreland flexural bulge: the Pantanal wetland, SW Brazil. Tectonics 18:25–39 Veloso HP (1972) Aspectos fito-ecológicos da Bacia do Alto Rio Paraguai. Instituto de Geografia São Paulo (USP), Brazil Volkmer-Ribeiro C (1992) The freshwater sponges in some peatbog ponds in Brazil. Amazoniana 12:317–335 Volkmer-Ribeiro C, Pauls SM (2000) Esponjas de água dulce (Porifera, Demospongiae) de Venezuela. Acta Biol Venez 20:1–28 Volkmer-Ribeiro C, Turcq B (1996) SEM analysis of siliceous spicules of a freshwater sponge indicate paleoenvironmental changes. Acta Microscópica 5:186–187 Wantzen KM, Machado FD, Voss M, Boriss H, Junk WJ (2002) Seasonal isotopic shifts in fish of the Pantanal wetland, Brazil. Aquat Sci 64:239–251 123 J Paleolimnol DOI 10.1007/s10933-011-9545-6 ERRATUM Erratum to: Limnogeology in Brazil’s ‘‘forgotten wilderness’’: a synthesis from the large floodplain lakes of the Pantanal Michael M. McGlue • Aguinaldo Silva • Fabrı́cio A. Corradini • Hiran Zani Mark A. Trees • Geoffrey S. Ellis • Mauro Parolin • Peter W. Swarzenski • Andrew S. Cohen • Mario L. Assine • ! Springer Science+Business Media B.V. 2011 Erratum to: J Paleolimnol (2011) 46:273–289 DOI 10.1007/s10933-011-9538-5 The caption for Fig. 5 is incorrect. It should read: Fig. 5 Common fluvial and lacustrine sponge spicules encountered in lake-bottom sediments from the study lakes. A = gemmosclere of the lotic sponge Corvospongilla secktii. B = gemmosclere of the lotic sponge Oncosclera navicella. C = gemmosclere of the lentic sponge Radiospongilla amazonensis. The online version of the original article can be found under doi:10.1007/s10933-011-9538-5. M. M. McGlue (&) ! M. A. Trees ! A. S. Cohen Department of Geosciences, The University of Arizona, 1040 East 4th Street, Tucson, AZ 85721, USA e-mail: mmmcglue@email.arizona.edu G. S. Ellis Energy Resources Program, U.S. Geological Survey, Denver, CO, USA e-mail: gsellis@usgs.gov M. A. Trees e-mail: treesma@email.arizona.edu M. Parolin Faculdade Estadual de Ciências e Letras de Campo Mourão, Av. Comendador Norberto Marcondes, 733, Campo Mourão, PR 87303-100, Brazil e-mail: mauroparolin@gmail.com A. S. Cohen e-mail: cohen@email.arizona.edu A. Silva Departamento de Ciências do Ambiente, Universidade Federal de Mato Grosso do Sul—UFMS-CPAN, Av. Rio Branco, 1270, Corumbá, MS 79304-902, Brazil e-mail: aguinald_silva@yahoo.com.br F. A. Corradini Faculdade de Geografia, Universidade Federal do ParáUFPA, Folha 31, Quadra 7, Lote Especial S/N, Marabá, PA 68501-970, Brazil e-mail: f_coradini@yahoo.com.br P. W. Swarzenski U.S. Geological Survey, Santa Cruz, CA, USA e-mail: pswarzen@usgs.gov M. L. Assine Departamento de Geologia Aplicada—IGCE, Universidade Estadual Paulista—UNESP/Campus Rio Claro, Av. 24-A, 1515, Rio Claro, SP 13506-900, Brazil e-mail: assine@rc.unesp.br H. Zani Divisão de Sensoriamento Remoto, Instituto Nacional de Pesquisas Espaciais—INPE, Av. dos Astronautas, 1758, Saõ José dos Campos, SP 12201-970, Brazil e-mail: hzani@dsr.inpe.br 123 J Paleolimnol D = gemmosclere of the lentic sponge Trochospongilla variabilis. E = gemmosclere of the lotic sponge Uruguaya corallioides. F = gemmosclere of the 123 lentic sponge Heteromeyenia sp. G = gemmosclere of the lentic sponge Corvoheteromeyenia sp. H = megasclere of the lentic sponge Metania spinata