Vous êtes sur la page 1sur 13

SEDGEO-04926; No of Pages 13

Sedimentary Geology xxx (2015) xxx–xxx

Contents lists available at ScienceDirect

Sedimentary Geology

journal homepage: www.elsevier.com/locate/sedgeo

Sediment generation in humid Mediterranean setting: Grain-size and


source-rock control on sediment geochemistry and mineralogy
(Sila Massif, Calabria)
Hilmar von Eynatten a,⁎, Raimon Tolosana-Delgado b, Volker Karius a, Kai Bachmann b, Luca Caracciolo c
a
Geowissenschaftliches Zentrum der Georg-August-Universität Göttingen, Abteilung Sedimentologie/Umweltgeologie, Göttingen, Germany
b
Helmholtz Institute Freiberg for Resource Technology, Freiberg, Germany
c
Chemostrat Ltd., Welshpool, UK

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

Article history: Grain-size control on sediment composition is investigated in modern proximal sediment from the Sila Massif,
Received 3 July 2015 where basic to felsic intrusive rocks are exposed in a Mediterranean humid–temperate upland climate. Samples
Received in revised form 15 October 2015 were taken from small creeks and weathering profiles from three areas reflecting different bed rock compositions.
Accepted 19 October 2015
Samples were separated into eleven grain size fractions from very coarse sand to clay and analyzed by (i) X-ray
Available online xxxx
fluorescence for chemical composition and (ii) X-ray diffraction and Mineral Liberation Analysis for mineralogical
Keywords:
composition. The chemical composition vs. grain size relations were modeled by compositional linear regression.
Geochemistry Mineralogical composition of selected samples is used to substantiate the interpretations based on geochemistry.
Mineralogy Results: reveal a high degree of chemical weathering with chemical index of alteration (CIA) up to 92. High CIA
Chemical weathering values are restricted to the fine-grained fractions, while sand-sized sediment average at low to moderate CIA values
Comminution (~60). Although strongly weathered, the three sample suites reflecting basic to felsic plutonic bed rock can be ef-
Provenance fectively discriminated across all grain-size classes using trace elements such as V, Rb, and Sr. Linear trend model-
Compositional linear regression ing and mineralogical data reflect similar patterns for all sample suites implying similar processes independent of
source rock composition. This includes overall decrease of quartz and K-feldspar over the full grain-size range from
very coarse sand to clay, which is contrasted by overall increase of sheet silicates from coarse to fine. Among the
latter, increase of clay minerals strongly outpaces the increase of micas in silt to clay fractions. A more complex
behavior is shown by plagioclase, which is most abundant in intermediate grain-size fractions for all sample suites.
This is likely caused by initial hydrolysis along cleavage and twinning planes and subsequent breakage of plagioclase
crystals into smaller fragments. Towards finer grain size, intense hydrolysis has destroyed most feldspars.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction as specific methodological constraints (e.g., microscopic determination


of framework grains or heavy minerals is restricted to certain grain-
The chemical and mineralogical compositions of sediments and sed- size ranges). Therefore, classic approaches to infer geological conditions
imentary rocks are strongly controlled by grain size (e.g. Blatt et al., from sediment composition typically rely on data obtained from a given
1972; Basu, 1976). Other factors include source rock composition, narrow grain size range (e.g. Basu et al., 1975; Ingersoll et al., 1984;
weathering, and physical as well as chemical processes that modify the Morton and Hallsworth, 1994). The shortcomings of such approaches
sediment while being transported from source to sink (e.g. Johnsson, have been recently outlined by, for instance, Garzanti et al. (2009).
1993). The effect of grain size is generally thought to exceed the impact Therefore, comparing data obtained from different grain size fractions
of all other control factors on the variability of sediment composition or from sediment samples with contrasting grain size distributions inev-
(e.g. Garzanti et al., 2011; Bloemsma et al., 2012). Nevertheless, many itably requires an understanding of the compositional relations across
studies inferring geologic or paleoclimatic conditions from sediment grain size grades. If these relations are controlled by purely physical pro-
characteristics do not consider grain size influence adequately. cesses such as hydrodynamics or mechanical comminution, relatively
The high relevance of grain-size for analyzing sediment composition simple empirical or numerical models can be used to correct for the
is contrasted by the fact that analyzing the entire grain-size spectrum is grain-size dependent control on composition (e.g. Garzanti et al., 2009;
often impossible, due to limitations of the accessible rock record as well von Eynatten et al., 2012). For complex interactions of chemical and
physical processes, which must be considered the rule rather than the
⁎ Corresponding author. exception in natural systems, Bloemsma et al. (2012) developed a statis-
E-mail address: hilmar.von.eynatten@geo.uni-goettingen.de (H. von Eynatten). tical method to separate the grain-size dependent part of compositional

http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
0037-0738/© 2015 Elsevier B.V. All rights reserved.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
2 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

variation from a residual part, which reflects other controlling factors 1994). Late Pleistocene to Holocene catchment-averaged erosion rates
such as provenance or diagenesis. Application of this tool revealed that for the flat uplands are low, scattering around 0.1 mm/a as inferred
trends of grain size vs. composition strongly depend on the specific geo- from cosmogenic nuclide data (Olivetti et al., 2012).
logical setting, implying that more empirical studies and models are The samples were collected in June 2010 (RT3) and September 2013
needed to better understand individual settings. (RT4). They derive from three different areas of the Sila Massif, which
The aim of the paper is to describe and quantify the effects of grain were selected to represent the diverse range of the Sila Batholith from
size and source-rock lithology on geochemical and mineralogical felsic to basic plutonic rocks. From north to south, the first area is locat-
compositions of sediments in a specific case study. The samples were ed to the northeast of Lago Cecita and exposes felsic rocks of granodio-
taken from three areas of the humid–temperate Sila Massif, southern rite to monzogranite composition (i.e. cordierite-bearing biotite–
Italy, composed of contrasting basic to felsic plutonic source rocks. A muscovite granodiorite to monzogranite according to Messina et al.,
broad grain size spectrum covering more than three orders of magni- 1991; peraluminous granite according to Graessner et al., 2000). These
tude from very coarse sand to clay is considered. Previous studies rocks range in SiO2 content from 68 to 74 wt.% (Ayuso et al., 1994)
have documented significant chemical weathering of soils and sediment and constitute the most felsic sample suite, termed F2 (Fig. 1). The sec-
in the Sila Massif (e.g. Scarciglia et al., 2007). In a first step, weathering ond area is located around Carlo Magno and Silvana Mansio north of
and provenance effects on sediment composition, as well as their varia- Lago Arvo. It belongs to the widely exposed hornblende and biotite-
tions across the grain size grades from coarse to fine are evaluated. In a bearing tonalites to granodiorites, which range in SiO2 content from
second step, the results are compared to a previous study (von Eynatten 60 to 71 wt.% (Messina et al., 1991; Ayuso et al., 1994). The selected
et al., 2012) from a highly contrasting climatic setting where chemical area constitutes the felsic to intermediate sample suite, termed F1
weathering is negligible and thus mechanical comminution controls (Fig. 1). Some bedrock samples of this area indicate a dominant compo-
sediment composition across grain-size grades. sition of F1 bedrocks in the less felsic part of the tonalites to granodio-
The study is essentially based on geochemical data because they can rites with SiO2 content around 61 to 62%. The third area is located
be obtained easily and precisely for the full grain size spectrum from directly northeast of Lago Arvo around the village of Rovale and exposes
very coarse sand to clay, which covers N 95% of the global clastic sedi- basic to intermediate rocks of the so-called Rovale Zone or Complex.
ment budget given that conglomerates and breccias comprise only a This small area (~15 km2) consists of gabbroic rocks (norites, amphibole
few percent of the total clastic sediment record (e.g. Pettijohn, 1975). gabbros; 42–49 wt.% SiO2, 18–26 wt.% Al2O3) in its western part close to
Mineralogical composition is obtained by X-ray diffraction for selected Rovale, while in its eastern part plagioclase and amphibole-rich diorites
samples and grain-size fractions to verify interpretations based on geo- and tonalites prevail (52–57 wt.% SiO2, 17–18 wt.% Al2O3), with small
chemical data. Moreover, Mineral Liberation Analysis (MLA) has been local gabbroic bodies (Caggianelli et al., 1994). This area constitutes
tested on some samples because it potentially provides a tool to analyze the basic to intermediate sampling area of the study, termed B (Fig. 1).
mineral composition over a large grain size range. Two types of samples have been collected: modern sediment from
small proximal creeks (Sed, N = 10) as well as material from
2. Study area and sampling strategy weathering profiles or grus exposed in roadcuts (WP, N = 6). Small
creek means channel width less than 2 m. Sampling was intended to
The study area is located in the uplands of the Sila Massif in northern cover a broad grain-size spectrum from very coarse sand to clay from
Calabria (southern Italy) at altitudes between approx. 1150 and each sampling site. Therefore, Sed-samples may represent a mixture
1600 masl, characterized by a smooth topography with gentle hills of up to three subsamples taken from a small area of no more than
and slightly incised streams in between. It has a typical Mediterranean approx. 20 m2. Grus samples were taken from the debris cones devel-
humid–temperate upland climate, characterized by mean annual pre- oped along roadcuts through deeply disintegrated granitoid bed rock.
cipitation of 1400 to 1800 mm and mean annual temperature of While sample suites F1 and F2 comprise both Sed and WP samples
10–12 °C, with mean monthly temperatures ranging between − 1 °C (altogether seven samples for each suite), only two sediment samples
and 18 °C (cf. Scarciglia et al., 2012). The climatic and geomorphological are available from sampling suite B (Table 1).
conditions have caused deep and intense weathering of the respective
bed rocks, which has been investigated in numerous weathering pro- 3. Methods
files (e.g. Le Pera and Sorriso-Valvo, 2000; Scarciglia et al., 2007), and
is reflected in the sediments exported from the Sila Massif through the Samples were separated in up to eleven grain-size fractions ranging
main drainage systems, i.e. the Crati and Neto River basins (e.g. Le from very coarse sand (1 to 2 mm, −1 b Φ b 0) to clay (b 2 μm, Φ N 9),
Pera et al., 2001). with Φ being the negative logarithm of the grain diameter d to the basis
The Calabrian Massif constitutes an allochthonous crustal segment of 2 (Φ = −log2 d). Grain size separation was achieved in one-Φ-unit
of the western Variscan belt in Europe, exposed in a structurally high steps by wet sieving of the sand-sized fractions (− 1 b Φ b 4), and by
position between the Alpine Mesozoic to Cenozoic sedimentary nappe gravity settling for all finer fractions (Φ N 4). Each separation step of
piles of the Apennines to the North and the Maghrebides in Sicily to the fine-grained fractions in Atterberg-cylinders was repeated 8 to 14
the Southwest (Ayuso et al., 1994; Graessner et al., 2000). The Sila times until quantitative separation of the respective grain-size fraction
Massif forms the northern part of the Calabrian Massif and is mainly was achieved. The remaining suspension was vacuum-filtered through
composed of high-grade (granulite facies) metapelites in the West 0.2 μm cellulose acetate filters. Therefore, the finest fraction actually
and Southwest, low to medium-grade metasedimentary rocks in the represents the grain size fraction 9 b Φ b 12.
East, and the Sila Batholith comprising most of the central part of the Each sediment grain-size fraction was powdered, fused using
Sila Massif (Fig. 1). Peak metamorphism is dated at around 300 Ma, lithium metaborate, and subsequently analyzed by X-ray fluorescence
roughly coeval to the Late Variscan intrusion ages of the batholith (XRF) method using a PANalytical AXIOS Advanced sequential X-ray
(Graessner et al., 2000). The Sila Batholith is a complex intrusive body spectrometer at the Geoscience Center Göttingen. All samples and
composed of variable plutonic rocks that range from granites and grano- grain size fractions were analyzed for ten major element oxides (SiO2,
diorites to tonalite, diorite, and even gabbroic rocks (Messina et al., Al2O3, TiO2, Fe2Otot
3 = total iron calculated as Fe2O3, MnO, MgO, CaO,
1991; Ayuso et al., 1994; Caggianelli et al., 1994). After intrusion the Na2O, K2O, P2O5), the loss on ignition (LOI), and 16 trace elements
plutons cooled and were exhumed to mid-crustal levels in Permian to (Ba, Co, Cr, Cu, Ga, Nb, Nd, Ni, Pb, Rb, Sc, Sr, V, Y, Zn, Zr). The RT4 samples
Triassic time (Ayuso et al., 1994; Graessner et al., 2000). Final exhuma- were additionally analyzed for Ce, Hf, La, Mo, Sm, Th, U, and Yb. The
tion of the Sila Massif is constrained by zircon and apatite fission-track elemental concentrations of all samples and grain-size fractions are
analysis to Oligocene to Mid-Miocene time (35 to 15 Ma; Thomson, available in the Supplementary data file.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 3

Fig. 1. Simplified geological map of the study area, modified from Messina et al. (1991) and Borelli et al. (2014).

Analytical precision is typically in the range of 1% for major elements of our samples. The only exception is Zr: although all four reference
and 5% for trace elements. Accuracy of measurements was tested materials have comparable Zr concentrations (96–180 ppm), the de-
against internationally certified reference materials with values for JB- viations for JA-2, JB-3 and JR-1 are only 2% and thus strongly contrast
3, JA-2, and JR-1 taken from Imai et al. (1995) and values for JG-2 the high deviation for JG-2.
taken from Guevara et al. (2001). Accuracy is better than 3% for all Selected samples were analyzed by X-ray diffraction (XRD) for identi-
major oxides except for TiO2 (JG-2, 6%), MnO (JA-2, 4%, JG-2, 14%), fication and quantification of mineral phases. Individual grain size frac-
MgO (JA-2, 4%, JR-1, 8%), P2O5 (JA-2, 5%, JG-2, 46%, JR-1, 14%) and tions were wet milled with a McCrone micronising mill in distilled
Fe2Otot
3 (JG-2, 7%), and better than 8% for all trace elements except water, evaporated to dryness at 60 °C, and pulverized using agate mortar.
for Ba (JR-1, 15%), Ce (JB-3, 12%), Ga (JR-1, 12%), Nb (JG-2, 31%), Rb 10% ZnO was admixed as internal standard. X-ray diffraction analyses
(JB-3, 14%), Sr (JG-2, 75%), Zn (JB-3, 9%) and Zr (JG-2, 45%). Observed were performed on a Philips X'Pert MPD, equipped with a PW3050
deviations higher than 10% are mostly related to concentrations near Goniometer, and CuKα radiation. Qualitative phase analysis was achieved
the detection limit and/or below or at the lower end of the data range using a search match routine of the program X'Pert Highscore (Version
2.2a) (Panalytical B.V., 2006). Quantitative phase analysis was performed
with Rietveld refinements based on the diffraction data using the pro-
gram AutoQuan (Version 2.80) (GE Inspection Technologies, 2014).
Table 1 All Mineral Liberation Analysis (MLA) measurements were run on a
List of samples including bed rock suite, sample type, coordinates and approximate eleva- scanning electron microscope FEI Quanta 650F equipped with two
tion (Sed = sediment from small creeks; WP = weathering profile or grus from roadcuts).
Bruker Quantax X-Flash 5030 Silicon Drift Detectors (SDD) and the spe-
Sample Suite Type N–coordinates–E Elevation cific software MLA Suite 3.1.4 for automated data acquisition. The MLA
RT3-1 F1 WP 39°16′06.17″ 16°32′01.42″ ~1480 masl measurements were carried out at the Geometallurgy Laboratory at
RT3-2 F1 Sed 39°16′06.61″ 16°31′59.36″ ~1480 masl the Helmholtz Institute Freiberg. More detailed information about the
RT3-3A F1 WP 39°16′49.60″ 16°32′18.91″ ~1570 masl functionality of the MLA system can be found in Fandrich et al. (2007).
RT3-3B F1 WP 39°16′49.60″ 16°32′18.91″ ~1570 masl The operating conditions used for this study are listed in Table 2. All
RT3-5A F2 Sed 39°23′14.85″ 16°33′12.04″ ~1170 masl
RT3-6 F2 Sed 39°22′12.59″ 16°31′30.34″ ~1150 masl
measured datasets were processed with the MLA Image Processing soft-
RT4-3 F1 Sed 39°16′48.28″ 16°34′36.48″ ~1580 masl ware, version 3.1.4. A coincidence threshold of 90% with the standard
RT4-4 F1 Sed 39°17′34.23″ 16°31′16.66″ ~1530 masl spectra was used for the classification of the measurements.
RT4-5 F1 Sed 39°17′36.39″ 16°31′23.51″ ~1520 masl The statistical techniques applied are fundamentally based on the
RT4-6 B Sed 39°15′10.33″ 16°32′10.70″ ~1330 masl
log-ratio approach to compositional data, as introduced by Aitchison
RT4-7 B Sed 39°15′17.02″ 16°33′39.36″ ~1380 masl
RT4-10 F2 WP 39°23′21.79″ 16°33′44.29″ ~1270 masl (1986). This includes the evaluation of the covariance structure via
RT4-11 F2 Sed 39°24′17.18″ 16°36′26.30″ ~1430 masl biplot analysis, the calculation of geometric means and predictive re-
RT4-12 F2 Sed 39°23′03.87″ 16°36′51.97″ ~1390 masl gions, and the linear regression model. All this is performed on trans-
RT4-13 F2 WP 39°23′42.77″ 16°36′03.80″ ~1320 masl formed data, i.e. the data are transferred from the constrained
RT4-15 F2 WP 39°23′22.18″ 16°33′43.33″ ~1270 masl
compositional space to the Euclidean space by centered log-ratio

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
4 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

Table 2 4. Results
Summary of MLA parameters for samples RT4-3, RT4-4, and RT4-7.

SEM parameters MLA parameters 4.1. Major and trace element concentrations and patterns
Mode GXMAP Scan speed 16
Voltage (kV) 25 Resolution (pixels) 500 × 500 The variability of the major element oxide concentrations for the
Working dist. (mm) 12 Pixel size (μm/px) 1.6 three sample suites B, F1, and F2 against the eleven separated grain-
Probe current (nA) 10 Acq. time (ms) 6 size fractions from very coarse sand to clay is illustrated in Fig. 2. Con-
Spot size 5.84 Min. EDX-count 2000
trasts between the three suites are clearly visible for some elements
HFW (μm) 800 Step size (px) 6×6
Brightness 83.5 GXMAP trigger 25–255 (e.g. Si, Ca, K, Fe) but typically converge or even disappear towards
Contrast 26.2 Min. particle size (px) 10 finer grain size. Some major element oxides show general trends such
BSE calib. Au 252 Min. grain size (px) 4 as overall decrease of SiO2 and increase of Al2O3 and loss on ignition
(LOI) with decreasing grain size. Other elements display specific pat-
terns for individual parts of the grain size spectrum for some or all of
the three sample suites (e.g. Na, Ca, P, Ti, Fe). For instance, CaO displays
transformation (clr; Aitchison, 1986). The transformation frees the data a more or less pronounced increase in the coarser fractions from very
from the constant-sum constraint implying, for instance, that incremen- coarse sand to very fine sand and a decrease in the fine fractions from
tal changes in concentrations in clr-scale are independent of the respec- very coarse silt to clay. The latter is very pronounced for suites B and
tive level of concentration, i.e. the clr-scale provides a translation- F1. Maximum CaO concentrations are thus measured in the very fine
invariant measure of the degree of change (e.g. von Eynatten et al., sand to very coarse silt fractions (3 b Φ b 5) for all sample suites
2003a). For linear regression, the transformation is followed by standard (Fig. 2). A similar observation is made for Na2O but is restricted to the
multiple regression techniques to estimate the coefficients of the compo- intermediate to felsic sample suites F1 and F2. In the basic sample
sitional linear trend. The trend may include steps if distinct breaks are ob- suite Fe2O3, TiO2, and MnO are peaking in the fine to very fine sand
served at certain grain-size thresholds. Methodological details are range (2 b Φ b 4), accompanied by a small trough in SiO2 and MgO in
described elsewhere (Tolosana-Delgado and von Eynatten, 2009; this range. P2O5 is peaking in very coarse silt (4 b Φ b 5) for all sample
Tolosana-Delgado and van den Boogart, 2011; von Eynatten et al., 2012; suites, decreases in the coarse to fine silt fractions and then again nota-
and references therein). bly increases in the clay fraction (Φ N 9) (Fig. 2). Upon visual inspection

Fig. 2. Major element raw data concentrations in oxide wt.% for all three sample suites of the Sila Massif (blue = B, gabbro to diorite; red = F1, tonalites to granodiorites; brown = F2,
granodiorite to monzogranite) versus grain size in Φ-grades from very coarse sand (−1 b Φ b 0) to clay (Φ N 9). The variability of individual samples of a grain-size class is illustrated
in box plots for suites F1 and F2 (N = 7 each), while sample suite B (N = 2) is represented by stippled lines with solid line reflecting the mean. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 5

of the raw percentage data K2O appears to discriminate best between all component (PC2, y-axis) captures ~ 22% of the total variability and is
suites across the full grain size range, while MgO separates best the basic characterized by strong positive loading of LOI and strong negative load-
suite B from the rest, and CaO separates best F1 from F2. ing of CaO. PC1 effectively separates sediments from basic source rocks
Trace element concentrations typically vary by an order of magni- from intermediate to felsic ones (i.e. sample suites B vs. F1 vs. F2;
tude across sample suites and grain-size range (Fig. 3). Some elements Fig. 4A). This, however, holds only for grain size fractions coarser than
show general increase with decreasing grain size for all sample suites medium silt (i.e. Φ ≤ 6). Towards finer grain size the contrast between
(Th, Pb, U, Ga) while three more elements show this trend for F1 and the sample suites decreases with decreasing scores of PC1 and increas-
F2 only (Zn, Cr, Nb). Zr and Hf peak in the very coarse silt fraction for ing scores of PC2. PC2 groups the very coarse silt to sand fractions for
all sample suites and Zr/Hf ratios scatter around 40 without grain-size each sample suite, and effectively orders the finer grain size fractions
dependence. Y and the light (LREE: La, Ce) and middle rare earth ele- by increasing PC2 scores (brown arrows in Fig. 4).
ments (MREE: Sm, Nd) all peak in the very fine sand to very coarse The mixed biplot of major elements and selected trace elements
silt fractions (3 b Φ b 5) for all sample suites, however, in F2 the concen- (Fig. 4B) reveals a roughly similar picture. While V behaves like the
trations of these elements stay high over the entire fine range (Fig. 3). V mafic elements Mg and Ti (strong negative loading on PC1), Rb, Y and
and Nb peak at 3 b Φ b 5 in the basic sample suite only. Upon visual in- Zr have positive loadings on PC1. Zr also has strong negative loading
spection Th, U, and Rb appear to discriminate best between the three on PC2. Zn appears to be strongly related to the high LOI of the finest
sample suites across the full grain size range. V nicely separates the fractions (positive loading on PC2). The separation of the three sample
basic suite from the rest, while Sr and Ba nicely separate F1 from F2 suites appears even more pronounced in the mixed biplot. Moreover,
across the full grain size range (Fig. 3). grain size separation is enhanced. Fine-grained fractions still have the
highest scores on PC2, while the coarser fractions are now better differ-
4.2. Covariance structure entiated by PC2 (Fig. 4B): the lowest scores in each suite are restricted to
very fine sand and very coarse silt (3 ≤ Φ b 5), while the coarser sand
Evaluation of the covariance structure of the dataset is performed fractions yield intermediate scores on PC2.
using compositional biplots, which provide a 2D-representation of Comparing only the two granitoid suites F1 and F2 allows for focus-
multivariate compositional data based on principal components (e.g. ing the biplot analysis on features controlled by parameters other than
Aitchison, 1990; von Eynatten et al., 2003b). The data set has been source rock, such as grain-size and/or sample type. A threefold structure
transformed using centered log-ratio transformation (clr) to account becomes visible that is predominantly controlled by grain size (Fig. 5):
for the specific compositional nature of geochemical data (Aitchison, (i) very coarse to medium sand (Φ ≤ 2) is mainly characterized by pos-
1986). This implies that the center of the biplot corresponds to the geo- itive scores on PC1 and negative scores on PC2 and PC3 translating to
metric mean of the dataset. high proportions of Si, Na, K, and Rb, (ii) very fine sand and very coarse
The major element biplot of the entire data set (Fig. 4A) reveals a silt (3 ≤ Φ b 5) is mainly characterized by positive scores on PC1, PC2
strong grouping depending on sample suite and grain size. The first and PC3 translating to high proportions of Ca, Zr, Y and P, and (iii)
principal component (PC1, x-axis) captures ~49% of the total variability medium silt to clay (Φ N 6) characterized by negative scores on PC1
and is characterized by strong positive loadings of K and Na and nega- while PC2 and PC3 are rather unspecific. The latter translates to
tive loadings of Mg, Ti, Fe, P, Ca, and LOI. The second principal high proportions of Zn, V, Ti, Mg, Fe, and P. In contrast, the sample

Fig. 3. Trace element raw data concentrations in ppm (note log-scale) for all three sample suites of the Sila Massif (blue = B, gabbro to diorite; red = F1, tonalites to granodiorites;
brown = F2, granodiorite to monzogranite) versus grain size in Φ-grades from very coarse sand (−1 b Φ b 0) to clay (Φ N 9). The variability of individual samples of a grain-size class
is illustrated in box plots for suites F1 and F2 (N = 7 each), while sample suite B (N = 2) is represented by stippled lines with solid line reflecting the mean. For key to box plots see
Fig. 2. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
6 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

Fig. 4. Compositional biplots of the clr-transformed geochemical data (Aitchison, 1990). Symbols reflect sample suite (B, F1, F2) and sample type (Sed = sediment, WP = weathering
profile), color coding refers to grain size class in Φ-units from very coarse sand (yellow; −1 ≤ Φ b 0) to clay (black; Φ N 9). A given value x thus represents the grain-size class of
x ≤ Φ b x + 1. Both biplots indicate sample position in the two dimensional projection based on the first (x-axis, PC-1) and second (y-axis, PC-2) principal components, as well as the var-
iables and their loadings on the respective principal component. Left side (A) shows the major element oxide data including loss on ignition (LOI). Right side (B) shows the same major
elements along with selected trace elements (Rb, Sr, V, Y, Zn, Zr). Numbers in italics refer to percentages of the total variability explained by PC1 and PC2. (For interpretation of the ref-
erences to color in this figure legend, the reader is referred to the web version of this article.)

type (i.e. sediment from creek vs. weathering profile) does not reveal explains the high variance for F2 samples regarding Y and REE concen-
systematic differences in the biplots. However, the fine-grained trations (Fig. 3, third row).
fractions from the weathering profile of suite F2 (WP, triangles
upside down) appear to be separated from their sediment equivalents 4.3. Linear regression model
(Sed, rhombs) by high proportions of Y relative to e.g. Ca. This contrast
The modeling has been restricted to the major elements (including
LOI) and the intermediate to felsic suites F1 and F2, as only these pro-
vide reasonable numbers of samples. Constraints on the linear regres-
sion model like (i) similar slope over the full grain-size range and (ii)
distinct steps at Φ = 4 and Φ = 8, as applied in a previous study (von
Eynatten et al., 2012), resulted in a poor fitting of the data especially
for Ca, Na, P as well as Si and K. The raw data already point at such misfit
because some elements show distinct breaks in increase/decrease pat-
terns at around Φ = 4 (Fig. 2). Moreover, the three-pole structure
governed by grain size as revealed in Fig. 5A calls for modifications
from a simple uniform linear trend over the full grain-size range. We,
therefore, chose model constraints which account for the observed
breaks in slope, i.e. the slopes are allowed to be different for the sand
fraction (Φ b 4) compared to the silt-clay fraction (Φ N 4). Further, a
possible step has been allocated at Φ = 4 but no further steps appear
necessary. The regressions are calculated separately for F1 and F2 with
intercepts as well as the step at Φ = 4 being independent, while the
slopes are forced to be equal for F1 and F2.
The resulting linear trend indicates a reasonable fit to the composi-
tional range of sample suites F1 and F2 for all major elements except
for P2O5 (Fig. 6). The relatively poor linear fit for P is due to its double
peaking in very coarse silt and clay with a significant trough in between.
Well adjusted negative slopes for the full grain-size range (i.e. decreas-
ing relative concentration with decreasing grain size) are obtained for
Si, Na, and K, while general positive slopes are obtained for Fe, Mg, Ti,
and LOI (Fig. 7). In contrast, Ca and Al reveal opposing slopes for sand
and silt-clay: Ca shows distinct increase from coarse sand to very fine
sand and very coarse silt, and then markedly decreases towards clay,
while Al behaves vice versa (Fig. 6). However, Al exhibits very small
slopes in both parts of the grain-size spectrum implying relatively
small changes over the full grain-size spectrum.
Fig. 5. Compositional biplots of major element oxides (excluding LOI) and selected trace
elements (Rb, Sr, V, Y, Zn, Zr) for the intermediate to felsic suites F1 and F2. The composi- 4.4. Mineralogical composition
tional contrast that still exists between F1 and F2 has been minimized by matching the
means of each suite (i.e. by perturbing one data set by the inverse geometric mean of The XRD analysis of three selected samples covers the grain-size
the other). Panel (A) shows the first (x-axis) vs. second (y-axis) principal components, spectrum from coarse sand to clay as well as all three sample suites B,
(B) shows the first (x-axis) vs. third (y-axis) principal components. Numbers in italics
refer to percentages of the total variability explained by PC1 and PC2 (A) or by PC1 and
F1, and F2. Inferred mineral compositions include quartz, K-feldspar
PC3 (B). For key to symbols and colors see Fig. 4. (For interpretation of the references to (orthoclase, microcline), plagioclase (almost pure albite to Ca-rich vari-
color in this figure legend, the reader is referred to the web version of this article.) eties), amphibole (hornblende), pyroxene (enstatite, augite), Ti-phases

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 7

Fig. 6. Linear regression model (solid lines) adjusted for the F1 (red) and F2 (brown) sample suites against grain size classes, displayed in clr-scale for each of the 10 variables included in
the model (9 major elements plus LOI) separately. For key to box plots see Fig. 2. (For interpretation of the references to color in this figure legend, the reader is referred to the web version
of this article.)

(ilmenite, rutile), muscovite, biotite, chlorite, clay minerals (illite, smec- chain silicates, total mica including chlorite, and total clay minerals
tite, kaolinite), and gibbsite. Quantitative phase models are well adjust- (Fig. 8). Moreover, gibbsite and Ti-phases are listed, if appropriate.
ed if the quality parameter 1–ρ of the Rietveld refinement is approx. 1%. In general, mineral proportions are highly variable with respect to
This criterion is reasonably fulfilled for the sample suites F1 and F2 (1–ρ: both sample suites and grain size. As expected, clay minerals and mica
0.8%–3.5%) but less well for sample suite B (1–ρ: 2.5%–6.6%). For this increase from coarse sand to clay. Increase of clay minerals is strongest
sample qualitative analysis revealed a Mg-rich amphibole phase (e.g. in the finest fraction (~60–75%). Only in the F2 samples micas (mainly
cummingtonite) which could not be quantified by AutoQuan due to muscovite) decrease relatively from fine silt to clay, which is, however,
the lack of an appropriate structural model. As a consequence, the amor- more than compensated by the strong increase of clay minerals (~60%)
phous content in this sample is calculated to ~10% within the sand and and gibbsite (~10%). Quartz content decreases continuously from coarse
coarse silt fractions which is likely related to the missing amphibole (~ 22% in B, ~ 40% in F1, F2) to fine (b 5%). The same is observed for
phase. Because structural phase models do not necessarily reflect the K-feldspar, starting from ~ 20%, except for the B-sample where
real structure of the mineral in the sample, especially in the case of high- K-feldspar is pretty low. Plagioclase content is highest (~ 20–30%) in
ly weathered materials, quantitative Rietveld modeling delivered partly the very fine sand and/or coarse silt fractions for all samples, and then
ambiguous results for these highly weathered materials, especially in decreases dramatically towards finer fractions. Amphibole and
the finer-grained fractions. We therefore present, besides quartz, quan- pyroxene decrease strongly from coarse (~ 25%) to fine (b 5%) for the
titative data for the summed values of the clearly determined and quan- B-sample, are relatively low in F1 (max. ~5%, mostly hornblende) and
tified phase groups, i.e. quartz, total K-feldspar, total plagioclase, total almost lacking in F2. Ti-phases are significant in the B-sample only
(max. ~5%), and peak in very fine sand (Fig. 8).
Results for Mineral Liberation Analysis (MLA) are restricted to the
sand-sized fractions of samples RT4-3 and RT4-4 (F1) as well as RT4-7
(B). The mineral phases (or groups) identified and quantified include
quartz, K-feldspar, albite, plagioclase, amphibole (mainly hornblende),
pyroxene (augite, pigeonite), biotite, chlorite, muscovite, clay minerals,
Ti-minerals (e.g. ilmenite, titanite, TiO2-polymorphs), apatite, zircon,
and other accessories (e.g. monazite, magnetite, carbonate). The miner-
al concentrations for each sample and the trends from coarse sand to
very fine sand are listed in Table 3. Separation between suites B and
F1 is straightforward due to strongly contrasting proportions of quartz,
K-feldspar, and chain silicates (amphiboles + pyroxenes). Very good
discrimination within the sand fractions is thus obtained by the quartz
to chain silicates and K-feldspar to plagioclase ratios (Table 3). Very
fine sand typically shows decrease of quartz and K-feldspar and increase
of plagioclase and amphibole compared to coarse sand. Heavy minerals
such as apatite, zircon, and Ti-minerals increase by factors of ~2 to 17 in
very fine sand compared to coarse sand.
Comparison of XRD vs. MLA data for samples RT4-3 (suite F1) and
RT4-7 (suite B), for which both methods were applied on two grain-
size fractions each, yields an overall good agreement of the respective
mineral phases (Fig. 9). RT4-3 yields good agreement for all phases ex-
cept for muscovite, which yields lower values with MLA. RT4-7 yields
Fig. 7. Graphical representation of the model parameters slope in the sand fraction (Φ b 4) very good agreement for quartz, amphibole, K-feldspar, and clay min-
vs. slope in the silt to clay fraction (Φ N 4). Dotted line reflects 1:1 relation of the slopes. erals, good agreement for plagioclase, biotite + chlorite, and muscovite,

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
8 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

Fig. 8. Quantitative X-ray diffraction (XRD) results of three samples, one from each of the sample suites B (RT4-7), F1 (RT4-3), and F2 (RT4-11). For each sample five grain-size fractions
from coarse sand (0 b Φ b 1) to clay (Φ N 9) were analyzed and quantified. Mica includes biotite, muscovite, and chlorite; claymin includes illite, smectite, and kaolinite; gib = gibbsite;
ti-phases include ilmenite and rutile. Error bars include 3 sigma errors of Rietveld refinements. Note different scale for concentrations N32%.

but some contrast in the case of pyroxene and Ti-minerals (especially original minerals into clay minerals, with kaolinite commonly reflecting
for coarse sand; Fig. 9). The higher values for pyroxene in the MLA intense weathering and complete removal of mobile elements including
dataset may derive from some problems in discriminating Al-poor am- potassium. Chemical weathering may also mobilize silica, including
phiboles from pyroxene with this technique. Alternatively, Mg-rich desilicification of clay minerals and formation of gibbsite (Vazques,
chain silicates have been observed in the XRD-spectra but could 1981). Among the vast number of proxies for chemical weathering the
not be quantified and may be as high as ~ 10% (see above). Taken chemical index of alteration (CIA; Nesbitt and Young, 1982) is by far
together, the data indicate reasonable consistency between the two the most popular. It essentially describes the degree of removal of
methods. Ca2+, Na+, and K+, normalized to Al, through transformation of feld-
spars to clay minerals. Unweathered source rocks typically have CIA of
5. Discussion 40–50 while extremely weathered materials entirely composed of, for
instance, kaolinite, gibbsite, and Fe-(hydr)oxides have CIA close to
5.1. Chemical weathering 100. Average pelitic rocks that typically experienced significant
weathering yield CIA of ~ 70 (e.g. CIA 70 for post-Archean Australian
Chemical weathering of silicate minerals such as feldspar or chain shale, Taylor and McLennan, 1985; CIA 69 for Phanerozoic average cra-
silicates leads to exchange of cations of alkaline and alkaline earth ele- tonic shale, Condie, 1993). It must be noted that CIA values or other
ments for H+ via hydrolysis (e.g. Bahlburg and Dobrzinski, 2011). Min- weathering indices do not exclusively depend on the degree of chemical
eralogically, this process is characterized by the transformation of the weathering. Other factors such as initial source rock composition

Table 3
Results of MLA analysis.

Sample RT4-3 RT4-4 RT4-7

Grain-size cs–vfsa cs–vfsa cs–vfsa

Quartz 40.5–20.2 38.2–19.2 22.7–11.3


K-feldspar 23.3–14.1 13.7–17.4 4.3–2.4
Albite 7.4–9.2 6.8–10.6 4.5–3.2
Plagioclase 5.1–11.9 9.6–16.1 (6.6)b 5.3–16.1
Amphibole 0.4–5.1 0.03–6.0 21.2–21.7 (26.3)b
Pyroxene 0.0–0.0 0.0–0.0 14.9–9.9
Biotite 7.4–11.4 10.8–7.1 (12.4)b 8.2–6.9
Chlorite 0.5–1.7 0.6–1.3 1.6–8.1
Muscovite 1.4–1.6 (2.4)b 2.5–1.7 2.8–1.7
Clay minerals 13.6–23.7 17.1–17.4 (14.1)b 11.6–13.5
Ti-minerals 0.10–0.39 0.25–0.86 2.2–4.5 (5.9)b
Apatite 0.14–0.29 0.20–1.25 0.44–0.63
Zircon 0.05–0.12 (0.03)b 0.03–0.54 0.02–0.06
Quartz/Σ feldspar 1.1–0.6 1.3–0.8 1.6–0.5
K-feldspar/plagioclasec 1.9–0.7 0.8–0.7 0.4–0.1
Quartz/chain silicatesd ~100–4.0 N100–3.2 0.6–0.4
Quartz/clay minerals 3.0–0.9 2.2–1.1 2.0–0.8
a
The values [wt.%] describe the trend from coarse sand (cs) to very fine sand (vfs).
b
Values in brackets refer to exceptions from the trend observed in medium or fine sand.
c
Albite included in plagioclase.
d
Chain silicates = amphibole + pyroxene.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 9

Calculated CIA values for the Sila samples show a very wide range
from 55 to 92, i.e. from almost “fresh” to heavily weathered materials
(Fig. 10). There is no clear distinction between the three sample suites.
Nevertheless, the most felsic suite F2 tends to slightly lower CIA at the
coarse and fine tails of the grain-size range and the basic suite shows
slightly higher CIA from coarse silt to very fine silt (5 b Φ b 9; Fig. 10).
In general, CIA values show a pronounced increase with decreasing
grain size. However, within the sand fraction there is only little increase
from CIA 55–61 for very coarse sand to CIA 56–68 for very fine sand. The
average CIA for all sand fractions from all samples is 60. Very coarse silt
(4 b Φ b 5) is still relatively low in CIA (57–68; medians still around 60)
while all finer fractions mark a continuous increase up to the clay
fraction, which range in CIA from 77 to 92 with mean values between
84 and 88 for the three suites (Fig. 10). Average CIA for all the fine frac-
tions from coarse silt to clay (Φ N 5) is 75.4, which is higher than average
shale (see above).
The strong chemical weathering in the Sila Massif has long been
known and is described elsewhere. It is characterized by the develop-
ment of deep weathering profiles on granitoid bedrock (e.g. Le Pera
and Sorriso-Valvo, 2000), dissolution microtextures including intensely
etched quartz grains (e.g. Scarciglia et al., 2007), and the occurrence of
various neo-formed clay minerals such as illite–smectite mixed layers,
kaolinite, and halloysite (e.g. Borelli et al., 2014). Our data support this
through (i) high CIA up to 92 and (ii) high proportions of minerals indi-
cating strong chemical weathering such as kaolinite and gibbsite in the
finest sediment fractions.
Grain size dependence of chemical weathering indices was de-
scribed earlier, but has been typically restricted to broader categories
such as sand(stone) vs. mud(stone) (e.g. Ohta, 2008). The detailed
grain-size resolution obtained here fully confirms that comparing
weathering proxies from samples with different grain size distributions
is likely to produce biased results. The effect appears less relevant for
different grades of sand(stone) but highly relevant for comparison of
sand vs. mud and for comparison of different grades and mixtures of
silt to clay. For the latter, our data suggest strong contrast in CIA al-
though the source area is similar in terms of both bedrock type (for
Fig. 9. Comparison of results on mineral composition from X-ray diffraction (XRD) vs. Min- each sample suite) and weathering conditions. This statement is crucial
eral Liberation Analysis (MLA) for two samples from suite F1 (RT4-3) and suite B (RT4-7). given that many case studies have based their interpretations on rela-
Two grain-size fractions from each sample are available for comparison (qtz = quartz, tive subtle changes in weathering proxies like, for instance, 5 to
pla = plagioclase incl. albite, kf = K-feldspar, am = amphibole, px = pyroxene, bio = bi-
15 units in CIA (e.g. Nesbitt and Young, 1982; Passchier and Krissek,
otite, chl = chlorite, clm = Σ clay minerals). For explanation see text.
2008; Bahlburg and Dobrzinski, 2011).

(Nesbitt and Young, 1984), grain size (von Eynatten et al., 2012), or 5.2. Source-rock discrimination
provenance settings including e.g. sheet-silicate rich units or sediment
recycling (Garzanti and Resentini, 2015) are prone to add significant The three sample suites B, F1 and F2 reflect marked contrasts in bed-
variation to the calculated CIA values. rock mineralogy and geochemistry. The question is whether the sedi-
ments reflect these contrasts across the full grain size range despite
intense chemical alteration? Raw data concentrations and covariance
structure suggest high discriminative potential for K2O, MgO, and CaO,
as well as Rb, Th, U, V, Sc, Co, Ba, and Sr (Figs. 2 to 5). Because some of
these elements are sensitive to chemical weathering (e.g. Ca, Sr, Rb, K,
U), this observation implies that chemical weathering was not extreme
enough to fully destroy source rock signatures. Typical element ratios
used to discriminate sediments derived from basic vs. sediments de-
rived from felsic rocks such as Ba/Co, La/Co, La/Sc, or Th/Sc (e.g. Bhatia
and Crook, 1986; Cullers et al., 1988) almost completely separate sam-
ple suite B from the more felsic suites, like does Rb/Sc and Rb/Co. The
best separation between suite B and all other samples based on single
ratios is obtained by Rb/V. All these ratios, however, do not sufficiently
separate F1 from F2 across all grain size grades. This can be achieved
by using Rb/Sr as additional discriminator (Fig. 5A).
Consequently, a very good discrimination of the three sample suites
is achieved in the ternary diagram with trace elements Rb, Sr and V,
Fig. 10. Chemical index of alteration (CIA) for each grain-size fraction and each sample
suite B, F1, and F2. The variability of individual samples of a grain-size class is illustrated
where relative V concentration effectively discriminates sample
in box plots for suites F1 and F2 (N = 7 each), while sample suite B (N = 2) is represented suite B from all the other samples (probability N90%) and Rb/Sr ratio
by stippled lines with solid line reflecting the mean. For key to box plots see Fig. 2. is the main discriminator between F1 and F2 (Fig. 11). The latter

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
10 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

fractions, clay minerals even increase relative to mica (i.e. stronger de-
crease of quartz/clay compared to quartz/mica; Fig. 12).
Chemically, sheet silicate enrichment is reflected by an overall de-
crease of SiO2 and increase of Al2O3 and LOI from very coarse sand to
clay (Fig. 2). All other major elements have a more complex pattern, es-
pecially in relative terms (clr-scale, Fig. 6) where even Al increases in
the Φ N 4 fractions only. Considering the sand-fractions separately, the
main trend is controlled by strong depletion of SiO2 and K2O while
CaO and P2O5 increase. All other major elements show only minor
changes (slopes b 0.05) except for LOI showing moderate increase
(Fig. 7). This pattern mostly reflects depletion of quartz and K-feldspar
and increase of plagioclase (and partly amphibole, RT4-3) in (very)
fine sand compared to (very) coarse sand, as corroborated by XRD
data (Fig. 8). Chemical weathering of feldspar grains preferentially
evolves at the surfaces as well as along cleavage and twinning planes,
the latter in turn accelerate mechanical comminution of mineral grains
(e.g. Berner and Holdren, 1979; Banfield and Eggleton, 1990). Among
Fig. 11. V–Rb–Sr ternary diagram for discrimination of the three sample suites B, F1, and
F2. Fields represent predictive regions for each suite with 67% (solid line) and 90% the feldspars, plagioclase weathers much faster than K-feldspar (e.g.
(stippled line) probability. White et al., 2001), implying that comminution of plagioclase is more
effective than K-feldspar at the initial stages of feldspar weathering.
Therefore, plagioclase is enriched over K-feldspar in fine sand to coarse
discrimination is less strict due to minor overlap between F1 and F2, silt fractions, where Ca peaks in both raw data and the linear model
which mainly derive from the finer fractions (Φ N 5) of one single sam- (Figs. 2, 6). This decrease in K-feldspar/plagioclase ratio from (very)
ple at the western boundary of suite F2 (RT3-6) that shows relatively coarse to very fine sand is strongly corroborated by both XRD data
low Rb/Sr ratios compared to the rest of F2. Replacing the trace elements (Fig. 12) and MLA data (Table 3). P2O5 reflects enrichment of apatite
Rb, Sr and V by the major element oxides K2O, CaO, and MgO, respec- and/or other phosphate minerals such as monazite in very fine sand/
tively, leads to almost the same level of discrimination. very coarse silt, consistent with general enrichment of heavy mineral
in these size fractions (e.g. zircon, Ti-minerals; Table 3).
In the silt to clay range, the linear model is characterized by strong
5.3. Grain-size control on composition increase in LOI and strong decrease of CaO and especially Na2O. More-
over, a moderate increase of Fe2O3, and MgO (slopes around 0.1) is ob-
Biplot analysis of major and trace elements as well as the linear re- served, and only minor changes for TiO2, P2O5, Al2O3, SiO2 and K2O
gression model reveals strong grain size control on sediment composi- (slopes at around ±0.03–0.07, Fig. 7). This observation reflects strong
tion. In mineralogical terms the general trend from coarse to fine is increase of clay minerals and mica compared to all other minerals
accompanied by the increase of sheet silicates at the expense of quartz (Figs. 8, 12) and is well in line with the strong increase of CIA obtained
and the silicates. The increase of sheet silicates is twofold: (i) newly for Φ N 5 (Fig. 10). Because some of these sheet silicates still keep potas-
formed clay minerals are preferentially concentrated in the finest frac- sium (illite, biotite, muscovite), the decrease of K2O is only minor com-
tions due to intense chemical weathering of feldspars and other sili- pared to CaO and Na2O, although most K-feldspar and plagioclase have
cates, and (ii) micas are enriched in the finer fractions due to enhanced disappeared in the finest fractions (Fig. 8). Relative increase of Fe2O3
chemical alteration and mechanical comminution along cleavage planes, and MgO also reflects increase of the respective sheet silicates (biotite,
as well as hydrodynamic sorting. This is reflected in the ratios of quartz chlorite, smectite). Fe-(hydr-)oxides appear negligible because Fe2O3
and feldspar over both micas and clay minerals, which strongly decrease and MgO are strongly coupled (Fig. 7).
from sand to silt-clay (Fig. 12). Decrease of the quartz/clay ratio from Interestingly, the least changes in the linear regression model over
coarse sand to clay may exceed two orders of magnitude. In the finest the full grain size range (i.e. smallest slopes regardless of sign; Fig. 7)
refer to Al2O3 (Figs. 6, 7). Obviously, Al increase in raw data concentra-
tions towards fine fractions (Fig. 2) but has low overall variability
(Figs. 4, 5) and only small (F1) or almost no (F2) increase from coarse
to fine in relative clr-scale (Fig. 6). Because Al is largely immobile during
weathering processes and is commonly enriched in most clay minerals,
Al typically increases towards fine fractions, even in clr-scale (e.g.
Bloemsma et al., 2012). This effect is quite small here, likely because
we included the loss on ignition (reflecting mainly the water content
of the sheet silicates), which captures most of the relative increase to-
wards the fine grain-size classes.
Regarding trace elements, the Zr/Zn ratio provides a straightforward
proxy for grain size, as already observed for basic and granitoid rocks in
glacial setting (von Eynatten et al., 2012). The new data corroborate this
finding, and extend its use to intensely weathered settings, where Zn is
similarly enriched in the finest fractions, regardless of source-rock li-
thology (Fig. 13). The source rocks, i.e. basic versus more felsic lithology,
are best discriminated by the Rb/V ratio (Figs. 3, 11, 13).

5.4. Comparison to non-weathered sediments


Fig. 12. Selected mineral ratios from coarse sand to clay for all three sample suites based on
X-ray diffraction data. Note logarithmic scale implying that, for instance, quartz to clay
mineral ratio (qtz/clay) decreases by more than two orders of magnitude from coarse In an analogous study, von Eynatten et al. (2012) investigated the in-
sand to clay. fluence of grain size and source rocks on sediment composition in

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 11

model constraints as for the Sila Massif, except for the step at Φ = 8
which is obviously necessary to describe the Aar Massif data set (von
Eynatten et al., 2012).
The comparison shows striking contrasts in the modeled linear re-
gression trends of the major elements against grain size (Fig. 14). The
main observations on contrasting trends for the Sila case vs. the Aar
case include (i) strong decrease of SiO2 already in the sand fraction
from very coarse to very fine sand, (ii) strong increase of CaO in the
sand fraction from coarse to fine, (iii) continuous and pronounced de-
crease for CaO and Na2O in the silt-clay fraction, (iv) continuous in-
crease of Al2O3 and Fe2O3 for the silt-clay fraction (i.e. no significant
steps at Φ = 8), and (v) no increase of K (i.e. no step) for the finest frac-
tions at Φ N 8. These observations mainly reflect the following processes,
which are prominent in the Sila case compared to the Aar case:

(1) Initial plagioclase weathering through hydrolysis along cleavage


and/or twinning planes causes breakage and preferential enrich-
ment of plagioclase over quartz and K-feldspar from coarse sand
to very fine sand, as reflected in Ca increase (observation ii) and
some Si decrease (observation i). The effect may be supported by
initial amphibole weathering in the case of F1.
(2) Intense chemical weathering mainly through hydrolysis leads to
the almost complete removal of feldspar and chain silicates from
the intermediate grain sizes (3 b Φ b 6) to the finest fractions
Fig. 13. Rb/V vs. Zr/Zn diagram showing all samples and all grain-size fractions. For expla- (observation iii).
nation see text.
(3) Enrichment of various Al- and/or Fe-rich sheet silicates (biotite,
chlorite, smectite, illite, kaolinite) towards the finer and finest
glacial settings where chemical weathering is negligible and mechanical fractions as a consequence of process no. 2 (observation iv).
comminution is the main process that governs the separation of mineral (4) Enrichment of K-poor clay minerals (kaolinite, smectite) relative
phases and, hence, their distribution in different grain-size fractions. to micas in the finest fractions (observation v). This reflects ongo-
Study area was granitic bedrock of the Aar Massif in the Central Alps ing removal of potassium, which is essential to obtain the high
and basic amphibolites from the Silvretta Massif in the eastern Alps. CIA values.
The Aar example is quite similar in bedrock composition to sample
suite F2 from the Sila Massif. Given the similar starting point (i.e. gran- The combination of processes (1) and (2) causes the smoothed Ca
itoid bedrock composition) the comparison of the two cases is actually a (Figs. 2, 6) and plagioclase (Fig. 8) peaks in the Sila case instead of the
comparison of the effects of strong vs. negligible chemical weathering. abrupt break (i.e. step) in Ca concentration of the glacial Aar case
For a reasonable comparison of the linear regression, we recalculated (Fig. 14), which is caused by mechanical comminution in combination
the model adjusted to the Aar Massif case study, applying the same with inherited mineral-specific grain size distributions (von Eynatten

Fig. 14. Comparison of the adjusted linear regression trends for the weathered case (Sila Massif, sample suites F1 and F2) against the glacial case (Aar Massif) for the most important major
elements. Minor differences in F1 and F2 values compared to Fig. 6 are caused by excluding loss on ignition (LOI) because these data are not available for the Aar Massif data set.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
12 H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx

et al., 2012). The latter combination was also found to be responsible for (6) Given the constantly contrasting behavior of SiO2 and Al2O3 the
quartz concentrations peaking in fine to very fine sand in the glacial case ratio Si/Al provides a reasonable proxy for grain size in the Sila
study (2 b Φ b 4), which is not the case in the Sila example. There, SiO2 case. This is not valid for the glacial case (Fig. 14). However, Zr/
continuously decreases over the full grain-size range from coarse to fine Zn provides a good proxy for grain size valid for both settings:
(observation i), which is caused by intense hydrolysis supported by high ratios indicate zircon concentration in the intermediate frac-
quartz leaching (Scarciglia et al., 2007) and, finally, desilicification of tions (mostly 3 b Φ b 5) while low ratios indicate sheet silicate en-
clay minerals and formation of gibbsite as observed in some samples richment in the fine fractions (e.g. Φ N 7).
(Fig. 8). Despite all these differences between the highly weathered
and the glacial case study, some elements converge towards the finest
fractions to very similar relative concentrations independent of similar The application of Mineral Liberation Analysis (MLA) yields some
(e.g. SiO2) or contrasting starting points (e.g. Fe2O3). encouraging results, namely (i) the apparent compatibility with X-ray
For the fine fractions, the elements showing most contrast between diffraction data for the main source-rock forming minerals quartz,
the highly weathered and the glacial case study are Al2O3 and Na2O K-feldspar, plagioclase, amphibole, and most sheet silicates, and (ii)
(Fig. 14), suggesting that the ratio of these two elements represents a the abundance and grain-size distribution of certain heavy minerals.
simple and straightforward proxy for chemical weathering, which The potential of the method with respect to the fine grain size ranges
avoids potential bias through K-metasomatism (Fedo et al., 1995) and especially highly weathered materials needs further exploration.
and/or the need for corrections for Ca associated with non-silicate Supplementary data to this article can be found online at http://dx.
phases (von Eynatten et al., 2003a). doi.org/10.1016/j.sedgeo.2015.10.008.

6. Conclusions Acknowledgments

Analysis and modeling of composition vs. grain size relations of Salvatore Critelli greatly supported us during field work in the Sila
proximal sediments in a highly weathered Mediterranean setting (Sila Massif and subsequent discussions. Sample preparation and grain-size
Massif) and its comparison to a previous study in glacial settings in fractionation were performed by Cornelia Friedrich. Journal reviewers
the European Alps (von Eynatten et al., 2012) reveal the following Abhijit Basu and Gert Jan Weltje and guest editor Eduardo Garzanti
main conclusions: are thanked for their careful and constructive handling of the
manuscript.
(1) The chemical index of alteration (CIA) corroborates strong chem-
ical weathering by individual values as high as 92 and average References
values of 84 to 88 for the finest fractions. Fractions coarser than Aitchison, J., 1986. The Statistical Analysis of Compositional Data. Monographs on Statis-
coarse silt (i.e. Φ b 5), however, reveal low to moderate values tics and Applied Probability. Chapman & Hall Ltd., London (UK) ((Reprinted in
in CIA (55 to 68) and an average of 60 for all sand fractions 2003 with additional material by The Blackburn Press). 416 pp.).
Aitchison, J., 1990. Relative variation diagrams for describing patterns of compositional
from all sample suites. The CIA value of a specific sample thus variability. Mathematical Geology 22, 487–511.
strongly depends on its grain-size distribution. Ayuso, R.A., Messina, A., De Vivo, B., Russo, S., Woodruff, L.G., Sutter, J.F., Belkin, H.E., 1994.
(2) Although strongly weathered, the three sample suites reflecting Geochemistry and argon thermochronology of the Variscan Sila batholith, southern
Italy: source rocks and magma evolution. Contributions to Mineralogy and Petrology
basic to intermediate (B), intermediate to felsic (F1), and felsic 117, 87–109.
(F2) plutonic bed rock can be effectively discriminated across Bahlburg, H., Dobrzinski, N., 2011. A review of the chemical index of alteration (CIA) and
all grain-size classes using trace elements such as V, Rb, and Sr. its application to the study of Neoproterozoic glacial deposits and climate transitions.
In: Arnaud, E., Halverson, G.P., Shields-Zhou, G. (Eds.), The Geological Record of
At least in the coarser fractions, major elements Mg, Ca, and K Neoproterozoic Glaciations. Geological Society, London, Memoirs 36, pp. 81–92.
do so, too. Banfield, J.F., Eggleton, R.A., 1990. Analytical transmission electron microscope studies of
(3) The chemical composition vs. grain-size relations can be accu- plagioclase, muscovite, and K-feldspar weathering. Clays and Clay Minerals 38,
77–89.
rately modeled by linear regression only if different slopes are Basu, A., 1976. Petrology of Holocene fluvial sand derived from plutonic source rocks: im-
allowed for the sand (Φ b 4) and the silt to clay (Φ N 4) fractions. plications to paleoclimatic interpretation. Journal of Sedimentary Petrology 46,
Similar slopes in the respective grain-size range for different 694–709.
Basu, A., Young, S.W., Suttner, L.J., James, W.C., Mack, G.H., 1975. Re-evaluation of the use
sample suites (i.e. intermediate and felsic source rocks F1 and
of undulatory extinction and polycrystallinity in detrital quartz for provenance inter-
F2, respectively) underline similar processes independent of pretation. Journal of Sedimentary Petrology 45, 873–882.
source rock composition. Berner, R.A., Holdren, G.R., 1979. Mechanism of feldspar weathering — II. Observations of
(4) The mineralogy behind the geochemical data reflects overall (i.e. feldspars from soils. Geochimica et Cosmochimica Acta 43, 1173–1186.
Bhatia, M., Crook, K.A.W., 1986. Trace element characteristics of greywackes and tectonic
valid for all three sample suites) decrease of quartz and K- setting discrimination of sedimentary basins. Contribution to Mineralogy and Petrol-
feldspar over the full grain-size range from coarse to fine. This ogy 92, 181–193.
is compensated by overall increase of sheet silicates from coarse Blatt, H., Middleton, G.V., Murray, R.C., 1972. Origin of Sedimentary Rocks. Prentice-Hall,
Englewood Cliffs, NJ (US) (634 pp.).
to fine, where the increase of clay minerals strongly outpaces the Bloemsma, M.R., Zabel, M., Stuut, J.B.W., Tjallingii, R., Collins, J.A., Weltje, G.J., 2012. Model-
increase of micas in the silt to clay fractions (i.e. Φ N 5). A more ling the joint variability of grain size and chemical composition in sediments. Sedi-
complex behavior is shown by plagioclase, which peaks in inter- mentary Geology 280, 135–148.
Borelli, L., Perri, F., Critelli, S., Gullà, G., 2014. Characterization of granitoid and gneiss
mediate grain-size fractions (i.e. 3 b Φ b 6) in all sample suites. weathering profiles of the Mucone River basin (Calabria, southern Italy). Catena
The latter is caused by initial hydrolysis along cleavage planes 113, 325–340.
and subsequent breakage of plagioclase crystals into smaller Caggianelli, A., Del Moro, A., Piccarreta, G., 1994. Petrology of basic and intermediate oro-
genic granitoids from the Sila Massif (Calabria, southern Italy). Geological Journal 29,
fragments, preferentially in the 125 μm to 16 μm range. 11–28.
(5) Mechanical comminution, mineral durability, and inherited Condie, K.C., 1993. Chemical composition and evolution of the upper continental crust:
grain-size distribution were found to be relevant parameters in contrasting results from surface samples and shales. Chemical Geology 104, 1–37.
Cullers, R.L., Basu, A., Suttner, L.J., 1988. Geochemical signature of provenance in sand-size
the glacial case study, responsible for peaking of, for instance,
material in soils and stream sediments near the Tobacco Root Batholith, Montana,
quartz and total SiO2 concentration in fine to very fine sand USA. Chemical Geology 70, 335–348.
(2 b Φ b 4). This is not observed in the highly weathered Sila Fandrich, R., Gu, Y., Burrows, D., Moeller, K., 2007. Modern SEM-based mineral liberation
case. Instead, both quartz and total SiO2 concentrations continu- analysis. International Journal of Mineral Processing 84, 310–320.
Fedo, C.M., Nesbitt, H.W., Young, G.M., 1995. Unravelling the effects of potassium metaso-
ously decrease from coarse to fine, reflecting the cumulative ef- matism in sedimentary rocks and paleosoils, with implications for paleoweathering
fects of minor mechanical forces, quartz leaching, and hydrolysis. and provenance. Geology 23, 921–924.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008
H. von Eynatten et al. / Sedimentary Geology xxx (2015) xxx–xxx 13

Garzanti, E., Resentini, A., 2015. Provenance control on chemical indices of weathering Olivetti, V., Cyr, A.J., Molin, P., Faccenna, C., Franger, D.E., 2012. Uplift history of the Sila
(Taiwan river sands). Sedimentary Geology, http://dx.doi.org/10.1016/j.sedgeo. Massif, southern Italy, deciphered from cosmogenic 10Be erosion rates and river lon-
2015.06.013 (this volume). gitudinal profile analysis. Tectonics 31, TC3007.
Garzanti, E., Andò, S., Vezzoli, G., 2009. Grain-size dependance of sediment composition Panalytical, B.V., 2006. X'pert Highscore 2.2a. Software for X-ray Diffraction Analysis.
and environmental bias in provenance studies. Earth and Planetary Science Letters Panalytical B.V., Almelo, The Netherlands.
277, 422–432. Passchier, S., Krissek, L.A., 2008. Oligocene–Miocene Antarctic continental weathering re-
Garzanti, E., Andò, S., France-Lanord, C., Censi, P., Vignola, P., Galy, V., Lupker, M., 2011. cord and paleoclimatic implications, Cape Roberts drilling project, Ross Seal,
Mineralogical and chemical variability of fluvial sediments. 2. Suspended-load silt Antarctica. Palaeogeography Palaeoclimatology Palaeoecology 260, 30–40.
(Ganga–Brahmaputra, Bangladesh). Earth and Planetary Science Letters 302, Pettijohn, F., 1975. Sedimentary Rocks. 3rd ed. Harper & Row, New York (628 pp.).
107–120. Scarciglia, F., Le Pera, E., Critelli, S., 2007. The onset of the sedimentary cycle in a mid-
GE Inspection Technologies GmbH, 2014. Rayflex — X-ray Diffraction Software, AutoQuan latitude upland environment: weathering, pedogenesis, and geomorphic processes
Version 2.80. Ahrensburg, Germany. on plutonic rocks (Sila Massif, Calabria). Geological Society of America, Special
Graessner, T., Schenk, V., Bröcker, M., Mezger, K., 2000. Geochronological constraints on Paper 420, 149–166.
the timing of granitoid magmatism, metamorphism and post-metamorphic cooling Scarciglia, F., Saporito, N., La Russa, M.F., Le Pera, E., Macchione, M., Puntillo, D., Crisci,
in the Hercynian crustal cross-section of Calabria. Journal of Metamorphic Geology G.M., Pezzino, A., 2012. Role of lichens in weathering of granodiorite in the Sila up-
18, 409–421. lands (Calabria, southern Italy). Sedimentary Geology 280, 119–134.
Guevara, M., Verma, S.P., Velasco-Tapia, F., 2001. Evaluation of GSJ intrusive rocks JG1, Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: Its Composition and Evolution.
JG2, JG3, JG1a, and JGb1 by an objective outlier rejection statistical procedure. Revista Blackwell Science, Oxford 315 pp.
Mexicana de Ciencias Geologicas 2001 (18), 74–88. Thomson, S.N., 1994. Fission track analysis of the crystalline basement rocks of the
Imai, N., Terashima, S., Itoh, S., Ando, A., 1995. 1994 Compilation of analytical data for Calabrian Arc, southern Italy: evidence of Oligo-Miocene late-orogenic extension
minor and trace elements in seventeen GSJ geochemical reference samples, “igneous and erosion. Tectonophysics 238, 331–352.
rock series”. Geostandards Newsletter 1995 (19), 135–213. Tolosana-Delgado, R., van den Boogart, K.G., 2011. Linear models with compositions in R.
Ingersoll, R.V., Bullard, T.F., Ford, R.L., Grimm, J.P., Pickle, J.D., Sares, S.W., 1984. The effect In: Pawlowsky-Glahn, V., Buccianti, A. (Eds.), Compositional Data Analysis: Theory
of grain size on detrital modes: a test of the Gazzi–Dickinson point-counting method. and Applications. John Wiley & Sons, Ltd., pp. 356–371 (Chapter 26).
Journal of Sedimentary Petrology 54, 103–116. Tolosana-Delgado, R., von Eynatten, H., 2009. Grain-size control on petrographic compo-
Johnsson, M.J., 1993. The system controlling the composition of clastic sediments. Geolog- sition of sediments: compositional regression and rounded zeroes. Mathematical
ical Society of America, Special Paper 284, 1–19. Geosciences 41, 869–886.
Le Pera, E., Sorriso-Valvo, M., 2000. Weathering and morphogenesis in a Mediterranean Vazques, F.M., 1981. Formation of gibbsite in soils and saprolites of temperate–humid
climate, Calabria, Italy. Geomorphology 34, 251–270. zones. Clay Minerals 16, 43–52.
Le Pera, E., Arribas, J., Critelli, S., Tortosa, A., 2001. The effects of source rocks and chemical von Eynatten, H., Barceló-Vidal, C., Pawlowsky-Glahn, V., 2003a. Modelling compositional
weathering on the petrogenesis of siliciclastic sand from the Neto River (Calabria, change: the example of chemical weathering of granitoid rocks. Mathematical Geol-
Italy). Sedimentology 48, 357–378. ogy 35, 231–251.
Messina, A., Compagnoni, R., De Vivo, B., Perrone, V., Russo, S., 1991. Geological and pet- von Eynatten, H., Barceló-Vidal, C., Pawlowsky-Glahn, V., 2003b. Sandstone composition
rochemical study of the Sila Massif plutonic rocks (northern Calabria, Italy). Bollettino and discrimination: a statistical evaluation of different analytical methods. Journal
della Societa Geologica Italiana 110, 165–206. of Sedimentary Research 73, 47–57.
Morton, A.C., Hallsworth, C.R., 1994. Identifying provenance-specific features of detrital von Eynatten, H., Tolosana-Delgado, R., Karius, V., 2012. Sediment generation in modern
heavy mineral assemblages in sandstones. Sedimentary Geology 90, 241–256. glacial settings: source-rock and grain-size control on sediment composition. Sedi-
Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred mentary Geology 280, 80–92.
from major element chemistry of lutites. Nature 299, 715–717. White, A.F., Bullen, T.D., Schulz, M.S., Blum, A.E., Huntington, T.G., Peters, N.E., 2001. Differ-
Nesbitt, H.W., Young, G.M., 1984. Prediction of some weathering trends of plutonic and ential rates of feldspar weathering in granitic regoliths. Geochimica et Cosmochimica
volcanic rocks based on thermodynamic and kinetic considerations. Geochimica et Acta 65, 847–869.
Cosmochimica Acta 48, 1523–1534.
Ohta, T., 2008. Measuring and adjusting the weathering and hydraulic sorting effects for
rigorous provenance analysis of sedimentary rocks: a case study from the Jurassic
Ashikita Group, south-west Japan. Sedimentology 55, 1687–1701.

Please cite this article as: von Eynatten, H., et al., Sediment generation in humid Mediterranean setting: Grain-size and source-rock control on
sediment geochemistry and mineralogy (Sila Massif, Calabria), Sedimentary Geology (2015), http://dx.doi.org/10.1016/j.sedgeo.2015.10.008

Vous aimerez peut-être aussi