Next Article in Journal
Lineaments in the Gravity Image of the Border Zone between the Central and Outer Carpathians
Next Article in Special Issue
A Review of the Mineralogy, Petrography, and Geochemistry of Serpentinite from Calabria Regions (Southern Italy): Problem or Georesource?
Previous Article in Journal
Nanoquartz in Late Cretaceous Deposits in the Lower Berezovskaya Subformation
Previous Article in Special Issue
Petrographic and Geochemical Inferences for Genesis of Terra Rossa: A Case Study from the Apulian Karst (Southern Italy)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Geochemistry as a Clue for Paleoweathering and Provenance of Southern Apennines Shales (Italy): A Review

Department of Sciences, University of Basilicata, 85100 Potenza, Italy
*
Author to whom correspondence should be addressed.
Minerals 2023, 13(8), 994; https://doi.org/10.3390/min13080994
Submission received: 28 June 2023 / Revised: 22 July 2023 / Accepted: 24 July 2023 / Published: 26 July 2023

Abstract

:
The southern Apennines (Italy) chain is a fold-and-thrust belt mainly derived from the deformation of the African–Apulian passive margin where shallow-water, basinal, and shelf-margin facies successions, including fine-grained sediments, occur. Here, we provide a review of the geochemistry of Meso–Cenozoic shales from the Lagonegro basin to elucidate provenance and paleoweathering. The different suites of these shales are dominated by 2:1 clay minerals and are Fe shales and shales. An R-mode factor analysis suggests Ti, Al, and LREE (F1) and K2O-MgO (F2) covariance, likely related to the illite → smectite → kaolinite evolution during weathering. HREE and Y are distributed by phosphate minerals, suggesting LREE/HREE fractionation. The CIA paleoweathering proxy rules out non-steady-state weathering conditions and indicates that the source area was affected by moderate to intense weathering. The paleoprecipitation values derived from the CIA-K and CALMAG indices show median values in the 1214–1610 mm/y range. The Eu/Eu*, Sm/Nd, and Ti/Al provenance ratios point toward a UCC-like source excluding any mafic supply and suggest that the Lagonegro basin was connected, through a southern area, with the African cratonic area. However, the Eu/Eu* median value of the southern Apennine shales is quite similar to the value of the Archean shales, possibly indicating a less differentiated component. This is consistent, in many samples, with the value of the (Gd/Yb)ch ratio, suggesting that the shales likely incorporated ancient sediments derived from African Archean terranes through a cannibalistic process.

1. Introduction

The chemical composition of siliciclastic sedimentary rocks is an important record of the geological evolution of the continental crust through time, because some elements are quantitatively transported in the terrigenous fine fraction [1]. Of these, the rare earth elements (REEs) are the most useful because their distribution is not affected by secondary processes, and the REE patterns of fine-grained siliciclastic sediments and some elemental ratios, especially Eu/Eu*, are assumed to reflect the exposed crustal abundance in the source area [2,3,4,5,6,7,8,9]. However, weathering conditions affect the leaching of elements during the path from the protolith(s) to the sediment(s). Low-field-strength elements have a high affinity for aqueous phases, although some of them can be further partially incorporated into secondary clay minerals (see [10,11], and references therein). Mobile elements can be used to evaluate the degree of chemical weathering [12,13,14,15,16] and to characterize the paleoclimate and paleoprecipitation [17,18,19,20,21,22,23]. Consequently, the distributions of both selected major and trace elements in fine-grained siliciclastic sediments are widely used to constrain the provenance, paleoweathering, and paleoclimate conditions.
The evolution of the geology of the Mediterranean area includes peculiar phenomena, such as continental rifting during the Triassic, oceanic spreading in the Jurassic–Early Cretaceous, the closure of the Tethys from the Late Cretaceous to the Tertiary, and finally, the continental collision between the Adriatic–African plate and the European plate. This work provides a review of the geochemistry of Meso–Cenozoic shales in the southern Apennines (Italy) to elucidate the parental affinity and paleoclimate conditions affecting shale formation to furnish a more comprehensive picture of the geological frame involved. As indicated in Table 1, all data in this paper are already published. The aim of this work was to discuss these data from different perspectives in order to give them new interpretations and to obtain new useful information.

2. Geological Framework

The southern Apennines is a fold-and-thrust belt of Adriatic origin, the formation of which mainly derived from the deformation of the African passive margin (Figure 1). The thrust belt formed during the Oligocene–Pleistocene and involved the early accretionary phases, sedimentary cover, and the ophiolitic suite of the Ligurian Ocean [24,25]. During the Miocene, the wedge consisted mainly of Meso–Cenozoic plate-forming sediments and deep-water environments and Neogene–Pleistocene foredeep deposits.
The paleogeography of the southern Apennines in the Mesozoic is a fundamental building block for understanding the geodynamics of the western Mediterranean [26]. Considering the simplest reconstruction of the pre-Orogenic paleogeography of the southern Apennines, we observe that the African (Apulian) passive margin was characterized by a Meso–Cenozoic pelagic basin (Lagonegro basin) belonging to the Ligurian oceanic domain and lying between two carbonate platforms of the same age (Figure 2) [27,28,29,30].
The pre-Orogenic successions of Lagonegro are composed of siliciclastic, carbonate, and siliceous sediments from the Lower–Middle Triassic to the Oligo–Miocene [28]. The oldest rocks consist of shallow-water siliciclastic sediments, organogenic limestones, and, upwards, of siliciclastic deposits that indicate a progressive deepening of the basin; these rocks make up the Monte Facito Fm (Lower–Middle Triassic). Superimposed on this formation, we observe pelagic succession characterized by predominant carbonate sedimentation until the late Triassic (“Calcari con selce” Fm) that was subsequently replaced by Jurassic siliceous sedimentation (“Scisti silicei” Fm).
A significant increase in tectonic activity along the platform–basin boundaries, which resulted in uplift and erosion of the platform margins between the Late Cretaceous and the Oligocene, was responsible for the deposition of coarse-grained calcareous–clastic sediments interbedded with reddish marls and shales (“Flysch Rosso” Fm) and varicolored shales and marls (“Argille Varicolori” Fm) with the latter accumulating in a depocentral area of the basin [28,29]. Near Campomaggiore and Vaglio di Basilicata, along the front of the southern Apennines chain, there are outcrop successions of “Flysch Rosso” Fm where shales are interbedded with siltstones and radiolarian cherts.

3. Sampling and Analytical Methods

For this study, a total of 52 samples were taken from four different suites, and most of them belonged to the “Flysch Rosso” Formation. Near the localities of Campomaggiore and Vaglio di Basilicata, 22 “Flysch Rosso” samples (FRA suite) were collected in a succession where shales are interbedded with siltstones and radiolarian cherts [31]. Sixteen samples of “Flysch Rosso” shales from boreholes crossing a sequence with minor calcarenite levels (FRB suite) [32] were sampled near the village of Monteverde, in the northeastern sector of the Lucanian Apennines. Finally, 14 samples (VC suite) of shales belonging to the “Argille Varicolori” Fm [3,33] were taken from a borehole crossing a pelitic sequence with rare limestones and marly limestones at the eastern boundary of the southern Apennine chain, near the village of Tolve.
Major elements were estimated by Philips (Netherland) PW 1410 X-ray fluorescence spectrometer on powdered samples using a matrix correction method [34,35,36]. The amount of trace elements, including REEs, was obtained by using ICP-MS at the Centre de Recherche Petrograpique et Geochimique of Vandoeuvre-les-Nancy [37]. The estimated precision and accuracy for trace element determinations were better than 5%, except for elements with a concentration of 10 ppm or lower (10%–15%). Total loss on ignition (LOI) was gravimetrically estimated after overnight heating at 950 °C. Average errors for trace elements were less than ±5%m except for elements present at 10 ppm or lower (±5–10%).
Finally, univariate, and multivariate statistics were performed with STATGRAPHICS 18 software.

4. Results and Discussion

4.1. Mineralogy and Geochemistry

The mineralogical composition of studied samples was quite homogeneous (Table 1), showing that the mineralogy was dominated by 2:1 clay minerals, such as illite and smectite, with the latter probably being derived from the weathering of illite, e.g., [38,39]. Kaolinite generally had a similar abundance to chlorite, whereas quartz and calcite were minor phases.
The results for the major and trace elements’ abundances and fractionation indexes are reported in Table 2; Table 3 and as box and whiskers plots in Figure 3, Figure 4 and Figure 5. Furthermore, The log(SiO2/Al2O3) vs. log(Fe2O3/K2O) classification diagram indicates that the majority of samples can be classified as shales and Fe shales (Figure 6).
Chemical analyses show that the most abundant major oxides are SiO2, (median suite FRA = 63.1 ± 4.39 wt.%; median suite FRB; 50.32 ± 1.22 wt.%; median suite VC = 50.38 ± 1.79 wt.%), Al2O3 (median suite FRA = 16. 52 ± 2.07 wt.%; median suite FRB = 20.71 ± 0.69 wt.%; median suite VC = 25.3 ± 1.94 wt.%), and Fe2O3 (median suite FRA = 7.17 ± 1.93 wt.%; median suite FRB = 9.2 ± 1.23 wt.%; median suite VC = 6.77 ± 1.57 wt.%) (Figure 4). The shales show lower SiO2 and K2O contents and higher Al2O3 and Fe2O3 contents with respect to the Upper Continental Crust (hereafter, the UCC, [1] and, consistently, are classified as Fe shales and shales with very few samples of the FRA subset, which result in wacke and Fe sand [40].
Regarding the trace elements, Ba shows median values of 173.5 ± 55.49 (FRA suite), 336.45 ± 84.97 ppm (FRB suite), and 169.0 ± 155.8 ppm (VC suite). The high value of the Ba standard deviation in the VC suite is related to a single sample (VC13) that shows Ba abundance of 758 ppm and more CaO enrichment with respect to other samples. Y displays median values of 22.5 ± 6.98 ppm, 24.5 ± 3.61 ppm, and 29.5 ± 4.68 ppm, respectively, for the FRA, FRB, and VC suites.
The ΣREE median value is 189.4 ± 40.5, and the chondrite normalized REE patterns (Figure 7) are characterized by a typical shale-like trend with enrichment of LREEs ((La/Yb)ch median = 12.70 ± 2.03), a flat distribution of HREEs ((Gd/Yb)ch median = 1.93 ± 0.27), the absence of Ce anomalies (Ce/Ce* median = 0.95 ± 0.08), and negative Eu anomalies (Eu/Eu* median = 0.70 ± 0.04). It must be noted that the lack of a positive Ce anomaly excludes prevailing intense and long-lasting weathering during subaerial oxidizing environmental conditions [38]. Further details can be found in [3,9,39,40,41,42].

4.2. Factors Affecting the Distribution of Elements

An R-mode factor analysis, classical type of factoring, was performed to evaluate interelemental relationships among the major oxides, Ba, Y, LREE, and HREE. Factors were extracted with the STATGRAPHICS 18 package. This operation was performed using a standardized correlation matrix and thereby weighting all the variables equally during factor calculations. The communalities provide an index of the efficiency of the proposed set of factors [43], and the magnitude of the communalities calculated in this study suggests that most of the original variance is still accounted for by the present set of factors. Three factors explain 91.7% of the total variance in the geochemical database (Table 4).
The first factor (F1; Var.% = 47.7) includes significant positive weightings for Al2O3, TiO2, and LREE and a negative weighting for SiO2. Titanium, Al, and LREE are low-solubility elements characterized by low to very low residence times in oceanic water and seawater to the upper crust partition coefficient [1]. The potential of these elements to be fractionated during the sedimentary process, especially in the source rock(s)–shale transformation, is thus negligible. Further, the LREE covariance with Al2O3 and TiO2 supports the idea that these elements in shales are controlled by clay minerals or occur as detrital tiny crystals in the clayey grain-size fraction [33,40]. Silica in our shales is mainly hosted in 2:1 and 1:1 clay minerals, and its negative weighting can be related to illite → smectite → kaolinite evolution during a more intense weathering stage, also promoting the residual accumulation of elements characterized by lower solubility, such as Ti, Al, and LREE.
The second factor (F2; Var.% = 28.2) has significant and positive weightings for MgO and K2O. In 2:1 clay minerals, which in these shales are the most abundant phases, K+ and Mg2+ occur in the interlayer spaces of illite and smectite, respectively. Further, Mg2+ may also substitute Al3+ in the octahedral sheets. Both of these cations can be released to the soil solutions during the illite → smectite → kaolinite weathering, thus explaining the K2O-MgO covariance associated with F2.
The third factor (F3; Var.% = 15.8) shows significant and positive weightings for P2O5, Y, and HREE. The yttrium geochemical affinity for the HREE is well known, and both may be distributed by phosphate minerals in a variety of sedimentary rocks [44,45] and shales [3,46]. F3, coupled with F1, indicates fractionation between LREE and HREE+Y in the southern Apennines shales, suggesting that an index such as (La/Yb)ch non-necessarily records the source area.

4.3. Paleoweathering

It is well known that alkali and alkaline–earth elements are mobile elements, and their depletion in siliciclastic and residual sediments is generally related to chemical weathering processes, producing a depletion of mobile elements (i.e., Ca, Na, and K) and an enrichment of immobile elements such as Al, and the relationship between mobile and immobile elements was used to reconstruct the weathering conditions affecting the source area. The chemical index of alteration (hereafter, CIA) [12] is the most commonly used index to define broad palaeoweathering conditions in siliciclastic and residual sediments e.g., [3,16,47,48]. The CIA was calculated using the formula (molecular proportions): CIA = [Al2O3/(Al2O3 + CaO* + Na2O + K2O)] × 100, where CaO* is the amount of CaO only regarding silicate minerals. The CIA value for non-weathered upper crustal rocks is approximately 50%, whereas highly weathered residual rocks have CIA values of higher than 90%.
In the analyzed samples, the median CIA values of the three subsets were in the 80–85% range (FRA: CIA = 85 ± 2; FRB: CIA = 80 ± 2; VC: CIA = 85 ± 5), suggesting that southern Apennines shales largely formed through prevailing moderate to intense weathering. Further, the reasonably homogeneous CIA values in the southern Apennines shales appear to rule out non-steady-state weathering conditions, usually related to active tectonism. The A-CN-K diagram (Figure 8) illustrates a composite trend starting from a UCC-like protolith, initially pointing toward an illite composition (FRB subset), and further moving toward the A apex and kaolinite formation (FRA and VC subsets).
The chemical composition of paleosoils can also be used to assess paleoprecipitation (see [49], and references therein), through the climofunctions CIA-K (100 × (Al/(Al + Ca + Na)) [50] and the related mean annual precipitation (MAP(mm/y) = 221.1e0.0197(CIA-K) [16] and CALMAG (100 × (Al/(Al + Ca + Mg)) [20,21], MAP(mm/y) = 22.69(CALMAG)–435.8 (Table 5)). The picture depicted by the CIA-K and CALMAG is consistent with the CIA and A-CN-K diagram involving higher median values for the FRA (CIA-K = 90 ± 2; MAP = 1309 ± 44 mm/y; CALMAG = 87 ± 2; MAP = 1548 ± 51 mm/y) and VC (CIA-K = 90 ± 45; MAP = 1286 ± 111 mm/y; CALMAG = 90 ± 4; MAP = 1610 ± 83 mm/y) subsets than the FRB subset (CIA-K = 87 ± 3. MAP = 1214 ± 76 mm/y; CALMAG = 82 ± 3; MAP = 1444 ± 59 mm/y), suggesting that the latter was affected by drier climatic conditions (Figure 9).

4.4. Provenance

Proxies based on low-mobility elements, affected by minor fractionation during intense weathering and recording chemical differentiation, are used to identify the parent rock(s) of shales [2,3,7,8,10,11,14,15]. Among them, the Eu anomaly is retained as the more conservative provenance proxy [3,10,11,14].
In the southern Apennine shales, the value of the Eu/Eu* is close to the UCC value (Eu/Eu* = 0.65), indicating crustal provenance and ruling out a significant mafic supply. The Sm/Nd ratio shows similar values to Eu/Eu*, reflecting that chemical differentiation and only minor fractionation of Sm and Nd occurs during weathering [51]. Further, the observed covariance between Al2O3 and TiO2 in these shales excludes any fractionation between Ti and Al, suggesting the effectiveness of the Ti/Al ratio as the provenance proxy. In the Eu/Eu* vs. Sm/Nd (Figure 10) and Eu/Eu* vs. Ti/Al diagrams (Figure 11) the shales fall far from the average value for basalt and quite close to the UCC values, thus excluding any mafic provenance. It is worth nothing that the Eu/Eu* median is, with respect to the UCC, slightly higher (0.70 ± 0.04), with a shift toward the value of the Archean shales (Eu/Eu* = 0.73; [52]), suggesting the presence of a less differentiated component. Several shales show (Gd/Yb)ch > 2, which is usually associated with Archean terranes (Figure 12), e.g., [52,53]. Thus, the values of the (Gd/Yb)ch ratio suggest that southern Apennine shales likely incorporated material derived from exposed African Archean terranes or, more likely, ancient sediments derived from African Archean terranes through a cannibalistic process [9] (Figure 12). These findings suggest that the Lagonegro basin was connected through a southern entry area with a continental margin, the African cratonic area, providing siliciclastic detritus according to what was proposed for other siliciclastic sediments of the Lagonegro basin (Figure 13) [41].

5. Conclusions

The different suites of southern Apennine shales largely share similar mineralogical and geochemical features. The samples are dominated by 2:1 clay minerals and are Fe shales and shales showing, with respect to the UCC, lower SiO2 and K2O and higher Al2O3 and Fe2O3 contents. The different factors extracted by R-mode factor analysis indicate Ti, Al, and LREE (F1) and K2O-MgO (F2) covariance, likely related to the illite → smectite → kaolinite evolution during weathering. HREE and Y are distributed by phosphate minerals, suggesting LREE/HREE fractionation. The CIA paleoweathering proxy rules out non-steady-state weathering conditions and indicates that the source area was affected by moderate to intense weathering, which is also consistent with what is depicted by the CIA-K and CALMAG paleoweathering proxies. The paleoprecipitation values derived from the CIA-K and CALMAG indices show median values in the 1214–1610 mm/y range with the FRB subset affected by drier climatic conditions.
The Eu/Eu*, Sm/Nd, and Ti/Al provenance ratios point toward an UCC-like source excluding any mafic supply and suggesting that the Lagonegro Basin was connected, through a southern entry area, with the African cratonic area. However, the Eu/Eu* median value of the southern Apennine shales is almost similar to the value of the Archean shales, possibly indicating a less differentiated component. In addition, in many samples, the (Gd/Yb)ch ratio is >2, indicating that the shales likely incorporated ancient sediments derived from African Archean terranes through a cannibalistic process.

Author Contributions

Conceptualization, R.B., G.R. and G.M.; sampling activity G.M.; laboratory analysis and methodology, G.M.; writing—original draft preparation, R.B., G.R. and G.M.; supervision, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Taylor, S.R.; McLennan, S.M. The Continental Crust: Its Composition and Evolution; Blackwell: Oxford, UK, 1985. [Google Scholar]
  2. McLennan, S.M.; Hemming, S.; McDaniel, D.K.; Hanson, G.N. Geochemical approaches to sedimentation, provenance, and tectonics. Geol. Soc. Am. Spec. Pap. 1993, 284, 21–40. [Google Scholar]
  3. Mongelli, G.; Critelli, S.; Perri, F.; Sonnino, M.; Perrone, V. Sedimentary recycling, provenance and paleoweathering from chemistry and mineralogy of Mesozoic continental red bed mudrocks, Peloritani Mountains, Southern Italy. Geochem. J. 2006, 40, 197–209. [Google Scholar] [CrossRef] [Green Version]
  4. Fedo, C.M.; Eriksson, K.A.; Krogstad, E.J. Geochemistry of shales from the Archean (~3.0 Ga) Buhwa Greenstone Belt, Zimbabwe: Implications for provenance and source-area weathering. Geochim. Cosmochim. Acta 1996, 60, 1751–1763. [Google Scholar] [CrossRef]
  5. Hassan, S.; Ishiga, H.; Roser, B.P.; Do Zen, K.; Naka, T. Geochemistry of Permian–Triassic shales in the Salt Range, Pakistan: Implications for provenance and tectonism at the Gondwana margin. Chem. Geol. 1999, 158, 293–314. [Google Scholar] [CrossRef]
  6. Bauluz, B.; Mayayo, M.J.; Fernandez-Nieto, C.; Gonzales Lopez, J.M. Geochemistry of Precambrian and Paleozoic siliciclastic rocks from the Iberian Range (NE Spain): Implications for source-area weathering, sorting, provenance, and tectonic setting. Chem. Geol. 2000, 168, 135–150. [Google Scholar] [CrossRef]
  7. Cullers, R.L.; Podkovyrov, V.N. Geochemistry of the Mesoproterozoic Lakhanda shales in southeastern Yakutia, Russia: Implications for mineralogical and provenance control, and recycling. Precambrian Res. 2000, 104, 77–93. [Google Scholar] [CrossRef]
  8. Condie, K.C.; Lee, D.; Farmer, G.L. Tectonic setting and provenance of the Neoproterozoic Uinta Mountain and Big Cottonwood groups, northern Utah: Constraints from geochemistry, Nd isotopes, and detrital modes. Sediment. Geol. 2001, 141, 443–464. [Google Scholar] [CrossRef]
  9. Mongelli, G. Rare earth elements in Oligo-Miocenic pelitic sediments from Lagonegro basin, Southern Apennines, Italy: Implication for provenance and source area weathering. Int. J. Earth Sci. 2004, 93, 612–620. [Google Scholar] [CrossRef]
  10. Sinisi, R.; Mongelli, G.; Mameli, P.; Oggiano, G. Did the Variscan relief influence the Permian climate of Mesoeurope? Insights from geochemical and mineralogical proxies from Sardinia (Italy). Palaeogeogr. Palaeoclimatol. Palaeoecol. 2014, 396, 132–154. [Google Scholar] [CrossRef]
  11. Mameli, P.; Mongelli, G.; Sinisi, R.; Oggiano, G. Weathering products of a dismantled Variscan basement. Minero-chemical proxies to insight on Cretaceous palaeogeography and Late Neogene palaeoclimate of Sardinia (Italy). Front. Earth Sci. 2020, 8, 290. [Google Scholar] [CrossRef]
  12. Nesbitt, H.W.; Young, G.M. Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 1982, 299, 715–717. [Google Scholar] [CrossRef]
  13. Harnois, L. The CIW index: A new chemical index of weathering. Sediment. Geol. 1988, 55, 319–322. [Google Scholar] [CrossRef]
  14. Fedo, C.M.; Eriksson, K.A.; Blenkinsop, T.G. Geologic history of the Archean Buhwa Greenstone Belt and surrounding granite–gneiss terrane, Zimbabwe, with implications for the evolution of the Limpopo Belt. Can. J. Earth Sci. 1995, 32, 1977–1990. [Google Scholar] [CrossRef]
  15. Perri, F. Reconstructing chemical weathering during the Lower Mesozoic in the Western-Central Mediterranean area: A review of geochemical proxies. Geol. Mag. 2018, 155, 944–954. [Google Scholar] [CrossRef]
  16. Perri, F. Chemical weathering of crystalline rocks in contrasting climatic conditions using geochemical proxies: An overview. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2020, 556, 109873. [Google Scholar] [CrossRef]
  17. Sheldon, N.D.; Retallack, G.J.; Tanaka, S. Geochemical climofunctions from North American soils and application to paleosols across the Eocene-Oligocene boundary in Oregon. J. Geol. 2002, 110, 687–696. [Google Scholar] [CrossRef] [Green Version]
  18. Rasmussen, C.; Tabor, N.J. Applying a quantitative pedogenic energy model across a range of environmental gradients. Soil Sci. Soc. Am. J. 2007, 71, 1719–1729. [Google Scholar] [CrossRef]
  19. Lukens, W.E.; Nordt, L.C.; Stinchcomb, G.E.; Driese, S.G.; Tubbs, J.D. Reconstructing pH of paleosols using geochemical proxies. J. Geol. 2018, 126, 427–449. [Google Scholar] [CrossRef]
  20. Lukens, W.E.; Stinchcomb, G.E.; Nordt, L.C.; Kahle, D.J.; Driese, S.G.; Tubbs, J.D. Recursive partitioning improves paleosol proxies for rainfall. Am. J. Sci. 2019, 319, 819–845. [Google Scholar] [CrossRef]
  21. Nordt, L.C.; Driese, S.D. New weathering index improves paleorainfall estimates from Vertisols. Geology 2010, 38, 407–410. [Google Scholar] [CrossRef]
  22. Nordt, L.C.; Driese, S.G. A modern soil characterization approach to reconstructing physical and chemical properties of paleo-Vertisols. Am. J. Sci. 2010, 310, 37–64. [Google Scholar] [CrossRef]
  23. Stinchcomb, G.E.; Nordt, L.C.; Driese, S.G.; Lukens, W.E.; Williamson, F.C.; Tubbs, J.D. A data-driven spline model designed to predict paleoclimate using paleosol geochemistry. Am. J. Sci. 2016, 316, 746–777. [Google Scholar] [CrossRef]
  24. Knott, S.D. The Liguride complex of southern Italy—A Cretaceous to Paleogene accretionary wedge. Tectonophysics 1987, 142, 217–226. [Google Scholar] [CrossRef]
  25. Monaco, C.; Tortorici, L. Tectonic role of ophiolite-bearing terranes in the development of the Southern Apennines orogenic belt. Terra Nova 1995, 7, 153–160. [Google Scholar] [CrossRef]
  26. Pescatore, T.; Renda, P.; Schiattarella, M.; Tramutoli, M. Stratigraphic and structural relationships between Meso-Cenozoic Lagonegro basin and coeval carbonate platforms in southern Apennines, Italy. Tectonophysics 1999, 315, 269–286. [Google Scholar] [CrossRef]
  27. Ogniben, L. Schema introduttiva alla geologia del confine Calabro-Lucano. Mem. Soc. Geol. Ital. 1969, 8, 453–763. [Google Scholar]
  28. Mostardini, F.; Merlini, S. Appennino Centro. Meridionale e Proposta di Modello Strutturale. Mem. Soc. Geol. Ital. 1986, 35, 177–202. [Google Scholar]
  29. Pescatore, T.; Renda, P.; Tramutoli, M. Relationship between the Lagonegro and Silicidi formations of the middle valley of the Basento River, Lucca, Southern Apennines. Mem. Soc. Geol. Ital. 1988, 41, 353–361. [Google Scholar]
  30. Marsella, E.; Bally, A.W.; D’Argenio, B.; Cippitelli, G.; Pappone, G. Tectonic history of the Lagonegro Domain and Southern Apennine thrust belt evolution. Tectonophysics 1995, 252, 307–330. [Google Scholar] [CrossRef]
  31. Fiore, S.; Piccarreta, G.; Santaloia, F.; Santarcangelo, R.; Tateo, F. The Flysch Rosso shales from the southern Apennines, Italy. 1. Mineralogy and geochemistry. Per. Mineral. 2000, 69, 63–68. [Google Scholar]
  32. Mongelli, G. Trace elements distribution and mineralogical composition in the <2-μm size fraction of shales from the Southern Apennines, Italy. Mineral. Petrol. 1995, 53, 103–114. [Google Scholar]
  33. Caggianelli, A.; Fiore, S.; Mongelli, G.; Salvemini, A. REE distribution in the clay fraction of pelites from the southern Apennines, Italy. Chem. Geol. 1992, 99, 253–263. [Google Scholar] [CrossRef]
  34. Franzini, M.; Leoni, L.; Saitta, M. A simple method to evaluate the matrix effects in X-ray fluorescence analysis. X-ray Spectrom. 1972, 1, 151–154. [Google Scholar] [CrossRef]
  35. Franzini, M.; Leoni, L.; Saitta, M. Revisione di una metodologia analitica per fluorescenza-X, basata sulla correzione completa degli effetti di matrice Rend. Soc. Ital. Mineral. Petrol. 1975, 31, 365–378. [Google Scholar]
  36. Leoni, L.; Saitta, M. Determination of yttrium and niobium on standard silicate rocks by X-ray fluorescence analyses Rend. Soc. Ital. Mineral. Petrol. 1976, 32, 497–510. [Google Scholar]
  37. Govindaraju, K.; Mevelle, G. Fully automated dissolution and separation methods for inductively coupled plasma atomic emission spectrometry rock analysis. Application to the determination of rare earth elements. Plenary lecture. J. Anal. Atom. Spectrom. 1987, 2, 615–621. [Google Scholar]
  38. Mongelli, G. Ce-anomalies in the textural components of Upper Cretaceous karst bauxites from the Apulian carbonate platform (southern Italy). Chem. Geol. 1997, 140, 69–79. [Google Scholar] [CrossRef]
  39. Herron, M.M. Mineralogy from geochemical well logging. J. Sedim. Petrol. 1986, 58, 820–829. [Google Scholar] [CrossRef]
  40. Fiore, S.; Mongelli, G. Hypothesis on the genesis of day minerals in the fine fraction of «Argille varicolori» from Andretta (southem Apennines). Miner. Petrog. Acta 1991, 34, 183–190. [Google Scholar]
  41. Di Leo, P.; Dinelli, E.; Mongelli, G.; Schiattarella, M. Geology and geochemistry of Jurassic pelagic sediments, Scisti silicei Formation, southern Apennines. Italy Sed. Geol. 2002, 150, 229–246. [Google Scholar] [CrossRef]
  42. Mongelli, G.; Dinelli, E. The geochemistry of shales from the “Frido Unit”, Liguride Complex, Lucanian Apennines, Italy: Implications for provenance and tectonic setting. Ofiotiti 2001, 26, 457–466. [Google Scholar]
  43. Davis, J.C. Statistics and Data Analysis in Geology; John Wiley & Sons: New York, NY, USA, 1986. [Google Scholar]
  44. Ahmadnejad, F.; Mongelli, G. Geology, geochemistry, and genesis of REY minerals of the late Cretaceous karst bauxite deposits, Zagros Simply Folded Belt, SW Iran: Constraints on the ore-forming process. J. Geochem. Explor. 2022, 240, 107030. [Google Scholar] [CrossRef]
  45. Ferhaoui, S.; Kechiched, R.; Bruguier, O.; Sinisi, R.; Kocsis, L.; Mongelli, G.; Bosch, D.; Ameur Zaimeche, O.; Laouar, R. Rare earth elements plus yttrium (REY) in phosphorites from the Tébessa region (Eastern Algeria): Abundance, geochemical distribution through grain size fractions, and economic significance. J. Geochem. Explor. 2022, 241, 107058. [Google Scholar] [CrossRef]
  46. Liu, Y.H.; Lee, D.C.; You, C.F.; Takahata, N.; Iizuka, Y.; Sano, Y.; Zhou, C. In-situ U–Pb dating of monazite, xenotime, and zircon from the Lantian black shales: Time constraints on provenances, deposition and fluid flow events. Precambrian Res. 2020, 349, 105528. [Google Scholar] [CrossRef]
  47. Perri, F.; Critelli, S.; Martín-Martín, M.; Montone, S.; Amendola, U. Unravelling hinterland and offshore palaeogeography from pre-to-syn-orogenic clastic sequences of the Betic Cordillera (Sierra Espuña), Spain. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2017, 468, 52–69. [Google Scholar] [CrossRef] [Green Version]
  48. Abedini, A.; Mongelli, G.; Khosravi, M.; Sinisi, R. Geochemistry and secular trends in the middle–late Permian karst bauxite deposits, northwestern Iran. Ore Geol. Rev. 2020, 124, 103660. [Google Scholar] [CrossRef]
  49. Michel, L.A.; Sheldon, N.D.; Myers, T.S.; Tabor, N.J. Assessment of pretreatment methods on CIA-K and CALMAG indices and the effects on paleoprecipitation estimates. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2022, 601, 111102. [Google Scholar] [CrossRef]
  50. Maynard, J.B. Chemistry of modern soils as a guide to interpreting Precambrian paleosols. J. Geol. 1992, 100, 279–289. [Google Scholar] [CrossRef]
  51. Viers, J.; Wasserburg, G.J. Behavior of Sm and Nd in a lateritic soil profile. Geochim. Et Cosmochim. Acta 2004, 68, 2043–2054. [Google Scholar] [CrossRef]
  52. Condie, K.C. Chemical composition and evolution of the upper continental crust: Contrasting results from surface samples and shales. Chem. Geol. 1993, 104, 1–37. [Google Scholar] [CrossRef]
  53. McLennan, S.M.; Taylor, S.R. Sedimentary rocks, and crustal evolution: Tectonic setting and secular trends. J. Geol. 1991, 99, 1–21. [Google Scholar] [CrossRef]
Figure 1. Geological map of southern Italy with sampling site localization.
Figure 1. Geological map of southern Italy with sampling site localization.
Minerals 13 00994 g001
Figure 2. Paleogeographic evolution of the Lagonegro basin during the Eocene–Oligocene. Acronyms: LU and SU = Ligurian ocean and internal basinal to shelf-margin domains (‘Liguride’ and ‘Sicilide’ units); CLP and APC = Campania-Lucania and Apulian platforms; G = Flysch Galestrino (Lower–Middle Cretaceous); VC = varicoloured clays (middle Cretaceous–Oligocene), FR = Flysch Rosso (Upper Cretaceous–Oligocene). Modified from [26].
Figure 2. Paleogeographic evolution of the Lagonegro basin during the Eocene–Oligocene. Acronyms: LU and SU = Ligurian ocean and internal basinal to shelf-margin domains (‘Liguride’ and ‘Sicilide’ units); CLP and APC = Campania-Lucania and Apulian platforms; G = Flysch Galestrino (Lower–Middle Cretaceous); VC = varicoloured clays (middle Cretaceous–Oligocene), FR = Flysch Rosso (Upper Cretaceous–Oligocene). Modified from [26].
Minerals 13 00994 g002
Figure 3. Box plot for the major oxide abundance (wt.%) in the studied samples. UCC values are taken from [1].
Figure 3. Box plot for the major oxide abundance (wt.%) in the studied samples. UCC values are taken from [1].
Minerals 13 00994 g003
Figure 4. Box plot of the trace element abundance (ppm) in the studied samples. UCC values are taken from [1].
Figure 4. Box plot of the trace element abundance (ppm) in the studied samples. UCC values are taken from [1].
Minerals 13 00994 g004
Figure 5. Box plot for Ce and Eu anomalies and La/Yb and Gd/Yb fractionation indices of the studied samples.
Figure 5. Box plot for Ce and Eu anomalies and La/Yb and Gd/Yb fractionation indices of the studied samples.
Minerals 13 00994 g005
Figure 6. Herron’s classification diagram [39] using log(SiO2/Al2O3)-log(Fe2O3/K2O).
Figure 6. Herron’s classification diagram [39] using log(SiO2/Al2O3)-log(Fe2O3/K2O).
Minerals 13 00994 g006
Figure 7. Chondrite-normalized distribution patterns of REEs in the studied samples.
Figure 7. Chondrite-normalized distribution patterns of REEs in the studied samples.
Minerals 13 00994 g007
Figure 8. A (Al2O3) − CN (CaO* + Na2O) − K (K2O) ternary diagram (molecular proportions; Nesbitt and Young, 1982). The UCC, andesite, and basalt values are taken from [1].
Figure 8. A (Al2O3) − CN (CaO* + Na2O) − K (K2O) ternary diagram (molecular proportions; Nesbitt and Young, 1982). The UCC, andesite, and basalt values are taken from [1].
Minerals 13 00994 g008
Figure 9. Box plot showing the values of CIA-K and CALMAG (mm/y) in the studied shales.
Figure 9. Box plot showing the values of CIA-K and CALMAG (mm/y) in the studied shales.
Minerals 13 00994 g009
Figure 10. Eu/Eu* vs. Ti/Al binary diagram. The UCC and basalt values were taken from [1]; granite values were taken from [52].
Figure 10. Eu/Eu* vs. Ti/Al binary diagram. The UCC and basalt values were taken from [1]; granite values were taken from [52].
Minerals 13 00994 g010
Figure 11. Eu/Eu* vs. Sm/Nd binary diagram. The UCC and basalt values were taken from [1]; granite values were taken from [52].
Figure 11. Eu/Eu* vs. Sm/Nd binary diagram. The UCC and basalt values were taken from [1]; granite values were taken from [52].
Minerals 13 00994 g011
Figure 12. (Gd/Yb)ch vs. Eu/Eu* classification diagram. The fields follow [53]. The UCC values were taken from [1]; granite and shale values were taken from [52].
Figure 12. (Gd/Yb)ch vs. Eu/Eu* classification diagram. The fields follow [53]. The UCC values were taken from [1]; granite and shale values were taken from [52].
Minerals 13 00994 g012
Figure 13. Schematic Eocene–Oligocene palaeogeography of the Apulian margin including the Lagonegro basin, showing the connection to the African cratonic area through a southern entry area. Modified from [41].
Figure 13. Schematic Eocene–Oligocene palaeogeography of the Apulian margin including the Lagonegro basin, showing the connection to the African cratonic area through a southern entry area. Modified from [41].
Minerals 13 00994 g013
Table 1. General features and the main mineralogical composition of the studied sample suites.
Table 1. General features and the main mineralogical composition of the studied sample suites.
LocalityLithologyAgeFormationSample CodeNo. of SamplesMineralogy
Campomaggiore and Vaglio di Basilicata [31]shales interbedded with siltstones and
radiolarian cherts
Late Cretaceous–Medium MioceneRed Flysch Fm.FRA22Illite > Smectite > Kaolinite, Chlorite > Quartz > Calcite
Monteverde
village [32]
shales with
minor calcarenite levels
Late Cretaceous–Medium MioceneRed Flysch Fm.FRB16Smectite > Illite > Kaolinite, Chlorite > Quartz > Calcite
Tolve village [3,33]pelitic sequence with rare limestones and marly limestonesUpper
Cretaceous–Lower Eocene
Argille Varicolori Fm.VC16Illite > Smectite > Kaolinite, Chlorite > Quartz > Calcite
Table 2. Major oxides (wt.%) of studied shale samples.
Table 2. Major oxides (wt.%) of studied shale samples.
SiO2TiO2Al2O3Fe2O3MnOMgOCaONa2OK2OP2O5LOI
FRA363.40.6914.46.030.771.570.541.011.40.0510.1
FRA458.90.7916.68.720.041.440.420.981.50.0710.6
FRA857.90.7716.09.780.061.480.431.071.40.0811.1
FRA10 58.50.8017.86.510.251.470.571.011.40.1011.7
FRA1557.30.7716.510.090.051.370.470.981.40.1011.8
FRA16 67.80.5512.76.360.031.410.440.991.00.118.86
FRA17 57.40.8619.09.00.112.211.130.411.90.066.90
FRA18 68.50.6913.77.990.041.470.361.021.50.085.41
FRA1962.60.8316.79.50.061.490.281.151.30.065.99
FRA20 59.20.9018.79.980.051.520.311.261.70.115.67
FRA21 66.70.6914.95.710.041.440.240.841.10.056.25
FRA22 65.80.7016.66.480.071.680.370.591.10.067.03
FRA23 60.40.8820.16.840.051.570.270.971.50.076.94
FRA24 68.20.6614.87.170.051.570.20.921.20.076.04
FRA25 61.70.8117.88.880.041.680.171.161.50.055.94
FRA26 63.10.7318.26.60.051.860.310.711.30.057.14
FRA27 55.30.8718.28.670.301.320.90.241.40.0712.8
FRA2967.90.5813.15.630.111.470.70.231.40.058.82
FRA3067.20.7015.32.780.021.210.740.281.30.0510.4
FRA3168.90.6513.64.120.091.240.720.251.30.059.10
FRA3263.40.7514.87.580.031.520.810.281.60.129.23
FRA3558.60.8917.95.790.042.20.880.352.00.0711.6
FRB148.80.9421.68.260.052.850.862.322.40.0811.8
FRB2 49.10.8920.612.530.052.80.381.353.10.099.05
FRB349.10.88229.130.042.650.421.762.80.0611.2
FRB451.90.92219.790.033.110.411.613.00.058.13
FRB550.30.7919.910.360.423.650.541.633.30.069.01
FRB650.70.8520.58.970.053.360.52.273.20.069.68
FRB751.60.8319.99.350.063.530.541.653.10.079.30
FRB849.60.8720.49.150.193.330.591.662.90.0711.10
FRB9 52.21.0821.19.140.043.280.511.783.00.107.77
FRB10 51.60.9520.89.260.043.160.631.722.80.098.97
FRB11 50.30.9420.410.860.042.910.541.712.70.099.65
FRB12 50.10.9220.48.620.343.091.162.282.30.0711.20
FRB13 52.21.07228.420.042.920.471.4.02.50.088.97
FRB14 49.70.9720.211.730.032.690.741.852.40.099.65
FRB15 48.51.1220.98.260.152.82.772.222.00.1011.20
FRB1750.51.1321.69.850.042.510.51.972.30.099.49
VC249.91.0624.57.370.091.832.041.522.60.128.99
VC446.41.2128.56.480.031.330.383.131.10.0911.30
VC650.70.9722.99.90.052.420.161.364.00.067.46
VC850.71.3025.96.240.301.910.351.772.40.079.30
VC949.31.0124.59.750.041.750.621.782.10.079.15
VC1252.21.2123.57.130.042.280.981.613.00.097.90
VC1349.41.2223.49.070.042.621.091.823.40.077.83
VC1450.11.5127.84.70.031.380.341.892.20.0610.0
VC1551.91.3827.25.20.031.620.361.412.50.068.39
VC1653.71.1824.66.30.051.860.721.362.50.087.68
VC1853.21.1327.24.960.041.070.350.631.00.0710.3
VC2050.91.1726.36.640.031.530.490.971.90.0510.1
VC2249.21.3828.26.890.041.520.271.121.30.0610.0
VC2349.51.22237.290.042.551.044.03.50.107.71
Table 3. Trace elements (ppm) of the studied shale samples.
Table 3. Trace elements (ppm) of the studied shale samples.
BaYLaCeNdSmEuGdTbDyHoErTmYbLuLREEHREEƩREECe/Ce*Eu/Eu*(La/Yb)cho(Gd/Yb)cho
FRA31512234.876.630.16.41.25.30.84.20.92.40.42.40.4147.917.98165.91.00.659.801.79
FRA41522438.58333.46.81.46.114.40.92.40.42.50.4161.719.52181.21.00.6810.411.98
FRA81412840.78734.88.11.76.8151.22.80.42.50.4170.621.7192.31.00.6911.002.20
FRA10 1722845.798.841.28.81.87.61.25.61.12.90.42.50.4194.523.52218.01.00.6712.352.46
FRA15160314593.240.28.81.97.81.25.91.23.10.42.80.4187.224.68211.91.00.6910.862.26
FRA16 1183132.288.234.18.71.97.81.25.41.12.70.42.40.4163.223.22186.41.20.709.072.63
FRA17 178223675.230.46.11.45.1-4.3-2.5-2.20.4147.715.88163.61.00.7410.911.86
FRA18 1902332.766.729.76.31.45.2-4.3-2.8 2.10.4135.416.12151.51.00.7310.331.95
FRA192402134.673.529.86.11.34.9-4.2-2.3 2.10.414415.12159.11.00.7411.031.85
FRA20 1853842.285.237.78.51.97.2-5.9-3-2.70.5173.621.07194.70.90.7310.762.21
FRA21 3251930.865.526.451.14-3.5 1.9 1.80.3127.712.59140.31.00.7211.311.76
FRA22 2042233.974.128.75.81.24.7-3.9-2.2-1.90.3142.514.26156.81.00.6911.811.98
FRA23 1652842.991.636.47.11.65.4-4.7-2.5 2.30.417816.83194.81.00.7812.391.86
FRA24 1751530.374.627.95.81.35-4 2.5-2.10.4138.615.22153.81.10.729.751.91
FRA25 1781233.171.328.45.61.24.3-4-2.4-2.20.4138.414.36152.81.00.7210.211.58
FRA26 1262231.270.125.35.114.2-3.4-2.1-1.80.4131.712.82144.521.070.6511.461.85
FRA27 3242640.983.635.78.11.87.114.812.40.42.40.4168.321.18189.480.960.7111.522.40
FRA292051928.464.324.9514.60.73.20.720.31.90.3122.614.74137.341.060.6610.101.96
FRA301422233.673.131.361.25.60.84.30.92.40.32.40.414418.24162.241.010.639.461.89
FRA311942132.768.3316.21.35.20.83.80.82.30.42.60.4138.217.54155.740.970.698.501.62
FRA321274341.190.9419.92.3101.57.41.53.80.53.40.5182.931.02213.921.010.718.172.41
FRA351262838.988.633.57.41.66.40.8512.60.42.60.4168.420.84189.241.070.7110.111.99
FRB136525.145.288.136.17.721.66.4-4.4-2.6-2.20.4177.217.57194.760.930.6914.022.36
FRB2 34124.145.791.236.17.631.66.3-4.2-2.6-2.10.4180.617.17197.770.950.6914.652.42
FRB335220.247.395.235.36.871.45.33.23.7-2.4-20.4184.618.32202.920.970.6816.052.16
FRB429719.145.491.937.47.881.662.64.2-2.4-2.10.4182.619.27201.830.960.7014.542.31
FRB537521.742.583.228.76.11.24.7-3.6-2.4-2.20.4160.514.47174.920.960.6713.371.77
FRB651417.140.678.428.35.71.24.82.83.5-2.2-2.10.415316.96169.930.940.7113.001.82
FRB748524.936.173.126.35.21.24.1-3.5-1.8-20.3140.712.96153.670.980.7611.961.64
FRB837420.640.880.329.16.031.34.9-3.7-2.2-20.4156.214.4170.620.960.7213.581.95
FRB9 33228.244.887.533.67.151.55.8-4.3-2.5-2.20.4173.116.81189.870.940.7213.532.10
FRB10 30825.642.781.832.66.641.45.62.94.1-2.3-2.20.4163.718.86182.520.920.7113.292.07
FRB11 29524.942.582.832.36.781.55.6-4.3-2.4-2.20.4164.316.32180.630.940.7413.172.07
FRB12 44224.140.478.730.86.451.45.3-4-2.3-20.4156.415.4171.770.940.7213.652.16
FRB13 29626.541.676.328.75.531.23.8-3.6-1.8-20.3152.212.75164.930.900.8013.771.52
FRB14 21223.843.183.233.76.821.55.72.54.2-2.4-2.10.4166.818.7185.510.930.7513.792.17
FRB15 24527.846.987.335.17.381.66.1-4.7-2.6-2.40.4176.717.74194.410.900.7413.322.06
FRB172353147.992.938.97.791.86.2-5-2.5-2.40.4187.418.24205.650.920.7813.252.07
VC21483551.399.640.58.291.76.80.85.1-3-2.50.4199.720.31220.030.930.6913.992.21
VC41083455.511345.58.851.86.80.95.6-3.1-30.5222.721.67244.320.970.6912.371.81
VC61982254.110742.77.331.45.60.84.2-2.6-2.70.5210.617.71228.350.940.6413.401.66
VC81982962.811847.69.231.76.40.75-2.9-2.80.5237.319.97257.30.900.6615.311.87
VC91672047.685.631.15.961.351.24.1-2.3-2.20.4170.216.52186.740.890.7414.341.79
VC122353357.6103469.281.87.30.95.7-3.2-2.90.521622.33238.370.850.6813.622.08
VC137583374.713857.910.61.97.50.95.8-3.3-3.40.6281.623.27304.880.890.6415.031.81
VC141453180.913662.4111.86.80.75.8-3.4-3.30.6290.422.41312.820.810.6516.571.67
VC152372875.713059.310.41.76.30.75.3-3.3-3.20.5274.921295.870.820.6516.231.63
VC161513060.197.745.18.361.66.30.95.1-3.1-2.80.5211.320.36231.650.780.6714.401.82
VC181302844.878.737.27.051.45.40.74.6-2.6-2.60.5167.817.8185.580.830.7011.561.66
VC202202456.410642.98.051.55.414.7-2.9-2.60.521318.53231.570.900.6814.901.72
VC22143315299.141.27.131.44.90.44.7-2.8-2.90.5199.317.49216.780.910.7012.231.38
VC231712266.71175410.31.97.60.86.2-3.7-3.40.6248.324.15272.490.840.6613.251.80
Note: Eu/Eu* = [Eun/√(Smn · Gdn)]; Ce/Ce* = [Cen/√(Lan · Prn)].
Table 4. Factor Loading Matrix After Varimax Rotation.
Table 4. Factor Loading Matrix After Varimax Rotation.
FactorFactorFactor
123
SiO2−0.72
TiO20.96
Al2O30.93
Fe2O3
MgO 0.98
CaO
Na2O
K2O 0.77
P2O5 0.93
Ba
Y 0.77
LREE0.82
HREE 0.72
Var.%47.728.215.8
Note: The numbers are the weights of the variables in the extracted factors. Variables with weights of less than 0.70 were omitted.
Table 5. CIA, CIA-K, and CALMAG values.
Table 5. CIA, CIA-K, and CALMAG values.
SamplesCIA (%)CIA-K (%)MAP-CIA-K (mm/year)CALMAG (%)MAP-CALMAG (mm/year)
FRA383871221.76851495.36
FRA485901282.73881566.31
FRA884881256.46881550.73
FRA10 86891271.69881559.37
FRA1585891273.24881566.33
FRA16 84861208.15851497.06
FRA17 85901295.90821434.69
FRA18 83881238.97861524.32
FRA1986891277.94891581.58
FRA20 85901283.48901599.20
FRA21 87911316.40881568.44
FRA22 89931363.13871545.49
FRA23 88921349.32901613.35
FRA24 86901307.39881556.48
FRA25 86911309.60891589.45
FRA26 89931367.47881554.74
FRA27 88921350.03871537.42
FRA2985911325.64831455.59
FRA3087921338.12871528.21
FRA3185911323.49851493.66
FRA3284911317.59841469.72
FRA3585911331.75831445.00
median85911308.50871548.11
st.dev.1.681.7443.992.2350.63
FRB179831129.91831448.89
FRB2 81901283.93851487.88
FRB382881242.23861515.03
FRB481881249.98841462.64
FRB578871216.99801385.25
FRB678841154.25821423.71
FRB779871214.73811395.72
FRB880871214.68821416.73
FRB9 80871218.04831439.33
FRB10 80861207.55821432.92
FRB11 81871213.38831456.84
FRB12 78811086.14801379.44
FRB13 84891281.11851487.09
FRB14 80851170.90831452.22
FRB15 7575970.04751267.54
FRB1782861203.67861513.96
median80871214.03831444.11
st.dev.1.933.4375.882.6058.94
VC280831136.43831458.25
VC486851181.96931682.57
VC681921334.40891572.55
VC885901289.39911622.88
VC985881245.81901598.06
VC1281871215.74861508.73
VC1379851181.38841469.89
VC1486901293.68931678.67
VC1586921338.78921653.84
VC1684901282.85891579.69
VC1893951435.72941700.33
VC2089931369.53921642.69
VC2291941389.33931676.12
VC237377996.64841475.38
median85901286.12901610.47
st.dev.5.104.67111.233.6582.78
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Buccione, R.; Rizzo, G.; Mongelli, G. Geochemistry as a Clue for Paleoweathering and Provenance of Southern Apennines Shales (Italy): A Review. Minerals 2023, 13, 994. https://doi.org/10.3390/min13080994

AMA Style

Buccione R, Rizzo G, Mongelli G. Geochemistry as a Clue for Paleoweathering and Provenance of Southern Apennines Shales (Italy): A Review. Minerals. 2023; 13(8):994. https://doi.org/10.3390/min13080994

Chicago/Turabian Style

Buccione, Roberto, Giovanna Rizzo, and Giovanni Mongelli. 2023. "Geochemistry as a Clue for Paleoweathering and Provenance of Southern Apennines Shales (Italy): A Review" Minerals 13, no. 8: 994. https://doi.org/10.3390/min13080994

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop