Basaltic Residual Soil-Waroszewski2019

You might also like

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

Geoderma 337 (2019) 983–997

Contents lists available at ScienceDirect

Geoderma
journal homepage: www.elsevier.com/locate/geoderma

Tracking textural, mineralogical and geochemical signatures in soils T


developed from basalt-derived materials covered with loess sediments (SW
Poland)

Jaroslaw Waroszewskia, , Tobias Sprafkeb, Cezary Kabałaa, Mirosław Kobierskic, Jakub Kierczakd,
Elżbieta Musztyfagaa, Aleksandra Lobaa, Ryszard Mazureke, Beata Łabaza
a
Wroclaw University of Environmental and Life Sciences, Institute of Soil Science and Environmental Protection, Grunwaldzka 53, 50-357 Wroclaw, Poland
b
University of Bern, Institute of Geography, Hallerstrasse 12, 3012 Bern, Switzerland
c
University of Science and Technology, Department of Soil Science and Soil Protection, Bernardyńska 6, 85-029 Bydgoszcz, Poland
d
University of Wroclaw, Institute of Geological Sciences, Cybulskiego 32, 50-205 Wroclaw, Poland
e
Institute of Soil Science and Agrophysics, Department of Soil Science and Soil Protection, Agricultural University of Krakow, Mickiewicza 21,31-120 Krakow, Poland

A R T I C LE I N FO A B S T R A C T

Handling Editor: Michael Vepraskas This study investigates the scale of inheritance of signatures typical of loess- and basalt-derived substrates in soils
Keywords: having both components present as parent material due to past periglacial processes. Based on field description,
Aeolian silts particle size distribution, mineralogy (heavy minerals and clay minerals) as well as geochemistry we track the
Basalt signal of both materials (local basalt-derived and loess-derived), estimate the maximum thickness of loess
Zr and Hf mantles and mixed zones (having both loess and basaltic components), and note the presence or absence of basal
Heavy minerals layers without allochtonous additions. The results show that geochemistry and heavy minerals are the most
Clay minerals reliable proxies for a verification of aeolian silt contributions and to define individual layers in complex soil
Pleistocene
parent materials.
Soil formation in the studied landscape depends on the type of slope sediments. Typical fine-grained
weathering of basalt without input from aeolian silt would promote rather vertic development. However, in thick
loess deposits as well as in thin silt-textured materials superimposed on clay loam beds, clay illuviation dom-
inates. Development of argic horizons, however, results in decreasing permeability, promoting water stagnation
and enhanced degradation processes in clay-rich subsoil. Instead of Luvisols, Stagnosols and Retisols developed
in the study regions. Basaltic block covers mixed with loess host cambic horizons. Based on our findings it seems
that at the edge of thin loess deposits no soils exist that developed exclusively from basaltic parent materials.

1. Introduction interrupted by erosional or depositional events (Boxleitner et al., 2017;


Gild et al., 2018; Stanley and Schaetzl, 2011) that strongly influence
Weathering processes lead to mechanical disintegration of rocks and observed weathering rates and further mineral transformations (Portes
provide suitable parent material for further pedogenetic processes (Egli et al., 2018).
et al., 2010; Kirkby, 2018; Norton et al., 2014) as well as determine the Basalt weathering in the temperate zone and in the tropics is widely
transformation of primary minerals into secondary phases (Egli et al., characterized in the literature (White and Blum, 1995; White and
2014, 2008a). Weathering intensity depends mostly on the suscept- Brantley, 2003; Soubrand-Colin et al., 2007; Soubrand-Colin et al.,
ibility of particular rock in terms of this process (Caner et al., 2014; 2009; Rasmussen et al., 2010). Basalts reveal patterns typical of mafic
Juilleret et al., 2016; Soubrand-Colin et al., 2005), time of exposure substrates, i.e. a relatively high share of MgO, Fe2O3, Al2O3, CaO and
(Egli et al., 2011), climate (Bojko and Kabala, 2016; Egli et al., 2010, undersaturation with SiO2. From a mineralogical and geochemical
2008b; Ponge et al., 2014) and type of vegetation (Egli et al., 2008b). In point of view, basalts contain easily weatherable minerals such as oli-
theory, weathering is a continuous in situ process that delivers weath- vine and pyroxenes, but are also important sources of Ti from ulvos-
ering products of local material which are transformed into soil by pinels, magnetite or titanomagnetite (Soubrand-Colin et al., 2009). The
further pedogenic processes. Sometimes, however, weathering can be clay fraction of soils developed from basalts is dominated by smectites


Corresponding author.
E-mail address: jaroslaw.waroszewski@upwr.edu.pl (J. Waroszewski).

https://doi.org/10.1016/j.geoderma.2018.11.008
Received 31 July 2018; Received in revised form 30 October 2018; Accepted 4 November 2018
Available online 23 November 2018
0016-7061/ © 2018 Elsevier B.V. All rights reserved.
J. Waroszewski et al. Geoderma 337 (2019) 983–997

in temperate zones (Bogda et al., 1998), while in tropical regions the Oligocene (25–31 Ma) and Miocene (18–22 Ma) due to active vol-
gibbsite and kaolinite prevail (Naidu et al., 1987; Noack et al., 1993; He canic processes (Birkenmajer et al., 2004) and later in the Pleistocene
et al., 2008; Lu et al., 2008; Morrison and Bonato, 2015). Usually soils exposed by intensive and selective denudation (Birkenmajer, 1967).
formed from basalt have a clay or clay loam texture (Soubrand-Colin The morphology of basaltic hills is characterized by the presence of
et al., 2007). rock outcrops (tors) and block covers (with occasionally smaller clasts)
The apparent insight into the mineralogy and geochemistry of ba- at the hilltops and the upper-slope positions, while, in the mid-slope
salt-derived weathering products and soils strives from studies of deep sections, heterogeneous covers were identified which were composed
weathering profiles and/or locations that were not obviously inter- from basaltic clasts mixed with sand and silt (Migoń et al., 2002). Lower
rupted by episodic geomorphic processes. Studies that quantify rates of Silesian basalts belong to the Central European volcanic province,
soil production usually attribute textural, morphological and geo- which reaches reaching from France through Germany and the Czech
chemical gradients or discontinuities in soil profiles to bioturbation Republic into Poland (Alibert et al., 1987). In our study, we use the
and/or soil creep (Johnson, 1993; Johnson et al., 2005), but also to term basalt as a common name for all mafic rocks having typical
incorporation of volcanic ash (e.g., Naidu et al., 1987; Morrison, 1991) aphanitic structure.
or the contribution of quarzitic components (Rex et al., 1969). There- Both study areas were completely covered by the Scandinavian ice
fore, in the areas influenced by glaciations or periglacial conditions the sheet during the Odra (Saalian, MIS 8/6) glaciation, while thick loess
common presence of slope cover beds with variable amounts of aeolian covers were deposited between MIS 4 and MIS 2 (Jary, 1999; Badura
silt (Kleber, 1997; Terhorst, 2000; Terhorst et al., 2013; Waroszewski et al., 2013). The presumed source of the loess in these areas is the
et al., 2018a) leads to the assumption that the complex or mixed slope Great Odra glacial valley, which was, between MIS4 and MIS3, supplied
covers may prevail in these zones also may prevail in basalts. by numerous non-glacial rivers, which delivered fine-textured weath-
More than 300 basaltic outcrops are identified in the Lower Silesia; ering material from the Sudeten Mts (Badura et al., 2013). The strati-
however, pure basalt-derived soils have not been discovered in the re- graphy of loess and loess derivates is not well-constrained especially in
gion to date (Kabala et al., 2015). All basalt outcrops in the region are the case of the Lubań region (Kida, 1999). The age of the oldest loess
located near (or are surrounded by) loess sheets; therefore, a significant sections (with features of colluviation) near Złotoryja, are estimated to
influence of aeolian silt to soils from basalt-derived materials is ex- be from the Wartha (Saalian II) stage, while the uppermost loess sec-
pected. Being dominated by quartz and feldspar, aeolian silts also tions (periglacially transformed) are considered to be Weichselian loess
contain carbonate, as well as a distinct suite of clays, e.g., mica-hydroxy (Raczkowski, 1969). Glacial deposits were displaced by erosion and
interlayered smectite (HIS), hydroxyinterlayered vermiculite (HIV), filled mostly valleys or remains (mostly fluvioglacial sands) in lower
chlorite) and heavy minerals (e.g., rutile, zircon). These components slope positions often mixed with basalt-derived slope sediments (Migoń
apparently differ from the minerals typically present in basalt and its et al., 2002).
regolith and may influence greatly the weathering and soil forming The climate in the Lower Silesia region is temperate, as it is in the
processes as well as the final properties of the soils. Still, a field iden- transitional zone between oceanic and continental climates. Mean an-
tification of aeolian substrates mixed with local bedrock (e.g., igneous, nual precipitation ranges from 800 to 900 mm in the Izerskie Upland to
mafic derivates) can be very complicated (e.g. Küfmann, 2003; around 700–750 mm (Pawlak, 2008). Mean annual air temperatures are
Waroszewski et al., 2018a); therefore, combined geochemical and mi- between 6 °C (Izerskie Upland) and 8 °C (Kaczawa Upland).
neralogical approaches could help in such differentiations (Muhs and Soils in the southern part of Lower Silesia developed predominantly
Benedict, 2006; Waroszewski et al., 2018a). This is necessary, as the developed from loess and loess derivates, and commonly have subsoil
sole presence of silt-sized particles can also be associated with weath- clay accumulation, which is typical for Luvisols (Kabala et al., 2015) or
ering (e.g., Szymański and Szkaradek, 2017). have humus-rich topsoil that fulfills the criteria of Chernozems and
In the study of parent materials and the direction of soil-forming Phaeozems (Labaz et al., 2018; Labaz and Kabala, 2016). Soils formed
processes on the slopes of Mt Sleza, Waroszewski et al. (2018a) dif- on consolidated ultramafic/mafic rocks (such as basalt, serpentine or
ferentiated aeolian mantles (practically free of coarse fractions and gabbro) are usually classified as Cambisols or Leptosols (Weber, 1982;
predominated by aeolian silt) from the mixed zones that contain both Pędziwiatr et al., 2018).
local and aeolian materials and the strata that are solely of local origin The forest vegetation in the Lubań transect is dominated by beech
(granite, gabbro, serpentinite regoliths). This study applies this concept trees (with the addition of birch and oak) in the upper- and mid-slope
to soils developed on basalt bedrock, where the deciphering of the sections (sites L1–L3), while, in the footslopes, spruce prevails locally
origin of slope deposits in the field is often complicated by high en- (site L4). In the Muchów transect, the forest soils (sites M1–M3) are
richment of the local weathering products with silt fraction. The key covered with under-mixed forests with oak, hornbeam and spruce. The
questions of this paper are as follows: (1) Are the geochemical and sites M4 and M5 are located on grass vegetation, mostly by Alopecurus
mineralogical signatures of aeolian silt and basalt derived-materials pratensis L. and Festuca rubra L.
recognizable; and do they indicate predomination of a particular
component? (2) Do different proxies (morphological, textural, miner- 3. Methods
alogical and geochemical) agree in the detection of lithological
boundaries within the soil profiles and specifically in detecting the 3.1. Soil sampling
maximum depth of the aeolian contribution? Additional issues are (1)
to track the direction of soil-forming processes in such a heterogeneous Nine soil profiles were studied in two separate slope catenas
environment (transformed by Pleistocene glaciations and loess deposi- (Table 1) reaching over basaltic bedrock. The catena Muchów with NW
tion) and (2) identification of the probable source (provenance) of the exposition consist of five soil pits (M1–5), while the transect in Lubań
loess covering basalt substrates. tends to SW and consists of four profiles (L1–4). The location and ar-
rangements of the soil catenas were selected to provide an insight into
2. Study area different configurations of loess deposits and mixed zones. Each soil pit
was dug until the underlying bedrock was reached or no changes in the
Two soil catenas were chosen for this study, one in the Kaczawa lithology were observed down to the depth of 150 cm. Soil layers and
Upland (Muchów transect) and the other in the Izerskie Upland (Lubań genetic horizons were distinguished and described according to the
transect) in the central and western parts of the Lower Silesia province Guidelines For Soil Description (FAO, 2006). Soil horizons were de-
(SW Poland), respectively (Fig. 1). Catenas were located on basalt re- scribed considering their depth, colour, texture, structure and abun-
sidual hills (deeper parts of volcanic chimneys) formed at the turn of dance of rock fragments. We paid special attention to the identification

984
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 1. Location of the two studied transects. A: Location of the two study areas Luban (B) and Muchow (C) within Lower Silesia. Note the mapped loess distribution
of Lower Silesia (Waroszewski et al., 2018a) within the European loess belt (map of loess sediments in Central Europe adapted from Sprafke and Obreht, 2016, after
Haase et al., 2007).

of lithological discontinuities and the occurrence of mixed zones be- surface and basal horizons to evaluate the presence of basalts.
tween loess and underlying substrates. Soils were classified according to
the FAO-WRB system (IUSS Working Group WRB, 2015). Approxi- 3.2. Description of rock thin sections
mately 1–2 kg of soil material was taken from each distinguished soil
horizon. Samples were air-dried, manually ground, and passed through Petrographic descriptions of polished thin sections of dominant
a 2 mm sieve. Additionally, rock fragments were taken from several rocks were carried out using a Leica DM2500P optical microscope in

Table 1
Main characteristics of studied sites.
Profile Latitude Elevation Slope inclination Slope position Parent material Land use WRB (IUSS Working Group 2015)
(m a.s.l.) (°)

M1 N 51°01′4.2″ 427 12 Shoulder/back Loess/basalt block Forest Eutric Amphiskeletic Folic Cambisol (Anosiltic,
E 16°01′50.6″ slope covers Ochric)
M2 N 51°01′10″ 402 8 Back slope Loess/basalt slope bed Forest Eutric Luvic Albic Folic Stagnosol (Pantosiltic,
E 16°01′40″ Ochric,)
M3 N 51°01′13″ 396 4 Back slope Loess/basalt slope bed Forest Eutric Luvic Albic Stagnosol (Anosiltic, Ochric,
E 16°01′34″ Raptic, Endoskeletic)
M4 N 51°01′16.5″ 390 3 Foot slope Loess/basalt slope bed Meadow Eutric Luvic Albic Glossic Stagnosol (Episiltic,
E 16°01′32.3″ Endoloamic, Aric, Ochric, Raptic, Endoskeletic)
M5 N 51°01′19″ 389 2 Foot slope Loess/basalt slope bed/ Arable land Eutric Luvic Albic Glossic Stagnosol (Episiltic,
E 16°01′31″ glacial materials Amphiloamic, Aric, Ochric, Raptic, Endoskeletic)
L1 N 51°03′41.7″ 376 15 Shoulder Loess/basalt block Forest Dystric Amphiskeletic Folic Cambisol (Anosiltic,
E 15°14′05.7″ covers Ochric)
L2 N 51°03′34.9″ 347 7 Back slope Loess/basalt block Forest Eutric Amphiskeletic Stagnic Cambisol (Anosiltic,
E 15°14′05.4″ covers Ochric)
L3 N 51°03′30.4″ 322 5 Foot slope Loess/basalt slope bed Forest Eutric Luvic Albic Stagnosol (Episiltic, Loamic,
E 15°14′03.8″ Raptic, Katoskeletic)
L4 N 51°02′58.7″ 307 2 Toe slope Loess Forest Dystric Albic Stagnic Retisol (Anosiltic, Cutanic,
E 15°13′10.9″ Ochric)

985
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Table 2
Morphology and basic properties of studied soils (in distinguished layers).
Profile Layers Depth (cm) Hosted horizons Texture in fine earths Rock fragments abundance (%) Clay skins CaCO3 (%)

M1 Mixed zone 0–80 A, Bw1, Bw2 SiL 15–90 − −


M2 Loess mantle 0–52 Ah, Eg SiL 10–15 − −
Mixed zone 52–110 Eg/Btg, 2Btg, 2CBg SiL, CL 50–80 + −
M3 Loess mantle 0–28 Ah, AEg, Eg SiL 10–15 − −
Mixed zone 28–100 Eg/2Btg1, 2Btg1, 2Btg2, 3CBg SiL, CL 30–80 ++ −
M4 Loess mantle 0–54 AEg, Eg/2Btg SiL 5–20 − −
Mixed zone 54–110 2Btg, 3CBg L 50–60 +++ −
M5 Loess mantle 0–32 Apg, Eg SiL 5–15 − −
Fluvioglacial material 32–100 Eg/2Btg, 2Btg/Eg, 2Btg, 3BCg SCL, SL 30–70 ++ −
L1 Mixed zone 0–80 AB, Bw1, Bw2, BC SiL 25–90 − −
L2 Mixed zone 0–75 Ah, Bw1, Bw2, BC SiL 80–90 − −
L3 Loess mantle 0–35 Ah, AE, E, EBtg SiL 10–15 − −
Mixed zone 35–90 Btg1, 2Btg2, 2Btg2, 3BCg2 SiL, L, CL 55–90 +++ −
L4 Loess mantle 0–140 Eg, Eg/Btg, Btg1/Eg, Btg2/Eg, BC SiL – ++ −

transmitted and reflected light. The modal composition of the studied analysed at 25 °C followed by ethylene glycol solvation. The air-dried K-
rocks was estimated by point counting using the freeware software saturated samples were analysed at 25 °C and then heated at 300 °C and
JMicroVision v1.2.7. 550 °C.
Clay minerals were identified based on the positions of basal peaks
3.3. Basic analyses (Brown and Brindley, 1980; Brindley, 1980) and the clay mineral
proportions were estimated from the EG solvated pattern. Semi-quan-
Soil pH was potentiometrically measured in a 1:2.5 (soil: distilled titative estimation of the clay mineral contribution was derived from
water) suspension. Soil organic carbon (SOC) content was determined integrated peak areas (Barnhisel and Bertsch 1989). The XRD patterns
by the dry combustion method using CO2 spectroscopic measurements were processed using the ORIGIN software.
(Ströhlein CS-mat 5500 analyser). After sample pretreatment with 3%
H2O2 and dispersion with hexametaphosphate, particle-size distribution 3.6. Heavy minerals preparation and identification
was performed using sieves (sand fraction) and the hydrometer method
for the silt and clay fraction (Van Reeuwijk, 2002). The heavy mineral distribution of selected soil horizons were
identified using the separation technique of Mange and Maurer (1992)
3.4. Geochemistry and determined in the 0.25–0.1 mm sand fractions. At least 300 grains
were separated using sodium polytungstate solution (2.97 g·cm−3).
The contents of major and trace elements (including REE) were Heavy minerals were identified using JEOL JSM-IT100 coupled with an
obtained using ICP-ES and ICP-MS after sample fusion with lithium Energy Dispersive Spectrometer (EDS) at high vacuum, 15–20 kV
borate and an alloy dissolution with nitric acid (ACME Labs, Canada). working mode and 40 s counting time in polished thin sections. Heavy
The analytical reproducibility (2-sigma), as estimated from replicate minerals were identified based on their geochemical compositions and
analyses of eight randomly chosen samples ranges from 0.3% (SiO2) to the spectra collection from Reed (2010).
11% (Zr) at 95% confidence limits. Analytical accuracy (2-sigma), as
estimated from four measurements of the reference sample STD SO-19 4. Results
is from 0.2 (MgO) to 4.3% (Hf) at 95% confidence limits. To evaluate
the aeolian silt contribution to soils and distinguished loess components 4.1. Identification of coarse fragments
from basalt- derived materials, we focused on relatively immobile ele-
ments (Ti, Hf, Zr,) that having relatively high ionic potential and are The rock samples from the Lubań transect have a porphyritic texture
considered as chemically immobile in most of the near-surface en- with subhedral to euhedral olivine and pyroxene phenocrysts of
vironments (Hutton, 1977; McLennan, 1989, 2001; Muhs and Benedict, average diameter ranging from 0.5 to 2 mm. The largest phenocrysts
2006). In pedogenesis studies, these elements are considered to be reach up to 3 mm in diameter. Large and spongy central parts of pyr-
markers of aeolian silt contribution (Scheib et al., 2014, Waroszewski oxene phenocrysts are surrounded by clear rims (Fig. 3a). Olivine
et al., 2018a). phenocrysts are locally zoned and partly or completely replaced by
orange-brown iddingsite (Fig. 3a). The groundmass of the rock is
3.5. Clay mineralogy composed of clinopyroxene, plagioclase and nepheline. Individual
grains are anhedral to subhedral and reach up to 0.05 mm in size.
The mineralogical composition of the clay fraction from selected Furthermore, euhedral crystals of opaque minerals (probably magne-
soil horizons was analysed by X-ray diffraction (XRD) using an X-Pert tite-ulvospinel) up to 0.05 mm in size are evenly distributed evenly in
PRO diffractometer with the Empyrean XRD tube Cu LFF DK 303072; the rock groundmass (Fig. 3a, b). Based on microscopic observations
and the X'celerator detector-type RTMS and the goniometer PW 3050/ and previous research by Puziewicz et al. (2011), the rocks at Lubań
60. CuKα radiation was used with an applied voltage of 40 kV and classify as basalts (basanite or nephelinite).
current of 20 mA. Oriented mounts of clay fraction (< 2.0 μm) were At Muchów, the rock fragments have a porphyritic texture with
scanned from 2 to 30 2θ at a counting time of 30 s per 0.02 2θ step. For subhedral to euhedral olivine phenocrysts of average diameter ranging
XRD analysis, the clay fraction was obtained according to the procedure from 0.25 to 0.5 mm. The largest phenocrysts reach up to 2 mm in
of Jackson (1969), including the organic matter and free iron oxides diameter. The phenocrysts are partly or completely replaced by orange-
removal (using 30% H2O2 and a citrate-bicarbonate-dithionite buffer, brown iddingsite and set in a fine-grained groundmass composed of
respectively). The samples were dispersed with Na-Amberlite. Clay unoriented, subhedral pyroxene crystals; anhedral plagioclase and/or
samples were saturated with Mg and K and prepared on glass slides for feldspathoids and euhedral to subhedral black opaque minerals (Fig. 3c,
X-ray diffraction analysis. The air-dried Mg-saturated samples were d). The average diameter or length of groundmass minerals is less than

986
J. Waroszewski et al. Geoderma 337 (2019) 983–997

(caption on next page)


987
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 2. Photos and sketches showing the detailed morphology of all presented soils profiles. The lithological boundaries are drawn according to the findings in this
study.

Table 3
Mean values of major and selected trace elements (Hf and Zr) in distinguished layers.
Profile Layers SiO2 Al2O3 Fe2O3 MgO CaO Na2O K2 O TiO2 P2O5 MnO Hf Zr

% ppm

M1 Mixed zone 75.8 8.45 3.62 0.90 0.70 0.82 1.77 0.87 0.07 0.08 10.8 430.0
M2 Loess mantle 79.2 8.58 2.99 0.57 0.50 0.89 2.19 0.81 0.06 0.03 12.4 493.1
Mixed zone 69.6 11.72 5.63 1.11 0.76 0.79 2.07 1.09 0.10 0.09 12.0 460.0
M3 Loess mantle 81.1 7.28 1.95 0.35 0.42 0.87 2.06 0.80 0.06 0.02 13.6 516.3
Mixed zone 72.4 10.88 4.89 0.96 0.66 0.82 2.12 0.94 0.07 0.06 11.1 424.6
M4 Loess mantle 78.5 8.57 2.60 0.48 0.63 0.88 2.21 0.82 0.10 0.06 12.5 491.9
Mixed zone 76.3 9.52 4.09 0.83 0.62 0.77 2.00 0.75 0.07 0.07 10.7 410.9
M5 Loess mantle 77.0 8.74 3.70 0.57 0.66 0.80 2.13 0.77 0.09 0.08 11.5 467.5
Fluvioglacial material 77.5 9.25 3.78 0.73 0.54 0.57 1.77 0.66 0.07 0.05 6.7 259.9
L1 Mixed zone 74.9 9.07 4.28 1.27 1.03 0.83 2.02 1.16 0.17 0.09 19.0 754.3
L2 Mixed zone 73.9 9.23 4.34 1.06 0.81 0.78 1.93 1.19 0.10 0.07 14.0 548.4
L3 Loess mantle 77.3 8.76 3.66 0.61 0.52 0.84 2.15 0.98 0.06 0.08 16.9 658.2
Mixed zone 74.4 10.71 3.90 0.93 0.62 0.87 2.69 0.73 0.10 0.04 10.7 407.1
L4 Loess mantle 78.9 8.96 3.13 0.57 0.37 0.84 2.38 0.76 0.07 0.04 14.9 599.5

Fig. 3. Microphotographs (PPL and XPL) of the rock fragments from the soil profiles: (A, B) basalt from Lubań transect; (C, D) basalt sampled in Muchów catena. Ol:
olivine, Px: pyroxene, Pl: plagioclase, Opq: opaque components.

0.2 mm. Phenocrysts make up about 20 vol% of rock and pyroxene, 4.2. Soil profile grouping across slope catena based on soil particle-size
plagioclase or feldspatoids and opaques constitute up to 50 vol%, 22 vol distribution and geochemistry
% and 8 vol% respectively. According to these properties, all evaluated
fragments classify as basalts. Based on morphological (Table 2 and Fig. 2), textural (Fig. 2) and
geochemical (Table 3, Fig. 2) data, four main cases of slope materials
were distinguished: (1) thick mantle loess — profile L4; (2) thin, silty-

988
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Table 4 increase at the contact with the Bw horizons, until they reach70–90%.
Content of clay minerals estimated based on XRD analysis. Silt dominates in the texture (74–50%), with a decreasing trend in
Profile/horizon Depth (cm) S Ch HIV K I I-S HIS Q V domination with soil depth (Fig. 3). Compared to M1, the profiles L1
and L2 contain more silt and very fine sand (23–16% very fine sand in
M1 Bw2 X XXX XX XX BC horizons). These differences are mirrored by the geochemistry. Hf
M3 Eg X XX XX X X
and Zr mean contents are considerably lower in M1 compared to L1 and
M3 2Btg1 XX X XX X XX
M3 3CBtg XXX XX X X
L2. Still, there is little doubt that M1 contains a significant amount of
M5 Eg X X XX XX X XX aeolian silt, as values of Hf and Zr are below 11 mg·kg−1 and
M5 2Btg/Eg XXX XX XX X 450 mg·kg−1, respectively. Especially in the Bw and BC horizons there
M5 3BCg XXX XX X X is an increasing influence of basalt traceable due to higher TiO2 con-
L1 Bw1 X XXX XX X
tents (Table 3). In all three profiles, well-structured cambic horizons
L1 BC X XX XX XX X
L3 E X XXX XX X X developed.
L3 2Btg1 XXX X X X
L3 2BCg2 XXXX X X X 4.2.4. Thin, silty-textured mantle separated abruptly from underlying
L4 Eg X XXX XX X
materials (aeolian mantle above fluvioglacial deposits)
L4 Btg1/Eg X XX XX X X X X
The loess mantle in the M5 profile overlies coarse-textured flu-
Explanations: XXXX > 70%, XXX 30–70%, XX 10–30%, X 5–10%. vioglacial sediment directly; there is no transitional layer, instead there
S — smectite, Ch — chlorite, HIV — hydroxy interlayered vermiculite, K — is an abrupt lithologic discontinuity present at the contact of those two
kaolinite, I — illite, I-S — illite-smectite, HIS — hydroxy interlayered smectite, substrates (Fig. 2). The surficial mantle is almost free from rock frag-
Q — quartz. ments (5–15%), while the subsoil contains 30 to 70% basaltic rock
fragments mixed with rounded gravels probably of fluvioglacial origin
textured mantle over mixed zone — profiles M2, M3, M4 and L3; (3) (granites, quartzites and so on). The topsoil layer, 32 cm thick, has
basalt block covers with silty infillings — profiles M1, L1 and L2; and texture of silt loam, with silt content ranging from 52 to 54%. The silt
(4) thin silty-textured mantle abruptly separated from underlying flu- content in the subsoil is much lower and does not exceed 32%. Abrupt
vioglacial materials — profile M5. differences between the two materials are also expressed in medium
and fine sand contents that are 2–3 times higher in the subsoil’s hor-
4.2.1. Footslope/toeslope thick loess covers izons (Fig. 3). Due to irregular contents of sand and clay, the texture of
Relatively thick loess mantles cover the plain surrounding the basalt the lowermost horizons varies from sandy clay loam to sandy loam.
hills near Lubań and their footslopes, as identified at the L4 site in the Dissimilarities between the two materials are also highlighted in geo-
Lubań transect (Fig. 1). Profile L4 does not contain coarse fragments chemical data, where Hf and Zr concentrations, are twofold higher in
throughout the entire column, until a depth of 140 cm and shows a the silt mantle compared to the glaciofluvial underlying sediments
constant silt content (70–65%) as well as stable ratio between coarse (Fig. 3). In addition, differences in the content of MgO, Na2O and K2O
and fine silt (Fig. 3). The homogenous nature of the substrate is also point out different lithologies.
expressed in concentrations of SiO2, TiO2 (0.58–0.79%), Hf
(15.9–13.2 mg·kg−1) and Zr (639–551 mg·kg−1). Starting from a depth 4.3. Clay mineralogy of studied soils
of 30 cm intensive tonguing appears (having variable diameters), that
creates a polygonal net in the vertical and lateral orientations, corre- Illite (peak 1.01 nm), hydroxy-interlayered vermiculite (peak near
sponding to retic properties (Fig. 2). The tongues are bleached and 1.4 nm; Mg- and K-saturation and EG-solvation), hydroxy-interlayed
contain substrate from eluvial horizon (Eg), which has a slightly higher smectite (peak ca. 1.5 nm) and chlorite (peak around 1.4 nm after
fine silt content (Fig. 3). Clay accumulation was detected between heating) are typical for soil horizons that contain a considerable share
depth of 20 and 100 cm, classifying this part as an argic horizon. of aeolian silt. These assemblages were detected in thick, homogenous
loess deposits (profile L4); thin, silty-textured loess mantles (uppermost
4.2.2. Thin aeolian silt mantle above mixed layers E horizons in profiles M3, M5 and L3) and loess mixed with basaltic
Profiles M2, M3, M4 and L3 have surficial silty-textured mantle block covers (profiles M1 and L1).
above a thick transition (mixed) zone. The aeolian silt mantles have a Mixed zones in profiles M3, M4 and L3 reveal very high amounts of
variable thicknesses between 28 and 54 cm and contain less than 20% smectites (peaks around 1.7, 1.4 after Mg-saturation and EG-solvation
of coarse fragments (only basaltic gravels and boulders). The content of and 1.2 after K-saturation) while the other secondary clay minerals are
silt fraction ranges between 56 and 72%, and the sand fraction is generally absent (Table 4). The underlying clay loam solifluction beds
dominated by very fine sand, as in other loess mantles (Fig. 3). In all from profiles M3 and L3 show similar pattern of clay mineral dis-
profiles, an abrupt increase in coarse basaltic rocks was observed at the tribution (absolute dominance of smectites, peaks at 1.7 and 1.71 nm
contact of the loess mantle and mixed zones. Below silt-textured man- after EG-solvation as well as 1.39 and 1.44 nm after Mg saturation), as
tles, in the mixed zone, the silt share decreases. The lowermost horizons in mixed zones (Fig. 6).
(3CBg) have silt loam/loam (M3 and M4, respectively) or clay loam Additionally, in all evaluated samples kaolinite (peak at 0.72 nm,
(M3, L3) textures and more than 80% of rock fragments are arranged in collapsing during heating under 500 °C) was detected (Table 4, Fig. 5
lateral position, while the structure in the subsoil is strong and platy. In and 6).
the lowermost horizon of the M4 profile, also some addition of glacial
material was noted. Geochemical data (Table 3) helps distinguished 4.4. Heavy minerals
aeolian mantles from mixed zones and notes that TiO2 increased with
soil depth while Hf (12.0–10.7 mg·kg−1) and Zr (460–407 mg·kg−1) Thick loess mantles (profile L4) exhibit the dominance of amphi-
decrease. boles, chlorite, epidote, ilmenite, rutile, saurolite and zircon (Fig. 7).
The signal from basalt materials is very low as the pyroxene content
4.2.3. Basalt block/stone covers with aeolian silt admixture (mixed layers) does not exceed 0.6%. In basaltic block, covers mixed with loess heavy
The profiles M1, L1 and L2 illustrate the cases where very weakly mineral distribution have a different trend; here, amphiboles are very
weathered basaltic rock fragments (gravels and boulder size) are ad- common (38%); however markers of basaltic materials such as ulvos-
mixed with finer silty substrate. The uppermost horizons (A, AB) consist pinel, pyroxene and olivine, make up 45% in total, proving significantly
of up to 15–20% of rocks, then coarse fragments increase abruptly the presence of mafic components. Soils with mixed zones (profiles M3

989
J. Waroszewski et al. Geoderma 337 (2019) 983–997

(caption on next page)


990
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 4. Depth plots showing basic (grain size distribution, pH and content of organic carbon) and geochemical data (Hf, Zr, TiO2) from the studied soil profiles.

and L3) have a very clear signal in BC horizons from basaltic compo- which to evaluate the thicknesses of loess mantles and mixed zones. The
nents: ulvospinel (25–40%), pyroxene (5–8%) and olivine (3–17%), but specific composition of aeolian silt materials (Martignier et al., 2015;
the aeolian admixture is also very clear (amphiboles, epidote, pum- Munroe et al., 2007; Portes et al., 2018) is very well reflected in thick
pellyite, zircon, rutile). In the Btg horizons, the impact of basaltic loess, loess mantles and block covers with loess input, as HIV, HIS,
material is decreasing, while components related to aeolian silt con- chlorite and illite together with kaolinite were detected in these units.
tribution are becoming more pronounced (Fig. 7). Aeolian silt mantles The differentiation of layers composed mainly of aeolian silts, such as
over mixed zones reveal a very high share of amphiboles, epidote, rutile thick loess mantles and thin silt-textured layers is possible (Fig. 4 and
and ilmenite (as in like profile L4), and, at the same time they hold very Fig. 8). However, difficulties appear in mixed zones, in which the
minor contents of pyroxene, ulvospinel and pyroxene. In almost all aeolian-silt-dominated substrate is incorporated into weathered basaltic
profiles, significant share of Fe-oxides was noted and interpreted as rocks. Smectitic components are detectable in mixed zones (profile M3,
being part of magnetite. horizon 2Btg/Eg) in medium amounts (Tab. 4). In contrast, swelling
phases predominate absolutely in the clay fraction of the lowermost
5. Discussion horizons (profile M5, Fig. 6; profile L3, Tab. 4) which is not typical for
aeolian silts. However, most of mixed zones hold Bt horizons, where
5.1. Inheritance of geochemical and mineralogical features smectites accumulated via illuviation, which is an obvious phenomenon
(Bogda et al., 1998; Drewnik et al., 2014). Hence, smectites cannot be
The degree of aeolian silt inclusion into residuum materials (local treated as a mineralogical marker of basalt-derived materials, although
bedrock) at the edge of thin loess patches is often hard to define in relatively young fine-grained basaltic saprolites, drilled in few sites
(Schaetzl and Luehmann, 2013, Waroszewski et al., 2018a). Also, the in the Lower Silesia (e.g. at Męcinka) smectites are abundant (Choma-
identification of thin loess deposits itself has proved to be problematic Moryl, 2007). They are, however, purely clay textured and represent
in such a diffuse contact zone (Schaetzl et al., 2018). Possible ways to warm and humid weathering, probably occurring during the Neogene
evaluate how strong the intermixing of loess (aeolian silts) with the and are much different from clay loam slope deposits formed in the
underlying sediments are: (1) field work observations and description Pleistocene. Nonetheless, the influence of weakly weathered basalts or
(Luehmann et al., 2013; Martignier et al., 2015; Schaetzl, 2008); (2) basaltic-derived materials, as smectite precursors, cannot be excluded.
tracking original loess texture with particle-size analyses (Schaetzl and Drawing a boundary between the mixed zone and basal layers (based
Luehmann, 2013), (3) conducting a mineralogical investigation, in- on clay mineralogy) is barely possible (Fig. 8).
cluding basic mineralogy with primary and secondary phases (Muhs The most convincing evidence with which to evaluate the degree of
and Benedict, 2006), as well as heavy fraction; and (4) through geo- aeolian influence down the profile and thus to discriminate mixed zones
chemistry (Muhs, 2018; Waroszewski et al., 2018a; Waroszewski et al., from basal layers are heavy minerals and geochemistry (Fig. 8). Soli-
2018b). fluction clay loam layers (profiles L3 and M3) are revealed by the
Thick loess deposits are easy to recognize (Küfmann, 2003; presence of zircon, rutile or monazite (Fig. 7) and also by the contents
Waroszewski et al., 2018a) as are aeolian drapes/loess mantles (Gild of Hf (oscillating from 14.1 to 10.0 mg·kg−1) and Zr (ranging from 551
et al., 2018; Waroszewski et al., 2018a) that directly overlie continuous to 381 mg·kg−1). These data show very weak but significant aeolian silt
rock or coarser/finer sediments directly where drawn lithological incorporation. Therefore, these clay loams need to be treated as part of
boundaries are sharp and not distracted. Typically, field observations of the multi-layered mixed zones. The geochemistry also supports this,
morphological features are sufficient to designating parent materials showing a still high content of Hf and Zr that ranges from 9.1 to
and confirming the presence of loess (Labaz et al., 2018; Stanley and 10.6 mg·kg−1 and from 363.5 to 414.8 mg·kg−1, respectively. The given
Schaetzl, 2011). concentrations are enough to link such substrates with loess and loess
In the case of the four categories of parent material configurations derivates (Scheib et al., 2014). Despite the high content of TiO2
defined herein, profile L4 developed in thick loess without any sig- (1.35–2.11%), which indicates minor basaltic influence, an aeolian silt
nificant basaltic component. The aeolian mantles of the profiles M2–5 incorporation is unquestionable. Based on the geochemical pattern,
and L3 and the basaltic blocks covered with fine substrate (L1, L2 and only in profile M5 should the material below the loess mantle be de-
M1) reveal clear aeolian silt signatures detectable in the field. However, signated as unaffected by aeolian components (Fig. 8), despite the
the thicknesses and maximum depths of mixed zones often remain aeolian contribution being significant.
unclear. Most important in this context is the differentiation of the At the outer lateral boundaries of thin loess deposits there are no
lower boundary of the influence for aeolian silt, which is the upper soils developed purely from basaltic materials; they always exhibit
boundary of the pure residual material and which also may show signs strong signals from aeolian silt substrates. This is reflected mostly by
of transport along slopes. This cover layer composed of local material geochemical data, but the heavy minerals assemblages also clearly in-
without aeolian addition is the basal layer in the German system of dicate deep mixing of loess materials with overlying sediments and lack
cover beds (Kleber and Terhorst, 2013). Defining the lower boundary of of basal layers without aeolian silt in most trenches.
aeolian influence is obviously challenging in our study area. Differences in heavy minerals and particle size and data between the
The reasons for difficulties to detecting the lithological boundaries Lubań and Muchów transects indicate two possible sources of aeolian
in the field, including the evaluation of grain size distribution, are due silts. Based on the abundance of heavy minerals abundance those en-
to (a) the prevalence of a silt loam texture even in the C horizons vironments are: (i) metamorphic and characterized by amphiboles,
containing a large share of basaltic clasts (e.g. profile M2 2BCg); (b) the epidote, titanite, pumpellyite and ilmenite; and (ii) igneous as re-
transformation of the silt loam texture into loam due to weathering (e.g. presented by monacite, zircon, staurolite and also partially by rutile.
profile M4, Fig. 4), blurring loess-like features and their proper re- The evident presence of two sources is expressed also by the relation of
cognition; and (c) the transition between silt loam materials (Btg hor- fine silt to coarse silt - in the soils of the Lubań transect the coarse silt
izons) hosting a high share of coarse basaltic material and horizons content is higher, while, in Muchów fine silt predominates. In addition,
(3BCg) having clay loam (profiles M3 and L3, Fig. 4). In addition to the very fine sand content is much higher at Lubań. The coarser mode of
field impressions and particle-size distributions (Fig. 8) further proof is loess deposits points to nearby sources of material (short distance
necessary in our study area to define the lithological boundaries. transport; Krawczyk et al., 2018). These components were probably
Clay mineralogy could serve only partially as an additional tool with deflated from glacial outwash plains (Badura et al., 2013; Luehmann

991
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 5. X-ray diffraction (XRD) patterns of clay minerals with all kinds of
treatments (EG: ethylene glycol, Mg & K exchange, heating) in soils developed
from block covers mixed with loess and thick loess deposits. The d-spacing are
given in nanometres.

et al., 2013; Nyland et al., 2018) mobilized by katabatic winds (Gild


et al., 2018; Martignier et al., 2015; Schaetzl and Attig, 2013). Loess
materials and loess derivates incorporated into the basaltic material at
the Lubań transect, reveal a bit more gneiss/mica schist character (e.g.,
staurolite), while Muchów has a more metamorphic type of source.
Therefore, these differences could be connected with the Great Odra
Valley delivery of weathered silty material via the rivers Kwisa and
Bóbr. This direct supply via mountain rivers of siliciclastic fine-grained
material (silt) probably took place during the Pleniweichselian (MIS4-
MIS2; Badura et al., 2013). The observed differences in particle dis-
tribution in loess suggest the involvement of various aeolian episodes/
sources, which are probably linked to changes in paleoenvironmental
conditions during loess deposition (Gild et al., 2018; Kels et al., 2014).

5.2. The influence of slope deposits on soil formation

In the deeper parts of the periglacial block and grus slope sediments,
the degree of aeolian silt contribution is weak or moderate (profiles L1,
L2 and M1) with a predominance of basaltic coarse fragments, while
upwards the presence of loess infillings increases until the surface,
where it predominates (profiles M2, M3, M4, M5 and L3). Only in-
dividual rock fragments are preserved in continuously deposited loess.
In places, loess deposits are so thick that underlying basaltic material
was completely covered, and no rock fragments are present at the
surface (profile L4). Therefore, four types of slope sediments were
distinguished on both investigated slopes (Fig. 9): (1) thick loess de-
posits in the toe slope positions; (2a + b) heterogeneous deposits
composed of fine clayey/sandy and silty (loess) material with the pre-
sence of larger clasts in high abundance and (3) block fields with
aeolian silt admixture directly below rock outcrops and cliffs. A de-
crease of pure loess deposits with altitude could be related to higher
wind speeds preventing accumulation (Nyland et al., 2018). The other
possible factors are erosion and slope processes (syn- and postdeposi-
tional) that led to the disappearance of an existing loess mantle, which
was not stabilized by basaltic block/grus covers. Stratification features
(e.g., lateral orientation of basalt fragments) recognized in hetero-
geneous deposits indicate the role of solifluction in their formation and
relocation on slopes (Migoń et al., 2002). These processes probably had
only short-range effects (Migoń and Kacprzak, 2014) and variable in-
tensities on the studied slopes.
Basaltic bedrock as the potential parent material for the studied sites
was influenced by an aeolian silt deposition (with syn- and post-sedi-
mentary processes) that differs strongly from old basalt saprolites which
has been never under the influence of glaciations or periglacial condi-
tions. Our findings indicate that the main soil-forming process is clay
translocation, which is not common in typical basalt-derived materials
(with expandable clays) in which vertic properties have been rather
common (Kabala et al., 2015).
Thick loess mantles and thin silty-textured layers contain material
with sufficient (not too high) content of clay minerals, including fine
clays (smectites) and have good water permeability. In a relatively
humid climate, the leaching of carbonates and mobilization of clays is
possible in a material that contains a clay fraction that is easy to
translocate. Domination of clay illuviation and formation of argic hor-
izons causes a reduction in permeability and, under constantly humid
conditions typical argic horizons are transformed via tonging into dis-
continuous/cracked vertical forms. Such degradation processes in illu-
vial-clay subsoil (Bockheim, 2015; Kels et al., 2014; Quénard et al.,
2011; Szymański et al., 2012, 2011) observed in deep loess mantles
(52–130 cm) result in the presence of retic or in the case of thick loess

992
J. Waroszewski et al. Geoderma 337 (2019) 983–997

(caption on next page)


993
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 6. X-ray diffraction (XRD) patterns of clay minerals with all kinds of treatments (EG: ethylene glycol, Mg & K exchange, heating) in soils having loess mantle and
variable thickness of mixing zone. The d-spacing are given in nanometres.

Fig. 7. Heavy minerals distribution in selected soil profiles referring to thick loess mantles (L1), block covers covered with loess (L1) and soils having different
thickness and morphology of mixed zones (M3 and L3).

994
J. Waroszewski et al. Geoderma 337 (2019) 983–997

Fig. 8. Evaluation of mixing different sediments and distinguishing layers (aeolian mantle, mixed zone, basal layer and fluvioglacial material) by applying different
proxies. The final interpretation is always consistent with geochemical data that serves as the most appropriate tool to track even weak aeolian input on underlying
sediments.

Fig. 9. Idealized catena for the basalt hills of Lower Silesia, in proximity to loess deposits without (fluvio-)glacial influence.

mantles covering dense fluvioglacial loams, glossic properties. These translocation result in a differentiated soil texture and the formation of
features lead to a classification of respective soils as Retisols (L4) and an argic horizon. At the same time, dense argic horizons and the low-
Glossic Stagnosols (M4). The intensity of leaching is expressed in Re- ermost clay loam horizons caused strong stagnation of rainwater
tisol with qualifier Alic. (Rubinić et al., 2015, Rubinić et al., 2017). Mixed zones and part of the
In the midslope sections, deposits having clay loam and silt clay loess mantles fulfil the criteria for stagnic colour properties and together
loam predominate, with clay contents of ca. 30%. Contribution of with reduction conditions, determines adherence to Stagnosols (Fig. 9),
aeolian silt dilutes the primary (clay loam) soil texture (Waroszewski instead of the expected Luvisols.
et al., 2018a) and determinates the formation of clay-illuviation fea- Assuming their formation throughout the Holocene, the soils de-
tures on soil peds and within root channels. Clay dispersion and clay veloped in the block fields with dust admixture are relatively weakly

995
J. Waroszewski et al. Geoderma 337 (2019) 983–997

weathered and reveal rather stable conditions (in terms of the present References
morphodynamic stability and weathering processes), as pyroxenes and
amphiboles are present in high amounts (Fig. 7). Within the clay frac- Alibert, C., Leterrier, J., Panasiuk, M., Zimmermann, J.L., 1987. Trace and isotope geo-
chemistry of the alkaline Tertiary volcanism in southwestern Poland. Lithos 20 (4),
tion, mixed-layered minerals dominate (almost no smectites). Aeolian
311–321.
additions to coarse basaltic regoliths (block fields) free of fine material Badura, J., Jary, Z., Smalley, I., 2013. Sources of loess material for deposits in Poland and
determine further soil forming processes. Silty strata provide a valuable part of Central Europe: the lost Big River. Quat. Int. 296, 15–22.
Barnhisel, R.I., Bertsch, P.M., 1989. Chlorite and hydroxy-interlayered vermiculite and
substrate for the development of well-structured horizons penetrated smectite. In: Dixon, J.B., Weed, S.B. (Eds.), Minerals in Soil Environments. SSSA Book
easily with roots, while low clay contents and fast water movement Ser., 1 1. pp. 729–788. https://doi.org/10.2136/sssabookser1.2ed.c15.
exclude the possibility of clay translocation, so no stagnic or argic hor- Birkenmajer, K., 1967. Bazalty dolnośląskie jako zabytki przyrody nieożywionej. Ochrona
Przyrody 32, 225–276 (in Polish).
izons were formed. Therefore, in profiles developed within block fields Birkenmajer, K., Pécskay, Z., Grabowski, J., Lorenc, M.W., Zagożdżon, P., 2004.
mixed with loess (M1, L1, L2) a large amounts of roots and the activity Radiometric dating of the Tertiary volcanic in Lower Silesia, Poland. IV. Further K-Ar
and paleo-magnetic data from late Oligocene to early Miocene basaltic rocks of the
of earthworms were responsible for the moderately pedogenic alter- foresudetic block. Ann. Soc. Geol. Pol. 74, 1–19.
nation and formation of young biogenic cambic horizons; therefore, the Bockheim, J.G., 2015. Distribution and origin of glossic horizons in soils of the western
investigated soils were classified as Cambisols. It cannot be excluded Great Lakes Region. Geoderma Reg. 5, 226–236. https://doi.org/10.1016/J.
GEODRS.2015.08.004.
that in the past, loess sediments were thicker in the upper slopes on Bogda, A., Chodak, T., Szerszeń, L., 1998. Properties and Composition of Clay Minerals in
basaltic block covers and soils with subsoil clay accumulation existed. Soils of the Lower Silesia. In: AR Wrocław, pp. 89 (in Polish).
Their erosion could be linked to the deforestation that took place on Bojko, O., Kabala, C., 2016. Transformation of physicochemical soil properties along a
mountain slope due to land management and climate changes — a case study from
most basaltic hills from the Middle Ages until the 19th century. the Karkonosze Mountains, SW Poland. Catena 140. https://doi.org/10.1016/j.
catena.2016.01.015.
Boxleitner, M., Musso, A., Waroszewski, J., Malkiewicz, M., Maisch, M., Dahms, D.,
Brandová, D., Christl, M., de Castro Portes, R., Egli, M., 2017. Late
6. Conclusions
Pleistocene–Holocene surface processes and landscape evolution in the central Swiss
Alps. Geomorphology 295. https://doi.org/10.1016/j.geomorph.2017.07.006.
This study is a guideline on how to investigate the scale of in- Brindley, G.W., 1980. Quantitative X-ray mineral analysis of clays. In: Brindley, G.W.,
Brown, G. (Eds.), Crystal Structures of Clay Minerals and Their X-ray Identification.
heritance of signatures typical of loess and basalt-derived substrates in Mineral. Soc. Monogr. 5, London, pp. 411–438. https://doi.org/10.1180/mono-5.7.
soils having both components as a parent material. Field and laboratory Brown, G., Brindley, G.W., 1980. X-ray diffraction procedures for clay mineral identifi-
data (profile morphology, granulometry, mineralogy and geochemistry) cation. In: Brindley, G.W., Brown, G. (Eds.), Crystal Structures of Clay Minerals and
Their X-ray Identification. Mineral. Soc. Monogr. 5, London, pp. 305–359. https://
were compared in terms of their significance in tracking the thickness of doi.org/10.1180/mono-5.5.
mixed zones between loess-dominated mantles and local substrate Caner, L., Radtke, L.M., Vignol-Lelarge, M.L., Inda, A.V., Bortoluzzi, E.C., Mexias, A.S.,
2014. Basalt and rhyo-dacite weathering and soil clay formation under subtropical
without aeolian silt incorporation. Our results show that geochemistry climate in southern Brazil. Geoderma. https://doi.org/10.1016/j.geoderma.2014.06.
and heavy minerals are the most reliable proxies for the verification of 024.
aeolian silt contributions and for defining individual layers (loess Choma-Moryl, K., 2007. Expansiveness of slected basaltoid wethering products in Lower
Silesia, Poland. Geol. Sudet. 39, 1–9.
mantle, mixed zone, basal layers) in complex soil parent materials. Only Drewnik, M., Skiba, M., Szymański, W., Żyła, M., 2014. Mineral composition vs. soil
by comparing all parameters, a universal understanding of their im- forming processes in loess soils — a case study from Kraków (Southern Poland).
portance to complex parent material arrangement be made possible. Catena 119, 166–173.
Egli, M., Merkli, C., Sartori, G., Mirabella, A., Plötze, M., 2008a. Weathering, miner-
Soil development on basaltic hills at the outer boundaries of loess alogical evolution and soil organic matter along a Holocene soil toposequence de-
deposits is controlled by the Pleistocene relief of the slopes and the way veloped on carbonate-rich materials. Geomorphology 97, 675–696. https://doi.org/
10.1016/J.GEOMORPH.2007.09.011.
aeolian silt is incorporated into slope deposits. On basaltic block covers Egli, M., Mirabella, A., Sartori, G., 2008b. The role of climate and vegetation in weath-
mixed with loess, Cambisols developed. Midslope sections are domi- ering and clay mineral formation in late Quaternary soils of the Swiss and Italian
nated by Stagnosols developed from loess mixed with clay loam rich in Alps. Geomorphology 102, 307–324. https://doi.org/10.1016/J.GEOMORPH.2008.
04.001.
basaltic clast layers. The lowermost part of the slope is characterized by Egli, M., Sartori, G., Mirabella, A., Giaccai, D., 2010. The effects of exposure and climate
the presence of thick loess deposits that give rise to Retisols formation on the weathering of late Pleistocene and Holocene Alpine soils. Geomorphology 114,
466–482. https://doi.org/10.1016/J.GEOMORPH.2009.08.008.
under humid conditions. Egli, M., Wernli, M., Burga, C., Kneisel, C., Mavris, C., Valboa, G., Mirabella, A., Plötze,
In the studied Cambisols and Retisols, all signatures typical for M., Haeberli, W., 2011. Fast but spatially scattered smectite-formation in the pro-
aeolian silts were detected, while the basaltic component played minor glacial area Morteratsch: an evaluation using GIS. Geoderma. https://doi.org/10.
1016/j.geoderma.2011.05.001.
role or was not found at all. Along the slopes impact of basalt-derived Egli, M., Dahms, D., Norton, K., 2014. Soil formation rates on silicate parent material in
slope sediments is clearer especially in the lowermost horizons, how- alpine environments: different approaches-different results? Geoderma. https://doi.
ever input of loess was also traceable even in the C horizons. Based on org/10.1016/j.geoderma.2013.08.016.
FAO (Ed.), 2006. Guidelines for Soil Description, 4th ed. FAO, Rome.
our findings it seems that at the edge of thin loess deposits no soils exist Gild, C., Geitner, C., Sanders, D., 2018. Discovery of a landscape-wide drape of late-
which were developed purely from a basaltic type of parent material; glacial Aeolian silt in the western Northern Calcareous Alps (Austria): first results and
implications. Geomorphology 301, 39–52. https://doi.org/10.1016/J.GEOMORPH.
they always revealed strong or medium geochemical and mineralogical 2017.10.025.
signals pointing to an aeolian silt contribution. Haase, D., Fink, J., Haase, G., Ruske, R., Pecsi, M., Richter, H., Altermann, M., Jager, K.D.,
2007. Loess in Europe e its spatial distribution based on a European loess map, scale
1:2,500,000. Quat. Sci. Rev. 26, 1301e1312.
He, Y., Li, D.C., Velde, B., Yang, Y.F., Huang, C.M., Gong, Z.T., Zhang, G.L., 2008. Clay
Acknowledgements minerals in a soil chronosequence derived from basalt on Hainan Island, China and its
implication for pedogenesis. Geoderma 148, 206–212.
This research was financed by National Science Center (Poland) Hutton, J.T., 1977. Titanium and zirconium minerals. In: Dixon, J.B., Weed, S.B. (Eds.),
Minerals in Soil Environments. Soil Science Society of America, Madison, Wisconsin,
projects Sonata 8 nr 2014/15/D/ST10/04087 and Opus 8 nr 2014/15/ pp. 673–688.
B/ST10/04606 as well as by Wroclaw University of Environmental and IUSS Working Group WRB, 2015. World Reference Base for Soil Resources 2014, update
2015 International Soil Classification System for Naming Soils and Creating Legends
Life Sciences, project numbers D210/0021/18 and B030/0020/18. The for Soil Maps. World Soil Resources Reports No. 106. FAO, Rome.
authors are indebted to two anonymous reviewers for their constructive Jackson, M.L., 1969. Soil chemical analysis: advanced course. Published by the author
remarks and comments on an earlier version of the manuscript. 2nd ed. Dept. of Soil Science, Univ. of Wisconsin, Madison, Wisconsin.
Jary, Z., 1999. Ostatni cykl lesowy w Polsce SW. In: Jary, Z. (Ed.), III Seminarium Lessowe
– Geneza i wiek pokrywowych utworów pylastych południowo-zachodniej Polski.
Uniwersytet Wrocławski, Wrocław-Bożków, Poland, pp. 35–37.
Appendix A. Supplementary data Johnson, D.L., 1993. Dynamic denudation evolution of tropical, subtropical and tempe-
rate landscapes with three-tiered soils: toward a general theory of landscape evolu-
tion. Quat. Int. 17, 67–78.
Supplementary data to this article can be found online at https:// Johnson, D.L., Domier, J.E.J., Johnson, D.N., 2005. Animating the biodynamics of soil
doi.org/10.1016/j.geoderma.2018.11.008. thickness using process vector analysis: A dynamic denudation approach to soil

996
J. Waroszewski et al. Geoderma 337 (2019) 983–997

formation. Geomorphology. https://doi.org/10.1016/j.geomorph.2004.08.014. 1016/j.geoderma.2014.02.022.


Juilleret, J., Dondeyne, S., Vancampenhout, K., Deckers, J., Hissler, C., 2016. Mind the Portes, R., Dahms, D., Brandová, D., Raab, G., Christl, M., Kühn, P., Ketterer, M., Egli, M.,
gap: a classification system for integrating the subsolum into soil surveys. Geoderma 2018. Evolution of soil erosion rates in alpine soils of the Central Rocky Mountains
264, 332–339. https://doi.org/10.1016/j.geoderma.2015.08.031. using fallout Pu and δ 13 C. Earth Planet. Sci. Lett. 496, 257–269. https://doi.org/10.
Kabala, C., Bekier, J., Bińczycki, T., Bogacz, A., Bojko, O., Cuske, M., Ćwieląg-Piasecka, I., 1016/j.epsl.2018.06.002.
et al., 2015. Soils of Lower Silesia: Origins, Diversity and Protection. PTG, PTSH, Puziewicz, J., Koepke, J., Gregoire, M., Ntaflos, T., Matusiak-Małek, M., 2011.
Wrocław. Lithospheric mantle modification during Cenozoic rifting in Central Europe: evidence
Kels, H., Protze, J., Sitlivy, V., Hilgers, A., Zander, A., Bertrams, M., Lehmkuhl, F., 2014. from the Księginki nephelinite (SW Poland) Xenolith Suite. J. Petrol. 52, 2107–2145.
Genesis of loess-like sediments and soils at the foothills of the Banat Mountains, Quénard, L., Samouëlian, A., Laroche, B., Cornu, S., 2011. Lessivage as a major process of
Romania — examples from the Paleolithic sites Româneşti and Coşava. Quat. Int. soil formation: a revisitation of existing data. Geoderma 167–168, 135–147. https://
351, 213–230. https://doi.org/10.1016/J.QUAINT.2014.04.063. doi.org/10.1016/J.GEODERMA.2011.07.031.
Kida, J., 1999. Stan badań lessów i osadów lessopodobnych na obszarze Dolnego Śląska. Raczkowski, W., 1969. Lessy i utwory pylaste Przedgórza Sudeckiego. Archiwum
In: Jary, Z. (Ed.), III Seminarium Lessowe – Geneza i wiek pokrywowych utworów Instytutu Geografii i Rozwoju Regionalnego. Uniwersytetu Wrocławskiego (149 pp).
pylastych południowo-zachodniej Polski. Uniwersytet Wrocławski, Wrocław-Bożków, Rasmussen, C., Dahlgren, R.A., Southard, R.J., 2010. Basalt weathering and pedogenesis
Poland, pp. 43–54. across and environmental range in the southern Cascade Range, California, USA.
Kirkby, M.J., 2018. A conceptual model for physical and chemical soil profile evolution. Geoderma 154, 473–485.
Geoderma. https://doi.org/10.1016/j.geoderma.2018.06.009. Reed, S.J.B., 2010. Electron microprobe analysis and scanning electron microscopy in
Kleber, A., 1997. Cover-beds as soil parent materials in midlatitude regions. Catena 30, geology. University of Cambridge, Cambridge University Press.
197–213. https://doi.org/10.1016/S0341-8162(97)00018-0. Rex, R.W., Syers, J.K., Jackson, M.L., Clayton, R.N., 1969. Eolian origin of quartz in soils
Kleber, A., Terhorst, B., 2013. Mid-latitude slope deposits (Cover Beds). In: Developments of the Hawaiian Islands and in Pacific pelagic sediments. Science 163, 277–279.
in Sedimentology. 66 Elsevier, Amsterdam. Rubinić, V., Galović, L., Husnjak, S., Durn, G., 2015. Climate vs. parent material — which
Krawczyk, M., Ryzner, K., Skurzyński, J., Jary, Z., 2018. Lithological indicators of loess is the key of Stagnosol diversity in Croatia? Geoderma 241–242, 250–261. https://
sedimentation of SW Poland. Contemp. Trends Geosci. 6 (2), 94–111. doi.org/10.1016/J.GEODERMA.2014.11.029.
Küfmann, C., 2003. Soil types and eolian dust in hight mountainous karst of the Northern Rubinić, V., Galović, L., Lazarević, B., Husnjak, S., Durn, G., 2017. Pseudogleyed loess
Calcareous Alps (Zugspitzplatt, Wetterstein Mountains, Germany). Catena 53, derivates — the most common soil parent materials in the Pannonian region of
211–227. Croatia. Quat. Int. https://doi.org/10.1016/J.QUAINT.2017.06.044.
Labaz, B., Kabala, C., 2016. Human-induced development of mollic and umbric horizons Schaetzl, R.J., 2008. The distribution of silty soils in the Grayling Fingers region of
in drained and farmed swampy alluvial soils. Catena 139, 117–126. https://doi.org/ Michigan: evidence for loess deposition onto frozen ground. Geomorphology 102,
10.1016/J.CATENA.2015.12.013. 287–296. https://doi.org/10.1016/J.GEOMORPH.2008.03.012.
Labaz, B., Musztyfaga, E., Waroszewski, J., Bogacz, A., Jezierski, P., Kabala, C., 2018. Schaetzl, R.J., Attig, J.W., 2013. The loess cover of northeastern Wisconsin. Quat. Res. 79,
Landscape-related transformation and differentiation of Chernozems — catenary 199–214. https://doi.org/10.1016/J.YQRES.2012.12.004.
approach in the Silesian Lowland, SW Poland. Catena 161. https://doi.org/10.1016/ Schaetzl, R.J., Luehmann, M.D., 2013. Coarse-textured basal zones in thin loess deposits:
j.catena.2017.10.003. products of sediment mixing and/or paleoenvironmental change? Geoderma 192,
Lu, S.G., Xue, Q.F., Zhu, L., Yu, J.Y., 2008. Mineral magnetic properties of a weathering 277–285. https://doi.org/10.1016/J.GEODERMA.2012.08.001.
sequence of soils derived from basalt in eastern China. Catena 73, 23–33. Schaetzl, R., Bettis, E., Crouvi, O., Fitzsimmons, K., Grimley, D., Hambach, U., ... Zech, R.,
Luehmann, M.D., Schaetzl, R.J., Miller, B.A., Bigsby, M.E., 2013. Thin, pedoturbated, and 2018. Approaches and challenges to the study of loess—introduction to the LoessFest
locally sourced loess in the western Upper Peninsula of Michigan. Aeolian Res. 8, Special Issue. Quat. Res. 89 (3), 563–618. https://doi.org/10.1017/qua.2018.15.
85–100. https://doi.org/10.1016/J.AEOLIA.2012.11.003. Scheib, A.J., Birke, M., Dinelli, E., Project Team, G.E.M.A.S., 2014. Geochemical evidence
Mange, M.A., Maurer, H.F.W., 1992. Heavy Minerals in Colour. Chapman and Hall, of aeolian deposits in European soils. Boreas 43, 175–192.
London (147 pp.). Soubrand-Colin, M., Bril, H., Neel, C., Courtin-Nomade, A., Martin, F., 2005. Weathering
Martignier, L., Nussbaumer, M., Adatte, T., Gobat, J.-M., Verrecchia, E.P., 2015. of basaltic rocks from the French Massif Central: origin and fate of Ni, Cr, Zn and Cu.
Assessment of a locally-sourced loess system in Europe: the Swiss Jura Mountains. Can. Mineral. https://doi.org/10.1016/j.geoderma.2006.08.017.
Aeolian Res. 18, 11–21. https://doi.org/10.1016/J.AEOLIA.2015.05.003. Soubrand-Colin, M., Neel, C., Bril, H., Grosbois, C., Caner, L., 2007. Geochemical beha-
McLennan, S.M., 1989. Rare earth elements in sedimentary rocks: influence of prove- viour of Ni, Cr, Cu, Zn and Pb in an Andosol-Cambisol climosequence on basaltic
nance and sedimentary processes. Rev. Mineral. 21, 169–200. rocks in the French Massif Central. Geoderma. https://doi.org/10.1016/j.geoderma.
McLennan, S.M., 2001. Relationships between trace elemental composition of sedimen- 2006.08.017.
tary rocks and upper continental crust. Geochem. Geophys. Geosyst. 2, Soubrand-Colin, M., Horen, H., Courtin-Nomade, A., 2009. Mineralogical and magnetic
2000GC000109. https://doi.org/10.1029/2000GC000109. characterisation of iron titanium oxides in soils developed on two various basaltic
Migoń, P., Kacprzak, A., 2014. Lateral diversity of regolith and soils under a mountain rocks under temperate climate. Geoderma. https://doi.org/10.1016/j.geoderma.
slope — implications for interpretation of hillslope materials and processes, central 2008.11.018.
Sudetes, SW Poland. Geomorphology 221, 69–82. Sprafke, T., Obreht, I., 2016. Loess: rock, sediment or soil — what is missing for its de-
Migoń, P., Maciejak, K., Zygmunt, M., 2002. Periglacial morphology of basalt hills in the finition? Quat. Int. 399, 198207.
Kaczawa Upland (Western Sudetes). Prz. Geol. 74, 4,491–508 (in Polish). Stanley, K.E., Schaetzl, R.J., 2011. Characteristics and paleoenvironmental significance of
Morrison, R.J., 1991. Variability of the Fagaaga Series soils of Western Samoa. Geoderma a thin, dual-sourced loess sheet, north-central Wisconsin. Aeolian Res. 2, 241–251.
49, 105–116. https://doi.org/10.1016/J.AEOLIA.2011.01.001.
Morrison, R.J., Bonato, J.A., 2015. Weathering and soil genesis from the Nasinu Basalt, Szymański, W., Szkaradek M. 2017. Andesite weathering and soil formation in a mod-
South-East Viti Levu, Fiji. S. Pac. J. Nat. Sci. 33 (2), 25–38. erately humid climate: a case study from the western Carpathians (southern Poland),
Muhs, D.R., 2018. The geochemistry of loess: Asian and North American deposits com- Carpath. J. Earth Environ. Sci.,13, No. 1, p. 93 – 105; DOI:10.26471/cjees/2018/
pared. J. Asian Earth Sci. 155, 81–115. https://doi.org/10.1016/J.JSEAES.2017.10. 013/010
032. Szymański, W., Skiba, M., Skiba, S., 2012. Origin of reversible cementation and brittle-
Muhs, D.R., Benedict, J.B., 2006. Eolian additions to late Quaternary alpine soils, Indian ness of the fragipan horizon in Albeluvisols of the Carpathian Foothills, Poland.
Peaks Wilderness Area, Colorado Front range. Arct. Antarct. Alp. Res. 38, 120–130. Catena 99, 66–74. https://doi.org/10.1016/j.catena.2012.07.012.
Munroe, J.S., Farrugia, G., Ryan, P.C., 2007. Parent material and chemical weathering in Terhorst, B., 2000. The influence of Pleistocene landforms on soil-forming processes and
alpine soils on Mt. Mansfield, Vermont, USA. Catena 70, 39–48. https://doi.org/10. soil distribution in a loess landscape of Baden–Württemberg (south-west Germany).
1016/J.CATENA.2006.07.003. Catena 41, 165–179. https://doi.org/10.1016/S0341-8162(00)00098-9.
Naidu, R., Kirkman, J.H., Morrison, R.J., 1987. Mineralogy of soils from basaltic ash, Terhorst, B., Kleber, A., Bibus, E., 2013. Relative dating with cover beds. Dev. Sedimentol.
Taveuni, Fiji. Geoderma 33, 181–192. 66, 229–251. https://doi.org/10.1016/B978-0-444-53118-6.00007-6.
Noack, Y., Colin, F., Nahon, D., Delvigne, J., Michaux, L., 1993. Secondary mineral for- Van Reeuwijk, L.P., 2002. Procedures for Soil Analysis, 6th ed. ISRIC, Wageningen,
mation during natural weathering of pyroxene: review and thermodynamic ap- Netherlands.
proach. Am. J. Sci. 293, 111–134. Waroszewski, J., Sprafke, T., Kabala, C., Musztyfaga, E., Labaz, B., Wozniczka, P., 2018a.
Norton, K.P., Molnar, P., Schlunegger, F., 2014. The role of climate-driven chemical Aeolian silt contribution to soils on the mountain slopes (Mt. Ślęża, SW Poland).
weathering on soil production. Geomorphology 204, 510–517. https://doi.org/10. Quat. Res. 89 (3), 702–717.
1016/J.GEOMORPH.2013.08.030. Waroszewski, J., Egli, M., Brandová, D., Christl, M., Kabala, C., Malkiewicz, M., Kierczak,
Nyland, K.E., Ignatov, A., Miller, B.A., 2018. A new depositional model for sand-rich loess J., Glina, B., Jezierski, P., 2018b. Identifying slope processes over time and their
on the Buckley Flats outwash plain, northwestern Lower Michigan. Aeolian Res. 31, imprint in soils of medium-high mountains of Central Europe (the Karkonosze
91–104. https://doi.org/10.1016/J.AEOLIA.2017.05.005. Mountains, Poland). Earth Surf. Process. Landf. 43 (6), 1195–1212.
Pawlak, W., 2008. Atlas of Lower and Opole Silesia, 2nd ed. Wroclaw University, Weber, J., 1982. Genesis and properties of soils derived from serpentinites in Lower
Wroclaw, Poland. Silesia. Part IV. Characteristics of colloidal fraction. Roczniki Gleboznawcze 33 (2),
Pędziwiatr, A., Kierczak, J., Waroszewski, J., Ratié, G., Quantin, C., Ponzever, E., 2018. 73–84 (in Polish).
Rock-type control of Ni, Cr, and Co phytoavailability in ultramafic soils. Plant Soil White, A.F., Blum, A.E., 1995. Effects of climate on chemical weathering in watersheds.
423, 339–362. Geochim. Cosmochim. Acta 59, 1729–1747.
Ponge, J.F., Sartori, G., Garlato, A., Ungaro, F., Zanella, A., Jabiol, B., Obber, S., 2014. White, A.F., Brantley, S.L., 2003. The effect of time on the weathering of silicate minerals:
The impact of parent material, climate, soil type and vegetation on Venetian forest why do weathering rates differ in the laboratory and field? Chem. Geol. 202 (3–4),
humus forms: a direct gradient approach. Geoderma 226–227. https://doi.org/10. 479–506.

997

You might also like