Next Article in Journal
A Topographic Perspective on the Propensity for Degradation of Plateau Swampy Meadows in Maduo County, West China
Next Article in Special Issue
Digital Mapping of Soil Properties Using Ensemble Machine Learning Approaches in an Agricultural Lowland Area of Lombardy, Italy
Previous Article in Journal
Establishment of a Monitoring Model for the Cotton Leaf Area Index Based on the Canopy Reflectance Spectrum
Previous Article in Special Issue
Insights into the Effects of Study Area Size and Soil Sampling Density in the Prediction of Soil Organic Carbon by Vis-NIR Diffuse Reflectance Spectroscopy in Two Forest Areas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pedodiversity and Organic Matter Stock of Soils Developed on Sandstone Formations in the Northern Apennines (Italy)

by
Livia Vittori Antisari
1,2,
William Trenti
1,
Alessandro Buscaroli
3,
Gloria Falsone
1,
Gilmo Vianello
2 and
Mauro De Feudis
1,*
1
Department of Agricultural and Food Sciences, Alma Mater Studiorum-University of Bologna, Via Fanin, 40, 40127 Bologna, Italy
2
Centro Sperimentale per lo Studio e l’Analisi del Suolo (CSSAS), Alma Mater Studiorum-University of Bologna, 40127 Bologna, Italy
3
Department of Biological, Geological, and Environmental Sciences, Alma Mater Studiorum-University of Bologna, Via S. Alberto 163, 48100 Ravenna, Italy
*
Author to whom correspondence should be addressed.
Land 2023, 12(1), 79; https://doi.org/10.3390/land12010079
Submission received: 2 December 2022 / Revised: 20 December 2022 / Accepted: 23 December 2022 / Published: 27 December 2022

Abstract

:
Pedodiversity is considered the cornerstone of biodiversity. This work aimed to (1) assess pedodiversity according to vegetation, topographic factors, and lithology and to (2) identify the major soil-forming factors on soil organic matter (SOM) stock at a 0–30 cm depth. These goals were reached using data from 147 georeferenced soil profiles distributed along 400–1000 m (≤1000) and 1000–2134 m (>1000) altitudinal gradients in the northern part of the Apennine chain in Italy. Soils showed mainly weak or incipient development (i.e., Entisols and Inceptisols), which could be attributed to sand-based lithology, high slope gradients, and low SOM accumulation rates, which promote soil erosion processes. However, higher pedodiversity was observed at >1000 m than at ≤1000 m, likely due to the higher vegetation cover diversity and climate variability; Spodosols and Mollisols were also found. A greater SOM stock was found at >1000 than ≤1000 m, and vegetation seemed to not affect SOM amounts, suggesting a greater influence of climate on SOM content compared to vegetation. Considering ecosystem conservation, the observed spatial pedodiversity could be considered a critical basis for the protection of soil resources and pedodiversity itself in mountain regions.

1. Introduction

Biodiversity evaluation, also referred to as spatial diversity, is a major topic in ecosystem studies and is mostly applied to aboveground flora and fauna [1,2] but also recently to belowground biodiversity [3,4]. Conversely, few investigations concerning pedodiversity were carried out [5,6,7], although pedodiversity is considered the cornerstone of biodiversity given the strong relationship between soil and flora and fauna [8,9].
Pedodiversity is extremely complex and driven by factors involved in several soil formation processes [10,11], making it difficult to quantify the influence of each factor on soil properties. Nowadays, because of the major role of soil in storing organic matter [12], several studies focused on identifying the factors that mostly significantly affect soil organic carbon (SOC) [13,14,15]. Indeed, the spatial changes in SOC amounts are dependent on biotic and abiotic factors such as climate, the amount and quality of plant debris, vegetation cover, and land use [16,17,18,19]. In addition, SOC stock is affected by soil development according to the different weight of each pedogenesis factor [20,21]. In this regard, recent papers highlighted the importance of soil type knowledge in terms of lithology, the physicochemical characterization of organic debris, and pedogenetic horizons to determine soil forming processes, in order to understand the distribution and quality of SOC stock [17,22].
In recent decades, there has been growing interest in the knowledge of mountain soils because montane ecosystems are characterized both to have high pedodiversity [23,24,25] and to be sensitive to climate change [26,27,28]. The soil development in mountain environments is mainly influenced by geomorphological dynamics [29]. The slope drives many processes, mainly associated with snow and water action. Erosion and soil accumulation processes shape the landscape and affect soil genesis, rejuvenating soils in erosion zones, making them thinner, or deepening soils in the accumulation areas [30]. Furthermore, mountain areas are currently experiencing strong climatic changes that are altering temperatures, precipitation amounts and patterns, and duration of the seasons. For example, reduced snow cover periods were reported to coincide with increasing occurrences of summer droughts and winter freeze–thaw cycles in soils [31]. Such climate changes modify the C amounts in mountain soils due to the modification of plant C inputs [32], leading to the alteration of SOC immobilization–mineralization processes [26] and further interfering in soil formation processes. Mountain ecosystems are often characterized by poorly developed, shallow, and skeletal soils [33]. Organic C content is particularly important for soil fertility and the efficient functioning of such ecosystems [34,35]. Although mountain soils store relevant amounts of organic C [22,26,36], the effect of the simultaneous impact of environmental factors on the amount and spatial dynamics of SOC in mountain areas is still underestimated. This knowledge gap is mainly related to the complexity of mountain ecosystems due to the climate conditioning of geobotanic zonality [37] and the vertical diversity of soil profiles and their biological activity [38,39,40,41]. Further, mountain regions are characterized by greater geodiversity [42] than any other landscape [43,44]. All these factors make soils of mountain areas highly diverse [33]. While elevation and, therefore, temperature differences were identified as the dominant controls on mountain SOC at regional scales [45,46,47], soil properties such as soil moisture, pH, and soil texture [27,45,48] and vegetation/land use [49,50,51] may introduce a large variability in mountain SOC at local scales.
As a whole, while several studies have reported the impact of lithology, vegetation type, and altitude on SOC stock, few studies have examined the effect of the aforementioned soil forming factors on soil N and other soil physical and chemical properties in mountain regions.
Given the little information about major factors affecting pedodiversity and soil properties (e.g., texture, pH, SOC, total N, exchangeable bases), changes in mountain ecosystems were determined using data from 147 soil profiles for georeferenced soils developed on sandstone formations and distributed in the northern part of the Apennine chain in Italy, at different altitudinal gradients from 400 to 2134 m above sea level (a.s.l.). The main aims of the present work were to: (1) assess the pedodiversity, meaning the classification of the soils themselves [52] and properties according to vegetation, morphological factors (i.e., altitude, slope, and exposure), and lithology and (2) identify the major soil forming factors on organic C and total N stocks at a 0–30 cm depth.

2. Materials and Methods

This study used a dataset of 147 georeferenced soil profiles (Table S1 of the Supplementary Materials) located in the Northern Apennines in Italy. Such soil profiles are the result of projects financed by public institutions and self-financed projects conducted in the last 15–20 years by the authors of the present manuscript whose results were published in several papers [53,54,55,56,57,58,59,60,61,62,63,64,65,66,67]. Specifically, the soil profiles were located at altitudes ranging from 200 to 2134 m a.s.l. and under different land uses, with 54 soil profiles under chestnut stands (CHE); 16 under conifer reforestation, mainly black pine (Pinus nigra), spruce (Picea abies), and Douglasia (Pseudotsuga menziesii) stands planted between World War II and the 1950s (CON); 30 under beech (Fagus sylvatica) forests (BEF); 18 under blueberry (Vaccinium myrtillus and Vaccinium ulginosum) habitats (BLB); 18 under mixed forests consisting mainly of abandoned chestnut stands invaded by other tree species (e.g., acacia, elm, oak, ash, and poplar) (MF); and 11 under grasslands (GS).

2.1. Parent Material

The sites under investigation were characterized by diverse sandstone types as parent rock materials (Figure 1) whose chemical features [68,69,70,71] are reported in Table A1 of Appendix A:
-
Monghidoro formation (MOH): Cretacic–Paleogene (External Ligurian Domain) thick-bedded turbidite sandstones, with microconglomerates at the base and alternating shales and siltstones;
-
Loiano formation (LOI): Paleogene (Epiligurian succession) thick-bedded, weakly cemented sandstones;
-
Castiglione dei Pepoli sandstones (GRA), dormant landslide deposits: Miocene (Cervarola–Falterona succession) turbiditic sandstones in shallow to medium-thick beds;
-
Mount Cervarola sandstones, Dardagna member (CEV): Neogene (Cervarola–Falterona succession) alternating arenaceous and pelitic turbidite deposits, with intercalations of marly-calcareous turbidites, silty marls, calcarenites, and black clays. In the investigated area, there is a thick bed of black pelites and black and grey sandstones;
-
Mount Modino sandstones (MOD): Paleogene–Neogene (Modino succession) arenaceous layers intercalated with thinly layered turbidites made of fine sandstones, siltstones, marls, or, rarely, shales;
-
Macigno sandstones (MAC): late Oligocene (Tuscan succession) thick turbidite arenaceous sandstones and siltstones with rare thin carbonate beds and chaotic beds.
-
Mount Adone formation (ADO): post-lower Pliocene (Pliocene and Pleistocene marine succession) prevailing sandstones, generally lightly cemented, with abundant silty-clay matrix and subordinate conglomerates, in medium layers to banks, alternating with pelitic sand with thin to thick stratifications.

2.2. Climate

The climate of the Northern Apennines is characterized by cold winters with frequent snowfalls and hot summers. The mean annual temperature of the study areas ranges from 14.6 °C at the lowest altitude to 3.3 °C at the highest (Table S2 of the Supplementary Materials). The lowest temperatures were recorded in January (1.18 °C on average) with the exception of the highest altitude, where the lowest temperatures were recorded in February (−3.3 °C), and the warmest ones were recorded in July and August. The mean annual cumulative precipitation was 1004 mm and ranged between 753 and 1703 mm (Table S1 of the Supplementary Materials) with an uneven distribution between the months. In addition, above 2000 m a.s.l., the snow covers the soils for about 6 months per year (from November to April), which also happens at Mount Cimone [64]. The mean air temperature and mean annual cumulative precipitation data were used to determine soil moisture and temperature regimes according to the Newhall Simulation Method [72]. With the approximations allowed by the Soil Taxonomy [52], it was possible to identify a mesic temperature regime between 200 and 1300 m a.s.l. and a frigid temperature regime between 1300 to 2000 m. The soil moisture regime was found to be udic between 200 and 1800 m a.s.l. and perudic over 1800 m of altitude (Table 1).

2.3. Soil Profiles

To evaluate the pedodiversity, soil profiles were dug until the parent materials were sampled and then analyzed via genetic horizons and classified according to Soil Survey Staff [52]. Due to the large spatial scale of the area considered in the present investigation, the soil profiles showed large morphological variability, with differences in the genetic horizon sequence, horizon thickness, and soil profile depth. To cope with these issues, in this study, the upper 30 cm of soil profiles was considered for evaluation of the soil physical–chemical properties, and the soil horizons included within this soil depth interval were grouped into three layers: layer A, which included the A horizons; layer B, which included the AB, BA, EA, EB, Bw, and BC horizons; and layer C, which included the CB and C horizons. As a whole, the minimum, maximum, and mean thickness was 1.0, 21.0, and 5.7 cm for layer A; 2.0, 23.0, and 10.1 cm for layer B; and 1.0, 27.0, and 14.1 cm for layer C, respectively.

2.4. Laboratory Analyses

The pH was determined potentiometrically in deionized water with a 1:2.5 soil:water ratio using a pH-meter electrode (Crison, Germany). The particle size distribution was determined via the pipette method described by Gee and Bauder [75], after the dispersion of soil samples with sodium hexametaphosphate solution. Organic C and total nitrogen (TOC and TN, respectively) contents were determined using a Thermo Scientific Lab EA-1110 dry combustion analyzer. Exchangeable Ca and other cations (Na, K, and Mg) and cation exchange capacity were determined through extraction with a 1 M NH4-acetate solution at pH 7 [76]. Cations were measured by inductively coupled plasma optical emission spectrometry (ICP-OES, Ametek, Arcos Spectro). Exchangeable Ca (Caexch) and the percentage of base saturation (BS) are reported in this study because of the well-known role of Ca exchange in promoting SOC storage [77] and the importance of BS for soil taxonomic classification and soil fertility.

2.5. Soil Organic Matter Stock Calculation

Organic C and TN stocks (SOC stock and TN stock, respectively) within a 0–30 cm soil depth were calculated by the sum of the SOC stock and TN stock of each horizon included within that interval depth using the equation in [78] and expressed as Mg ha−1:
SOC   stock   and   TN   stock   =   TOC   or   TN × BD × thickness × ( 1 sk 100 )
where TOC or TN represents the contents of total organic carbon (%) and total nitrogen (%), BD is the bulk density (g cm−3), and sk is the coarse fragment content (Ø > 2 mm; %) by volume.
BD was estimated according to the pedotransfer function proposed by Hollis et al. [79] and broadly used for soils located within the Italian landscapes [80,81,82,83,84]. The presence of coarse fragments was recorded during the field survey [85].

2.6. Statistical Method

To better detect the influence of the soil formation factors related to altitude, the dataset was divided into two distinct ones: a dataset that grouped the soils located between 400 and 1000 m a.s.l. (hereafter ≤ 1000 m a.s.l.) and another including soils from 1000 to 2134 m a.s.l. (hereafter > 1000 m a.s.l.). Further, a clear vegetation type change was observed at 1000 m a.s.l. Indeed, below 1000 of altitude, CHE and MF generally shared the same altitudinal range. Similarly, the vegetation types found between 1000 and 2000 m of altitude (i.e., CON, BEF, BLB, and GS), in most of the cases, shared the same altitude.
Within each altitudinal range and without considering the soil types (pedodiversity), a Kruskal–Wallis one-way ANOVA [86] was used to compare the effects of land-use and sandstone formation type on the investigated soil parameters. The Kruskal–Wallis test [86] was also used to verify the effect of soil types (pedodiversity) on soil parameters without considering vegetation cover.
Hierarchical cluster analysis (CA) using Euclidean distance and Ward grouping methods [87] was performed on the experimental data of the two altitudinal ranges (i.e., ≤1000 m a.s.l. and >1000 m a.s.l.) to cluster the collected soils in groups sharing similarities. The k-means method was then employed [87], dividing the soil samples into a predetermined number of clusters, with the smallest difference observed between elements within each cluster and the largest one between the elements of different clusters. The groups obtained by CA were processed using the Kruskal–Wallis test to identify differences in the investigated soil parameters. To identify the main drivers affecting the groups obtained by CA, principal component analysis (PCA) was carried out.
A Kruskal–Wallis test was used to evaluate the statistical diversity of OC and TN stocks among the different uses of soil and between the two altitude ranges (i.e., ≤1000 m a.s.l. and >1000 m a.s.l.).
The statistical analysis was conducted using the R software (version 4.2.1) [88].

3. Results

3.1. Pedodiversity and Physicochemical Properties at Lower and Higher Altitude Range

Most of the investigated soils were classified as Inceptisols and Entisols (Table S1 of the Supplementary Materials) [52]. Alfisols at low altitude (200 m a.s.l.), developed on the ADO sandstone type (Figure 1), were also found under CHE. However, because of the scarce number of replicates for soils from 200 to 400 m a.s.l., these samples were deleted from dataset and, therefore, their data were not further elaborated.

3.1.1. Lower Altitude Range

Within the ≤1000 m a.s.l. altitudinal range, soils were classified as Entisols (18%) and Inceptisols (82%). Most of the Entisols were Typic Udorthents (61%), which were mainly found within CHE, while Lithic Udorthents were less frequently found (39%) and were mainly observed under MF and within fields with slopes ≥ 30°. Most of the Inceptisols were classified as Typic Dystrudepts (74%) and mainly recorded under CHE. About 18% of Inceptisols were Lithic Dystrudepts, with most of them under CHE, and about 8% were Oxyaquic Dystrudepts, which were found under both CHE and MF. Eutrudepts were also found (18%), all of them under MF.
Taking into account the vegetation cover, CHE presented lower pH, BS values, and contents of Caexch, sand, and clay than MF (Table 2). No significant differences for the amounts of TOC and TN were observed (Table 2).
Comparing the soil subgroups, most of the investigated soils had a loam soil texture class (Table 3), with a trend toward a higher amount of sand observed among the Lithic Udorthents and Lithic Eutrudepts (sandy loam), with a higher amount of silt and lower amount of sand in Lithic Dystrudepts (silty loam) and a higher clay content and lower sand content in Typic Eutrudepts (sandy clay loam).
The soil subgroups identified at a ≤1000 m a.s.l. altitudinal range did not show differences in TOC and TN, although a higher C:N ratio was observed in Typic Dystrudepts compared to Lithic Dystrudepts and Lithic Udorthents. For the Caexch, BS, and pH, their values were generally lower in Oxyaquic and Typic Dystrudepts than in Eutrudepts and Entisols.

3.1.2. Higher Altitude Range

At >1000 m a.s.l., we observed higher pedodiversity compared to that of the lower altitudinal range dataset (≤1000 m a.s.l.) (Table 2). Soils were classified as Entisols, Inceptisols, and Spodosols (24, 53, and 8%, respectively). Mollisols were also found under mountain grassland (GS) and represented 4% of soils investigated within the >1000 m a.s.l. altitudinal range (Table 3). Specifically, Lithic Hapludolls were found under GS, showing only a mollic epipedon and no other diagnostic horizons due to the very shallow soil. Entisols were mainly represented by soils with udic and perudic moisture regimes and a temperature regime from frigid to cryic. Lithic Udorthents (27%) were found under CON and BEF, while Typic Udorthents (9%) were observed under all vegetation cover types (Table 2). A cryic soil temperature regime was observed under GS, where Lithic Cryorthents (4%) were identified.
Most of the Inceptisols had a dystric feature, namely, a base saturation of less than 60 percent at all subsurface horizons between 25 and 75 cm below the soil surface. Furthermore, about 25% of Inceptisols belonged to Typic Dystrudepts, and about 7% were Dystrudepts featuring a lithic contact within 50 cm of the soil surface and/or with an umbric or mollic epipedon less than 50 cm thick (Lithic Dystrudepts, Humic Dystrudepts and Humic Lithic Dystrudepts). Under BEF and CON, Eutrudepts (Humic Lithic Eutrudepts and Lithic Eutrudepts) were also found.
Taking into consideration the vegetation cover, the differences observed for both soil particle size distribution and TOC content were negligible. Notably, the lowest TN content and, as a consequence, the highest C:N ratio were found under BEF. BLB, instead, presented the lowest values of both Caexch content and BS. Finally, the highest pH values were observed under GS (Table 3).
Soils located at >1000 m a.s.l. belonged mostly to the sandy loam texture class (Table 4). Like at the ≤1000 m a.s.l. altitudinal range, no differences in TOC and TN were observed. The C:N ratio showed some differences, but no clear trend related to soil type was evident. However, as expected, Spodosols presented lower pH, Caexch, and BS than Mollisols.

3.2. Cluster Analysis (CA) and Principal Component Analysis (PCA)

3.2.1. Lower Altitude Range

CA was used to cluster the soil samples collected from each horizon of the study sites located at ≤1000 m a.s.l. into three clusters (Figure S1 of the Supplementary Materials). One cluster (hereafter called Cluster 1) included 45 statistical units mostly belonging to layer A. The largest cluster (hereafter called Cluster 2) had 145 statistical units including soil samples from layer B and layer C. The smallest cluster (hereafter called Cluster 3) mainly included soil samples collected under CHE, although the A horizons of CHE were also found. The Kruskal–Wallis test carried-out on the identified clusters showed the highest values of pH (5.9), TOC content (58.3 g kg−1), TN content (3.7 g kg−1), Caexch content (13.6 cmol (+) kg−1), BS content (66.4%), and clay content (15%) for the soil samples of Cluster 1. Between Cluster 2 and 3, some further differences were observed: TOC and TN contents were higher in Cluster 3 than in Cluster 2, while Cluster 2 had higher contents of sand and silt than Cluster 3 (Table 5).
The first two principal components of PCA, performed on the dataset at below 1000 m a.s.l. explained 51.0% of the total variance (Figure 2). PCA allowed us to identity pH, OC, TN, sand, silt, Caexch, and BS as the main soil parameters in the separation of the clusters (Figure 2).

3.2.2. Higher Altitude Range

The CA performed on soil horizons of the study sites located at >1000 m a.s.l. produced three clusters (Figure S2 of the Supplementary Materials). The smallest cluster (hereafter called Cluster I) had 35 statistical units and included mainly horizon A of the soils characterized by MOD and CERV sandstone types. The largest cluster (hereafter called Cluster II) had 130 statistical units and included deeper horizons. The third cluster (hereafter called Cluster III) had 54 statistical units and included the A horizons of soils developed on the other sandstone types. The highest pH value was observed in Cluster I (5.7) and the lowest one in Cluster III (4.3) (Table 6). Both TOC and TN showed the highest contents in Cluster III (81.4 and 5.4 g kg−1, respectively) and the lowest in Cluster II. Caexch and BS showed the highest values in Cluster I (13.2 cmol (+) kg−1 and 88.3%, respectively) and the lowest ones in Cluster II (3.0 cmol (+) kg−1 and 26.6%, respectively). For the particle size distribution, Cluster II showed the lowest sand content and the highest clay content (57 and 11%, respectively).
For the above 1000 m a.s.l. altitudinal range, the first two principal components of the PCA explained 47.1% of the total variance (Table 3). PCA showed the great influence of pH, TN, TOC, sand, silt, Caexch, and BS on the distinctions among the clusters identified by CA (Figure 3).

3.3. Organic C and Total N Stocks

SOC and TN stocks differed within altitude classes and vegetation covers (Table 7). Both SOC and TN stocks showed the lowest values at the ≤1000 m a.s.l. altitudinal range. These differences were confirmed by the vegetation covers of CHE and MF. The main covers at ≤1000 m a.s.l. showed lower stocks of SOC and TN compared to the vegetation types located at a higher altitudinal range (73.3 and 4.6, and 70.8 and 4.7 Mg ha−1 for CHE and MF, respectively). Within the higher altitudinal interval, the SOC stock did not differ, while a lower TN stock (7.2 Mg ha−1) was found under BEF than under BLB, GS, and CON.

4. Discussion

4.1. Factors Affecting Pedodiversity

The investigated soils are located in a large altitudinal range, and, as a consequence, different hydrological and temperature regimes are present (from udic to perudic and from mesic to cryic, respectively), defining the different climates and phytoclimatic zones (Figure 1 and Table 1).
In the investigated sites, the soils were mainly characterized by weak or incipient development, mainly including Entisols and Inceptisols (Table 2), which could be attributed to the sand-based lithology, high slope gradients, and low organic matter accumulation rates promoting soil erosion processes and, therefore, weak soil development. Specifically, sandstone weathering processes generate soils with low clay and base contents and a weak soil structure [61,67,89], making them vulnerable to erosion. With organic matter serving as the major actor for the formation and stabilization of soil aggregates [90,91], such vulnerability was further boosted by low clay content, which likely limited organic carbon retention within the mineral horizons [22,92,93] and led to thin A horizons (5.5 cm, on average). However, the incorporation of organic matter within the mineral horizons was likely further prevented by acidic conditions resulting in the accumulation of organic matter on the soil surface as organic horizons [94]. The weak soil development could be also attributed to the steepness of the study sites since most of the investigated soils had a slope gradient greater than 11°, which is considered the threshold for the occurrence of intense soil erosion [95]. Therefore, the sandstone lithology, weak soil organic matter accumulation, and high slope gradients of these mountain areas of the Apennines promoted the formation of young and poorly developed soils (i.e., Entisols and Inceptisols). Conversely, where erosion is limited, and favorable development conditions occur, such as the study sites at the highest elevations with north–northeast exposure, more intense pedogenesis occurs and is mainly driven by podzolization processes (i.e., Spodosols).
Taking into consideration the altitudinal ranges, higher pedodiversity was observed at both the order and subgroup levels in a range of >1000 m a.s.l. compared to that found in the range of ≤1000 m a.s.l. (Table 2). Because of the role of vegetation type on soil properties and development [96,97,98], the highest pedodiversity found at >1000 m a.s.l. might be attributable to the higher vegetation cover diversity compared to that at ≤1000 m a.s.l., as discussed in the following paragraphs.
However, the highest vegetation cover diversity observed at >1000 m a.s.l. could be, in turn, due to the higher climate variability in terms of the mean air temperature, which is in accordance with several previous studies [99,100], demonstrating the great influence of climate changes related to altitude on vegetation placement within the different phytoclimatic zones. Indeed, while from about 400 to 1000 m a.s.l., the mean air temperature decreased by about 2 °C, within the >1000 m a.s.l. altitudinal range, the temperature change was about 6 °C (Table S2 of the Supplementary Materials). Additionally, the different mean air temperatures recorded along the elevation gradient affected soil processes with the resulting development of different soil types [101,102,103].

4.2. Properties of Soils at 400–1000 m a.s.l.

The investigated area ranging between 400 and 1000 m a.s.l. belongs to the Castanetum phytoclimatic zone (Table 1), which has been populated since ancient times by chestnut stands [104]. Here, suitable soils for chestnut stands are present and are formed on sedimentary or siliceous soils [105] where acidic conditions predominate [106]. Acidic soils are, in fact, favorable for the growth of such plant species [22,107,108], with chestnut being susceptible to the presence of the carbonates that typically come from limestone and marl [109]. The investigated CHE soils were sampled from a mountain farm and had suitable physicochemical characteristics to produce chestnut plants with a good vegetative and adaptive state [110]. Indeed, marked dystric characteristics (Lithic, Oxyaquic, and Typic Dystrudepts) characterized by low values of pH, Caexch content, BS, and amount of the finest soil particles were monitored under chestnut stands (Table 2). Further, CHE was mainly found on soils developed from parent materials, which were mainly poor in Ca and clay particles (i.e., GRA and MOH). Thus, under CHE, very weak soil variability was observed, mainly because of the ancient human-based selection of soils for plant chestnut stands. Unlike CHE, natural or semi-natural MFs were more evenly distributed among the different investigated sandstone formations. Thus, even if sandstone lithology limited diversity at the level of soil order, the presence of marly pelitic intercalations within the sandstone formation could generate differences at the soil subgroup level. Besides dystric features, under MF, eutric features were found (Lithic and Typic Eutrudepts) in soil developed on marl-enriched sandstones. The presence of such CaCO3-rich intercalation promoted the development of soils with high pH values, Caexch content, and BS percentages (Table 3).
Independently from sandstone formations, nonparametric statistical investigations (i.e., CA and PCA) highlighted that the highest values of BS, TOC, TN, and Caexch were observed in the A horizon, (Table 5 and Figure 2). This would suggest intense processes within this horizon related to the organic C cycle. Specifically, despite the parent material, the higher values of Caexch content and BS found in the A horizon could be attributed to the degradation of plant residues [111], as supported by the higher TOC content [112,113]. Thus, organic C cycle occurring in the A horizon allowed us to enhance nutrient (i.e., Ca) concentrations at the soil surface in soils that developed from nutrient-poor parent materials. The A horizons, however, were quite thin, with their mean values of thickness being 5.5 cm (from 1 to 12 cm). The key role of the A horizons in improving soil properties coupled with their thin thickness highlighted the vulnerability of these soils to degradation in areas such as mountainous regions, where soil erosion processes can be intense.

4.3. Properties of Soils at the 1000–2134 m a.s.l. Altitudinal Range

The altitudinal range from 1000 to 2134 m a.s.l. included Fagetum, Picetum, and Alpinetum phytoclimatic zones. Different moisture and temperature regimes were recorded from udic to perudic and from mesic to frigid and cryic, respectively (Table 1). The large variability in phytoclimatic zones, soil moisture, and temperature regimes further promoted pedodiversity within the >1000 m a.s.l. altitudinal range. In the northern Apennines, above the timberline and close to the top of the mountains, the primary grasslands occurred as the end of the elevation sequence of both primary heaths and secondary grasslands [114]. The highest land here includes ridges and steep slopes lashed by winds with high speed that carry away the snowpack accumulating in other areas, increasing the risk of frozen of soils with a consequent effect on the vegetation [115]. Three soil profiles were studied in the Alpinetum phytoclimatic zone of Mount Cimone under Nardetum GS [116]. These profiles were classified as Lithic Cryorthents, characterized by BS > 70% due to the presence of Caexch on the soil exchange complex. Under Nardetum GS located at the lower altitudes of Corno alle Scale, Lithic Mollisols were found. In Mollisols, it was possible to highlight a greater thickness of the A horizon compared to that in Entisols, but the physicochemical features of these soils were very similar (Table 4). Above the tree line, shrub communities with a predominance of Ericaceae dominated the mountain heathland (BLB, Vaccinium myrtillus and Vaccinium gaultherioides Bijelow) [117,118]. BLB soils showed marked dystric characteristics that were affected by strong leaching (low Caexch and BS), which, in some cases, resulted in spodic characteristics (Spodic Lithic Dystrudepts) at Corno alle Scale, while Spodosols (Typic Haplorthod) were found at Mount Prado and Mount Cusna. At Mount Cimone, Vittori Antisari et al. [64] investigated paleosols containing, at varying depths (but never exceeding 30 cm), a carbon-enriched horizon (Aub) that separated the top of the soil profile, called modern, from the buried soil, called ancient, showing well-expressed spodic properties. Therefore, in the present study, only modern-developed soils were taken into consideration, thereby providing a rather simple sequence of horizons (A-Bw-C).
It was unexpected to observe the highest C:N ratio of soils under BEF (Table 2). Indeed, the relatively low C:N ratio of beech residues is commonly recognized [119,120]. The high C:N ratio due to the lowest TN content found in the soils under BEF might be attributable to the fast degradation rate of such residues [121,122]. Specifically, the high rates of degradation mobilize a large amount of N, which, if not promptly uptaken by plants, can be lost through leaching.
Similar to the comparison between vegetation types, the soil subgroups identified at >1000 m a.s.l. did not show differences in soil texture (Table 4) likely due to the even distribution of such soil types on the different lithologies. Taking into consideration the investigated soil chemical properties (i.e., TOC, TN, C:N, pH, BS, and Caexch), although some differences could be observed among the soil subgroups, a clear trend was not identified (Table 4) due to the wide distribution of these subgroups within the study area in terms of vegetation, topography, and climate.
As observed for soils at the ≤1000 m a.s.l. altitudinal range, the CA performed for the soils at higher altitudes showed differentiation between the A horizons and the subsurface ones (Table 6). As already observed for the lower altitude, in A horizons at higher altitudes (Cluster I and III), the C cycling facilitated improvement of soil nutrient content, increasing Ca and BS values compared to deeper horizons (Cluster II). However, unlike the comparison between the soil subgroup vegetation types, the CA highlighted the influence of lithology on the A horizon’s development. Indeed, the A horizons of soils developed on MOD and CERV showed the highest values of Caexch content, BS, and pH in accordance with the chemical composition of MOD and CERV (Table A1) compared to that of the other lithologies found at >1000 m a.s.l.

4.4. Organic C and Total N Stocks within 0–30 CM Soil Depth

According to previous studies [22,123], within the study area, greater SOC and TN stocks were found at >1000 m a.s.l. compared to ≤1000 (Table 7). Such results were confirmed by vegetation type; indeed, the lowest SOC and TN stocks were observed under CHE and MF, both located at ≤1000 m a.s.l. These findings indicate the greater influence of climate on SOC and TN contents compared to vegetation, as observed by previous studies [47,124]. Indeed, although the vegetation affected the organic matter quality [125,126], until now, the lower mean air temperatures occurring at higher elevations strongly reduced organic matter mineralization by the microbial community [127,128]. Overall, despite the higher SOC and TN stocks at >1000 m a.s.l. compared to those at ≤1000 m, it is important to highlight that our study sites featured soils highly vulnerable to organic C loss due to the aforementioned topographic and edaphic factors, namely, the low amount of clay particles and the generally high slope gradient [129,130,131,132].

5. Conclusions

The present study highlighted relatively young soils (mainly Entisols and Inceptisols) in the study area. The wide distribution of Entisols and Inceptisols could be attributable to the area’s sand-based lithology, high slope gradients, and low organic matter accumulation rates, which make soils vulnerable to erosion processes and, therefore, do not promote soil development. Taking into consideration the subgroup level of the Soil Taxonomy classification, higher pedodiversity was found at altitudes >1000 m a.s.l. compared to those at ≤1000 m, likely due to the higher vegetation diversity and larger mean air temperature ranges at the former altitude compared to the latter.
Within the ≤1000 m a.s.l. altitudinal range, chestnut stands were present in the Castanetum phytoclimatic zone and were characterized by soils with dystric features (Lithic, Oxyaquic, and Typic Dystrudepts) and low values of pH, Caexch content, BS, and the finest soil particles. Unlike chestnut, the mixed forests contained soils with both dystric and eutric features because such soils also developed on marl-enriched sandstone.
The lower mean air temperatures recorded at >1000 m a.s.l. compared to the lower altitudinal range promoted the accumulation of soil organic matter, which likely reduced the erosion processes and, therefore, promoted soil development. Indeed, at >1000 m, Mollisols and Spodosols were found in addition to Entisols and Inceptisols.
Taking into consideration the SOM stock, the present study demonstrated the greater role of climate (i.e., temperature) on soil organic matter compared to vegetation. Indeed, a clear difference was observed between the investigated altitudinal ranges, while the SOM stock did not show differences between the vegetation types included within each altitudinal range. However, we are aware that the results of the present study need to be interpreted with caution, as we did not compare similar vegetation at different altitudes since vegetation is affected by climate change with respect to altitude.
The spatial pedodiversity observed in the northern part of the Apennine chain in Italy could be considered a critical basis for the protection of soil resources and pedodiversity itself in mountain regions, which are characterized by high complexity related to topography, lithology, and vegetation.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/land12010079/s1, Figure S1: Dendrogram obtained via hierarchical cluster analysis performed on soil samples from 400–1000 m above sea level; Figure S2: Dendrogram obtained via hierarchical cluster analysis performed on soil samples at 1000–2134 m above sea level; Table S1: Soil type, coordinates, aspect, slope gradient, vegetation type, and altitude of plots under investigation; Table S2: Monthly and yearly mean air temperatures and cumulative precipitation.

Author Contributions

Conceptualization, L.V.A. and G.V.; methodology, L.V.A. and G.V.; validation, L.V.A., G.F. and M.D.F.; formal analysis, L.V.A. and W.T.; investigation, L.V.A., G.V., G.F. and A.B.; resources, L.V.A.; data curation, W.T.; writing—original draft preparation, M.D.F.; writing—review and editing, L.V.A., G.F. and M.D.F.; visualization, L.V.A. and M.D.F.; supervision, L.V.A. and M.D.F.; funding acquisition, L.V.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by “Apennine Project, grant number RFOVittori”, “CASTANI-CO, Emilia Romagna Regional rural development program 2014–2020–number 5015571”, “Boschiamo Emilia Romagna Regional rural development program 2014–2020–number 5112703”, “Biodiversamente Castagno, Emilia Romagna Regional rural development program 2014–2020–number 5015570”, and “Castagni parlanti, Emilia Romagna Regional rural development program 2014–2020–number 5111733”.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We are grateful to Andrea Simoni for his technical support.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Mean chemical composition of the sandstone formations according to Dinelli et al. [68], van de Kamp and Leake [69], and Gazzi [70,71].
Table A1. Mean chemical composition of the sandstone formations according to Dinelli et al. [68], van de Kamp and Leake [69], and Gazzi [70,71].
MACMODCEVGRAMOHLOIADO
SiO2 (%)67.162.2760.5763.4670.3372.4358.46
TiO2 (%)0.490.50.50.710.30.230.69
Al2O3 (%)12.9411.2512.5016.5713.7914.0113.11
Fe2O3 (%)4.24.113.914.872.021.435.27
MnO (%)0.070.070.10.130.050.020.13
MgO (%)3.463.433.084.631.091.072.12
CaO (%)2.335.087.090.350.730.217.78
Na2O (%)2.762.682.72.382.273.022.59
K2O (%)2.192.292.072.483.64.312.64
P2O5 (%)0.130.140.130.220.090.050.15
MAC: Macigno sandstones; MOD: Mount Modino sandstones; CEV: Mount Cervarola sandstones; GRA: Castiglione dei Pepoli sandstones; MOH: Monghidoro formation; LOI: Loiano formation; ADO: Mount Adone formation.

References

  1. Schneider, F.D.; Brose, U.; Rall, B.C.; Guill, C. Animal diversity and ecosystem functioning in dynamic food webs. Nat. Commun. 2016, 7, 12718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hanz, D.M.; Cutts, V.; Barajas-Barbosa, M.P.; Algar, A.C.; Beierkuhnlein, C.; Fernández-Palacios, J.M.; Field, R.; Kreft, H.; Steinbauer, M.J.; Weigelt, P.; et al. Climatic and biogeographical drivers of functional diversity in the flora of the Canary Islands. Glob. Ecol. Biogeogr. 2022, 31, 1313–1331. [Google Scholar] [CrossRef]
  3. Inkotte, J.; Bomfim, B.; da Silva, S.C.; Valadão, M.B.X.; da Rosa, M.G.; Viana, R.B.; D’Ângelo Rios, P.; Gatto, A.; Pereira, R.S. Linking soil biodiversity and ecosystem function in a Neotropical savanna. Appl. Soil Ecol. 2022, 169, 104209. [Google Scholar] [CrossRef]
  4. Peng, Y.; Holmstrup, M.; Schmidt, I.K.; De Schrijver, A.; Schelfhout, S.; Heděnec, P.; Zheng, H.; Bachega, L.R.; Yue, K.; Vesterdal, L. Litter quality, mycorrhizal association, and soil properties regulate effects of tree species on the soil fauna community. Geoderma 2022, 407, 115570. [Google Scholar] [CrossRef]
  5. Costantini, E.A.C.; Dazzi, C. The Soils of Italy; Springer: New York, NY, USA, 2013; pp. 295–302. ISBN 978-94-007-5641-0. [Google Scholar]
  6. Guo, Y.; Gong, P.; Amundson, R. Pedodiversity in the United States of America. Geoderma 2003, 117, 99–115. [Google Scholar] [CrossRef] [Green Version]
  7. Toomanian, N.; Jalalian, A.; Khademi, H.; Eghbal, M.K.; Papritz, A. Pedodiversity and pedogenesis in Zayandeh-rud Valley, Central Iran. Geomorphology 2006, 81, 376–393. [Google Scholar] [CrossRef]
  8. Sokołowska, J.; Józefowska, A.; Woźnica, K.; Zaleski, T. Interrelationship between soil depth and soil properties of Pieniny National Park forest (Poland). J. Mt. Sci. 2019, 16, 1534–1545. [Google Scholar] [CrossRef]
  9. Casalini, A.I.; Bouza, P.J.; Bisigato, A.J. Geomorphology, soil and vegetation patterns in an arid ecotone. Catena 2019, 174, 353–361. [Google Scholar] [CrossRef]
  10. Jenny, H. Factors of Soil Formation: A System of Quantitative Pedology; McGraw-Hill: New York, NY, USA, 1941. [Google Scholar]
  11. Bockheim, J.G.; Gennadiyev, A.N.; Hartemink, A.E.; Brevik, E.C. Soil-forming factors and Soil Taxonomy. Geoderma 2014, 226–227, 231–237. [Google Scholar] [CrossRef]
  12. Sorokin, A.; Owens, P.; Láng, V.; Jiang, Z.D.; Michéli, E.; Krasilnikov, P. “Black soils” in the Russian Soil Classification system, the US Soil Taxonomy and the WRB: Quantitative correlation and implications for pedodiversity assessment. Catena 2021, 196, 104824. [Google Scholar] [CrossRef]
  13. Jiménez-González, M.A.; Álvarez, A.M.; Carral, P.; Almendros, G. Influence of soil forming factors on the molecular structure of soil organic matter and carbon levels. Catena 2020, 189, 104501. [Google Scholar] [CrossRef]
  14. Gray, J.M.; Bishop, T.F.A.; Wilson, B.R. Factors controlling soil organic carbon stocks with depth in Eastern Australia. Soil Sci. Soc. Am. J. 2015, 79, 1741–1751. [Google Scholar] [CrossRef] [Green Version]
  15. Johnson, K.D.; Harden, J.; McGuire, A.D.; Bliss, N.B.; Bockheim, J.G.; Clark, M.; Nettleton-Hollingsworth, T.; Jorgenson, M.T.; Kane, E.S.; Mack, M.; et al. Soil carbon distribution in Alaska in relation to soil-forming factors. Geoderma 2011, 167–168, 71–84. [Google Scholar] [CrossRef] [Green Version]
  16. Adiyah, F.; Michéli, E.; Csorba, A.; Gebremeskel Weldmichael, T.; Gyuricza, C.; Ocansey, C.M.; Dawoe, E.; Owusu, S.; Fuchs, M. Effects of landuse change and topography on the quantity and distribution of soil organic carbon stocks on Acrisol catenas in tropical small-scale shade cocoa systems of the Ashanti region of Ghana. Catena 2022, 216, 106366. [Google Scholar] [CrossRef]
  17. Kögel-Knabner, I.; Amelung, W. Soil organic matter in major pedogenic soil groups. Geoderma 2021, 384, 114785. [Google Scholar] [CrossRef]
  18. Weissert, L.F.; Salmond, J.A.; Schwendenmann, L. Variability of soil organic carbon stocks and soil CO2 efflux across urban land use and soil cover types. Geoderma 2016, 271, 80–90. [Google Scholar] [CrossRef]
  19. Boudjabi, S.; Chenchouni, H. Soil fertility indicators and soil stoichiometry in semi-arid steppe rangelands. Catena 2022, 210, 105910. [Google Scholar] [CrossRef]
  20. Voltr, V.; Menšík, L.; Hlisnikovský, L.; Hruška, M.; Pokorný, E.; Pospíšilová, L. The soil organic matter in connection with soil properties and soil inputs. Agronomy 2021, 11, 779. [Google Scholar] [CrossRef]
  21. Gonçalves, D.R.P.; Sá, J.C.d.M.; Mishra, U.; Cerri, C.E.P.; Ferreira, L.A.; Furlan, F.J.F. Soil type and texture impacts on soil organic carbon storage in a sub-tropical agro-ecosystem. Geoderma 2017, 286, 88–97. [Google Scholar] [CrossRef]
  22. De Feudis, M.; Falsone, G.; Vianello, G.; Agnelli, A.; Vittori Antisari, L. Soil organic carbon stock assessment in forest ecosystems through pedogenic horizons and fixed depth layers sampling: What’s the best one? Land Degrad. Dev. 2022, 33, 1446–1458. [Google Scholar] [CrossRef]
  23. Philippis, J.D.; Marion, D.A. Biomechanical effects, lithological variations, and local pedodiversity in some forest soils of Arkansas. Geoderma 2004, 124, 73–89. [Google Scholar] [CrossRef]
  24. Ibáñez, J.J.; Feoli, E. Global Relationships of Pedodiversity and Biodiversity. Vadose Zone J. 2013, 12, vzj2012.0186. [Google Scholar] [CrossRef]
  25. Costantini, E.A.C.; L’Abate, G. Beyond the concept of dominant soil: Preserving pedodiversity in upscaling soil maps. Geoderma 2016, 271, 243–253. [Google Scholar] [CrossRef]
  26. Gutiérrez-Girón, A.; Díaz-Pinés, E.; Rubio, A.; Gavilán, R.G. Both altitude and vegetation affect temperature sensitivity of soil organic matter decomposition in Mediterranean high mountain soils. Geoderma 2015, 237–238, 1–8. [Google Scholar] [CrossRef]
  27. Djukic, I.; Zehetner, F.; Tatzber, M.; Gerzabek, M.H. Soil organic-matter stocks and characteristics along an alpine elevation gradient. J. Plant Nutr. Soil Sci. 2010, 173, 30–38. [Google Scholar] [CrossRef]
  28. Britton, A.J.; Helliwell, R.C.; Lilly, A.; Dawson, L.; Fisher, J.M.; Coull, M.; Ross, J. An integrated assessment of ecosystem carbon pools and fluxes across an oceanic alpine toposequence. Plant Soil 2011, 345, 287–302. [Google Scholar] [CrossRef]
  29. Masseroli, A.; Bollati, I.M.; Proverbio, S.S.; Pelfini, M.; Trombino, L. Soils as a useful tool for reconstructing geomorphic dynamics in high mountain environments: The case of the Buscagna stream hydrographic basin (Lepontine Alps). Geomorphology 2020, 371, 107442. [Google Scholar] [CrossRef]
  30. Egli, M.; Poulenard, J. Soils of Mountainous Landscapes. In International Encyclopedia of Geography: People, the Earth, Environment and Technology; Wiley: Hoboken, NJ, USA, 2016; pp. 1–10. [Google Scholar] [CrossRef]
  31. Gobiet, A.; Kotlarski, S.; Beniston, M.; Heinrich, G.; Rajczak, J.; Stoffel, M. 21st century climate change in the European Alps-A review. Sci. Total Environ. 2014, 493, 1138–1151. [Google Scholar] [CrossRef]
  32. Hagedorn, F.; Gavazov, K.; Alexander, J.M. Above- And belowground linkages shape responses of mountain vegetation to climate change. Science 2019, 365, 1119–1123. [Google Scholar] [CrossRef]
  33. Romeo, R.; Vita, A.; Manuelli, S.; Zanini, E.; Freppaz, M.; Stanchi, S. Understanding Mountain Soils: A contribution from Mountain Areas to the International Year of Soils 2015; The Food and Agriculture Organization (FAO): Rome, Italy, 2015; pp. 1–169. ISBN 978-92-5-108804-3. [Google Scholar]
  34. Prietzel, J.; Christophel, D. Organic carbon stocks in forest soils of the German alps. Geoderma 2014, 221–222, 28–39. [Google Scholar] [CrossRef]
  35. Szewczyk, A.; Kaniuczak, J.; Hajduk, E.; Knap, R. Physical and chemical properties of selected soils from the surroundings of the Magura National Park (southern Poland). Soil Sci. Annu. 2015, 66, 36–44. [Google Scholar] [CrossRef] [Green Version]
  36. Garcia-Pausas, J.; Casals, P.; Camarero, L.; Huguet, C.; Sebastià, M.T.; Thompson, R.; Romanyà, J. Soil organic carbon storage in mountain grasslands of the Pyrenees: Effects of climate and topography. Biogeochemistry 2007, 82, 279–289. [Google Scholar] [CrossRef]
  37. Gutman, G.; Ohring, G.; Joseph, J.H. Interaction between the geobotanic state and climate: A suggested approach and a test with a zonal model. J. Atmos. Sci. 1984, 41, 2663–2678. [Google Scholar] [CrossRef]
  38. Ma, H.-P.; Yang, X.-L.; Guo, Q.-Q.; Zhang, X.-J.; Zhou, C.-N. Soil organic carbon pool along different altitudinal level in the Sygera Mountains, Tibetan Plateau. J. Mt. Sci. 2016, 13, 476–483. [Google Scholar] [CrossRef]
  39. Ghosh, B.N.; Sharma, N.K.; Alam, N.M.; Singh, R.J.; Juyal, G.P. Elevation, slope aspect and integrated nutrient management effects on crop productivity and soil quality in North-west Himalayas, India. J. Mt. Sci. 2014, 11, 1208–1217. [Google Scholar] [CrossRef]
  40. Jobbagy, E.G.; Jackson, R.B. The vertical distribution of soil organic carbon and its relation to climate and vegetation. Ecol. Appl. 2000, 10, 423. [Google Scholar] [CrossRef]
  41. Józefowska, A.; Miechówka, A. Enzymatic activity and enchytraeids abundance in agricultural mountain soils. Soil Sci. Annu. 2015, 66, 133–138. [Google Scholar] [CrossRef] [Green Version]
  42. Gray, M. Geodiversity: A significant, multi-faceted and evolving, geoscientific paradigm rather than a redundant term. Proc. Geol. Assoc. 2021, 132, 605–619. [Google Scholar] [CrossRef]
  43. Slaymaker, O.; Spencer, T.; Embleton-Hamann, C. Geomorphology and Global Environmental Change; Cambridge University Press: Cambridge, UK, 2015; ISBN 9780511627057. [Google Scholar]
  44. Johnstone, J. Alpine plant life: Functional plant ecology of high mountain ecosystems. By Christian Körner. Mt. Res. Dev. 2021, 41, M1–M2. [Google Scholar] [CrossRef]
  45. Bargali, V.K.; Bargali, S.S. Effect of size and altitude on soil organic carbon stock in homegarden agroforestry system in Central Himalaya, India. Acta Ecol. Sin. 2020, 40, 483–491. [Google Scholar] [CrossRef]
  46. Bangroo, S.A.; Najar, G.R.; Achin, E.; Truong, P.N. Application of predictor variables in spatial quantification of soil organic carbon and total nitrogen using regression kriging in the North Kashmir forest Himalayas. Catena 2020, 193, 104632. [Google Scholar] [CrossRef]
  47. Massaccesi, L.; De Feudis, M.; Leccese, A.; Agnelli, A. Altitude and vegetation affect soil organic carbon, basal respiration and microbial biomass in apennine forest soils. Forests 2020, 11, 710. [Google Scholar] [CrossRef]
  48. Leifeld, J.; Bassin, S.; Fuhrer, J. Carbon stocks in Swiss agricultural soils predicted by land-use, soil characteristics, and altitude. Agric. Ecosyst. Environ. 2005, 105, 255–266. [Google Scholar] [CrossRef]
  49. Canedoli, C.; Ferrè, C.; Abu El Khair, D.; Comolli, R.; Liga, C.; Mazzucchelli, F.; Proietto, A.; Rota, N.; Colombo, G.; Bassano, B.; et al. Evaluation of ecosystem services in a protected mountain area: Soil organic carbon stock and biodiversity in alpine forests and grasslands. Ecosyst. Serv. 2020, 44, 101135. [Google Scholar] [CrossRef]
  50. Yimer, F.; Ledin, S.; Abdelkadir, A. Soil organic carbon and total nitrogen stocks as affected by topographic aspect and vegetation in the Bale Mountains, Ethiopia. Geoderma 2006, 135, 335–344. [Google Scholar] [CrossRef]
  51. Gebrehiwot, K.; Desalegn, T.; Woldu, Z.; Demissew, S.; Teferi, E. Soil organic carbon stock in Abune Yosef afroalpine and sub-afroalpine vegetation, northern Ethiopia. Ecol. Process. 2018, 7, 6. [Google Scholar] [CrossRef] [Green Version]
  52. Soil survey staff. Keys to Soil Taxonomy, 12th ed.; USDA—Natural Resources Conservation Service: Washington, DC, USA, 2014. [Google Scholar]
  53. Brombin, V.; Mistri, E.; De Feudis, M.; Forti, C.; Salani, G.M.; Natali, C.; Falsone, G.; Vittori Antisari, L.; Bianchini, G. Soil carbon investigation in three pedoclimatic and agronomic settings of northern Italy. Sustainability 2020, 12, 539. [Google Scholar] [CrossRef]
  54. Vittori Antisari, L.; Vianello, G.; Pontalti, F.; Lorito, S.; Gherardi, M. Land use effects on organic matter in brown soils of the emilian apennines. In The Soils of Tomorrow: Soils Changing in a Changing World; CAB International: Wallingford, UK, 2008; pp. 311–328. [Google Scholar]
  55. Falsone, G.; Marinari, S.; Vittori Antisari, L.; Vianello, G. Soil processes related to organic matter modifications following Douglas-fir mature reforestation. Biol. Fertil. Soils 2015, 51, 277–287. [Google Scholar] [CrossRef]
  56. De Feudis, M.; Falsone, G.; Vianello, G.; Vittori Antisari, L. The conversion of abandoned chestnut forests to managed ones does not affect the soil chemical properties and improves the soil microbial biomass activity. Forests 2020, 11, 786. [Google Scholar] [CrossRef]
  57. De Feudis, M.; Falsone, G.; Vianello, G.; Vittori Antisari, L. Stable organic carbon pool rises in soil under chestnut (Castanea sativa Mill.) forest for timber production after 15 years since grafting onto satin-cut stumps. EQA—Int. J. Environ. Qual. 2020, 40, 1–10. [Google Scholar]
  58. De Feudis, M.; Selmi, C.; Falsone, G.; Missere, D.; Di Bonito, M.; Vittori Antisari, L. The importance of incorporating soil in the life cycle assessment procedure to improve the sustainability of agricultural management. Catena 2022, 218, 106563. [Google Scholar] [CrossRef]
  59. Dell’Abate, M.T.; Giovannini, C.; Pontalti, F.; Vianello, G.; Vittori Antisari, L. Caratterizzazione quali-quantitativa della sostanza organica degli epipedon di suoli bruni acidi dell’appennino emiliano sotto bosco di latifoglie e di conifere. In Atti Convegno Nazionale Società Italiana di Scienze del Suolo “Suolo Ambiente Paesaggio”; Gessa, C., Lorito, S., Vianello, G., Vittori Antisari, L., Eds.; Società Italiana di Scienza del Suolo: Palermo, Italy, 2007; Volume 1, pp. 332–342. ISBN 978.88.902831-1-6. [Google Scholar]
  60. Vittori Antisari, L.; Dell’Abate, M.T.; Buscaroli, A.; Gherardi, M.; Nisini, L.; Vianello, G. Role of soil organic matter characteristics in a pedological survey: “Bosco Frattona” natural reserve (Site of Community Importance, Italy) case study. Geoderma 2010, 156, 302–315. [Google Scholar] [CrossRef]
  61. Vittori Antisari, L.; Marinari, S.; Dell’Abate, M.T.; Baffi, C.; Vianello, G. Plant cover and epipedon SOM stability as factors affecting brown soil profile development and microbial activity. Geoderma 2011, 161, 212–224. [Google Scholar] [CrossRef]
  62. Vittori Antisari, L.; Falsone, G.; Carbone, S.; Vianello, G. Short-term effects of forest recovery on soil carbon and nutrient availability in an experimental chestnut stand. Biol. Fertil. Soils 2013, 49, 165–173. [Google Scholar] [CrossRef]
  63. Vittori Antisari, L.; Falsone, G.; Carbone, S.; Marinari, S.; Vianello, G. Douglas-fir reforestation in North Apennine (Italy): Performance on soil carbon sequestration, nutrients stock and microbial activity. Appl. Soil Ecol. 2015, 86, 82–90. [Google Scholar] [CrossRef]
  64. Vittori Antisari, L.; Agnelli, A.; Corti, G.; Falsone, G.; Ferronato, C.; Marinari, S.; Vianello, G. Modern and ancient pedogenesis as revealed by Holocene fire—Northern Apennines, Italy. Quat. Int. 2018, 467, 264–276. [Google Scholar] [CrossRef]
  65. Vittori Antisari, L.; Trenti, W.; De Feudis, M.; Bianchini, G.; Falsone, G. Soil quality and organic matter pools in a temperate climate (Northern italy) under different land uses. Agronomy 2021, 11, 1815. [Google Scholar] [CrossRef]
  66. Marinari, S.; Bonifacio, E.; Moscatelli, M.C.; Falsone, G.; Vittori Antisari, L.; Vianello, G. Soil development and microbial functional diversity: Proposal for a methodological approach. Geoderma 2013, 192, 437–445. [Google Scholar] [CrossRef]
  67. Marinari, S.; Vittori Antisari, L. Effect of lithological substrate on microbial biomass and enzyme activity in brown soil profiles in the northern Apennines (Italy). Pedobiologia 2010, 53, 313–320. [Google Scholar] [CrossRef]
  68. Dinelli, E.; Lucchini, F.; Mordenti, A.; Paganelli, L. Geochemistry of Oligocene-Miocene sandstones of the northern Apennines (Italy) and evolution of chemical features in relation to provenance changes. Sediment. Geol. 1999, 127, 193–207. [Google Scholar] [CrossRef]
  69. van de Kamp, P.C.; Leake, B.E. Petrology and geochemistry of siliciclastic rocks of mixed feldspathic and ophiolitic provenance in the Northern Apennines, Italy. Chem. Geol. 1995, 122, 1–20. [Google Scholar] [CrossRef]
  70. Gazzi, P. Ordine di apparizione dei minerali pesanti nella formazione di Monghidoro e nella serie di Loiano (Appennino Bolognese). Mineral. Petrogr. Acta 1963, 9, 79–95. [Google Scholar]
  71. Gazzi, P. Le arenarie del flysch sopracretaceo dell’Appennino modenese; correlazioni con il flysch di Monghidoro. Mineral. Petrogr. Acta 1966, 12, 69–97. [Google Scholar]
  72. Van Wambeke, A.; Hastings, P.; Tolomeo, M. Newhall Simulation Model: A BASIC Program for the IBM PC; Dept. of Agronomy, Cornell University: Ithaca, NY, USA, 1986. [Google Scholar]
  73. De Philippis, A. Forest ecology and phytoclimatology. Unasylva 1951, 5, 1. [Google Scholar]
  74. De Philippis, A. Classificazioni ed eindici del clima, in rapporto alla vegetazione forestale italiana. Plant Biosyst. 1937, 44, 1–169. [Google Scholar]
  75. Gee, G.W.; Bauder, J.W. Methods of Soil Analysis: Part 1—Physical and Mineralogical Methods; SSSA Book Series; Soil Science Society of America, American Society of Agronomy: Madison, WI, USA, 1986; Volume sssabookse, ISBN 978-0-89118-864-3. [Google Scholar]
  76. Soil Survey Laboratory Staff. Soil Survey Laboratory Staff. Soil survey laboratory methods manual. In Soil Survey Investigations Report; USDA-SCS: Washington, DC, USA, 2004; Volume 42, p. 700. [Google Scholar]
  77. Rowley, M.C.; Grand, S.; Verrecchia, É.P. Calcium-mediated stabilisation of soil organic carbon. Biogeochemistry 2018, 137, 27–49. [Google Scholar] [CrossRef] [Green Version]
  78. FAO. Measuring and modelling soil carbon stocks and stock changes in livestock production systems: Guidelines for assessment (version 1). In Livestock Environmental Assessment and Performance (LEAP) Partnership; FAO: Rome, Italy, 2019; 170p. [Google Scholar]
  79. Hollis, J.M.; Hannam, J.; Bellamy, P.H. Empirically-derived pedotransfer functions for predicting bulk density in European soils. Eur. J. Soil Sci. 2012, 63, 96–109. [Google Scholar] [CrossRef]
  80. Garlato, A.; Obber, S.; Vinci, I.; Sartori, G.; Manni, G. Stock attuale di carbonio organico nei suoli di montagna del Veneto. In Studi Trentini di Scienze Naturali; Museo Tridentino di Scienze Naturali: Trento, Italy, 2009; Volume 85, pp. 69–81. [Google Scholar]
  81. Conforti, M.; Matteucci, G.; Buttafuoco, G. Organic carbon and total nitrogen topsoil stocks, biogenetic natural reserve ‘Marchesale’ (Calabria region, southern Italy). J. Maps 2016, 13, 92–99. [Google Scholar] [CrossRef]
  82. Brenna, S.; Stolbovoy, V.; Rocca, R.; Sciaccaluga, M. Potential carbon sequestration of Lombardy soils (Italy). In Proceedings of the 19th World Congress of Soil Science, Soil Solutions for a Changing World, Brisbane, Australia, 1–6 August 2010; pp. 70–73. [Google Scholar]
  83. Ungaro, F.; Staffilani, F.; Tarocco, P. Assessing and mapping topsoil organic carbon stock at regional scale: A scorpan kriging approach conditional on soil map delineations and land use. Land Degrad. Dev. 2010, 21, 565–581. [Google Scholar] [CrossRef]
  84. Faggian, V.; Bini, C.; Zilioli, D.M. Carbon stock evaluation from topsoil of forest stands in ne Italy. Int. J. Phytoremediat. 2012, 14, 415–428. [Google Scholar] [CrossRef] [Green Version]
  85. Food and Agriculture Organization of the United Nations (FAO). Guidelines for Soil Description, 4th ed.; FAO: Rome, Italy, 2006. [Google Scholar]
  86. Hollander, M.; Wolfe, D.A.; Chicken, E. Nonparametric Statistical Methods; John Wiley & Sons: New York, NY, USA, 2013. [Google Scholar]
  87. Murtagh, F.; Legendre, P. Ward’s hierarchical agglomerative clustering method: Which algorithms implement ward’s criterion? J. Classif. 2014, 31, 274–295. [Google Scholar] [CrossRef] [Green Version]
  88. RStudio. Rstudio Team Rstudio: Integrated Development Environment for R 2021. 2021. Available online: https://www.rstudio.com/products/rstudio/ (accessed on 18 October 2022).
  89. Evans, D.L.; Quinton, J.N.; Tye, A.M.; Rodés; Rushton, J.C.; Davies, J.A.C.; Mudd, S.M. How the composition of sandstone matrices affects rates of soil formation. Geoderma 2021, 401, 115337. [Google Scholar] [CrossRef]
  90. Sithole, N.J.; Magwaza, L.S.; Thibaud, G.R. Long-term impact of no-till conservation agriculture and N-fertilizer on soil aggregate stability, infiltration and distribution of C in different size fractions. Soil Tillage Res. 2019, 190, 147–156. [Google Scholar] [CrossRef]
  91. Sarker, T.C.; Incerti, G.; Spaccini, R.; Piccolo, A.; Mazzoleni, S.; Bonanomi, G. Linking organic matter chemistry with soil aggregate stability: Insight from 13C NMR spectroscopy. Soil Biol. Biochem. 2018, 117, 175–184. [Google Scholar] [CrossRef]
  92. Islam, M.R.; Singh, B.; Dijkstra, F.A. Stabilisation of soil organic matter: Interactions between clay and microbes. Biogeochemistry 2022, 160, 145–158. [Google Scholar] [CrossRef]
  93. Schweizer, S.A.; Bucka, F.B.; Graf-Rosenfellner, M.; Kögel-Knabner, I. Soil microaggregate size composition and organic matter distribution as affected by clay content. Geoderma 2019, 355, 113901. [Google Scholar] [CrossRef]
  94. Zanella, A.; Ponge, J.F.; Jabiol, B.; Sartori, G.; Kolb, E.; Gobat, J.M.; Le Bayon, R.C.; Aubert, M.; De Waal, R.; Van Delft, B.; et al. Humusica 1, article 4: Terrestrial humus systems and forms—Specific terms and diagnostic horizons. Appl. Soil Ecol. 2018, 122, 56–74. [Google Scholar] [CrossRef]
  95. García-Ruiz, J.M.; Beguería, S.; Nadal-Romero, E.; González-Hidalgo, J.C.; Lana-Renault, N.; Sanjuán, Y. A meta-analysis of soil erosion rates across the world. Geomorphology 2015, 239, 160–173. [Google Scholar] [CrossRef] [Green Version]
  96. D’Amico, M.E.; Freppaz, M.; Filippa, G.; Zanini, E. Vegetation influence on soil formation rate in a proglacial chronosequence (Lys Glacier, NW Italian Alps). Catena 2014, 113, 122–137. [Google Scholar] [CrossRef]
  97. Maxbauer, D.P.; Feinberg, J.M.; Fox, D.L.; Nater, E.A. Response of pedogenic magnetite to changing vegetation in soils developed under uniform climate, topography, and parent material. Sci. Rep. 2017, 7, 17575. [Google Scholar] [CrossRef] [Green Version]
  98. Cerqueira, B.; Vega, F.A.; Silva, L.F.O.; Andrade, L. Effects of vegetation on chemical and mineralogical characteristics of soils developed on a decantation bank from a copper mine. Sci. Total Environ. 2012, 421–422, 220–229. [Google Scholar] [CrossRef] [PubMed]
  99. Bruun, H.H.; Moen, J.; Virtanen, R.; Grytnes, J.-A.; Oksanen, L.; Angerbjörn, A. Effects of altitude and topography on species richness of vascular plants, bryophytes and lichens in alpine communities. J. Veg. Sci. 2006, 17, 37. [Google Scholar] [CrossRef]
  100. Gastauer, M.; Thiele, J.; Porembski, S.; Neri, A.V. How do altitude and soil properties influence the taxonomic and phylogenetic structure and diversity of Brazilian páramo vegetation? J. Mt. Sci. 2020, 17, 1045–1057. [Google Scholar] [CrossRef]
  101. Pintaldi, E.; Viglietti, D.; D’Amico, M.E.; Magnani, A.; Freppaz, M. Abiotic parameters and pedogenesis as controlling factors for soil C and N cycling along an elevational gradient in a subalpine Larch forest (NW Italy). Forests 2019, 10, 614. [Google Scholar] [CrossRef]
  102. Egli, M.; Mirabella, A.; Mancabelli, A.; Sartori, G. Weathering of soils in Alpine areas as influenced by climate and parent material. Clays Clay Miner. 2004, 52, 287–303. [Google Scholar] [CrossRef]
  103. De Feudis, M.; Cardelli, V.; Massaccesi, L.; Bol, R.; Willbold, S.; Cocco, S.; Corti, G.; Agnelli, A. Effect of beech (Fagus sylvatica L.) rhizosphere on phosphorous availability in soils at different altitudes (Central Italy). Geoderma 2016, 276, 53–63. [Google Scholar] [CrossRef]
  104. Pezzi, G.; Maresi, G.; Conedera, M.; Ferrari, C. Woody species composition of chestnut stands in the Northern Apennines: The result of 200 years of changes in land use. Landsc. Ecol. 2011, 26, 1463–1476. [Google Scholar] [CrossRef]
  105. Pereira, M.G.; Caramelo, L.; Gouveia, C.; Gomes-Laranjo, J.; Magalhães, M. Assessment of weather-related risk on chestnut productivity. Nat. Hazards Earth Syst. Sci. 2011, 11, 2729–2739. [Google Scholar] [CrossRef] [Green Version]
  106. Arrobas, M.; Afonso, S.; Rodrigues, M.Â. Diagnosing the nutritional condition of chestnut groves by soil and leaf analyses. Sci. Hortic. 2018, 228, 113–121. [Google Scholar] [CrossRef]
  107. De Feudis, M.; Falsone, G.; Vittori Antisari, L. Mid-term (30 years) changes of soil properties under chestnut stands due to organic residues management: An integrated study. Catena 2021, 198, 105021. [Google Scholar] [CrossRef]
  108. Melicharová, L.; Vizoso-Arribe, O. Situation of sweet chestnut (Castanea sativa Mill.) in Spain, galicia: A review. Sci. Agric. Bohem. 2012, 2012, 78–84. [Google Scholar]
  109. Papaioannou, E.; Kostopoulou, S.; Stefanou, S. The effect of the conversion of chestnut (Castanea sativa Mill.) forests to orchards on soil fertility and nutrient content in leaves. Catena 2022, 211, 105948. [Google Scholar] [CrossRef]
  110. Mellano, M.G.; Beccaro, G.L.; Donno, D.; Marinoni, D.T.; Boccacci, P.; Canterino, S.; Cerutti, A.K.; Bounous, G. Castanea spp. biodiversity conservation: Collection and characterization of the genetic diversity of an endangered species. Genet. Resour. Crop. Evol. 2012, 59, 1727–1741. [Google Scholar] [CrossRef]
  111. Wang, L.; Chen, Y.; Zhou, Y.; Zheng, H.; Xu, Z.; Tan, B.; You, C.; Zhang, L.; Li, H.; Guo, L.; et al. Litter chemical traits strongly drove the carbon fractions loss during decomposition across an alpine treeline ecotone. Sci. Total Environ. 2021, 753, 142287. [Google Scholar] [CrossRef] [PubMed]
  112. Gruba, P.; Mulder, J. Tree species affect cation exchange capacity (CEC) and cation binding properties of organic matter in acid forest soils. Sci. Total Environ. 2015, 511, 655–662. [Google Scholar] [CrossRef]
  113. Ramos, F.T.; Dores, E.F.d.C.; Weber, O.L.d.S.; Beber, D.C.; Campelo, J.H.; Maia, J.C.d.S. Soil organic matter doubles the cation exchange capacity of tropical soil under no-till farming in Brazil. J. Sci. Food Agric. 2018, 98, 3595–3602. [Google Scholar] [CrossRef] [PubMed]
  114. Pignatti, S. The climax vegetation above timberline in the northern and central Apennines. Fitosociologia 1994, 26, 5–17. [Google Scholar]
  115. Tomaselli, M.; Carbognani, M.; Foggi, B.; Petraglia, A.; Rossi, G.; Lombardi, L.; Gennai, M. The primary grasslands of the northern Apennine summits (N-Italy): A phytosociological and ecological survey. Tuexenia 2019, 39, 181–213. [Google Scholar] [CrossRef]
  116. Benatti, A.; Bal, M.; Allée, P.; Bosi, G.; Mercuri, A.M. Plant landscape reconstruction above the current timberline at the Monte Cimone and Corno alle Scale mountain areas (Northern Apennines, Italy) during the Late Holocene: The evidence from soil charcoal. Holocene 2019, 29, 1767–1781. [Google Scholar] [CrossRef]
  117. Tomaselli, M.; Del Prete, C.; Manzini, M.L. Parco Regionale Dell’alto Appennino Modenese: L’ambiente Vegetale; Regione Emilia-Romagna: Bologna, Italy, 1996. [Google Scholar]
  118. Alessandrini, A.; Foggi, B.; Rossi, G.; Tomaselli, M. La Flora di Altitudine dell’Appennino Tosco-Emiliano; Regione Emilia-Romagna: Bologna, Italy, 2003; p. 329. [Google Scholar]
  119. Lorenz, M.; Derrien, D.; Zeller, B.; Udelhoven, T.; Werner, W.; Thiele-Bruhn, S. The linkage of 13C and 15N soil depth gradients with C:N and O:C stoichiometry reveals tree species effects on organic matter turnover in soil. Biogeochemistry 2020, 151, 203–220. [Google Scholar] [CrossRef]
  120. Spohn, M.; Chodak, M. Microbial respiration per unit biomass increases with carbon-to-nutrient ratios in forest soils. Soil Biol. Biochem. 2015, 81, 128–133. [Google Scholar] [CrossRef]
  121. Osei, R.; Titeux, H.; Bielak, K.; Bravo, F.; Collet, C.; Cools, C.; Cornelis, J.T.; Heym, M.; Korboulewsky, N.; Löf, M.; et al. Tree species identity drives soil organic carbon storage more than species mixing in major two-species mixtures (pine, oak, beech) in Europe. For. Ecol. Manage. 2021, 481, 118752. [Google Scholar] [CrossRef]
  122. Mora, J.L.; Molina-Clerencia, M.; Girona-García, A.; Martí-Dalmau, C.; Badía-Villas, D. Factors controlling the buildup of humus and particulate organic matter in European beech and Scots pine stands at their southernmost distribution limits (Moncayo Massif, Spain). Geoderma 2021, 401, 115211. [Google Scholar] [CrossRef]
  123. Tashi, S.; Singh, B.; Keitel, C.; Adams, M. Soil carbon and nitrogen stocks in forests along an altitudinal gradient in the eastern Himalayas and a meta-analysis of global data. Glob. Chang. Biol. 2016, 22, 2255–2268. [Google Scholar] [CrossRef] [PubMed]
  124. Oueslati, I.; Allamano, P.; Bonifacio, E.; Claps, P. Vegetation and topographic control on spatial variability of soil organic carbon. Pedosphere 2013, 23, 48–58. [Google Scholar] [CrossRef] [Green Version]
  125. Díaz-Pinés, E.; Rubio, A.; Van Miegroet, H.; Montes, F.; Benito, M. Does tree species composition control soil organic carbon pools in Mediterranean mountain forests? For. Ecol. Manag. 2011, 262, 1895–1904. [Google Scholar] [CrossRef]
  126. Kumar, A.; Sharma, M.P.; Taxak, A.K. Effect of vegetation communities and altitudes on the soil organic carbon stock in Kotli Bhel-1A catchment, India. Clean—Soil Air Water 2017, 45, 1600650. [Google Scholar] [CrossRef]
  127. Tewksbury, C.E.; Van Miegroet, H. Soil organic carbon dynamics along a climatic gradient in a southern Appalachian spruce-fir forest. Can. J. For. Res. 2007, 37, 1161–1172. [Google Scholar] [CrossRef]
  128. Wiesmeier, M.; Urbanski, L.; Hobley, E.; Lang, B.; von Lützow, M.; Marin-Spiotta, E.; van Wesemael, B.; Rabot, E.; Ließ, M.; Garcia-Franco, N.; et al. Soil organic carbon storage as a key function of soils—A review of drivers and indicators at various scales. Geoderma 2019, 333, 149–162. [Google Scholar] [CrossRef]
  129. Clark, K.E.; Stallard, R.F.; Murphy, S.F.; Scholl, M.A.; González, G.; Plante, A.F.; McDowell, W.H. Extreme rainstorms drive exceptional organic carbon export from forested humid-tropical rivers in Puerto Rico. Nat. Commun. 2022, 13, 2058. [Google Scholar] [CrossRef]
  130. Angst, G.; Pokorný, J.; Mueller, C.W.; Prater, I.; Preusser, S.; Kandeler, E.; Meador, T.; Straková, P.; Hájek, T.; van Buiten, G.; et al. Soil texture affects the coupling of litter decomposition and soil organic matter formation. Soil Biol. Biochem. 2021, 159, 108302. [Google Scholar] [CrossRef]
  131. La Manna, L.; Tarabini, M.; Gomez, F.; Rostagno, C.M. Changes in soil organic matter associated with afforestation affect erosion processes: The case of erodible volcanic soils from Patagonia. Geoderma 2021, 403, 115265. [Google Scholar] [CrossRef]
  132. Heger, A.; Becker, J.N.; Vásconez Navas, L.K.; Eschenbach, A. Factors controlling soil organic carbon stocks in hardwood floodplain forests of the lower middle Elbe River. Geoderma 2021, 404, 115389. [Google Scholar] [CrossRef]
Figure 1. Location of the study sites comprising the database of the investigation.
Figure 1. Location of the study sites comprising the database of the investigation.
Land 12 00079 g001
Figure 2. Principal component analysis (PCA) obtained from the dataset at a 400–1000 m altitudinal range. The distinct colors indicate the three clusters distributed along the principal component (Dim) 1 and 2. The sample codes included the sampling area, profile number, layers, and vegetation cover. The variables are in black. BS = base saturation; Ca = exchangeable Ca; TN = total N; OC = soil organic C; CN = OC–to–N ratio; cm = microbial biomass C.
Figure 2. Principal component analysis (PCA) obtained from the dataset at a 400–1000 m altitudinal range. The distinct colors indicate the three clusters distributed along the principal component (Dim) 1 and 2. The sample codes included the sampling area, profile number, layers, and vegetation cover. The variables are in black. BS = base saturation; Ca = exchangeable Ca; TN = total N; OC = soil organic C; CN = OC–to–N ratio; cm = microbial biomass C.
Land 12 00079 g002
Figure 3. Principal component analysis (PCA) obtained from the dataset at a 1000–2134 m altitudinal range. The distinct colors identified the three clusters distributed along principal components (Dim) 1 and 2. The sample codes included the sampling area, profile number, layer, and vegetation cover. The variables are in black. BS = base saturation; Ca = exchangeable Ca; N = total N; OC = soil organic C; CN = OC–to–N ratio; cm = microbial biomass C.
Figure 3. Principal component analysis (PCA) obtained from the dataset at a 1000–2134 m altitudinal range. The distinct colors identified the three clusters distributed along principal components (Dim) 1 and 2. The sample codes included the sampling area, profile number, layer, and vegetation cover. The variables are in black. BS = base saturation; Ca = exchangeable Ca; N = total N; OC = soil organic C; CN = OC–to–N ratio; cm = microbial biomass C.
Land 12 00079 g003
Table 1. Altitudinal range, soil temperature regime, and soil moisture regime of the phytoclimatic zones and subzones of the northern Apennine chain, Italy, according to De Philippis [73,74] and Soil Survey Staff [52] and the location of soil profiles within the altitudinal ranges. The capital letters refer to soil sampling areas, as shown in Figure 1.
Table 1. Altitudinal range, soil temperature regime, and soil moisture regime of the phytoclimatic zones and subzones of the northern Apennine chain, Italy, according to De Philippis [73,74] and Soil Survey Staff [52] and the location of soil profiles within the altitudinal ranges. The capital letters refer to soil sampling areas, as shown in Figure 1.
Phytoclimatic aAltitude Range aSoil Temperature Regime bSoil Moisture Regime bSoil
Profile
Soil Profile
Altitude range
ZoneSubzone(m a.s.l.)Number
Alpinetum >2000CryicPerudic3Land 12 00079 i001
Picetumcold1800–2000FrigidPerudic6
warm1500–1800FrigidUdic43
Fagetumcold1300–1500FrigidUdic16
warm1000–1300MesicUdic11
Castanetumcold600–1000MesicUdic13
warm400–600MesicUdic57
a De Philippis [73,74]; b Soil Survey Staff [52].
Table 2. Mean ± standard deviation of pH; base saturation (BS) and C:N ratio values; and concentrations of total organic carbon (TOC), total nitrogen (TN), exchangeable Ca (Caexch), sand, silt, and clay at a 0–30 cm soil depth, as well as identified soil types according to Soil Survey Staff [52] of chestnut groves (CHE), mixed forests (MF), beech forests (BEF), blueberry (BLB), coniferous forests (CON), and grasslands (GS) located within 400–1000 m and 1000–2134 m altitudinal ranges of the northeastern Apennine chain (Italy). Within each altitudinal range, different letters indicate significant differences (Kruskal–Wallis test, p < 0.05) among the means.
Table 2. Mean ± standard deviation of pH; base saturation (BS) and C:N ratio values; and concentrations of total organic carbon (TOC), total nitrogen (TN), exchangeable Ca (Caexch), sand, silt, and clay at a 0–30 cm soil depth, as well as identified soil types according to Soil Survey Staff [52] of chestnut groves (CHE), mixed forests (MF), beech forests (BEF), blueberry (BLB), coniferous forests (CON), and grasslands (GS) located within 400–1000 m and 1000–2134 m altitudinal ranges of the northeastern Apennine chain (Italy). Within each altitudinal range, different letters indicate significant differences (Kruskal–Wallis test, p < 0.05) among the means.
400–1000 m Altitudinal Range1000–2134 m Altitudinal Range
CHEMFBEFBLBCONGS
pH5.1 ± 0.7 b5.8 ± 1.2 a4.6 ± 0.5 b4.5 ± 0.3 b4.8 ± 0.6 b5.4 ± 0.8 a
TOC (g kg−1)28.9 ± 28.130.7 ± 28.847.4 ± 30.0 a52.1 ± 24.1 a45.8 ± 24.4 a45.8 ± 21.8 a
TN (g kg−1)1.7 ± 1.51.8 ± 1.42.8 ± 1.7 b3.5 ± 1.7 a3.2 ± 1.3 a3.9 ± 1.6 a
Caexch (cmol(+) kg−1)5.1 ± 7.6 b9.1 ± 6.0 a5.7 ± 6.2 a2.3 ± 2.7 b7.8 ± 7.1 a6.3 ± 5.2 a
BS (%)31.3 ± 27.3 b59.7 ± 29.4 a37.3 ± 27.9 b21.9 ± 16.2 c46.4 ± 32.3 b61.6 ± 33.5 a
C:N ratio16.2 ± 6.317.1 ± 5.617.2 5.3 a15.4 ± 4.8 b14.1 ± 4.8 b11.7 ± 1.8 c
Sand (%)49.0 ± 15.0 b57.0 ± 15.0 a61.0 ± 13.8 ab61.0 ± 15.0 ab53.0 ± 20.8 b63.0 ± 16.2 a
Silt (%)40.0 ± 15.0 a27.0 ± 12.0 b30.0 ± 11.1 a31.0 ± 11.9 a36.0 ± 19.5 a28.0 ± 12.4 a
Clay (%)10.0 ± 4.9 b16.0 ± 8.4 a10.0 ± 6.9 ab8.0 ± 5.4 b11.0 ± 5.7 a9.0 ± 6.8 ab
Soil Taxonomy [52]OrdersSubgroups Subgroups
EntisolsLithic Udorthents
Typic Udorthents
Lithic Udorthents
Typic Udorthents
Lithic Udorthents
Typic Udorthents
Typic UdorthentsLithic Udorthents
Typic Udorthents
Lithic Cryorthents
Typic Udorthents
Mollisols Lithic Hapludolls
InceptisolsLithic Dystrudepts
Oxyaquic Dystrudepts
Typic Dystrudepts
Lithic Eutrudepts
Typic Eutrudepts
Lithic Dystrudepts
Oxyaquic Dystrudepts
Typic Dystrudepts
Humic Lithic Dystrudepts
Humic Lithic Eutrudepts
Humic Dystrudepts
Oxyaquic Dystrudepts
Typic Dystrudepts
(Spodic) Lithic Dystrudepts
Lithic Dystrudepts
Humic Lithic Dystrudepts
Humic Dystrudepts
Typic Dystrudepts
Lithic Dystrudepts
Lithic Eutrudepts
Humic Lithic Dystrudepts
Humic Lithic Eutrudepts
Typic Dystrudepts
Typic Eutrudepts
Lithic Dystrudepts
Spodosols Lithic HaplohumodsTypic Haplorthods Lithic Haplorthods
Table 3. Mean ± standard deviation of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, base saturation (BS), and pH values, and texture class of soils identified at 400–1000 m above sea level and classified according to Soil Survey Staff [52]. Round brackets identify the number of soil profiles for each soil type. Within each column, different letters indicate significant differences among the values (Kruskal–Wallis test, p < 0.05).
Table 3. Mean ± standard deviation of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, base saturation (BS), and pH values, and texture class of soils identified at 400–1000 m above sea level and classified according to Soil Survey Staff [52]. Round brackets identify the number of soil profiles for each soil type. Within each column, different letters indicate significant differences among the values (Kruskal–Wallis test, p < 0.05).
Soil OrderSoil SubgroupSandSiltClayTextureTOCTNC/NCaexchBSpH
%%% g kg−1g kg−1 cmol(+) kg−1%
EntisolLithic Udorthents (5)65 ± 10 a21 ± 8 c14 ± 7 abSL31.1 ± 27.72.0 ± 1.716.2 ± 6.0 bc9.4 ± 9.4 ab65.7 ± 22.8 a5.7 ± 1.0 ab
Typic Udorthents (8)51 ± 16 cd40 ± 17 b10 ± 4 bL18.6 ± 18.51.2 ± 1.215.1 ± 6.3 ab11.3 ± 10.4 ab45.8 ± 27.8 b5.7 ± 1.3 ab
InceptisolLithic Dystrudepts (9)36 ± 18 e50 ± 19 a14 ± 6 abSiL27.6 ± 28.62.1 ± 1.912.4 ± 2.8 c5.4 ± 3.8 bc31.3 ± 17.6 bc5.1 ± 0.4 bc
Lithic Eutrudepts (5)57 ± 14 bc28 ± 10 c15 ± 11 abSL29.9 ± 26.91.9 ± 1.515.5 ± 4.2 ab18.8 ± 12.4 a77.4 ± 20.4 a6.7 ± 1.2 a
Oxyaquic Dystrudepts (4)52 ± 13 cd29 ± 10 c19 ± 7 aL35.6 ± 40.12.1 ± 2.115.4 ± 3.8 ab5.1 ± 6.6 bc39.9 ± 28.9 bc5.1 ± 0.6 c
Typic Dystrudepts (39)52 ± 13 cd39 ± 11 b9 ± 4 bL30.9 ± 28.41.6 ± 1.418.1 ± 7.0 a3.3 ± 4.4 c26.6 ± 21.5 c4.9 ± 0.5 c
Typic Eutrudepts (2)52 ± 7 ce27 ± 4 c21 ± 6 aSCL29.9 ± 30.91.7 ± 1.615.8 ± 3.6 ab9.2 ± 6.1 ab72.1 ± 21.4 ab6.3 ± 1.2 ab
SL is sandy loam; L is loam; SiL is silty loam; SCL is sandy clay loam.
Table 4. Mean ± standard deviation of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, base saturation (BS), and pH values, and texture class soils identified within the 1000–2134 m above sea level altitudinal range and classified according to Soil Survey Staff [52]. Round brackets indicate the number of soil profiles identified for each soil type. Within each column, different letters indicate significant differences among the values (Kruskal–Wallis test, p < 0.05).
Table 4. Mean ± standard deviation of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, base saturation (BS), and pH values, and texture class soils identified within the 1000–2134 m above sea level altitudinal range and classified according to Soil Survey Staff [52]. Round brackets indicate the number of soil profiles identified for each soil type. Within each column, different letters indicate significant differences among the values (Kruskal–Wallis test, p < 0.05).
Soil OrderSoil SubgroupSandSiltClayTextureTOCTNC/NCaexchBSpH
%%% g kg−1g kg−1 cmol(+) kg−1%
EntisolLithic Cryorthent (3)54 ± 18 d32 ± 11 ad14 ± 9 abSL41.5 ± 14.03.6 ± 1.111.5 ± 2.1 d8.2 ± 6.0 ab71.5 ± 23.2 a5.6 ± 1.0 ab
Lithic Udorthent (6)58 ± 13 d27 ± 11 bc14 ± 7 aSL40.9 ± 27.12.6 ± 1.616.8 ± 7.1 ab6.4 ± 5.6 bc51.7 ± 20.1 bc4.7 ± 0.4 cd
Typic Udorthent (7)56 ± 23 d36 ± 21 ab8 ± 4 bdSL54.2 ± 21.23.9 ± 1.514.1 ± 4.1 bd5.2 ± 6.2 cd46.5 ± 31.4 bc5.0 ± 0.5 bc
Inceptisol(Spodic) Lithic Dystrudepts (2)77 ± 6 a19 ± 4 d4 ± 3 dLS59.3 ± 31.73.6 ± 2.516.9 ± 3.9 ab2.6 ± 2.3 d37.6 ± 18.8 ce4.9 ± 0.3 bc
Humic Dystrudepts (4)52 ± 13 d38 ± 11 a10 ± 5 abL42.3 ± 27.02.5 ± 1.316.7 ± 7.0 ab1.5 ± 0.8 d17.2 ± 6.2 fg4.6 ± 0.3 cd
Humic Lithic Dystrudepts (9)60 ± 15 d31 ± 11 ac9 ± 7 bdSL47.5 ± 22.13.3 ± 1.414.3 ± 2.1 ac2.0 ± 2.7 d25.8 ± 18.8 eg4.6 ± 0.3 cd
Humic Lithic Eutrudepts (4)70 ± 13 ac23 ± 10 cd7 ± 4 cdSL34.3 ± 22.52.5 ± 1.313.2 ± 4.7 cd10.3 ± 8.9 ab69.1 ± 30.6 ab5.4 ± 0.7 ab
Lithic Dystrudepts (5)73 ± 8 ab20 ± 8 d6 ± 4 dSL39.5 ± 20.53.2 ± 1.812.7 ± 2.1 cd3.5 ± 4.7 d39.5 ± 35.3 df5.0 ± 0.5 b
Lithic Eutrudepts (3)42 ± 24 d45 ± 26 ac14 ± 6 aL47.9 ± 23.23.4 ± 1.414.7 ± 7.2 bd12.6 ± 5.2 a73.9 ± 16.9 a5.3 ± 0.7 ab
Typic Dystrudepts (19)59 ± 13 d32 ± 10 ab8 ± 5 bdSL52.1 ± 29.33.1 ± 1.816.8 ± 4.1 ab6.3 ± 6.4 bc24.3 ± 20.8 eg4.4 ± 0.3 e
SpodosolTypic Haplorthod (4)62 ± 11 cd28 ± 9 ac10 ± 4 acSL57.1 ± 36.53.8 ± 2.515.4 ± 5.2 ac2.4 ± 1.7 d19.1 ± 16.1 g4.5 ± 0.4 de
Lithic Haplohumod (2)58 ± 15 d37 ± 10 ab4 ± 3 dSL64.1 ± 31.23.5 ± 1.817.7 ± 5.7 a3.6 ± 3.0 cd39.9 ± 20.6 ce4.4 ± 0.3 de
MollisolLithic Hapludoll (3)57 ± 25 bd36 ± 20 ac6 ± 6 dSL46.4 ± 22.33.8 ± 1.211.6 ± 2.3 cd9.0 ± 4.9 ab93.5 ± 6.3 a6.2 ± 0.2 a
Table 5. Mean of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, and BS, TOC–to–TN ratio (C:N), and pH values of the clusters obtained via cluster analysis on the set of data below 1000 m above sea level. Within each column, values sharing similar letters do not indicate significant differences (Kruskal–Wallis test, p < 0.05).
Table 5. Mean of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, and BS, TOC–to–TN ratio (C:N), and pH values of the clusters obtained via cluster analysis on the set of data below 1000 m above sea level. Within each column, values sharing similar letters do not indicate significant differences (Kruskal–Wallis test, p < 0.05).
ClusterSandSiltClayTOCTNCaexchBSC:NpH
%%%g kg−1 Cmol (+) kg−1%
1 (45)54 a31 b15 a58.3 a3.7 a13.6 a66.4 a16.35.9 a
2 (145)55 a35 a10 b19.1 c1.2 c3.4 b29.8 b15.55.0 b
3 (23)25 b65 c10 b38.9 b1.3 b8.4 b31.5 b21.75.3 b
Table 6. Mean of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, and BS, TOC–to–TN ratio (C:N), and pH values of the clusters obtained via cluster analysis on the set of data for the 1000 m above sea level altitudinal range. Within each column, values sharing similar letters do not indicate significant differences (Kruskal–Wallis test, p < 0.05).
Table 6. Mean of sand, silt, clay, total organic carbon (TOC), total nitrogen (TN), and exchangeable Ca (Caexch) contents, and BS, TOC–to–TN ratio (C:N), and pH values of the clusters obtained via cluster analysis on the set of data for the 1000 m above sea level altitudinal range. Within each column, values sharing similar letters do not indicate significant differences (Kruskal–Wallis test, p < 0.05).
ClusterSandSiltClayTOCTNCaexchBSC:NpH
%%%g kg−1 cmol(+) kg−1%
I (33)60 a34 a6 b44.2 b3.4 b13.2 a88.3 a12.75.7 a
II (130)57 b32 ab11 a35.6 c2.3 c3.0 c26.6 c16.04.7 b
III (54)65 a28 b7 b81.4 a5.4 a6.1 b37.0 b15.44.3 c
Table 7. Mean ± standard error of soil organic carbon stock (SOC stock) and total nitrogen stock (TN stock) in chestnut groves (CHE), coniferous forests (CON), beech forests (BEF), blueberry habitat (BLB), mixed forests (MF), and grasslands (GS) and at 400–1000 and 1000–2134 m above sea level (a.s.l.). For the vegetation types and altitudinal ranges, different letters within each column indicated significant differences (Kruskal–Wallis test, p < 0.05).
Table 7. Mean ± standard error of soil organic carbon stock (SOC stock) and total nitrogen stock (TN stock) in chestnut groves (CHE), coniferous forests (CON), beech forests (BEF), blueberry habitat (BLB), mixed forests (MF), and grasslands (GS) and at 400–1000 and 1000–2134 m above sea level (a.s.l.). For the vegetation types and altitudinal ranges, different letters within each column indicated significant differences (Kruskal–Wallis test, p < 0.05).
SOC StockTN Stock
Mg ha−1
VegetationBEF118.4 ± 30.0 a7.2 ± 2.5 b
typeBLB141.1 ± 42.2 a9.6 ± 2.2 a
CHE73.3 ± 36.1 b4.6 ± 1.7 c
CON117.9 ± 24.7 a9.3 ± 2.5 a
GS128.7 ± 46.5 a11.4 ± 3.9 a
MF70.8 ± 25.9 b4.7 ± 1.9 c
Altitudinal≤100072.7 ± 33.9 b4.6 ± 1.7 b
range (m a.s.l.)>1000125.3 ± 38.5 a8.8 ± 3.0 a
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

Vittori Antisari, L.; Trenti, W.; Buscaroli, A.; Falsone, G.; Vianello, G.; De Feudis, M. Pedodiversity and Organic Matter Stock of Soils Developed on Sandstone Formations in the Northern Apennines (Italy). Land 2023, 12, 79. https://doi.org/10.3390/land12010079

AMA Style

Vittori Antisari L, Trenti W, Buscaroli A, Falsone G, Vianello G, De Feudis M. Pedodiversity and Organic Matter Stock of Soils Developed on Sandstone Formations in the Northern Apennines (Italy). Land. 2023; 12(1):79. https://doi.org/10.3390/land12010079

Chicago/Turabian Style

Vittori Antisari, Livia, William Trenti, Alessandro Buscaroli, Gloria Falsone, Gilmo Vianello, and Mauro De Feudis. 2023. "Pedodiversity and Organic Matter Stock of Soils Developed on Sandstone Formations in the Northern Apennines (Italy)" Land 12, no. 1: 79. https://doi.org/10.3390/land12010079

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