Next Article in Journal
The Slowdown in China’s Energy Consumption Growth in the “New Normal” Stage: From Both National and Regional Perspectives
Next Article in Special Issue
Climate Change and Silvopasture: The Potential of the Tree and Weather to Modify Soil Carbon Balance
Previous Article in Journal
Analyzing Safety Concerns of (e-) Bikes and Cycling Behaviors at Intersections in Urban Area
Previous Article in Special Issue
Livestock Management for the Delivery of Ecosystem Services in Fire-Prone Shrublands of Atlantic Iberia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Responses in Soil Carbon and Nitrogen Fractionation after Prescribed Burning in the Montseny Biosphere Reserve (NE Iberian Peninsula)

by
Sangita Chowdhury
1,*,
José Manjón-Cabeza
1,
Mercedes Ibáñez
1,2,
Christian Mestre
1,
Maria José Broncano
3,
María Rosa Mosquera-Losada
4,
Josefina Plaixats
3 and
M.-Teresa Sebastià
1,2
1
Group GAMES, Department HBJ, School of Agrifood and Forestry Science and Engineering (ETSEA), University of Lleida (UdL), 25198 Lleida, Spain
2
Laboratory of Functional Ecology and Global Change (ECOFUN), Forest Science and Technology Centre of Catalonia (CTFC), 25280 Solsona, Spain
3
Grup de Recerca en Remugants, Departament de Ciència Animal i dels Aliments, Universitat Autònoma de Barcelona, 08193 Bellaterra, Spain
4
Department of Crop Production and Engineering Projects, High Polytechnic School, University of Santiago de Compostela, 15782 Lugo, Spain
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(7), 4232; https://doi.org/10.3390/su14074232
Submission received: 18 November 2021 / Revised: 28 March 2022 / Accepted: 29 March 2022 / Published: 2 April 2022

Abstract

:
Prescribed fire is one of the most widely-used management tools to recover encroached rangelands. Fire has been reported to cause changes in the soil physical and chemical properties. However, the legacy effects of former plant species on soil responses to fire remains unknown. The legacy effect of the former extant plant species on soil carbon (C) and nitrogen (N) fractionation distribution after prescribed burning in topsoil (0–5 cm and 5–10 cm) was investigated in Mediterranean shrublands in Montseny. We sampled soils under five vegetation patch types: Cytisus scoparius L., Calluna vulgaris L., Erica arborea L., Pteridium aquilinum L., and Cladonia biocrusts, pre- and post-burning. Multivariate analysis on soil C and N fractions showed that soils under the legume Cytisus and the biocrust were the most differentiated. Vegetation patch types tended to respond differently to burning, soils under Cytisus, Cladonia and Calluna showing the strongest response. Total C and N, and C and N in sand decreased after burning in the 0–5 cm soil layer. Conversely, C in silt, as well as N in clay and silt, increased with soil depth after burning. This study will be helpful for understanding ecological legacy effects and their possible consequences when planning prescribed burning.

1. Introduction

Fire is an important driver of environmental changes in an ecosystem, responsible for altering nutrient pools by changing the physical, chemical and biological properties of soils and nutrient cycling [1,2]. Prescribed burning is considered as the deliberate application of fire characterized by lower temperature, intensity and severity compared to wildfires [3,4]. This practice is a widely-used to reduce the risk of wildfire and recover shrublands [5].
Soils are considered as a major reservoir of C in terrestrial ecosystems [6]. Soil includes different carbon (C) and nitrogen (N) pools where soil labile C and N pools are characterized by their small size and fast turnover rates, with large size and slow turnover rates, recalcitrant organic C and N pools are physically or chemically protected [7]. Soil organic carbon (SOC) is responsible for soil fertility, and it is used as an indicator of soil health. Decreased soil C/N ratio led to a decline in SOC and an increase in soil N [8,9]. The influence of prescribed fires on soil C and N depends on fire frequency and intensity, in addition to climatic factors [3,10]. Furthermore, prescribed burning is responsible to affect only the upper centimetres of the soil [4]. Prescribed fires may also affect microbial community composition and function that lead to altered C and N cycling in an ecosystem [10,11]. Some studies have described increased upper soil C and N content after low intensity fires because of the incorporation of unburned or partially unburned slash fragments into the soil [12,13,14], while other studies report no change [15]. Soil organic matter decreased with short-term low intensity fire in topsoil in Mediterranean Shrubland [16,17]. [18] observed decreases in soil nutrient concentrations with greater fire intensity on topsoil in a mixed Chaparral in USA. Soil carbon and nitrogen were lost from the first 5 cm with high intensity and long-term burning in a shrubland of Eastern Spain [19].
Soil particle size fractions are helpful in maintaining the stability of ecosystems [20].
Soil organic matter within the sand fraction is allocated to the active (labile) pool, and that in the silt and clay fractions, to the passive (recalcitrant) pool. The labile pool is easily affected by fluctuation in environmental conditions as well as decomposed rapidly and becomes oxidized easily with changes in land use [21]. The passive/non-labile pool is more stable and recalcitrant, and therefore this fraction is decomposed slowly by microbial activity [22].
Furthermore, plants determine the quantity and the quality of residues, soil organic matter, as well as soil structure [23]. Thus, soil functions are also affected by plant functional diversity [24]. Plant functional types (PFTs) have proved to be a useful tool for predicting soil processes including C and N cycles [25]. Legumes have the potential to modify soil nutrient availability as legumes have the capacity to fix symbiotic N [26]. In addition, biological soil crusts (BSCs), which are assemblages of lichens, fungi, cyanobacteria, and mosses that colonize the soil surface, play a key role in the N cycle, because N-fixing lichens and free-living heterotrophic bacteria forming part of BSCs are able to fix substantial amounts of atmospheric N [27].
Many studies have focused on the dynamics of topsoil C and N stocks after prescribed burning. There are several studies conducted on fire effects on seed germination, natural regeneration, and changes in vegetation composition. However, very few studies have focused on the effects of fire on soil C and N fractionation, that is, C and N contents in clay, sand, and silt fractions [16,28]. Furthermore, to our knowledge, this is the first study that examines the legacy effects of the prior plant species on soil carbon and nitrogen particle size fractions after prescribed burning. Ecological legacy is a concept focused mainly on community or ecosystem-level phenomena, related to memories of the ecosystem to past events [29]. To understand present biodiversity patterns and predict the future ecological impacts of ongoing human practices on ecosystem services and functions, it is important to understand the carry-overs, or legacies, of the past events [30]. Therefore, the aim of this study was to assess the impact of prescribed burning on the C and N contents in the different fractions in topsoil (0–5 cm and 5–10 cm) in Pla de la Llacuna, Montseny, particularly to examine the legacy effect of the former extant plant species on soil carbon and nitrogen fractions after prescribed burning. We hypothesise that due to low to moderate prescribed fire, (1) the C and N contents in different soil fractions will be increased because of accumulation of burnt material; (2) patches dominated by different plant functional types will show variability in the soil C and N contents, with soil under legume containing higher C and N contents compared with other species; and (3) burning effects on soil C and N distribution will be modified by prior plant species, that is, there will be ecological legacy effects of former extant species.

2. Materials and Methods

2.1. Study Area

The study was conducted at the Pla de la Llacuna (longitude 2°18′ to 2°22′ east, latitude 41°44′ to 41°47′ north), located in the Pla de la Calma, an elevated plateau that ranges between 1000 and 1350 m a. s. l. This occupies an irregular area of about 974 hectares in the Montseny Natural Park, in the Northern Catalan Pre-Littoral Range, north-eastern Iberian Peninsula (Figure 1). This park comprises Mediterranean and Central European landscapes including different biomes, having local influence of metropolitan conurbations nearby.
The plateau is characterized by humid Mediterranean climate; mean annual precipitation approximately ranges between 700 and 1000 mm, where snow accounts for around 10% of the annual precipitation. The mean annual temperature is 11 °C according to the Tagamanent meteorological station (https://www.meteo.cat accessed on 17 November 2021). Bedrock is a metamorphic schist where the major minerals are quartz albite, muscovite and chlorite. Soils are acidic, with a pH of 4.5 to 5.5, and characterised by a sandy-loam texture [31]. The topographic location and humidity circulation lead to a general situation that favors the presence of Atlantic vegetation, where the shrubs Erica scorparia L., Erica arborea L. and Calluna vulgaris L. are widespread. The area is covered principally by shrubs and grass [32], including biological crusts dominated by various species of the lichen Cladonia sp pl., usually accompanied by some grasses. For centuries, the hills of the Montseny mountains have been used as pastures. Montseny was declared as a Natural Park in 1977; since then many traditional practices have declined, and some, such as shepherds’ burning, have completely ceased. Many small flocks composed mainly of sheep and goats grazed in Montseny in former times. The lack of direct grazing has caused a change in the vegetation structure and open grasslands are now covered by shrubs [33].

2.2. Field Sampling and Laboratory Determinations

The area selected for the prescribed burning experiment had a surface of 1.7 ha (Figure 1). A smaller area of 30 × 90 m within this surface was selected for the soil sampling (Figure 1). The burning event was conducted in 28 February 2019. Two samplings were performed in the study area, one before the burning event (pre-burning) which was carried out in 30 January 2019; and a re-sampling afterwards (post-burning) in 5 March 2019. The fire intensity was low to moderate in our study area. There was a small rain after burning in February 21 (0.1 mm).
Five vegetation patch types were initially identified as the most abundant vegetation patches in the shrubland. Each type was dominated by different species: the vascular plants Calluna vulgaris L., Erica arborea L., Cytisus scoparius L. and Pteridium aquilinum L.; and the biocrust dominated by lichens mostly of the genus Cladonia. These vegetation patches belong to different plant functional types, where the shrubs C. vulgaris and E. arborea are Ericaceae and C. scoparius is a legume; P. aquilinum is a fern; and the Cladonia patch is a lichen-dominated biocrust, often mixed with grasses. Henceforth, we will use the terminology: Calluna (CV) for Calluna vulgaris patches; Erica (EA) for Erica arborea patches; Cytisus (CS) for Cytisus scoparius patches; Pteridium (PA) for Pteridium aquilinum; and Cladonia (CSP) for the biocrusts with Cladonia.
The sampling was conducted using a stratified directed sampling. Soil samples were extracted in two different soil layers (0–5 cm and 5–10 cm) using a 4 × 4 cm2 coring probe. The treatments used for stratification (burning, vegetation patch and soil layer) were replicated six times across the sampling area, resulting in 120 samples. Sampling points were georeferenced throughout the sampling area with a highly precise GNSS Leica Zeno 20 (Leica Geosystems AG, Heerbrugg, Switzerland) with a differential correction Real Time Kinematic (RTK) broadcast system that was connected to the RTKAT service.
In addition, we measured the slope at each soil sampling point using the 5 m DEM (Digital Elevation Model) from the IGN (www.ign.es accessed on 17 November 2021). Slope was calculated in QGIS based on the DEM provided by the Institut Cartogràfic Geogràfic Català (ICGC, 2021), and a slope value provided for each sampling point. Slope was included in the modelling as a way of controlling the possible effects of spatial microtopographical heterogeneity.
Afterwards soils were transported to the laboratory and oven dried at 60 °C until constant weight. Soil samples were physically fractionated using the method developed by Six et al. [34]. The three soil fraction samples of known moisture content were analysed by a LECO C.N.H.S. Elemental Analyzer for the percentage of C and N. Total percentage of soil C and N was also measured in the two different soil depth layers. This results in eight variables including: total C, C in clay, C in silt, C in sand, as well as total N, N in clay, N in silt, N in sand. The C/N ratio was also calculated for each of the fractions described above.

2.3. Data Analysis

Multivariate indirect ordination analysis was applied to the ensemble of the soil variables analysed in this study. In particular, we applied Principal Component Analysis (PCA). Multivariate analysis was conducted with CANOCO 5.1 [35].
In addition to multivariate analysis, we performed univariate statistical regression on each study variable, using linear mixed effect models, with the identity of the sampling point as random factor, and with burning, vegetation patch type, soil layers and slope as fixed factors. We included the interactions among those factors, except for slope, which was used as a covariate. Then the best model was selected according to AIC (Akaike Information Criterion) and stepwise method (both forward and backward). We did the normality test of all variables by Shapiro-Wilk Normality Test. All the variables followed the normal distribution except C/N ratio in the clay fraction. In this case, we used a generalized linear mixed model with inverse gaussian distribution. Then we performed Tukey’s post hoc analysis at the p < 0.05 significance level. All statistical analyses were conducted in the R version 4.0.2 [36].

3. Results

In the study area, silt was the fraction including the highest soil C and N contents (Table 1 and Table 2). We found that most of the study variables increased with slope as well as soil depth (Table 1, Table 2 and Table 3).

3.1. Overall Soil C and N Distribution

Considering together all the C and N variables, including C and N content in the total soil and in the three soil fractions, the Principal Component Analysis (PCA) explained 65% of the total variability. PCA axis 1 explained 88% of the explained variability and mainly separated samples according to the plant species under which the soil was originally extracted (Figure 2A). Soils under the legume shrub Cytisus were distributed along the most positive part on PCA1, followed by those under Pteridium (Figure 2A); while soils under Cladonia distributed mostly on the negative side (Figure 2A). Soils under the two Ericaceae had low responses to this axis, suggesting intermediate soil C and N trends (Figure 2A). In addition, PCA axis 2 added 6% to the explained variability. This axis mostly explained the variability due to burning (Figure 2A). Soils under Cytisus and Cladonia followed by Calluna were the ones showing a higher response to burning compared to the soil under the other species (Figure 2A). Finally, PCA axis 3 explained 3% of the total explained variability. Axes PCA2 and PCA3 suggest differences among vegetation patches in overall soil C and N responses to burning, soils under Cytisus, Calluna and Cladonia being the ones most differentiated between unburned and burned conditions (Figure 2A,B).

3.2. Total and Fractional Soil C

Both total C (Figure 3A; Table 1) and C in the sand fraction in the 0–5 cm soil layer (Figure 3D; Table 1) decreased after burning; conversely, C in the silt fraction (Figure 3C; Table 1) increased with prescribed burning in the 5–10 cm layer, according to the regression model. However, this was not captured by the Tukey test, suggesting that the increase of C in the silt fraction with burning is relatively weak.
The highest C content was found in the soils under Cytisus patches, while the lowest C content was found under Cladonia patches in the sand fraction (Figure 4D; Table 1). No significant interactions were found between plant species and burning in the total carbon, neither in the three soil C fractions (Table 1). However, we could detect tendencies in the particular response of some plant species to burning (Figure S1 in Supplementary Materials). After burning, total soil C in the 0–5 cm soil layer showed a decreasing trend in all species, which was more pronounced in Cladonia and Erica patches compared to the other patch types (Figure S1A in Supplementary Materials). C in sand also decreased in all the species, but especially in the Cytisus and Calluna patches in the 0–5 cm soil layer (Figure S1D in Supplementary Materials). In contrast, soil C increased after burning more remarkably in Cytisus and Calluna in the silt fraction in the 5–10 cm layer (Figure S1G in Supplementary Materials).
Soil C content in the 5–10 cm layer was less than half than that in the 0–5 cm layer under Cytisus, Erica and Pteridium. While soil C content in the 5–10 cm layer was half than that in the 0–5 cm layer under Cladonia and Calluna (Figure 5; Table 1).

3.3. Total and Fractional soil N

Total soil N (Figure 3E; Table 2), and N in the sand fraction in the 0–5 cm soil layer (Figure 3H; Table 2) decreased after burning. Conversely, soil N in both the clay (Figure 3F; Table 2) and the silt fractions (Figure 3G; Table 2) significantly increased with prescribed burning in the 5–10 cm layer.
Significantly highest and lowest N content were found in sand fraction under Cytisus and Cladonia patches, respectively, compared to the other vegetation patches (Figure 4H; Table 2). There were no significant interactions between plant species and burning in the total soil N, neither in the three soil N fractions (Table 2). However, there were some tendencies for vegetation patch types to respond differently to burning (Figure S2 in Supplementary Materials). Total soil N decreased in almost all vegetation patch types, but more remarkably in Cladonia in the 0–5 cm layer after burning (Figure S2A in Supplementary Materials). N in sand decreased in all species, but mainly in Calluna, followed by Cytisus and Erica in the 0–5 cm soil layer after burning (Figure S2D in Supplementary Materials). On the other hand, N in the clay fraction increased especially in Cytisus, but also slightly in Cladonia and Calluna after burning in the 5–10 cm layer (Figure S2F in Supplementary Materials). This increasing trend was also found for the silt fraction, where N increased after burning more remarkably in Cytisus and Calluna in the 5–10 cm layer (Figure S2G in Supplementary Materials).
Table 2. Final models for all N variables in the study. p-values from the mixed model regressions on N variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Table 2. Final models for all N variables in the study. p-values from the mixed model regressions on N variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Explanatory Variables Nitrogen Variables
Total Soil NitrogenNitrogen in ClayNitrogen in SiltNitrogen in Sand
Slope0.001 **0.003 **0.003 **<0.001 ***
Burning0.028 *0.047 *0.002 **0.108
Species0.0750.0880.1020.013 *
Soil depth<0.001 ***<0.001 ***<0.001 ***<0.001 ***
Burning * Soil depth--0.022 *0.082
p < 0.1; p < 0.05 *; p < 0.01 **; p < 0.001 ***.

3.4. Total and Fractional Soil C/N Ratio

The soil C/N ratio decreased with soil depth (Figure 3I). In addition, there was a tendency for total soil C/N to decrease after burning, particularly from the soil under Cladonia biocrusts, followed by Erica. patches (significant burning x species interaction; Figure 6; Table 3).
Prescribed fire significantly decreased the C/N ratio of total soil (Figure 3I) and sand fraction (Figure 3L) in the 0–5 cm layer, as well as in clay (Figure 3J) and silt fractions (Figure 3K) (Table 3).
Table 3. Final models for all C/N variables in the study. p-values from the mixed model regressions on C/N variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Table 3. Final models for all C/N variables in the study. p-values from the mixed model regressions on C/N variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Explanatory Variables C/N Ratio
Total C/N RatioC/N Ratio in ClayC/N Ratio in SiltC/N Ratio in Sand
Slope-0.031 *-0.104
Burning0.002 **0.024 *<0.001 ***0.415
Species0.821---
Soil depth<0.001 ***0.014<0.001 ***<0.001 ***
Burning * Soil depth0.051--0.033 *
Burning * Species0.023 *---
p < 0.1; p < 0.05 *; p < 0.01 **; p < 0.001 ***.

4. Discussion

Our results show an important effect of burning and species composition on total and fractional soil C and N dynamics (Figure 2 and Figure 3) (Figures S1 and S2 in Supplementary Materials) which was unknown in previous literature. Part of the variability in the soil C and N contents can be attributed to the specific vegetation patch, with soils under the legume species performing higher C and N contents compared with other species. Furthermore, after burning, we found a significant decrease in soil C and N content in the coarser soil fractions in the 0–5 cm soil layer but an increase in the finer fractions in the 5–10 cm layer.
We found that most of the study variables increased with slope (Table 1, Table 2 and Table 3) maybe due to the spatial microtopographical heterogeneity in the distribution of soil parameters [37].

4.1. Total and Fractional Soil C and N Distribution before and after Burning

Total soil C (Figure 3A; Table 1) and N (Figure 3E; Table 2), and C (Figure 3D; Table 1) and N (Figure 3H; Table 2) in the sand fraction in the 0–5 cm soil layer decreased after burning. The soil C and N can be substantially decreased with low to moderate fire intensity [2], which may explain the remarkable decrease of soil C and N observed at 0–5 cm depth in Pla de la Llacuna. The decrease in total C in low severity prescribed burning might be attributed to C loss as CO2 into the atmosphere, while the decrease of total N might be attributed to N loss as volatilization of N [38], as well as direct convective transfer of ash [39]. Ref. [17] also found a significant decrease in total soil organic C (SOC) and in total N content in the uppermost soil layer immediately after prescribed burning. In our study, the reduction of soil C and N during the fire affected mainly the sand fraction, maybe because combustion can be more intense in this size range due to the oxygen present in macropores [40]. Our findings contrast the results obtained in a study by [13], where SOC at 0–5 cm depth increased immediately after low-intensity prescribed burning in Mediterranean grassland in the northeastern Iberian Peninsula.
In contrast, soil N in the clay fraction (Figure 3F; Table 2) as well as C (Figure 3C; Table 1) and N (Figure 3G; Table 2) in the silt fraction significantly increased with prescribed burning in the 5–10 cm layer. This suggests the redistribution of C and N in soil fractions after prescribed burning by promoting C and N enrichment in finer fractions in the study area. The increase of N in the clay as well as C and N in the silt fraction may be due to the downwards translocation and accumulation of C and N at 5–10 cm layer compared to the 0–5 cm layer [41].
These results partly support the first hypothesis that the C and N contents in different soil fractions will be increased due to the accumulation of burnt material. However, we have found that coarser fractions generally losing C and N after fire; while finer fractions tend to increase their C and N content (Figure 3; Table 1 and Table 2).

4.2. Plant Species and Species Legacy Effects on Total and Fractional Soil C and N Distribution after Burning

PCA1 differentiated vegetation patches according to initial soil C and N conditions (Figure 2A), while PCA2 (Figure 2A), and PCA3 (Figure 2B) showed that patches separated differently on those axes according to burning. This suggests that species had very strong effects on soils before burning but left imprint on soil fractional C and N after burning in dissimilar intensity, suggesting that species legacy effects were uneven after burning. This is shown by the observed tendency of vegetation patch types to respond differently to burning (Figures S1 and S2 in Supplementary Materials). Furthermore, the effects of species were not as strong as the effects of other treatment factors (Table 1, Table 2 and Table 3).
The legume shrub Cytisus showed the highest differentiation when considering overall composition of soil C and N parameters compared to other species, both before and after burning (PCA1; Figure 2A). Legumes are capable to fix atmospheric N and allocate it to the plant in exchange for carbohydrates. Therefore, legumes can strongly enhance the input of N into the soil ecosystem [42]. Due to their effectiveness in transferring aminoacids between nodules and roots, legumes favour organic N sources compared to other plant functional types (PFT) [43]. In addition, legumes possess higher leaf nitrogen content and higher specific leaf area compared with other plant functional types, traits related with increased photosynthetic rates, which increase net CO2 uptake [44,45]. Thus, legumes enhance plant productivity, which in turn lead to increased C sequestration in soil [42]. Those legumes’ traits result in higher litter quality and litter decomposition rates than non-legumes due to symbiotic relationships [46]; which in turn can result in higher soil C and N. The study by [47] agreed with our results, showing how soil C and N pools were enhanced by the presence of two legume species in an experiment carried out over 2 years. However, the enhancement effect on soil C has been reported to disappear at high legume proportions [48].
Cladonia patches also showed important differences in soil C and N distribution compared to the other patch types (PCA1; Figure 2A). A study conducted in Siberian forests by [49] found that soil C storage was lower in lichen patches, likely due to lower rates of C fixation [50], or higher rates of decomposition of lichen patches [23]. Ref. [51] found that NO3 was lower in the site dominated by lichens than those composed by other plant functional types.
Soils under Cytisus and Cladonia, followed by Calluna, were the soils showing a higher response to burning compared to the soil under the other species (PCA2 and PCA3; Figure 2A,B). Burned residues of N-rich legume can stimulate N mineralization and nitrification after low intensity fires [52]. Legumes have also been reported to respond positively to fire in other ecosystems, as for instance tallgrass prairies and pine forests of the southeastern US [53,54]. On the other hand, maybe because lichens have lower accumulation rates of organic matter content and mineral N, predominantly ammonium, compared with grasses and mosses, [55] found that these soil characteristics were also maintained after fire. According to [56], a large proportion of the nutrients can be mobilized with such fire intensity from the soil under Calluna after burning as smoke, where smoke is defined as including the gases and volatile products of combustion together with suspended solid particles.
These results are in the agreement with our second hypothesis, which states that different species-dominated patches will have different soil C and N total and fractional contents, and legumes will contain higher C and N proportions than other species. Furthermore, they also agree with our third hypothesis that burning effects on soil C and N distribution will be modified by former plant species. However, the legacy effects of prior vegetation patches were not equally intense in all patches, neither they were necessarily linked to prior soil differences among plants (Figure 2)

4.3. Total and Fractional Soil C/N Ratio

The decreasing trend of the C/N ratio after burning found in this study agrees with what has been traditionally reported in the literature, generally attributed to preferential immobilization of N over C after prescribed burning [1,57]. However, to our knowledge, no study has previously reported the dependency that this parameter showed on prior vegetation patch, according to our results (Figure 6). The total soil C/N ratio was the soil parameter showing the strongest legacy effect compared to other studied soil C and N variables, strongly supporting our third hypothesis.

5. Conclusions

We have found that soil C and N contents are dependent on soil fractions, with the sand fraction being more vulnerable to lose C and N than clay and silt. In particular, C and N in total soil and in the sand fraction in the 0–5 cm layer decreased after short-term prescribed burning but increased in silt and clay fractions in the 5–10 cm layer, which is likely due to downwards translocation and accumulation of C and N from coarse fraction in 0–5 cm soil to fine fractions in 5–10 cm soil layer. Hence, it is a recommendation for further study to assess the translocation process of C and N in the different fractions in this ecosystem. The decreasing trend of the C/N ratio in this study suggests that soil C in the study site could be more labile than N. There were differences both in the way different species distributed C and N among fractions and total C and N, and in how different patches responded to burning. That means the composition of species matters. The species legacy effects in the soil C and N responses to burning were best revealed when analysing jointly all the C and N variables in the study by multivariate analysis, with total soil C/N being the single variable most influenced by ecological legacies. The legume shrub Cytisus showed the highest differential overall composition of soil C and N parameters compared to other species as legumes possess more C and N than other species. So, legumes can help in maintaining soil fertility in the site [58], as well as provide support in the harsh conditions and compensate nutrient loss [59]. Soils under Cytisus and Calluna were the most responsive to burning compared to other species. Therefore, the effect of short-term prescribed burning is dependent on the species composition of the ecosystem, and it is important in ecological as well as in management aspects.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/su14074232/s1, Figure S1: Mean ± 1 SE total and fractional (top to bottom) soil C distribution in the 0–5 cm and 5–10 cm soil layers per vegetation patch type (Cytisus scoparius (CS); Cladonia (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA) and burning treatment, Figure S2: Mean ± 1 SE total and fractional (top to bottom) soil N distribution in the 0–5 cm and 5–10 cm soil layers per vegetation patch type (Cytisus scoparius (CS); Cladonia (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA) and burning treatment, Figure S3: Mean ± 1 SE total and fractional (top to bottom) soil C/N ratio distribution in the 0–5 cm and 5–10 cm soil layers per vegetation patch type (Cytisus scoparius (CS); Cladonia (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA) and burning treatment.

Author Contributions

Conceptualization, M.-T.S., J.P., M.J.B. and M.I.; methodology, M.-T.S., J.P., M.J.B. and M.I.; formal analysis, S.C., J.M.-C. and M.I.; investigation, M.-T.S., S.C., J.M.-C. and M.I.; resources, M.-T.S. and J.P.; data curation, S.C.; writing—original draft preparation, S.C., M.-T.S., M.I. and J.M.-C.; writing—review and editing, M.-T.S., M.I., J.P., C.M., M.J.B., M.R.M.-L., J.M.-C. and S.C.; supervision, M.-T.S., M.I. and J.M.-C.; project administration, M.-T.S. and J.P.; funding acquisition, M.-T.S. and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This study was developed through projects OPEN2PRESERVE (SOE2/P5E0804), from the EU SUDOE; and IMAGINE (CGL2017-85490-R), from the Spanish Science Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available upon request to MTS under given circumstances.

Acknowledgments

We would like to thank Antonio Rodríguez and David Estany for helping in field soil sampling. Divina Váquez, from the University of Santiago de Compostela, and Alex Escolà, from the University of Lleida, are acknowledged for helping with the sample processing in the laboratory and facilitating field precision instrumentation, respectively. We acknowledge the coordinator of the OPEN2PRESERV project Rosa Maria Canals for her overall support in this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. González-Pérez, J.A.; González-Vila, F.J.; Almendros, G.; Knicker, H. The effect of fire on soil organic matter—A review. Environ. Int. 2004, 30, 855–870. [Google Scholar] [CrossRef] [PubMed]
  2. Certini, G. Effects of fire on properties of forest soils: A review. Oecologia 2005, 143, 1–10. [Google Scholar] [CrossRef] [PubMed]
  3. Fernandes, P.M.; Davies, G.M.; Ascoli, D.; Fernández, C.; Moreira, F.; Rigolot, E.; Stoof, C.R.; Vega, J.A.; Molina, D. Prescribed burning in southern Europe: Developing fire management in a dynamic landscape. Front. Ecol. Environ. 2013, 11. [Google Scholar] [CrossRef] [Green Version]
  4. San Emeterio, L.; Múgica, L.; Gutiérrez, R.; Juaristi, A.; Pedro, J.; Canals Tresserras, R. Cambios en el nitrógeno edáfico tras la realización de quemas controladas para mejora de pastos pirenaicos. Pastos Rev. la Soc. Española para el Estud. los Pastos 2013, 43, 44–53. [Google Scholar]
  5. Reverchon, F.; Xu, Z.; Blumfield, T.J.; Chen, C.; Abdullah, K.M. Impact of global climate change and fire on the occurrence and function of understorey legumes in forest ecosystems. J. Soils Sediments 2012, 12, 150–160. [Google Scholar] [CrossRef] [Green Version]
  6. Ramachandran Nair, P.K.; Nair, V.D.; Mohan Kumar, B.; Showalter, J.M. Carbon sequestration in agroforestry systems. Adv. Agron. 2010, 108, 237–307. [Google Scholar] [CrossRef]
  7. McLauchlan, K.K.; Hobbie, S.E. Comparison of Labile Soil Organic Matter Fractionation Techniques. Soil Sci. Soc. Am. J. 2004, 68, 1616–1625. [Google Scholar] [CrossRef]
  8. Snider, M.J.; Reinhardt, L.; Wolfenden, R.; Cleland, W.W. 15N kinetic isotope effects on uncatalyzed and enzymatic deamination of cytidine. Biochemistry 2002, 41, 415–421. [Google Scholar] [CrossRef]
  9. Dijkstra, P.; Menyailo, O.V.; Doucett, R.R.; Hart, S.C.; Schwartz, E.; Hungate, B.A. C and N availability affects the 15N natural abundance of the soil microbial biomass across a cattle manure gradient. Eur. J. Soil Sci. 2006, 57, 468–475. [Google Scholar] [CrossRef]
  10. Artz, R.R.E.; Reid, E.; Anderson, I.C.; Campbell, C.D.; Cairney, J.W.G. Long term repeated prescribed burning increases evenness in the basidiomycete laccase gene pool in forest soils. FEMS Microbiol. Ecol. 2009, 67, 397–410. [Google Scholar] [CrossRef]
  11. Davies, G.M.; Domènech, R.; Gray, A.; Johnson, P.C.D. Vegetation structure and fire weather influence variation in burn severity and fuel consumption during peatland wildfires. Biogeosciences 2016, 13, 389–398. [Google Scholar] [CrossRef] [Green Version]
  12. Soto, B.; Diaz-Fierros, F. Interactions between plant ash leachates and soil. Int. J. Wildl. Fire 1993, 3, 207–216. [Google Scholar] [CrossRef]
  13. Úbeda, X.; Lorca, M.; Outeiro, L.R.; Bernia, S.; Castellnou, M. Effects of prescribed fire on soil quality in Mediterranean grassland (Prades Mountains, north-east Spain). Int. J. Wildl. Fire 2005, 14, 379–384. [Google Scholar] [CrossRef]
  14. Roaldson, L.M.; Johnson, D.W.; Miller, W.W.; Murphy, J.D.; Walker, R.F.; Stein, C.M.; Glass, D.W.; Blank, R.R. Prescribed Fire and Timber Harvesting Effects on Soil Carbon and Nitrogen in a Pine Forest. Soil Sci. Soc. Am. J. 2014, 78, S48–S57. [Google Scholar] [CrossRef]
  15. Boyer, W.D.; Miller, J.H. Effect of burning and brush treatments on nutrient and soil physical properties in young longleaf pine stands. For. Ecol. Manag. 1994, 70, 311–318. [Google Scholar] [CrossRef]
  16. Girona-García, A.; Ortiz-Perpiñá, O.; Badía-Villas, D.; Martí-Dalmau, C. Effects of prescribed burning on soil organic C, aggregate stability and water repellency in a subalpine shrubland: Variations among sieve fractions and depths. Catena 2018, 166, 68–77. [Google Scholar] [CrossRef] [Green Version]
  17. Armas-Herrera, C.M.; Martí, C.; Badía, D.; Ortiz-Perpiñá, O.; Girona-García, A.; Porta, J. Immediate effects of prescribed burning in the Central Pyrenees on the amount and stability of topsoil organic matter. Catena 2016, 147, 238–244. [Google Scholar] [CrossRef]
  18. Marion, G.M.; Moreno, J.M.; Oechel, W.C. Fire Severity, Ash Deposition, and Clipping Effects on Soil Nutrients in Chaparral. Soil Sci. Soc. Am. J. 1991, 55, 235–240. [Google Scholar] [CrossRef]
  19. Rovira, P.; Romanyà, J.; Duguy, B. Long-term effects of wildfires on the biochemical quality of soil organic matter: A study on Mediterranean shrublands. Geoderma 2012, 179–180, 9–19. [Google Scholar] [CrossRef]
  20. Six, J.; Bossuyt, H.; Degryze, S.; Denef, K. A history of research on the link between (micro)aggregates, soil biota, and soil organic matter dynamics. Soil Tillage Res. 2004, 79, 7–31. [Google Scholar] [CrossRef]
  21. Haynes, R.J. Labile Organic Matter Fractions as Central Components of the Quality of Agricultural Soils: An Overview. Adv. Agron. 2005, 85, 221–268. [Google Scholar] [CrossRef]
  22. Wiesenberg, G.L.B.; Dorodnikov, M.; Kuzyakov, Y. Source determination of lipids in bulk soil and soil density fractions after four years of wheat cropping. Geoderma 2010, 156, 267–277. [Google Scholar] [CrossRef]
  23. Wardle, D.A.; Nilsson, M.C.; Zackrisson, O.; Gallet, C. Determinants of litter mixing effects in a Swedish boreal forest. Soil Biol. Biochem. 2003, 35, 827–835. [Google Scholar] [CrossRef]
  24. Debouk, H.; Emeterio, L.S.; Marí, T.; Canals, R.M.; Sebastià, M.T. Plant functional diversity, climate and grazer type regulate soil activity in natural grasslands. Agronomy 2020, 10, 1291. [Google Scholar] [CrossRef]
  25. De Kauwe, M.G.; Medlyn, B.E.; Zaehle, S.; Walker, A.P.; Dietze, M.C.; Wang, Y.P.; Luo, Y.; Jain, A.K.; El-Masri, B.; Hickler, T.; et al. Where does the carbon go? A model-data intercomparison of vegetation carbon allocation and turnover processes at two temperate forest free-air CO2 enrichment sites. New Phytol. 2014, 203, 883–899. [Google Scholar] [CrossRef] [Green Version]
  26. Craine, J.M.; Tilman, D.; Wedin, D.; Reich, P.; Tjoelker, M.; Knops, J. Functional traits, productivity and effects on nitrogen cycling of 33 grassland species. Funct. Ecol. 2002, 16, 563–574. [Google Scholar] [CrossRef] [Green Version]
  27. Evans, R.D.; Ehleringer, J.R. A break in the nitrogen cycle in aridlands? Evidence from δp15N of soils. Oecologia 1993, 94, 314–317. [Google Scholar] [CrossRef]
  28. Granged, A.J.P.; Jordán, A.; Zavala, L.M.; Muñoz-Rojas, M.; Mataix-Solera, J. Short-term effects of experimental fire for a soil under eucalyptus forest (SE Australia). Geoderma 2011, 167–168, 125–134. [Google Scholar] [CrossRef]
  29. Crow, T.R. Ecosystems: Balancing Science with Management. Restor. Ecol. 2000, 8, 99–101. [Google Scholar] [CrossRef]
  30. Vellend, M.; Brown, C.D.; Kharouba, H.M.; Mccune, J.L.; Myers-Smith, I.H. Historical ecology: Using unconventional data sources to test for effects of global environmental change. Am. J. Bot. 2013, 100, 1294–1305. [Google Scholar] [CrossRef] [Green Version]
  31. Belillas, C.M.; Rodà, F. Nutrient budgets in a dry heathland watershed in northeastern Spain. Biogeochemistry 1991, 13, 137–157. [Google Scholar] [CrossRef]
  32. Pié Valls, G.; Vilar Sais, L. Corologia de la flora vascular d’interès de conservació al Parc Natural del Montseny. Butlletí la Inst. Catalana d’Història Nat. 2014, 3987, 65–74. [Google Scholar] [CrossRef]
  33. Bartolomé, J.; Plaixats, J.; Fanlo, R.; Boada, M. Conservation of isolated Atlantic heathlands in the Mediterranean region: Effects of land-use changes in the Montseny biosphere reserve (Spain). Biol. Conserv. 2005, 122, 81–88. [Google Scholar] [CrossRef]
  34. Six, J.; Conant, R.T.; Paul, E.A.; Paustian, K. Stabilization mechanisms of SOM implications for C saturation of soils.pdf. Plant Soil 2002, 241, 155–176. [Google Scholar] [CrossRef]
  35. Ter Braak, C.J.F.; Smilauer, P. Canoco (Version 5.10): Canoco Reference Manual and User’s Guide: Software for Ordination; Biometris, Wageningen University & Research: Wageningen, The Netherlands, 2018. [Google Scholar]
  36. R Core Team. R: A Language and Environment for Statistical Computing; R Found. Stat. Comput.: Vienna, Austria, 2020. [Google Scholar]
  37. Sebastiá, M.T. Role of topography and soils in grassland structuring at the landscape and community scales. Basic Appl. Ecol. 2004, 5, 331–346. [Google Scholar] [CrossRef]
  38. Handbook of chemistry and physics: 1st student edition. Trends Biochem. Sci. 1988, 13, 116. [CrossRef]
  39. Smith, D.W. Concentrations of soil nutrients before and after fire. Can. J. Soil Sci. 1970, 50, 17–29. [Google Scholar] [CrossRef]
  40. Jordán, A.; Zavala, L.M.; Mataix-Solera, J.; Nava, A.L.; Alanís, N. Effect of fire severity on water repellency and aggregate stability on Mexican volcanic soils. Catena 2011, 84, 136–147. [Google Scholar] [CrossRef]
  41. Leal, O.A.; Dick, D.P.; Costa, F.S.; Knicker, H.; de Carvalho Júnior, J.A.; Santos, J.C. Carbon in physical fractions and organic matter chemical composition of an acrisol after Amazon forest burning and conversion into pasture. J. Braz. Chem. Soc. 2019, 30, 413–424. [Google Scholar] [CrossRef]
  42. Fornara, D.A.; Banin, L.; Crawley, M.J. Multi-nutrient vs. nitrogen-only effects on carbon sequestration in grassland soils. Glob. Change Biol. 2013, 19, 3848–3857. [Google Scholar] [CrossRef] [Green Version]
  43. Moran-Zuloaga, D.; Dippold, M.; Glaser, B.; Kuzyakov, Y. Organic nitrogen uptake by plants: Reevaluation by position-specific labeling of amino acids: Reevaluation of organic N uptake by plants by position-specific labeling. Biogeochemistry 2015, 125, 359–374. [Google Scholar] [CrossRef]
  44. Reich, P.B.; Ellsworth, D.S.; Walters, M.B. Leaf structure (specific leaf area) modulates photosynthesis-nitrogen relations: Evidence from within and across species and functional groups. Funct. Ecol. 1998, 12, 948–958. [Google Scholar] [CrossRef]
  45. Ibanez, M.; Altimir, N.; Ribas, A.; Eugster, W.; Sebastia, M.T. Phenology and plant functional type dominance drive CO2 exchange in seminatural grasslands in the Pyrenees. J. Agric. Sci. 2020, 158, 3–14. [Google Scholar] [CrossRef] [Green Version]
  46. Fornara, D.A.; Tilman, D. Plant functional composition influences rates of soil carbon and nitrogen accumulation. J. Ecol. 2008, 96, 314–322. [Google Scholar] [CrossRef]
  47. De Deyn, G.B.; Quirk, H.; Yi, Z.; Oakley, S.; Ostle, N.J.; Bardgett, R.D. Vegetation composition promotes carbon and nitrogen storage in model grassland communities of contrasting soil fertility. J. Ecol. 2009, 97, 864–875. [Google Scholar] [CrossRef]
  48. Rodríguez, A.; Canals, R.M.; Sebastià, M.T. Positive Effects of Legumes on Soil Organic Carbon Stocks Disappear at High Legume Proportions Across Natural Grasslands in the Pyrenees. Ecosystems 2021. Available online: https://link.springer.com/article/10.1007/s10021-021-00695-9 (accessed on 17 November 2021). [CrossRef]
  49. Webb, E.E.; Heard, K.; Natali, S.M.; Bunn, A.G.; Alexander, H.D.; Berner, L.T.; Kholodov, A.; Loranty, M.M.; Schade, J.D.; Spektor, V.; et al. Variability in above- and belowground carbon stocks in a Siberian larch watershed. Biogeosciences 2017, 14, 4279–4294. [Google Scholar] [CrossRef] [Green Version]
  50. Turetsky, M.R.; Mack, M.C.; Hollingsworth, T.N.; Harden, J.W. The role of mosses in ecosystem succession and function in Alaska’s boreal forest. Can. J. For. Res. 2010, 40, 1237–1264. [Google Scholar] [CrossRef] [Green Version]
  51. Castillo-Monroy, A.P.; Maestre, F.T.; Delgado-Baquerizo, M.; Gallardo, A. Biological soil crusts modulate nitrogen availability in semi-arid ecosystems: Insights from a Mediterranean grassland. Plant Soil 2010, 333, 21–34. [Google Scholar] [CrossRef]
  52. Madritch, M.D.; Cardinale, B.J. Impacts of tree species diversity on litter decomposition in northern temperate forests of Wisconsin, USA: A multi-site experiment along a latitudinal gradient. Plant Soil 2007, 292, 147–159. [Google Scholar] [CrossRef]
  53. Hendricks, J.J.; Boring, L.R. N2-fixation by native herbaceous legumes in burned pine ecosystems of the southeastern United States. For. Ecol. Manag. 1999, 113, 167–177. [Google Scholar] [CrossRef]
  54. Newland, J.A.; DeLuca, T.H. Influence of fire on native nitrogen-fixing plants, and soil nitrogen status in ponderosa pine—Douglas-fir forests in western Montana. Can. J. For. Res. 2000, 30, 274–282. [Google Scholar] [CrossRef]
  55. Sedia, E.G.; Ehrenfeld, J.G. Differential effects of lichens, mosses and grasses on respiration and nitrogen mineralization in soils of the New Jersey Pinelands. Oecologia 2005, 144, 137–147. [Google Scholar] [CrossRef] [PubMed]
  56. Evans, C.C.; Allen, S.E. Nutrient Losses in Smoke Produced during Heather Burning. Oikos 1971, 22, 149. [Google Scholar] [CrossRef]
  57. Certini, G.; Nocentini, C.; Knicker, H.; Arfaioli, P.; Rumpel, C. Wildfire effects on soil organic matter quantity and quality in two fire-prone Mediterranean pine forests. Geoderma 2011, 167–168, 148–155. [Google Scholar] [CrossRef]
  58. Barthès, B.; Barthès, B.; Azontonde, A.; Blanchart, E.; Girardin, C.; Villenave, C.; Lesaint, S.; Oliver, R.; Feller, C. Effect of a legume cover crop (Mucuna pruriens var. utilis) on soil carbon in an Ultisol under maize cultivation in southern Benin. Soil Use Manag. 2004, 20, 231–239. [Google Scholar] [CrossRef]
  59. Chaer, G.M.; Resende, A.S.; Campello, E.F.C.; De Faria, S.M.; Boddey, R.M.; Schmidt, S. Nitrogen-fixing legume tree species for the reclamation of severely degraded lands in Brazil. Tree Physiol. 2011, 31, 139–149. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) Study area located in the NE of the Iberian Peninsula, (B,C) in the Natural Park of Montseny (Catalonia). (D) RGB orthomosaics representing the pre-burned and burned area of the Pla de la Llacuna experimental area (courtesy of the OPEN2PRESERVE project, https://open2preserve.eu/ accessed on 17 November 2021).
Figure 1. (A) Study area located in the NE of the Iberian Peninsula, (B,C) in the Natural Park of Montseny (Catalonia). (D) RGB orthomosaics representing the pre-burned and burned area of the Pla de la Llacuna experimental area (courtesy of the OPEN2PRESERVE project, https://open2preserve.eu/ accessed on 17 November 2021).
Sustainability 14 04232 g001
Figure 2. Samples distribution along the two first axes (PCA1 and PCA2 (A); and the first and third axes, PCA1 and PCA3 (B) of Principal Component Analysis (PCA) performed on the overall soil C and N variables (total and fractional) in 0–5 and 5–10 soil layers. Samples are clumped into groups according to vegetation patch type and prescribed burning treatment. The mean value ± 1 standard error of each axis variable is represented for each species and treatment by different symbols and whiskers.
Figure 2. Samples distribution along the two first axes (PCA1 and PCA2 (A); and the first and third axes, PCA1 and PCA3 (B) of Principal Component Analysis (PCA) performed on the overall soil C and N variables (total and fractional) in 0–5 and 5–10 soil layers. Samples are clumped into groups according to vegetation patch type and prescribed burning treatment. The mean value ± 1 standard error of each axis variable is represented for each species and treatment by different symbols and whiskers.
Sustainability 14 04232 g002
Figure 3. Mean ± 1 SE total and fractional (top to bottom) soil C (AD), N (EH) and C/N ratio (IL) distribution in the 0–5 cm and 5–10 cm soil layers. Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (soil layers and burning). Sample size n = 120.
Figure 3. Mean ± 1 SE total and fractional (top to bottom) soil C (AD), N (EH) and C/N ratio (IL) distribution in the 0–5 cm and 5–10 cm soil layers. Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (soil layers and burning). Sample size n = 120.
Sustainability 14 04232 g003
Figure 4. Mean ± 1 SE total and fractional (top to bottom) soil C (AD) and N (EH) distribution per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA). Different letters indicate significant differences (p < 0.05) among treatments according to Tukey mean comparison tests. Sample size n = 120.
Figure 4. Mean ± 1 SE total and fractional (top to bottom) soil C (AD) and N (EH) distribution per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA). Different letters indicate significant differences (p < 0.05) among treatments according to Tukey mean comparison tests. Sample size n = 120.
Sustainability 14 04232 g004
Figure 5. Mean ± 1 SE soil C distribution in sand fraction in 0–5 cm and 5–10 cm soil layers per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA). Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (soil layers and species). Sample size n = 120.
Figure 5. Mean ± 1 SE soil C distribution in sand fraction in 0–5 cm and 5–10 cm soil layers per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA). Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (soil layers and species). Sample size n = 120.
Sustainability 14 04232 g005
Figure 6. Mean ± 1 SE soil C/N ratio distribution in total soil per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA) and burning treatment. Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (burning and species). Sample size n = 120.
Figure 6. Mean ± 1 SE soil C/N ratio distribution in total soil per vegetation patch type; Cytisus scoparius (CS); Cladonia biocrust (CSP); Calluna vulgaris (CV); Erica arborea (EA); Pteridium aquilinum (PA) and burning treatment. Different letters indicate significant differences (p < 0.05) among treatments according to multiple Tukey mean comparison tests (burning and species). Sample size n = 120.
Sustainability 14 04232 g006
Table 1. Final models for all C variables in the study. p-values from the mixed model regressions on C variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Table 1. Final models for all C variables in the study. p-values from the mixed model regressions on C variables, for the following tested explanatory variables: slope, burning, species, and soil depth.
Explanatory Variables Carbon Variables
Total Soil CarbonCarbon in ClayCarbon in SiltCarbon in Sand
Slope0.001 **0.001 **0.002 **<0.001 ***
Burning0.004 **0.1750.1910.085
Species0.0950.0610.0760.024 *
Soil depth<0.001 ***<0.001 ***<0.001 ***<0.001 ***
Burning * Soil depth0.021 *-0.039 *0.034 *
Species * Soil depth---0.076
p < 0.1; p < 0.05 *; p < 0.01 **; p < 0.001 ***.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chowdhury, S.; Manjón-Cabeza, J.; Ibáñez, M.; Mestre, C.; Broncano, M.J.; Mosquera-Losada, M.R.; Plaixats, J.; Sebastià, M.-T. Responses in Soil Carbon and Nitrogen Fractionation after Prescribed Burning in the Montseny Biosphere Reserve (NE Iberian Peninsula). Sustainability 2022, 14, 4232. https://doi.org/10.3390/su14074232

AMA Style

Chowdhury S, Manjón-Cabeza J, Ibáñez M, Mestre C, Broncano MJ, Mosquera-Losada MR, Plaixats J, Sebastià M-T. Responses in Soil Carbon and Nitrogen Fractionation after Prescribed Burning in the Montseny Biosphere Reserve (NE Iberian Peninsula). Sustainability. 2022; 14(7):4232. https://doi.org/10.3390/su14074232

Chicago/Turabian Style

Chowdhury, Sangita, José Manjón-Cabeza, Mercedes Ibáñez, Christian Mestre, Maria José Broncano, María Rosa Mosquera-Losada, Josefina Plaixats, and M.-Teresa Sebastià. 2022. "Responses in Soil Carbon and Nitrogen Fractionation after Prescribed Burning in the Montseny Biosphere Reserve (NE Iberian Peninsula)" Sustainability 14, no. 7: 4232. https://doi.org/10.3390/su14074232

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