Next Article in Journal
Anthropogenic Pressure and Climate Change Could Severely Hamper the Avian Metacommunity of the Sicilian Wetlands
Next Article in Special Issue
Microbial Communities of Ferromanganese Sedimentary Layers and Nodules of Lake Baikal (Bolshoy Ushkany Island)
Previous Article in Journal
Species Distribution Modeling Reveals Recent Shifts in Suitable Habitat for Six North American Cypripedium spp. (Orchidaceae)
Previous Article in Special Issue
Bacterial and Archaeal Water and Sediment Communities of Two Hot Spring Streams in Tengchong, Yunnan Province, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Warmer and Poorer: The Fate of Alpine Calcareous Grasslands in Central Apennines (Italy)

by
Marco Varricchione
1,
Maria Laura Carranza
1,*,
Valter Di Cecco
2,
Luciano Di Martino
2 and
Angela Stanisci
1
1
EnvixLab, Department of Biosciences and Territory, University of Molise, C. da Fonte Lappone. 86090, Pesche and Via Duca degli Abruzzi 67, 86039 Termoli, Italy
2
Maiella Seed Bank, Maiella National Park, Loc. Colle Madonna, 66010 Lama dei Peligni, Italy
*
Author to whom correspondence should be addressed.
Diversity 2022, 14(9), 695; https://doi.org/10.3390/d14090695
Submission received: 28 July 2022 / Revised: 19 August 2022 / Accepted: 19 August 2022 / Published: 23 August 2022
(This article belongs to the Special Issue Diversity in Extreme Environments)

Abstract

:
Global change threatens alpine biodiversity and its effects vary across habitat types and biogeographic regions. We explored vegetation changes over the last 20 years on two Mediterranean alpine calcareous grasslands in central Apennines (Italy): stripped grasslands (EUNIS code E4.436) with Sesleria juncifolia growing on steep slopes, and wind edge swards (EUNIS code E4.42) with Carex myosuroides. Based on a re-visitation of 25 vegetation plots of 4 × 4 m, we assessed changes in overall and endemic plant species cover and richness by nonparametric Kruskal–Wallis test. We explored changes in structure and ecology using growth forms and Landolt indicators for temperatures. We identified species’ contribution to temporal changes using the similarity percentage procedure (SIMPER). The results evidenced a significant decline in all species cover and richness on both plant communities with a significant decline in alpine and endemic species and in hemicryptophytes with rosette and scapose ones on stripped grasslands, as well as a decline in subalpine and suffruticose chamaephytes species on wind edge swards. Such biodiversity loss, so far observed only in the warmest and Southern Mediterranean summits of Europe, is likely attributable to the combined effect of higher temperatures; the increase in the vegetative period; and the decrease in water availability, which is particularly severe in calcareous regions. Our study suggested the vulnerability of the analyzed alpine ecosystems to global change and the importance of monitoring activities to better understand vegetation trends and adaptation strategies in subalpine, alpine, and nival ecosystems.

1. Introduction

Alpine ecosystems, distributed over the limit of forests and beyond to the snow line, are among the ecosystems most sensitive to and threatened by climate change (e.g., raising temperatures and change in precipitation patterns) [1,2,3]. Alpine plant communities are mainly shaped by temperatures and hydrologic features, which in turn derive from the interaction of winter precipitation, wind exposure, landscape position [4], snowpack accumulation [5], snowmelt patterns [6], and summer soil moisture [7], overall influencing the duration of the vegetative period [8].
The alpine life zone encompasses low stature vegetation above the climatic tree line worldwide and accounts for approximately 3% of the vegetated land area on Earth [9,10] and, due to their environmental and biogeographic peculiarities, they contain 4% of the Earth’s flora (8000–10,000 species of plants) [4]. Alpine ecosystems in Europe host 20% of the native plants, conforming a hotspot of biodiversity [11,12,13]. Moreover, the complex mosaic of mountain habitats and the presence of dispersal barriers have promoted the presence of highly specialized communities [14], particularly rich in endemic and rare vascular plants [15,16]. Mediterranean summits in Europe deserve particular attention because of the presence of a high amount of endemism [17,18]. Mediterranean alpine ecosystems, having been only partially glaciated during the ice age of Pleistocene [19] and remaining isolated for very long periods, constitute a major refuge of plant species, several of which are cryophilic endemic [20].
Climate change shapes alpine vegetation and plant diversity in different ways. For instance, the increase in temperatures recorded in the last decades [21,22] has consistently altered alpine and nival species’ distribution. In several European alpine plant communities, a general increase in species richness [23,24,25] was registered and it was explained as a “thermophilization process”, which consists of the upward shifting of thermophilic plants towards higher altitudes [23,26,27,28,29,30,31,32,33,34,35], or as a “range-filling process”, performed by species dispersing from existing neighbor communities within the same elevation belt [36,37].
Climate warming may likewise lead to a decline in cryophilic endemic taxa, which become particularly vulnerable because of their specialized habitat requirements, narrow distribution ranges, and low capacity to modify their geographic distribution [38]. Alpine endemics on south Europe, being distributed in small populations with low genetic diversity, are even more vulnerable to global warming [39,40,41].
In addition to the ongoing increase in temperatures, a reduction in winter rainfalls over the central and southern Mediterranean basin, including Italian Apennines, has been registered during the second half of the 20th century [42,43]. The combination of rising temperatures and decreasing precipitation is most likely related to a steep biodiversity decline documented by recent studies on some Mediterranean summits [23,44,45,46] and on some calcareous grasslands of the Alps [47].
Considering the consistent and heterogeneous changes ongoing on alpine ecosystems, likely related with both the increase in temperatures [21,22] and the reduction in annual rainfalls [1,42,48,49,50,51], the present work sets out to explore vegetation dynamics on alpine Mediterranean calcareous grasslands and swards in central Apennines (Italy). Through a re-visitation vegetation study (after 18 years), we explored the temporal changes in two alpine communities in the Maiella National Park (MNP) that are representative of the central Apennine’s alpine belt: Apennine stripped grasslands, growing on steep slopes, and wind edge swards, both included in the 6170 EU Habitat “alpine and subalpine calcareous grasslands” [52,53].
We addressed the following questions: (i) what happened during the last 18 years on Apennine stripped grasslands and wind edge swards in terms of species composition, structure (growth forms), and ecology (Landolt indicator values)? (ii) Do the endemic species cover and richness changed over time?

2. Materials and Methods

2.1. Study Area

The study area, located in central Italy, includes the higher sectors (from ~2400 up to 2793 m a.s.l.) of the Maiella National Park (Figure 1), characterized by a large limestone summit plateau covered by a thick (1–3 m) mantle of debris, modelled by periglacial processes in which tectonic-karst depressions are surrounded by steep slopes [54,55]. The most prominent soils on the plateau are those with a patterned ground surface of either micro-sorted circles or micro-sorted stripes [56]. The presence of the second highest peak in the Apennines (Mount Amaro, 2793 m a.s.l.) and the great extension of the alpine ecosystems above 2000 m of elevation (59 km2) [55] ensure that alpine vegetation is well expressed in terms of its typical flora [57,58].
For these reasons, it is considered an adequate site for investigating climate change effects on Mediterranean alpine vegetation [31,59,60].

2.2. Climate Features

From a bioclimatic point of view, the analyzed summits are included in the alpine biogeographical region [61] and the climate corresponds to the subalpine–alpine humid type. As in other alpine ecosystems, the climate in central Apennines has consistently changed during the last 60 years, with a significant increase in mean annual temperatures (+1.7 °C, amounting to 0.26 °C per decade) [35,62], spring and winter minimum temperatures, and summer mean temperatures [37]. At the same time, a decrease in annual precipitations (−30 mm) and an increase in the number of very dry years per decade were also recorded [62]. Steinbauer et al. [63] in a recent paper evidenced an increase in the in situ mean soil temperatures, for central Apennines summits, of 0.5 °C per decade (since 2001), values that are above the global average warming [64].
The analysis of local time series of soil temperatures registered by data-loggers in the Majella summits in the context of GLORIA monitoring network (Global Observation Research Initiative in Alpine Environments; “IT CAM FEM”: 2405 m a.s.l.) [65] evidenced a similar increase in temperatures. Specifically, we have registered a significant increase during the last two decades in the mean annual and mean autumn soil temperatures (from 2.85 °C to 3.73 °C and from 3.48 °C to 4.77 °C, respectively; Figure 2) and an increase in the vegetative period (e.g., the number of days with temperatures above 2 °C), which, during in the last twenty years, have raised from 176 to 200 days per year (Figure 2).

2.3. Plant Communities

We analyzed two alpine plant communities: (a) the Apennine stripped grasslands (EUNIS code E4.436) [66], represented by Sesleria juncifolia scraped and garland grasslands, and (b) the wind edge swards (EUNIS code E4.42) [66], characterized by meso-xerophile and relatively continuous Carex myosuroides grasslands. Both communities are quite widespread in central Apennine summit areas and are of conservation concern in Europe (Habitat 6170 EU: “alpine and subalpine calcareous grasslands”) [52,53]. Apennine stripped grasslands with Sesleria juncifolia grow on steep slopes with rendzina soils poor in organic matter and on prevalent E-SE aspects [67]. The wind edge swards with Carex myosuroides grow on more evolved calcareous cryoturbated soils rich in organic matter (brown rendzina) [67]. Both communities are characterized by the abundance of endemic, sub-endemic taxa, Mediterranean-montane species, and southern European orophytes [67].
Apennine stripped grasslands and wind edge swards grasslands in the study area have a very low anthropo-zoogenic pressure as they are in the core area of the Majella national park, where the livestock has been strictly regulated since more than 50 years ago [18]. The alpine zone of the park hosts the chamois (Rupicapra pyrenaica ornata), which, however, occurs preferentially on other environmental units as cliffs, screes, and snowbeds.

2.4. Vegetation Data

We re-visited, in the year 2021 (hereafter T2), 25 georeferenced phytosociological relevés previously sampled with a random-stratified sampling protocol in the year 2003 (hereafter T1). Re-visited vegetation plots are stored in VIOLA Database [60,68] and are distributed as follows: 12 on Apennine stripped grasslands and 13 on wind edge swards. Detailed local information of each revisited plot is reported in Appendix A.
Re-visitation of relevés were conducted in the same season (July–August) in order to remove the effects of phenological variability [69] and following the same sampling protocol (e.g., plot size; previous species lists; and plot headers’ information: geographic coordinates, slope, altitude, and aspect) [70]. For each georeferenced vegetation plot of 16 m2 (4 × 4 m), all plant species and their cover/abundance in compliance with the Braun–Blanquet scale [71,72] were registered. Vegetation plots were permanent and marked by metal pegs. The nomenclature follows the updated checklists of the vascular flora native to Italy [73]. Growth forms and chorotypes (e.g., endemic species) conform the “Flora d’Italia” [74]. We assigned to each taxa the species’ ecological indicator value for temperature (T) according to Landolt [75], which ranges from 1 to 5 according to the elevation distribution of the species, which is related to the average air temperature during the growing season. We adopted Landolt T values because of their demonstrated usefulness in depicting the response of vegetation to local topographic micro-thermal heterogeneity [74]. We classified species into three groups according to the average altitudinal distribution of the species sensu Landolt [73]: alpine (T: 1–1.5), subalpine (T: 2–2.5), and montane (T: 3–3.5). For the endemic Mediterranean flora, without Landolt indicators, values were assigned by the authors after checking the requirements of the species in the literature [58,76]. For further quantitative analysis, we rescaled the original Braun–Blanquet cover values for each species on the percentage cover scale [77].

2.5. Statistical Analysis

After a brief inspection of the species list on T1 and T2, we compared for both of the analyzed communities the number and cover of overall and endemic species over time by the nonparametric Kruskal–Wallis test for equal medians on ranked data (cover and richness).
Furthermore, we explored the changes in vegetation structure and ecology using the growth forms [78] and Landolt indicators values for temperatures.
We also identified the species that contribute most consistently to differences between the two temporal steps (T1 and T2) using a similarity percentage procedure (SIMPER) [79].
All analyses were performed in R statistical computing program [80] using the Vegan package [81] and PAST (paleontological statistics software for education and data analysis) [82].

3. Results

The recorded species list included 91 vascular plant species and subspecies mainly belonging to Asteraceae, Fabaceae and Poaceae families—72 on the Apennine stripped grasslands and 76 on the wind edge swards. A total of 23 of the total species were endemic (about 25%), with 21 occurring on stripped grasslands (29%) and 17 on wind edge swards (22%) (Table 1). Most of the endemic species were scapose hemicryptophytes (48%) and suffruticose chamaephytes (22%; Table 1).
The temporal comparison (T1 vs. T2) of species richness and cover per plot for the entire pool by Kruskal–Wallis test evidenced similar temporal trends on both of the studied plant communities. We registered a significant decrease in overall species cover and richness per plot on stripped grassland (Psame < 0.01 and Psame = 0.03, respectively) and on wind edge swards (Psame = 0.04 and Psame < 0.01, respectively) (Figure 3).
Similarly, we registered a significant decline in endemic species on stripped grassland (cover Psame < 0.001; richness Psame = 0.04) (Figure 3).
Concerning the average altitudinal distribution of the species, we detected on stripped grassland a significant decrease in alpine species per plot (cover Psame < 0.01 and richness Psame = 0.03) and, on wind edge swards, a decline in subalpine species (cover Psame = 0.02 and richness Psame < 0.01) (Figure 4).
We also observed some changes in vegetation structure, with stripped grasslands registering a significant decrease in cover for the hemicryptophytes, in particular for the hemicryptophytes with rosette (H ros) and scapose hemicryptophytes (H scap) (Psame < 0.01 and Psame = 0.04, respectively), and the wind edge swards showing a significant reduction in cover for the suffruticose chamaephytes (CH suffr) species over time (Psame < 0.01) (Figure 5).
According to the similarity percentage analysis (SIMPER) (Table 2), 8 of the 72 species for Apennine stripped grasslands and 7 of the 76 species for wind edge swards contributed to 50% of the observed temporal differences (T1 vs. T2) in vegetation composition. Two of these declining species are endemics (Armeria gracilis subsp. majellensis and Festuca violacea subsp. italica).
In the stripped grasslands, we observed the reduction in cover on six of these eight species and four of them are alpine (T = 1–1.5). Only two species, the montane Anthyllis vulneraria subsp. pulchella and the subalpine Carex kitaibeliana, increased in cover over time.
In the wind edge swards, the species that contributed most to the decreasing plant cover value over time were the subalpine (T = 2–2.5) Carex kitaibeliana and Helianthemum oelandicum subsp. alpestre. Only two species, the alpine Salix retusa and Carex myosuroides, increased their cover and the montane Anthyllis vulneraria subsp. pulchella remained stable over time.

4. Discussion

The re-visitation after 18 years of stripped grasslands and wind edge swards vegetation plots in central Apennines evidenced an important decline in plant diversity with a significant decrease in species richness and cover and the loss of some endemic species. It also evidenced significant alterations in plant communities’ structure and ecology with a reduction in alpine and hemicryptophytes in stripped grasslands with Sesleria juncifolia and in subalpine and suffruticose chamaephytes in wind edge swards with Carex myosuroides.
Similar decreasing trends in species richness and cover over time were observed in other Mediterranean and alpine summits in Europe. For instance, a decrease in alpine plants was observed in Mediterranean summits of Corsica [23], Sierra Nevada [44], and Crete [45], and those changes were attributed by the authors to biological effects of climate warming and the consequent decrease in water availability [44].
Similarly, Porro et al. [46] have observed a decline in plant diversity in northern Apennines and Nicklas et al. [47] described a slow and progressive loss of species richness in the calcareous grasslands of the Alps. Such biodiversity decline registered on some mountain summits is most likely a result of a combination of rising temperatures and decreasing summer precipitations occurring in southern Europe during the last decades [41,83,84,85]. Indeed, during the last five decades, a general increase in annual mean temperatures [86,87,88] with a decrease in annual precipitations and an extension of the dry summer season [43,89,90] across the Italian peninsula and Europe were recorded.
In this light, the increase in soil temperatures registered in the study area (by the GLORIA network data loggers), in particular in autumn, as well as the prolongation of the growing season (e.g., later accumulation of snow and earlier meltdown), should be related with the decline in species’ diversity registered in both stripped grasslands and wind edge swards. For instance, the warmer conditions registered in autumn could stimulate the premature germination of seeds [91] and the production of early seedlings that are not able to survive to the following harsh winter conditions (e.g., steep diurnal temperature range and the alternation of warm and very cold days) [92,93].
Notice that the decline in species richness registered during the last 18 years in the stripped grasslands is a result of the drop in alpine (e.g., Plantago atrata subsp. atrata and Poa alpina subsp. alpina) and endemic plants (e.g., Festuca violacea subsp. italica). Alpine species are cryophilous taxa, which are vulnerable to the increase in temperatures and the decrease in soil water availability [29,30,33,35,85]. Concerning the observed decline in endemic plants, similar results were reported in recent studies carried out in the southern Mediterranean summits of Europe such as Sierra Nevada [44] and Crete [45], and those changes were interpreted by the authors as a reaction of vegetation to climate change. We also observed in stripped grasslands a decline in hemicryptophytes with rosette (H ros) and scapose hemicryptophytes (H scap) (e.g., Plantago atrata subsp. atrata, Armeria gracilis subsp. majellensis, Ranunculus pollinensis, Scorzoneroides montana subsp. Breviscarpa and Viola eugeniae subsp. eugeniae), which are common in continuous grasslands [35,94] and are weakly adapted to cope with the new climatic conditions (e.g., higher temperatures, the prolongation of the growing season, and the consequent increased water stress) [63].
The decline in diversity values observed in the wind edge swards is due to the decrease in subalpine species (e.g., Carex kitaibeliana, Helianthemum oelandicum subsp. alpestre), which may be attributable to the harsher environmental conditions registered during the last decades [2]. Indeed, the prolongation of the growing season (e.g., later accumulation of snow and earlier meltdown) likely promoted in wind edge swards steeper diurnal temperature ranges and higher winter frost stress, limiting both the development and survival of several subalpine species [62,95]. On the other hand, the stability of the stress-tolerant and cold-adapted alpine species, in which wind edge swards in central Apennines are rich [96], is likely due to their ability to cope, at least so far, with such effects of climate change. The observed species cover changes in the wind edge swards vegetation structure are due to the decrease in suffruticose chamaephytes (CH suffr; e.g., Helianthemum oelandicum subsp. alpestre, Iberis saxatilis subsp. saxatilis, Noccaea stylosa). As observed by Steinbauer et al. [63] on alpine ridges across southern European summits, in our case too, climate warming seems to induce a decrease in cushion plants cover.
The observed decline in plant species richness and cover registered on central Apennines stripped grasslands and wind edge formations is most likely related to the rise in temperatures, the increased vegetative period, and the reduction in soil water content, which are impinging European summits and are particularly severe in calcareous regions [47]. The temperature rise registered in the study area in combination with a series of exceptionally dry summers (2003, 2015, 2021) [97], frequent summer storms with high runoff particularly detrimental on a karstic bedrock [98], and the convention winds promoting high evapotranspiration, likely all contribute to the observed change in plant diversity values.
Our study evidenced a particular process of diversity loss ongoing on alpine stripped and wind edge grasslands, which had not been observed so far on other plant communities in central Apennines. In fact, previous studies on the Mediterranean Apennines registered a stability of species richness and an increase in plant cover on Pinus mugo scrubs [37,76] on scree and snowbeds vegetation [35,94]. Such a temporal increases in plant cover and richness have been attributed to range filling processes (e.g., greening) and to the upward shift of montane/subalpine species towards the alpine and nival belts (e.g., thermophilization process), both promoted by climatic warming [33,76,99]. Southern mountains of the northern hemisphere (such as those in the Mediterranean area) host several species’ range-edge populations [41,47,100], which are more vulnerable to climatic changes than core populations [51,101,102,103,104].
The observed decline in plant species diversity is particularly worrying if accounting for the most updated climate change models that, for mountain areas in south Europe, have predicted increasing temperatures, decreasing precipitations, and a lengthening summer season [43,64,89,90]. Similarly, recent climatic projections for the next 50 years on Mediterranean mountains have predicted an increase in temperatures (+1.4 °C) and a reduction in precipitation (−4.8%) [42]. Altered climatic conditions that are shrinking the European mountain flora [23] are even more dangerous on Mediterranean areas, like the analyzed alpine stripped and wind edge grasslands, which, being rich in endemics [23,105,106,107], should have permanent damage on biodiversity.
These endemic alpine plant populations already placed at maximum local elevational gradients could face “mountaintop extinction” [104] and survive only in “warm-stage microrefugia”, geomorphologic niches that constantly maintain cold-air pooling and temperature inversions [108], with implications for the conservation of regional biodiversity.
Our results evidenced the urgent need for identifying “warm-stage microrefugia” in alpine protected areas able to support “climate-smart” in situ conservation planning for the Mediterranean mountains [106,109]. Such microrefugia could form a network of conservation sites of taxonomic and genetic diversity (e.g., alpine endemics), which should be strictly protected [108,110]. Those measures should be integrated with ex situ conservation practices, such as those locally implemented in the Majella Seed Bank [111], both complementary cost-effective actions [112].

5. Conclusions

The present re-visitation study evidenced significant changes over the last 18 years in the stripped grassland and wind edge swards plant communities in central Apennines.
Both plant communities showed a significant decrease in plant species cover and richness. As on Apennine stripped grassland, we observed a loss in alpine and endemic species, on wind edge swards, a reduction in subalpine species was registered. These changes are likely attributable to the combined effect of higher temperatures, the increase in the growing period (e.g., later accumulation of snow and earlier meltdown), and the decrease in soil water availability triggered by global change.
This study evidenced an ongoing biodiversity loss on calcareous alpine grasslands of mountaintops in central Apennines, which had so far been observed until now only in the warmest and Southern Mediterranean summits of Europe. Furthermore, in our study area, the stripped and garland calcareous grasslands (especially those dominated by Sesleria juncifolia) have been revealed to be the most sensitive to this effect.
The current re-visitation study contributed to increasing our knowledge about vegetation changes in central Mediterranean mountain top plant communities and underlined the need for monitoring activities on these sentinel ecosystems, which are increasingly threatened by global change.
Analogous monitoring research should be carried in other Mediterranean high mountains, where consistent effects of climate change are forecasted in order to better understand local vegetation trends and adaptation strategies in subalpine, alpine, and nival ecosystems.

Author Contributions

Conceptualization, M.V., A.S. and M.L.C.; methodology, M.V., A.S. and M.L.C.; formal analysis, M.V. and M.L.C.; data curation, M.V. and V.D.C.; writing—original draft preparation, M.V., A.S. and M.L.C.; writing—review and editing, M.V., V.D.C., A.S., L.D.M. and M.L.C.; supervision, A.S. and M.L.C. All authors have read and agreed to the published version of the manuscript.

Funding

The research was partially funded, as a subcontractor of the University of Vienna, by European Research Council Executive Agency Grant Agreement Number 883669, project MICROCLIM “A microscale perspective on alpine floras under climate change”.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The research took advantage of facilities of the LTER network, as it was carried out in LTER IT01-001-T Appennino centro-meridionale: Majella-Matese https://deims.org/c85fc568-df0c-4cbc-bd1e-02606a36c2bb (accessed on 1 February 2022). We are grateful to the reviewers and the editors for their valuable comments, which help us to improve the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Headers information of each re-sampled plot in both the studied plant communities.
Table A1. Headers information of each re-sampled plot in both the studied plant communities.
Plant CommunityRelevé NumberRelevé Area (m2)Altitude (m a.s.l.)AspectSlope (Degree)
Apennine stripped grasslands1162555E15
2162563E18
3162601NW15
4162483S13
5162558SW5
6162614S20
7162539S15
8162572S22
9162501S34
10162400N6
11162554NE7
12162567E7
Wind edge swards1162658 7
2162524NW10
3162529SW10
4162525NW7
5162479S5
6162481S7
7162481E4
8162610SW22
9162617E10
10162665SE16
11162597NE18
12162598E12
13162536SE20

References

  1. Chelli, S.; Wellstein, C.; Campetella, G.; Canullo, R.; Tonin, R.; Zerbe, S.; Gerdol, R. Climate change response of vegetation across climatic zones in Italy. Clim. Res. 2017, 71, 249–262. [Google Scholar] [CrossRef]
  2. Rogora, M.; Buzzi, F.; Dresti, C.; Leoni, B.; Lepori, F.; Mosello, R.; Patelli, M.; Salmaso, N. Climatic effects on vertical mixing and deep-water oxygen content in the subalpine lakes in Italy. Hydrobiologia 2018, 824, 33–50. [Google Scholar] [CrossRef]
  3. Verrall, B.; Pickering, C.M. Alpine vegetation in the context of climate change: A global review of past research and future directions. Sci. Total Environ. 2020, 748, 141344. [Google Scholar] [CrossRef]
  4. Körner, C. Alpine Plant Life, 2nd ed.; Springer: Berlin, Germany, 2003. [Google Scholar]
  5. Carlson, B.Z.; Choler, P.; Renaud, J.; Dedieu, J.P.; Thuiller, W. Modelling snow cover duration improves predictions of functional and taxonomic diversity for alpine plant communities. Ann. Bot. 2015, 116, 1023–1034. [Google Scholar] [CrossRef] [PubMed]
  6. Green, K.; Pickering, C. Vegetation, microclimate and soils associated with the latestlying snowpatches in Australia. Plant Ecol. Divers. 2009, 2, 289–300. [Google Scholar] [CrossRef]
  7. Fisk, M.C.; Schmidt, S.K.; Seastedt, T.R. Topographic patterns of above-and belowground production and nitrogen cycling in alpine tundra. Ecology 1998, 79, 2253–2266. [Google Scholar] [CrossRef]
  8. Hua, X.; Ohlemüller, R.; Sirguey, P. Differential effects of topography on the timing of the growing season in mountainous grassland ecosystems. Environ. Adv. 2022, 8, 100234. [Google Scholar] [CrossRef]
  9. Heywood, V.H. Global Biodiversity Assessment; Cambridge Press: Cambridge, UK, 1995. [Google Scholar]
  10. Körner, C. The grand challenges in functional plant ecology. Front. Plant Sci. 2011, 2, 1–3. [Google Scholar] [CrossRef]
  11. Médail, F.; Quézel, P. Biodiversity hotspots in the Mediterranean Basin: Setting global conservation priorities. Conserv. Biol. 1999, 13, 1510–1514. [Google Scholar] [CrossRef]
  12. Väre, H.; Lampinen, R.; Humphries, C.; Williams, P. Taxonomic diversity of vascular plants in the European alpine areas. In Alpine Biodiversity in Europe; Nagy, L., Grabherr, G., Körner, C., Thompson, D.B.A., Eds.; Springer: Berlin, Germany, 2003; pp. 133–148. [Google Scholar]
  13. Aeschimann, D.; Rasolofo, N.; Theurillat, J.-P. Analyse de la flore des Alpes. 5: Milieux et phytosociologie. Candollea 2013, 68, 5–27. [Google Scholar] [CrossRef]
  14. Myers, N.; Mittermeier, R.A.; Mittermeier, C.G.; da Fonseca, G.A.B.; Kent, J. Biodiversity hotspots for conservation priorities. Nature 2000, 403, 853–858. [Google Scholar] [CrossRef] [PubMed]
  15. Körner, C. Impact of atmospheric changes on alpine vegetation: The ecophysiological perspective. In Potential Ecological Impacts of Climate Change in the Alps and Fennoscandian Mountains; Guisan, A., Holten, J.I., Spichiger, R., Tessier, L., Eds.; Conservatoire et Jardin Botaniques de Genève: Geneva, Switzerland, 1995; pp. 113–120. [Google Scholar]
  16. Testolin, R.; Attorre, F.; Borchardt, P.; Brand, R.F.; Bruelheide, H.; Chytrý, M.; De Sanctis, M.; Dolezal, J.; Finckh, M.; Haider, S.; et al. Global patterns and drivers of alpine plant species richness. Global Ecol. Biogeogr. 2021, 30, 1218–1231. [Google Scholar] [CrossRef]
  17. Favarger, C. Endemism in the Montane Floras of Europe. In Taxonomy Phytogeography and Evolution; Valentine, D.H., Ed.; Academic Press: London, UK; New York, NY, USA, 1972; pp. 191–204. [Google Scholar]
  18. Gils, H.V.; Conti, F.; Ciaschetti, G.; Westinga, E. Fine resolution distribution modelling of endemics in Majella National Park, Central Italy. Plant Biosyst. 2012, 146, 276–287. [Google Scholar] [CrossRef]
  19. Hughes, P.D.; Woodward, J.C. Timing of glaciation in the Mediterranean mountains during the last cold stage. J. Quat. Sci. 2008, 23, 575–588. [Google Scholar] [CrossRef]
  20. Catonica, C.; Manzi, A. L’influenza della storia climatica e geologica recente sulla Flora d’alta quota dei gruppi montuosi del Gran Sasso e della Majella (Appennino Centrale). Riv. Piemont. Stor. Nat. 2002, 23, 19–29. [Google Scholar]
  21. Stocker, T.F.; Qin, D.; Plattner, G.-K.; Alexander, L.V.; Allen, S.K.; Bindoff, N.L.; Bréon, F.-M.; Church, J.A.; Cubasch, U.; Xie, S.-P. Technical Summary. In Climate Change 2013: The Physical Science Basis; Contribution of working group to the fifth assessment report of the intergovernmental panel on climate change; Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M., Eds.; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2013; pp. 33–109. [Google Scholar]
  22. Klein, G.; Rebetez, M.; Rixen, C.; Vitasse, Y. Unchanged risk of frost exposure for subalpine and alpine plants after snowmelt in Switzerland despite climate warming. Int. J. Biometeorol. 2018, 62, 1755–1762. [Google Scholar] [CrossRef]
  23. Pauli, H.; Gottfried, M.; Dullinger, S.; Abdaladze, O.; Akhalkatsi, M.; Alonso, J.L.B.; Coldea, G.; Dick, J.; Erschbamer, B.; Calzado, R.F.; et al. Recent plant diversity changes on Europe’s mountain summits. Science 2012, 336, 353–355. [Google Scholar] [CrossRef]
  24. Matteodo, M.; Ammann, K.; Verrecchia, E.P.; Vittoz, P. Snowbeds are more affected than other subalpine–alpine plant communities by climate change in the Swiss Alps. Ecol. Evol. 2016, 6, 6969–6982. [Google Scholar] [CrossRef]
  25. Steinbauer, M.J.; Grytnes, J.-A.; Jurasinski, G.; Kulonen, A.; Lenoir, J.; Pauli, H.; Rixen, C.; Winkler, M.; Bardy-Durchhalter, M.; Barni, E.; et al. Accelerated increase in plant species richness on mountain summits is linked to warming. Nature 2018, 556, 231–234. [Google Scholar] [CrossRef]
  26. Cannone, N.; Sgorbati, S.; Guglielmin, M. Unexpected impacts of climate change on alpine vegetation. Front. Ecol. Environ. 2007, 5, 360–364. [Google Scholar] [CrossRef]
  27. Holzinger, B.; Hulber, K.; Camenisch, M.; Grabherr, G. Changes in plant species richness over the last century in the eastern Swiss Alps: Elevational gradient, bedrock effects and migration rates. Plant Ecol. 2008, 195, 179–196. [Google Scholar] [CrossRef]
  28. Parolo, G.; Rossi, G. Upward migration of vascular plants following a climate warming trend in the Alps. Basic Appl. Ecol. 2008, 9, 100–107. [Google Scholar] [CrossRef]
  29. Britton, A.J.; Beale, C.M.; Towers, W.; Hewison, R.L. Biodiversity gains and losses: Evidence for homogenisation of Scottish alpine vegetation. Biol. Conserv. 2009, 142, 1728–1739. [Google Scholar] [CrossRef]
  30. Engler, R.; Randin, C.F.; Thuiller, W.; Dullinger, S.; Zimmermann, N.E.; Araújo, M.B.; Pearman, P.B.; Le Lay, G.; Piedallu, C.; Albert, C.H.; et al. 21st Century Climate Change Threatens Mountain Flora Unequally Across Europe. Glob. Change Biol. 2011, 17, 2330–2341. [Google Scholar] [CrossRef]
  31. Palombo, C.; Chirici, G.; Marchetti, M.; Tognetti, R. Is land abandonment affecting forest dynamics at high elevation in Mediterranean mountains more than climate change? Plant Biosyst. 2013, 147, 1–11. [Google Scholar] [CrossRef]
  32. Erschbamer, B.; Unterluggauer, P.; Winkler, E.; Mallaun, M. Changes in plant species diversity revealed by long-term monitoring on mountain summits in the Dolomites (northern Italy). Preslia 2011, 83, 387–401. [Google Scholar]
  33. Gottfried, M.; Pauli, H.; Futschik, A.; Akhalkatsi, M.; Barančok, P.; Benito Alonso, J.L.; Coldea, G.; Dick, J.; Erschbamer, B.; Fernández Calzado, M.R.; et al. Continent-wide response of mountain vegetation to climate change. Nat. Clim. Chang. 2012, 2, 111–115. [Google Scholar] [CrossRef]
  34. Stanisci, A.; Frate, L.; Di Cella, M.U.; Pelino, G.; Petey, M.; Siniscalco, C.; Carranza, M.L. Short-Term Signals of Climate Change in Italian Summit Vegetation: Observations at Two GLORIA Sites. Plant Biosyst. Int. J. Deal. All Asp. Plant Biol. 2014, 150, 227–235. [Google Scholar] [CrossRef]
  35. Evangelista, A.; Frate, L.; Carranza, M.L.; Attorre, F.; Pelino, G.; Stanisci, A. Changes in composition, ecology and structure of high-mountain vegetation: A re-visitation study over 42 years. AoB Plants 2016, 8, 1–11. [Google Scholar] [CrossRef]
  36. Cannone, N.; Pignatti, S. Ecological responses of plant species and communities to climate warming: Upward shift or range filling processes? Clim. Chang. 2014, 123, 201–214. [Google Scholar] [CrossRef]
  37. Calabrese, V.; Carranza, M.L.; Evangelista, A.; Marchetti, M.; Stinca, A.; Stanisci, A. Long-term changes in the composition, ecology, and structure of Pinus mugo scrubs in the Apennines (Italy). Diversity 2018, 10, 70. [Google Scholar] [CrossRef]
  38. Erschbamer, B.; Kiebache, T.; Mallaun, M.; Unterluggauer, P. Short-term signals of climate change along an altitudinal gradient in the South Alps. Plant Ecol. 2009, 202, 79–89. [Google Scholar] [CrossRef]
  39. Schwartz, M.W.; Iverson, L.R.; Prasad, A.M.; Matthews, S.N.; O’Connor, R.J. Predicting extinctions as a result of climate change. Ecology 2006, 87, 1611–1615. [Google Scholar] [CrossRef]
  40. Dirnbock, T.; Essl, F.; Rabitsch, W. Disproportional risk for habitat loss of high-altitude endemic species under climate change. Glob. Chang. Biol. 2011, 17, 990–996. [Google Scholar] [CrossRef]
  41. Fernández-Calzado, R.; Molero, M.J.; Merzouki, A.; Casares, P. Vascular plant diversity and climate change in the upper zone of Sierra Nevada, Spain. Plant Biosyst. 2012, 146, 1044–1053. [Google Scholar] [CrossRef]
  42. Nogués-Bravo, D.; Araújo, M.B.; Lasanta, T.; Moreno, J.I.L. Climate Change in Mediterranean Mountains during the 21st Century. Ambio 2008, 37, 280–285. [Google Scholar] [CrossRef]
  43. Cherif, S.; Doblas-Miranda, E.; Lionello, P.; Borrego, C.; Giorgi, F.; Iglesias, A.; Jebari, S.; Mahmoudi, E.; Moriondo, M.; Pringault, O.; et al. Climate and Environmental Change in the Mediterranean Basin–Current Situation and Risks for the Future. In Drivers of Change; Cramer, W., Guiot, J., Marini, K., Eds.; Union for the Mediterranean, Plan Bleu, UNEP/MAP: Marseille, France, 2021; pp. 59–180. [Google Scholar]
  44. Lamprecht, A.; Pauli, H.; Fernández Calzado, M.R.; Lorite, J.; Molero, M.J.; Steinbauer, K.; Winkler, M. Changes in plant diversity in a water-limited and isolated high-mountain range (Sierra Nevada, Spain). Alp. Bot. 2021, 131, 17–39. [Google Scholar] [CrossRef]
  45. Kazakis, G.; Ghosn, D.; Remoundou, I.; Nyktas, P.; Talias, M.A.; Vogiatzakis, I.N. Altitudinal Vascular Plant Richness and Climate Change in the Alpine Zone of the Lefka Ori, Crete. Diversity 2021, 13, 22. [Google Scholar] [CrossRef]
  46. Porro, F.; Tomaselli, M.; Abeli, T.; Gandini, M.; Gualmini, M.; Orsenigo, S.; Petraglia, A.; Rossi, G.; Carbognani, M. Could plant diversity metrics explain climate-driven vegetation changes on mountain summits of the GLORIA network? Biodivers. Conserv. 2019, 28, 3575–3596. [Google Scholar] [CrossRef]
  47. Nicklas, L.; Walde, J.; Wipf, S.; Lamprecht, A.; Mallaun, M.; Rixen, C.; Steinbauer, K.; Theurillat, J.-P.; Unterluggauer, P.; Vittoz, P.; et al. Climate Change Affects Vegetation Differently on Siliceous and Calcareous Summits of the European Alps. Front. Ecol. Evol. 2021, 9, 642309. [Google Scholar] [CrossRef]
  48. Turco, M.; Vezzoli, R.; De Ronco, P.; Mercogliano, P. Variation in Discharge, Precipitation and Temperature in Po River and Tributaries Basins; no. 185; Centro Euro-Mediterraneo sui Cambiamenti Climatici (CMCC): Lecce, Italy, 2013. [Google Scholar]
  49. Sillmann, J.; Kharin, V.V.; Zwiers, F.W.; Zhang, X.; Bronaugh, D. Climate extremes indices in the CMIP5 multimodel ensemble: Part 2. Future climate projections. J. Geophys. Res. Atmos. 2013, 118, 2473–2493. [Google Scholar] [CrossRef]
  50. Barredo, J.I.; Mauri, A.; Caudullo, G.; Dosio, A. Assessing Shifts of Mediterranean and arid climates under RCP4.5 and RCP8.5 climate projections in Europe. Pure Appl. Geophys. 2018, 175, 3955–3971. [Google Scholar] [CrossRef]
  51. Di Nuzzo, L.; Vallese, C.; Benesperi, R.; Giordani, P.; Chiarucci, A.; Di Cecco, V.; Di Martino, L.; Di Musciano, M.; Gheza, G.; Lelli, C.; et al. Contrasting multitaxon responses to climate change in Mediterranean mountains. Sci. Rep. 2021, 11, 4438. [Google Scholar] [CrossRef]
  52. Biondi, E.; Blasi, C.; Burrascano, S.; Casavecchia, S.; Copiz, R.; Del Vico, E.; Galdenzi, D.; Gigante, D.; Lasen, C.; Spampinato, G.; et al. Manuale Italiano di Interpretazione Degli Habitat Della Direttiva 92/43/CEE.; Ministero dell’Ambiente e della Tutela del Territorio e del Mare, Direzione per la Protezione della Natura: Rome, Italy, 2009; pp. 1–16.
  53. European Commission DG Environment. Interpretation Manual of European Union Habitats. [Eur 28. Nature ENV B.3]. 2013. Available online: http://ec.europa.eu/environment/nature/legislation/habitatsdirective/docs/Int_Manual_EU28.pdf (accessed on 26 June 2019).
  54. Jaurand, E. Les Heritages Glaciaire de l’Apennin. Ph.D. Thesis, Éditions de la Sorbonne Paris, Paris, France, 1994. [Google Scholar]
  55. Giraudi, C. Nuovi dati sul glacialismo della montagna della Majella (Abruzzo, Italia centrale). Ital. J. Quat. Sci. 1998, 11, 265–271. [Google Scholar]
  56. 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; FAO: Rome, Italy, 2015. [Google Scholar]
  57. Blasi, C.; Di Pietro, R.; Fortini, P.; Catonica, C. The main Plant community types of the alpine belt of the Apennine chain. Plant Biosyst. 2003, 137, 83–110. [Google Scholar] [CrossRef]
  58. Stanisci, A.; Carranza, M.L.; Pelino, G.; Chiarucci, A. Assessing the diversity pattern of cryophilous plant species in high elevation habitats. Plant Ecol. 2010, 212, 595–600. [Google Scholar] [CrossRef]
  59. Stanisci, A.; Pelino, G.; Blasi, C. Vascular plant diversity and climate change in the alpine belt of the central Apennines (Italy). Biodivers. Conserv. 2005, 14, 1301–1318. [Google Scholar] [CrossRef]
  60. Stanisci, A.; Evangelista, A.; Frate, L.; Stinca, A.; Carranza, M.L. VIOLA—Database of high mountain vegetation of central Apennines. Phytocoenologia 2016, 46, 231–232. [Google Scholar] [CrossRef]
  61. Cervellini, M.; Zannini, P.; Di Musciano, M.; Fattorini, S.; Jiménez-Alfaro, B.; Rocchini, D.; Field, R.; Vetaas, O.R.; Severin, D.H.I.; Beierkuhnlein, C.; et al. A grid-based map for the Biogeographical Regions of Europe. Biodivers. Data J. 2020, 8, e53720. [Google Scholar] [CrossRef] [PubMed]
  62. Petriccione, B.; Bricca, A. Thirty years of ecological research at the Gran Sasso d’Italia LTER site: Climate change in action. Nat. Cons. 2019, 34, 9–39. [Google Scholar] [CrossRef]
  63. Steinbauer, K.; Lamprecht, A.; Winkler, M.; Di Cecco, V.; Fasching, V.; Ghosn, D.; Maringer, A.; Remoundou, I.; Suen, M.; Stanisci, A.; et al. Recent changes in high-mountain plant community functional composition in contrasting climate regimes. Sci. Total Environ. 2022, 829, 154541. [Google Scholar] [CrossRef] [PubMed]
  64. Intergovernmental Panel on Climate Change. Sixth Assessment Report. Climate Change 2022: Impacts, Adaptation and Vulnerability; Cambridge University Press: Cambridge, UK, 2022. [Google Scholar]
  65. Pauli, H.; Gottfried, M.; Lamprecht, A.; Niessner, S.; Rumpf, S.; Winkler, M.; Steinbauer, K.; Grabherr, G. The GLORIA Field Manual; Global Observation Research Initiative in Alpine Environments: Vienna, Austria, 2015. [Google Scholar]
  66. Chytrý, M.; Tichý, L.; Hennekens, S.M.; Knollová, I.; Janssen, J.A.M.; Rodwell, J.S.; Peterka, T.; Marcenò, C.; Landucci, F.; Danihelka, J.; et al. EUNIS Habitat Classification: Expert system, characteristic species combinations and distribution maps of European habitats. Appl. Veg. Sci. 2020, 23, 648–675. [Google Scholar] [CrossRef]
  67. Biondi, E.; Blasi, C.; Allegrezza, M.; Anzellotti, I.; Azzella, M.M.; Carli, E.; Casavecchia, S.; Copiz, R.; Del Vico, E.; Facioni, L.; et al. Plant communities of Italy: The Vegetation Prodrome. Plant Biosyst. 2014, 148, 728–814. [Google Scholar] [CrossRef]
  68. Evangelista, A.; Frate, L.; Stinca, A.; Carranza, M.L.; Stanisci, A. VIOLA-The vegetation database of the central Apennines: Structure, current status and usefulness for monitoring Annex i EU habitats (92/43/EEC). Plant Sociol. 2016, 53, 47–58. [Google Scholar] [CrossRef]
  69. Vymazalová, M.; Axmanová, I.; Tichý, L. Effect of intra-seasonal variability on vegetation data. J. Veg. Sci. 2012, 23, 978–984. [Google Scholar] [CrossRef]
  70. Chytrý, M.; Tichý, L.; Hennekens, S.M.; Schaminée, J.H.J. Assessing vegetation change using vegetation-plot databases: A risky business. Appl. Veg. Sci. 2014, 17, 32–41. [Google Scholar] [CrossRef]
  71. Braun-Blanquet, J. Pflanzensoziologie: Grundzüge der Vegetationskunde; Springer: Berlin, Germany, 1964; ISBN 3709140781. [Google Scholar]
  72. Westhoff, V.; Van Der Maarel, E. The braun-blanquet approach. In Classification of Plant Communities; Springer: Berlin/Heidelberg, Germany, 1978; pp. 287–399. [Google Scholar]
  73. Bartolucci, F.; Peruzzi, L.; Galasso, G.; Albano, A.; Alessandrini, A.; Ardenghi, N.M.G.; Astuti, G.; Bacchetta, G.; Ballelli, S.; Banfi, E.; et al. An updated checklist of the vascular flora native to Italy. Plant Biosyst. 2018, 152, 179–303. [Google Scholar] [CrossRef]
  74. Pignatti, S.; Guarino, R.; La Rosa, M. Flora d’Italia; Edagricole: Rome, Italy, 2017; Volume 1, ISBN 8850652429. [Google Scholar]
  75. Landolt, E.; Bäumler, B.; Erhardt, A.H.O.; Klötzli, F.; Lämmler, W.; Nobis, M.; Rudmann-Maurer, K.; Schweingruber, F.H.; Theurillat, J.-P.; Urmi, E.; et al. Flora Indicativa. Ecological Indicator Values and Biological Attributes of the Flora of Switzerland and the Alps, 2nd ed.; Haupt Verlag: Berna, Switzerland, 2010. [Google Scholar]
  76. Frate, L.; Carranza, M.L.; Evangelista, A.; Stinca, A.; Schaminée, J.H.J.; Stanisci, A. Climate and land use change impacts on Mediterranean high-mountain vegetation in the Apennines since the 1950s. Plant Ecol. Divers. 2018, 11, 85–96. [Google Scholar] [CrossRef]
  77. Tichý, L.; Hennekens, S.M.; Novák, P.; Rodwell, J.S.; Schaminée, J.H.J.; Chytrý, M. Optimal transformation of species cover for vegetation classification. Appl. Veg. Sci. 2020, 23, 710–717. [Google Scholar] [CrossRef]
  78. Raunkiær, C.C. The Life Forms of Plants and Statistical Plant Geography; Oxford University Press: Oxford, UK, 1934. [Google Scholar]
  79. Clarke, K.R. Non-parametric multivariate analyses of changes in community structure. Aust. J. Ecol. 1993, 18, 117–143. [Google Scholar] [CrossRef]
  80. R Core Team. A Language and Environment for Statistical Computing; R Foundation for Statistical Computing: Vienna, Austria, 2020; Available online: http://www.r-project.org/index.html (accessed on 7 February 2021).
  81. Oksanen, A.J.; Blanchet, F.G.; Friendly, M.; Kindt, R.; Legendre, P.; Mcglinn, D.; Minchin, P.R.; Hara, R.B.O.; Simpson, G.L.; Solymos, P.; et al. Vegan Package 2020. Available online: https://github.com/vegandevs/vegan (accessed on 7 February 2021).
  82. Hammer, Ø.; Harper, D.A.T.; Ryan, P.D. Past Palaeontological Statistics, Ver. 1.79. 2001, pp. 1–88. Available online: https://palaeo-electronica.org/2001_1/past/issue1_01.htm#:~{}:text=The%20program%2C%20called%20PAST%20(PAleontological,plotting%2C%20and%20simple%20phylogenetic%20analysis (accessed on 10 June 2021).
  83. Mariotti, A.; Zeng, N.; Yoon, J.-H.; Artale, V.; Navarra, A.; Alpert, P.; Li, L.Z.X. Mediterranean water cycle changes: Transition to drier 21st century conditions in observations and CMIP3 simulations. Environ. Res. Lett. 2008, 3, 044001. [Google Scholar] [CrossRef]
  84. Toreti, A.; Fioravanti, G.; Percontia, W.; Desiatoa, F. Annual and seasonal precipitation over Italy from 1961 to 2006. Int. J. Climatol. 2009, 29, 1976–1987. [Google Scholar] [CrossRef]
  85. Fernández Calzado, M.R.; Molero, J. Changes in the summit flora of a Mediterranean mountain (Sierra Nevada, Spain) as a possible effect of climate change. Lazaroa 2013, 34, 65–75. [Google Scholar] [CrossRef]
  86. Brunetti, M.; Buffoni, L.; Mangianti, F.; Maugeri, M.; Nanni, T. Temperature, precipitation and extreme events during the last century in Italy. Glob. Planet Chang. 2004, 40, 141–149. [Google Scholar] [CrossRef]
  87. Elguindi, N.; Rauscher, S.A.; Giorgi, F. Historical and future changes in maximum and minimum temperature records over Europe. Clim. Chang. 2013, 117, 415–431. [Google Scholar] [CrossRef]
  88. Simolo, C.; Brunetti, M.; Maugeri, M.; Nanni, T. Increasingly warm summers in the Euro-Mediterranean zone: Mean temperatures and extremes. Reg. Environ. Chang. 2014, 14, 1825–1832. [Google Scholar] [CrossRef]
  89. Christensen, J.H.; Hewiston, B.; Busuioc, A.; Chen, A.; Gao, X.; Held, I.; Jones, R.; Kolli, R.K.; Kwon, W.T.; Laprise, R.; et al. AR4 Climate Change 2007: The Physical Science Basis; Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge Univ. Press: Cambridge, UK, 2007; pp. 847–940. [Google Scholar]
  90. Benito, B.; Lorite, J.; Penas, J. Simulating potential effects of climatic warming on altitudinal patterns of key species in Mediterranean-alpine ecosystems. Clim. Chang. 2011, 108, 471–483. [Google Scholar] [CrossRef]
  91. Mondoni, A.; Rossi, G.; Orsenigo, S.; Probert, R.J. Climate warming could shift the timing of seed germination in alpine plants. Ann. Bot. 2012, 110, 155–164. [Google Scholar] [CrossRef]
  92. Giménez-Benavides, L.; Escudero, A.; Pérez-García, F. Seed germination of high mountain Mediterranean species: Altitudinal, interpopulation and interannual variability. Ecol. Res. 2005, 20, 433–444. [Google Scholar] [CrossRef]
  93. Orsenigo, S.; Abeli, T.; Rossi, G.; Bonasoni, P.; Pasquaretta, C.; Gandini, M.; Mondoni, A. Effects of autumn and spring heat waves on seed germination of high mountain plants. PLoS ONE 2015, 10, e0133626. [Google Scholar] [CrossRef]
  94. Varricchione, M.; Di Cecco, V.; Santoianni, L.A.; Stanisci, A.; Di Febbraro, M.; Di Martino, L.; Carranza, M.L. Diagnostic Species Diversity Pattern Can Provide Key Information on Vegetation Change: An Insight into High Mountain Habitats in Central Apennines. J. Zool. Bot. Gard. 2021, 2, 453–472. [Google Scholar] [CrossRef]
  95. Edwards, A.C.; Scalenghe, R.; Freppaz, M. Changes in the seasonal snow cover of alpine regions and its effect on soil processes: A review. Quat. Int. 2007, 162–163, 172–181. [Google Scholar] [CrossRef]
  96. Stanisci, A.; Bricca, A.; Calabrese, V.; Cutini, M.; Pauli, H.; Steinbauer, K.; Carranza, M.L. Functional composition and diversity of leaf traits in subalpine versus alpine vegetation in the Apennines. AoB Plants 2020, 12, 1–11. [Google Scholar] [CrossRef]
  97. Haslinger, K.; Schöner, W.; Anders, I. Future drought probabilities in the Greater Alpine Region based on COSMO-CLM experiments–Spatial patterns and driving forces. Meteorol. Z. 2016, 25, 137–148. [Google Scholar] [CrossRef]
  98. Cutini, M.; Marzialetti, F.; Giuliana, B.; Rianna, G.; Theurillat, J.-P. Bioclimatic pattern in a Mediterranean mountain area: Assessment from a classification approach on a regional scale. Int. J. Biometeorol. 2021, 65, 1085–1097. [Google Scholar] [CrossRef]
  99. Chen, I.-C.; Hill, J.K.; Ohlemüller, R.; Roy, D.B.; Thomas, C.D. Rapid Range Shifts of Species Associated with High Levels of Climate Warming. Science 2011, 333, 1024–1026. [Google Scholar] [CrossRef]
  100. Rehm, E.M.; Olivas, P.; Stroud, J.; Feeley, K.J. Losing your edge: Climate change and the conservation value of range-edge populations. Ecol. Evol. 2015, 5, 4315–4326. [Google Scholar] [CrossRef]
  101. Jump, A.S.; Peñuelas, J. Running to stand still: Adaptation and the response of plants to rapid climate change. Ecol. Lett. 2005, 8, 1010–1020. [Google Scholar] [CrossRef]
  102. Hoffmann, S.; Irl, S.D.H.; Beierkuhnlein, C. Predicted climate shifts within terrestrial protected areas worldwide. Nat. Commun. 2019, 10, 1–10. [Google Scholar] [CrossRef]
  103. Manes, S.; Costello, M.J.; Beckett, H.; Debnath, A.; Devenish-Nelson, E.; Grey, K.-A.; Jenkins, R.; Ming Khan, T.; Kiessling, W.; Krause, C.; et al. Endemism increases species’ climate change risk in areas of global biodiversity importance. Biol. Conserv. 2021, 25, 109270. [Google Scholar] [CrossRef]
  104. Watts, S.H.; Mardon, D.K.; Mercer, C.; Watson, D.; Cole, H.; Shaw, R.F.; Jump, A.S. Riding the elevator to extinction: Disjunct arctic-alpine plants of open habitats decline as their more competitive neighbors expand. Biol. Conserv. 2022, 272, 109620. [Google Scholar] [CrossRef]
  105. Nagy, L.; Grabherr, G.; Körner, C.; Thompson, D.B.A. Alpine Biodiversity in Europe; Springer: Berlin, Germany, 2003. [Google Scholar]
  106. Kougioumoutzis, K.; Kokkoris, I.P.; Panitsa, M.; Trigas, P.; Strid, A.; Dimopoulos, P. Plant Diversity Patterns and Conservation Implications under Climate-Change Scenarios in the Mediterranean: The Case of Crete (Aegean, Greece). Diversity 2020, 12, 270. [Google Scholar] [CrossRef]
  107. Di Biase, L.; Pace, L.; Mantoni, C.; Fattorini, S. Variations in Plant Richness, Biogeographical Composition, and Life Forms along an Elevational Gradient in a Mediterranean Mountain. Plants 2021, 10, 2090. [Google Scholar] [CrossRef] [PubMed]
  108. Gentili, R.; Baroni, C.; Caccianiga, M.; Armiraglio, S.; Ghiani, A.; Citterio, S. Potential warm-stage microrefugia for alpine plants: Feedback between geomorphological and biological processes. Ecol. Complex. 2015, 21, 87–99. [Google Scholar] [CrossRef]
  109. Veron, S.; Faith, D.P.; Pellens, R.; Pavoine, S. Priority areas for phylogenetic diversity: Maximising gains in the mediterranean basin. In Phylogenetic Diversity: Applications and Challenges in Biodiversity Science; Springer International Publishing: Cham, Switzerland, 2018; pp. 145–166. [Google Scholar]
  110. Haight, J.; Hammill, E. Protected areas as potential refugia for biodiversity under climatic change. Biol. Conserv. 2020, 241, 108258. [Google Scholar] [CrossRef]
  111. Di Cecco, V.; Di Santo, M.; Di Musciano, M.; Manzi, A.; Di Cecco, M.; Ciaschetti, G.; Marcantonio, G.; Di Martino, L. The Majella National Park: A case study for the conservation of plant biodiversity in the Italian Apennines. Ital. Bot. 2020, 10, 1–24. [Google Scholar] [CrossRef]
  112. Hawkes, J.G.; Maxted, N.; Ford-Lloyd, B.V. The Ex Situ Conservation of Plant Genetic Resources; Springer Science and Business Media: Berlin, Germany, 2012. [Google Scholar]
Figure 1. Study area (Maiella National Park) along with the distribution of vegetation plots on Apennine stripped grasslands (EUNIS code E4.436) and wind edge swards (EUNIS code E4.42) and of the GLORIA data-loggers measuring soil temperatures on Mt. Femmina Morta.
Figure 1. Study area (Maiella National Park) along with the distribution of vegetation plots on Apennine stripped grasslands (EUNIS code E4.436) and wind edge swards (EUNIS code E4.42) and of the GLORIA data-loggers measuring soil temperatures on Mt. Femmina Morta.
Diversity 14 00695 g001
Figure 2. Annual and autumn (September–November) mean soil temperatures (°C) and vegetative period (e.g., the number of days with temperatures above 2 °C) for the period 2002 to 2020. Asterisks indicate significant trends. Data were retrieved from data-loggers of GLORIA monitoring network (Global Observation Research Initiative in Alpine Environments) IT CAM FEM (2405 m a.s.l.).
Figure 2. Annual and autumn (September–November) mean soil temperatures (°C) and vegetative period (e.g., the number of days with temperatures above 2 °C) for the period 2002 to 2020. Asterisks indicate significant trends. Data were retrieved from data-loggers of GLORIA monitoring network (Global Observation Research Initiative in Alpine Environments) IT CAM FEM (2405 m a.s.l.).
Diversity 14 00695 g002
Figure 3. Boxplot comparing cover (grey) and species richness per 4 × 4 m plot (white) of overall and endemic species in two time steps (T1: 2003 and T2: 2021) in Apennine stripped grasslands (A) and wind edge swards (B). Only the significant results are reported. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01, *** p < 0.001).
Figure 3. Boxplot comparing cover (grey) and species richness per 4 × 4 m plot (white) of overall and endemic species in two time steps (T1: 2003 and T2: 2021) in Apennine stripped grasslands (A) and wind edge swards (B). Only the significant results are reported. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01, *** p < 0.001).
Diversity 14 00695 g003
Figure 4. Boxplot comparing cover (grey) and richness per 4 × 4 m plot (white) of alpine species (T: 1–1.5) and subalpine species (T: 2–2.5) in two time steps (T1: 2003 and T2: 2021) in Apennine stripped grassland (A) and wind edge swards (B). Only the significant results are reported. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01).
Figure 4. Boxplot comparing cover (grey) and richness per 4 × 4 m plot (white) of alpine species (T: 1–1.5) and subalpine species (T: 2–2.5) in two time steps (T1: 2003 and T2: 2021) in Apennine stripped grassland (A) and wind edge swards (B). Only the significant results are reported. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01).
Diversity 14 00695 g004
Figure 5. Boxplot comparing cover of growth forms (sensu Pignatti et al. [74]) in two time steps (T1: 2003 and T2: 2021 in Apennine stripped grasslands (A) and wind edge swards (B). H ros: hemicryptophytes with rosette; H scap: scapose hemicryptophytes; CH suffr: suffruticose chamaephytes. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01).
Figure 5. Boxplot comparing cover of growth forms (sensu Pignatti et al. [74]) in two time steps (T1: 2003 and T2: 2021 in Apennine stripped grasslands (A) and wind edge swards (B). H ros: hemicryptophytes with rosette; H scap: scapose hemicryptophytes; CH suffr: suffruticose chamaephytes. Asterisks indicate significant differences according to the Kruskal–Wallis test for equal medians (* p < 0.05, ** p < 0.01).
Diversity 14 00695 g005
Table 1. Endemic species present in the analyzed alpine plant communities along with their relative growth form (sensu Pignatti et al. [74]), average altitudinal distribution of the species, and species’ ecological indicator value for temperature (EIVs T) sensu Landolt [75]. CH pulv: pulvinate chamaephytes; CH suffr: suffruticose chamaephytes; H caesp: caespitose hemicryptophytes; H ros: hemicryptophytes with rosette; H scap: scapose hemicryptophytes; and T scap: scapose therophytes.
Table 1. Endemic species present in the analyzed alpine plant communities along with their relative growth form (sensu Pignatti et al. [74]), average altitudinal distribution of the species, and species’ ecological indicator value for temperature (EIVs T) sensu Landolt [75]. CH pulv: pulvinate chamaephytes; CH suffr: suffruticose chamaephytes; H caesp: caespitose hemicryptophytes; H ros: hemicryptophytes with rosette; H scap: scapose hemicryptophytes; and T scap: scapose therophytes.
SpeciesGrowth FormAverage Altitudinal DistributionEIVs T
Achillea barrelieri subsp. barrelieriH scapAlpine1.5
Alyssum diffusum subsp. diffusumCH suffrMontane3
Androsace vitaliana subsp. praetutianaCH suffrAlpine1.5
Armeria gracilis subsp. majellensisH rosSubalpine2
Carduus chrysacanthusH scapSubalpine2.5
Cerastium thomasiiCH suffrAlpine1.5
Erysimum majellenseH scapSubalpine2.5
Festuca violacea subsp. italicaH caespAlpine1.5
Galium magellenseH scapAlpine1.5
Helictochloa praetutiana subsp. praetutianaH caespSubalpine2.5
Leontopodium nivaleH scapAlpine1.5
Myosotis grauiH scapSubalpine2.5
Noccaea stylosaCH suffrSubalpine2
Pedicularis elegansH rosSubalpine2.5
Phyllolepidum rupestreCH suffrAlpine1.5
Ranunculus pollinensisH scapSubalpine2.5
Rhinanthus wettsteiniiT scapSubalpine2
Saxifraga italicaCH pulvAlpine1
Scorzoneroides montana subsp. breviscapaH rosAlpine1.5
Trifolium pratense subsp. semipurpureumH scapMontane3.5
Valeriana saliuncaH scapAlpine1
Viola eugeniae subsp. eugeniaeH scapSubalpine2.5
Viola magellensisH scapAlpine1
Table 2. Plant species that contribute up to the 50% to floristic changes over time (T1: 2003 and T2: 2021) assessed by the similarity percentage procedure (SIMPER) [79]. For each taxon, the growth form, the average altitudinal distribution of the species (Range), and species’ ecological indicator value for temperature (EIVs T) sensu Landolt [75], as well as the specific and cumulative contribution and the mean cover on T1 and T2, are also reported.
Table 2. Plant species that contribute up to the 50% to floristic changes over time (T1: 2003 and T2: 2021) assessed by the similarity percentage procedure (SIMPER) [79]. For each taxon, the growth form, the average altitudinal distribution of the species (Range), and species’ ecological indicator value for temperature (EIVs T) sensu Landolt [75], as well as the specific and cumulative contribution and the mean cover on T1 and T2, are also reported.
Plant CommunitySpeciesGrowth FormRangeEIVs Species Contribution (%)Cumulative Contribution (%)Mean Cover
TT1T2
Apennine stripped grasslandsPlantago atrata subsp. atrataH rosAlpine1.5101020.58.8
Anthyllis vulneraria subsp. pulchellaH scapMontane37.917.911.715.4
Carex kitaibelianaH caespSubalpine272518.520.6
° Armeria gracilis subsp. majellensisH rosSubalpine26.731.714.83.1
° Festuca violacea subsp. italicaH caespAlpine1.56.43813.97.3
Poa alpina subsp. alpinaH caespAlpine1.55.643.6122.6
Helianthemum oelandicum subsp. alpestreCH suffrSubalpine2.5548.674.5
Silene acaulis subsp. bryoidesCH pulvAlpine14.152.75.95.6
Wind edge swardsCarex kitaibelianaH caespSubalpine212.412.425.120.3
Helianthemum oelandicum subsp. alpestreCH suffrSubalpine2.511.123.539.423.3
Salix retusaCH frutAlpine1.57.931.45.913.3
° Festuca violacea subsp. italicaH caespAlpine1.55.737.110.74.9
Carex myosuroidesH caespAlpine1.55.442.54.38.5
Oxytropis campestrisH scapAlpine1.54.246.75.92.6
Anthyllis vulneraria subsp. pulchellaH scapMontane34.250.91414
° Endemic species.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Varricchione, M.; Carranza, M.L.; Di Cecco, V.; Di Martino, L.; Stanisci, A. Warmer and Poorer: The Fate of Alpine Calcareous Grasslands in Central Apennines (Italy). Diversity 2022, 14, 695. https://doi.org/10.3390/d14090695

AMA Style

Varricchione M, Carranza ML, Di Cecco V, Di Martino L, Stanisci A. Warmer and Poorer: The Fate of Alpine Calcareous Grasslands in Central Apennines (Italy). Diversity. 2022; 14(9):695. https://doi.org/10.3390/d14090695

Chicago/Turabian Style

Varricchione, Marco, Maria Laura Carranza, Valter Di Cecco, Luciano Di Martino, and Angela Stanisci. 2022. "Warmer and Poorer: The Fate of Alpine Calcareous Grasslands in Central Apennines (Italy)" Diversity 14, no. 9: 695. https://doi.org/10.3390/d14090695

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