Next Article in Journal
Distribution Patterns and Multilevel Factors of the Innovation Activities of China’s New Energy Vehicle Industry
Previous Article in Journal
A Study of Diffusion Equation-Based Land-Use/Land-Cover Change Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Remote Sensing Methods for the Biophysical Characterization of Protected Areas Globally: Challenges and Opportunities

1
Soil Erosion and Conservation Research Group, Spanish Research Council (CEBAS-CSIC), E-30100 Murcia, Spain
2
Joint Research Centre (JRC), European Commission, I-21027 Ispra, Italy
3
School of Natural Science, Bangor University, Bangor LL57 2DG, UK
4
Rothamsted Research, Harpenden AL5 2JQ, UK
5
Arcadia Sistemi Informativi Territoriali Sr., I-27029 Vigevano, Italy
*
Author to whom correspondence should be addressed.
ISPRS Int. J. Geo-Inf. 2021, 10(6), 384; https://0-doi-org.brum.beds.ac.uk/10.3390/ijgi10060384
Submission received: 24 March 2021 / Revised: 3 May 2021 / Accepted: 28 May 2021 / Published: 4 June 2021
(This article belongs to the Special Issue Conservation Applications of Spatial Analysis and Remote Sensing)

Abstract

:
Protected areas (PAs) are a key strategy to reverse global biodiversity declines, but they are under increasing pressure from anthropogenic activities and concomitant effects. Thus, the heterogeneous landscapes within PAs, containing a number of different habitats and ecosystem types, are in various degrees of disturbance. Characterizing habitats and ecosystems within the global protected area network requires large-scale monitoring over long time scales. This study reviews methods for the biophysical characterization of terrestrial PAs at a global scale by means of remote sensing (RS) and provides further recommendations. To this end, we first discuss the importance of taking into account the structural and functional attributes, as well as integrating a broad spectrum of variables, to account for the different ecosystem and habitat types within PAs, considering examples at local and regional scales. We then discuss potential variables, challenges and limitations of existing global environmental stratifications, as well as the biophysical characterization of PAs, and finally offer some recommendations. Computational and interoperability issues are also discussed, as well as the potential of cloud-based platforms linked to earth observations to support large-scale characterization of PAs. Using RS to characterize PAs globally is a crucial approach to help ensure sustainable development, but it requires further work before such studies are able to inform large-scale conservation actions. This study proposes 14 recommendations in order to improve existing initiatives to biophysically characterize PAs at a global scale.

1. Introduction

Protected areas (PAs) are one of the main conservation strategies to counter the current biodiversity crisis [1]. However, PAs are under ongoing social, economic and environmental threats, and so the conservation of biodiversity within PAs and the restoration of PAs constitute one of the main current sociopolitical challenges [2]. The long-term conservation benefits of PAs depend on timely management actions based on relevant data and models that can predict the responses of ecosystems to various stress factors [3,4].
Anthropogenic activities and the changes in land use they generate have an impact on how efficient PAs are at protecting biodiversity globally [5,6]. Moreover, climate change impacts severely affect PAs, including increased frequency of flooding, soil erosion and plant water stress [7]. It is increasingly recognized that even large, historically stable ecosystems (such as the Amazon) are threatened and could undergo regime shifts to alternative ecosystems within 50 years [8].
The 2030 Agenda for Sustainable Development (https://www.un.org/sustainabledevelopment/sustainable-development-goals/; accessed 1 June 2021) recognizes that social and economic development depends on the sustainable management of our planet’s natural resources. The new EU Biodiversity Strategy for 2030 has further set ambitious goals and objectives regarding PAs, which include (1) legally protecting 30% of the EU’s land area and 30% of the EU’s sea area, (2) strictly protecting at least a third of the EU’s PAs and (3) effectively managing and monitoring all PAs [9]. The forthcoming new global strategy of the UN Convention on Biological Diversity (CBD) is also likely to set targets that are more ambitious than those set for 2020. Addressing these challenges requires large-scale integrated studies that characterize PAs as well as knowledge sharing platforms where scientists, managers and policy makers can work together to address the challenges mentioned above [10,11].
Recently, there has been an increase in the attention paid not only to the conservation of biodiversity within PAs but also to the preservation of important habitat and ecosystem functions and services [12,13,14,15]. Indeed, natural ecosystems provide us, among others, with drinking water, timber, food, pollination and carbon storage, as well as cultural and spiritual services. This was examined in detail in the Millennium Ecosystem Assessment [16] and was further reflected in Aichi Target 11 (https://www.cbd.int/aichi-targets/target/11; accessed 1 June 2021), adopted by CBD parties in 2010. Moreover, habitat and ecosystem characterizations can provide important complementary insights to the more commonly used species-based approaches to conservation [17,18,19].
Remote sensing is considered a valuable source of information for the management of natural resources and landscapes [20,21,22], as well as for the development of indicators for monitoring progress toward international environmental targets such as the Sustainable Development Goals (SDGs) [23,24]. Available time series allow, among others, the monitoring of vegetation conditions, landscape and habitat changes, land degradation, the assessment of ecosystem services, the identification of disturbed areas and the monitoring of the spread of invasive species [25,26,27,28]. They thus help to understand an ecosystem’s response and resilience to multiple stressors [29]. In this regard, remote sensing has revolutionized our ability to monitor PAs over the past decade [20,30,31,32,33].
Several broad types of applications can be supported by RS data and models in relation to PAs. A first type would be the near real-time monitoring of biodiversity, pressures and threats, environmental anomalies (such as weather and vegetation) and events such as fires, floods and storms, all highly relevant to inform, among other things, day-to-day PA management, enforcement and risk management [34,35,36]. A second type of application would be the mapping and assessment of specific habitats and ecosystems, relevant for, for example, management plans, monitoring strategies or condition assessment. This latter type of study paves the way for a third type of application that extends specific habitat or ecosystem mapping and assessment methods and integrates this information to systematically characterize PAs based on their ecological complexity, relevant for, for example, zoning plans, assessment of representativeness, prioritization of PAs or the identification of new areas requiring protection [37,38,39,40]. This paper focuses on these biophysical characterization applications.
While there have been a few attempts to characterize landscapes from an ecological perspective from local to regional scales [41,42,43,44,45], global characterization of PAs is urgently needed for the identification of gaps in current protection efforts, the systematic design of complementary PAs, raising awareness about the ecological values of PAs and to support international policy initiatives aimed at preserving biodiversity and ensuring a high provision of ecosystem services [46]. Moreover, global biophysical characterization of PAs can also facilitate and complement biodiversity-based protection initiatives and characterizations [47,48,49]. As an example of previous global efforts, the “terrestrial ecoregions of the world” [6,50] represent a set of large, ecologically meaningful regions at the global scale, containing distinct assemblages of natural communities and species, but they do not provide additional information on ecosystems contained within those ecoregions and have instead been used to prioritize the importance of conservation of larger regions [51].
This study seeks to provide recommendations for the biophysical characterization of terrestrial PAs at the global scale by means of RS. To this end, in Section 2, we discuss the importance of taking into account structural and functional attributes, as well as integrating a broad spectrum of variables, to account for the different ecosystem and habitat types within PAs, reviewing examples at the local and regional scales. In Section 3, we discuss potential candidate input variables at the global scale for the characterization of PAs, as well as the challenges and limitations of existing global environmental stratifications and the biophysical characterization of PAs, and offer recommendations. Computational and interoperability issues are also discussed, as well as the potential for cloud-based platforms linked to earth observations to support large-scale characterization of PAs. Finally, Section 4 provides a summary list of recommendations. Although focusing on terrestrial areas, we also mention a few examples of RS data used to characterize marine protected areas (MPAs).

2. Relevant Ecological Units and Descriptors

In order to comprehend the ecological complexity in PAs, biophysical characterization of PAs should take into account the different ecosystem and habitat types that are present within them and distinguish their ecological attributes as much as possible, including structural and functional attributes. To this end, a wide range of environmental descriptors should be included in the analysis, including drivers that ultimately shape ecosystems (Figure 1).
The assessment of structural attributes, such as vegetation height or heterogeneity by means of RS, helps distinguish characteristic ecosystems and habitats within PAs such as forests, wetlands, grasslands, shrublands, dunes and riparian habitats, among others. Furthermore, RS variables related to functional attributes, such as vegetation phenology or energy fluxes, have been proven to complement and improve habitat and ecosystem classifications based only on structural features by capturing the occurrence of natural disturbances and vegetation productivity, among others [52,53,54]. Several studies have reviewed the use of RS for assessing habitat and ecosystem structure, functions and condition in PAs at the local and regional scales [24,55,56,57,58,59,60,61,62].
With regard to structural attributes, wetlands, riparian forests and dune habitats, for example, have been mapped by means of texture- and object-based RS data analysis and machine learning algorithms in order to characterize and monitor changes in PAs [63,64,65,66,67,68,69,70,71,72,73]. Grasslands have been accurately mapped using time series of RS data [74]. Forest and shrubland structure has been mapped by means of very high resolution imagery [75,76,77]. Tree species richness across the tropics has been mapped by means of full-waveform lidar data [78]. Vegetation structure has been mapped at the local and regional levels in PAs by means of manned and unmanned aerial vehicles carrying airborne LiDAR and multi- and hyperspectral sensors [79,80,81]. Chetan and Dornik [82] quantified changes in vegetation greenness and structures within Natura 2000 sites over 20 years. The vegetation heterogeneity and pattern was characterized by means of image texture measures (e.g., gray-level co-occurrence matrix) derived from RS data [83,84,85,86,87,88,89].
In relation to functional attributes, several studies have quantified vegetation productivity over time by means of remote sensing-derived indices and found correlation with biodiversity patterns [90,91,92,93]. Moreover, the effect of disturbances such as post-fire forest vegetation regrowth was studied by means of different RS vegetation indices [94,95]. For a recent review of methods, sensors and ecosystem structural and functional attributes assessed by means of RS in PAs, see [33].
Furthermore, given the inherent ecological complexity that can be found within PAs, their systematic characterization needs to extend specific habitat or ecosystem mapping and assessment methods so that all habitat and ecosystem components that are present within them are taken into account [96,97]. By stratifying the natural landscape into homogeneous regions defining ecological units, the complexity of PAs can be converted into something that is more manageable and understandable [98]. For example, if a protected landscape contains both a lake and mountains, separating both elements cartographically would help inform and support adaptive management. In this regard, methods to characterize PAs should rely on a comprehensive list of environmental quantitative descriptors based on RS data, which could be categorized into different topics: (1) vegetation, including structure, phenology and disturbances; (2) climate; (3) water budget; (4) energy exchanges; (5) terrain and (6) soil, among others.
As was previously mentioned, vegetation-related variables, such as the amount of woody and herbaceous biomass or different vegetation indices, can help us distinguish between broad ecosystem types (such as forests, grasslands or wetlands) by capturing their structure, phenology and productivity [99]. Climatic descriptors, such as precipitation and temperature, are also important variables to be included in biophysical assessments to represent seasonality, extremes and limiting climatic factors [100,101,102,103]. Topographic gradients drive many patterns and processes in hydrology and ecology and are key to understanding the variation of habitats and biodiversity [104,105]. Water-related variables are also a good proxy for plant water stress and the presence of aquatic ecosystems and can therefore supplement the information on climate and vegetation by distinguishing differing responses to available water [106,107,108]. Variables that describe the energy exchanges between the land surface and the atmosphere, as well as the partition of energy into ground and vegetation, are also essential for ecological assessment and modeling [109].
Soil data are often ignored when characterizing PAs, but more than 25% of the Earth’s species live only in the soil [110]. Aside from that, soils form the foundation for many vegetation types and provide key supporting ecosystem services that are crucial for the maintenance of other types of services [111]. Given that soil biodiversity cannot be directly monitored by RS, soil descriptors that can be directly or indirectly monitored by RS and modeling can act as proxies [112,113]. In this regard, soil organic carbon appears to be one of the main drivers of soil microbial biodiversity at the global scale [114,115,116], particularly in extreme environments with low net primary productivity such as polar [117] and dryland regions [118]. Soil texture is also a relevant descriptor, since previous research has demonstrated that soil biota abundance and biodiversity, particularly soil microorganisms, increase with a decreasing soil particle size [119].

3. Global Characterization of Protected Areas

3.1. Global Input Variables and Data Sources

In the previous section, we reviewed the importance of taking into account the structural and functional attributes as well as integrating a broad spectrum of variables to account for the different ecosystem and habitat types within PAs. In this subsection, we give a list of potential candidate input variables mapped at global scale for global characterizations of PAs and discuss some limitations and recommendations.
Data sources presenting time series and regular updates at the global scale should be favored over single records in time to allow for the assessment of change over time and identify reference conditions. When correlated variables are used, principal component analysis can often be applied in order to compress them and use the resulting uncorrelated axes as input for the models to avoid redundant predictors [120].
Given that global RS data usually shows greater inaccuracies than local or regional datasets, the use of ensembles of different input data or models corresponding to the same variable might be advantageous, providing more accurate outputs as well as better conveying uncertainty [121,122,123,124,125]. Aside from that, many biophysical variables mapped at the local or regional scales are not available at a global scale, which might limit the relevance of global analyses for local scale management. Therefore, global characterization of PAs should be primarily aimed at informing larger scale conservation and management actions and plans, unless no better information is available at the local or regional scales.
Table 1 lists a set of recommended variables that can be used at a global scale for the biophysical characterization of terrestrial PAs. The list is not exhaustive, but it provides a wide range of relevant variables, including potential data sources. A more comprehensive list of potential variables can be found at the Global Climate Observing System Programme (https://public.wmo.int/en/programmes/global-climate-observing-system/essential-climate-variables; accessed 1 June 2021) or the Copernicus Global Land Service (http://land.copernicus.vgt.vito.be/PDF/portal/Application.html#Home; accessed 1 June 2021). A table with additional information, including the URLs of data sources, can be found as supplementary material. For MPAs, previous studies have highlighted candidate variables measurable by RS which are relevant to characterizing marine habitats [126,127,128,129,130]. They include, among others, bathymetry, the concentration of chlorophyll-a, sea surface temperature and sea surface salinity. A comprehensive list of these marine variables—together with access to the RS measurements of these variables—can be also found at the Copernicus Marine Service (https://resources.marine.copernicus.eu/?option=com_csw&task=results; accessed 1 June 2021) and the Living Wales Geoportal (https://wales.livingearth.online/data/environmental-variables/marine/; accessed 1 June 2021).

3.2. Global Environmental Stratifications

There are several biophysical characterizations available at the global scale, partially or totally based on RS data and modeling. Metzger et al. [137,138,139] used a broad set of bioclimatic variables to stratify the world in 18 environmental zones in order to support global ecosystem research and monitoring. Ivits et al. [53] mapped global ecosystem functional types using vegetation phenology and productivity variables by means of principal components and cluster analysis. Sayre [140] developed a map of global ecological land units using bioclimate, landforms, lithology and land cover variables. Tuanmu and Jetz [141] developed 14 remote sensing-based metrics to characterize habitat heterogeneity at a 1 km resolution at a global scale based on textural information extracted from the enhanced vegetation index (EVI) [142] and found out that bird species richness was strongly associated with habitat heterogeneity. Jung et al. [17] developed a global map of terrestrial habitat types following the IUCN habitat classification scheme (https://www.iucnredlist.org/resources/habitat-classification-scheme; accessed 1 June 2021) based on land cover, climate and land use data. Sayre et al. [51] developed a global classification of world climate regions and world ecosystems based on environmental descriptors, such as landforms, moisture, temperature, vegetation type and land use. Finally, in [143], a global ecosystem typology, including indicative distribution maps based on a large set of different environmental descriptors, existing global occurrence maps of specific ecosystem types and previous global environmental characterizations, was developed. They used a hierarchical classification system that first characterized ecosystems by their ecological functions and then distinguished ecosystems with contrasting species assemblages.
These global stratification initiatives are not limited to PAs and are indeed useful for prioritizing the conservation importance of larger regions. However, RS and modeling efforts specifically aimed to systematically characterize PAs could provide more relevant information needed to inform several policy initiatives as well as support management applications in PAs at regional or global scales, such as the assessment of ecological representativeness, the prioritization of PAs, connectivity assessments and the mapping of new areas requiring protection.

3.3. Global Characterization of Protected Areas

In relation to global characterizations within PAs by means of RS and modeling, in [144], the EODHaM system for characterizing habitats in PAs and surrounding areas was developed using earth observation data and expert knowledge. They used a semi-automated statistical procedure based on data related to the terrain, vegetation, water balance and land use. As part of the Digital Observatory for Protected Areas (DOPA) [145], in [120], PAs were systematically stratified globally into different habitat functional types based on remote sensing data and modeling, allowing for the quantification of the similarity between a reference area (representing a habitat functional type) and the surrounding areas based on a set of ecological indicators [146,147,148]. The method also graphically compares the ecological features of each habitat functional type found in a PA to help identify their main characteristics and understand the main biophysical gradients that occur at the PA level (Figure 2). The methodology uses a combination of several multivariate statistical analyses based on different global predictors that account for the climate, topography, vegetation and water exchanges. One of the advantages of this methodology is that the analysis is fully automated and can be performed at different spatial resolutions, which is especially important when dealing with smaller PAs. Furthermore, the similarity maps that are produced can also be used to identify new potential areas to be protected to strengthen ecological connectivity. When used in conjunction with forecasted bioclimatic data, the approach can further help identify new areas for conservation by considering current and climate change scenarios [147].
When prioritizing and ranking PAs, most studies have focused on species diversity to measure uniqueness [149,150]. However, biophysical characterizations have also been used along with biotic variables to perform gap and representation analyses in PAs [51,151]. Dubois et al. [147] proposed a methodology to assess the uniqueness of PAs based on biophysical variables. However, they lacked the means to decompose each analyzed area into areas with similar ecological features. The methodology proposed by [120] partially solved this issue by identifying habitat functional types and mapping similar areas at the ecoregion scale. This approach could be used to further create a composite indicator for each PA that reflects the biophysical richness of PAs and the uniqueness of their habitats. Coastal PAs in particular should be taken into account when developing indices of this type, given their inherent complexity as ecotones and the higher pressures they are exposed to because of human developments that are often concentrated along the coasts [152,153,154,155,156].
Perhaps the main limitation of global biophysical assessments using RS is the lack of ground truthing and comparison maps in order to evaluate results [157]. In this regard, resulting habitat and ecosystem types based on RS methods could be classified according to existing global typologies in order to serve and support different initiatives of habitat and ecosystem monitoring globally. For example, a hierarchical classification framework could be applied to the ecological features resulting from the methodology developed in [120], in which some key variables guide the first broad set of typologies and other variables help distinguish more specific subclasses according to existing typologies. The previously mentioned recent global environmental stratification initiatives already provide potential comparison maps, such as the IUCN Global Ecosystem Typology [158] and the set of world climate regions and world ecosystems [51]. The proposed approach would allow for taking similar regional features into consideration as well as going deeper into a specific global ecosystem type (e.g., tropical moist forests and mangroves).
In relation to the marine realm, current efforts to globally characterize PAs by means of RS have focused on the use of bathymetry. As such, DOPA uses a model of global bathymetry that is partially based on RS data to compute a marine habitat diversity index for MPAs [145]. The fact that (1) most RS methods can only derive information from the upper layer of the ocean (with the exception of altimeters for coarse scale bathymetry), (2) the spatial resolution of available RS data may be too coarse to characterize MPAs and (3) RS-based management of MPAs requires large financial and human resources constitutes major impediments to the use of RS data for characterizing MPAs [130]. This may explain why global characterization of MPAs using RS is limited. However, initiatives to characterize PAs using a broader set of RS-measured variables are more numerous at the regional [130,159,160] and local scales [161,162]. Beyond the characterization of MPAs, RS data have been used to assess the connectivity of MPA networks [154,163] and to delineate bioregions that can be further used as a basis to inform the design of MPA networks [164,165,166,167].

3.4. Computing Infrastructures

Computational capacity is another important limitation when characterizing PAs at the global scale. Most models and processing workflows developed so far are limited by the fact that there is no direct integration with external data sources and models, with most of them being standalone desktop or server applications. In this regard, large computational advances have occurred in recent years based on cloud-based infrastructures that support remote sensing data acquisition and processing [168]. Several tools have already been developed at a global scale to serve different purposes, such as the Global Surface Water Explorer (GSWE; https://global-surface-water.appspot.com/map; accessed 1 June 2021; [169]), the Map of Life (https://www.mol.org/; accessed 1 June 2021), the Global Forest Watch (https://www.globalforestwatch.org/map; accessed 1 June 2021), the remote sensing application for land cover classification and monitoring (https://remap-app.org/remap; accessed 1 June 2021), EarthMap (http://earthmap.org; accessed 1 June 2021) and the Living Atlas of the World (https://livingatlas.arcgis.com/en/home/; accessed 1 June 2021). Bastin et al. [170] used the GSWE to assess the level of protection of inland open surface waters and their trends within PAs globally.
Among other options, Google Earth Engine (GEE; [171]), ArcGIS Online (https://www.esri.com/en-us/arcgis/products/arcgis-online/overview; accessed 1 June 2021) and European Copernicus Data and Information Access Services (DIAS; https://www.copernicus.eu/en/access-data/dias; accessed 1 June 2021) offer data and services for cloud-based processing and remote sensing on large scales. Typical environmental applications include detecting deforestation, classifying land cover, estimating forest biomass and carbon or mapping the world’s roadless areas [172]. The advantage of using those services lies in the easy data access (including time series), the possibility to create graphical user interfaces and their remarkable computation speed, as processing is outsourced to cloud servers. Moreover, OpenEO (https://openeo.org/; accessed 1 June 2021) allows interoperability with big earth observation cloud backends for several programming languages.

4. Concluding Remarks and Recommendations

While the methods for mapping and assessing habitats and ecosystems are equally useful within and outside PAs, integrated assessment methods that systematically characterize and measure the diversity of habitats and ecosystems within a region are especially relevant when applied within PAs at a global scale. The global characterization of PAs can provide multiple benefits and applications, such as (1) supporting short-, medium- and long-term management actions, especially at the regional and global scales, that can ensure the maintenance of biodiversity and maximize the provision of ecosystem services [173,174], (2) evaluating the effects of climate change in PAs [175] and (3) informing policy initiatives, such as the European Biodiversity Strategy or the post-2020 Global Biodiversity Framework, on how to develop monitoring tools and indicators to promote sustainable management of PAs [176]. These kinds of analyses not only need to be done at a global scale, but also repeated if possible (e.g., annually) to document the changes that occur [177]. In this regard, the use of variables representing longer-term periods is also useful for capturing the presence of potential habitats and ecosystems, which can then be used as a reference for monitoring and condition assessment purposes. Furthermore, although locally derived variables are better descriptors of the ecosystems, global data sources are needed in order to systematically compare PAs across the globe and inform larger scale conservation actions.
In the last decade, cloud-based infrastructures have greatly improved the access to time series of relevant earth observation variables, which are crucial to the proper monitoring and assessment of ecosystems [178], bringing new opportunities for the global characterization of PAs. However, it is also necessary to translate the results from the global characterization of PAs into information that can be used in the real world, such as by sharing all data and models generated using online interoperable tools [179,180,181,182]. As an example of this, DOPA provides access to various global datasets and indicators that can inform decision-making and PA management [148], such as climate and topographic statistics, information about pressures, the occurrence of extreme events, land cover, land degradation and fragmentation, ecosystem services and species. Moreover, the Protected Planet website allows for exploring the World Database on Protected Areas (WDPA), maintained by the UN Environment Programme World Conservation Monitoring Centre (UNEP-WCMC). The CBD-mandated WDPA is the key reference dataset for any global protected area analysis, and it includes both spatial (mapped boundary or point location) and nonspatial (e.g., name, type, size, age and status) information for over 230,000 protected areas worldwide [183]. Despite accelerated efforts to improve the global PA data, the quality of the WDPA data still varies greatly between countries and regions, and this should be acknowledged in any analysis using the WDPA. Only limited information related to the systematic global biophysical characterization of PAs can be found online thus far, such as the terrestrial habitat diversity index in DOPA [145].
Systematic information related to the uniqueness or the importance of PAs based on biophysical variables could, among other things, further support the ranking and prioritization of PAs based on the diversity of their habitats and ecosystems. Biophysical studies also allow us to study the role of habitats and ecosystems in maintaining biodiversity in the context of climate change, since species populations can adapt to changes by moving to new areas that meet their ecological requirements [146]. Several applications of habitat models have shown a high correlation between biodiversity and the diversity of habitat types, and they can help identify potential new areas that should be protected in order to maintain species protection into the future [120,184,185].
Table 2 gives an overview of applications of different environmental descriptors, including methods and data, which are relevant for the biophysical characterization of PAs, highlighting the importance of taking into account structural and functional attributes as well as integrating a broad spectrum of environmental descriptors in the global biophysical characterization of PAs.
Finally, we give a summary of the recommendations proposed to improve the global biophysical characterization of PAs in relation to different aspects, which are listed below.
Environmental attributes and descriptors:
  • The structural and functional attributes of ecosystems and habitats within PAs should be addressed;
  • A broad set of variables representative of key biophysical quantitative descriptors should be used to produce integrated assessments, potentially including vegetation, energy, climate, water, terrain and soil.
Data sources and processing:
  • Global data sources presenting time series and regular updates should be preferred;
  • Dimensionality reduction techniques are often used to deal with correlated input variables;
  • The use of ensembles of different input data or models corresponding to the same variable is recommended to provide more accurate outputs and deal with uncertainty.
Methods:
  • The use of interoperable RS cloud-based infrastructures is recommended for large-scale processing;
  • Analyses should be regularly repeated to document changes;
  • The analysis should extend beyond specific habitat or ecosystem mapping and assessment methods so that a variety of habitats and ecosystem types can be identified;
  • The resulting habitat and ecosystem types within PAs should be, to the greatest extent possible, comparable with existing global typologies;
  • There is a clear need and potential to develop methodologies for assessing the biophysical uniqueness of PAs that could support prioritization analyses;
  • Methods should allow the prediction of climate change impacts on ecosystems by using forecasted bioclimatic data.
Application in policy and practice:
  • Translate the results into information that can be used by policy and decision-makers;
  • Ensure transparency and reproducibility by sharing all data and models generated using online interoperable tools;
  • Global characterization of PAs should be specifically aimed at informing larger scale conservation and management actions and plans, unless no better information is available at the local or regional scales.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijgi10060384/s1. Table S1: Supporting information for Table 1 including URLs with relevant data sources.

Author Contributions

Writing—original draft, Javier Martínez-López; writing—review and editing, Bastian Bertzky, Simon Willcock, Marine Robuchon, María Almagro, Giacomo Delli and Grégoire Dubois. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by a European Commission Expert Contract (CT-EX2020D381533-101). The development and maintenance of the Digital Observatory for Protected Areas are supported mainly by the institutional activities of Directorate D (Sustainable Resources) at the Joint Research Centre of the European Commission as well as by the Biodiversity and Protected Areas Management (BIOPAMA) EU-ACP programme, an initiative of the African, Caribbean and Pacific (ACP) Group of States financed by the 10th and 11th European Development Funds of the European Union (EU).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable. No new data were created or analyzed in this study.

Acknowledgments

The authors want to thank three anonymous reviewers and the editor for their constructive comments on an earlier version of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maxwell, S.L.; Cazalis, V.; Dudley, N.; Hoffmann, M.; Rodrigues, A.S.L.; Stolton, S.; Visconti, P.; Woodley, S.; Kingston, N.; Lewis, E.; et al. Area-based conservation in the twenty-first century. Nature 2020, 586, 217–227. [Google Scholar] [CrossRef]
  2. Mappin, B.; Chauvenet, A.L.M.; Adams, V.M.; Di Marco, M.; Beyer, H.L.; Venter, O.; Halpern, B.S.; Possingham, H.P.; Watson, J.E.M. Restoration priorities to achieve the global protected area target. Conserv. Lett. 2019, e12646. [Google Scholar] [CrossRef] [Green Version]
  3. Thomas, C.D.; Gillingham, P.K. The performance of protected areas for biodiversity under climate change. Biol. J. Linn. Soc. Lond. 2015, 115, 718–730. [Google Scholar] [CrossRef]
  4. Elsen, P.R.; Monahan, W.B.; Merenlender, A.M. Global patterns of protection of elevational gradients in mountain ranges. Proc. Natl. Acad. Sci. USA 2018, 115, 6004–6009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Pascual, U.; Balvanera, P.; Díaz, S.; Pataki, G.; Roth, E.; Stenseke, M.; Watson, R.T.; Başak Dessane, E.; Islar, M.; Kelemen, E.; et al. Valuing nature’s contributions to people: The IPBES approach. Curr. Opin. Environ. Sustain. 2017, 26–27, 7–16. [Google Scholar] [CrossRef] [Green Version]
  6. Dinerstein, E.; Olson, D.; Joshi, A.; Vynne, C.; Burgess, N.D.; Wikramanayake, E.; Hahn, N.; Palminteri, S.; Hedao, P.; Noss, R.; et al. An Ecoregion-Based Approach to Protecting Half the Terrestrial Realm. Bioscience 2017, 67, 534–545. [Google Scholar] [CrossRef]
  7. Grêt-Regamey, A.; Weibel, B. Global assessment of mountain ecosystem services using earth observation data. Ecosyst. Serv. 2020, 46, 101213. [Google Scholar] [CrossRef]
  8. Cooper, G.S.; Willcock, S.; Dearing, J.A. Regime shifts occur disproportionately faster in larger ecosystems. Nat. Commun. 2020, 11, 1175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. EC. EU Biodiversity Strategy for 2030 Bringing Nature Back into Our Lives; (COM/2020/380 final); Official Journal of the European Union: Brussels, Belgium, 2020. [Google Scholar]
  10. Saarikoski, H.; Mustajoki, J.; Barton, D.N.; Geneletti, D.; Langemeyer, J.; Gomez-Baggethun, E.; Marttunen, M.; Antunes, P.; Keune, H.; Santos, R. Multi-Criteria Decision Analysis and Cost-Benefit Analysis: Comparing alternative frameworks for integrated valuation of ecosystem services. Ecosyst. Serv. 2016, 22, 238–249. [Google Scholar] [CrossRef]
  11. Belle, E.; Kingston, N.; Burgess, N.; Sandwith, T.; Ali, N.; MacKinnon, K. Protected Planet Report 2018: Tracking Progress Towards Global Targets for Protected Areas; UNEP-WCMC, IUCN and NGS: Cambridge, UK; Gland, Switzerland; Washington, DC, USA, 2018. [Google Scholar]
  12. Bonet-García, F.J.; Pérez-Luque, A.J.; Moreno-Llorca, R.A.; Pérez-Pérez, R.; Puerta-Piñero, C.; Zamora, R. Protected areas as elicitors of human well-being in a developed region: A new synthetic (socioeconomic) approach. Biol. Conserv. 2015, 187, 221–229. [Google Scholar] [CrossRef]
  13. Moreno-Llorca, R.; Vaz, A.S.; Herrero, J.; Millares, A.; Bonet-García, F.J.; Alcaraz-Segura, D. Multi-scale evolution of ecosystem services’ supply in Sierra Nevada (Spain): An assessment over the last half-century. Ecosyst. Serv. 2020, 46, 101204. [Google Scholar] [CrossRef]
  14. Naidoo, R.; Gerkey, D.; Hole, D.; Pfaff, A.; Ellis, A.M.; Golden, C.D.; Herrera, D.; Johnson, K.; Mulligan, M.; Ricketts, T.H.; et al. Evaluating the impacts of protected areas on human well-being across the developing world. Sci. Adv. 2019, 5, eaav3006. [Google Scholar] [CrossRef] [Green Version]
  15. Bunce, R.G.H.; Bogers, M.M.B.; Evans, D.; Halada, L.; Jongman, R.H.G.; Mucher, C.A.; Bauch, B.; de Blust, G.; Parr, T.W.; Olsvig-Whittaker, L. The significance of habitats as indicators of biodiversity and their links to species. Ecol. Indic. 2013, 33, 19–25. [Google Scholar] [CrossRef]
  16. Reid, W.V.; Mooney, H.A.; Cropper, A.; Capistrano, D.; Carpenter, S.R.; Chopra, K.; Dasgupta, P.; Dietz, T.; Duraiappah, A.K.; Hassan, R.; et al. Millennium Ecosystem Assessment. Ecosystems and Human Well-Being: Synthesis; Island Press: Washington, DC, USA, 2005. [Google Scholar]
  17. Jung, M.; Dahal, P.R.; Butchart, S.H.M.; Donald, P.F.; De Lamo, X.; Lesiv, M.; Kapos, V.; Rondinini, C.; Visconti, P. A global map of terrestrial habitat types. Sci. Data 2020, 7, 256. [Google Scholar] [CrossRef]
  18. Pereira, H.M.; Ferrier, S.; Walters, M.; Geller, G.N.; Jongman, R.H.G.; Scholes, R.J.; Bruford, M.W.; Brummitt, N.; Butchart, S.H.M.; Cardoso, A.C.; et al. Essential Biodiversity Variables. Science 2013, 339, 277–278. [Google Scholar] [CrossRef] [Green Version]
  19. Mairota, P.; Cafarelli, B.; Boccaccio, L.; Leronni, V.; Labadessa, R.; Kosmidou, V.; Nagendra, H. Using landscape structure to develop quantitative baselines for protected area monitoring. Ecol. Indic. 2013, 33, 82–95. [Google Scholar] [CrossRef]
  20. Dash, J.; Ogutu, B.O. Recent advances in space-borne optical remote sensing systems for monitoring global terrestrial ecosystems. Prog. Phys. Geogr. 2016, 40, 322–351. [Google Scholar] [CrossRef]
  21. Buchanan, G.M.; Beresford, A.E.; Hebblewhite, M.; Escobedo, F.J.; De Klerk, H.M.; Donald, P.F.; Escribano, P.; Koh, L.P.; Martínez-López, J.; Pettorelli, N.; et al. Free satellite data key to conservation. Science 2018, 361, 139–140. [Google Scholar] [CrossRef]
  22. Rose, R.A.; Byler, D.; Eastman, J.R.; Fleishman, E.; Geller, G.; Goetz, S.; Guild, L.; Hamilton, H.; Hansen, M.; Headley, R.; et al. Ten ways remote sensing can contribute to conservation. Conserv. Biol. 2015, 29, 350–359. [Google Scholar] [CrossRef] [Green Version]
  23. O’Connor, B.; Moul, K.; Pollini, B.; de Lamo, X.; Simonson, W.; Allison, H.; Albrecht, F.; Guzinski, R.M.; Larsen, H.; McGlade, J.; et al. Earth Observation for SDG—Compendium of Earth Observation Contributions to the SDG Targets and Indicators; European Space Agency: Paris, France, 2020; p. 165. [Google Scholar]
  24. Petrou, Z.I.; Manakos, I.; Stathaki, T. Remote sensing for biodiversity monitoring: A review of methods for biodiversity indicator extraction and assessment of progress towards international targets. Biodivers. Conserv. 2015, 24, 2333–2363. [Google Scholar] [CrossRef]
  25. Latombe, G.; Pyšek, P.; Jeschke, J.M.; Blackburn, T.M.; Bacher, S.; Capinha, C.; Costello, M.J.; Fernández, M.; Gregory, R.D.; Hobern, D.; et al. A vision for global monitoring of biological invasions. Biol. Conserv. 2017, 213, 295–308. [Google Scholar] [CrossRef]
  26. De Araujo Barbosa, C.C.; Atkinson, P.M.; Dearing, J.A. Remote sensing of ecosystem services: A systematic review. Ecol. Indic. 2015, 52, 430–443. [Google Scholar] [CrossRef]
  27. Dubovyk, O. The role of Remote Sensing in land degradation assessments: Opportunities and challenges. Eur. J. Remote Sens. 2017, 50, 601–613. [Google Scholar] [CrossRef]
  28. Wang, Y.; Yésou, H. Remote sensing of floodpath lakes and wetlands: A challenging frontier in the monitoring of changing environments. Remote Sens. 2018, 10, 1955. [Google Scholar] [CrossRef] [Green Version]
  29. Wang, Y.; Lu, Z.; Sheng, Y.; Zhou, Y. Remote sensing applications in monitoring of protected areas. Remote Sens. 2020, 12, 1370. [Google Scholar] [CrossRef]
  30. Mao, L.; Li, M.; Shen, W. Remote sensing applications for monitoring terrestrial protected areas: Progress in the last decade. Sustainability 2020, 12, 5016. [Google Scholar] [CrossRef]
  31. Pettorelli, N.; Wegmann, M.; Gurney, L.; Dubois, G. Monitoring Protected Areas from Space. In Protected Areas: Are They Safeguarding Biodiversity? Joppa, L.N., Baillie, J.E.M., Robinson, J.G., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2016; pp. 242–259. [Google Scholar]
  32. Gillespie, T.W.; Willis, K.S.; Ostermann-Kelm, S. Spaceborne remote sensing of the world’s protected areas. Prog. Phys. Geogr. Earth Environ. 2015, 39, 388–404. [Google Scholar] [CrossRef]
  33. Duan, P.; Wang, Y.; Yin, P. Remote sensing applications in monitoring of protected areas: A bibliometric analysis. Remote Sens. 2020, 12, 772. [Google Scholar] [CrossRef] [Green Version]
  34. Tsyganskaya, V.; Martinis, S.; Marzahn, P. Flood Monitoring in Vegetated Areas Using Multitemporal Sentinel-1 Data: Impact of Time Series Features. Water 2019, 11, 1938. [Google Scholar] [CrossRef] [Green Version]
  35. Liu, C.-C.; Shieh, M.-C.; Ke, M.-S.; Wang, K.-H. Flood prevention and emergency response system powered by google earth engine. Remote Sens. 2018, 10, 1283. [Google Scholar] [CrossRef] [Green Version]
  36. Nemani, R.; Hashimoto, H.; Votava, P.; Melton, F.; Wang, W.; Michaelis, A.; Mutch, L.; Milesi, C.; Hiatt, S.; White, M. Monitoring and forecasting ecosystem dynamics using the Terrestrial Observation and Prediction System (TOPS). Remote Sens. Environ. 2009, 113, 1497–1509. [Google Scholar] [CrossRef]
  37. Wiens, J.; Sutter, R.; Anderson, M.; Blanchard, J.; Barnett, A.; Aguilar-Amuchastegui, N.; Avery, C.; Laine, S. Selecting and conserving lands for biodiversity: The role of remote sensing. Remote Sens. Environ. 2009, 113, 1370–1381. [Google Scholar] [CrossRef] [Green Version]
  38. Wilson, M.C.; Chen, X.-Y.; Corlett, R.T.; Didham, R.K.; Ding, P.; Holt, R.D.; Holyoak, M.; Hu, G.; Hughes, A.C.; Jiang, L.; et al. Habitat fragmentation and biodiversity conservation: Key findings and future challenges. Landsc. Ecol. 2016, 31, 219–227. [Google Scholar] [CrossRef] [Green Version]
  39. Wang, R.; Gamon, J.A. Remote sensing of terrestrial plant biodiversity. Remote Sens. Environ. 2019, 231, 111218. [Google Scholar] [CrossRef]
  40. Jetz, W.; Cavender-Bares, J.; Pavlick, R.; Schimel, D.; Davis, F.W.; Asner, G.P.; Guralnick, R.; Kattge, J.; Latimer, A.M.; Moorcroft, P.; et al. Monitoring plant functional diversity from space. Nat. Plants 2016, 2, 16024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Rolf, W.; Lenz, R.; Peters, D. Development of a quantitative “bioassay” approach for ecosystem mapping. Int. J. Biodivers. Sci. Ecosyst. Serv. Manag. 2012, 8, 71–79. [Google Scholar] [CrossRef]
  42. Mücher, C.A.; Klijn, J.A.; Wascher, D.M.; Schaminée, J.H.J. A new European Landscape Classification (LANMAP): A transparent, flexible and user-oriented methodology to distinguish landscapes. Ecol. Indic. 2010, 10, 87–103. [Google Scholar] [CrossRef]
  43. Hargrove, W.W.; Hoffman, F.M. Potential of multivariate quantitative methods for delineation and visualization of ecoregions. Environ. Manag. 2004, 34 (Suppl. 1), S39–S60. [Google Scholar] [CrossRef]
  44. Metzger, M.J.; Bunce, R.G.H.; Jongman, R.H.G.; Mücher, C.A.; Watkins, J.W. A climatic stratification of the environment of Europe. Glob. Ecol. Biogeogr. 2005, 14, 549–563. [Google Scholar] [CrossRef]
  45. Sayre, R.; Comer, P.; Warner, H.; Cress, J. A New Map of Standardized Terrestrial Ecosystems of the Conterminous United States; U.S. Geological Survey Professional Paper; U.S. Geological Survey: Reston, VA, USA, 2009; p. 17. [Google Scholar]
  46. Felipe-Lucia, M.R.; Soliveres, S.; Penone, C.; Fischer, M.; Ammer, C.; Boch, S.; Boeddinghaus, R.S.; Bonkowski, M.; Buscot, F.; Fiore-Donno, A.M.; et al. Land-use intensity alters networks between biodiversity, ecosystem functions, and services. Proc. Natl. Acad. Sci. USA 2020, 117, 28140–28149. [Google Scholar] [CrossRef]
  47. Pettorelli, N.; Owen, H.J.F.; Duncan, C. How do we want Satellite Remote Sensing to support biodiversity conservation globally? Methods Ecol. Evol. 2016, 7, 656–665. [Google Scholar] [CrossRef]
  48. Skidmore, A.K.; Pettorelli, N.; Coops, N.C.; Geller, G.N.; Hansen, M.; Lucas, R.; Mücher, C.A.; O’Connor, B.; Paganini, M.; Pereira, H.M.; et al. Environmental science: Agree on biodiversity metrics to track from space. Nature 2015, 523, 403–405. [Google Scholar] [CrossRef] [Green Version]
  49. Vihervaara, P.; Mononen, L.; Auvinen, A.-P.; Virkkala, R.; Lü, Y.; Pippuri, I.; Packalen, P.; Valbuena, R.; Valkama, J. How to integrate remotely sensed data and biodiversity for ecosystem assessments at landscape scale. Landsc. Ecol. 2015, 30, 501–516. [Google Scholar] [CrossRef]
  50. Olson, D.M.; Dinerstein, E.; Wikramanayake, E.D.; Burgess, N.D.; Powell, G.V.N.; Underwood, E.C.; D’amico, J.A.; Itoua, I.; Strand, H.E.; Morrison, J.C.; et al. Terrestrial Ecoregions of the World: A New Map of Life on Earth: A new global map of terrestrial ecoregions provides an innovative tool for conserving biodiversity. BioScience 2001, 51, 933–938. [Google Scholar] [CrossRef]
  51. Sayre, R.; Karagulle, D.; Frye, C.; Boucher, T.; Wolff, N.H.; Breyer, S.; Wright, D.; Martin, M.; Butler, K.; Van Graafeiland, K.; et al. An assessment of the representation of ecosystems in global protected areas using new maps of World Climate Regions and World Ecosystems. Glob. Ecol. Conserv. 2020, 21, e00860. [Google Scholar] [CrossRef]
  52. Pérez-Hoyos, A.; Martínez, B.; García-Haro, F.; Moreno, Á.; Gilabert, M. Identification of Ecosystem Functional Types from Coarse Resolution Imagery Using a Self-Organizing Map Approach: A Case Study for Spain. Remote Sens. 2014, 6, 11391–11419. [Google Scholar] [CrossRef] [Green Version]
  53. Ivits, E.; Cherlet, M.; Horion, S.; Fensholt, R. Global biogeographical pattern of ecosystem functional types derived from earth observation data. Remote Sens. 2013, 5, 3305–3330. [Google Scholar] [CrossRef] [Green Version]
  54. Bastos, R.; Monteiro, A.T.; Carvalho, D.; Gomes, C.; Travassos, P.; Honrado, J.P.; Santos, M.; Cabral, J.A. Integrating land cover structure and functioning to predict biodiversity patterns: A hierarchical modelling framework designed for ecosystem management. Landsc. Ecol. 2016, 31, 701–710. [Google Scholar] [CrossRef]
  55. Nagendra, H.; Lucas, R.; Honrado, J.P.; Jongman, R.H.G.; Tarantino, C.; Adamo, M.; Mairota, P. Remote sensing for conservation monitoring: Assessing protected areas, habitat extent, habitat condition, species diversity, and threats. Ecol. Indic. 2013, 33, 45–59. [Google Scholar] [CrossRef]
  56. Corbane, C.; Lang, S.; Pipkins, K.; Alleaume, S.; Deshayes, M.; García Millán, V.E.; Strasser, T.; Vanden Borre, J.; Toon, S.; Michael, F. Remote sensing for mapping natural habitats and their conservation status—New opportunities and challenges. Int. J. Appl. Earth Obs. Geoinf. 2015, 37, 7–16. [Google Scholar] [CrossRef]
  57. Villoslada, M.; Bunce, R.G.H.; Sepp, K.; Jongman, R.H.G.; Metzger, M.J.; Kull, T.; Raet, J.; Kuusemets, V.; Kull, A.; Leito, A. A framework for habitat monitoring and climate change modelling: Construction and validation of the Environmental Stratification of Estonia. Reg. Environ. Chang. 2017, 17, 335–349. [Google Scholar] [CrossRef] [Green Version]
  58. Jongman, R.H.G.; Mücher, C.A.; Bunce, R.G.H.; Lang, M.; Sepp, K. A Review of Approaches for Automated Habitat Mapping and their Potential Added Value for Biodiversity Monitoring Projects. J. Landsc. Ecol. 2019, 12, 53–69. [Google Scholar] [CrossRef] [Green Version]
  59. Lang, M.; Vain, A.; Bunce, R.G.H.; Jongman, R.H.G.; Raet, J.; Sepp, K.; Kuusemets, V.; Kikas, T.; Liba, N. Extrapolation of in situ data from 1-km squares to adjacent squares using remote sensed imagery and airborne lidar data for the assessment of habitat diversity and extent. Environ. Monit. Assess. 2015, 187, 76. [Google Scholar] [CrossRef] [PubMed]
  60. Vaz, A.S.; Marcos, B.; Gonçalves, J.; Monteiro, A.; Alves, P.; Civantos, E.; Lucas, R.; Mairota, P.; Garcia-Robles, J.; Alonso, J.; et al. Can we predict habitat quality from space? A multi-indicator assessment based on an automated knowledge-driven system. Int. J. Appl. Earth Obs. Geoinf. 2015, 37, 106–113. [Google Scholar] [CrossRef]
  61. Mairota, P.; Cafarelli, B.; Didham, R.K.; Lovergine, F.P.; Lucas, R.M.; Nagendra, H.; Rocchini, D.; Tarantino, C. Challenges and opportunities in harnessing satellite remote-sensing for biodiversity monitoring. Ecol. Inform. 2015, 30, 207–214. [Google Scholar] [CrossRef]
  62. Pettorelli, N.; Schulte to Bühne, H.; Tulloch, A.; Dubois, G.; Macinnis-Ng, C.; Queirós, A.M.; Keith, D.A.; Wegmann, M.; Schrodt, F.; Stellmes, M.; et al. Satellite remote sensing of ecosystem functions: Opportunities, challenges and way forward. Remote Sens. Ecol. Conserv. 2017, 4, 1–23. [Google Scholar] [CrossRef]
  63. Chatziantoniou, A.; Psomiadis, E.; Petropoulos, G. Co-Orbital Sentinel 1 and 2 for LULC Mapping with Emphasis on Wetlands in a Mediterranean Setting Based on Machine Learning. Remote Sens. 2017, 9, 1259. [Google Scholar] [CrossRef] [Green Version]
  64. Chavez, L.J. Identifying Dune Habitat through the Use of Remote Sensing Classifications. Ph.D. Thesis, Texas State University, San Marcos, TA, USA, 2019. [Google Scholar]
  65. Mao, D.; Wang, Z.; Du, B.; Li, L.; Tian, Y.; Jia, M.; Zeng, Y.; Song, K.; Jiang, M.; Wang, Y. National wetland mapping in China: A new product resulting from object-based and hierarchical classification of Landsat 8 OLI images. ISPRS J. Photogramm. Remote Sens. 2020, 164, 11–25. [Google Scholar] [CrossRef]
  66. Campbell, A.; Wang, Y. High spatial resolution remote sensing for salt marsh mapping and change analysis at fire island national seashore. Remote Sens. 2019, 11, 1107. [Google Scholar] [CrossRef] [Green Version]
  67. Szantoi, Z.; Escobedo, F.J.; Abd-Elrahman, A.; Pearlstine, L.; Dewitt, B.; Smith, S. Classifying spatially heterogeneous wetland communities using machine learning algorithms and spectral and textural features. Environ. Monit. Assess. 2015, 187, 262. [Google Scholar] [CrossRef] [PubMed]
  68. Kollár, S.; Vekerdy, Z.; Márkus, B. Forest habitat change dynamics in a riparian wetland. Procedia Environ. Sci. 2011, 7, 371–376. [Google Scholar] [CrossRef] [Green Version]
  69. Szantoi, Z.; Escobedo, F.; Abd-Elrahman, A.; Smith, S.; Pearlstine, L. Analyzing fine-scale wetland composition using high resolution imagery and texture features. Int. J. Appl. Earth Obs. Geoinf. 2013, 23, 204–212. [Google Scholar] [CrossRef]
  70. Lane, C.; Liu, H.; Autrey, B.; Anenkhonov, O.; Chepinoga, V.; Wu, Q. Improved Wetland Classification Using Eight-Band High Resolution Satellite Imagery and a Hybrid Approach. Remote Sens. 2014, 6, 12187–12216. [Google Scholar] [CrossRef] [Green Version]
  71. Strasser, T.; Lang, S. Object-based class modelling for multi-scale riparian forest habitat mapping. Int. J. Appl. Earth Obs. Geoinf. 2015, 37, 29–37. [Google Scholar] [CrossRef]
  72. Johansen, K.; Coops, N.C.; Gergel, S.E.; Stange, Y. Application of high spatial resolution satellite imagery for riparian and forest ecosystem classification. Remote Sens. Environ. 2007, 110, 29–44. [Google Scholar] [CrossRef]
  73. Wendelberger, K.S.; Gann, D.; Richards, J.H. Using Bi-Seasonal WorldView-2 Multi-Spectral Data and Supervised Random Forest Classification to Map Coastal Plant Communities in Everglades National Park. Sensors 2018, 18, 829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Rapinel, S.; Mony, C.; Lecoq, L.; Clément, B.; Thomas, A.; Hubert-Moy, L. Evaluation of Sentinel-2 time-series for mapping floodplain grassland plant communities. Remote Sens. Environ. 2019, 223, 115–129. [Google Scholar] [CrossRef]
  75. Zhang, L.; Li, X.; Lu, S.; Jia, K. Multi-scale object-based measurement of arid plant community structure. Int. J. Remote Sens. 2016, 37, 2168–2179. [Google Scholar] [CrossRef]
  76. Silveyra Gonzalez, R.; Latifi, H.; Weinacker, H.; Dees, M.; Koch, B.; Heurich, M. Integrating LiDAR and high-resolution imagery for object-based mapping of forest habitats in a heterogeneous temperate forest landscape. Int. J. Remote Sens. 2018, 39, 8859–8884. [Google Scholar] [CrossRef]
  77. Stabach, J.A.; Dabek, L.; Jensen, R.; Wang, Y.Q. Discrimination of dominant forest types for Matschie’s tree kangaroo conservation in Papua New Guinea using high-resolution remote sensing data. Int. J. Remote Sens. 2009, 30, 405–422. [Google Scholar] [CrossRef]
  78. Marselis, S.M.; Abernethy, K.; Alonso, A.; Armston, J.; Baker, T.R.; Bastin, J.; Bogaert, J.; Boyd, D.S.; Boeckx, P.; Burslem, D.F.R.P.; et al. Evaluating the potential of full-waveform lidar for mapping pan-tropical tree species richness. Glob. Ecol. Biogeogr. 2020. [Google Scholar] [CrossRef]
  79. Jiménez López, J.; Mulero-Pázmány, M. Drones for conservation in protected areas: Present and future. Drones 2019, 3, 10. [Google Scholar] [CrossRef] [Green Version]
  80. Onojeghuo, A.O.; Blackburn, G.A. Optimising the use of hyperspectral and LiDAR data for mapping reedbed habitats. Remote Sens. Environ. 2011, 115, 2025–2034. [Google Scholar] [CrossRef]
  81. Guo, X.; Coops, N.C.; Tompalski, P.; Nielsen, S.E.; Bater, C.W.; John Stadt, J. Regional mapping of vegetation structure for biodiversity monitoring using airborne lidar data. Ecol. Inform. 2017, 38, 50–61. [Google Scholar] [CrossRef]
  82. Cheţan, M.A.; Dornik, A. 20 years of landscape dynamics within the world’s largest multinational network of protected areas. J. Environ. Manag. 2020, 111712. [Google Scholar] [CrossRef]
  83. Mairota, P.; Cafarelli, B.; Labadessa, R.; Lovergine, F.; Tarantino, C.; Lucas, R.M.; Nagendra, H.; Didham, R.K. Very high resolution Earth observation features for monitoring plant and animal community structure across multiple spatial scales in protected areas. Int. J. Appl. Earth Obs. Geoinf. 2015, 37, 100–105. [Google Scholar] [CrossRef]
  84. Park, Y.; Guldmann, J.-M. Measuring continuous landscape patterns with Gray-Level Co-Occurrence Matrix (GLCM) indices: An alternative to patch metrics? Ecol. Indic. 2020, 109, 105802. [Google Scholar] [CrossRef]
  85. Ozdemir, I.; Mert, A.; Ozkan, U.Y.; Aksan, S.; Unal, Y. Predicting bird species richness and micro-habitat diversity using satellite data. For. Ecol. Manag. 2018, 424, 483–493. [Google Scholar] [CrossRef]
  86. St-Louis, V.; Pidgeon, A.M.; Clayton, M.K.; Locke, B.A.; Bash, D.; Radeloff, V.C. Satellite image texture and a vegetation index predict avian biodiversity in the Chihuahuan Desert of New Mexico. Ecography 2009, 32, 468–480. [Google Scholar] [CrossRef]
  87. Wood, E.M.; Pidgeon, A.M.; Radeloff, V.C.; Keuler, N.S. Image texture as a remotely sensed measure of vegetation structure. Remote Sens. Environ. 2012, 121, 516–526. [Google Scholar] [CrossRef]
  88. Farwell, L.S.; Elsen, P.R.; Razenkova, E.; Pidgeon, A.M.; Radeloff, V.C. Habitat heterogeneity captured by 30-m resolution satellite image texture predicts bird richness across the United States. Ecol. Appl. 2020, 30, e02157. [Google Scholar] [CrossRef] [PubMed]
  89. Farwell, L.S.; Gudex-Cross, D.; Anise, I.E.; Bosch, M.J.; Olah, A.M.; Radeloff, V.C.; Razenkova, E.; Rogova, N.; Silveira, E.M.O.; Smith, M.M.; et al. Satellite image texture captures vegetation heterogeneity and explains patterns of bird richness. Remote Sens. Environ. 2021, 253, 112175. [Google Scholar] [CrossRef]
  90. Hobi, M.L.; Dubinin, M.; Graham, C.H.; Coops, N.C.; Clayton, M.K.; Pidgeon, A.M.; Radeloff, V.C. A comparison of Dynamic Habitat Indices derived from different MODIS products as predictors of avian species richness. Remote Sens. Environ. 2017, 195, 142–152. [Google Scholar] [CrossRef]
  91. Berry, S.; Mackey, B.; Brown, T. Potential applications of remotely sensed vegetation greenness to habitat analysis and the conservation of dispersive fauna. Pac. Conserv. Biol. 2007, 13, 120. [Google Scholar] [CrossRef]
  92. Madonsela, S.; Cho, M.A.; Ramoelo, A.; Mutanga, O. Remote sensing of species diversity using Landsat 8 spectral variables. ISPRS J. Photogramm. Remote Sens. 2017, 133, 116–127. [Google Scholar] [CrossRef] [Green Version]
  93. Ribeiro, I.; Proença, V.; Serra, P.; Palma, J.; Domingo-Marimon, C.; Pons, X.; Domingos, T. Remotely sensed indicators and open-access biodiversity data to assess bird diversity patterns in Mediterranean rural landscapes. Sci. Rep. 2019, 9, 6826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Chu, T.; Guo, X.; Takeda, K. Remote sensing approach to detect post-fire vegetation regrowth in Siberian boreal larch forest. Ecol. Indic. 2016, 62, 32–46. [Google Scholar] [CrossRef]
  95. Fernandez-Manso, A.; Quintano, C.; Roberts, D.A. Burn severity influence on post-fire vegetation cover resilience from Landsat MESMA fraction images time series in Mediterranean forest ecosystems. Remote Sens. Environ. 2016, 184, 112–123. [Google Scholar] [CrossRef]
  96. Alcaraz, D.; Paruelo, J.; Cabello, J. Identification of current ecosystem functional types in the Iberian Peninsula. Glob. Ecol. Biogeogr. 2006, 15, 200–212. [Google Scholar] [CrossRef]
  97. Schirpke, U.; Leitinger, G.; Tasser, E.; Rüdisser, J.; Fontana, V.; Tappeiner, U. Functional spatial units are fundamental for modelling ecosystem services in mountain regions. Appl. Geogr. 2020, 118, 102200. [Google Scholar] [CrossRef]
  98. Keith, D.A.; Rodríguez, J.P.; Rodríguez-Clark, K.M.; Nicholson, E.; Aapala, K.; Alonso, A.; Asmussen, M.; Bachman, S.; Basset, A.; Barrow, E.G.; et al. Scientific foundations for an IUCN Red List of ecosystems. PLoS ONE 2013, 8, e62111. [Google Scholar] [CrossRef] [Green Version]
  99. Xie, Y.; Sha, Z.; Yu, M. Remote sensing imagery in vegetation mapping: A review. J. Plant Ecol. 2008, 1, 9–23. [Google Scholar] [CrossRef]
  100. Hijmans, R.J.; Cameron, S.E.; Parra, J.L.; Jones, P.G.; Jarvis, A. Very high resolution interpolated climate surfaces for global land areas. Int. J. Climatol. 2005, 25, 1965–1978. [Google Scholar] [CrossRef]
  101. Fick, S.E.; Hijmans, R.J. WorldClim 2: New 1-km spatial resolution climate surfaces for global land areas. Int. J. Climatol. 2017, 37, 4302–4315. [Google Scholar] [CrossRef]
  102. Holdridge, L.R. Life Zone Ecology; Tropical Science Center: San Jose, CA, USA, 1967; p. 206. [Google Scholar]
  103. Beck, H.E.; Zimmermann, N.E.; McVicar, T.R.; Vergopolan, N.; Berg, A.; Wood, E.F. Present and future Köppen-Geiger climate classification maps at 1-km resolution. Sci. Data 2018, 5, 180214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Amatulli, G.; Domisch, S.; Tuanmu, M.-N.; Parmentier, B.; Ranipeta, A.; Malczyk, J.; Jetz, W. A suite of global, cross-scale topographic variables for environmental and biodiversity modeling. Sci. Data 2018, 5, 180040. [Google Scholar] [CrossRef] [Green Version]
  105. Körner, C.; Paulsen, J.; Spehn, E.M. A definition of mountains and their bioclimatic belts for global comparisons of biodiversity data. Alp. Bot. 2011, 121, 73–78. [Google Scholar] [CrossRef] [Green Version]
  106. Linke, S.; Lehner, B.; Ouellet Dallaire, C.; Ariwi, J.; Grill, G.; Anand, M.; Beames, P.; Burchard-Levine, V.; Maxwell, S.; Moidu, H.; et al. Global hydro-environmental sub-basin and river reach characteristics at high spatial resolution. Sci. Data 2019, 6, 283. [Google Scholar] [CrossRef] [Green Version]
  107. Gao, H. Satellite remote sensing of large lakes and reservoirs: From elevation and area to storage. WIREs Water 2015, 2, 147–157. [Google Scholar] [CrossRef]
  108. Pickens, A.H.; Hansen, M.C.; Hancher, M.; Stehman, S.V.; Tyukavina, A.; Potapov, P.; Marroquin, B.; Sherani, Z. Mapping and sampling to characterize global inland water dynamics from 1999 to 2018 with full Landsat time-series. Remote Sens. Environ. 2020, 243, 111792. [Google Scholar] [CrossRef]
  109. Li, Z.-L.; Tang, B.-H.; Wu, H.; Ren, H.; Yan, G.; Wan, Z.; Trigo, I.F.; Sobrino, J.A. Satellite-derived land surface temperature: Current status and perspectives. Remote Sens. Environ. 2013, 131, 14–37. [Google Scholar] [CrossRef] [Green Version]
  110. Beach, T.; Luzzadder-Beach, S.; Dunning, N.P. Out of the Soil: Soil (Dark Matter Biodiversity) and Societal “Collapses” from Mesoamerica to Mesopotamia and Beyond. In Biological Extinction: New Perspectives; Dasgupta, P., Raven, P., McIvor, A., Eds.; Cambridge University Press: Cambridge, UK, 2019; pp. 138–174. [Google Scholar]
  111. Drobnik, T.; Schwaab, J.; Grêt-Regamey, A. Moving towards integrating soil into spatial planning: No net loss of soil-based ecosystem services. J. Environ. Manag. 2020, 263, 110406. [Google Scholar] [CrossRef] [PubMed]
  112. Smith, P.; Soussana, J.-F.; Angers, D.; Schipper, L.; Chenu, C.; Rasse, D.P.; Batjes, N.H.; van Egmond, F.; McNeill, S.; Kuhnert, M.; et al. How to measure, report and verify soil carbon change to realize the potential of soil carbon sequestration for atmospheric greenhouse gas removal. Glob. Chang. Biol. 2020, 26, 219–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Shoshany, M.; Goldshleger, N.; Chudnovsky, A. Monitoring of agricultural soil degradation by remote-sensing methods: A review. Int. J. Remote Sens. 2013, 34, 6152–6181. [Google Scholar] [CrossRef]
  114. Delgado-Baquerizo, M.; Maestre, F.T.; Reich, P.B.; Trivedi, P.; Osanai, Y.; Liu, Y.-R.; Hamonts, K.; Jeffries, T.C.; Singh, B.K. Carbon content and climate variability drive global soil bacterial diversity patterns. Ecol. Monogr. 2016, 86, 373–390. [Google Scholar] [CrossRef]
  115. Fierer, N.; Jackson, R.B. The diversity and biogeography of soil bacterial communities. Proc. Natl. Acad. Sci. USA 2006, 103, 626–631. [Google Scholar] [CrossRef] [Green Version]
  116. Bastida, F.; Eldridge, D.J.; García, C.; Kenny Png, G.; Bardgett, R.D.; Delgado-Baquerizo, M. Soil microbial diversity-biomass relationships are driven by soil carbon content across global biomes. ISME J. 2021. [Google Scholar] [CrossRef]
  117. Siciliano, S.D.; Palmer, A.S.; Winsley, T.; Lamb, E.; Bissett, A.; Brown, M.V.; van Dorst, J.; Ji, M.; Ferrari, B.C.; Grogan, P.; et al. Soil fertility is associated with fungal and bacterial richness, whereas pH is associated with community composition in polar soil microbial communities. Soil Biol. Biochem. 2014, 78, 10–20. [Google Scholar] [CrossRef]
  118. Maestre, F.T.; Delgado-Baquerizo, M.; Jeffries, T.C.; Eldridge, D.J.; Ochoa, V.; Gozalo, B.; Quero, J.L.; García-Gómez, M.; Gallardo, A.; Ulrich, W.; et al. Increasing aridity reduces soil microbial diversity and abundance in global drylands. Proc. Natl. Acad. Sci. USA 2015, 112, 15684–15689. [Google Scholar] [CrossRef] [Green Version]
  119. Weil, R.R.; Brady, N.C. Nature and Properties of Soils, 15th ed.; Pearson Education: London, UK, 2016; p. 192. [Google Scholar]
  120. Martínez-López, J.; Bertzky, B.; Bonet-García, F.; Bastin, L.; Dubois, G. Biophysical Characterization of Protected Areas Globally through Optimized Image Segmentation and Classification. Remote Sens. 2016, 8, 780. [Google Scholar] [CrossRef] [Green Version]
  121. Pessôa, A.C.M.; Anderson, L.O.; Carvalho, N.S.; Campanharo, W.A.; Junior, C.H.L.S.; Rosan, T.M.; Reis, J.B.C.; Pereira, F.R.S.; Assis, M.; Jacon, A.D.; et al. Intercomparison of burned area products and its implication for carbon emission estimations in the amazon. Remote Sens. 2020, 12, 3864. [Google Scholar] [CrossRef]
  122. Chen, F.; Crow, W.T.; Ciabatta, L.; Filippucci, P.; Panegrossi, G.; Marra, A.C.; Puca, S.; Massari, C. Enhanced Large-Scale Validation of Satellite-Based Land Rainfall Products. J. Hydrometeor 2021, 22, 245–257. [Google Scholar] [CrossRef]
  123. Cunningham, D.; Cunningham, P.; Fagan, M.E. Identifying biases in global tree cover products: A case study in costa rica. Forests 2019, 10, 853. [Google Scholar] [CrossRef] [Green Version]
  124. Willcock, S.; Hooftman, D.A.P.; Blanchard, R.; Dawson, T.P.; Hickler, T.; Lindeskog, M.; Martinez-Lopez, J.; Reyers, B.; Watts, S.M.; Eigenbrod, F.; et al. Ensembles of ecosystem service models can improve accuracy and indicate uncertainty. Sci. Total Environ. 2020, 747, 141006. [Google Scholar] [CrossRef]
  125. Tuanmu, M.-N.; Jetz, W. A global 1-km consensus land-cover product for biodiversity and ecosystem modelling. Glob. Ecol. Biogeogr. 2014, 23, 1031–1045. [Google Scholar] [CrossRef]
  126. Muller-Karger, F.E.; Miloslavich, P.; Bax, N.J.; Simmons, S.; Costello, M.J.; Sousa Pinto, I.; Canonico, G.; Turner, W.; Gill, M.; Montes, E.; et al. Advancing marine biological observations and data requirements of the complementary essential ocean variables (eovs) and essential biodiversity variables (ebvs) frameworks. Front. Mar. Sci. 2018, 5. [Google Scholar] [CrossRef]
  127. Muller-Karger, F.E.; Hestir, E.; Ade, C.; Turpie, K.; Roberts, D.A.; Siegel, D.; Miller, R.J.; Humm, D.; Izenberg, N.; Keller, M.; et al. Satellite sensor requirements for monitoring essential biodiversity variables of coastal ecosystems. Ecol. Appl. 2018, 28, 749–760. [Google Scholar] [CrossRef] [PubMed]
  128. Miloslavich, P.; Bax, N.J.; Simmons, S.E.; Klein, E.; Appeltans, W.; Aburto-Oropeza, O.; Andersen Garcia, M.; Batten, S.D.; Benedetti-Cecchi, L.; Checkley, D.M.; et al. Essential ocean variables for global sustained observations of biodiversity and ecosystem changes. Glob. Chang. Biol. 2018. [Google Scholar] [CrossRef]
  129. El Mahrad, B.; Newton, A.; Icely, J.D.; Kacimi, I.; Abalansa, S.; Snoussi, M. Contribution of remote sensing technologies to a holistic coastal and marine environmental management framework: A review. Remote Sens. 2020, 12, 2313. [Google Scholar] [CrossRef]
  130. Kachelriess, D.; Wegmann, M.; Gollock, M.; Pettorelli, N. The application of remote sensing for marine protected area management. Ecol. Indic. 2014, 36, 169–177. [Google Scholar] [CrossRef]
  131. Wilson, A.M.; Jetz, W. Remotely Sensed High-Resolution Global Cloud Dynamics for Predicting Ecosystem and Biodiversity Distributions. PLoS Biol. 2016, 14, e1002415. [Google Scholar] [CrossRef] [PubMed]
  132. Tansey, K.; Grégoire, J.-M.; Defourny, P.; Leigh, R.; Pekel, J.-F.; van Bogaert, E.; Bartholomé, E. A new, global, multi-annual (2000–2007) burnt area product at 1 km resolution. Geophys. Res. Lett. 2008, 35. [Google Scholar] [CrossRef]
  133. Giglio, L.; Csiszar, I.; Justice, C.O. Global distribution and seasonality of active fires as observed with the Terra and Aqua Moderate Resolution Imaging Spectroradiometer (MODIS) sensors. J. Geophys. Res. 2006, 111. [Google Scholar] [CrossRef]
  134. Carmona-Moreno, C.; Belward, A.; Malingreau, J.-P.; Hartley, A.; Garcia-Alegre, M.; Antonovskiy, M.; Buchshtaber, V.; Pivovarov, V. Characterizing interannual variations in global fire calendar using data from Earth observing satellites. Glob. Chang. Biol. 2005, 11, 1537–1555. [Google Scholar] [CrossRef]
  135. Potapov, P.; Li, X.; Hernandez-Serna, A.; Tyukavina, A.; Hansen, M.C.; Kommareddy, A.; Pickens, A.; Turubanova, S.; Tang, H.; Silva, C.E.; et al. Mapping global forest canopy height through integration of GEDI and Landsat data. Remote Sens. Environ. 2020, 112165. [Google Scholar] [CrossRef]
  136. Lehner, B.; Grill, G. Global river hydrography and network routing: Baseline data and new approaches to study the world’s large river systems. Hydrol. Process. 2013, 27, 2171–2186. [Google Scholar] [CrossRef]
  137. Hansen, M.C.; Potapov, P.V.; Moore, R.; Hancher, M.; Turubanova, S.A.; Tyukavina, A.; Thau, D.; Stehman, S.V.; Goetz, S.J.; Loveland, T.R.; et al. High-resolution global maps of 21st-century forest cover change. Science 2013, 342, 850–853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ermida, S.L.; Soares, P.; Mantas, V.; Göttsche, F.-M.; Trigo, I.F. Google Earth Engine Open-Source Code for Land Surface Temperature Estimation from the Landsat Series. Remote Sens. 2020, 12, 1471. [Google Scholar] [CrossRef]
  139. Metzger, M.J.; Bunce, R.G.H.; Jongman, R.H.G.; Sayre, R.; Trabucco, A.; Zomer, R. A high-resolution bioclimate map of the world: A unifying framework for global biodiversity research and monitoring. Glob. Ecol. Biogeogr. 2013, 22, 630–638. [Google Scholar] [CrossRef] [Green Version]
  140. Sayre, R.; Dangermond, J.; Frye, C.; Vaughan, R.; Aniello, P.; Breyer, S.; Cribbs, D.; Hopkins, D.; Naumann, R.; Derrenbacher, B. A New Map of Global Ecological Land Units: An Ecophysiographic Stratification Approach; American Association Of Geographers: Washington, DC, USA, 2014; p. 46. [Google Scholar]
  141. Tuanmu, M.-N.; Jetz, W. A global, remote sensing-based characterization of terrestrial habitat heterogeneity for biodiversity and ecosystem modelling. Glob. Ecol. Biogeogr. 2015, 24, 1329–1339. [Google Scholar] [CrossRef]
  142. Huete, A.; Didan, K.; Miura, T.; Rodriguez, E.P.; Gao, X.; Ferreira, L.G. Overview of the radiometric and biophysical performance of the MODIS vegetation indices. Remote Sens. Environ. 2002, 83, 195–213. [Google Scholar] [CrossRef]
  143. Keith, D.A.; Ferrer-Paris, J.R.; Nicholson, E.; Kingsford, R.T. (Eds.) IUCN Global Ecosystem Typology 2.0: Descriptive Profiles for Biomes and Ecosystem Functional Groups; IUCN, International Union for Conservation of Nature: Gland, Switzerland, 2020. [Google Scholar]
  144. Lucas, R.; Blonda, P.; Bunting, P.; Jones, G.; Inglada, J.; Arias, M.; Kosmidou, V.; Petrou, Z.I.; Manakos, I.; Adamo, M.; et al. The Earth Observation Data for Habitat Monitoring (EODHaM) system. Int. J. Appl. Earth Obs. Geoinf. 2015, 37, 17–28. [Google Scholar] [CrossRef]
  145. Dubois, G.; Bastin, L.; Bertzky, B.; Mandrici, A.; Conti, M.; Saura, S.; Cottam, A.; Battistella, L.; Martínez-López, J.; Boni, M.; et al. Integrating Multiple Spatial Datasets to Assess Protected Areas: Lessons Learnt from the Digital Observatory for Protected Areas (DOPA). ISPRS Int. J. Geoinf. 2016, 5, 242. [Google Scholar] [CrossRef] [Green Version]
  146. Brink, A.; Martínez-López, J.; Szantoi, Z.; Moreno-Atencia, P.; Lupi, A.; Bastin, L.; Dubois, G. Indicators for assessing habitat values and pressures for protected areas—An integrated habitat and land cover change approach for the udzungwa mountains national park in tanzania. Remote Sens. 2016, 8, 862. [Google Scholar] [CrossRef] [Green Version]
  147. Dubois, G.; Schulz, M.; Skøien, J.; Bastin, L.; Peedell, S. eHabitat, a multi-purpose Web Processing Service for ecological modeling. Environ. Model. Softw. 2013, 41, 123–133. [Google Scholar] [CrossRef] [Green Version]
  148. Dubois, G.; Bastin, L.; Martínez-López, J.; Cottam, A.; Temperley, W.; Bertzky, B.; Graziano, M. The Digital Observatory for Protected Areas (DOPA) Explorer 1.0; EUR 27162 EN, Publications Office of the European Union: Luxembourg, 2015. [Google Scholar]
  149. Hoffmann, S.; Beierkuhnlein, C.; Field, R.; Provenzale, A.; Chiarucci, A. Uniqueness of protected areas for conservation strategies in the european union. Sci. Rep. 2018, 8, 6445. [Google Scholar] [CrossRef]
  150. Ejrnæs, R.; Frøslev, T.G.; Høye, T.T.; Kjøller, R.; Oddershede, A.; Brunbjerg, A.K.; Hansen, A.J.; Bruun, H.H. Uniquity: A general metric for biotic uniqueness of sites. Biol. Conserv. 2018, 225, 98–105. [Google Scholar] [CrossRef]
  151. Forero-Medina, G.; Joppa, L. Representation of global and national conservation priorities by Colombia’s Protected Area Network. PLoS ONE 2010, 5, e13210. [Google Scholar] [CrossRef]
  152. Martínez-López, J.; Bergillos, R.J.; Bonet, F.J.; de Vente, J. Connecting research infrastructures, scientific and sectorial networks to support integrated management of Mediterranean coastal and rural areas. Environ. Res. Lett. 2019, 14, 115001. [Google Scholar] [CrossRef]
  153. Bennett, M.K.; Younes, N.; Joyce, K. Automating drone image processing to map coral reef substrates using google earth engine. Drones 2020, 4, 50. [Google Scholar] [CrossRef]
  154. Ospina-Alvarez, A.; de Juan, S.; Davis, K.J.; González, C.; Fernández, M.; Navarrete, S.A. Integration of biophysical connectivity in the spatial optimization of coastal ecosystem services. Sci. Total Environ. 2020, 733, 139367. [Google Scholar] [CrossRef]
  155. Giardino, C.; Brando, V.E.; Gege, P.; Pinnel, N.; Hochberg, E.; Knaeps, E.; Reusen, I.; Doerffer, R.; Bresciani, M.; Braga, F.; et al. Imaging spectrometry of inland and coastal waters: State of the art, achievements and perspectives. Surv. Geophys. 2018, 40, 1–29. [Google Scholar] [CrossRef] [Green Version]
  156. de la Fuente, B.; Bertzky, B.; Delli, G.; Mandrici, A.; Conti, M.; Florczyk, A.J.; Freire, S.; Schiavina, M.; Bastin, L.; Dubois, G. Built-up areas within and around protected areas: Global patterns and 40-year trends. Glob. Ecol. Conserv. 2020, 24, e01291. [Google Scholar] [CrossRef] [PubMed]
  157. Willcock, S.; Hooftman, D.A.P.; Balbi, S.; Blanchard, R.; Dawson, T.P.; O’Farrell, P.J.; Hickler, T.; Hudson, M.D.; Lindeskog, M.; Martinez-Lopez, J.; et al. A Continental-Scale Validation of Ecosystem Service Models. Ecosystems 2019, 1–16. [Google Scholar] [CrossRef] [Green Version]
  158. Keith, D.A.; Ferrer-Paris, J.R.; Nicholson, E.; Bishop, M.J.; Polidoro, B.A.; Ramirez-Llodra, E.; Tozer, M.G.; Nel, J.L.; Mac Nally, R.; Gregr, E.J.; et al. Indicative distribution maps for Ecosystem Functional Groups—Level 3 of IUCN Global Ecosystem Typology. Zenodo 2020. [Google Scholar] [CrossRef]
  159. Lightfoot, P. Object-Based Mapping of Temperate Marine Habitats from Multi-Resolution Remote Sensing Data. Ph.D. Thesis, Newcastle University, Newcastle, UK, 2018. [Google Scholar]
  160. Sagar, S.; Falkner, I.; Dekker, A.; Huang, Z.; Blondeau-Patissier, D.; Phillips, C.; Przeslawski, R. Earth Observation for monitoring of Australian Marine Parks and Other off-Shore Marine Protected Areas; Report to the National Environmental Science Program—Marine Biodiversity Hub; Geoscience Australia: Canberra, Australia, 2020. [Google Scholar]
  161. Innangi, S.; Tonielli, R.; Romagnoli, C.; Budillon, F.; Di Martino, G.; Innangi, M.; Laterza, R.; Le Bas, T.; Lo Iacono, C. Seabed mapping in the Pelagie Islands marine protected area (Sicily Channel, southern Mediterranean) using Remote Sensing Object Based Image Analysis (RSOBIA). Mar. Geophys. Res. 2018, 1–23. [Google Scholar] [CrossRef] [Green Version]
  162. Hogg, O.T.; Huvenne, V.A.I.; Griffiths, H.J.; Linse, K. On the ecological relevance of landscape mapping and its application in the spatial planning of very large marine protected areas. Sci. Total Environ. 2018, 626, 384–398. [Google Scholar] [CrossRef]
  163. Assis, J.; Fragkopoulou, E.; Serrão, E.A.; Horta e Costa, B.; Gandra, M.; Abecasis, D. Weak biodiversity connectivity in the European network of no-take marine protected areas. Sci. Total Environ. 2021, 145664. [Google Scholar] [CrossRef] [PubMed]
  164. U.S. Geological Survey; Sayre, R.; Wright, D.; Breyer, S.; Butler, K.; Van Graafeiland, K.; Costello, M.; Harris, P.; Goodin, K.; Guinotte, J.; et al. A Three-Dimensional Mapping of the Ocean Based on Environmental Data. Oceanography 2017, 30, 90–103. [Google Scholar] [CrossRef]
  165. Longhurst, A.R. Ecological Geography of the Sea; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  166. Roberson, L.A.; Lagabrielle, E.; Lombard, A.T.; Sink, K.; Livingstone, T.; Grantham, H.; Harris, J.M. Pelagic bioregionalisation using open-access data for better planning of marine protected area networks. Ocean Coast. Manag. 2017, 148, 214–230. [Google Scholar] [CrossRef]
  167. Godet, C.; Robuchon, M.; Leroy, B.; Cotté, C.; Baudena, A.; Da Silva, O.; Fabri-Ruiz, S.; Lo Monaco, C.; Sergi, S.; Koubbi, P. Matching zooplankton abundance and environment in the South Indian Ocean and Southern Ocean. Deep Sea Res. Part I Oceanogr. Res. Pap. 2020, 163, 103347. [Google Scholar] [CrossRef]
  168. Tamiminia, H.; Salehi, B.; Mahdianpari, M.; Quackenbush, L.; Adeli, S.; Brisco, B. Google Earth Engine for geo-big data applications: A meta-analysis and systematic review. ISPRS J. Photogramm. Remote Sens. 2020, 164, 152–170. [Google Scholar] [CrossRef]
  169. Pekel, J.-F.; Cottam, A.; Gorelick, N.; Belward, A.S. High-resolution mapping of global surface water and its long-term changes. Nature 2016, 540, 418–422. [Google Scholar] [CrossRef]
  170. Bastin, L.; Gorelick, N.; Saura, S.; Bertzky, B.; Dubois, G.; Fortin, M.-J.; Pekel, J.-F. Inland surface waters in protected areas globally: Current coverage and 30-year trends. PLoS ONE 2019, 14, e0210496. [Google Scholar] [CrossRef]
  171. Gorelick, N.; Hancher, M.; Dixon, M.; Ilyushchenko, S.; Thau, D.; Moore, R. Google Earth Engine: Planetary-scale geospatial analysis for everyone. Remote Sens. Environ. 2017. [Google Scholar] [CrossRef]
  172. Kumar, L.; Mutanga, O. Google earth engine applications since inception: Usage, trends, and potential. Remote Sens. 2018, 10, 1509. [Google Scholar] [CrossRef] [Green Version]
  173. Tallis, H.; Mooney, H.; Andelman, S.; Balvanera, P.; Cramer, W.; Karp, D.; Polasky, S.; Reyers, B.; Ricketts, T.; Running, S.; et al. A global system for monitoring ecosystem service change. Bioscience 2012, 62, 977–986. [Google Scholar] [CrossRef] [Green Version]
  174. Vihervaara, P.; Auvinen, A.-P.; Mononen, L.; Törmä, M.; Ahlroth, P.; Anttila, S.; Böttcher, K.; Forsius, M.; Heino, J.; Heliölä, J.; et al. How Essential Biodiversity Variables and remote sensing can help national biodiversity monitoring. Glob. Ecol. Conserv. 2017, 10, 43–59. [Google Scholar] [CrossRef]
  175. Pettorelli, N.; Chauvenet, A.L.M.; Duffy, J.P.; Cornforth, W.A.; Meillere, A.; Baillie, J.E.M. Tracking the effect of climate change on ecosystem functioning using protected areas: Africa as a case study. Ecol. Indic. 2012, 20, 269–276. [Google Scholar] [CrossRef]
  176. O’Connor, B.; Secades, C.; Penner, J.; Sonnenschein, R.; Skidmore, A.; Burgess, N.D.; Hutton, J.M. Earth observation as a tool for tracking progress towards the Aichi Biodiversity Targets. Remote Sens. Ecol. Conserv. 2015, 1, 19–28. [Google Scholar] [CrossRef]
  177. Szantoi, Z.; Brink, A.; Buchanan, G.; Bastin, L.; Lupi, A.; Simonetti, D.; Mayaux, P.; Peedell, S.; Davy, J. A simple remote sensing based information system for monitoring sites of conservation importance. Remote Sens. Ecol. Conserv. 2016, 2, 16–24. [Google Scholar] [CrossRef]
  178. Campbell, A.D.; Wang, Y. Salt marsh monitoring along the mid-Atlantic coast by Google Earth Engine enabled time series. PLoS ONE 2020, 15, e0229605. [Google Scholar] [CrossRef] [Green Version]
  179. Hobern, D.; Apostolico, A.; Arnaud, E.; Bello, J.C.; Canhos, D.; Dubois, G.; Field, D.; Alonso Garcia, E.; Hardisty, A.; Harrison, J.; et al. Global Biodiversity Informatics Outlook: Delivering Biodiversity Knowledge in the Information Age; GBIF Secretariat: Copenhagen, Denmark, 2013. [Google Scholar]
  180. Kissling, W.D.; Hardisty, A.; García, E.A.; Santamaria, M.; De Leo, F.; Pesole, G.; Freyhof, J.; Manset, D.; Wissel, S.; Konijn, J.; et al. Towards global interoperability for supporting biodiversity research on essential biodiversity variables (EBVs). Biodiversity 2015, 16, 99–107. [Google Scholar] [CrossRef] [Green Version]
  181. Hardisty, A.R.; Michener, W.K.; Agosti, D.; Alonso García, E.; Bastin, L.; Belbin, L.; Bowser, A.; Buttigieg, P.L.; Canhos, D.A.L.; Egloff, W.; et al. The Bari Manifesto: An interoperability framework for essential biodiversity variables. Ecol. Inform. 2019, 49, 22–31. [Google Scholar] [CrossRef]
  182. Kissling, W.D.; Ahumada, J.A.; Bowser, A.; Fernandez, M.; Fernández, N.; García, E.A.; Guralnick, R.P.; Isaac, N.J.B.; Kelling, S.; Los, W.; et al. Building essential biodiversity variables (EBVs) of species distribution and abundance at a global scale. Biol. Rev. Camb. Philos. Soc. 2018, 93, 600–625. [Google Scholar] [CrossRef] [Green Version]
  183. Bingham, H.C.; Juffe Bignoli, D.; Lewis, E.; MacSharry, B.; Burgess, N.D.; Visconti, P.; Deguignet, M.; Misrachi, M.; Walpole, M.; Stewart, J.L.; et al. Sixty years of tracking conservation progress using the World Database on Protected Areas. Nat. Ecol. Evol. 2019, 3, 737–743. [Google Scholar] [CrossRef] [PubMed]
  184. Signorello, G.; Prato, C.; Marzo, A.; Ientile, R.; Cucuzza, G.; Sciandrello, S.; Martínez-López, J.; Balbi, S.; Villa, F. Are protected areas covering important biodiversity sites? An assessment of the nature protection network in Sicily (Italy). Land Use Policy 2018, 78, 593–602. [Google Scholar] [CrossRef]
  185. Willcock, S.; Martínez-López, J.; Hooftman, D.A.P.; Bagstad, K.J.; Balbi, S.; Marzo, A.; Prato, C.; Sciandrello, S.; Signorello, G.; Voigt, B.; et al. Machine learning for ecosystem services. Ecosyst. Serv. 2018, 33, 165–174. [Google Scholar] [CrossRef]
Figure 1. Overview of the different elements that need to be included and analyzed in the biophysical characterization of PAs.
Figure 1. Overview of the different elements that need to be included and analyzed in the biophysical characterization of PAs.
Ijgi 10 00384 g001
Figure 2. Example map of the habitat functional types (HFTs) identified in Udzungwa Mountains National Park (Tanzania) and normalized mean values of the biophysical variables used in the eHabitat+ model (EPSG:4326). NDVI stands for the normalized difference vegetation index, and NDWI stands for the normalized difference water index. A detailed description of the study variables and the methodology followed can be found in [120].
Figure 2. Example map of the habitat functional types (HFTs) identified in Udzungwa Mountains National Park (Tanzania) and normalized mean values of the biophysical variables used in the eHabitat+ model (EPSG:4326). NDVI stands for the normalized difference vegetation index, and NDWI stands for the normalized difference water index. A detailed description of the study variables and the methodology followed can be found in [120].
Ijgi 10 00384 g002
Table 1. Relevant biophysical input variables that can be used for the characterization of terrestrial protected areas at a global scale. Acronyms used: NASA; National Snow and Ice Data Center (NSIDC); U.S. Geological Survey (USGS); European Space Agency (ESA); Global Land Analysis and Discovery (GLAD); hydrological data and maps based on Shuttle elevation derivatives at multiple scales (HydroSHEDS); General Bathymetric Chart of the Oceans (GEBCO).
Table 1. Relevant biophysical input variables that can be used for the characterization of terrestrial protected areas at a global scale. Acronyms used: NASA; National Snow and Ice Data Center (NSIDC); U.S. Geological Survey (USGS); European Space Agency (ESA); Global Land Analysis and Discovery (GLAD); hydrological data and maps based on Shuttle elevation derivatives at multiple scales (HydroSHEDS); General Bathymetric Chart of the Oceans (GEBCO).
TopicVariableBased on RSTemporal Extent and ResolutionSpatial ResolutionProducer
ClimateWorldClim bioclimatic variables (a set of temperature and rainfall variables specifically developed for ecological modeling)NoMonthly average climate datasets from the period 1970–2000 and future climate data1 kmWorldClim version 2.1: [101]
ClimateMean annual precipitation
ClimatePotential evapotranspirationYesMulti-daily datasets from 2001 to present500 mUSGS
ClimateCloud coverYesMonthly average from a 15-year period (2000–2014)1 kmEarthEnv [131]
VegetationFire frequencyYesMonthly data from 2001 to present250 mESA Copernicus [132,133,134]
VegetationPercentage of woody vegetation coverYesYearly datasets from 2000 to 2020250 mUSGS
VegetationPercentage of grassland cover
VegetationMean of the maximum and minimum normalized difference vegetation indexYesMulti-daily datasets from 2000 to present250 mUSGS
VegetationLeaf area indexYesMulti-daily datasets from 2014 to present300 mESA Copernicus
VegetationVegetation heightYes201930 mGLAD [135]
SoilSurface soil moistureYesDaily datasets from 1978 to present27.75 kmESA Copernicus
SoilSoil organic carbonNoReference period: 1905–2016250 mSoilGrids
SoilSoil texture
SoilSoil acidity
TerrainSlope, elevation and aspectPartially2020500 mGEBCO
TerrainModified topographic index (can be derived from flow accumulation)Partially2008500 mHydroSHEDS [136]
WaterMean Normalized Difference Water Index (can be derived from surface reflectance composites)YesDaily datasets from 2000 to present500 mUSGS
WaterWater seasonalityYesReference period: 1999–201830 mGLAD [108]
WaterSnow water equivalent (amount of water contained within the snowpack)YesDaily datasets from 2002 to 201125 kmNSIDC
WaterSnow cover fraction or frequencyYesDaily datasets from 2000 to present500 mNSIDC
EnergySurface albedo
EnergyLand surface temperature (LST, a mixture of vegetation and soil temperature)YesMulti-daily datasets from 2000 to present5.6 kmUSGS
EnergyMean solar radiationNoMonthly average climate datasets from the period 1970 to 20001 kmWorldClim version 2.1: [101]
Table 2. Summary table with example applications of different environmental descriptors, including data and methods, that are relevant for the biophysical characterization of PAs. Acronyms used: object-based image analysis (OBIA); normalized difference vegetation index (NDVI); normalized difference water index (NDWI); machine learning (ML); principal components analysis (PCA); light detection and ranging (LiDAR); digital elevation model (DEM); normalized difference blue/red ratio (NDBR); wide dynamic range vegetation index (WDRVI); soil-adjusted vegetation index (SAVI); green–red vegetation index (GRVI); plant senescence reflectance index (PSRI); and water band index (WBI).
Table 2. Summary table with example applications of different environmental descriptors, including data and methods, that are relevant for the biophysical characterization of PAs. Acronyms used: object-based image analysis (OBIA); normalized difference vegetation index (NDVI); normalized difference water index (NDWI); machine learning (ML); principal components analysis (PCA); light detection and ranging (LiDAR); digital elevation model (DEM); normalized difference blue/red ratio (NDBR); wide dynamic range vegetation index (WDRVI); soil-adjusted vegetation index (SAVI); green–red vegetation index (GRVI); plant senescence reflectance index (PSRI); and water band index (WBI).
ApplicationEnvironmental DescriptorsRS and Ancillary DataMethods
Wetland and dune habitat mapping [63,64,65,66,67,68,69,70,73,178]
  • Vegetation greenness
  • Vegetation and soil water content
  • Water seasonality
  • Topography
  • Soil
RS-based vegetation (NDVI, WDRVI, SAVI) and water (NDWI) indices; LiDAR- or radar-derived DEMs; soil depth layer interpolated from ground collected data points; modeled spatial and temporal distribution of water.OBIA; ML; PCA; texture analysis; cluster analysis
Riparian, forest, grassland and shrubland habitat mapping [71,72,75,76,77,81]
  • Vegetation greenness
  • Vegetation height
  • Topography
RS-based vegetation indices (NDVI, EVI; GRVI); LiDAR-derived vegetation height; radar-derived DEM.OBIA; texture analysis; PCA; ML; cluster analysis
Assessment of habitat quality, diversity and extent [59,60,83,85,88]
  • Vegetation greenness
  • Vegetation height
  • Primary productivity
  • Vegetation seasonality
  • Canopy shadow fraction (CSF)
  • Vegetation and soil water content
  • Topography
RS-based vegetation indices (NDVI, EVI); LiDAR-derived vegetation height; Slope derived from a DEM; RS-based water index (NDWI); CSF from RS-based NDBR; vegetation seasonality and productivity products derived from the analysis of temporal dynamics of seasonal changes in NDVI;Cluster and landscape pattern analysis; texture analysis; PCA
Environmental stratifications [17,51,57,139,140,141]
  • Vegetation greenness
  • Bioclimatic variables
  • Altitude
  • Geomorphology and landforms
  • Land cover
  • Lithology
RS-based vegetation indices (EVI); long-term average climate data, such as temperature, precipitation and aridity, interpolated from meteorological stations; geomorphological, landforms and altitude data from a LiDAR- or radar-derived DEM; global lithology map integrating existing surficial lithology maps; land cover classes interpreted from satellite data.Cluster analysis; PCA; texture analysis
Mapping of ecosystem and habitat functional types [52,53,96,120,144]
  • Vegetation greenness
  • Vegetation and soil water content
  • Vegetation phenology and productivity
  • Vegetation structure
  • Land surface temperature (LST)
  • Albedo
  • Soil moisture (SM)
  • Bioclimatic variables
  • Topography
RS-based vegetation (NDVI, PSRI) and water (WBI, NDWI) indices; vegetation phenology and productivity products derived from the analysis of temporal dynamics of seasonal changes in NDVI; LST derived from satellite thermal infrared bands, such as MODIS; RS-derived albedo; RS-based soil moisture products, such as the ESA CCI Soil Moisture; slope derived from a DEM; RS-based percentage of woody and grassland vegetation cover; long-term average climate data, such as temperature, precipitation and aridity, interpolated from meteorological stations.PCA; cluster and landscape pattern analysis; ML; OBIA
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Martínez-López, J.; Bertzky, B.; Willcock, S.; Robuchon, M.; Almagro, M.; Delli, G.; Dubois, G. Remote Sensing Methods for the Biophysical Characterization of Protected Areas Globally: Challenges and Opportunities. ISPRS Int. J. Geo-Inf. 2021, 10, 384. https://0-doi-org.brum.beds.ac.uk/10.3390/ijgi10060384

AMA Style

Martínez-López J, Bertzky B, Willcock S, Robuchon M, Almagro M, Delli G, Dubois G. Remote Sensing Methods for the Biophysical Characterization of Protected Areas Globally: Challenges and Opportunities. ISPRS International Journal of Geo-Information. 2021; 10(6):384. https://0-doi-org.brum.beds.ac.uk/10.3390/ijgi10060384

Chicago/Turabian Style

Martínez-López, Javier, Bastian Bertzky, Simon Willcock, Marine Robuchon, María Almagro, Giacomo Delli, and Grégoire Dubois. 2021. "Remote Sensing Methods for the Biophysical Characterization of Protected Areas Globally: Challenges and Opportunities" ISPRS International Journal of Geo-Information 10, no. 6: 384. https://0-doi-org.brum.beds.ac.uk/10.3390/ijgi10060384

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