Next Article in Journal
Urban Expansion and Carbon Emission Logistic Curve Hypothesis and Its Verification: A Case Study of Jiangsu Province
Next Article in Special Issue
Impact of Urbanization through High-Speed Rail on Regional Development with the Interaction of Socioeconomic Factors: A View of Regional Industrial Structure
Previous Article in Journal
Volcanic Landforms and Landscapes of the East Carpathians (Romania) and Their Geoheritage Values
Previous Article in Special Issue
Impacts of Zagreb’s Urban Development on Dynamic Changes in Stream Landscapes from Mid-Twentieth Century
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Land Change Science and the STEPLand Framework: An Assessment of Its Progress

1
Grumets Research Group, Departament of Geography, Building B, Universitat Autònoma de Barcelona, 08193 Bellaterra, Catalonia, Spain
2
Department of Methods and Models for Economics, Territory and Finance, Faculty of Economics, Sapienza University of Rome, Via del Castro Laurenziano 9, I-00161 Rome, Italy
*
Author to whom correspondence should be addressed.
Land 2022, 11(7), 1065; https://doi.org/10.3390/land11071065
Submission received: 10 June 2022 / Revised: 8 July 2022 / Accepted: 9 July 2022 / Published: 13 July 2022

Abstract

:
This contribution assesses a new term that is proposed to be established within Land Change Science: Spatio-TEmporal Patterns of Land (‘STEPLand’). It refers to a specific workflow for analyzing land-use/land cover (LUC) patterns, identifying and modeling driving forces of LUC changes, assessing socio-environmental consequences, and contributing to defining future scenarios of land transformations. In this article, we define this framework based on a comprehensive meta-analysis of 250 selected articles published in international scientific journals from 2000 to 2019. The empirical results demonstrate that STEPLand is a consolidated protocol applied globally, and the large diversity of journals, disciplines, and countries involved shows that it is becoming ubiquitous. In this paper, the main characteristics of STEPLand are provided and discussed, demonstrating that the operational procedure can facilitate the interaction among researchers from different fields, and communication between researchers and policy makers.

1. Introduction

Landscape and ecosystem transformations, and their implications for global environmental change and sustainability, are a major research challenge for both ecological and social sciences [1]. As a result, the analysis of changes in land use and land cover is considered essential for researching major environmental issues such as desertification, eutrophication, acidification, greenhouse effects, biodiversity loss, and climate warming [2]. The progress in these studies was extensively coordinated from 1994 to 2005 by the Land-Use and Land-Cover Change (LUCC) core project of the International Geosphere-Biosphere Programme (IGBP) and the International Human Dimensions Programme (IHDP) [3,4], which represent the starting point of other global and regional initiatives, such as the ESA GlobCover [5] and the NASA LCLUC Programme [6]. In 2006, an even broader initiative, the Global Land Project, was established with the aim of synthesizing and integrating insights, knowledge, and research methodologies, thereby identifying scientific priorities and a research agenda [7]. Two major challenges faced by this initiative were (i) a refined understanding of complex feedbacks between societal and environmental components of the integrated land system, and (ii) up-scaling local and regional processes of change to reach a “truly global” understanding [8]. Consequently, a new research field, called Land Change Science (LCS), has attracted increasing interest and efforts to better understand LUCC patterns and dynamics that affect the structure and functioning of Earth systems [9]. This perspective requires using specific analysis tools, e.g., Geographical Information Systems (GIS), to integrate specific knowledge in socio-economic and ecological subsystems.
LCS has carried out research at many spatial and temporal scales with the objectives of explaining human–environmental system dynamics that generate landscape changes; improving spatially explicit LUCC models that are compatible with Earth systems models; and, finally, assessing system intrinsic properties and the related outcomes, such as vulnerability, resilience, and sustainability [1]. The refined analysis of interactions between biophysical and socio-economic subsystems has implied the active involvement of researchers from natural, social, and spatial disciplines. The complexity of integrating diverse data, space–time patterns, and socio-biophysical processes has led to some methodological issues related to merging analytical traditions, mainly when people, places and environment are linked [10]. Spatial accuracy and uncertainty, trans-scalar mismatches, and the choice of aggregation levels (from county/district to households/individuals) are examples of these issues [11].
A seminal contribution by [7] defined LCS as an interdisciplinary research field that engages scientists across the social, economic, geographical, and natural sciences, and is conceptually and operationally separate from the broader discipline of Land System Science. This distinction was motivated by the increasing interest in (i) the role of drivers and impacts, (ii) the interactions between socio-ecological systems, and (iii) the connection between world regions, cities, and their rural hinterlands. Changes over time in human interactions with the natural surroundings, including land management and the provisioning of ecosystem services, are the focus of emerging socio-ecological interpretative paradigms [8]. The usual aims of LUCC studies are to identify drivers of change and, based on this knowledge, envisage future scenarios, proposed in both individual (local) cases and more globalized research. Nevertheless, generalization and validation processes can sometimes be difficult due to the diversity of driving forces, study areas, indicators, and modeling approaches [12].
Giving value to long-lasting research, LCS is a (rapidly evolving) disciplinary field that has produced a wealth of methodological innovations and empirical observations when it has been used to assess and interpret LUCC patterns, drivers, and inter-linkages [13]. From this perspective, its development has been characterized by a focus on local case studies and a specific emphasis on methodological developments that have improved remote sensing tools and geo-spatial analysis. It has also promoted the adoption of theoretical and empirical frameworks derived from disciplines such as geography, landscape ecology, and regional science.
After more than 25 years of LUCC research, our hypothesis is that the progress of LCS seems unquestionable in many social and natural disciplines, and it has become relatively ubiquitous because it is applied by experts worldwide. Based on an original meta-analysis of selected literature, our main objective is to assess a specific workflow resulting from LCS that we propose to call STEPLand (Spatio-TEmporal Patterns of Land). By delineating specific analysis criteria of the reference sample, we summarize the main attributes, data typology, and methodological steps, in addition to technical details and issues of practical applications, typical of LCS approaches. Our review reports on the reliability, robustness, and global applicability of operational definitions and frameworks oriented toward the STEPLand perspective, demonstrating the specific contribution of this approach to selecting appropriate methods and models for LUCC research.

2. STEPLand Framework

STEPLand is defined as an operation framework typical of LCS, and composed of four steps, moving from broader to narrower epistemological perspectives (Figure 1):
1. LUCC patterns and dynamics. The main objective of the first step is LUCC assessment using, e.g., remote techniques and geo-spatial approaches, and answering, at least, the following questions: “What LUC are changing?”, “Where?”, “When?”.
2. LUCC modeling. The main objective is to describe the core drivers of LUCC using qualitative or quantitative methods and answering the following questions: “Which causal factors are involved in LUC changes?”, “Why?”, “How?”.
3. Assessment of socio-environmental consequences. The aim is to delineate and understand the socio-environmental consequences of LUCC using qualitative or quantitative methodologies, and explicitly considering landscape transformations and ecosystem service dynamics to answer the question: “And then?”.
4. Compiling future scenarios. The goal of this last step is to establish future LUCC projections, answering questions such as “What happens if (…)?”.

2.1. LUCC Patterns and Dynamics

The first step of STEPLand includes a formal assessment and characterization of LUCC, mainly based on Earth Observation (EO) techniques, e.g., remote sensing (RS). EO and GIS data facilitate trans-sectoral research and provide a platform for integrating multiple information layers that may include ancillary (or validation) data [14,15]. RS tools provide valuable multitemporal data for monitoring LUCC patterns and processes at different spatial scales (local, regional, national, continental, and global), and GIS techniques make it possible to analyze them explicitly [16,17,18]. Spatial, spectral, and radiometric RS resolutions affect the size, condition, and precision of the explicit features to be discriminated in a landscape scene, whereas the temporal resolution of satellite imagery determines the intrinsic territorial, environmental, and socio-economic dynamics of local systems [10,19,20,21]. Both dimensions represent large technical challenges. During the past 40 years, significant advances in sensor technologies have improved the spatial, spectral, radiometric, and temporal resolutions, and the coverage, of satellite imagery [22]. The classification phase is the process of identifying spectral similarities in the multidimensional spectral space and linking them to LUC categories. This involves assigning pixels to different LUC classes identified in a given study. Diverse classification taxonomy options are now available but, traditionally, these have been divided according to the pixel technique used (per-pixel or sub-pixel), or according to the application of training samples (supervised methods or unsupervised methods), among others [23,24,25].
Similarity in spectral reflectance properties of natural surfaces at a given moment often impedes consistent identification and mapping of a large range of LUC, such as agricultural crops or individual communities constituted by natural vegetation. Furthermore, the spectral response of many cover types varies throughout the year: LUC categories that appear similar in spring may become distinguishable at earlier or later stages of the annual cycle. Therefore, a multi-temporal approach based on large RS databases is more suitable for revealing the complex LUC patterns characteristic, for instance, of the Mediterranean Basin [26,27,28]. In order to minimize the classification errors due to seasonal changes, an appropriate strategy is to select images from the same month or, at least, from the same season [29,30]. After a classification is obtained, the next step is to assess the accuracy or how well a classification has worked. This assessment measures the correlation between satellite imagery classification and ground reference samples to quantify the overall agreement between RS classification (global accuracy and user and producer accuracy by categories) and ground truthing data [31].
Once LUC maps are generated and tested, among the available change detection techniques, the most usual method is the post-classification or map-to-map comparison (i.e., a comparative analysis of independently produced LUC classifications from different dates) associated with a transition or change matrix [32,33]. According to [34], post-classification is an accurate procedure that bypasses the difficulties associated with analyzing images acquired at different times of the year or by different sensors, and thus delineating the nature of change [35]. Nevertheless, this method has two well-known critical issues [36,37]: first, misregistration of the polygon boundaries (locational inaccuracy) in the different classifications, and, therefore, the presence of border pixels with false positive or negative changes. With rasters, this is basically the consequence of non-matching pixels [38]; with vector shapes, this problem is known as “sliver”, i.e., narrow polygons of uncertain interpretation [39,40,41]. The second issue relates to problems derived from classification errors: a false positive change may be recorded when no change has taken place because a pixel in one map (or in both maps) is misclassified, or false negative changes, when no change is identified but a change has taken place in the field. Therefore, this approach requires a good accuracy level in both classification maps because the change map accuracy is the product of the accuracies of the individual classifications, and therefore, it is subject to error propagation [42,43,44].

2.2. LUCC Modeling

Modeling is a process that provides a platform for encoding inferred (or deduced) relationships, thus allowing simulations and projections based on mathematical (algorithmic) specifications or procedures [45]. LUCC modeling is a tool for supporting planners and policymakers in developing robust policies and decisions. At the same time, models can be used to provide an ex ante assessment of policies or serve as an early warning system for environmental impacts. Six conceptual dimensions are considered to be particularly important for carrying out LUCC modeling [46]: (i) analysis level, (ii) cross-scale dynamics, (iii) driving forces, (iv) spatial interactions and neighborhood effects, (v) temporal dynamics, and (vi) level of data integration. In general terms, LUCC is often modeled as a function of a selection of socio-economic and biophysical variables acting as driving forces that shape land change [2,47,48,49].
Modeling LUCC driving forces comprises a wide variety of methodological approaches, which can be classified in different ways. For instance, [3] divided driving forces into three groups: socio-economic drivers, biophysical drivers, and land management variables; [50] differentiated proximate causes, those actions that directly affect land use, such as wood extraction, from underlying causes, i.e., those “fundamental forces” that underpin the proximate causes, including demographic, economic, technological, institutional, and cultural factors [51]. In general, Ref. [52] identified five driving forces: (i) biophysical constraints and potentials, (ii) economic factors, (iii) social factors, (iv) spatial policies, and (v) spatial interactions and neighborhood characteristics. Others [48,53,54] differentiated actors of change from driving forces. Actors are the decision-making and mediating agents, including individuals (e.g., farmers), households, neighborhoods, agencies (e.g., planning organizations), and institutions, whereas driving forces are the (sometimes materialized) expression of their decisions or acts, e.g., through laws, subsidies, or incentives [55]. Moreover, Ref. [53] provided a specific definition of the “spatial domain” as the institutional (and geographical) context where a given agent interacts with the landscape (e.g., countries, regions, prefectures/provinces, municipalities). Representation of the domain can be facilitated in a geographically explicit model using boundary maps or vector layers. The extent of the study area also influences the selection: larger areas imply a diversification of LUC contexts and may require analyzing a larger variety of driving forces. Therefore, the analysis scale (local, regional, national, or global) can produce a specific representation based on different driving forces. For instance, the presence of small ecologically valuable areas can be the main determinant of LUCC patterns at a local scale, whereas the distance from the market can be a more important factor at a regional scale [46].
Specification and quantification of the intrinsic relationship between driving forces and LUCC are particularly important in model implementation. According to [56], modelers should select the drivers or explanatory variables that are supposed to play an active role in land change. Even in automated approaches, input variables are selected based on expert knowledge, although data availability (e.g., the lack of data for some economic variables, such as land ownership, or, for earlier times, e.g., gross domestic product in 1850) is often a major limitation, as documented in [57].
According to [45], models used for LUCC analysis range from those oriented towards pattern description/recognition to those quantifying and interpreting dynamics. One study [58] provided a more generalized classification, arguing that models can be static or dynamic, spatial or non-spatial, inductive or deductive, and/or agent-based or pattern-based (e.g., emulation of individual decision makers vs. inference of the underlying behavior derived from LUCC patterns). Models can use a large range of information (satellite imagery, official statistics, maps or field surveys, among others) often implemented in GIS and eventually combined in composite indicators. Actor-based, bottom-up models based on household surveys represent the land change agent explicitly by emulating individual decision makers through agent-based approaches. Land evaluation pattern-based top-down models use RS and census data to simulate LUCC through parameterized transition equations that convert land from one cover type to another [59]. Finally, Ref. [60] separated inductive pattern-based models from cellular automata approaches, sector-based economic models, spatially disaggregated economic approaches, and agent-based models.
Therefore, driving forces form a tangled system of interactions that affect multiple temporal and spatial levels, making it difficult to carry out adequate analyses and obtain representation systems. Combining data from the social and natural sciences is a particularly complicated task due to the different operational scales, the complexity in relating social science data to a specific geographic place, and the difficulty involved in integrating qualitative data, which is more common in social science, and less common in ecological disciplines [47]. Moreover, driving forces—and not only LUCC—are also subject to changes, which influence the identification of representative study periods, and thus affects the model’s results [61]. Precision of statistical methodologies and data availability are also important. As a summary, identification of LUCC drivers (sensu [26]) implies (i) the clarification of latent relationships between landscape patterns and driving forces (explorative models), and (ii) the projection of future landscapes under different scenarios (predictive models).
Two of the main issues arising when socio-economic and biophysical variables are combined spatially, as a process of data integration or data equalization [62,63,64] characteristic of STEPLand, are the different data formats and spatial scales. Socio-economic data are usually available from official statistics in tabular format at some administrative boundaries: neighborhoods, districts, census tracks, municipalities, regions, and countries, among others [65,66]. By comparison, biophysical variables are mainly extracted from EO sources, from spatial interpolation techniques (e.g., exploiting point data derived from climate stations) or from other sources (e.g., rasterization of archive maps) having a specific pixel size [67]. When the two types of data are combined, these issues may lead to a loss in spatial precision [68], given the need to integrate data at one specific scale, namely, according to administrative boundaries (i.e., native vector file) or to lattices (i.e., raster file with a given pixel size). As socio-economic data are mostly available for administrative areas, an option is to convert biophysical data into class intervals and calculate the area (or percentage) occupied by each interval within the appropriate spatial (polygon) domain. For instance, Ref. [69] adopted the area option, whereas [70] calculated the proportion of three different intervals of slopes within each district, and adopted a similar solution for soil moisture. Another study [71] clipped LUC data to the boundaries of each of the 25 watersheds considered in the study, and then calculated zonal statistics for each watershed to extract the relative proportions of each LUC type for subsequent use in statistical analyses. Another option is to calculate the mean value of cardinal variables per spatial (polygon) domain, thus producing a high generalization. It is also possible to assign central (median) or dominant (mode) values for each administrative area. This latter option is appropriate for discrete variables, such as soil type.
When the option is to work at a pixel size, the problem is the inverse, as it affects the socio-economic data: a rasterization is required in order to match the pixel size of biophysical variables and LUCC. The main issue in this case is whether a statistical analysis is applied because all the pixels included in the administrative boundary have the same socio-economic value, showing a maximum spatial autocorrelation, which leads to a violation of the assumption of independent residuals. To minimize this situation, a convenient “solution” is to apply a “reduced factor” through, for instance, a stratified random sampling at a lower number of pixels. This option was applied in [72] using a “contraction factor” amounting to 10. Another option is “data generalization”, e.g., using a grid lattice having larger pixel sizes, which is representative of a lower spatial resolution; for example, in [73], all the variables describing deforestation and the respective drivers of change were aggregated to 25 km grid cells.

2.3. Socio-Environmental Consequences

The third step of STEPLand takes the socio-environmental consequences of LUCC into account. For instance, LUCC may affect weather and climate variability by altering biophysical, biogeochemical, and energy exchange processes at local, regional, and global scales. Therefore, the consequences of these processes are scale-dependent because some of them affect the local environment (e.g., local water quality), whereas other impacts extend far beyond the location where they arise (e.g., carbon cycle, climate change). Because not all LUCCs have global effects, and LUCCs are not irreversible, there are several multi-directional impacts that can reinforce, mitigate, or offset multiple consequences, enriching the debate regarding on-site and off-site factors of change [74,75].
One specific repercussion considered in this review is related to landscape, given its particular nature because it contributes significantly to well-being and quality of life. Quantification of spatio-temporal landscape dynamics and the underlying drivers is key for planning appropriate decisions in this field [76]. Landscape metrics are common tools for measuring spatial changes in landscape composition and configuration, including fragmentation and diversity [77]. From a STEPLand perspective, they can be applied to RS and GIS data and simultaneously used with LUCC models and statistical methods. There are several quantitative measures, for example, those in [78], that can be used to assess landscape composition, such as the number of patches (patch richness) and uniformity and variety (evenness and diversity), and to assess the configuration or the spatial distribution of patches in the landscape (landscape pattern); these include patch shape, isolation, spread between classes (contagion), mean patch size, and density.
Another additional consequence considered in STEPLand is the impact on ecosystem services because, in recent years, these services have attracted the increasing attention of researchers, policymakers, and other stakeholders worldwide [79]. The main reason for this is that, when land is used, society changes and modifies the quantity and quality of the provision of these services. According to [80], supply and demand of ecosystem services can be assessed at, and transferred to, different spatial and temporal scales by linking LUCC (e.g., extracted from remote sensing) with other data (e.g., obtained from interviews). Their results reveal patterns of human activities over time and space, and the capacities of different ecosystems to provide ecosystem services under changing LUC. However, Ref. [81] reviewed the “ecosystem services” concept and the various methods applied for mapping and assessing quantitative methods, and the significant problem of there not being a clear distinction between services, functions, and benefits.

2.4. Futures Scenarios

The last stage included in the STEPLand perspective corresponds to the simulation and prediction of future LUCC scenarios. This is a significant process for policy makers because it enables them to better anticipate actions, especially in the context of urban planning and the protection of natural land [82]. A previous study [83] provided a literature review of models applied to predict LUCC, including Markov chains (MCs), landscape models, CLUE-S models, cellular automata (CA), integration of Markov chains and cellular automata (MC-CA), and artificial neural networks. Markov chains, for instance, quantify LUCC probabilities between different states that are recorded in a transition matrix [84]. This matrix is the result of cross-tabulation between satellite images derived from two sequential dates, adjusted by proportional error and translated into a set of probability images, one for each LUC category [85,86]. In addition, CA models are one of the most relevant tools for understanding complex systems, particularly LUC patterns, given their intrinsic sensitivity to both spatial configuration and neighborhood relationships. The future state of the cells is determined by the current state of the cell itself and that of the neighborhood cells, following transition functions based on a set of rules. A variety of methods has been used for calculating transition rules in CA, such as logistic regression, multinomial logit, linear and geometric formulations, support vector machines, and, more recently, artificial neural networks [87].

3. Methodology

Results of the meta-analysis presented in this chapter aim to provide a snapshot of the STEPLand framework, rather than a complete and systematic overview of the whole body of literature, which is particularly vast. The chosen option was to review in great detail, from 2000 to 2019 (20 years of analysis), a relatively large sample of scientific articles, published in journals having an impact factor, with the aim of extracting key variables and attributes useful for assessing the evolution of the STEPLand framework. To identify articles representative of the STEPLand workflow (excluding books), we first selected published papers from the Web of Science database according to pre-determined criteria, such as target keywords including “land-use”, “land-cover”, and “changes”. The search engine reported a total of 9574 entries. Figure 2 shows the total number of publications by year, outlining an exponential increase since 2000. From the total sample, we selected 250 articles. This sample size was considered significant and representative given the deep reading required to extract the information that was the object of this review. These entries were extracted in four stages with the objective of assessing the STEPLand framework progress. These steps were to determine: (i) the presence of target keywords in the title or in the keywords; and (ii) the presence of the main objectives and applied methodology that are considered characteristic of STEPLand (based on a deep, manual reading). A third criterion was to choose papers published in different scientific journals and years, with the objective of determining the evolution of the STEPLand framework. In the last step, we chose papers with a high number of Web of Science citations (having at least 50 citations until 2007) (Figure 3). The full list of articles can be found in Appendix A.
From the 250 publications selected, and after a deep review, the following data for the first step (LUCC patterns and dynamics) were extracted: time period; the country or countries of the study area; number of LUC categories; RS classification method, if applied, and its accuracy; and, finally, whether an overlay of individual classifications was applied (Table 1).
In the second step (“LUCC modeling”), considering the previous work by [88], our classification of statistical and empirical techniques for identifying the main driving forces of LUCC included: type of analysis (exploratory, training, regression) and objective (data reduction or structure detection, classification, association, graphical model, summarization and data transformation); type of dependent and independent variables (driving forces considered); scale of analysis; and model type (Table 2).
In the third step, the main objective was to report the nature of the socio-environmental consequences of LUCC (e.g., water quality or climate), differentiating a quantitative spatial model from more qualitative patterns (Table 3). If the focus was on landscape, then two variables were analyzed: the software used, and the metrics quantified at class (or landscape) level. Finally, a third option was whether the focus was on ecosystem services (Table 3). In this case, various approaches have been applied, including market prices, productivity, and benefit transfer methods. This latter option is the procedure used to estimate the economic benefits gathered from one site (e.g., considering each biome or LUC type) and applying them to another site with related demographic, economic, and ecological characteristics [79]. The early valuation study by [89] used this approach to extrapolate the overall economic value of 17 ecosystem services provided by 16 main biomes. These estimates were updated based on a larger database of more than 300 case studies from all over the world [90]. Some value modifications are often made according to expert opinion surveys or based on statistical models [91].
Finally, in the fourth step, the future scenarios, the main objective was to report which forecasting model was applied and for which horizon (short, medium, or long-term).

4. Results

The results from the entire sample of manuscripts (250) showed that the partial approaches most common in STEPLand include step 1 and 2 together (LUCC patterns and modeling), followed by step 1 and 3 together (LUCC patterns and socio-environmental consequences), and step 1 alone (LUCC patterns) (Figure 4). Only a few studies adopted a more complete framework, integrating step 1 with 2 and 3 (13%); and only 4% of the sample adopted steps 1, 2, and 4 synergistically. Interestingly, none of the 250 reviewed papers included all the steps.

4.1. Step 1: LUC Changes

Figure 5 shows that 68.0% of the first authors were affiliated with geography, environmental sciences, forestry, ecology, rural development, natural resources, or land management. Therefore, STEPLand can honestly be considered a framework that has been followed mainly by geographers, environmentalists, and ecologists, since the purest social scientists (including economists, for instance) were represented in only 4.4% of manuscripts, included in 32% of “Others” category in the figure. Nonetheless, the sample papers were published in 65 scientific journals. Figure 6 shows the top ten journals having the most publications, totaling 164 papers (65.6%); four journals accounted for 71.9%, whereas the impact factor quartiles of all ten were Q1 and Q2.
From a geographical point of view, the results showed that 74 countries were included in the reviewed articles. Most American, European, and Asian countries were represented, and Africa was the continent that was least represented (Figure 7).
Figure 8 documents that most of the 250 manuscripts were developed using only RS data as an EO source (51.2% of the total), followed by cartographic databases produced by scientific/governmental agencies (22.4%). The remaining works used aerial photographs only, associated with other sources, and cartographic databases based on other sources. A combination of aerial photographs with satellite imagery was quite common during the first analysis period before the 1970s. From the 128 papers that adopted only satellite images, most of them used Landsat exclusively (78.3%) or combined with other sensors (15.5%).
The temporal extent analysis showed that long-term analysis (more than 30 years) was the most usual situation, accounting for more than 80 papers (34.8%), followed by the group between 16 and 20 years (16.4), equivalent to the medium term. However, from the total number of works that used methods for identifying and classifying LUC with RS, 8.0% did not explain the methods used, whereas, in 14.4%, the method used was photointerpretation of satellite images. The rest (77.6%) carried out automatic classifications, with diverse options. The supervised methods were the most common, as they were used in 45.6% of the total. The most common supervised classification was maximum likelihood (equivalent to 71.9% of all the supervised classifiers) (Figure 9), whereas the second most common method for automatic classification was the object-based classifier (OBIA) (11.2% of the total).
The most common numbers of LUC categories used in RS classifications were between six and ten (60.0%) and below six (28.8%). Working with a rather low number of categories was, in many cases, the result of reclassification processes. Considering the accuracy assessment, 53% of works did not mention any value (133 entries); this percentage includes the studies working with cartographical databases produced by scientific/governmental agencies (e.g., CORINE). In terms of the percentage of overall accuracy extracted from the automatic classifications mentioned before, the groups between 81% and 90% and above 90% were the most frequent (36.8% and 24.0%, respectively).

4.2. Step 2: LUCC Modeling

A total of 129 manuscripts included LUCC modeling (51.6%), and, of these, 74 used a statistical modeling technique, whereas 55 presented a qualitative analysis, in some cases supported by interviews and field surveys. The most-used quantitative model was regression (38.0%) followed by exploratory data reduction (7.7%), such as factor analysis, canonical analysis, or correspondence analysis (Figure 10). It is necessary to highlight the conversion of land use and its effects (CLUE) framework [92]. This framework is based on an empirical spatial analysis for identifying the most important biophysical and socio-economic drivers and their relationships with the extension and intensity of LUCC. A previous study [93] provided core publications and examples of applications for the different CLUE versions (CLUE, CLUE-S, Dyna-CLUE, and CLUE-Scanner). Figure 11 shows that the most-used regression technique was LOR, including multiple logistic regression (53.1%) and LIR (28.6%). GWR and BRT were the other important methods, mainly since 2016 for the latter.
The analysis of the dependent (or response) variable showed that the most-used variable corresponded to LUC (25.3%), followed by LUCC (21.1%) and, to a lesser degree, to forestry (19.7%), specifically related to deforestation, afforestation, and/or reforestation. The group of driving forces most used in modeling techniques was the combination of biophysical, socio-economic, and distance variables (40.8%), followed by biophysical and socio-economic variables (21.1%), and socio-economic variables only (18.3%). The pixel was the most usual spatial domain (61.9%), followed by administrative boundaries (25.3%). Finally, the word “drivers”, together with “driving forces”, was used from the first work, in 2000, until the last year of analysis (2019), showing that it has spread successfully.

4.3. Step 3: Socio-Environmental Consequences

A total of 105 manuscripts (42.0% of the total) presented an analysis related to socio-environmental consequences of LUCC. Of these, 41.9% analyzed the socio-environmental consequences using both quantitative (43.2%) and qualitative (56.8%) tools, whereas 43.8% were linked to landscape dynamics, and the rest were mainly connected to ecosystem services (14.3%). Of the total number of manuscripts that analyzed socio-environmental consequences (44), 55.8% corresponded to biodiversity and disasters, followed by agriculture (23.2%). Qualitative analysis was used in 100% of studies of the socio-environmental consequences of LUCC in agriculture, 76.9% of studies of biodiversity, and 45.4% of studies of disasters, whereas analyses of water, climate, weather, and energy were 100% quantitative. The models used in the quantitative analysis were very diverse; some used export coefficients, whereas others used operational indications provided by the Panel on Climate Change, the Unit Stream Power Erosion and Deposition model (USPED), the Soil and Water Assessment Tool (SWAT), the Revised Wind Erosion Equation (RWEQ), or investment in carbon storage.
In the case of landscape consequences, 93.5% of manuscripts calculated some metrics, and the qualitative option was used in the minority. In 63.6% of these papers, FRAGSTATS software was used, followed by Patch Analysis in 13.6% of cases. For both landscape and class levels, the most-used metrics are shown in Figure 12. Those corresponding to configuration were, by order of importance: number of patches, mean patch size, patch density, largest patch index, edge density, landscape shape index and the interspersion and juxtaposition index. For composition, the most-used metric was Shannon’s diversity index.
The results of ecosystem services models showed that the benefit transfer method was clearly the most-used option (66.7%). According to some researchers, this is a commonly applied method for assessing ecosystem services due to its feasibility for making a quick assessment, at regional and national scales, and the low cost of collecting the required primary data; however, the local ecosystem conditions are an evident limitation [91]. InVEST was the second most-used option (20%). This is a set of spatial models that analyze and predict the provision of ES from LUCC maps and related biophysical, economic, and institutional data for a given region [94,95,96].

4.4. Step 4: Future Scenarios

The results of future scenario modeling showed an increasing interest for this topic. A clear predominance of Markov chains as the computation method was evident during the initial years, from 2000 to 2008, and MC-CA dominated the literature in the latter years, with a very fragmented use of the remaining methodologies (Figure 13). The dominance of CA is even greater if we consider (i) the Monitoring Land Cover/Use Dynamics (MOLAND) model, an improved CA version developed by [97]; and (ii) the artificial neural network-based cellular automata (ANN-CA) model, because CA is used to model the LUCC by applying the transition probabilities from the ANN learning process [98].
Another choice is the Land Change Modeler (LCM), which provides a tool for evaluating and modeling LUCC by performing three steps: change analysis, transition sub-models, and LUCC predictions [99]. From the three separate empirical models, the multi-layer perceptron (MLP) neural network is the only procedure that can model multiple transitions at the same time, and it was the selected option in most of the papers analyzed.
Other options that should be highlighted are those models developed specifically for future scenario analysis, such as, for instance, CLUE-S, which is an updated version of the early CLUE model. CLUE-S is based on an empirical analysis of location suitability combined with the dynamic simulation of competition and interactions between the spatiotemporal dynamics of LUC systems, and was specifically developed for the spatially explicit simulation of LUCC [100]. The model includes spatial and non-spatial modules that combine statistical analyses and decision rules that determine the sequence of LUC types [101]. Another alternative is the CLUMondo model, which is the latest addition to the CLUE series, and is a spatially explicit and dynamic land system change model based on the highest transition potential. This variable is the result of a sum of local suitability, conversion resistance, and competitive advantages of a land system [102]. The final parameter analyzed was the temporal extent of future scenarios. The majority was between 11 and 20 years (50%), followed by less than 10 years (26.9%).

5. Discussion

The main objective of this research was to assess the progress of a specific workflow that we proposed to call STEPLand (Spatio-TEmporal Patterns of Land), which involves four identified steps. The study provides a comprehensive picture of the STEPLand framework by analyzing a sample of 250 scientific articles published during the past 20 years. The requirements for inclusion in the analysis were that the papers were published in impact factor journals and had at least 50 citations (until 2007), with the presence of target keywords (land use, land cover, changes) and objectives characteristic of STEPLand (LUCC patterns, spatial modeling of driving forces, socio-environmental consequences, and future scenarios). The analysis also attempted to cover a broad sample of journals. These criteria provided a provisional framework delineating the STEPLand framework that can be improved in future works by changing the conditions if required, and increasing or decreasing the minimum number of citations, for instance. This assumption is made even clearer in the map of analyzed countries (Figure 7), where parts of Africa and the Middle East are underrepresented, and other countries such as New Zealand or Malaysia are missing. Therefore, future work is needed to try to include all countries, while substantially increasing the number of entries.
Once the best indicators, according to the authors’ point of view, to characterize the STEPLand framework were established, data extraction was undertaken based on an in-depth reading analysis by both researchers, which was an empirical task for quantifying the results.
Step 1, LUCC patterns and dynamics, is the basis of the STEPLand framework. From the 250 publications analyzed, and according to the affiliation of the first author and the journal titles, a robust outcome was that geography and ecology were the most common fields. Therefore, the “pure” social vision was much less represented. The citation impact of the first ten journals was high, and was mostly Q1 and Q2.
According to EO sources, the use of satellite images was the predominant option, but it should be noted that a non-negligible percentage of cases (22.4%) were carried out using cartographic databases already produced by scientific or governmental agencies. This option is clearly less time consuming because an RS classification can be a complicated task, especially for non-experts, and it also allows other research steps to be carried out more comfortably. Nevertheless, this situation can only be useful if the temporal extent, the spatial resolution, and the number of LUC categories from cartographic databases are suitable for the specific research target.
Landsat images were the most used for analyzing LUCC having a temporal extent longer than 30 years. Therefore, the medium spatial resolution (30 m × 30 m) was the preferred resolution compared to other options such as MODIS, SPOT, or IKONOS. In fact, this outcome was expected given the difficulties involved in obtaining uninterrupted and continuous time series of data from 1972 onwards, which is the longest-running archive [20,103]. Without doubt, the change to an open (free) data policy in 2012 benefitted its use. However, the use of optical imagery is affected by cloud cover, and, consequently, all of the publications mentioned that images that are free of clouds should be selected as much as possible.
In addition to photointerpretation, supervised classifications were frequently chosen and maximum likelihood was the most-used parametric classifier. This finding strongly agrees with a remote sensing literature meta-analysis carried out by [104], in which 32% of the reviewed articles employed maximum likelihood algorithms, although the highest accuracies were obtained with machine learning methods. This dominance is attributed to the wide availability of these algorithms in conventional remote sensing image-processing software packages. This argument was verified by [105]: these authors stated that the uncertainties regarding the effective use and implementation of machine learning techniques were the main barrier to their use.
Another clear outcome is that researchers adopting the first step of the STEPLand workflow seemed to be aware of the classification issues in differentiating a high number of LUC categories. Conversely, according to the classification accuracy, some LUCC maps using the post-classification technique showed low accuracies (below 80%). This is a technical issue that merits a more in-depth discussion.
A characteristic of the second step, LUCC modeling, is the inclusion, in the materials or data sections, of a table showing the explanatory or independent variables or driving forces used in the model. From our analyzed papers, the table in [106] includes the variable name and a small description indicating the corresponding unit of measurement (proportion, percentage, or index, among others). The tables of other later works also provided the source, the scale, and the model used (raster or vector).
Our review indicates that LUCC modeling is a very important stage in STEPLand because 51.6% of works applied quantitative techniques for identifying the relationship between LUCC and the driving forces [107,108]. The most-used quantitative procedure was regression (38.0%), of which logistic regression was the most frequent, followed by linear regression. Our results agree with earlier studies; as stated by [109], multiple regression is frequently used by ecologists and biologists to identify factors influencing response variables, such as species richness or occurrence. Another study [110] noted that regression models are one of the most-common techniques for identifying relationships between variables, including drivers, and modeling transition potentials. They also stated that logistic regressions were the dominant statistical tool for describing relationships between a dichotomous variable (the categorical nature of LUCC) and a set of explanatory drivers. Some solving methods were applied to avoid criticisms about multicollinearity and spatial autocorrelation [60], such as random selection of pixels, generalization/aggregation of pixel size (coarser resolution), correlation coefficients, or analysis of variance [111].
Furthermore, this review found that the combined uses of biophysical, socio-economic, and distance variables are the most-common drivers applied at the pixel scale. Nevertheless, some studies do not explain, in detail or at all, the combined use of different sources, units of measurement, and scales, i.e., the interrelationships between the driving forces. This is even worse when administrative boundaries are involved [35], although the method chosen to combine them is determinant in the acquisition of useful results.
In step 3, socio-environmental consequences were mainly analyzed, followed by landscape and ecosystem services. Biodiversity and disasters were the topics most studied using generally qualitative models, whereas quantitative models were predominant in the remaining contributions. In ecosystem services, the benefit transfer method was the most applied, followed by the InVEST model. According to [112], this latter model, together with others (e.g., ARIES or CLASSIC), has two problems: it fails to fully consider or simplify the ecological process, and its evaluation accuracy is still controversial.
Finally, in future scenarios, in 2005, Ref. [113] asserted that the Markov-type models were some of the most common methods for predicting change among various categorical states. This clear preponderance can be explained by their spatial and temporal robustness because GIS and RS data can be efficiently incorporated to define the initial conditions to parameterize Markov models, calculate transition probabilities, and determine the neighborhood rules with transitional potential maps [114].

6. Conclusions

STEPLand is a consolidated framework within LCS that has been applied by experts worldwide, mainly from the geographical and natural sciences. The high diversity of journals and countries involved in the research shows that it has become relatively ubiquitous. Issues related to LUCC have attracted interest among a wide variety of researchers, ranging from those who wish to model spatial and temporal patterns of land conversion, to those who aim to understand causes and consequences of land transformation. Therefore, STEPLand facilitates communications and interactions among researchers from different fields, and between researchers and policy makers.
Based on this review, we conclude that step 1 of STEPLand research is characterized by:
  • Researchers mainly involved in geographic or environmental sciences.
  • Articles published in journals mainly specialized in ecology, geography, or environmental sciences, often having a high impact factor.
  • RS data as the main EO source used to obtain patterns, of which, Landsat images were the most used, after reducing the number of images included in the study due to the presence of clouds. Use of the Landsat open (free) archive has been essential to the spread of the STEPLand framework.
  • A relatively long horizon (over 30 years), which was the most-used temporal extent.
  • Supervised methods, in general, and the maximum likelihood classifier, in particular, which were predominant in the automatic classifications, with LUC legends primarily composed of 6 to 10 classes.
  • Post-classification methods (with transition matrix), which were the predominant tool in LUCC analysis.
Step 2 was characterized by:
  • The combination of analysis of patterns and dynamics, and modeling.
  • Quantitative spatial modeling based on statistics, mainly using regressions, headed by LOR and LIR.
  • A dependent variable extracted from LUC or LUCC, with a considerable weight assigned to forest variables.
  • The combination of biophysical, socio-economic, and distance variables as the main driving forces, in which data integration was very important.
  • Pixels from regular lattices as the main spatial domain.
Step 3 was characterized by:
  • Analysis of general socio-environmental consequences using qualitative methods, mainly related to biodiversity and natural disasters, and quantitative models, although very diverse, applied in the rest of topics. Surveys have been used to determine driving forces and the perception of socio-environmental consequences for local people. Spatially combining RS with other data can be difficult.
  • A high weight attached to the analysis of LUCC landscape consequences, mainly using the FRAGSTATS software for calculating, at least in many cases, the number of patches and the mean patch size.
  • In the analysis of ecosystem services, the application of the benefit transfer method, which was the most common model given its feasibility at regional or national scales to develop a quick assessment and the low cost of collecting the required primary data.
Finally, step 4 was characterized by:
  • A clear predominance of Markov chain (between 2000 and 2008) and Markov chain–cellular automata models.
  • A substantial weight of CA in other options, such as in the MOLAND model.
  • The more recent application of other models, such as CLUMondo or Land Change Modeler.
To conclude, a future scenario for STEPLand includes using improved spatial and temporal resolutions of satellite imagery, which may be implemented by adopting the free Sentinel imagery or with the contributions of drones. This may provide a more refined modeling of LUCC and the resulting socio-environmental consequences. Another challenge is the synergistic application of the four STEPLand stages, which was not documented in the sample of papers reviewed in this work.

Author Contributions

All authors were involved in the research conceptualization, data processing, and manuscript writing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Appendix A contains the references of the 250 selected articles analyzed in this review.
  • Abd El-Kawya, O.R.; Rød, J.K.; Ismail, H.A.; Suliman, A.S. Land use and land cover change detection in the western Nile delta of Egypt using remote sensing data. Appl. Geogr. 2011, 31, 483–494.
  • Abrantes, P.; Fontes, I.; Gomes, E.; Rocha, J. Compliance of land cover changes with municipal land use planning: Evidence from the Lisbon metropolitan region (1990–2007). Land Use Pol. 2016, 51, 120–134.
  • Acheampong, M.; Yu, Q.; Enomah, L.D.; Anchang, J.; Eduful, M. Land use/cover change in Ghana’s oil city: Assessing the impact of neoliberal economic policies and implications for sustainable development goal number one—A remote sensing and GIS approach. Land Use Pol. 2018, 73, 373–384.
  • Adhikari, A., Hansen, A.J. Land use change and habitat fragmentation of wildland ecosystems of the North Central United States. Landsc. Urban Plann. 2018, 177, 196–216.
  • Akhtar, F.; Awan, U.K.; Tischbein, B.; Liaqat, U.W. A phenology based geo-informatics approach to map land use and land cover (2003–2013) by spatial segregation of large heterogenic river basins. Appl. Geogr. 2017, 88, 48–61.
  • Akinyemi, F.O. Land change in the central Albertine rift: Insights from analysis and mapping of land use-land cover change in north-western Rwanda. Appl. Geogr. 2017, 87, 127–138.
  • Alphan, H. Classifying land cover conversions in coastal wetlands in the Mediterranean: pairwise comparisons of Landsat images. Land Degrad. Dev. 2012, 23, 278–292.
  • Alphan, H. Land-use change and urbanization of Adana, Turkey. Land Degrad. Dev. 2003, 14, 575–586.
  • Alvarez-Martinez, J.M.; Suárez-Seoanea, S.; De Luis Calabuig, E. Modelling the risk of land cover change from environmental and socio-economic drivers in heterogeneous and changing landscapes: The role of uncertainty. Landsc. Urban Plann. 2011, 101, 108–119.
  • Amsalu, A.; Stroosnijder, L.; de Graaff, J. Long-term dynamics in land resource use and the driving forces in the Beressa watershed, highlands of Ethiopia. J. Environ. Manag. 2007, 83, 448–459.
  • Ariti, A.T.; van Vliet, J.; Verburg, P.H. Land-use and land-cover changes in the Central Rift Valley of Ethiopia: Assessment of perception and adaptation of stakeholders. Appl. Geogr. 2015, 65, 28–37.
  • Arowolo, A.O.; Deng, X. Land use/land cover change and statistical modelling of cultivated land change drivers in Nigeria. Reg. Environ. Change 2018, 18, 247–259.
  • Arowolo, A.O.; Deng, X.; Olatunji, O.A.; Obayelu, A.E. (2018) Assessing changes in the value of ecosystem services in response to land-use/land-cover dynamics in Nigeria. Sci. Total Environ. 2018, 636, 597–609.
  • Auch, R.F.; Napton, D.E.; Kambly, S.; Moreland, T.R.; Sayler, K.L. The driving forces of land change in the Northern Piedmont of the United States. Geogr. Rev. 2012, 102, 53–75.
  • Bakker, M.M.; Govers, G.; Kosmas, C.; Vanacker, V., van Oost, K., Rounsevell, M. (2005) Soil erosion as a driver of land-use change. Agric. Ecosyst. Environ. 2005, 105, 467–481.
  • Bakr, N.; Weindorf, D.C.; Bahnassy, M.H.; Marei, S.M.; El-Badawi, M.M. Monitoring land cover changes in a newly reclaimed area of Egypt using multi-temporal Landsat data. Appl. Geogr. 2010, 30, 592–605.
  • Başkent, E.Z.; Kadioğullari, A.I. Spatial and temporal dynamics of land use pattern in Turkey: A case study in Inegöl. Landsc. Urban Plann. 2007, 81, 316–327.
  • Basse, R.M.; Omrani, H.; Charif, O.; Gerber, P.; Bódis, K. Land use changes modelling using advanced methods: Cellular automata and artificial neural networks. The spatial and explicit representation of land cover dynamics at the cross-border region scale. Appl. Geogr. 2014, 53, 160–171.
  • Batunacun; Nendel, C.; Hu, Y.; Lakes, T. Land-use change and land degradation on the Mongolian Plateau from 1975 to 2015—A case study from Xilingol, China. Land Degrad. Dev. 2018, 29, 1595–1606.
  • Batunacun; Wieland, R.; Lakes, T.; Yunfeng, H.; Nendel, C. Identifying drivers of land degradation in Xilingol, China, between 1975 and 2015. Land Use Pol. 2019, 83, 543–559.
  • Baumann, M.; Kuemmerle, T.; Elbakidze, M.; Ozdogan, M.; Radeloff, V.; Keuler, N.S.; Prishchepov, A.V.; Kruhlov, I.; Hostert, P. Patterns and drivers of post-socialist farmland abandonment in Western Ukraine. Land Use Pol. 2011, 28, 552–562.
  • Beilin, R.; Lindborg, R.; Stenseke, M.; Pereira, H.M.; Llausàs, A.; Slätmo, E.; Cerqueira, Y.; Navarro, L.; Rodrigues, P.; Reichelt, N.; Munro, N.; Queiroz, C. Analysing how drivers of agricultural land abandonment affect biodiversity and cultural landscapes using case studies from Scandinavia, Iberia and Oceania. Land Use Pol. 2014, 36, 60–72.
  • Bekele, B.; Wu, W.; Legesse, A.; Temesgen, H.: Yirsaw, E. Random and systematic land use/land cover transitions in semi-arid landscapes of Ethiopian Central Rift Valley lakes region (East Africa). Appl. Ecol. Environ. Res. 2018, 16, 3993–4014.
  • Bekele, B.; Wu, W.; Legesse, A.; Temesgen, H.; Yirsaw, E. Socio-environmental impacts of land use/cover change in Ethiopian Central Rift Valley lakes region, East Africa. Appl. Ecol. Environ. Res. 2018, 16, 6607–6632.
  • Bender, O.; Boehmer, H.J.; Jens, D.; Schumacher, K.P. Using GIS to analyse long-term cultural landscape change in Southern Germany. Landsc. Urban Plann. 2005, 70, 111–125.
  • Berihun, M.L.; Tsunekawa, A.; Haregeweyn, N.; Meshesha, D.T.; Adgo, E.; Tsubo, M.; Masunaga, T.; Fenta, A.A.; Sultan, D.; Yibeltala, M. Exploring land use/land cover changes, drivers and their implications in contrasting agro-ecological environments of Ethiopia. Land Use Pol. 2019, 87. https://doi.org/10.1016/j.landusepol.2019.104052.
  • Beuchle, R.; Grecchi, R.C.; Shimabukuro, Y.E.; Seliger, R.; Eva, H.D.; Sano, E.; Achard, F. Land cover changes in the Brazilian Cerrado and Caatinga biomes from 1990 to 2010 based on a systematic remote sensing sampling approach. Appl. Geogr. 2015, 58, 116–127.
  • Biazin, B.; Sterk, G. Drought vulnerability drives land-use and land cover changes in the Rift Valley dry lands of Ethiopia. Agric. Ecosyst. Environ. 2013, 164, 100–113.
  • Bičıík, I.; Jeleček, L.; Štěpánek, V. Land-use changes and their social driving forces in Czechia in the 19th and 20th Centuries. Land Use Pol. 2001, 18, 65–73.
  • Bieling, C.; Plieninger, T.; Schaich, H. Patterns and causes of land change: empirical results and conceptual considerations derived from a case study in the Swabian Alb, Germany. Land Use Pol. 2013, 35, 192–203.
  • Bracchetti, L.; Carotenuto, L.; Catorci, A. Land-cover changes in a remote area of central Apennines (Italy) and management directions. Landsc. Urban Plann. 2012, 104, 157–170.
  • Braimoh, A.K.; Onishi, T. Spatial determinants of urban land use change in Lagos, Nigeria. Land Use Pol. 2007, 24, 502–515.
  • Brown, D.G.; Pijanowski, B.C.; Duh, J.D. Modeling the relationship between land use and land cover on private lands in the Upper Midwest, USA. J. Environ. Manag. 2000, 59, 247–263.
  • Brunori, E.; Salvati, L.; Mancinelli, R.; Smiraglia, D.; Biasi, R. Multi-temporal land use and cover changing analysis: the environmental impact in Mediterranean area. Int. J. Sustain. Dev. World Ecol. 2017, 24, 276–288.
  • Bürgi, M.; Turner, M.G. Factors and processes shaping land cover and land cover changes along the Wisconsin River. Ecosystems 2002, 5, 184–201.
  • Cabral, A.I.R.; Lagos Costa, F. Land cover changes and landscape pattern dynamics in Senegal and Guinea Bissau borderland. Appl. Geogr. 2017, 82, 115–128.
  • Calvo-Iglesias, M.S.; Fra-Paleo, U.; Díaz-Valera, R.A. Changes in farming system and population as drivers of land cover and landscape dynamics: the case of enclosed and semi-openfield systems in Northern Galicia (Spain). Landsc. Urban Plann. 2009, 90, 168–177.
  • Campbell, D.J.; Lusch, D.P.; Smucker, T.A.; Wangui, E.E. (2005) Multiple methods in the study of driving forces of land use and land cover change: a case study of SE Kajiado District, Kenya. Hum. Ecol. 2005, 33, 763–794.
  • Carmel, Y.; Kadmon, R.; Nirel, R. Spatiotemporal predictive models of Mediterranean vegetation dynamics. Ecol. Appl. 2001, 11, 268–280.
  • Carmona, A.; Nahuelhual, L. Combining land transitions and trajectories in assessing forest cover change. Appl. Geogr. 2012, 32, 904–915.
  • Chakraborty, A.; Sachdeva, K.; Joshi, P.K. Mapping long-term land use and land cover change in the central Himalayan region using a tree-based ensemble classification approach. Appl. Geogr. 2016, 74, 136–150.
  • Cegielska, K.; Noszczyk, T.; Kukulska, A.; Szylar, M.; Hernik, J.; Dixon-Gough, R.; Jombach, S.; Valánszki, I.; Kovács, K.F. Land use and land cover changes in post-socialist countries: some observations from Hungary and Poland. Land Use Pol. 2018, 78, 1–18.
  • Celine, E.; Philippe, M.; Astrid, V.; Catherine, B.; Musampa, C.; Pierre, D. National forest cover change in Congo Basin: deforestation, reforestation, degradation and regeneration for the years 1990, 2000 and 2005. Glob. Change Biol. 2013, 19, 1173–1187. https://doi.org/10.1111/gcb.12092.
  • Celio, E.; Koellner, T.; Grêt-Regamey, A. Modeling land use decisions with Bayesian networks: Spatially explicit analysis of driving forces on land use change. Environ. Model Softw. 2014, 52, 222–233.
  • Chen, W.; Chi, G.; Li, J. The spatial association of ecosystem services with land use and land cover change at the county level in China, 1995–2015. Sci. Total Environ. 2019, 669, 459–470.
  • Chen, Y.; Li, X.; Liu, X.; Ai, B. (2013) Analyzing land-cover change and corresponding impacts on carbon budget in a fast developing sub-tropical region by integrating MODIS and Landsat TM/ETM+ images. Appl. Geogr. 2013, 45, 10–21.
  • Chi, W.; Zhao, Y.; Kuang, W., He, H. Impacts of anthropogenic land use/cover changes on soil wind erosion in China. Sci. Total Environ. 2019, 668, 204–215.
  • Cifaldi, R.L.; Allan, J.D.; Duh, J.D.; Brown, D.G. Spatial patterns in land cover of exurbanizing watersheds in southeastern Michigan. Landsc. Urban Plan. 2004, 66, 107–123.
  • Clement, F.; Orange, D.; Williams, M.; Mulley, C.; Epprecht, M. Drivers of afforestation in Northern Vietnam: Assessing local variations using geographically weighted regression. Appl. Geogr. 2009, 29, 561–576.
  • Cousins, S.A.O. Analysis of land-cover transitions based on 17th and 18th century cadastral maps and aerial photographs. Landsc. Ecol. 2001, 16, 41–54.
  • Crespin, S.J.; Simonetti, J.A. Loss of ecosystem services and the decapitalization of nature in El Salvador. Ecosyst. Serv. 2016, 17, 5–13.
  • Da, F.; Chen, X.; Qi, J. Spatiotemporal characteristic of land use/land cover change in the Middle and Lower reaches of Shule River Basin based on an intensity analysis. Sustainability 2019, 11, 1360. https://doi.org/10.3390/su11051360.
  • Dadashpoor, H.; Azizi, P.; Moghadas, M. Land use change, urbanization, and change in landscape pattern in a metropolitan area. Sci. Total Environ. 2019, 655, 707–719.
  • Daunt, A.B.P.; Silva, T.S.F. Beyond the park and city dichotomy: land use and land cover change in the northern coast of São Paulo (Brazil). Landsc. Urban Plann. 2019, 189, 352–361.
  • Devátý, J.; Dostál, T.; Hösl, R.; Krása, J.; Strauss, P. Effects of historical land use and land pattern changes on soil erosion—Case studies from Lower Austria and Central Bohemia. Land Use Pol. 2019, 82, 674–685.
  • Dewan, A.M.; Yamaguchi, Y.; Rahman, M.Z. Dynamics of land use/cover changes and the analysis of landscape fragmentation in Dhaka Metropolitan, Bangladesh. GeoJournal 2012, 77, 315–330.
  • Dewan, A.M.; Yamaguchi, Y. Land use and land cover change inn Greater Dhaka, Bangladesh: using remote sensing to promote sustainable urbanization. Appl. Geogr. 2009, 29, 390–401.
  • Díaz, G.I.; Nahuelhual, L.; Echeverría, C.; Marín, S. Drivers of land abandonment in Southern Chile and implications for landscape planning. Landsc. Urban Plann. 2011, 99, 207–217.
  • Díaz-Caravantes, R.E.; Sánchez-Flores, E. Water transfer effects on peri-urban land use/land cover: A case study in a semi-arid region of Mexico. Appl. Geogr. 2011, 31, 413–425.
  • Disperati, L.; Virdis, S.G.P. Asessment of land-use and land-cover changes from 1965 to 2014 in Tam Ging-Cau Hai Lagoon, central Vietnam. Appl. Geogr. 2015, 58, 48–64.
  • Dou, P.; Chen, Y. Dynamic monitoring of land-use/land-cover change and urban expansion in Shenzhen using Landsat imagery from 1988 to 2015. Int. J. Remote Sens. 2017, 38, 5388–5407.
  • Ektvedt, T.M.; Vetaas, O.R.; Lundberg, A. Land-cover changes during the past 50 years in the semi-arid tropical forest region of Northern Peru. Erdkunde 2012, 66, 57–75.
  • Espindola, G.M.; de Aguiar, A.P.D.; Pebesma, E.; Câmara, G.; Fonseca, L. Agricultural land use dynamics in the Brazilian Amazon based on remote sensing and census data. Appl. Geogr. 2012, 32, 240–252.
  • Falcucci, A.; Maiorano, L.; Boitani, L. Changes in land-use/land-cover patterns in Italy and their implications for biodiversity conservation. Landsc. Ecol. 2007, 22, 617–631.
  • Fasona, M.; Omojola, A. Land cover change and land degradation in parts of the southwest coast of Nigeria. Afr. J. Ecol. 2009, 47 (Suppl. 1), 30–38.
  • Figueroa, F.; Sánchez-Cordero, V.; Meave, J.A., Trejo, I. Socioeconomic context of land use and land cover change in Mexican biosphere reserves. Environ. Conserv. 2009, 36, 180–191.
  • Flamenco-Sandoval, A.; Martinez Ramos, M.; Masera, O.R. Assessing implications of land-use and land-cover change dynamics for conservation of a highly diverse tropical rain forest. Biol. Conserv. 2007, 138, 131–145.
  • Frondoni, R.; Mollo, B.; Capotorti, G. A landscape analysis of land cover change in the Municipality of Rome (Italy): spatio-temporal characteristics and ecological implications of land cover transitions from 1954 to 2001. Landsc. Urban Plan. 2011, 100, 117–128.
  • Gao, J.; Li, F.; Gao, H.; Zhou, C.; Zhang, X. The impact of land-use change on water-related ecosystem services: a study of the Guishui River basin, Beijing, China. J. Clean. Prod. 2017, 163, S148–S155.
  • Gao, J.; Liu, Y.; Chen, Y. Land cover changes during agrarian restructuring in Northeast China. Appl. Geogr. 2006, 26, 312–322.
  • Gao, C.; Zhou, P. Spatial driving forces of dominant land use/land cover transformations in the Dongjiang River watershed, Southern China. Environ. Monit. Assess. 2016, 188, 84. https://doi.org/10.1007/s10661-015-5088-z.
  • Garcia-Frapolli, E.; Ayala-Orozco, B.; Bonilla-Moheno, M.; Espadas-Manrique, C.; Ramos-Fernández, G. Biodiversity conservation, traditional agriculture and ecotourism: land cover/land use change projections for a natural protected area in the northeastern Yucatan Peninsula, Mexico. Landsc. Urban Plan. 2007, 83, 137–153.
  • Gautam, A.; Webb, E.L.; Shivakoti, G.P.; Zoebisch, M.A. Land use dynamics and landscape change pattern in a mountain watershed in Nepal. Agric. Ecosyst. Environ. 2003, 99, 83–96.
  • Gellrich, M.; Baur, P.; Robinson, B.H.; Bebi, P. Combining classification tree analyses with interviews to study why sub-alpina grasslands sometimes revert to forest: a case study from the Swiss Alps. Agric. Syst. 2008, 96, 124–138.
  • Gellrich, M.; Zimmermann, N.E. Investigating the regional-scale pattern of agricultural land abandonment in the Swiss mountains: a spatial statistical modelling approach. Landsc. Urban Plan. 2007, 79, 65–76.
  • Gellrich, M.; Baur, P.; Koch, B.; Zimmermann, N.E. Agricultural land abandonment and natural forest re-growth in the Swiss mountains: A spatially explicit economic analysis. Agric. Ecosyst. Environ. 2007, 118, 93–108.
  • Geoghegan, J.; Cortina Villar, S.; Klepeis, P.; Mendoza, P.M.; Ogneva-Himmelberger, Y.; Chowdhury, R.R.; Turner II, B.L.; Vance, C. Modeling tropical deforestation in the southern Yucatán peninsular region: comparing survey and satellite data. Agric. Ecosyst. Environ. 2001, 85, 25–46.
  • Getnet, M.; Hengsdijk, H.; van Ittersum, M. Disentangling the impacts of climate change, land use change and irrigation on the Central Rift Valley water system of Ethiopia. Agric. Water Manag. 2014, 137, 104–115.
  • Gilani, H.; Shrestha, H.L.; Murthy, M.S.R.; Phuntso, P.; Pradhan, S.; Bajracharya, B.; Shrestha, B. Decadal land cover change dynamics in Bhutan. J. Environ. Manag. 2015, 148, 91–100.
  • Gómez-Mendoza, L.; Vega-Peña, E.; Ramirez, M.I.; Palacio-Prieto, J.L.; Galicia, L. Projecting land-use change processes in the Sierra Norte of Oaxaca, Mexico. Appl. Geogr. 2006, 26, 276–290.
  • Gong, C.; Yu, S.; Joesting, H.; Chen, J. Determining socioeconomic drivers of urban forest fragmentation with historical remote sensing images. Landsc. Urban Plan. 2013, 117, 57–65.
  • Grecchi, R.C.; Gwyn, Q.H.; Bénie, G.B.; Formaggio, A.R.; Fahl, F.C. Land use and land cover changes in a Brazilian Cerrado: a multidisciplinary approach to assess the impacts of agricultural expansion. Appl. Geogr. 2014, 55, 300–312.
  • Guida-Johnson, B.; Zuleta, G.A. Land-use and land-cover change and ecosystem loss in the Espinal ecoregion, Argentina. Agric. Ecosyst. Environ. 2013, 181, 31–40.
  • Haregeweyn, N.; Fikadu, G.; Tsunekawa, A.; Tsubo, M.; Meshesha, D.T. The dynamics of urban expansion and its impacts on land use/land cover change and small-scale farmers living near the urban fringe: a case study of Bahir Dar, Ethiopia. Landsc. Urban Plan. 2012, 106, 149–157.
  • Hasan, S.; Shi, W.; Zhu, X.; Abbas, S. Monitoring of land use/land cover and socioeconomic changes in South China over the last three decades using Landsat and nightime light data. Remote Sens. 2019, 11, 1658.
  • Hauser, L.T.; Vu, G.N.; Nguyen, B.A.; Dade, E.; Nguyen, H.M.; Nguyen, T.T.Q.; Le, T.Q.; Vu, L.H.; Tong, A.T.H.; Pham, H.V. Uncovering the spatio-temporal dynamics of land cover change and fragmentation of mangroves in the Ca Mau peninsula, Vietnam using multi-temporal SPOT satellite imagery (2004–2013). Appl. Geogr. 2017, 86, 197–207.
  • Heidarlou, H.B.; Shafiei, A.B.; Erfanian, M.; Tayyebi, A.; Alijanpour, A. Effects of preservation policy on land use changes in Iranian Northern Zagros forests. Land Use Pol. 2019, 81, 76–90.
  • Hersperger, A.M.; Bürgi, M. Going beyond landscape change description: quantifying the importance of driving forces of landscape change in a Central Europe case study. Land Use Pol. 2009, 26, 640–648.
  • Hietel, E.; Waldhardt, R.; Otte, A. Linking socio-economic factors, environment and land cover in the German Highlands, 1945–1999. J. Environ. Manag. 2005, 75, 133–143.
  • Hu, Y.; Batunacun; Zhen, L.; Zhuang, D. Assessment of land-use and land-cover change in Guangxi, China. Sci. Rep. 2019, 9, 2189. https://doi.org/10.1038/s41598-019-38487-w.
  • Huq, N.; Bruns, A.; Ribbe, L. Interactions between freshwater ecosystem services and land cover changes in southern Bangladesh: A perspective from short-term (seasonal) and long-term (1973–2014) scale. Sci. Total Environ. 2019, 650 Pt 1, 132–143.
  • Ierodiaconou, D.; Laurenson, L.; Leblanc, M.; Stagnitti, F.; Duff, G.; Salzman, S.; Versace, V. The consequences of land use change on nutrient exports: a regional scale assessment in south-west Victoria, Australia. J. Environ. Manag. 2005, 74, 305–316.
  • Jenerette, G.D.; Wu, J. Analysis and simulation of land-use change in the central Arizona-Phoenix region, USA. Landsc. Ecol. 2001, 16, 611–626.
  • Jepson, W. A disappearing biome? Reconsidering land-cover change in the Brazilian savanna. Geogr. J. 2005, 171, 99–111.
  • Jiyuan, L.; Zengxiang, Z.; Xinliang, X.; Wenhui, K.; Wancun, Z.; Shuwen, Z.; Rendong, L.; Changzhen, Y.; Dongsheng, Y.; Shixin, W.; Nan, J. Spatial patterns and driving forces of land use change in China during the early 21st century. J. Geogr. Sci. 2010, 20, 483–494.
  • Kamusoko, C.; Aniya, M.; Adi, B.; Manjoro, M. Rural sustainability under threat in Zimbabwe—Simulation of future land use/cover changes in the Bindura district based on the Markov-cellular automata model. Appl. Geogr. 2009, 29, 435–447.
  • Kamusoko, C.; Aniya, M. Land use/cover change and landscape fragmentation analysis in the Bindura district, Zimbabwe. Land Degrad. Dev. 2007, 18, 221–233.
  • Kelarestaghi, A.; Jafarian Jeloudar, Z. Land use/cover change and driving force a198-207nalyses in parts of northern Iran using RS and GIS techniques. Arab. J. Geosci. 2011, 4, 401–411. https://doi.org/10.1007/s12517-009-0078-5.
  • Kertész, A.; Nagy, L.A.; Balázs, B. Effect of land use change on ecosystem services in Lake Balaton Catchment. Land Use Pol. 2019, 80, 430–438.
  • Killeen, T.J.; Calderon, V.; Soria, L.; Quezada, B.; Steininger, M.K.; Harper, G.; Solórzano, L.A.; Tucker, C.J. Thirty Years of Land-cover Change in Bolivia. Ambio 2007, 36, 600–606.
  • Kindu, M.; Schneider, T.; Teketay, D.; Knoke, T. Land use/land cover change analysis using object-based classification approach in Munessa-Shashemene landscape of the Ethiopian Highlands. Remote Sens. 2013, 5, 2411–2435.
  • Kindu, M.; Schneider, T.; Teketay, D.; Knoke, T. Changes of ecosystem service values in response to land use/land cover dynamics in Munessa-Shashemene landscape of the Ethiopian highlands. Sci. Total Environ. 2016, 547, 137–147.
  • Kindu, M.; Schneider, T.; Teketay, D.; Knoke, T. Drivers of land use/land cover changes in Munessa-Shashemene landscape of the south-central highlands of Ethiopia. Environ. Monit. Assess. 2015, 187, 452.
  • Kleemann, J.; Baysal, G.; Bulley, H.N.N.; Fürst, C. Assessing driving forces of land use and land cover change by a mixed-method approach in north-eastern Ghana, West Africa. J. Environ. Manag. 2017, 196, 411–442.
  • Kolb, M.; Mas, J.F.; Galicia, L. Evaluating drivers of land-use change and transition potential models in a complex landscape in Southern Mexico. Int. J. Geogr. Inform. Sci. 2013, 27, 1804–1827.
  • Kolb, M.; Galicia, L. Challenging the linear forestation narrative in the Neo-tropic: regional patterns and processes of deforestation and regeneration in southern Mexico. Geogr J. 2012, 178, 147–161.
  • Kolios, S.; Stylos, C.D. Identification of land cover/land use changes in the greater area of the Prevenza peninsula in Greece using Landsat satellite data. Appl. Geogr. 2013, 40, 150–160.
  • Krajewski, P.; Solecka, I.; Mrozik, K. Forest landscape change and preliminary study on its driving forces in Ślęża Landscape Park (Southwestern Poland) in 1883–2013. Sustainability 2018, 10, 4526. https://doi.org/10.3390/su10124526.
  • Kuang, W.; Liu, J.; Dong, J.; Chi, W.; Zhang, C. The rapid and massive urban and industrial land expansion in China between 1990–2010: a CLUD-based analysis of their trajectories, patterns and drivers. Landsc. Urban Plan. 2016, 145, 21–33.
  • Kukkonen, M.; Käyhkö, N. Spatio-temporal analysis of forest change in contrasting land use regimes of Zanzibar, Tanzania. Appl. Geogr. 2014, 55, 193–202.
  • Lai, S.; Leone, F.; Zoppi, C. Land cover changes and environmental protection: A study based on transition matrices concerning Sardinia (Italy). Land Use Pol. 2017, 67, 126–150.
  • Leh, M.; Bajwa, S.; Chaubey, I. Impact of land use change on erosion risk: an integrated remote sensing, geographic information system and modeling methodology. Land Degrad. Dev. 2013, 24, 409–421.
  • Li, X.; Wang, Y.; Li, J.; Lei, B. Physical and socioeconomic driving factors of land-use and land-cover changes: a case study of Wuhan city, China. Discrete Dyn. Nat. Soc. 2016, 2016, 8061069. https://doi.org/10.1155/2016/8061069.
  • Li, G.; Sun, S.; Fang, C. The varying driving forces of urban expansion in China: insights from a spatial-temporal analysis. Landsc. Urban Plann. 2018, 174, 63–77.
  • Lieskovský, J.; Bezák, P.; Spulerová, J.; Lieskovský, T.; Koleda, P.; Dobrovodská, M., Bürgi, M.; Gimmi, U. The abandonment of traditional agricultural landscape in Slovakia—Analysis of extent and driving forces. J. Rural Stud. 2015, 37, 75–84.
  • Lindström, S.; Mattsson, E.; Nissanka, S.P. Forest cover change in Sri Lanka: the role of small scale farmers. Appl. Geogr. 2012, 34, 680–692.
  • Liping, C.; Yujun, S.; Saeed, S. Monitoring and predicting land use and land cover changes using remote sensing and GIS techniques—A case study of a hilly area, Jiangle, China. PLoS One 2018, 13. https://doi.org/10.1371/journal.pone.0200493.
  • Liu, J.; Zhan, J.; Deng, X. Spatio-temporal patterns and driving forces of urban land expansion in China during the economic reform era. Ambio 2005, 34, 450–455.
  • Liu, M.; Tian, H. China’s land cover and land use from 1700 to 2005: estimations from high-resolution satellite data and historical archives. Glob. Biogeochem. Cycles 2010, 24. https://doi.org/10.1029/2009GB003687.
  • Liu, J.; Kuang, W.; Zhang, Z.; Xu, X.; Qin, Y.; Ning, J.; Zhou, W.; Zhang, S.; Li, R.; Yan, C.; Wu, S.; Shi, X.; Jiang, N.; Yu, D.; Pan, X.; Chi, W. Spatiotemporal characteristics, patterns, and causes of land-use changes in China since the late 1980s. J. Geogr. Sci. 2014, 24, 195–210.
  • Liu, T.; Yang, X. Monitoring land changes in an urban area using satellite imagery, GIS and landscape metrics. Appl. Geogr. 2015, 56, 42–54.
  • Liu, Y.; Luo, T.; Liu, Z.; Kong, X.; Li, J.; Tan, R. A comparative analysis of urban and rural construction land use change and driving forces: Implications for urban-rural coordination development in Wuhan, Central China. Habitat Int. 2015, 47, 113–125.
  • Long, H.; Tang, G.; Li, X.; Heiling, G.K. Socio-economic driving forces of land-use change in Kunshan, the Yangtze River Delta economic area of China. J. Environ. Manag. 2007, 83, 351–364.
  • López, E.; Bocco, G.; Mendoza, M.; Duhau, E. Predicting land-cover and land-use change in the urban fringe—A case in Morelia city, Mexic. Landsc. Urban Plann. 2001, 55, 271–285.
  • Lorena, R.B.; Lambin, E.F. The spatial dynamics of deforestation and agent use in Amazon. Appl. Geogr. 2009, 29, 171–181.
  • Lundberg, A. Climate and land-use change as driving forces in lowland seminatural vegetation dynamics. Erdkunde 2011, 65, 335–353.
  • Luo, G.; Yin, C.; Chen, X.; Xu, W.; Lu, L. Combining system dynamics model and CLUE-S model to improve land use scenario analyses at regional scale: a case study of Sangong watershed in Xinjiang, China. Ecol. Complex 2010, 7, 198–207.
  • Ma, T.; Li, X.; Ding, S.; Zhou, F.; Cui, B. Four decades’ dynamics of coastal blue carbon storage driven by land use/land cover transformation under natural and anthropogenic processes in the Yellow River Delta, China. Sci. Total Environ. 2019, 655, 741–750.
  • Mallinis, G.M.; Koutsias, N.; Arianoutsou, M. Monitoring land use/land cover transformations from 1945 to 2007 in two peri-urban mountainous areas of Athens metropolitan area. Sci. Total Environ. 2014, 490, 262–278.
  • Mallinis, G.; Emmanoloudis, D.; Giannakopoulos, V.; Maris, F.; Koutsias, N. Mapping and interpreting historical land cover/land use changes in a Natura 2000 site using earth observation data: the case of Nestos delta, Greece. Appl. Geogr. 2011, 31, 312–320.
  • Mapedza, E.; Wright, J.; Fawcett, R. An investigation of land cover change in Mafungautsi Forest, Zimbabwe, using GIS and participatory mapping. Appl. Geogr. 2003, 23, 1–21.
  • Marshall, E.; Randhir, T.O. Spatial modeling of land cover change and watershed response using Markovian cellular automata and simulation. Water Resour. Res. 2008, 44. https://doi:10.1029/2006WR005514.
  • Mas, J-F.; Nogueira de Vasconcelos, R.; Franca-Rocha, W. Analysis of high temporal resolution land use/land cover trajectories. Land 2019, 8, 30. https://doi.org/10.3390/land8020030.
  • Mas, J-F.; Velázquez, A.; Reyes Díaz-Gallegos, J.; Mayorga-Saucedo, R.; Alcántara, C.; Bocco, G.; Castro, R.; Fernández, T.; Pérez-Vega, A. Assessing land use/cover changes: a nationwide multidate spatial database for Mexico. Int. J. Appl. Earth Obs. Geoinf. 2004, 5, 249–261.
  • Matasov, V.; Prishchepov, A.; Jepsen, M.R.; Müller, D. Spatial determinants and underlying drivers of land-use transitions in European Russia from 1770 to 2010. J. Land Use Sci. 2019, 14, 4–6, 362–377.
  • Mena, C.F.; Bilsborrow, R.E.; McClain, M.E. Socioeconomic drivers of deforestation in the Northern Ecuadorian Amazon. Environ. Manag. 2006, 37, 802–815.
  • Mendoza, J.E.; Etter, A. Multitemporal analsysis (1940–1996) of land cover changes in the southwestern Bogotá highplain (Colombia). Landsc. Urban Plann. 2002, 59, 147–158.
  • Mendoza, M.E.; López Granados, E.; Geneletti, D.; Pérez-Salicrup, D.R.; Salinas, V. Analysing land cover and land use change processes at watershed level: a multitemporal study in the Lake Cuitzeo Watershed, Mexico (1975–2003). Appl. Geogr. 2011, 31, 237–250.
  • Mendoza-González, G.; Martínez, M.L.; Lithgow, D.; Pérez-Maqueo, O.; Simonin, P. Land use change and its effects on the value of ecosystem services along the coast of the Gulf of Mexico. Ecol. Econ. 2012, 82, 23–32.
  • Meneses, B.M.; Reis, E.; Pereira, S.; Vale, M.J.; Reis, R. Understanding driving forces and implications associated with the land use and land cover changes in Portugal. Sustainability 2017, 9, 351.
  • Meneses, B.M.; Reis, R.; Vale, M.J.; Saraiva, R. Land use and land cover changes in Zêzdere watershed (Portugal)—Water quality implications. Sci. Total Environ. 2015, 527–528, 439–447.
  • Meshesha, D.T.; Tsunekawa, A.; Tsubo, M.; Ali, S.A.; Haregeweyn, N. Land-use change and its socio-environmental impact in Eastern Ethiopia’s highland. Reg. Environ. Change 2014, 14, 757–768.
  • Millington, J.D.A.; Perry, G.L.W.; Romero-Calcerrada, R. Regression techniques for examining land use/cover change: a case study of a Mediterranean landscape. Ecosystems 2007, 10, 562–578.
  • Mireille, N.M.; Mwangi, H.M.; Mwangi, J.K.; Gathenya, J.M. Analysis of land use change and its impact on the hydrology of Kakia and Esamburmbur sub-watersheds of Narok county, Kenya. Hydrology 2019, 6, 86. https://doi.org/10.3390/hydrology6040086.
  • Monteiro, A.T.; Fava, F.; Hiltbrunner, E.; Della Marianna, G.; Bocchi, S. Assessment of land cover changes and spatial drivers behind loss of permanent meadows in the lowlands of Italian Alps. Landsc. Urban Plann. 2011, 100, 287–294.
  • Mottet, A.; Ladet, S.; Coque, N.; Gibon, A. Agricultural land-use change and its drivers in mountain landscapes: A case study in the Pyrenees. Agric. Ecosyst. Environ. 2006, 114, 296–310.
  • Msofe, N.K.; Sheng, L.; Lyimo, J. Land use change trends and their driving forces in the Kilombero valley floodplain, southeastern Tanzania. Sustainability 2019, 11, 505. https://doi.org/10.3390/su11020505.
  • Mukhopadhyay, A.; Mondal, A.; Mukherjee, S.; Khatua, D.; Ghosh, S.; Mitra, D.; Ghosh, T. Forest cover change prediction using hybrid methodology of geoinformatics and Markov chain model: a case study on sub-Himalayan town Gangtok, India. J. Earth Syst. Sci. 2014, 123, 1349–1360.
  • Müller, D.; Sikor, T. Effects of postsocialist reforms on land cover and land use in South-Eastern Albania. Appl. Geogr. 2006, 26, 175–191.
  • Müller, D.; Zeller, M. Land use dynamics in the central highlands of Vietnam: a spatial model combining village survey data with satellite imagery interpretation. Agric. Econ. 2002, 27, 333–354.
  • Munthali, M.G.; Davis, N.; Adeola, A.M.; Botai, J.O.; Kamwi, J.M.; Chisale, H.L.W.; Orimoogunje, O.O.I. Local Perception of Drivers of Land-Use and Land-Cover Change Dynamics across Dedza District, Central Malawi Region. Sustainability 2019, 11, 832. https://doi.org/10.3390/su11030832.
  • Muriuki, G.; Seabrook, L.; McAlpine, C.; Jacobson, C.; Price, B.; Baxter, G. Land cover change under unplanned human settlements: a study of the Chyulu Hills squatters, Kenia. Landsc. Urban Plann. 2011, 99, 154–165.
  • Nagabhatla, N.; Padmanabhan, M.; Kühle, P.; Vishnudas, S.; Betz, L.; Niemeyer, B. LCLU as an entry point for transdisciplinary research—Reflections from an agriculture land use change study in South Asia. J. Environ. Manag. 2015, 148, 42–52.
  • Nahuelhual, L.; Carmona, A.; Lara, A.; Echeverría, C.; González, M.E. Land-cover change to forest plantations: proximate causes and implications for the landscape in south-central Chile. Landsc. Urban Plann. 2012, 107, 12–20.
  • Napton, D.E.; Auch, R.F.; Headley, R.; Taylor, J.L. Land changes and their driving forces in the Southeastern United States. Reg. Environ. Change 2010, 10, 37–53.
  • Narumalani, S.; Mishra, D.R.; Rothwell, R.G. Change detection and landscape metrics for inferring anthropogenic processes in the greater EFMO area. Remote Sens. Environ. 2004, 91, 478–489.
  • Nath, B.; Niu, Z.; Singh, R.P. Land use and land cover changes, and environment and risk evaluation of Dujiangyan city (SW China) using remote sensing and GIS techniques. Sustainability 2018, 10, 4631. https://doi.org/10.3390/su10124631.
  • Newman, M.E.; McLaren, K.P.; Wilson, B.S. Long-term socio-economic and spatial pattern drivers of land cover change in a Caribbean tropical moist forest, the Cockpit Country, Jamaica. Agric. Ecosyst. Environ. 2014, 186, 185–200.
  • Ningal, T.; Hartemink, A.E.; Bregt, A.K. Land use change and population growth in the Morobe Province of Papua New Guinea between 1975 and 2000. J. Environ. Manag. 2008, 87, 117–124.
  • Ouedraogo, I.; Barron, J.; Tumbo, S.D.; Kahimba, F.C. Land cover transition in Northern Tanzania. Land Degrad. Dev. 2016, 27, 682–692.
  • Ouedrago, I.; Mbow, C.; Balinga, M.; Neufeldt, H. Transitions in land use architecture under multiple human driving forces in a semi-arid zone. Land 2015, 4, 560–577.
  • Padonou, E.A.; Lykke, A.M.; Bachmann, Y.; Idohou, R.; Sinsin, B. Mapping changes in land use/land cover and prediction of future extension of bowé in Benin, West Africa. Land Use Pol. 2017, 69, 85–92.
  • Pan, Y.; Roth, A.; Yu, Z.; Doluschitz, R. The impact of variation in scale on the behavior of a cellular automata used for land use change modeling. Comput. Environ. Urban Syst. 2010, 34, 400–408.
  • Parcerisas, LL.; Marull, J.; Pino, J.; Tello, E.; Coll, F.; Basnou, C. (2012) Land use changes, landscape ecology and their socioeconomic driving forces in the Spanish Mediterranean coast (El Maresme County, 1850–2005). Environ. Sci. Policy 2012, 23, 120–132.
  • Parmenter, A.W.; Hansen, A.; Kennedy, R.E.; Cohen, W.; Langner, U.; Lawrence, R.; Maxwell, B.; Gallant, A.; Aspinall, R. Land use and land cover change in the Greater Yellowstone Ecosystem: 1975–1995. Ecol. Appl. 2003, 13, 687–703.
  • Pauleit, S.; Ennos, R.; Golding, Y. Modeling the environmental impacts of urban land use and land cover change—A study in Merseyside, UK. Landsc. Urban Plan. 2005, 71, 295–310.
  • Pazúr, R.; Bolliger, J. Land changes in Slovakia: past processes and future directions. Appl. Geogr. 2017, 85, 163–175.
  • Pelorosso, R.; Leone, A.; Boccia, L. Land cover and land use change in the Italian central Apennines: a comparison of assessment methods. Appl. Geogr. 2009, 29, 35–48.
  • Pérez-Vega, A.; Mas, J.F.; Ligmann-Zielinska, A. Comparing two approaches to land use/cover change modeling and their implications for the assessment of biodiversity loss in a deciduous tropical forest. Environ. Model Softw. 2012, 29, 11–23.
  • Petrov, L.O.; Lavalle, C.; Kasanko, M. Urban land use scenarios for a tourist region in Europe: applying the MOLAND model to Algarve, Portugal. Landsc. Urban Plan. 2009, 92, 10–23.
  • Phiri, D.; Morgenroth, J.; Xu, C. Long-term land cover change in Zambia: an assessment of driving forces. Sci. Total Environ. 2019, 697, 134206. https://doi.org/10.1016/j.scitotenv.2019.134206.
  • Phiri, D.; Morgenroth, J.; Xu, C. Four decades of land cover and forest connectivity study in Zambia—An object-based image analysis approach. Int. J. Appl. Earth Obs. Geoinf. 2019, 79, 97–109.
  • Pineda Jaimes, N.B.; Bosque Sendra, J.; Gómez Delgado, M.; Franco Plata, R. Exploring the drivers forces behind deforestation in the state of Mexico (Mexico) using geographically weighted regression. Appl. Geogr. 2010, 30, 576–591.
  • Pôças, I.; Cunha, M.; Pereira, L.S. Remote sensing based indicators of changes in a mountain rural landscape of Northeast Portugal. Appl. Geogr. 2011, 31, 871–880.
  • Polasky, S.; Nelson, E.; Pennington, D.; Johnson, K.A. The Impact of Land-Use Change on Ecosystem Services, Biodiversity and Returns to Landowners: A Case Study in the State of Minnesota. Environ. Resour. Econ. 2011, 48, 219–242.
  • Poyatos, R.; Latron, J.; Llorens, P. Land use and land cover change after agricultural abandonment. The case of a Mediterranean mountain area (Catalan Pre-Pyrenees). Mt. Res. Dev. 2003, 23, 362–368.
  • Popovici, E.A.; Kucsicsa, G.; Bălteanu, D.; Grigorescu, I.; Mitrică, B.; Dumitraşcu, M.; Damian, N. Past and future land use/cover flows related to agricultural lands in Romania. An assessment using CLUE-S model and CORINE land cover database. Carpathian J. Earth Environ. Sci. 2018, 13, 613–628.
  • Qasim, M.; Hubacek, K.; Termansen, M.; Khan, A. Spatial and temporal dynamics of land use pattern in District Swat, Hindu Kush Himalayan region of Pakistan. Appl. Geogr. 2011, 31, 820–828.
  • Quintero-Gallego, M.E.; Quintero-Angel, M.; Vila-Ortega, J.J. Exploring land use/land cover change and drivers in Andean mountains in Colombia: a case in rural Quindío. Sci. Total Environ. 2018, 634, 1288–1299.
  • Rasmussen, L.V.; Reenberg, A. Land use rationales in desert fringe agriculture. Appl. Geogr. 2012, 34, 595–605.
  • Reger, B.; Otte, A.; Waldhardt, R. Identifying patterns of land-cover change and their physical attributes in a marginal European landscape. Landsc. Urban Plann. 2007, 81, 104–113.
  • Regos, A.; Ninyerola, M.; Moré, G.; Pons, X. Linking land cover dynamics with driving forces in mountain landscape of the Northwestern Iberian Peninsula. Int. J. Appl. Earth Obs. Geoinf. 2015, 38, 1–14.
  • Reid, R.S.; Kruska, R.L.; Muthui, N.; Taye, A.; Wotton, S.; Wilson, C.J.; Mulatu, W. Land-use and land-cover dynamics in response to changes in climatic, biological and socio-political forces: the case of southwestern Ethiopia. Landsc. Ecol. 2000, 15, 339–355.
  • Rhemtulla, J.M.; Mladenoff, D.J.; Clayton, M.K. Regional land-cover conversion in the U.S. upper Midwest: magnitude of change and limited recovery (1850–1935–1993). Landsc. Ecol. 2007, 22, 57–75.
  • Röder, A.; Pröpper, M.; Stellmes, M.; Schneibel, A.; Hill, J. Assessing urban growth and rural land use transformations in across-border situation in Northern Namibia and Southern Angola. Land Use Pol. 2015, 42, 340–354.
  • Rojas, C.; Pino, J.; Basnou, C.; Vivanco, M. Assessing land-use and -cover changes in relation to geographic factors and urban planning in the metropolitan area of Concepción (Chile). Implications for biodiversity conservation. Appl. Geogr. 2013, 39, 93–103.
  • Romero-Calcerrada, R.; Perry, G.L.W. The role of land abandonment in landscape dynamics in the SPA ‘Encinares del río Alberche y Cofio’, Central Spain, 1984–1999. Landsc. Urban Plann. 2004, 66, 217–232.
  • Roy Chowdhury, R. Landscape change in the Calakmul Biosphere Reserve, Mexico: modeling the driving forces of smallholder deforestation in land parcels. Appl. Geogr. 2006, k, 129–152.
  • Rutherford, G.N.; Bebi, P.; Edwards, P.J.; Zimmermann, N.E. Assessing land-use statistics to model land cover change in a mountainous landscape in the European Alps. Ecol. Model. 2008, 212, 460–471.
  • Rutherford, G.N.; Guisan, A.; Zimmermann, N.E. Evaluating sampling strategies and logistic regression methods for modelling complex land cover changes. J. Appl. Ecol. 2007, 44, 414–424.
  • Salvati, L.; De Zuliani, E.; Sabbi, A.; Cancellieri, L.; Tufano, M.; Caneva, G.; Savo, V. Land-cover changes and sustainable development in a rural cultural landscape of central Italy: classical trends and counter-intuitive results. Int. J. Sustain. Dev. World Ecol. 2017, 24, 27–36. https://doi.org/10.1080/13504509.2016.1193778.
  • Salvati, L.; Munafo, M.; Gargiulo Morelli, V.; Sabbi, A. Low-density settlements and land use changes in a Mediterranean urban region. Landsc. Urban Plann. 2012, 105, 43–52.
  • Salvati, L.; Sabbi, A. Exploring long-term land cover changes in an urban region of southern Europe. Int. J. Sustain. Dev. World Ecol. 2011, 18, 273–282.
  • Saputra, M.H.; Lee, H.S. Prediction of land use and land cover changes for North Sumatra, Indonesia, using an artificial neural-network-based cellular automaton. Sustainability 2019, 11, 3024. https://doi.org/10.3390/su11113024.
  • Schneeberger, N.; Bürgi, M.; Hersperger, A.M.; Ewald, K.C. Driving forces and rates of landscape change as a promising combination for landscape change research—An application on the northern fringe of the Swiss Alps. Land Use Pol. 2007, 24, 349–361.
  • Schweizer, P.E.; Matlack, G.R. Factors driving land use change and forest distribution on the coastal plain of Mississippi, USA. Landsc. Urban Plan. 2014, 121, 55–64.
  • Schulz, J.J.; Cayuela, L.; Echevarria, C.; Salas, J.; Rey Benayas, J.M. Monitoring land cover change of the dryland forest landscape of Central Chile (1975–2008). Appl. Geogr. 2010, 30, 436–447.
  • Serneels, S.; Lambin, E.F. Proximate causes of land-use change in Narok District, Kenya: a spatial statistical model. Agric. Ecosyst. Environ. 2011, 85, 65–81.
  • Serra, P.; Pons, X.; Saurí, D. Land-cover and land-use change in a Mediterranean landscape: A spatial analysis of driving forces integrating biophysical and human factors. Appl. Geogr. 2008, 28, 189–209.
  • Seto, K.C.; Kaufmann, R.K. Modeling the drivers of urban land use change in the Pearl River Delta, China: integrating remote sensing with socioeconomic data. Land Econ. 2003, 79, 106–121.
  • Seto, K.C.; Woodcock, C.E.; Song, C.; Huang, X.; Lu, J.; Kaufmann, R.K. Monitoring land-use change in the Pearl River Delta using Landsat TM. Int. J. Remote Sens. 2002, 23, 1985–2004.
  • Shalaby, A.; Tateishi, R. Remote sensing and GIS for mapping and monitoring land cover and land-use changes in the Northwestern coastal zone of Egypt. Appl. Geogr. 2007, 27, 28–41.
  • Shen, X.; Liu, B.; Lu, X. Effects of land use/land cover on diurnal temperature range in the temperate grassland region of China. Sci. Total Environ. 2017, 575, 1211–1218.
  • Shi, G.; Ye, P.; Ding, L.; Quinones, A.; Li, Y.; Jiang, N. Spatio-temporal patterns of land use and cover change from 1990 to 2010: a case study of Jiangsu province, China. Int. J. Environ. Res. Public Health 2019, 16, 907. https://doi.org/10.3390/ijerph16060907.
  • Shifaw, E.; Sha, J.; Li, X.; Bao, Z.; Zhou, Z. An insight into land-cover changes and their impacts on ecosystem services before and after the implementation of a comprehensive experimental zone plan in Pingtan island, China. Land Use Pol. 2019, 82, 6331–642.
  • Shoyama, K.; Braimoh, A.K. Analyzing about sixty years of land-cover change and associated landscape fragmentation in Shiretoko Peninsula, Northern Japan. Landsc. Urban Plan. 2011, 101, 22–29.
  • Shupe, S. Statistical and spatial analysis of land cover impact on selected Metro Vancouver, British Columbia watersheds. Environ. Manag. 2013, 51, 18–31.
  • Sluiter, R.; de Jong, S.M. Spatial patterns of Mediterranean land abandonment and related land cover transitions. Landsc. Ecol. 2007, 22, 559–576.
  • Smaliychuk, A.; Müller, D.; Prishchepov, A.V.; Levers, C.; Kruhlov, I.; Kuemmerle, T. Recultivation of abandoned agricultural lands in Ukraine: Patterns and drivers. Glob. Environ. Change 2016, 38, 70–81.
  • Smiraglia, D.; Ceccarelli, T.; Bajocco, S.; Salvati, L.; Perini, L. Linking trajectories of land change, land degradation processes and ecosystem services. Environ. Res. 2016, 147, 590–600.
  • Smiraglia, D.; Ceccarelli, T.; Bajocco, S.; Perini, L.; Salvati, L. Unraveling landscape complexity: land use/land cover changes and landscape pattern dynamics (1954–2008) in contrasting peri-urban and agro-forest regions of Northern Italy. Environ. Manag. 2015, 56, 916–932.
  • Song, W.; Deng, X. Land-use/land-cover change and ecosystem service provision in China. Sci. Total Environ. 2017, 576, 705–719.
  • Sonter, L.J.; Moran, C.J.; Barrett, D.J.; Soares-Filho, B.S. Processes of land use change in mining reasons. J. Clean. Prod. 2014, 84, 494–501.
  • Stefanski, J.; Chaskovskyy, O.; Waske, B. Mapping and monitoring of land use changes in post-Soviet western Ukraine using remote sensing data. Appl. Geogr. 2014, 55, 155–164.
  • Szabó, S.; Bertalan, L.; Kerekes, A.; Novák, T.J. Possibilities of land use change analysis in a mountainous rural area: a methodological approach. Int. J. Geogr. Inf. Sci. 2016, 30, 708–726.
  • Taillefumier, F.; Piégay, H. Contemporary land use changes in prealpine Mediterranean mountains: a multivariate GIS-based approach applied to two municipalities in the Southern French Prealps. Catena 2003, 51, 267–296.
  • Tavares, A.O.; Monteiro, M.; Barros, J.L.; Santos, P.P. Long-term land-use changes in small/medium sized cities. Enhancing the general trends and local characteristics. Eur. Plan. Stud. 2019, 27, 1432–1459.
  • Tavares, A.O.; Pato, R.L.; Magalhães, M.C. Spatial and temporal land use change and occupation over the last half century in a peri-urban area. Appl. Geogr. 2012, 34, 432–444.
  • Teixeira, Z.; Teixeira, H.; Marques, J.C. Systematic processes of land use/land cover change to identify relevant driving forces: implications on water quality. Sci. Total Environ. 2014, 470–471, 1320–1335.
  • Temesgen, H.; Nyssen, J.; Zenebe, A.; Haregeweyn, N.; Kindu, M.; Lemenih, M.; Haile, M. Ecological succession and land use changes in a lake retreat area (Main Ethiopia Rift Valley). J. Arid Environ. 2013, 91, 53–60.
  • Tomaz, C.; Alegria, C.; Massano Monteiro, J.; Canavarro Teixeira, M. Land cover change and afforestation of marginal and abandoned agricultural land: a 10 year analysis in a Mediterranean region. For. Ecol. Manag. 2013, 308, 40–49.
  • Tovar, C.; Seijmonsbergen, A.C.; Duivenvoorden, J.F. Monitoring land use and land cover change in mountain regions: an example in the Jalca grasslands of the Peruvian Andes. Landsc. Urban Plan. 2013, 112, 40–49.
  • Trincsi, K.; Pham, T-T-H.; Turner, S. Mapping moutain diversity: Ethnic minorities and land use land cover change in Vietnam’s borderlands. Land Use Pol. 2014, 41, 484–497.
  • Tsegaye, D.; Moe, S.R.; Veded, P.; Aynekulu, E. Land-use/cover dynamics in Northern Afar rangelands, Ethiopia. Agric. Ecosyst. Environ. 2010, 139, 174–180.
  • Twisa, S.; Buchroithner, M.F. Land-use and land-cover (LULC) change detection in Wami river basin, Tanzania. Land 2019, 8, 136. https://doi.org/10.3390/land8090136.
  • Uddin, K.; Chaudhary, S.; Chettri, N.; Kotru, R.; Murthy, M.; Chaudhary, R.P.; Ning, W.; Shrestha, S.M.; Gautam, S.K. The changing land cover and fragmenting forest on the roof of the world: a case study in Nepal’s Kailash sacred landscape. Landsc. Urban Plann. 2015, 141, 1–10.
  • Vasconcelos, M.J.P.; Mussá Biai, J.C.; Araújo, A.; Diniz, M.A. Land cover change in two protected areas of Guinea-Bissau (1956–1998). Appl. Geogr. 2002, 22, 139–156.
  • Verburg, P.H.; Overmars, K.P.; Huigen, M.G.A.; de Groot, W.T.; Veldkamp, A. Analysis of the effects of land use change on protected areas in the Philippines. Appl. Geogr. 2006, 26, 153–173.
  • Verburg, PL.H.; Ritsema van Eck, J.R.; de Nijs, T.C.M.; Dijst, M.J.; Schot, P. Determinants of land-use change patterns in the Netherlands. Environ. Plann. B Plann. Des. 2004, 31, 125–150.
  • Vidal-Macua, J.J.; Ninyerola, M.; Zabala, A.; Domingo-Marimon, C.; Pons, X. Factors affecting forest dynamics in the Iberian Peninsula from 1987 to 2012. The role of topography and drought. For. Ecol. Manag. 2017, 406, 290–306.
  • Wakeel, A.; Rao, K.S.; Maikhuri, R.K.; Saxena, K.G. Forest management and land use/cover changes in a typical micro watershed in the mid elevation zone of Central Himalaya, India. For. Ecol. Manag. 2005, 213, 229–242.
  • Wang, X.; Zheng, D.; Shen, Y. Land use change and its driving forces on the Tibetan Plateau during 1990–2000. Catena 2008, 72, 56–66.
  • Wang, Y.C.; Hu, B.K.H.; Myint, S.W.; Feng, C.C.; Chow, W.T.L.; Passy, P.F. Patterns of land change and their potential impacts on land surface temperature change in Yangon, Myanmar. Sci. Total Environ. 2018, 643, 738–750.
  • Were, K.O.; Dick, Ø.B.; Singh, B.R. Remotely sensing the spatial and temporal land cover changes in Eastern Mau forest reserve and Lake Nakuru drainage basin, Kenya. Appl. Geogr. 2013, 41, 75–86.
  • Wilson, C.O. Land use/land cover water quality nexus: quantifying anthropogenic influences on surface water quality. Environ. Monit. Assess. 2015, 187, 424.
  • Wolde Yohannes, A.; Cotter, M.; Kelboro, G.; Dessalegn, W. Land use and land cover changes and their effects on the landscape of Abaya-Chamo basin, Southern Ethiopia. Land 2018, 7, 2. https://doi.org/10.3390/land7010002.
  • Wood, E.C.; Tappan, G.G.; Hadj, A. Understanding the drivers of agricultural land use change in south-central Senegal. J. Arid Environ. 2004, 59, 565–582.
  • Wu, Y.; Li, S.; Yu, S. Monitoring urban expansion and its effects on land use and land cover changes in Guangzhou city, China. Environ. Monit. Assess. 2016, 188, article 54.
  • Wyman, M.S.; Stein, T.V. Modeling social and land-use/land-cover change data to assess drivers of smallholder deforestation in Belize. Appl. Geogr. 2010, 30, 329–342.
  • Yi, H.; Güneralp, B.; Filippi, A.M.; Kreuter, U.P.; Güneralp, I. Impacts of land change and ecosystem services in the San Antonio river basin, Texas, from 1984 to 2010. Ecol. Econ. 2017, 135, 125–135.
  • Yirsaw, E.; Wu, W.; Temesgen, H.; Bekele, B. Socioeconomic drivers of spatio-temporal land use/land cover changes in a rapidly urbanizing area of China, the Su-Xi-Chang region. Appl. Ecol. Environ. Res. 2017, 15, 809–827.
  • Yu, W.; Zang, S.; Wu, C.; Liu, W.; Na, X. Analyzing and modeling land use land cover change (LUCC) in the Daqing City, China. Appl. Geogr. 2011, 31, 600–608.
  • Yuan, F.; Sawaya, K.E.; Loeffelholz, B.C.; Bauer, M.E. Land cover classification and change analysis of the Twin Cities (Minnesota) Metropolitan Area by multitemporal Landsat remote sensing. Remote Sens. Environ. 2005, 98, 317–328.
  • Xiao, J.; Shen, Y.; Ge, J.; Tateishi, R.; Tang, C.; Liang, Y.; Huang, Z. Evaluating urban expansion and land use change in Shijiazhuang, China, by using GIS and remote sensing. Landsc. Urban Plann. 2006, 75, 69–80.
  • Xu, Z.; Fan, W.; Wei, H.; Zhang, P.; Ren, J.; Gao, Z.; Ulgiati, S.; Kong, W.; Dong, X. Evaluation and simulation of the impact of land use change on ecosystem services based on a carbon flow model: a case study of the Manas River Basin of Xinjiang, China. Sci. Total Environ. 2019, 652, 117–133.
  • Xystrakis, F.; Psarras, T.; Koutsias, N. A process-based land use/land cover change assessment on a mountainous area of Greece during 1945–2009: Signs of socio-economic drivers. Sci. Total Environ. 2017, 587–588, 360–370.
  • Zanganeh, S.; Saurí, D.; Serra, P.; Modugno, S.; Seifolddini, F.; Pourahma, A. Urban sprawl pattern and land-use change detection in Yazd, Iran. Habitat Int. 2011, 35, 521–528.
  • Zhang, F.; Yushanjiang, A.; Jing, Y. Assessing and predicting changes of the ecosystem service values based on land use/cover change in Ebinur Lake Wetland National Nature Reserve, Xinjiang, China. Sci. Total Environ. 2019, 656, 1133–1144.
  • Zheng D.; Heath, L.S.; Ducey, M.J. Satellite detection of land-use change and effects on regional forest aboveground biomass estimates. Environ. Monit. Assess. 2008, 144, 67–79.
  • Zomeni, M.; Tzanopoulos, J.; Pantis, J.D. Historical analysis of landscape change using remote sensing techniques: an explanatory tool for agricultural transformation in Greek rural areas. Landsc. Urban Plan. 2008, 86, 38–46.

References

  1. Turner, B.L., II; Lambin, E.F.; Reenberg, A. The emergence of land change science for global environmental change and sustainability. Proc. Natl. Acad. Sci. USA 2007, 104, 20666–20671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Brissoulis, H. Analysis of Land Use Change: Theoretical and Modeling Approaches, 2nd ed.; Loveridge, S., Jackson, R., Eds.; Regional Research Institute, West Virginia University: Morgantown, WV, USA, 2020; Available online: https://researchrepository.wvu.edu/rri-web-book/3 (accessed on 8 May 2020).
  3. Turner, B.L., II; Skole, D.; Sanderson, S.; Fischer, G.; Fresco, L.; Leemans, R. Land-Use and Land-Cover Change: Science/Research Plan; IGBP Report 35; IGBP/HDP; Royal Swedish Academy of Sciences: Stockholm, Sweden; Geneva, Switzerland, 1995. [Google Scholar]
  4. Seitzinger, S.P.; Gaffney, O.; Brasseur, G.; Broadgate, W.; Ciais, P.; Claussen, M.; Erisman, J.W.; Kiefer, T.; Lancelot, C.; Monks, P.S.; et al. International Geosphere–Biosphere Programme and Earth system science: Three decades of co-evolution. Anthropocene 2015, 12, 3–16. [Google Scholar] [CrossRef]
  5. Arino, O.; Bicheron, P.; Achard, F.; Latham, J.; Witt, R.; Weber, J.L. Globcover: The Most Detailed Portrait of Earth; European Space Agency Bulletin: Frascati, Italy, 2008; Volume 136. [Google Scholar]
  6. Justice, C.; Gutman, G.; Vadrevu, K.P. NASA Land Cover and Land Use Change (LCLUC): An interdisciplinary research program. J. Environ. Manag. 2015, 148, 4–9. [Google Scholar] [CrossRef] [PubMed]
  7. Verburg, P.H.; Crossman, N.; Ellis, E.C.; Heinimann, A.; Hostert, P.; Mertz, O.; Nagendra, H.; Sikor, T.; Erb, K.-H.; Golubiewski, N.; et al. Land system science and sustainable development of the earth system: A global land project perspective. Anthropocene 2015, 12, 29–41. [Google Scholar] [CrossRef] [Green Version]
  8. Rounsevell, M.D.A.; Pedroli, B.; Erb, K.-H.; Gramberger, M.; Busck, A.G.; Haberl, H.; Kristensen, S.; Kuemmerle, T.; Lavorel, S.; Lindner, M.; et al. Challenges for land system science. Land Use Policy 2012, 29, 899–910. [Google Scholar] [CrossRef]
  9. Gutman, G.; Janetos, A.C.; Justice, C.O.; Moran, E.F.; Mustard, J.F.; Rindfuss, R.R.; Skole, D.; Turner, B.L., II; Cochrane, M.A. Land Change Science: Observing, Monitoring and Understanding Trajectories of Change on the Earth’s Surface; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2004. [Google Scholar]
  10. Rindfuss, R.R.; Walsh, S.J.; Turner, B.L., II; Fox, J.; Mishra, V. Developing a science of land change: Challenges and methodological issues. Proc. Natl. Acad. Sci. USA 2004, 101, 13976–13981. [Google Scholar] [CrossRef] [Green Version]
  11. Bajocco, S.; Dragoz, E.; Gitas, I.; Smiraglia, D.; Salvati, L.; Ricotta, C. Mapping forest fuels through vegetation phenology: The role of coarse-resolution satellite time-series. PLoS ONE 2015, 10, e0119811. [Google Scholar] [CrossRef] [Green Version]
  12. Rindfuss, R.R.; Entwisle, B.; Walsh, S.J.; An, L.; Badenoch, N.; Brown, D.G.; Deadman, P.; Evans, T.P.; Fox, J.; Geoghegan, J.; et al. Land use change: Complexity and comparisons. J. Land Use Sci. 2008, 3, 1–10. [Google Scholar] [CrossRef] [Green Version]
  13. Meyfroidt, P.; Chowdhury, R.R.; de Bremond, A.; Ellis, E.C.; Erb, K.-H.; Filatova, T.; Garrett, R.D.; Grove, J.M.; Heinimann, A.; Kuemmerle, T.; et al. Middle-range theories of land system change. Glob. Environ. Change 2018, 53, 52–67. [Google Scholar] [CrossRef]
  14. Baynard, C.W. Remote sensing applications: Beyond land-use and land-cover change. Adv. Remote Sens. 2013, 2, 228–241. [Google Scholar] [CrossRef] [Green Version]
  15. Nagabhatla, N.; Padmanabhan, M.; Kühle, P.; Vishnudas, S.; Betz, L.; Niemeyer, B. LCLUC as an entry point for transdisciplinary research–Reflections from an agriculture land use change study in South Asia. J. Environ. Manag. 2015, 148, 42–52. [Google Scholar] [CrossRef] [PubMed]
  16. Long, H.; Tang, G.; Li, X.; Heilig, G.K. Socio-economic driving forces of land-use change in Kunshan, the Yangtze River Delta economic area of China. J. Environ. Manag. 2007, 83, 351–364. [Google Scholar] [CrossRef] [PubMed]
  17. Uddin, K.; Chaudhary, S.; Chettri, N.; Kotru, R.; Murthy, M.; Chaudhary, R.P.; Ning, W.; Shrestha, S.M.; Gautam, S.K. The changing land cover and fragmenting forest on the Roof of the World: A case study in Nepal’s Kailash Sacred Landscape. Landsc. Urban Plan. 2015, 141, 1–10. [Google Scholar] [CrossRef] [Green Version]
  18. Gómez, C.; White, J.C.; Wulder, M.A. Optical remotely sensed time series data for land cover classification: A review. ISPRS J. Photogramm. Remote Sens. 2016, 116, 55–72. [Google Scholar] [CrossRef] [Green Version]
  19. Serra, P.; Pons, X. Monitoring farmers’ decisions on Mediterranean irrigated crops using satellite image time series. Int. J. Remote Sens. 2008, 29, 2293–2316. [Google Scholar] [CrossRef]
  20. Wulder, M.A.; Masek, J.G.; Cohen, W.B.; Loveland, T.R.; Woodcock, C.E. Opening the archive: How free data has enabled the science and monitoring promise of Landsat. Remote Sens. Environ. 2012, 122, 2–10. [Google Scholar] [CrossRef]
  21. Kuemmerle, T.; Erb, K.; Meyfroidt, P.; Muller, D.; Verburg, P.H.; Estel, S.; Haberl, H.; Hostert, P.; Jepsen, M.R.; Kastner, T.; et al. Challenges and opportunities in mapping land use intensity globally. Curr. Opin. Environ. Sustain. 2013, 5, 484–493. [Google Scholar] [CrossRef]
  22. Sishodia, R.; Ray, R.L.; Singh, S.K. Applications of Remote Sensing in Precision Agriculture: A Review. Remote Sens. 2020, 12, 3136. [Google Scholar] [CrossRef]
  23. Lu, D.; Weng, Q. A survey of image classification methods and techniques for improving classification performance. Int. J. Remote Sens. 2007, 28, 823–870. [Google Scholar] [CrossRef]
  24. Gilani, H.; Shrestha, H.L.; Murthy, M.S.R.; Phuntso, P.; Pradhan, S.; Bajracharya, B.; Shrestha, B. Decadal land cover change dynamics in Bhutan. J. Environ. Manag. 2015, 148, 91–100. [Google Scholar] [CrossRef]
  25. Phiri, D.; Morgenroth, J. Developments in Landsat land cover classification methods: A review. Remote Sens. 2017, 9, 967. [Google Scholar] [CrossRef] [Green Version]
  26. Alvarez-Martinez, J.M.; Suárez-Seoanea, S.; De Luis Calabuig, E. Modelling the risk of land cover change from environmental and socio-economic drivers in heterogeneous and changing landscapes: The role of uncertainty. Landsc. Urban Plan. 2011, 101, 108–119. [Google Scholar] [CrossRef]
  27. Serra, P.; Pons, X. Two Mediterranean irrigation communities in front of water scarcity: A comparison using satellite image time series. J. Arid Environ. 2013, 98, 41–51. [Google Scholar] [CrossRef]
  28. Kibret, K.S.; Marohn, C.; Cadisch, G. Assessment of land use and land cover change in South Central Ethiopia during four decades based on integrated analysis of multi-temporal images and geospatial vector data. Remote Sens. Appl. Soc. Environ. 2016, 3, 1–19. [Google Scholar] [CrossRef]
  29. Clement, F.; Orange, D.; Williams, M.; Mulley, C.; Epprecht, M. Drivers of afforestation in Northern Vietnam: Assessing local variations using geographically weighted regression. Appl. Geogr. 2009, 29, 561–576. [Google Scholar] [CrossRef]
  30. Ariti, A.T.; van Vliet, J.; Verburg, P.H. Land-use and land-cover changes in the Central Rift Valley of Ethiopia: Assessment of perception and adaptation of stakeholders. Appl. Geogr. 2015, 65, 28–37. [Google Scholar] [CrossRef]
  31. Foody, G.M. Status of land cover classification accuracy assessment. Remote Sens. Environ. 2002, 80, 185–201. [Google Scholar] [CrossRef]
  32. Pontius, R.G.; Shusas, E.; McEachern, M. Detecting important categorical land changes while accounting for persistence. Agric. Ecosyst. Environ. 2004, 101, 251–268. [Google Scholar] [CrossRef]
  33. Cuba, N. Research note: Sankey diagrams for visualizing land cover dynamics. Landsc. Urban Plan. 2015, 139, 163–167. [Google Scholar] [CrossRef]
  34. Mas, J.F. Monitoring land-cover changes: A comparison of change detection techniques. Int. J. Remote Sens. 1999, 20, 139–152. [Google Scholar] [CrossRef]
  35. Dewan, A.M.; Yamaguchi, Y.; Rahman, M.Z. Dynamics of land use/cover changes and the analysis of landscape fragmentation in Dhaka Metropolitan, Bangladesh. GeoJournal 2012, 77, 315–330. [Google Scholar] [CrossRef]
  36. Aspinall, R.J.; Hill, M.J. Land cover change: A method for assessing the reliability of land cover changes measured from remotely-sensed data. In Proceedings of the International Geoscience and Remote Sensing Symposium, IGARSS ’97, Singapore, 4–8 August 1997; pp. 269–271. [Google Scholar]
  37. Serra, P.; Pons, X.; Saurí, D. Post-classification change detection with data from different sensors. Some accuracy considerations. Int. J. Remote Sens. 2003, 24, 3311–3340. [Google Scholar] [CrossRef]
  38. Szabó, S.; Bertalan, L.; Kerekes, A.; Novák, T.J. Possibilities of land use change analysis in a mountainous rural area: A methodological approach. Int. J. Geogr. Inform. Sci. 2016, 30, 708–726. [Google Scholar] [CrossRef] [Green Version]
  39. Chrisman, N.R. The accuracy of map overlays: A reassessment. Landsc. Urban Plan. 1987, 14, 427–439. [Google Scholar] [CrossRef]
  40. Goodchild, M.F.; Guoqing, S.; Shiren, Y. Development and test of an error model for categorical data. Int. J. Geogr. Inf. Systems 1992, 6, 87–104. [Google Scholar] [CrossRef]
  41. Frondoni, R.; Mollo, B.; Capotorti, G. A landscape analysis of land cover change in the Municipality of Rome (Italy): Spatio-temporal characteristics and ecological implications of land cover transitions from 1954 to 2001. Landsc. Urban Plan. 2011, 100, 117–128. [Google Scholar] [CrossRef]
  42. Singh, A. Digital change detection techniques using remotely-sensed data. Int. J. Remote Sens. 1989, 10, 989–1003. [Google Scholar] [CrossRef] [Green Version]
  43. Lambin, E.F.; Strahler, A.H. Change vector analysis in multitemporal space: A tool to detect and categorize land-cover change processes using high temporal resolution satellite data. Remote Sens. Environ. 1994, 48, 231–244. [Google Scholar] [CrossRef]
  44. Were, K.O.; Dick, Ø.B.; Singh, B.R. Remotely sensing the spatial and temporal land cover changes in Eastern Mau forest reserve and Lake Nakuru drainage basin, Kenya. Appl. Geogr. 2013, 41, 75–86. [Google Scholar] [CrossRef]
  45. Brown, D.G.; Verburg, P.H.; Pontius, R.G., Jr.; Lange, M. Opportunities to improve impact, integration, and evaluation of land change models. Curr. Opin. Environ. Sustain. 2013, 5, 452–457. [Google Scholar] [CrossRef]
  46. Verburg, P.; Ritsema van Eck, J.; de Nijs, T.; Dijst, M.; Schot, P. Determinants of land use change patterns in the Netherlands. Environ. Plan. B Urban Anal. City Sci. 2004, 31, 125–150. [Google Scholar] [CrossRef] [Green Version]
  47. Bürgi, M.; Hersperger, A.M.; Schneeberger, N. Driving forces of landscape change–Current and new directions. Landsc. Ecol. 2004, 19, 857–868. [Google Scholar] [CrossRef]
  48. Hersperger, A.M.; Gennaio, M.P.; Verburg, P.H.; Bürgi, M. Linking land change with driving forces and actors: Four conceptual models. Ecol. Soc. 2010, 15, 1. Available online: http://www.ecologyandsociety.org/vol15/iss4/art1/ (accessed on 25 February 2020). [CrossRef] [Green Version]
  49. Dang, A.N.; Kawasaki, A. A review of methodological integration in land-use change models. Int. J. Agric. Environ. Inf. Syst. 2016, 7, 1–25. [Google Scholar] [CrossRef]
  50. Geist, H.J.; Lambin, E.F. What Drives Tropical Deforestation? A Meta-Analysis of Proximate and Underlying Causes of Deforestation Based on Subnational Case Study Evidence; Land-Use and Land-Cover Change (LUCC). Project IV. International Human Dimensions Programme on Global Environmental Change (IHDP); International Geosphere-Biosphere Programme (IGBP); LUCC International Project Office: Louvain-la-Neuve, Belgium, 2001. [Google Scholar]
  51. Serneels, S.; Lambin, E.F. Proximate causes of land-use change in Narok District, Kenya: A spatial statistical model. Agric. Ecosyst. Environ. 2001, 85, 65–81. [Google Scholar] [CrossRef] [Green Version]
  52. Verburg, P.H.; Schot, P.P.; Dijst, M.J.; Veldkamp, A. Land use change modelling: Current practice and research priorities. GeoJournal 2004, 61, 309–324. [Google Scholar] [CrossRef]
  53. Agarwal, C.; Green, G.L.; Grove, M.; Evans, T.; Schweik, C. A Review and Assessment of Land-Use Change Models: Dynamics of Space, Time and Human Choice; General Technical Report NE-297; United States Department of Agriculture, Indiana University: Bloomington, IN, USA, 2000. [Google Scholar]
  54. Schneeberger, N.; Bürgi, M.; Hersperger, A.M.; Ewald, K.C. Driving forces and rates of landscape change as a promising combination for landscape change research—An application on the northern fringe of the Swiss Alps. Land Use Policy 2007, 24, 349–361. [Google Scholar] [CrossRef]
  55. Valbuena, D.; Verburg, P.H.; Bregt, A.K.; Ligtenberg, A. An agent-based approach to model land-use change at a regional scale. Landsc. Ecol. 2010, 25, 185–199. [Google Scholar] [CrossRef] [Green Version]
  56. Mas, J.F.; Paegelow, M.; Camacho Olmedo, M.T. LUCC modeling approaches to calibration. In Geomatic Approaches for Modeling Land Change Scenarios; Camacho Olmedo, M.T., Paegelow, M., Mas, J.-F., Escobar, F., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 11–25. [Google Scholar]
  57. Parcerisas, L.L.; Marull, J.; Pino, J.; Tello, E.; Coll, F.; Basnou, C. Land use changes, landscape ecology and their socioeconomic driving forces in the Spanish Mediterranean coast (El Maresme County, 1850–2005). Environ. Sci. Policy 2012, 23, 120–132. [Google Scholar] [CrossRef]
  58. Mas, J.F.; Kolb, M.; Paegelow, M.; Camacho Olmedo, M.T.; Houet, T. Inductive pattern-based land use/cover change models: A comparison of four software packages. Environ. Model. Softw. 2014, 51, 94–111. [Google Scholar] [CrossRef] [Green Version]
  59. Pérez-Vega, A.; Mas, J.F.; Ligmann-Zielinska, A. Comparing two approaches to land use/cover change modeling and their implications for the assessment of biodiversity loss in a deciduous tropical forest. Environ. Model. Softw. 2012, 29, 11–23. [Google Scholar] [CrossRef]
  60. Camacho Olmedo, M.T.; Paegelow, M.; Mas, J.-F.; Escobar, F. Geomatic Approaches for Modeling Land Change Scenarios; Springer International Publishing: Cham, Switzerland, 2018. [Google Scholar]
  61. Kolb, M.; Mas, J.F.; Galicia, L. Evaluating drivers of land-use change and transition potential models in a complex landscape in Southern Mexico. Int. J. Geogr. Inform. Sci. 2013, 27, 1804–1827. [Google Scholar] [CrossRef]
  62. Entwisle, B.; Walsh, S.J.; Rindfuss, R.R.; Chamratrithirong, A. Land-use/land-cover and population dynamics, Nang Rong, Thailand. In People and Pixels. Linking Remote Sensing and Social Science; Liverman, D., Moran, E.F., Rindfuss, R.R., Stern, P.C., Eds.; National Academy Press: Washington, DC, USA, 1998; pp. 121–144. [Google Scholar]
  63. Walsh, S.J.; Crews-Meyer, K.A. Linking People, Place and Policy: A GIScience Approach; Kluwer Academic Publishers: Norwell, MA, USA, 2002. [Google Scholar]
  64. Fox, J.; Rindfuss, R.R.; Walsh, S.J.; Mishra, V. People and the Environment; Kluwer Academic Publishers: Boston, MA, USA, 2004. [Google Scholar]
  65. Zambon, I.; Colantoni, A.; Carlucci, M.; Morrow, N.; Sateriano, A.; Salvati, L. Land quality, sustainable development and environmental degradation in agricultural districts: A computational approach based on entropy indexes. Environ. Impact Assess. Rev. 2017, 64, 37–46. [Google Scholar] [CrossRef]
  66. Salvati, L.; Ciommi, M.T.; Serra, P.; Chelli, F.M. Exploring the spatial structure of housing prices under economic expansion and stagnation: The role of socio-demographic factors in metropolitan Rome, Italy. Land Use Policy 2019, 81, 143–152. [Google Scholar] [CrossRef]
  67. Bajocco, S.; Ceccarelli, T.; Smiraglia, D.; Salvati, L.; Ricotta, C. Modeling the ecological niche of long-term land use changes: The role of biophysical factors. Ecol. Indic. 2016, 60, 231–236. [Google Scholar] [CrossRef]
  68. Wood, C.H.; Skole, D. Linking satellite, census, and survey data to study deforestation in the Brazilian Amazon. In People and Pixels. Linking Remote Sensing and Social Science; Liverman, D., Moran, E.F., Rindfuss, R.R., Stern, P.C., Eds.; National Academy Press: Washington, DC, USA, 1998; pp. 70–93. [Google Scholar]
  69. Perz, S.G.; Skole, D.L. Social determinants of secondary forests in the Brazilian Amazon. Soc. Sci. Res. 2003, 32, 25–60. [Google Scholar] [CrossRef]
  70. Reger, B.; Otte, A.; Waldhardt, R. Identifying patterns of land-cover change and their physical attributes in a marginal European landscape. Landsc. Urban Plan. 2007, 81, 104–113. [Google Scholar] [CrossRef]
  71. Shupe, S. Statistical and Spatial Analysis of Land Cover Impact on Selected Metro Vancouver, British Columbia Watersheds. Environ. Manag. 2013, 51, 18–31. [Google Scholar] [CrossRef]
  72. Monteiro, A.T.; Fava, F.; Hiltbrunner, E.; Marianna, G.D.; Bocchi, S. Assessment of land cover changes and spatial drivers behind loss of permanent meadows in the lowlands of Italian Alps. Landsc. Urban Plan. 2011, 100, 287–294. [Google Scholar] [CrossRef]
  73. Espindola, G.M.; de Aguiar, A.P.D.; Pebesma, E.; Câmara, G.; Fonseca, L. Agricultural land use dynamics in the Brazilian Amazon based on remote sensing and census data. Appl. Geogr. 2012, 32, 240–252. [Google Scholar] [CrossRef]
  74. Chhabra, A.; Geist, H.; Houghton, R.A.; Haberl, H.; Braimoh, A.K.; Vlek, P.L.G.; Patz, J.; Xu, J.; Ramankutty, N.; Coomes, O.; et al. Multiple impacts pof land-use/cover change. In Land-Use and Land-Cover Change. Local Processes and Global Impacts; Lambin, E.F., Geist, H., Eds.; Springer: Heidelberg, Germany, 2006; pp. 71–116. [Google Scholar]
  75. Salvati, L.; Gemmiti, R.; Perini, L. Land degradation in Mediterranean urban areas: An unexplored link with planning? Area 2012, 44, 317–325. [Google Scholar] [CrossRef]
  76. Dadashpoor, H.; Azizi, P.; Moghadas, M. Land use change, urbanization, and change in landscape pattern in a metropolitan area. Sci. Total Environ. 2019, 655, 707–719. [Google Scholar] [CrossRef] [PubMed]
  77. Wu, Y.; Li, S.; Yu, S. Monitoring urban expansion and its effects on land use and land cover changes in Guangzhou city, China. Environ. Monit. Assess. 2016, 188, 54. [Google Scholar] [CrossRef] [PubMed]
  78. McGarigal, K.; Cushman, S.; Ene, E. FRAGSTATS v4: Spatial Pattern Analysis Program for Categorical and Continuous Maps. 2012. Available online: https://www.umass.edu/landeco/research/fragstats/fragstats.html (accessed on 25 May 2021).
  79. Arowolo, A.O.; Deng, X.; Olatunji, O.A.; Obayelu, A.E. Assessing changes in the value of ecosystem services in response to land-use/land-cover dynamics in Nigeria. Sci. Total Environ. 2018, 636, 597–609. [Google Scholar] [CrossRef] [PubMed]
  80. Burkhard, B.; Kroll, F.; Nedkov, S.; Müller, F. Mapping ecosystem service supply, demand and budgets. Ecol. Indic. 2012, 21, 17–29. [Google Scholar] [CrossRef]
  81. Kertész, A.; Nagy, L.A.; Balázs, B. Effect of land use change on ecosystem services in Lake Balaton Catchment. Land Use Policy 2019, 80, 430–438. [Google Scholar] [CrossRef]
  82. Basse, R.M.; Omrani, H.; Charif, O.; Gerber, P.; Bódis, K. Land use changes modelling using advanced methods: Cellular automata and artificial neural networks. The spatial and explicit representation of land cover dynamics at the cross-border region scale. Appl. Geogr. 2014, 53, 160–171. [Google Scholar] [CrossRef]
  83. Aburas, M.M.; Ahamad, M.S.S.; Omar, N.Q. Spatio-temporal simulation and prediction of land-use change using conventional and machine learning models: A review. Environ. Monit. Assess. 2019, 205, 191–205. [Google Scholar] [CrossRef]
  84. Hu, Y.; Batunacun; Zhen, L.; Zhuang, D. Assessment of land-use and land-cover change in Guangxi, China. Sci. Rep. 2019, 9, 2189. [Google Scholar] [CrossRef] [Green Version]
  85. Garcia-Frapolli, E.; Ayala-Orozco, B.; Bonilla-Moheno, M.; Espadas-Manrique, C.; Ramos-Fernández, G. Biodiversity conservation, traditional agriculture and ecotourism: Land cover/land use change projections for a natural protected area in the northeastern Yucatan Peninsula, Mexico. Landsc. Urban Plan. 2007, 83, 137–153. [Google Scholar] [CrossRef]
  86. Marshall, E.; Randhir, T.O. Spatial modeling of land cover change and watershed response using Markovian cellular automata and simulation. Water Resour. Res. 2008, 44, 1–10. [Google Scholar] [CrossRef] [Green Version]
  87. Omrani, H.; Charif, O.; Gerber, P.; Bódis, K.; Basse, R.M. Simulation of Land Use Changes Using Cellular Automata and Artificial Neural Network; Working Paper No. 2012-01; CEPS/INSTEAD: Luxembourg, 2012. [Google Scholar]
  88. Lesschen, J.P.; Verburg, P.H.; Staal, S.J. Statistical Methods for Analysing the Spatial Dimension of Changes in Land Use and Farming Systems; LUCC Report Series No. 7; The International Livestock Research Institute: Nairobi, Kenya; LUCC Focus 3 Office, Wageningen University: Wageningen, The Netherlands, 2005. [Google Scholar]
  89. Costanza, R.; d’Arge, R.; de Groot, R.; Farber, S.; Grasso, M.; Hannon, B.; Limburg, K.; Naeem, S.; O’Neill, R.V.; Paruelo, J.; et al. The value of the world’s ecosystem services and natural capital. Nature 1997, 387, 253–260. [Google Scholar] [CrossRef]
  90. Costanza, R.; de Groot, R.; Sutton, P.; van der Ploeg, S.; Anderson, S.J.; Kubiszewski, I.; Farber, S.; Turner, R.K. Changes in the global value of ecosystem services. Glob. Environ. Change 2014, 26, 152–158. [Google Scholar] [CrossRef]
  91. Song, W.; Deng, X. Land-use/land-cover change and ecosystem service provision in China. Sci. Total Environ. 2017, 576, 705–719. [Google Scholar] [CrossRef] [PubMed]
  92. Verburg, P.H.; de Koning, G.H.J.; Kok, K.; Veldkamp, A.; Bouma, J.A. spatial explicit allocation procedure for modelling the pattern of land use change based upon actual land use. Ecol. Model. 1999, 116, 45–61. [Google Scholar] [CrossRef]
  93. Institute for Environmental Studies CLUE Model. 2021. Available online: http://www.ivm.vu.nl/en/Organisation/departments/spatial-analysis-decision-support/Clue/ (accessed on 15 July 2021).
  94. Tallis, H.; Ricketts, T.; Guerry, A.; Wood, S.; Sharp, R.; Nelson, E.; Ennaanay, D.; Wolny, S.; Olwero, N.; Vigerstol, K.; et al. InVEST 2.5.6 User’s Guide; The Natural Capital Project: Stanford, CA, USA, 2013. [Google Scholar]
  95. Gao, J.; Li, F.; Gao, H.; Zhou, C.; Zhang, X. The impact of land-use change on water-related ecosystem services: A study of the Guishui River basin, Beijing, China. J. Clean. Prod. 2017, 163, S148–S155. [Google Scholar] [CrossRef]
  96. Jiang, W.; Deng, Y.; Tang, Z.; Lei, X.; Chen, Z. Modelling the potential impacts of urban ecosystem changes oncarbon storage under different scenarios by linking the CLUE-S andthe InVEST models. Ecol. Model. 2017, 345, 30–40. [Google Scholar] [CrossRef]
  97. White, R.; Engelen, G.; Uljee, I. The use of constrained cellular automata for high-resolution modelling of urban land use dynamics. Environ. Plan. B Urban Anal. City Sci. 1997, 24, 323–343. [Google Scholar] [CrossRef]
  98. Saputra, M.H.; Lee, H.S. Prediction of land use and land cover changes for North Sumatra, Indonesia, using an artificial neural-network-based cellular automaton. Sustainability 2019, 11, 3024. [Google Scholar] [CrossRef] [Green Version]
  99. Eastman, J.R.; Toledano, J. A short presentation of the land change modeler (LCM). In Geomatic Approaches for Modeling Land Change Scenarios; Camacho Olmedo, M.T., Paegelow, M., Mas, J.-F., Escobar, F., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 499–505. [Google Scholar]
  100. Verburg, P.H.; Soepboer, W.; Limpiada, R.; Espaldon, M.V.O.; Sharifa, M.A.; Veldkamp, A. Modelling the spatial dynamics of regional land use: The CLUE-S model. Environ. Manag. 2002, 30, 391–405. [Google Scholar] [CrossRef]
  101. Schaldach, R.; Priess, J.A. Integrated models of the Land System: A review of modelling approaches on the regional to global scale. Living Rev. Landsc. Res. 2008, 2, 1. Available online: http://www.livingreviews.org/lrlr-2008-1 (accessed on 28 February 2020). [CrossRef] [Green Version]
  102. Van Asselen, S.; Verburg, P.H. Land cover change or land-use intensification: Simulating land system change with a global-scale land change model. Glob. Change Biol. 2013, 19, 3648–3667. [Google Scholar] [CrossRef] [PubMed]
  103. Chakraborty, A.; Sachdeva, K.; Joshi, P.K. Mapping long-term land use and land cover change in the central Himalayan region using a tree-based ensemble classification approach. Appl. Geogr. 2016, 74, 136–150. [Google Scholar] [CrossRef]
  104. Yu, L.; Liang, L.; Wang, J.; Zhao, Y.; Cheng, Q.; Hu, L.; Liu, S.; Yu, L.; Wang, X.; Zhu, P.; et al. Meta-discoveries from a synthesis of satellite-based land-cover mapping research. Int. J. Remote Sens. 2014, 35, 4573–4588. [Google Scholar] [CrossRef]
  105. Maxwell, A.E.; Warner, T.A.; Fang, F. Implementation of machine-learning classification in remote sensing: An applied review. Int. J. Remote Sens. 2018, 39, 2784–2817. [Google Scholar] [CrossRef] [Green Version]
  106. Carmel, Y.; Kadmon, R.; Nirel, R. Spatiotemporal predictive models of Mediterranean vegetation dynamics. Ecol. Appl. 2001, 11, 268–280. [Google Scholar] [CrossRef]
  107. Minetos, D.; Polyzos, S. Multivariate statistical methodologies for testing hypothesis of land-use change at the regional level. A review and evaluation. J. Environ. Prot. Ecol. 2009, 10, 834–866. [Google Scholar]
  108. Noszczyk, T. A review of approaches to land use changes modeling. Hum. Ecol. Risk Assess. 2019, 25, 1377–1405. [Google Scholar] [CrossRef]
  109. MacNally, R. Multiple regression and inference in ecology and conservation biology: Further comments on identifying important predictor variables. Biodivers. Conserv. 2002, 11, 1397–1401. [Google Scholar] [CrossRef]
  110. Lakes, T.; Müller, D.; Krüger, C. Cropland change in southern Romania: A comparison of logistic regressions and artificial neural networks. Landsc. Ecol. 2009, 24, 1195–1206. [Google Scholar] [CrossRef]
  111. Lorena, R.B.; Lambin, E.F. The spatial dynamics of deforestation and agent use in the Amazon. Appl. Geogr. 2009, 29, 171–181. [Google Scholar] [CrossRef]
  112. Xu, Z.; Fan, W.; Wei, H.; Zhang, P.; Ren, J.; Gao, Z.; Ulgiati, S.; Kong, W.; Dong, X. Evaluation and simulation of the impact of land use change on ecosystem services based on a carbon flow model: A case study of the Manas River Basin of Xinjiang, China. Sci. Total Environ. 2019, 652, 117–133. [Google Scholar] [CrossRef]
  113. Pontius, G.R.; Malanson, J. Comparison of the structure and accuracy of two land change models. Int. J. Geogr. Inform. Sci. 2005, 19, 243–265. [Google Scholar] [CrossRef]
  114. Kamusoko, C.; Aniya, M.; Adi, B.; Manjoro, M. Rural sustainability under threat in Zimbabwe-Simulation of future land use/cover changes in the Bindura district based on the Markov-cellular automata model. Appl. Geogr. 2009, 29, 435–447. [Google Scholar] [CrossRef]
Figure 1. A workflow from broader to narrower knowledge issues illustrating the STEPLand perspective. Source: own elaboration.
Figure 1. A workflow from broader to narrower knowledge issues illustrating the STEPLand perspective. Source: own elaboration.
Land 11 01065 g001
Figure 2. Total number of articles identified by year from Web of Science as potential examples of the STEPLand framework and the 250 manuscripts analyzed.
Figure 2. Total number of articles identified by year from Web of Science as potential examples of the STEPLand framework and the 250 manuscripts analyzed.
Land 11 01065 g002
Figure 3. Total number of citations per year.
Figure 3. Total number of citations per year.
Land 11 01065 g003
Figure 4. Percentage of articles categorized according to STEPLand steps. Source: own elaboration.
Figure 4. Percentage of articles categorized according to STEPLand steps. Source: own elaboration.
Land 11 01065 g004
Figure 5. Affiliation discipline of the first author from all papers. Source: own elaboration.
Figure 5. Affiliation discipline of the first author from all papers. Source: own elaboration.
Land 11 01065 g005
Figure 6. The top ten journals, totaling 164 papers, with most publications analyzed in this review. Source: own elaboration.
Figure 6. The top ten journals, totaling 164 papers, with most publications analyzed in this review. Source: own elaboration.
Land 11 01065 g006
Figure 7. Study areas per country from the 250 papers analyzed. Source: own elaboration.
Figure 7. Study areas per country from the 250 papers analyzed. Source: own elaboration.
Land 11 01065 g007
Figure 8. EO data sources from all papers. Source: own elaboration.
Figure 8. EO data sources from all papers. Source: own elaboration.
Land 11 01065 g008
Figure 9. Methods used for obtaining LUC maps only including those developed with RS (128 papers). NE: not explained; Photo: photointerpretation of satellite images; OBIA: object-based image analysis; Superv: supervised classification, without more specifications; Unsuperv: unsupervised classification; Hybrid: hybrid classifier; ML: maximum likelihood; SVM: support vector machine; kNN: k-nearest neighbors; Tree reg: tree regression analysis. Source: own elaboration.
Figure 9. Methods used for obtaining LUC maps only including those developed with RS (128 papers). NE: not explained; Photo: photointerpretation of satellite images; OBIA: object-based image analysis; Superv: supervised classification, without more specifications; Unsuperv: unsupervised classification; Hybrid: hybrid classifier; ML: maximum likelihood; SVM: support vector machine; kNN: k-nearest neighbors; Tree reg: tree regression analysis. Source: own elaboration.
Land 11 01065 g009
Figure 10. Statistical techniques used for analyzing driving forces of LUCC (129 papers). CLUE: conversion of land use and its effects; EDR: exploratory data reduction; EDT: exploratory data transformation; EC: exploratory classification; EGM: exploratory graphical model; ES: exploratory summarization; TR: training; REG: regression; QUAL: qualitative analysis. Source: own elaboration.
Figure 10. Statistical techniques used for analyzing driving forces of LUCC (129 papers). CLUE: conversion of land use and its effects; EDR: exploratory data reduction; EDT: exploratory data transformation; EC: exploratory classification; EGM: exploratory graphical model; ES: exploratory summarization; TR: training; REG: regression; QUAL: qualitative analysis. Source: own elaboration.
Land 11 01065 g010
Figure 11. Regression techniques (49 papers). GWR: geographically weighted regression; PLSR: partial least squares regression; LOR: logistic regression; LIR: linear regression; SUR: seemingly unrelated regression; BRT: boosted regression trees. Source: own elaboration.
Figure 11. Regression techniques (49 papers). GWR: geographically weighted regression; PLSR: partial least squares regression; LOR: logistic regression; LIR: linear regression; SUR: seemingly unrelated regression; BRT: boosted regression trees. Source: own elaboration.
Land 11 01065 g011
Figure 12. Most-used calculated landscape metrics by number of manuscripts. Own elaboration.
Figure 12. Most-used calculated landscape metrics by number of manuscripts. Own elaboration.
Land 11 01065 g012
Figure 13. Number of manuscripts (29 in total) including models of future scenarios, by year. CLUE-S: conversion of land use and its effects at small regional extent; Land CM: Land Change Modeler; MLP + ANN: multi-layer perceptron neural network; ANN-CA: artificial neural network-based cellular automata; CA: cellular automata.
Figure 13. Number of manuscripts (29 in total) including models of future scenarios, by year. CLUE-S: conversion of land use and its effects at small regional extent; Land CM: Land Change Modeler; MLP + ANN: multi-layer perceptron neural network; ANN-CA: artificial neural network-based cellular automata; CA: cellular automata.
Land 11 01065 g013
Table 1. Parameters collected in the first STEPLand step, corresponding to “LUCC patterns and dynamics”.
Table 1. Parameters collected in the first STEPLand step, corresponding to “LUCC patterns and dynamics”.
StepSTEPLand AimParameters Reviewed
1stLUCC patterns and dynamicsAffiliation of first author or co-author
Journal of publication
EO source:
Land-cover maps produced by scientific or governmental agencies
The same but combined with aerial photographs or remote sensing imagery
Only aerial photographs
Aerial photographs and remote sensing imagery
Only remote sensing imagery
Temporal extent:
Up 10 years
From 11 to 15 years
From 16 to 20 years
From 21 to 25 years
From 26 to 30 years
More than 30 years
Country/ies of the study area
Number of LULC categories:
From 1 to 5
From 6 to 10
From 11 to 15
From 16 to 20
More than 20
Classification method:
Photointerpretation
Pixel classification, differentiating unsupervised from supervised (maximum likelihood, support vector machine; kNN; decision tree or random forest; artificial neural network)
Subpixel classification
Object classification
LULC map accuracy of automatic classifications:
Between 60 and 70%
Between 71 and 80%
Between 81 and 90%
More than 90%
Overlay method
Table 2. Parameters collected in the second STEPLand step, corresponding to “LUCC modeling”.
Table 2. Parameters collected in the second STEPLand step, corresponding to “LUCC modeling”.
StepSTEPLand AimDefinition
2ndLUCC modellingType and objective of modelling techniques:
CLUE
Exploratory data reduction
Data transformation
Data classification
Data graphical model
Data summarization
Data training
Regression
Type of explanatory driving forces:
Only biophysical or geophysical
Only socioeconomic
Socioeconomic and biophysical
Socioeconomic, biophysical and distances
Biophysical and distances
Type of dependent or response variables:
LUC
LUCC
Spatial scale of analysis or domain:
Plots, parcels o fields
Grids, cells or pixels
Administrative boundaries
Causal nomenclature:
Factors
Causes
Drivers
Determinants
Table 3. Parameters considered in the third step, corresponding to “Socio-environmental consequences of LUCC”.
Table 3. Parameters considered in the third step, corresponding to “Socio-environmental consequences of LUCC”.
StepSTEPLand AimDefinition
3rdLUCC consequencesSocio-environmental:
Water quality
Agriculture and forestry (agricultural land degradation)
Biodiversity (habitats, native species, …)
Natural disasters (forest fires, droughts, land or soil degradation, …)
Climate
Weather
Energy
Landscape:
Software used
Main metrics at class level or/and landscape level
Ecosystem services:
Evaluation model
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Serra, P.; Salvati, L. Land Change Science and the STEPLand Framework: An Assessment of Its Progress. Land 2022, 11, 1065. https://doi.org/10.3390/land11071065

AMA Style

Serra P, Salvati L. Land Change Science and the STEPLand Framework: An Assessment of Its Progress. Land. 2022; 11(7):1065. https://doi.org/10.3390/land11071065

Chicago/Turabian Style

Serra, Pere, and Luca Salvati. 2022. "Land Change Science and the STEPLand Framework: An Assessment of Its Progress" Land 11, no. 7: 1065. https://doi.org/10.3390/land11071065

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