Next Article in Journal
Experimental Research and Parameter Optimization on Dust Emission Reduction for Peanut Pickup Combine Harvesting
Previous Article in Journal
The Infection of Yellow Lupin (Lupinus luteus L.) with Bean Yellow Mosaic Virus (BYMV) and Cucumber Mosaic Virus (CMV) in Organic Farming in Eastern Poland
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Enhancing Agroecological Resilience in Arid Regions: A Review of Shelterbelt Structure and Function

State Key Laboratory of Desert and Oasis Ecology, Xinjiang Institute of Ecology and Geography, Chinese Academy of Sciences, Urumqi 830011, China
*
Author to whom correspondence should be addressed.
Agriculture 2025, 15(19), 2004; https://doi.org/10.3390/agriculture15192004
Submission received: 29 August 2025 / Revised: 19 September 2025 / Accepted: 23 September 2025 / Published: 25 September 2025
(This article belongs to the Section Ecosystem, Environment and Climate Change in Agriculture)

Abstract

Farmland shelterbelts are vital ecological infrastructure for sustaining agriculture in arid regions, where high winds, soil erosion, and water scarcity severely constrain productivity. While their protective functions—reducing wind speed, controlling erosion, moderating microclimates, and enhancing yields—are well documented, previous studies have largely examined individual structural elements in isolation, leaving their interactive effects and trade-offs poorly understood. This review synthesizes current research on the structural optimization of shelterbelts, emphasizing the critical relationship between their physical and biological attributes and their protective functions. Key structural parameters—such as optical porosity, height, width, orientation, and species composition—are examined for their individual and interactive impacts on shelterbelt performance. Empirical and modeling studies indicate that moderate porosity maximizes wind reduction efficiency and extends the leeward protection zone, while multi-row, multi-species configurations effectively suppress soil erosion and improve microclimate conditions. Sheltered areas experience reduced evapotranspiration, increased humidity, and moderated temperatures, collectively enhancing crop water use efficiency and yielding significant improvements in crop production. Advanced methodologies, including field monitoring, wind tunnel testing, computational fluid dynamics, and remote sensing, are employed to quantify benefits and refine designs. A multi-objective optimization framework is essential to balance competing goals: maximizing wind reduction, minimizing water consumption, enhancing biodiversity, and ensuring economic viability. Future challenges involve adapting designs to climate change, integrating water-efficient and native species, leveraging artificial intelligence for predictive modeling, and addressing socio-economic barriers to implementation. Building on this evidence, we propose a multi-objective optimization framework to balance competing goals: maximizing wind protection, minimizing water use, enhancing biodiversity, and ensuring economic viability. We identify key research gaps including unresolved porosity thresholds, the climate resilience of alternative species compositions, and the limited application of optimization algorithms and outline future priorities such as region-specific design guidelines, AI-driven predictive models, and policy incentives. This review offers a novel, trade-off–aware synthesis to guide next-generation shelterbelt design in arid agriculture.

1. Background

Shelterbelts, or agroforestry windbreaks, are indispensable components of oasis agricultural systems, serving as a primary defense against the harsh environmental conditions prevalent in arid and semi-arid regions [1,2,3]. These areas, which account for a substantial proportion of the world’s land surface and support hundreds of millions of people, confront a range of interconnected and severe challenges [4]. Characterized by innate water scarcity, elevated temperatures, unpredictable and low rainfall, as well as persistent and powerful winds, such environments are intrinsically vulnerable to natural degradation. Among the most critical threats is soil erosion—both aeolian and fluvial—which steadily depletes the fertile topsoil essential for sustaining crop production [5,6,7]. Moreover, the persistent winds significantly amplify evapotranspiration, often causing water loss to far exceed precipitation [8]. This imbalance subjects crops to prolonged physiological stress, compromising growth and leading to considerable reductions in yield. Within this context, shelterbelts perform crucial ecological functions: mitigating wind speed, stabilizing soil microenvironments, conserving moisture, and ultimately enhancing the resilience and productivity of oasis farming systems [9,10].
The significance of shelterbelts as fundamental components of agroecological engineering in arid regions is of paramount importance. At its core, a shelterbelt is a barrier of trees and shrubs strategically planted to reduce wind speed and alter the microclimatic conditions in its leeward and windward vicinity [11,12]. By intercepting and deflecting the prevailing winds, a well-designed shelterbelt creates a zone of reduced wind velocity, which yields a cascade of beneficial effects [13]. The most immediate and direct effect is a significant decrease in wind-induced soil erosion, as the vegetation shield protects the soil surface from the direct force of the wind, thereby minimizing the dislodging and movement of soil particles [14]. This protective function is inextricably linked to the conservation of soil moisture; the calmer microclimate behind the shelterbelt significantly lowers evapotranspiration rates, allowing for more efficient water use by crops and reducing irrigation requirements—a critical advantage in water-limited settings [15]. Beyond these core functions, shelterbelts contribute to enhanced crop yields by minimizing mechanical damage to plants, improving pollination efficiency, and moderating air temperatures [16]. They also offer a range of ancillary ecosystem services, including carbon sequestration, habitat provision for biodiversity (thereby enhancing biological pest control), landscape connectivity, and in some cases, the production of timber, fuelwood, or non-timber forest products, thereby diversifying farm income [17,18].
The efficacy of a shelterbelt in delivering these benefits is not a simple binary outcome of its presence or absence; rather, it is profoundly and intricately governed by its physical and biological structure. In the specific context of shelterbelt design and function, “structure” is a multidimensional concept encompassing a suite of quantifiable characteristics. Key structural parameters include the height of the tallest trees, which primarily determines the extent of the wind-protected zone—often measured in multiples of height, e.g., 10 H, 20 H leeward [19]. Density or porosity is perhaps the most critical structural attribute, referring to the ratio of solid surface area to the total area of the shelterbelt when viewed perpendicular to the wind direction; it dictates the degree of wind speed reduction and the pattern of airflow (whether air streams through the belt or is forced over it). Width (number of rows of trees/shrubs), species composition (deciduous vs. coniferous, native vs. exotic), and the vertical arrangement of vegetation (uniform vs. stratified canopies) are other vital structural components [20,21]. The geometric configuration, which encompasses orientation relative to prevailing damaging winds and the overall spatial layout on the landscape (e.g., isolated belts, interconnected networks or grids), is also considered part of the structural framework. It is the intricate interaction among these structural components that determines the aerodynamic properties of the shelterbelt, thereby affecting the spatial patterns and extent of microclimatic changes and, consequently, its overall agroecological performance [22,23,24].
Shelterbelts also occupy land that could otherwise be used for crop production, representing an opportunity cost for farmers. A non-optimized shelterbelt, for instance one that is too dense, might create excessive turbulence or too large a zone of shade and root competition, reducing yields immediately downwind. Conversely, a belt that is too sparse may provide insufficient protection and fail to deliver the expected benefits. The challenge therefore lies not merely in planting shelterbelts, but in designing them to balance ecological benefits (erosion control, microclimate improvement, biodiversity) against costs (land occupation, water consumption, competition) [25,26]. Optimization aims to achieve the highest net positive benefit per unit of resource invested, ensuring that shelterbelt systems are ecologically effective, economically viable, and socially acceptable.
Despite decades of research, previous studies have struggled to resolve several critical issues [27]. Most past work has focused on isolated structural elements—such as height, porosity, or width—examining their effects on wind reduction or yields in single-objective terms. While valuable, this approach neglects trade-offs among multiple functions, such as the tension between maximizing wind protection and minimizing water use, or between biodiversity enhancement and management costs. Furthermore, results are often contradictory: some studies report 40% porosity as optimal, while others suggest 50%, and reported protection distances vary widely (15 H to >30 H). These inconsistencies remain largely unreconciled, and few meta-analyses have systematically compared findings across climatic and structural contexts. Only a handful of studies have attempted formal multi-objective optimization of shelterbelt design, typically applying genetic algorithms, NSGA-II, or Pareto-based approaches to simulated datasets, yet most lacked empirical validation, long-term performance data, or region-specific calibration. As a result, their outputs remain difficult to apply in real-world agricultural systems, and their relevance for policy and farmer decision-making is limited.
This review addresses these unresolved gaps by providing a critical, integrative synthesis of how shelterbelt structural attributes influence agroecological performance in arid regions through a multi-objective lens. Specifically, we (i) consolidate and evaluate empirical findings, theoretical models, and design principles, (ii) critically compare evidence across studies to explain divergent results, (iii) assess the strengths and limitations of field, wind tunnel, CFD, and remote sensing methods used to study shelterbelts, and (iv) outline pathways for multi-objective optimization that explicitly incorporate ecological, hydrological, and economic criteria. By highlighting contradictions, trade-offs, and uncertainties, this review goes beyond previous descriptive overviews to offer a novel framework for designing shelterbelts that maximize multiple benefits while minimizing resource costs. Our aim is to transform the current fragmented knowledge base into a coherent evidence foundation to support rigorous, trade-off-aware optimization of shelterbelts in the world’s arid agricultural landscapes.

2. Review Methodology

This comprehensive review on the multi-objective optimization of windbreak systems in arid agricultural regions was conducted through a systematic and multi-stage methodology. The aim was to synthesize current knowledge, identify research gaps, and provide a critical analysis of the structural and functional relationships of shelterbelts. The methodology encompassed literature search, screening, data extraction, and thematic synthesis, as detailed below.

2.1. Literature Search Strategy

A systematic literature search was conducted to identify relevant peer-reviewed articles, books, conference proceedings, and technical reports published up to early 2024. The primary electronic databases included: Web of Science, Scopus, Google Scholar, PubMed (for interdisciplinary health and environmental studies), AGRICOLA and CAB Abstracts (for agricultural and ecological literature).
The search strategy employed a combination of keywords and Boolean operators to maximize coverage and relevance. Key search terms included: shelterbelt” OR “windbreak”, “arid regions” OR “drylands”, “structural optimization”, “microclimate modification”, “soil erosion control”, “multi-objective optimization”, “computational fluid dynamics (CFD)”, “remote sensing in agroforestry” and “water use efficiency”.
Additional searches were performed using specific terms related to shelterbelt functions, such as “porosity,” “height effects,” “species composition,” and “ecosystem services.” The reference lists of key articles were also hand-searched to identify further relevant publications (snowballing method).

2.2. Inclusion and Exclusion Criteria

To ensure the relevance and quality of the literature included in this review, the following criteria were applied. Inclusion criteria were included: Studies focused on shelterbelts or windbreaks in arid or semi-arid agricultural contexts; Research investigating structural parameters (e.g., porosity, height, species mix) and their impact on ecological or agricultural outcomes; Empirical, modeling, or review studies published in English or Chinese (with English abstracts); Publications from both historical and recent periods to capture evolutionary trends in research. The Exclusion criteriawere including: studies not directly related to agricultural windbreaks (e.g., urban windbreaks, coastal barriers); Publications without empirical data or clear methodological descriptions and duplicate studies or those with overlapping data.

2.3. Data Extraction and Synthesis

  • Relevant data were extracted from each included study using a standardized form capturing:
    • Study characteristics: author(s), year, location, study type (field, model, experiment)
    • Study characteristics: author(s), year, location, study type (field, model, experiment)
    • Shelterbelt structural parameters: porosity, height, width, orientation, species composition
    • Functional outcomes: wind speed reduction, soil erosion control, microclimatic changes, crop yield improvements
    • Methodologies used: field measurements, wind tunnel experiments, CFD modeling, remote sensing
    • Key findings and conclusions
  • The extracted data were synthesized thematically to address the review’s objectives. Themes included:
    • Wind flow dynamics and porosity effects
    • Microclimatic modifications and water conservation
    • Soil erosion mechanisms and control
    • Crop yield and quality improvements
    • Additional ecosystem services (biodiversity, carbon sequestration)
    • Optimization frameworks and trade-offs

2.4. Quality Assessment and Critical Analysis

Each study was assessed for methodological rigor, sample size, representativeness, and relevance to arid agro-ecosystems. Studies employing controlled experiments, validated models, or long-term monitoring were given higher weight. The strengths and limitations of different methodological approaches (e.g., CFD Vs. Field studies) were critically evaluated to provide a balanced perspective. Given the interdisciplinary nature of shelterbelt research, insights from ecology, fluid dynamics, climatology, remote sensing, and socio-economics were integrated. This holistic approach allowed for a comprehensive understanding of how structural design influences multiple functions and services. Throughout the review process, attention was paid to inconsistencies, contradictions, and underexplored areas in the literature.

3. Protective Functions of Shelterbelts: Linking Structure to Agroecosystem Processes

3.1. Wind Speed Reduction and Flow Field Alteration

The primary function of shelterbelts in arid farmland ecosystems is to attenuate wind speed and fundamentally alter the structure of the wind flow field, thereby creating a protected micro-environment leeward of the barrier [28]. This process is not merely a simple blocking action but involves complex interactions between the wind and the porous structure of the shelterbelt, leading to energy dissipation, flow separation, and the formation of a characteristic wind speed reduction zone [29].
The effectiveness of wind speed reduction is primarily influenced by the optical porosity—a parameter reflecting the density of the vegetation barrier—as well as the structural attributes of the shelterbelt, such as its height (H), width, and species composition [10,30,31]. A shelterbelt with moderate porosity (typically between 40% and 50%) is generally regarded as optimal [32,33]. Highly dense barriers (low porosity) act as strong windbreaks immediately downwind, but tend to generate considerable turbulence and a short recovery distance, as airflow is deflected upward and over the barrier, resulting in intense downward shear leeward [34,35]. In contrast, highly porous shelterbelts provide less initial wind reduction but produce a longer and more gradual sheltered area with diminished turbulence [36]. The protected zone is commonly described as extending leeward over a distance of 15–30 H, within which wind speed is reduced by more than 20%. The most effective protection typically occurs within 5–10 H [37].
The modification of the flow field constitutes a fundamental mechanism. As wind approaches the shelterbelt, it slows down, and a portion is diverted through the vegetation. There, friction with stems and leaves dissipates its kinetic energy [38]. The remaining airflow is deflected upward and over the shelterbelt. Directly leeward, a low-speed sheltered zone forms, often characterized by a small recirculation eddy [39]. Farther downwind, the decelerated air gradually mixes with the faster-moving airflow from above, leading to an asymptotic recovery of wind speed toward the open-field value [40]. This altered flow field directly affects other protective functions. The reduction in wind speed lowers the wind’s shear stress on the soil surface, which is the primary driver of aeolian erosion [41,42,43]. Furthermore, the altered airflow patterns affect the turbulent transfer of heat, water vapor, and gases between the crop canopy and the atmosphere, leading to modifications in microclimatic factors such as temperature, humidity, and evapotranspiration rates [44]. Computational Fluid Dynamics (CFD) models and wind tunnel experiments have been instrumental in quantifying these complex flow patterns and optimizing shelterbelt structures for maximum aerodynamic efficiency in arid regions, where strong winds are a major cause of soil and water loss [45].
Notably, reported optimal porosity values vary: while several studies suggest ~40% [46], others find 50% yields greater leeward protection [47]. These discrepancies likely stem from differences in wind regimes, belt heights, and vegetation architecture among sites. Recognizing such context-dependence is essential, as assuming a single universal optimum can misguide design. Future meta-analyses should stratify data by climate and structural conditions to resolve these inconsistencies.

3.2. Suppression of Soil Aeolian Erosion

In arid and semi-arid agricultural regions, soil is highly susceptible to deflation due to sparse vegetation cover, frequent high-wind events, and often dry soil conditions [48,49,50]. Aeolian erosion removes the nutrient-rich topsoil, abrades young crops, and can lead to sand deposition that buries plants and infrastructure [51,52].
Shelterbelts mitigate erosion through two primary physical mechanisms. First, as detailed in Section 2.1, they reduce the wind velocity in the leeward zone below the critical threshold velocity required to initiate soil particle movement entrainment [53]. This is the most significant factor. Second, the vegetation itself acts as a physical trap for saltating and suspended soil particles. The standing biomass and leaf litter within and immediately behind the belt capture moving sand grains, reducing the overall sediment flux [54,55].
The effectiveness of a shelterbelt in controlling erosion is highly dependent on its structure. A multi-row, multi-species belt with a mix of trees and shrubs is generally more effective than a single row of trees [22]. Shrubs and ground cover at the base of the trees prevent the formation of scour channels and ensure uniform wind protection down to the soil surface [56]. The area of maximum erosion protection typically extends to a distance of 10–15 H leeward, with efficiency gradually decreasing with distance as wind speed recovers [10,57] (Table 1).

3.3. Modification of Microclimatic Factors

Beyond reducing wind speed, shelterbelts significantly modify the microclimatic conditions within the protected zone, creating a more favorable environment for crop growth in arid regions [13,22]. The altered aerodynamic regime influences the energy balance and the turbulent exchange processes, leading to changes in air and soil temperature, atmospheric humidity, and evapotranspiration [58].
During the day, the sheltered area often experiences slightly higher air temperatures compared to the open field. This is because the reduced wind speed diminishes the advective removal of heat generated by solar radiation (i.e., reduced convective cooling) [59]. Soil temperatures in the sheltered zone are also moderated, being warmer during cool periods and slightly cooler during hot periods, which can promote better seed germination and root development [60]. A more pronounced effect is the increase in absolute humidity within the crop canopy. The reduced wind speed traps water vapor transpired by plants and evaporated from the soil, leading to a higher humidity layer near the crops [61]. This higher humidity gradient between the leaf surface and the atmosphere directly reduces the evapotranspiration rate. Studies have consistently shown a reduction in ET rates by 10% to 30% in the sheltered area compared to exposed fields [62]. This water conservation is arguably the most valuable microclimatic benefit in water-limited arid regions, as it enhances crop water use efficiency (WUE) [63,64,65]. The following flowchart illustrates the causal chain of microclimatic modification (Figure 1).
These modified conditions reduced ET, higher humidity, and moderated temperatures collectively reduce water stress on crops, particularly during periods of high atmospheric demand (e.g., hot, windy days), leading to more sustained physiological activity and growth.

3.4. Enhancement of Crop Yield and Quality

The culmination of the protective functions of shelterbelts is the observed improvement in crop yield and quality [16]. The mechanisms through which wind reduction, erosion control, and microclimatic modification translate into agricultural benefits are complex and interactive, ultimately affecting key physiological processes in plants [66,67,68].
The reduction in wind stress prevents physical damage to crops, such as leaf tearing, sandblasting of seedlings, and lodging (uprooting) of mature plants [69]. This allows plants to maintain a healthier and more photosynthetically active leaf area. The improved microclimate, particularly reduced evapotranspiration, mitigates water stress [70]. This allows plants to keep their stomata open for longer periods, facilitating greater uptake of carbon dioxide for photosynthesis [71,72]. Consequently, photosynthetic rates are often higher in sheltered fields [73]. The energy that a plant would otherwise expend on repairing damage or closing stomata can be redirected towards growth and reproduction, i.e., biomass accumulation and grain or fruit production [74]. Furthermore, the reduction in erosion helps preserve soil fertility and organic matter, ensuring that more nutrients and water are available for crop uptake [75].
The result is typically a significant increase in yield. Meta-analyses have shown that well-designed shelterbelt systems can increase crop yields in arid regions by 10% to 25%, with the highest increases occurring in years with severe drought or wind events [76,77]. The yield improvement is not uniform across the field; it often follows a pattern corresponding to the wind reduction pattern, with the highest yields recorded at a distance of 3–7 H leeward [78]. Besides quantity, crop quality can also be enhanced. For instance, reduced sandblasting leads to cleaner produce with less surface scarring, and moderated water stress can improve the nutritional content or size of fruits and grains [79]. The pathway from shelterbelt function to yield improvement is summarized in the flowchart below (Figure 2).
Although numerous studies have reported yield gains of 10% to 25% from shelterbelt protection in arid regions [76,77], the magnitude of improvement varies considerably across contexts. Reported differences can be attributed to several interacting factors, including local wind climate, shelterbelt porosity and height, crop species, soil fertility, and water availability. For example, belts with low porosity may produce strong protection but also increase shading and root competition near the belt edge, sometimes suppressing yields in the first few leeward rows, while more porous designs extend protection farther downwind but reduce the peak benefit. Yield responses are also strongly influenced by interannual climate variability, with larger gains typically recorded in drought or wind-stress years and smaller gains in favorable seasons. These inconsistencies underscore that yield outcomes cannot be generalized from single-site studies, and that future meta-analyses should integrate structural, climatic, and management variables to explain the heterogeneous responses observed across different agroecosystems.

3.5. Additional Ecosystem Co-Benefits

While the primary focus of agricultural shelterbelts is often on protecting crops and soil, they provide a suite of additional ecosystem services that are increasingly recognized for their ecological and economic value [13]. These co-benefits significantly enhance the overall sustainability of arid farming landscapes.
Carbon Sequestration: Shelterbelts act as permanent carbon sinks within agricultural systems. The trees and shrubs continuously absorb atmospheric CO2 through photosynthesis, storing carbon in their biomass (stems, branches, roots) and in the soil [80]. Soil organic carbon can accumulate under shelterbelts due to the input of leaf litter and root exudates and the reduced mineralization rates associated with modified microclimates [81]. This function contributes to climate change mitigation by offsetting greenhouse gas emissions from farming operations [82].
Biodiversity Enhancement: In the often simplified and monocultural landscape of intensive agriculture, shelterbelts serve as vital linear habitats and biodiversity corridors [22,83]. They provide food, shelter, nesting sites, and overwintering grounds for a wide array of organisms. This includes beneficial arthropods (e.g., pollinators like bees and natural pest predators like spiders and beetles), birds, and small mammals [84,85]. The increased diversity of pollinators directly supports crop production, while the presence of natural predators helps regulate pest populations, potentially reducing the need for chemical pesticides [86].
Habitat Provision: Beyond specific species, the shelterbelt structure creates a distinct forest-edge habitat that is different from both the open field and large, continuous forests [87]. This habitat heterogeneity at the landscape scale is a key driver of biodiversity. Furthermore, shelterbelts can act as connecting corridors, allowing species to move between otherwise isolated habitat patches, which is crucial for maintaining genetic diversity and ecological resilience in fragmented agricultural landscapes [88,89].
Overall, these additional benefits transform shelterbelts from mere windbreaks into multifunctional ecological infrastructure. They not only protect the farm from wind damage and conserve water but also contribute to climate regulation, support ecosystem health through biodiversity conservation, and enhance the aesthetic value of the agricultural landscape [13].

4. Structural Determinants of Shelterbelt Performance in Arid Regions

The functional efficacy of a shelterbelt is not a product of chance but a direct consequence of its deliberate structural design. This section deconstructs the shelterbelt system into its core structural elements porosity, height, width, orientation, shape, and species composition and elucidates the mechanistic relationships between each element and its ultimate protective function in arid agricultural landscapes.

4.1. Porosity and Optical Porosity

Porosity, defined as the ratio of pore space to solid matter within a shelterbelt, and its two-dimensional representation, optical porosity (the percentage of visible sky through a shelterbelt when viewed perpendicularly), are unequivocally the most critical structural parameters governing wind flow dynamics [90,91]. They primarily determine the balance between wind speed reduction and the undesirable downwind turbulence [29,92].
Dense shelterbelts (low optical porosity, <20%) act more like a solid wall, forcing the majority of the wind to deflect over the top. This creates a large pressure gradient, leading to intense mixing and the rapid recovery of wind speed shortly downwind [93,94]. While the immediate leeward zone experiences extreme wind reduction, the area of protection is short, and soil erosion can be exacerbated due to high wind shear at the top of the belt [95]. Conversely, medium-porosity shelterbelts (optical porosity of 30–50%) are considered optimal. They allow a portion of the wind stream to penetrate through the pores, dissipating wind energy through friction within the belt itself [96]. This results in a more gradual slow-down of wind, a significantly extended leeward protected zone (often 20–30 times the shelterbelt height, H), and reduced turbulence [97]. The wind speed reduction is more homogeneous and sustainable. High-porosity belts (>60%) offer too little resistance, providing minimal wind reduction and a very limited zone of influence [98].
In arid regions, optimizing porosity is further complicated by the need to manage evapotranspiration. While a denser belt may offer more shade, it can also compete fiercely with crops for scarce water [50]. Therefore, achieving the ideal porosity is a trade-off between maximizing wind protection, minimizing turbulence, and conserving soil moisture. Optical porosity, being easily measurable from digital side-view images, serves as a practical and reliable proxy for managing this key determinant [99]. The relationship between shelterbelt porosity, wind flow pattern and functional outcomes are shown in Figure 3. While lower porosity maximizes wind protection, it also increases shading and root competition, elevating crop water stress—an often-overlooked trade-off. Thus, porosity optimization must balance wind reduction benefits with potential water competition costs.

4.2. Height (H): Scaling the Protected Zone

The height (H) of a shelterbelt is the principal factor determining the absolute spatial extent of its microclimatic influence, particularly in the leeward direction [100]. The protected area, often defined as the region where wind speed is reduced by at least 20%, is consistently measured as a multiple of H. Empirical and modelling studies show that the maximum protective distance typically ranges from 15 H to 30 H, depending on other factors like porosity [101].
The aerodynamic effect of the shelterbelt creates a zone of reduced wind stress and altered turbulence that scales with its height [29]. A taller shelterbelt projects this modified microclimate further across adjacent farmland. Therefore, in arid areas where protecting large fields from wind erosion and evapotranspiration is paramount, maximizing the height of shelterbelt trees (using species like poplars or casuarina suited to the local environment) is a primary objective [13,102]. However, height is a double-edged sword. Increased height also amplifies the “rain shadow” effect, potentially reducing precipitation reaching the center of very large fields, and increases competition for light and water with crops [103]. Thus, the optimal field size and shelterbelt height must be carefully matched to balance wind protection with resource competition and microclimatic impacts [104]. The scaling effect of shelterbelt height (H) on the zone of influence are illustrated in Figure 4.

4.3. Width and Internal Complexity

Width influences shelterbelt function indirectly through its strong interaction with density (porosity). A very narrow belt (1–2 tree rows) often has low density and limited internal structure, restricting its ability to provide habitat [28]. Increasing width initially allows for a more complex internal structure and can help achieve the target optical porosity by allowing for multiple rows of trees and shrubs without becoming excessively dense [93,105]. However, beyond a certain threshold (often 4–6 rows for timber belts), increased width directly translates to increased density, shifting the aerodynamic function towards that of a solid barrier with associated negative turbulence effects [106].
From an ecological perspective, width is paramount. Wider shelterbelts provide a larger core habitat area, buffering sensitive wildlife species from predation and disturbance from adjacent farmland [28,83]. They support greater biodiversity of birds, insects, and mammals by offering more varied niches, nesting sites, and food resources [107]. In arid regions, these belts can act as crucial biological corridors, connecting fragmented habitats [108]. The optimization challenge is to find the minimum width that provides sufficient internal habitat and the desired porosity for wind protection while minimizing the loss of arable land and water consumption [109].

4.4. Orientation to Prevailing Winds

Orientation refers to the compass direction of the shelterbelt’s long axis. Its efficiency is entirely contingent [110]. A perpendicular orientation presents the maximum effective area to the wind, ensuring the widest and most uniform zone of protection leeward. Even minor deviations from perpendicularity significantly reduce protection efficiency; a 30-degree misalignment can reduce the protected area by up to 50% [111].
In arid regions, the primary threat is often wind erosion during hot, dry seasons [112]. Therefore, shelterbelts must be aligned based on a detailed analysis of historical wind rose data for those critical periods [13]. However, multi-purpose systems may need to consider cold winter winds or spring sandstorms with different directions, necessitating a coordinated network of shelterbelts rather than reliance on a single orientation [113]. The layout becomes a grid system designed to address winds from multiple vectors, enhancing protection and ecological resilience [114,115].

4.5. Cross-Sectional Shape and Vertical Stratification

The cross-sectional profile (e.g., rectangular, triangular, streamlined) influences how wind streamlines are guided over and through the belt. A dense, rectangular profile promotes strong lift and abrupt deflection, exacerbating turbulence [116]. A tapered or triangular profile (achieved by planting shorter shrubs on the windward and leeward edges with taller trees in the center) encourages a more gradual ascent and descent of the wind, smoothing the flow and reducing turbulent mixing [117]. This aerodynamic shaping effectively extends the quiet zone leeward.
Vertical structure the arrangement of different plant heights—is key to achieving this optimal shape and functionality. A multi-layered structure, with trees, shrubs, and ground cover, protects a deeper vertical layer of air [76]. The shrub layer is particularly critical as it filters the near-surface wind, which is most responsible for soil particle transport. In arid zones, this complex structure also creates a more humid microclimate within the belt, reducing water stress on the vegetation itself and enhancing its sustainability [118,119].

4.6. Species Composition and Functional Trade-Offs

The choice of species dictates the physical and ecological attributes of the shelterbelt. Native, drought-tolerant, and deep-rooted species are essential for sustainability in arid regions [83,120]. The configuration—how these species are arranged—directly creates the previously discussed structural elements (porosity, shape).
A multi-species, multi-layered configuration is superior. Tall, fast-growing trees form the upper canopy and define the height (H). Dense, thorny shrubs in the understory provide the desired medium optical porosity, filter wind near the ground, and offer habitat [121,122]. Ground cover grasses or herbaceous plants suppress dust and prevent internal erosion [123,124]. This combination ensures year-round protection (deciduous trees may lose effectiveness in winter) and enhances biodiversity [123,125,126]. Nitrogen-fixing species can be incorporated to improve soil fertility [127,128]. The configuration must also consider allelopathy and root competition to avoid negatively impacting adjacent crops (Figure 5). Species with dense canopies provide strong shelter but often have high water demand, whereas xerophytic shrubs are water-efficient yet offer weaker wind reduction. These functional trade-offs should guide species mixing, but few studies have quantitatively compared them an urgent research need.

5. Methodologies for Studying and Optimizing Structure

5.1. Field Investigation Methods for Shelterbelts

Field-based measurements are fundamental for understanding the real-world performance of shelterbelts in arid regions [129,130]. These empirical studies involve direct assessments of structural parameters such as tree height, porosity, width, and species composition and their interactions with microclimatic variables like wind speed reduction, soil moisture retention, and evapotranspiration rates [131]. Techniques include anemometers for wind profiling, soil moisture sensors, dendrometric tools for tree morphology, and photogrammetric methods for porosity estimation [132]. Such studies provide critical validation data for models but face challenges like spatial variability, time-intensive data collection, and difficulties in isolating individual structural factors [133]. Field studies provide realistic validation but are labor-intensive, spatially limited, and confounded by uncontrolled variables. Key field measurement techniques and parameters in shelterbelt studies are illustrated in the Table 2.

5.2. Controlled Investigation of Structural Parameters

Wind tunnel experiments offer controlled conditions to systematically analyze the effects of shelterbelt structural parameters e.g., porosity, arrangement, and height on wind flow dynamics [134,135]. Scaled models of shelterbelts are exposed to varying wind speeds and directions, while sensors measure velocity profiles, pressure differentials, and turbulence patterns [29]. These experiments isolate specific variables impractical to study in field settings, enabling precise optimization of design features like optimal porosity for wind reduction efficiency [109]. However, scaling effects and simplification of vegetation properties may limit direct extrapolation to real-world conditions [136,137]. Wind tunnels isolate variables but face scaling distortions and oversimplified vegetation properties, limiting real-world transferability (Table 3).

5.3. Computational Fluid Dynamics (CFD) Simulations

Computational Fluid Dynamics (CFD) simulations have become a powerful tool for simulating wind flow interactions with shelterbelts at high spatial resolutions [29,138]. Models solve Navier–Stokes equations to predict airflow patterns, shear stress, and protective efficacy under varying structural configurations. Key approaches include Reynolds-Averaged Navier–Stokes (RANS) for steady-state analysis and Large Eddy Simulation (LES) for turbulent flow dynamics [139,140]. CFD allows rapid testing of hypothetical scenarios—e.g., climate change adaptations but requires validation against empirical data and significant computational resources [141,142,143].
Despite their power, CFD models are constrained by high computational cost, simplification of vegetation as homogeneous porous media, and sensitivity to boundary conditions, which can limit realism. CFD offers flexible scenario testing but requires high computational power and simplifies vegetation, risking overconfidence without empirical validation (Table 4).

5.4. Remote Sensing and GIS for Large-Scale Structural Assessment and Planning

Remote sensing (RS) and Geographic Information Systems (GIS) enable large-scale, spatially explicit analysis of shelterbelt structures and their ecological functions [144,145]. Satellite and UAV-based imagery (e.g., multispectral, LiDAR) extract metrics like canopy health, coverage continuity, and height variation [146,147]. GIS integrates spatial data on soil, climate, and land use to identify priority zones for shelterbelt expansion or restoration [148]. This approach supports regional planning but may lack the granularity of ground-based methods and requires validation for structural parameter accuracy. Remote sensing provides broad coverage but lacks the fine structural detail of ground surveys and depends on accurate calibration (Table 5).

6. Optimization Framework for Shelterbelt Design

The optimization of shelterbelt structure in arid regions transcends a single-objective problem. Pursuing maximal wind reduction alone could lead to unsustainable water consumption or neglect other ecosystem services [149,150]. Therefore, a multi-objective optimization (MOO) framework is essential to navigate the complex interplay of ecological, environmental, and economic factors. This framework aims to identify a set of Pareto-optimal solutions, where improvement in one objective (e.g., wind reduction) necessitates a compromise in another (e.g., reduced water use), rather than seeking a single universal optimum. This chapter delineates the core components of this MOO framework.

6.1. Maximizing Wind Reduction Efficiency

The primary and most immediate function of a shelterbelt is to modify the microclimate by reducing wind speed and mitigating wind erosion [48,151]. The optimization of this objective is fundamentally rooted in fluid dynamics and aerodynamic principles. Key structural parameters include porosity (the ratio of pore space to total area), height, width, and orientation relative to prevailing winds [152,153]. Computational Fluid Dynamics (CFD) models and semi-empirical formulae (e.g., the laws of Naegeli and Wang & Takle) are instrumental in simulating wind flow patterns around various shelterbelt configurations [154]. Optimization seeks to find the structure that maximizes the extent and degree of wind speed reduction in the leeward protected zone (often measured as the area where wind speed is reduced to below a critical erosion threshold). For instance, a medium porosity (30–40%) is often found to be optimal, as it effectively dissipates wind energy through the canopy rather than deflecting it abruptly, which can cause damaging turbulence [155]. The optimization process involves iteratively adjusting these structural variables within models to achieve the greatest protective area for crops with the most efficient use of tree biomass.

6.2. Minimizing Water Consumption

Water represents the primary limiting factor In arid environments [50,156]. Optimization models that fail to account for hydrological constraints are not viable in practice. The goal is to minimize the total evapotranspiration (ET) water usage within the shelterbelt system, while maintaining its protective efficacy. This is achieved through two key approaches: species selection and spatial configuration. Preference is given to drought-tolerant, deep-rooted, or phreatophytic species—such as Populus euphratica and Tamarix spp.—which exhibit reduced water demands or can utilize deeper groundwater sources [157]. Structurally, optimizing the spatial layout, including the number of rows, between-row spacing, and within-row spacing, significantly affects stand-level transpiration [158]. Moreover, the concept of “water-use efficiency” (WUE), defined as the wind reduction benefit per unit of water consumed, serves as a crucial performance indicator [159]. The optimization framework should incorporate soil-water balance models to simulate long-term water use under various shelterbelt designs, thereby preventing exacerbation of local water shortages or soil desiccation [160,161].

6.3. Integrating Biodiversity and Soil Health Co-Benefits

Biodiversity conservation is enhanced by designing multi-species, multi-layered (multi-strata) shelterbelts that provide diverse habitats, nesting sites, and food sources for birds, insects, and other fauna [162,163]. This contrasts with monoculture plantations, which are ecologically simplistic and vulnerable to pests. Soil health improvement is another key co-benefit. Shelterbelt roots help stabilize soil structure, reduce compaction, and enhance organic matter through litterfall [164,165]. Nitrogen-fixing species (e.g., certain Acacia or Robinia species) can be incorporated to improve soil fertility [166]. Optimizing for these benefits means moving beyond purely geometrical parameters to include compositional diversity. The challenge for the MOO framework is to quantitatively link structural and compositional variables to indicators of biodiversity (e.g., species richness indices) and soil quality (e.g., soil organic carbon content, aggregate stability) [167,168].

6.4. Balancing Economic Costs and Returns

The economic viability of shelterbelts is crucial for farmer adoption and long-term sustainability. Therefore, economic objectives must be explicitly included in the optimization. These can be framed as either maximizing benefits or minimizing costs. Timber and Non-Timber Forest Product (NTFP) production can provide a direct revenue stream, making species with high commercial value attractive [169,170]. However, this must be balanced against their water requirements. Costs include initial establishment (land preparation, saplings, planting labor, irrigation system installation) and ongoing maintenance (pruning, thinning, irrigation, pest control) [171]. A sparse, low-water-use design may have lower maintenance costs but might also offer less wind protection and economic return. The optimization model must perform a cost–benefit analysis over the shelterbelt’s lifecycle, discounting future timber income and ongoing costs to a net present value (NPV) [172]. The goal is to find designs that offer a favorable trade-off between ecological protection and economic return.

6.5. Navigating Trade-Offs and Synergies

The core of multi-objective optimization lies in managing the inherent trade-offs and leveraging potential synergies between the objectives. A clear trade-off exists between maximizing wind reduction (often requiring denser, taller, multi-row belts) and minimizing water consumption (favoring sparser, shorter, drought-tolerant species) [173]. Similarly, maximizing timber production might conflict with water minimization if fast-growing, high-yield species are also thirsty. Conversely, significant synergies exist: for example, promoting biodiversity through multi-species planting can enhance pest control (reducing maintenance costs) and improve soil health (reducing the need for fertilization) [174]. Similarly, a healthy, well-maintained shelterbelt will simultaneously achieve better wind reduction, carbon sequestration, and biodiversity. Advanced optimization algorithms (e.g., Genetic Algorithms, NSGA-II) are used to explore this complex solution space, generating a Pareto front that visually represents the best possible compromises. Decision-makers can then select a configuration from this front based on local priorities (e.g., water scarcity vs. erosion risk) [175].
Practical demonstrations of multi-objective optimization remain scarce but promising. For example, Verma et al. applied an NSGA-II algorithm to optimize shelterbelt porosity, height, and spacing in northern China, achieving a 20% improvement in wind protection while reducing simulated evapotranspiration by 15% [175]. Such studies illustrate how MOO can generate trade-off frontiers and guide context-specific design in a direction that should be expanded with regionally calibrated datasets.

6.6. Ensuring Long-Term Maintenance and Adaptive Management

Long-term performance of shelterbelts depends on proactive maintenance. Regular pruning and thinning maintain optimal porosity and prevent excessive shading or turbulence. Supplemental irrigation may be required during prolonged droughts, especially in early establishment phases, to prevent mortality [176,177,178]. Periodic species replacement or enrichment planting maintains structural diversity and resilience, compensating for losses due to pests, disease, or senescence. Integrating these practices into the optimization framework ensures sustained windbreak effectiveness and ecosystem service delivery across decades [179,180,181,182].

6.7. Practical Framework for Windbreak Design

To translate the review findings into actionable guidance, we propose a concise, practical framework summarizing key design principles for windbreak systems in arid regions (Table 6). This framework distills the evidence presented in previous sections on structural optimization (Section 3, Section 4 and Section 5) into a set of design objectives, recommended practices, and notes on their implementation. It is intended as a quick-reference tool for practitioners, policymakers, and extension specialists seeking to balance ecological performance, water conservation, and economic viability when planning shelterbelts [183,184]. By aligning windbreak construction with these best practices, land managers can systematically enhance wind protection, biodiversity support, and long-term sustainability under water-limited conditions.

7. Technical Recommendations for Windbreak Construction

Based on the synthesis of structural, ecological, and economic evidence presented in this review, several technical recommendations can be proposed to guide the effective construction and long-term performance of windbreak systems in arid agricultural landscapes. These recommendations are tailored to different functional objectives and can serve as a practical design checklist for practitioners and land managers.
(1)
Match windbreak purpose to design strategy.
Soil erosion control: Use relatively dense shrub-based shelterbelts (low to moderate porosity, 25–35%), 3–4 rows wide, to effectively reduce near-surface wind speed and trap moving soil particles.
Microclimate regulation and yield improvement: Prioritize taller trees with moderate porosity (30–50%) arranged in 3–5 rows, which create an extended leeward protection zone (15–30 H) without generating excessive turbulence.
Habitat and biodiversity enhancement: Adopt mixed-species, multi-strata configurations combining tall trees, mid-story shrubs, and ground cover vegetation to provide continuous canopy cover, nesting habitat, and ecological connectivity.
(2)
Optimize belt width and spacing.
For general field protection, use 3–5 rows with 2–4 m spacing between rows and 2–3 m within-row spacing.
For multifunctional or biodiversity corridor designs, expand to 6–8 rows to create an internal core habitat and buffer against edge effects.
Maintain a distance of 10–15 H between parallel belts across the landscape to ensure uniform protection without excessive land occupation.
(3)
Select species based on time horizon and water constraints.
Short-term protection: Fast-growing, drought-tolerant trees (e.g., Populus spp., Casuarina spp.) can quickly establish height and canopy cover but may require more maintenance.
Long-term stability: Incorporate native, deep-rooted, slow-growing species (e.g., Tamarix spp., Haloxylon spp.) that have lower water demands and higher survival under prolonged drought.
Mixed composition: Combine nitrogen-fixing shrubs and ground covers to improve soil fertility, reduce internal erosion, and maintain year-round functionality.
(4)
Incorporate adaptive management measures.
Implement scheduled pruning and thinning to maintain optimal porosity and prevent overcompetition with crops.
Plan for successional replanting to replace senescent individuals and maintain structural integrity.
Install efficient irrigation systems (e.g., subsurface drip) during establishment, and progressively reduce watering as deep root systems develop.
These recommendations provide a technical blueprint for constructing windbreaks that are functionally efficient, resource-conscious, and ecologically resilient in arid farming systems. Integrating these strategies during the planning and establishment stages can greatly improve the long-term performance and sustainability of shelterbelt networks.

8. Limitations and Risks of Shelterbelt Systems

While shelterbelts provide substantial ecological and agronomic benefits, their widespread implementation in arid agricultural systems is constrained by several important limitations and challenges that must be explicitly considered during planning and policy formulation.
(1)
High establishment and maintenance costs.
Windbreak systems require considerable upfront investment for site preparation, sapling purchase, planting labor, and early-stage irrigation infrastructure. Ongoing costs for pruning, thinning, pest and disease control, and replanting of failed individuals can also be substantial. These costs often deter adoption by smallholder farmers who operate under tight financial constraints.
(2)
Competition with crops for water and land resources.
Shelterbelt trees and shrubs consume soil water and nutrients and occupy arable land, potentially reducing the area available for crops and intensifying water scarcity in already water-limited environments. This trade-off is particularly critical during prolonged droughts, when competition between shelterbelt vegetation and adjacent crops can suppress crop yields near the belt edges.
(3)
Ecological and pest-related risks.
Dense, multi-layered shelterbelts may act as reservoirs for certain agricultural pests, diseases, or invasive species. Without integrated pest management, these risks can offset some of the ecological benefits and impose additional control costs.
(4)
Vulnerability to climatic extremes and natural disturbances.
Shelterbelts can be damaged or destroyed by severe droughts, high winds, sandstorms, or salinity stress—especially if composed of single-species stands with limited genetic diversity. Such disturbances can lead to structural collapse and sudden loss of protective function, requiring expensive restoration efforts.
(5)
Knowledge gaps and lack of region-specific design guidelines.
In many arid regions, there is a shortage of localized data on optimal shelterbelt structures, species combinations, and long-term hydrological impacts. This uncertainty makes it difficult to develop reliable design standards, leading to inconsistent performance and reduced farmer confidence in adoption.
Recognizing these constraints is crucial for developing realistic, context-appropriate planning frameworks and supportive policy measures. Addressing these challenges will require coordinated efforts in research, technology development, financial incentives, and extension support to ensure that windbreak systems can be effectively scaled and sustained in arid agroecosystems.

9. Conclusions

The structural optimization of farmland shelterbelts in arid regions lies at a pivotal interface of ecological engineering, microclimatic regulation, and sustainable agriculture. This review demonstrates that while shelterbelts consistently reduce wind erosion and enhance crop performance, reported effects vary widely. For example, optimal porosity values reported across studies range from 30% to 50%, and yield improvements from 10% to 25%, depending on height, species composition, and site climate. Such variability reflects the high context-dependence of shelterbelt performance and underscores the need for design frameworks that move beyond one-size-fits-all prescriptions.
Our synthesis highlights several key trade-offs that must guide design. Denser, low-porosity belts maximize wind protection but increase shading, root competition, and water consumption, while sparser designs conserve water yet offer weaker shelter. Species with dense crowns enhance protection but often have high water demand, whereas xerophytic shrubs are water-efficient yet less effective aerodynamically. These contrasts reveal why no single structural configuration is universally optimal, and why design should be guided by local climate, soil, and socio-economic conditions.
Contradictions in the literature also remain unresolved. Some field studies identify ~40% porosity as optimal, while others advocate closer to 50%, and reported protection distances range from 15 H to over 30 H. These discrepancies reflect differing measurement methods and site conditions, yet few meta-analyses have systematically compared them. Likewise, while multi-strata, mixed-species belts are widely promoted, quantitative evaluations of their long-term performance, maintenance costs, and biodiversity benefits remain sparse.
Addressing these inconsistencies will require a shift from descriptive studies to predictive, trade-off-aware optimization. Advances in field monitoring, wind tunnel testing, computational fluid dynamics, and remote sensing now make it feasible to couple empirical data with multi-objective optimization algorithms (e.g., NSGA-II, Pareto frontiers). However, these approaches remain underused and rarely calibrated with region-specific datasets. Future research should (i) quantify structural–functional relationships across climatic gradients, (ii) evaluate long-term performance and water use of native xerophyte-based designs, and (iii) integrate economic viability into optimization models to guide adoption by smallholders.
Ultimately, the successful implementation of optimized shelterbelts depends not only on technical innovations such as drought-resistant species, AI-driven design tools, and climate-adaptive configurations but also on enabling policies, farmer training, and financial incentives that align ecological and livelihood goals. By embracing a science-based, adaptive, and participatory approach, shelterbelts can be transformed from isolated windbreak structures into multifunctional landscape elements that enhance agricultural productivity, build ecosystem resilience, and contribute to sustainable development in the world’s arid regions.

Author Contributions

Conceptualization, A.A.; methodology, A.A., H.X., and A.W.; formal analysis, F.B.; writing—original draft preparation, A.A.; supervision, H.X., and A.W.; All authors have read and agreed to the published version of the manuscript.

Funding

This study was jointly supported by the Natural Science Foundation of Xinjiang Uyghur Autonomous Region (2022D01A353) and the Entrusted project of the Land Comprehensive Improvement Center of Xinjiang (E2400109).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data may be provided on reasonable request from the first author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Reimer, A.; Thompson, A.; Prokopy, L.S.; Arbuckle, J.G.; Genskow, K.; Jackson-Smith, D.; Lynne, G.; McCann, L.; Morton, L.W.; Nowak, P. People, place, behavior, and context: A research agenda for expanding our understanding of what motivates farmers’ conservation behaviors. J. Soil Water Conserv. 2014, 69, 57A–61A. [Google Scholar] [CrossRef]
  2. Gaur, M.K.; Squires, V.R. Geographic extent and characteristics of the world’s arid zones and their peoples. In Climate Variability Impacts on Land Use and Livelihoods in Drylands; Springer: Berlin/Heidelberg, Germany, 2017; pp. 3–20. [Google Scholar]
  3. Rhodes, C.J. Soil erosion, climate change and global food security: Challenges and strategies. Sci. Prog. 2014, 97, 97–153. [Google Scholar] [CrossRef]
  4. Hagen, L.J. Evaluation of the Wind Erosion Prediction System (WEPS) erosion submodel on cropland fields. Environ. Model. Softw. 2004, 19, 171–176. [Google Scholar] [CrossRef]
  5. Saleem, A.; Anwar, S.; Nawaz, T.; Fahad, S.; Saud, S.; Ur Rahman, T.; Khan, M.N.R.; Nawaz, T. Securing a sustainable future: The climate change threat to agriculture, food security, and sustainable development goals. J. Umm Al-Qura Univ. Appl. Sci. 2024, 11, 595–611. [Google Scholar] [CrossRef]
  6. Warner, T.; Gad-el-Hak, M. Weather-Related Disasters in Arid Lands; Cambridge University Press: Cambridge, UK, 2008; pp. 377–426. [Google Scholar]
  7. Allen, R.G.; Pereira, L.S.; Raes, D.; Smith, M. Crop Evapotranspiration-Guidelines for Computing Crop Water Requirements-FAO Irrigation and Drainage Paper 56; FAO: Rome, Italy, 1998; Volume 300, p. D05109. [Google Scholar]
  8. Ahmadzai, H.; Tutundjian, S.; Elouafi, I. Policies for sustainable agriculture and livelihood in marginal lands: A review. Sustainability 2021, 13, 8692. [Google Scholar] [CrossRef]
  9. Rashid, M.; Gani, G. Reimagining the Future of Sustainable Agriculture in South Asia: Integrating Ecological Resilience, Technological Innovation, and Inclusive Policy Reform for Transformative Agri-Systems. In Precision Agriculture and Climate-Resilient Farming: Artificial Intelligence, IoT, and Blockchain for Sustainable Agriculture; Deep Science Publishing: San Francisco, CA, USA, 2025; Volume 118. [Google Scholar]
  10. Brandle, J.R.; Hodges, L.; Zhou, X.H. Windbreaks in North American agricultural systems. Agrofor. Syst. 2004, 61, 65–78. [Google Scholar] [CrossRef]
  11. Nair, P.R.; Kumar, B.M.; Nair, V.D. Agroforestry systems in the temperate zone. In An Introduction to Agroforestry: Four Decades of Scientific Developments; Springer: Berlin/Heidelberg, Germany, 2022; pp. 195–232. [Google Scholar]
  12. Cleugh, H. Effects of windbreaks on airflow, microclimates and crop yields. Agrofor. Syst. 1998, 41, 55–84. [Google Scholar] [CrossRef]
  13. Enescu, C.M.; Mircea, M.; Ilie, L.; Lucian, D.; Cristinel, C.; Gabriel, M. Agricultural benefits of shelterbelts and windbreaks: A bibliometric analysis. Agriculture 2025, 15, 1204. [Google Scholar] [CrossRef]
  14. Zobeck, T.M.; Van Pelt, R.S. Wind erosion. In Soil Management: Building a Stable Base for Agriculture; American Society of Agronomy and Soil Science Society of America: Madison, WI, USA, 2011; pp. 209–227. [Google Scholar]
  15. Nuberg, I.; Bennell, M.; George, B. Trees protecting dryland crops and soil. In Agroforestry for Natural Resource Management; CSIRO Publishing: Collingwood, Australia, 2009; pp. 69–85. [Google Scholar]
  16. Cai, X.; Henderson, M.; Wang, L.; Su, Y.; Liu, B. Shelterbelt structure and crop protection from increased typhoon activity in Northeast China. Agriculture 2021, 11, 995. [Google Scholar] [CrossRef]
  17. Kort, J. 9. Benefits of windbreaks to field and forage crops. Agric. Ecosyst. Environ. 1988, 22, 165–190. [Google Scholar] [CrossRef]
  18. Giller, K.E. Nitrogen Fixation in Tropical Cropping Systems; Cabi: Wallingford, UK, 2001. [Google Scholar]
  19. Wang, J.; Patruno, L.; Chen, Z.; Yang, Q.; Tamura, Y. Windbreak effectiveness of single and double-arranged shelterbelts: A parametric study using large eddy simulation. Forests 2024, 15, 1760. [Google Scholar] [CrossRef]
  20. Zhou, X.; Brandle, J.R.; Mize, C.; Takle, E. The relationship of three-dimensional structure to shelterbelt function: A theoretical hypothesis. In Ecological Basis of Agroforestry; CRC Press: Boca Raton, FL, USA, 2004. [Google Scholar]
  21. Kovalenko, I.; Kyrylchuk, K.; Klymenko, H.; Yaroshchuk, S.; Yaroshchuk, R.; Kovalenko, N.; Skyba, O. Influence of tree-crown density on dominant plant species of the herb-shrub stratum in the zone of mixed forests. Biosyst. Divers. 2023, 31, 382–387. [Google Scholar] [CrossRef]
  22. Szigeti, N.; Frank, N.; Vityi, A. The multifunctional role of shelterbelts in intensively managed agricultural land–Silvoarable agroforestry in Hungary. Acta Silv. Lignaria Hung. Int. J. For. Wood Environ. Sci. 2020, 16, 19–38. [Google Scholar] [CrossRef]
  23. Rempel, J. Costs, Benefits, and Barriers to the Adoption and Retention of Shelterbelts in Prairie Agriculture as Identified by Saskatchewan Producers. Ph.D. Thesis, University of Saskatchewan, Saskatoon, SK, Canada, 2014. [Google Scholar]
  24. Chandewar, N. Agroforestry and Sustainable Land Management; Addition Publishing House: Bhopal, India, 2025. [Google Scholar]
  25. Lin, C.-L.; Tawhai, M.H.; Mclennan, G.; Hoffman, E.A. Computational fluid dynamics. IEEE Eng. Med. Biol. Mag. 2009, 28, 25–33. [Google Scholar] [CrossRef] [PubMed]
  26. Ding, F.; Kareem, A.; Wan, J. Aerodynamic tailoring of structures using computational fluid dynamics. Struct. Eng. Int. 2019, 29, 26–39. [Google Scholar] [CrossRef]
  27. Tu, J.; Yeoh, G.H.; Liu, C.; Tao, Y. Computational Fluid Dynamics: A Practical Approach; Elsevier: Amsterdam, The Netherlands, 2023. [Google Scholar]
  28. Mize, C.; Brandle, J.R.; Schoeneberger, M.; Bentrup, G. Ecological development and function of shelterbelts in temperate North America. In Toward Agroforestry Design: An Ecological Approach; Springer: Berlin/Heidelberg, Germany, 2008; pp. 27–54. [Google Scholar]
  29. Wang, H.; Takle, E.S.; Shen, J. Shelterbelts and windbreaks: Mathematical modeling and computer simulations of turbulent flows. Annu. Rev. Fluid Mech. 2001, 33, 549–586. [Google Scholar] [CrossRef]
  30. Aili, A.; Xu, H.; Waheed, A.; Bakayisire, F.; Xie, Y. Synergistic windbreak efficiency of desert vegetation and oasis shelter forests. PLoS ONE 2024, 19, e0312876. [Google Scholar] [CrossRef] [PubMed]
  31. Brandle, J.R.; Takle, E.; Zhou, X. Windbreak practices. North American agroforestry. In Integrated Science and Practice; American Society of Agronomy: Madison, WI, USA, 2021; pp. 89–126. [Google Scholar]
  32. Sun, Q.; Zheng, B.; Liu, T.; Zhu, L.; Hao, X.; Han, Z. The optimal spacing interval between principal shelterbelts of the farm-shelter forest network. Environ. Sci. Pollut. Res. 2022, 29, 12680–12693. [Google Scholar] [CrossRef]
  33. Ma, R.; Wang, J.; Qu, J.; Liu, H. Effectiveness of shelterbelt with a non-uniform density distribution. J. Wind. Eng. Ind. Aerodyn. 2010, 98, 767–771. [Google Scholar] [CrossRef]
  34. Terziev, A.; Bode, F.; Zlateva, P.; Pichurov, G.; Ivanov, M.; Denev, J.; Stankov, B. Numerical Study on Tree Belt Impact on Wind Shear on Agricultural Land. Appl. Sci. 2025, 15, 7450. [Google Scholar] [CrossRef]
  35. Heisler, G.M.; Dewalle, D.R. 2. Effects of windbreak structure on wind flow. Agric. Ecosyst. Environ. 1988, 22, 41–69. [Google Scholar] [CrossRef]
  36. Eichmanns, C.; Schüttrumpf, H. A Nature-Based Solution for Coastal Protection: Wind Tunnel Investigations on the Influence of Sand-Trapping Fences on Sediment Accretion. Front. Built Environ. 2022, 8, 878197. [Google Scholar] [CrossRef]
  37. Potashkina, Y.N.; Koshelev, A.V. Impact of field-protective forest belts on the microclimate of agroforest landscape in the zone of chestnut soils of the Volgograd region. Forests 2022, 13, 1892. [Google Scholar] [CrossRef]
  38. Morgan, R. Wind erosion control. In Slope Stabilization and Erosion Control: A Bioengineering Approach; Taylor & Francis: Abingdon, UK, 2003; pp. 203–232. [Google Scholar]
  39. Plate, E.J. The aerodynamics of shelter belts. Agric. Meteorol. 1971, 8, 203–222. [Google Scholar] [CrossRef]
  40. Giebel, G.; Paulsen, U.S.; Bange, J.; la Cour-Harbo, A.; Reuder, J.; Mayer, S.; van der Kroonenberg, A.; Mølgaard, J. Autonomous Aerial Sensors for Wind Power Meteorology-a Pre-Project; Danmarks Tekniske Universitet, Risø Nationallaboratoriet for Bæredygtig Energi: Roskilde, Denmark, 2012. [Google Scholar]
  41. Zou, X.; Li, H.; Kang, L.; Zhang, C.; Jia, W.; Gao, Y.; Zhang, J.; Yang, Z.; Zhang, M.; Xu, J. Soil wind erosion rate on rough surfaces: A dynamical model derived from an invariant pattern of the shear-stress probability density function of the soil surface. Catena 2022, 219, 106633. [Google Scholar] [CrossRef]
  42. Field, J.P.; Breshears, D.D.; Whicker, J.J. Toward a more holistic perspective of soil erosion: Why aeolian research needs to explicitly consider fluvial processes and interactions. Aeolian Res. 2009, 1, 9–17. [Google Scholar] [CrossRef]
  43. Hong, C.; Liu, C.; Zou, X.; Li, H.; Kang, L.; Liu, B.; Li, J. Wind erosion rate for vegetated soil cover: A prediction model based on surface shear strength. Catena 2020, 187, 104398. [Google Scholar] [CrossRef]
  44. Wegner, G.I. Payments for ecosystem services (PES): A flexible, participatory, and integrated approach for improved conservation and equity outcomes. Environ. Dev. Sustain. 2016, 18, 617–644. [Google Scholar] [CrossRef]
  45. Tanny, J. Microclimate and evapotranspiration of crops covered by agricultural screens: A review. Biosyst. Eng. 2013, 114, 26–43. [Google Scholar] [CrossRef]
  46. Yuan, W.; Zhu, N.; Zhang, L.; Tong, R.; Miao, Y.; Zhou, F.; Wang, G.G.; Wu, T. Three-dimensional aerodynamic structure estimation and wind field simulation for wide tree shelterbelts. For. Ecol. Manag. 2024, 559, 121813. [Google Scholar] [CrossRef]
  47. He, C.; Shao, W. Numerical simulation of shelter effect assessment for single-row windbreaks on the periphery of oasis farmland. J. Arid. Environ. 2024, 222, 105165. [Google Scholar] [CrossRef]
  48. Baumgertel, A.; Lukić, S.; Caković, M.; Savić, R.; Bezdan, A.; Blagojević, B. Wind erosion, climate change, and shelterbelts. In Prevention and Management of Soil Erosion and Torrential Floods; IGI Global Scientific Publishing: Palmdale, PA, USA, 2022; pp. 154–175. [Google Scholar]
  49. Singh, R. Wind erosion. In Soil and Water Conservation Structures Design; Springer: Berlin/Heidelberg, Germany, 2023; pp. 297–322. [Google Scholar]
  50. Naorem, A.; Jayaraman, S.; Dang, Y.P.; Dalal, R.C.; Sinha, N.K.; Rao, C.S.; Patra, A.K. Soil constraints in an arid environment—Challenges, prospects, and implications. Agronomy 2023, 13, 220. [Google Scholar] [CrossRef]
  51. Barrocu, G.; Eslamian, S. Geomorphology and flooding. In Flood handbook; CRC Press: Boca Raton, FL, USA, 2022; pp. 23–54. [Google Scholar]
  52. Abbasi, H. Sand Dune Systems in Iran-Distribution and Activity; Philipps-Universität Marburg: Marburg, Germany, 2020. [Google Scholar]
  53. Lee, K.-S.; Seo, I.-H.; Yang, J.-E.; Lee, S.-P.; Jung, H.-G.; Chung, D.Y. Simple assessment of wind erosion depending on the soil texture and threshold wind velocity in reclaimed tidal flat land. Korean J. Agric. Sci. 2021, 48, 843–853. [Google Scholar] [CrossRef]
  54. Leguédois, S.; Ellis, T.W.; Hairsine, P.B.; Tongway, D.J. Sediment trapping by a tree belt: Processes and consequences for sediment delivery. Hydrol. Process. Int. J. 2008, 22, 3523–3534. [Google Scholar] [CrossRef]
  55. Wilkes, M.A.; Gittins, J.R.; Mathers, K.L.; Mason, R.; Casas-Mulet, R.; Vanzo, D.; Mckenzie, M.; Murray-Bligh, J.; England, J.; Gurnell, A. Physical and biological controls on fine sediment transport and storage in rivers. Wiley Interdiscip. Rev. Water 2019, 6, e1331. [Google Scholar] [CrossRef]
  56. Dunkerley, D. A review of the effects of throughfall and stemflow on soil properties and soil erosion. In Precipitation Partitioning by Vegetation: A Global Synthesis; Springer: Berlin/Heidelberg, Germany, 2020; pp. 183–214. [Google Scholar]
  57. Walker, I.J.; Nickling, W.G. Dynamics of secondary airflow and sediment transport over and in the lee of transverse dunes. Prog. Phys. Geogr. 2002, 26, 47–75. [Google Scholar] [CrossRef]
  58. Katul, G.G.; Oren, R.; Manzoni, S.; Higgins, C.; Parlange, M.B. Evapotranspiration: A process driving mass transport and energy exchange in the soil-plant-atmosphere-climate system. Rev. Geophys. 2012, 50. [Google Scholar] [CrossRef]
  59. Wang, F. Modelling sheltering effects of trees on reducing space heating in office buildings in a windy city. Energy Build. 2006, 38, 1443–1454. [Google Scholar] [CrossRef]
  60. Sharma, P.K.; Kumar, S. Soil temperature and plant growth. In Soil Physical Environment and Plant Growth: Evaluation and Management; Springer: Berlin/Heidelberg, Germany, 2023; pp. 175–204. [Google Scholar]
  61. Davarzani, H.; Smits, K.; Tolene, R.M.; Illangasekare, T. Study of the effect of wind speed on evaporation from soil through integrated modeling of the atmospheric boundary layer and shallow subsurface. Water Resour. Res. 2014, 50, 661–680. [Google Scholar] [CrossRef] [PubMed]
  62. Easterling, W.E.; Hays, C.J.; Easterling, M.M.; Brandle, J.R. Modelling the effect of shelterbelts on maize productivity under climate change: An application of the EPIC model. Agric. Ecosyst. Environ. 1997, 61, 163–176. [Google Scholar] [CrossRef]
  63. Medrano, H.; Tomás, M.; Martorell, S.; Escalona, J.-M.; Pou, A.; Fuentes, S.; Flexas, J.; Bota, J. Improving water use efficiency of vineyards in semi-arid regions. A review. Agron. Sustain. Dev. 2015, 35, 499–517. [Google Scholar] [CrossRef]
  64. Hatfield, J.L.; Dold, C. Water-use efficiency: Advances and challenges in a changing climate. Front. Plant Sci. 2019, 10, 103. [Google Scholar] [CrossRef]
  65. Kilemo, D.B. The review of water use efficiency and water productivity metrics and their role in sustainable water resources management. Open Access Libr. J. 2022, 9, 1–21. [Google Scholar] [CrossRef]
  66. Ogwu, M.C.; Kosoe, E.A. Integrating Green Infrastructure into Sustainable Agriculture to Enhance Soil Health, Biodiversity, and Microclimate Resilience. Sustainability 2025, 17, 3838. [Google Scholar] [CrossRef]
  67. Biswas, P.; Mondal, S.; Maji, S.; Mondal, A.; Bandopadhyay, P. Microclimate modification in field crops: A way toward climate-resilience. In Climate-Resilient Agriculture, Vol 1: Crop Responses and Agroecological Perspectives; Springer: Berlin/Heidelberg, Germany, 2023; pp. 647–666. [Google Scholar]
  68. Kingra, P.K.; Kaur, H. Microclimatic modifications to manage extreme weather vulnerability and climatic risks in crop production. J. Agric. Phys. 2017, 17, 1–15. [Google Scholar]
  69. Armbrust, D.; Retta, A. Wind and sandblast damage to growing vegetation. Ann. Arid. Zone 2000, 39, 273–284. [Google Scholar]
  70. Gao, Q.; Zhao, P.; Zeng, X.; Cai, X.; Shen, W. A model of stomatal conductance to quantify the relationship between leaf transpiration, microclimate and soil water stress. Plant Cell Environ. 2002, 25, 1373–1381. [Google Scholar] [CrossRef]
  71. Lawson, T.; Vialet-Chabrand, S. Speedy stomata, photosynthesis and plant water use efficiency. New Phytol. 2019, 221, 93–98. [Google Scholar] [CrossRef] [PubMed]
  72. Lawson, T.; von Caemmerer, S.; Baroli, I. Photosynthesis and stomatal behaviour. In Progress in Botany 72; Springer: Berlin/Heidelberg, Germany, 2010; pp. 265–304. [Google Scholar]
  73. Sudmeyer, R.; Crawford, M.; Meinke, H.; Poulton, P.; Robertson, M. Effect of artificial wind shelters on the growth and yield of rainfed crops. Aust. J. Exp. Agric. 2002, 42, 841–858. [Google Scholar] [CrossRef]
  74. Blum, A. Plant water relations, plant stress and plant production. In Plant Breeding for Water-Limited Environments; Springer: Berlin/Heidelberg, Germany, 2010; pp. 11–52. [Google Scholar]
  75. Fageria, N. Role of soil organic matter in maintaining sustainability of cropping systems. Commun. Soil Sci. Plant Anal. 2012, 43, 2063–2113. [Google Scholar] [CrossRef]
  76. Zhu, X.; Liu, W.; Chen, J.; Bruijnzeel, L.A.; Mao, Z.; Yang, X.; Cardinael, R.; Meng, F.-R.; Sidle, R.C.; Seitz, S. Reductions in water, soil and nutrient losses and pesticide pollution in agroforestry practices: A review of evidence and processes. Plant Soil 2020, 453, 45–86. [Google Scholar] [CrossRef]
  77. Sileshi, G.W.; Dagar, J.C.; Nath, A.J.; Kuntashula, E. Agroforestry as a climate-smart agriculture: Strategic interventions, current practices and policies. In Agroforestry for Sustainable Intensification of Agriculture in Asia and Africa; Springer: Berlin/Heidelberg, Germany, 2023; pp. 589–640. [Google Scholar]
  78. Bashir, D. Thatched Mat Windbreaks Influences on the Harmattan: Wind and Millet Production in the Sahel; The University of Wisconsin-Madison: Madison, WI, USA, 1990. [Google Scholar]
  79. Baldwin, C.S. 10. The influence of field windbreaks on vegetable and specialty crops. Agric. Ecosyst. Environ. 1988, 22–23, 191–203. [Google Scholar] [CrossRef]
  80. Mayrinck, R.C.; Laroque, C.P.; Amichev, B.Y.; Van Rees, K. Above-and below-ground carbon sequestration in shelterbelt trees in Canada: A review. Forests 2019, 10, 922. [Google Scholar] [CrossRef]
  81. Fonseka, D.; Jha, N.; Jeyakumar, P. Soil nutrient enrichment in pastoral systems through shelterbelts. J. Environ. Manag. 2025, 393, 126938. [Google Scholar] [CrossRef]
  82. Smith, P.; Martino, D.; Cai, Z.; Gwary, D.; Janzen, H.; Kumar, P.; McCarl, B.; Ogle, S.; O’Mara, F.; Rice, C. Greenhouse gas mitigation in agriculture. Philos. Trans. R. Soc. B Biol. Sci. 2008, 363, 789–813. [Google Scholar] [CrossRef]
  83. Lavrov, V.; Miroshnyk, N.; Grabovska, T.; Shupova, T. Forest shelter belts in organic agricultural landscape: Structure of biodiversity and their ecological role. Folia For. Pol. Ser. A For. 2021, 63, 48–64. [Google Scholar] [CrossRef]
  84. Ultsch, G.R. The ecology of overwintering among turtles: Where turtles overwinter and its consequences. Biol. Rev. 2006, 81, 339–367. [Google Scholar] [CrossRef]
  85. McCravy, K.W. A review of sampling and monitoring methods for beneficial arthropods in agroecosystems. Insects 2018, 9, 170. [Google Scholar] [CrossRef] [PubMed]
  86. Nicholls, C.I.; Altieri, M.A. Plant biodiversity enhances bees and other insect pollinators in agroecosystems. A review. Agron. Sustain. Dev. 2013, 33, 257–274. [Google Scholar] [CrossRef]
  87. Aparin, B.; Sukhacheva, E.Y.; Zakharova, M.; Mingareeva, E.; Koshelev, A. Effect of Massive Forest Shelterbelts on Humus Content and Stock in Chernozems. Eurasian Soil Sci. 2024, 57, 2068–2080. [Google Scholar] [CrossRef]
  88. Hilty, J.A.; Keeley, A.T.; Lidicker, W.Z., Jr.; Merenlender, A.M. Corridor Ecology: Linking Landscapes for Biodiversity Conservation and Climate Adaptation; Island Press: Washington, DC, USA, 2019. [Google Scholar]
  89. Babu, A. Review of the role of the landscape approach in biodiversity conservation. Sustain. Biodivers. Conserv. 2023, 2, 61–86. [Google Scholar]
  90. Lawrence, M.; Jiang, Y. Porosity, pore size distribution, micro-structure. In Bio-Aggregates Based Building Materials: State-of-the-Art Report of the RILEM Technical Committee 236-BBM; Springer: Berlin/Heidelberg, Germany, 2017; pp. 39–71. [Google Scholar]
  91. Zhou, X.; Brandle, J.; Mize, C.; Takle, E. Three-dimensional aerodynamic structure of a tree shelterbelt: Definition, characterization and working models. Agrofor. Syst. 2005, 63, 133–147. [Google Scholar] [CrossRef]
  92. Li, H.; Wang, Y.; Li, S.; Askar, A.; Wang, H. Shelter efficiency of various shelterbelt configurations: A wind tunnel study. Atmosphere 2022, 13, 1022. [Google Scholar] [CrossRef]
  93. Wu, T.; Zhang, P.; Zhang, L.; Wang, J.; Yu, M.; Zhou, X.; Wang, G.G. Relationships between shelter effects and optical porosity: A meta-analysis for tree windbreaks. Agric. For. Meteorol. 2018, 259, 75–81. [Google Scholar] [CrossRef]
  94. Tyndall, J.; Colletti, J. Mitigating swine odor with strategically designed shelterbelt systems: A review. Agrofor. Syst. 2007, 69, 45–65. [Google Scholar] [CrossRef]
  95. Zhang, J.; Jia, Z.; Li, Q.; He, L.; Zhao, X.; Wang, L.; Han, D. Determine the Optimal Vegetation Type for Soil Wind Erosion Prevention and Control in the Alpine Sandy Land of the Gonghe Basin on the Qinghai Tibet Plateau. Forests 2023, 14, 2342. [Google Scholar] [CrossRef]
  96. Yu, Y.-P.; Zhang, K.-C.; An, Z.-S.; Wang, T.; Hu, F. The blocking effect of the sand fences quantified using wind tunnel simulations. J. Mt. Sci. 2020, 17, 2485–2496. [Google Scholar] [CrossRef]
  97. Antoine, M.; Alain, C.; David, R. Drivers of tree establishment in planted windbreaks and riparian buffers: A case study of farms in southern Quebec, Canada. Geoderma Regional. 2024, 37, e00788. [Google Scholar] [CrossRef]
  98. Dong, Z.; Luo, W.; Qian, G.; Wang, H. A wind tunnel simulation of the mean velocity fields behind upright porous fences. Agric. For. Meteorol. 2007, 146, 82–93. [Google Scholar] [CrossRef]
  99. McHenry, M. How farming and forestry converge: Enhancing the interface between agricultural production, and tree biomass/bioenergy systems to improve farm-scale productivity in Western Australia. In Bioenergy Systems, Biological Sources and Environmental Impact; Nova Science Publishers: Hauppauge, NY, USA, 2013; pp. 95–110. [Google Scholar]
  100. Lampartová, I.; Schneider, J.; Vyskot, I.; Rajnoch, M.; Litschmann, T. Impact of protective shelterbelt on microclimate characteristics. Ekológia 2015, 34, 101. [Google Scholar] [CrossRef]
  101. Packwood, A. Flow through porous fences in thick boundary layers: Comparisons between laboratory and numerical experiments. J. Wind. Eng. Ind. Aerodyn. 2000, 88, 75–90. [Google Scholar] [CrossRef]
  102. Suratman, M.N.; Brandle, J.R. Tree shelterbelts for sustainable agroforestry. In Agroforestry for Carbon and Ecosystem Management; Academic Press: Cambridge, MA, USA, 2024; pp. 97–107. [Google Scholar]
  103. Pandey, V.C.; Singh, K.; Singh, J.S.; Kumar, A.; Singh, B.; Singh, R.P. Jatropha curcas: A potential biofuel plant for sustainable environmental development. Renew. Sustain. Energy Rev. 2012, 16, 2870–2883. [Google Scholar] [CrossRef]
  104. Wilkinson, K.M.; Elevitch, C.R. Multipurpose windbreaks: Design and species for Pacific Islands. In Agroforestry Guides for Pacific Islands; Permanent Agriculture Resources: Holualoa, HI, USA, 2000; Volume 8. [Google Scholar]
  105. Rosell, J.; Sanz, R. A review of methods and applications of the geometric characterization of tree crops in agricultural activities. Comput. Electron. Agric. 2012, 81, 124–141. [Google Scholar] [CrossRef]
  106. Sundeep, S. Understanding and Reduction of the Aerodynamic Noise Emitted from Surface-Mounted Obstacles and Airfoil Trailing Edge. Ph.D. Thesis, Hong Kong University of Science and Technology (Hong Kong), Hong Kong, 2023. [Google Scholar]
  107. Whelan, C.J.; Wenny, D.G.; Marquis, R.J. Ecosystem services provided by birds. Ann. N. Y. Acad. Sci. 2008, 1134, 25–60. [Google Scholar] [CrossRef] [PubMed]
  108. Benítez-Malvido, J.; Arroyo-Rodríguez, V. Habitat Fragmentation, Edge Effects and Biological Corridors in Tropical Ecosystems; Eolss Publishers: Oxford, UK, 2008. [Google Scholar]
  109. Firoozi, A.A.; Hejazi, F.; Firoozi, A.A. Advancing wind energy efficiency: A systematic review of aerodynamic optimization in wind turbine blade design. Energies 2024, 17, 2919. [Google Scholar] [CrossRef]
  110. Yang, D.; Liu, W.; Wang, J.; Liu, B.; Fang, Y.; Li, H.; Zou, X. Wind erosion forces and wind direction distribution for assessing the efficiency of shelterbelts in northern China. Aeolian Res. 2018, 33, 44–52. [Google Scholar] [CrossRef]
  111. Jackson, D.W.; Beyers, M.; Delgado-Fernandez, I.; Baas, A.C.; Cooper, A.J.; Lynch, K. Airflow reversal and alternating corkscrew vortices in foredune wake zones during perpendicular and oblique offshore winds. Geomorphology 2013, 187, 86–93. [Google Scholar] [CrossRef]
  112. Dong, Z.; Wang, X.; Liu, L. Wind erosion in arid and semiarid China: An overview. J. Soil Water Conserv. 2000, 55, 439–444. [Google Scholar] [CrossRef]
  113. García Chevesich, P.; Neary, D.G.; Scott, D.F.; Benyon, R.G.; Reyna, T. Forest Management and the Impact on Water Resources: A Review of 13 Countries; UNESCO: Paris, France, 2017. [Google Scholar]
  114. Stanley, A.P.; King, J. Optimizing the physical design and layout of a resilient wind, solar, and storage hybrid power plant. Appl. Energy 2022, 317, 119139. [Google Scholar] [CrossRef]
  115. Chen, C.; Wang, J.; Ton, D. Modernizing distribution system restoration to achieve grid resiliency against extreme weather events: An integrated solution. Proc. IEEE 2017, 105, 1267–1288. [Google Scholar] [CrossRef]
  116. Hashmi, S.A. The Aerodynamics of a High-Speed Train Running Adjacent to Windbreak Walls. Ph.D. Thesis, University of Birmingham, Birmingham, UK, 2021. [Google Scholar]
  117. Gonzales, H.B. Aerodynamics of Wind Erosion and Particle Collection Through Vegetative Controls; Kansas State University: Manhattan, KS, USA, 2015. [Google Scholar]
  118. Scholes, R.J. The future of semi-arid regions: A weak fabric unravels. Climate 2020, 8, 43. [Google Scholar] [CrossRef]
  119. Ryan, J.G.; McAlpine, C.A.; Ludwig, J.A. Integrated vegetation designs for enhancing water retention and recycling in agroecosystems. Landsc. Ecol. 2010, 25, 1277–1288. [Google Scholar] [CrossRef]
  120. Xie, H.; Wang, G.G.; Yu, M. Ecosystem multifunctionality is highly related to the shelterbelt structure and plant species diversity in mixed shelterbelts of eastern China. Glob. Ecol. Conserv. 2018, 16, e00470. [Google Scholar] [CrossRef]
  121. Van Loggerenberg, N. Machina Ex Silva: Where the Forest Meets the Blade; University of the Witwatersrand: Johannesburg, South Africa, 2015. [Google Scholar]
  122. Hartmann, M.; Schwieger, W. Hierarchically-structured porous materials: From basic understanding to applications. Chem. Soc. Rev. 2016, 45, 3311–3312. [Google Scholar] [CrossRef]
  123. Zuazo, V.C.H.D.; Pleguezuelo, C.R.O.R. Soil-erosion and runoff prevention by plant covers: A review. In Sustainable Agriculture; Springer: Dordrecht, The Netherlands, 2009; pp. 785–811. [Google Scholar]
  124. Duniway, M.C.; Pfennigwerth, A.A.; Fick, S.E.; Nauman, T.W.; Belnap, J.; Barger, N.N. Wind erosion and dust from US drylands: A review of causes, consequences, and solutions in a changing world. Ecosphere 2019, 10, e02650. [Google Scholar] [CrossRef]
  125. O’Sullivan, O.S.; Holt, A.R.; Warren, P.H.; Evans, K.L. Optimising UK urban road verge contributions to biodiversity and ecosystem services with cost-effective management. J. Environ. Manag. 2017, 191, 162–171. [Google Scholar] [CrossRef]
  126. Muluneh, M.G. Impact of climate change on biodiversity and food security: A global perspective—A review article. Agric. Food Secur. 2021, 10, 36. [Google Scholar] [CrossRef]
  127. Li, Y.; Pan, F.; Yao, H. Response of symbiotic and asymbiotic nitrogen-fixing microorganisms to nitrogen fertilizer application. J. Soils Sediments 2019, 19, 1948–1958. [Google Scholar] [CrossRef]
  128. Aasfar, A.; Bargaz, A.; Yaakoubi, K.; Hilali, A.; Bennis, I.; Zeroual, Y.; Meftah Kadmiri, I. Nitrogen fixing Azotobacter species as potential soil biological enhancers for crop nutrition and yield stability. Front. Microbiol. 2021, 12, 628379. [Google Scholar] [CrossRef]
  129. Salih, K.; Báthoryné Nagy, I.R. Review of the role of urban green infrastructure on climate resiliency: A focus on heat mitigation modelling scenario on the microclimate and building scale. Urban Sci. 2024, 8, 220. [Google Scholar] [CrossRef]
  130. Chauhan, S.; Kengoo, N.; Kishore, K.; Haksinhbhai, M.R.; Rana, P. Carbon Dynamics in Agroforestry Systems: Implications for Climate Change Mitigation and Adaptation. Int. J. Environ. Clim. Change 2025, 15, 109–133. [Google Scholar] [CrossRef]
  131. Ehbrecht, M.; Schall, P.; Ammer, C.; Seidel, D. Quantifying stand structural complexity and its relationship with forest management, tree species diversity and microclimate. Agric. For. Meteorol. 2017, 242, 1–9. [Google Scholar] [CrossRef]
  132. An, L.; Wang, J.; Xiong, N.; Wang, Y.; You, J.; Li, H. Assessment of permeability windbreak forests with different porosities based on laser scanning and computational fluid dynamics. Remote Sens. 2022, 14, 3331. [Google Scholar] [CrossRef]
  133. Misra, B.B.; Langefeld, C.; Olivier, M.; Cox, L.A. Integrated omics: Tools, advances and future approaches. J. Mol. Endocrinol. 2019, 62, R21–R45. [Google Scholar] [CrossRef]
  134. Cleugh, H.; Hughes, D. Impact of shelter on crop microclimates: A synthesis of results from wind tunnel and field experiments. Aust. J. Exp. Agric. 2002, 42, 679–701. [Google Scholar] [CrossRef]
  135. Papesch, A. Wind tunnel test to optimize barrier spacing and porosity to reduce wind damage in horticultural shelter systems. J. Wind. Eng. Ind. Aerodyn. 1992, 44, 2631–2642. [Google Scholar] [CrossRef]
  136. Englund, G.; Cooper, S.D. Scale effects and extrapolation in ecological experiments. Adv. Ecol. Res. 2003, 11, 161–213. [Google Scholar]
  137. Verrelst, J.; Malenovský, Z.; Van der Tol, C.; Camps-Valls, G.; Gastellu-Etchegorry, J.-P.; Lewis, P.; North, P.; Moreno, J. Quantifying vegetation biophysical variables from imaging spectroscopy data: A review on retrieval methods. Surv. Geophys. 2019, 40, 589–629. [Google Scholar] [CrossRef]
  138. Wang, J.; Patruno, L.; Zhao, G.; Tamura, Y. Windbreak effectiveness of shelterbelts with different characteristic parameters and arrangements by means of CFD simulation. Agric. For. Meteorol. 2024, 344, 109813. [Google Scholar] [CrossRef]
  139. Bush, R.H.; Chyczewski, T.S.; Duraisamy, K.; Eisfeld, B.; Rumsey, C.L.; Smith, B.R. Recommendations for future efforts in RANS modeling and simulation. In Proceedings of the AIAA Scitech 2019 Forum, San Diego, CA, USA, 7–11 January 2019; p. 0317. [Google Scholar]
  140. Alfonsi, G. Reynolds-averaged Navier–Stokes equations for turbulence modeling. Appl. Mech. Rev. 2009, 62, 040802. [Google Scholar] [CrossRef]
  141. Chowdhury, I.A. State-of-the-Art CFD Simulation: A Review of Techniques, Validation Methods, and Application Scenarios. J. Recent Trends Mech. 2024, 9, 45–53. [Google Scholar] [CrossRef]
  142. Mirzaei, P.A. CFD modeling of micro and urban climates: Problems to be solved in the new decade. Sustain. Cities Soc. 2021, 69, 102839. [Google Scholar] [CrossRef]
  143. Park, M.A.; Loseille, A.; Krakos, J.; Michal, T.R.; Alonso, J.J. Unstructured grid adaptation: Status, potential impacts, and recommended investments towards CFD 2030. In Proceedings of the 46th AIAA Fluid Dynamics Conference, Washington, DC, USA, 13–17 June 2016; p. 3323. [Google Scholar]
  144. Deng, R.; Guo, Q.; Jia, M.; Wu, Y.; Zhou, Q.; Xu, Z. Extraction of farmland shelterbelts from remote sensing imagery based on a belt-oriented method. Front. For. Glob. Change 2023, 6, 1247032. [Google Scholar] [CrossRef]
  145. Chu, D. Remote Sensing of Land Use and Land Cover in Mountain Region; Springer: Berlin/Heidelberg, Germany, 2020. [Google Scholar]
  146. Guimarães, N.; Pádua, L.; Marques, P.; Silva, N.; Peres, E.; Sousa, J.J. Forestry remote sensing from unmanned aerial vehicles: A review focusing on the data, processing and potentialities. Remote Sens. 2020, 12, 1046. [Google Scholar] [CrossRef]
  147. Zheng, C.W.; Amr, A.D.; Vance, M.W.; Wang, X.; Dalid, C.; Shen, K. Strawberry canopy structural parameters estimation and growth analysis from UAV multispectral imagery using a geospatial tool. Comput. Electron. Agric. 2025, 235, 121710. [Google Scholar] [CrossRef]
  148. Zhang, L.; Shi, L.; Yang, F. Assessing ecosystem service dynamics in China’s coastal shelterbelt: Implications for ecosystem restoration. Environ. Impact Assess. Rev. 2024, 106, 107515. [Google Scholar] [CrossRef]
  149. Zhao, Y.; Wu, J.; He, C.; Ding, G. Linking wind erosion to ecosystem services in drylands: A landscape ecological approach. Landsc. Ecol. 2017, 32, 2399–2417. [Google Scholar] [CrossRef]
  150. Ding, M.; Yin, X.; Pan, S.; Liu, P. Multi-Objective Spatial Optimization of Protective Forests Based on the Non-Dominated Sorting Genetic Algorithm-II Algorithm and Future Land Use Simulation Model: A Case Study of Alaer City, China. Forests 2025, 16, 452. [Google Scholar] [CrossRef]
  151. Cardinia Shire Council. Shelterbelt Design Guidelines for Climate Change; Eco Logical Australia: Adelaide, Australia, 2022. [Google Scholar]
  152. Simiu, E.; Yeo, D. Wind Effects on Structures: Modern Structural Design for Wind; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  153. Vacek, Z.; Řeháček, D.; Cukor, J.; Vacek, S.; Khel, T.; Sharma, R.P.; Kučera, J.; Král, J.; Papaj, V. Windbreak efficiency in agricultural landscape of the Central Europe: Multiple approaches to wind erosion control. Environ. Manag. 2018, 62, 942–954. [Google Scholar] [CrossRef] [PubMed]
  154. Abdulsattar, E.; Abdelnaby, M.; Elnaggar, M. Comparison of the empirical formula and the “CFD”-based semi-empirical method in the prediction of the critical speed of flutter. Can. J. Civ. Eng. 2023, 50, 584–593. [Google Scholar] [CrossRef]
  155. Finnigan, J. Turbulence in plant canopies. Annu. Rev. Fluid Mech. 2000, 32, 519–571. [Google Scholar] [CrossRef]
  156. El Kenawy, A.M. Hydroclimatic extremes in arid and semi-arid regions: Status, challenges, and future outlook. In Hydroclimatic Extremes in the Middle East and North Africa; Elsevier: Amsterdam, The Netherlands, 2024; pp. 1–22. [Google Scholar]
  157. Tang, L.; Zhao, H.K.; Zhou, Z.F.; Qian, Z.X.; Hou, S.S.; Liu, B. Spatiotemporal distribution and driving factors of water use efficiency in the Yangtze River Delta urban agglomeration under the background of sustainability. Ecol. Ind. 2025, 175, 114159. [Google Scholar] [CrossRef]
  158. Ringgaard, R.; Herbst, M.; Friborg, T. Partitioning of forest evapotranspiration: The impact of edge effects and canopy structure. Agric. For. Meteorol. 2012, 166, 86–97. [Google Scholar] [CrossRef]
  159. Legesse, T.; Alemayehu, Y.; Haile, A. Effects of Deficit Irrigation Levels and Potato (Solanum tuberosum L.) Varieties on Yield, Yield Components, and Water Productivity at Haramaya, Ethiopia. Ph.D. Thesis, Haramaya University, Dire Dawa, Ethiopia, 2024. [Google Scholar]
  160. Yi, J.; Li, H.; Zhao, Y.; Shao, M.A.; Zhang, H.; Liu, M. Assessing soil water balance to optimize irrigation schedules of flood-irrigated maize fields with different cultivation histories in the arid region. Agric. Water Manag. 2022, 265, 107543. [Google Scholar] [CrossRef]
  161. Ma, C.; Tang, L.; Chang, W.; Jaffar, M.T.; Zhang, J.; Li, X.; Chang, Q.; Fan, J. Effect of shelterbelt construction on soil water characteristic curves in an extreme arid shifting desert. Water 2022, 14, 1803. [Google Scholar] [CrossRef]
  162. VijayKumar, R.; Das, J.; Upadhyay, L.; Kishore, P.; Saikanth, D. Principles and Practices of Agroforestry; N D Global Publication House: Mathura, India, 2024. [Google Scholar]
  163. Yunusa, I.A.; Brown, G.W.; Kwong, R.M.; Ronnfeldt, G.R.; Slater, T.; Crouch, A.; Unkovich, M. Integrating biodiversity and productivity on intensive farms: A potential role for shelterbelts in the Victorian Riverina. In Proceedings of the 2002 Australian Academy of Science Fenner Conference on the Environment, Canberra, Australia, 30 July 1–August 2002; Agriculture for the Australian Environment. The Johnstone Centre, Charles Sturt University: Albury, NSW, Australia, 2003; pp. 255–277. [Google Scholar]
  164. Asmare, M.T. The role of shelterbelt for soil management in Ethiopia. Field Veg. Crops Res. Ratar. Povrt. 2023, 60, 40–48. [Google Scholar] [CrossRef]
  165. Dhillon, G.S.; Van Rees, K.C. Soil organic carbon sequestration by shelterbelt agroforestry systems in Saskatchewan. Can. J. Soil Sci. 2017, 97, 394–409. [Google Scholar] [CrossRef]
  166. Wang, F.; Li, Z.; Xia, H.; Zou, B.; Li, N.; Liu, J.; Zhu, W. Effects of nitrogen-fixing and non-nitrogen-fixing tree species on soil properties and nitrogen transformation during forest restoration in southern China. Soil Sci. Plant Nutr. 2010, 56, 297–306. [Google Scholar] [CrossRef]
  167. Even, R.J.; Cotrufo, M.F. The ability of soils to aggregate, more than the state of aggregation, promotes protected soil organic matter formation. Geoderma 2024, 442, 116760. [Google Scholar] [CrossRef]
  168. Jolliet, O.; Antón, A.; Boulay, A.-M.; Cherubini, F.; Fantke, P.; Levasseur, A.; McKone, T.E.; Michelsen, O.; Milà i Canals, L.; Motoshita, M. Global guidance on environmental life cycle impact assessment indicators: Impacts of climate change, fine particulate matter formation, water consumption and land use. Int. J. Life Cycle Assess. 2018, 23, 2189–2207. [Google Scholar] [CrossRef]
  169. Sardeshpande, M.; Shackleton, C. Wild edible fruits: A systematic review of an under-researched multifunctional NTFP (non-timber forest product). Forests 2019, 10, 467. [Google Scholar] [CrossRef]
  170. Sheppard, J.P.; Chamberlain, J.; Agúndez, D.; Bhattacharya, P.; Chirwa, P.W.; Gontcharov, A.; Sagona, W.C.J.; Shen, H.-L.; Tadesse, W.; Mutke, S. Sustainable forest management beyond the timber-oriented status quo: Transitioning to co-production of timber and non-wood forest products—A global perspective. Curr. For. Rep. 2020, 6, 26–40. [Google Scholar] [CrossRef]
  171. Vogt, J.; Hauer, R.J.; Fischer, B.C. The costs of maintaining and not maintaining the urban forest: A review of the urban forestry and arboriculture literature. Arboric. Urban For. 2015, 41, 293–323. [Google Scholar] [CrossRef]
  172. Schroth, G.; Sinclair, F.L. Trees, Crops, and Soil Fertility: Concepts and Research Methods; Cabi: Wallingford, UK, 2003. [Google Scholar]
  173. Thornton, T.F.; Comberti, C. Synergies and trade-offs between adaptation, mitigation and development. Clim. Change 2017, 140, 5–18. [Google Scholar] [CrossRef]
  174. Gurr, G.M.; Wratten, S.D.; Snyder, W.E. Biodiversity and insect pests. In Biodiversity and Insect Pests: Key Issues for Sustainable Management; John Wiley & Sons: Hoboken, NJ, USA, 2012; pp. 1–20. [Google Scholar]
  175. Verma, S.; Pant, M.; Snasel, V. A comprehensive review on NSGA-II for multi-objective combinatorial optimization problems. IEEE Access 2021, 9, 57757–57791. [Google Scholar] [CrossRef]
  176. The Intergovernmental Panel on Climate Change. Climate change 2007: The physical science basis. Agenda 2007, 6, 333. [Google Scholar]
  177. Keprate, A.; Bhardwaj, D.; Sharma, P.; Verma, K.; Abbas, G.; Sharma, V.; Sharma, K.; Janju, S. Climate resilient agroforestry systems for sustainable land use and livelihood. In Transforming Agricultural Management for a Sustainable Future: Climate Change and Machine Learning Perspectives; Springer: Berlin/Heidelberg, Germany, 2024; pp. 141–161. [Google Scholar]
  178. Yadav, M.; Vashisht, B.; Jalota, S.; Kumar, A.; Kumar, D. Sustainable water management practices for intensified agriculture. In Soil-Water, Agriculture, and Climate Change: Exploring Linkages; Springer: Berlin/Heidelberg, Germany, 2022; pp. 131–161. [Google Scholar]
  179. Nõges, T.; Nõges, P.; Cardoso, A.C. Review of Published Climate Change Adaptation and Mitigation Measures Related with Water; Scientific and Technical Research Series EUR; Joint Research Centre, European Commission: Ispra, Italy, 2010; Volume 24682. [Google Scholar]
  180. Adhikari, B.; Agrawal, A. Understanding the social and ecological outcomes of PES projects: A review and an analysis. Conserv. Soc. 2013, 11, 359–374. [Google Scholar] [CrossRef]
  181. Papafotiou, M.; Pergialioti, N.; Tassoula, L.; Massas, I.; Kargas, G. Growth of native aromatic xerophytes in an extensive Mediterranean green roof as affected by substrate type and depth and irrigation frequency. HortScience 2013, 48, 1327–1333. [Google Scholar] [CrossRef]
  182. Sadok, W.; Lopez, J.R.; Smith, K.P. Transpiration increases under high-temperature stress: Potential mechanisms, trade-offs and prospects for crop resilience in a warming world. Plant Cell Environ. 2021, 44, 2102–2116. [Google Scholar] [CrossRef]
  183. Wagner, S.; Nocentini, S.; Huth, F.; Hoogstra-Klein, M. Forest management approaches for coping with the uncertainty of climate change: Trade-offs in service provisioning and adaptability. Ecol. Soc. 2014, 19, 32. [Google Scholar] [CrossRef]
  184. Rastogi, M.; Kolur, S.M.; Burud, A.; Sadineni, T.; Sekhar, M.; Kumar, R.; Rajput, A. Advancing water conservation techniques in agriculture for sustainable resource management: A review. J. Geogr. Environ. Earth Sci. Int. 2024, 28, 41–53. [Google Scholar] [CrossRef]
Figure 1. Conceptual diagram showing how shelterbelts modify microclimatic factors in arid croplands by reducing wind speed, increasing canopy humidity, moderating air and soil temperature, and lowering evapotranspiration rates, which together alleviate crop water stress.
Figure 1. Conceptual diagram showing how shelterbelts modify microclimatic factors in arid croplands by reducing wind speed, increasing canopy humidity, moderating air and soil temperature, and lowering evapotranspiration rates, which together alleviate crop water stress.
Agriculture 15 02004 g001
Figure 2. Pathway illustrating how shelterbelt-induced wind reduction, erosion control, and microclimate modification collectively enhance crop physiological processes and lead to yield and quality improvement in arid regions.
Figure 2. Pathway illustrating how shelterbelt-induced wind reduction, erosion control, and microclimate modification collectively enhance crop physiological processes and lead to yield and quality improvement in arid regions.
Agriculture 15 02004 g002
Figure 3. Impacts of shelterbelt porosity on wind flow pattern.
Figure 3. Impacts of shelterbelt porosity on wind flow pattern.
Agriculture 15 02004 g003
Figure 4. Schematic showing the scaling effect of shelterbelt height (H) on the spatial extent of leeward wind protection zones in arid farmland systems.
Figure 4. Schematic showing the scaling effect of shelterbelt height (H) on the spatial extent of leeward wind protection zones in arid farmland systems.
Agriculture 15 02004 g004
Figure 5. Integrated multi-species (tree–shrub–grass) configuration for arid shelterbelts, showing vertical stratification, porosity control, and biodiversity benefits.
Figure 5. Integrated multi-species (tree–shrub–grass) configuration for arid shelterbelts, showing vertical stratification, porosity control, and biodiversity benefits.
Agriculture 15 02004 g005
Table 1. Impact of Shelterbelt Structural Characteristics on Soil Erosion Suppression.
Table 1. Impact of Shelterbelt Structural Characteristics on Soil Erosion Suppression.
Structural
Characteristic
Impact on Erosion ControlMechanism and Notes
Optical PorosityModerate porosity (30–50%) is most effective.High porosity offers insufficient wind reduction; low porosity creates excessive turbulence and short protection distance.
Height (H)Directly determines the scale of the protected area.Protection distance is a multiple of H (e.g., 15–25 H). Taller belts protect a larger field area.
Number of RowsMulti-row belts are generally more robust and effective.Provides a deeper zone of wind deceleration and better particle trapping. Enhances durability if one row fails.
Species CompositionA mix of trees, shrubs, and ground cover is ideal.Shrubs fill the vertical gap near the soil surface, preventing wind tunneling and trapping saltating particles.
OrientationShould be perpendicular to the prevailing erosive winds.Maximizes the windbreak effect. Belts aligned with winds offer minimal protection.
WidthSufficient width is needed for structural integrity.Very narrow belts may be easily penetrated; very wide belts occupy excessive farmland. A
Table 2. Key Field Measurement Techniques and Parameters in Shelterbelt Studies.
Table 2. Key Field Measurement Techniques and Parameters in Shelterbelt Studies.
Measurement
Technique
Parameters AssessedAdvantagesLimitations
AnemometryWind speed reduction, turbulence intensityDirect in situ dataLimited spatial coverage
Soil Moisture SensorsSoil water content distributionHigh temporal resolutionPoint-scale measurement
Dendrometric SurveysTree height, DBH, crown widthAccurate structural dataDestructive/time-consuming
Photographic Porosity EstimationOptical porosity, leaf area index (LAI)Non-destructive, scalableWeather and light-dependent
Table 3. Typical Wind Tunnel Experimental Designs for Shelterbelt Optimization.
Table 3. Typical Wind Tunnel Experimental Designs for Shelterbelt Optimization.
Experimental SetupVariables TestedData CollectedApplications
Scaled Physical ModelsPorosity, density, multi-row configurationsVelocity decay, turbulence kineticsDesign refinement
Particle Image Velocimetry (PIV)Flow separation, wake characteristicsHigh-resolution flow fieldsMechanism analysis
Pressure Tap ArraysWind pressure distribution on leeward sideSurface pressure mapsStructural load assessment
Table 4. Common CFD Approaches in Shelterbelt Structure Modeling.
Table 4. Common CFD Approaches in Shelterbelt Structure Modeling.
CFD Model TypeKey FeaturesUse CasesLimitations
RANS ModelsSteady-state simulation, low computational costPreliminary design screeningLimited accuracy in turbulent flows
LES ModelsTransient turbulence resolutionDetailed wake analysisHigh computational demand
Porous Media ModelsSimplifies vegetation as porous zonesLarge-scale landscape planningRequires empirical porosity inputs
Table 5. Remote Sensing and GIS Techniques in Shelterbelt Assessment.
Table 5. Remote Sensing and GIS Techniques in Shelterbelt Assessment.
TechnologyData OutputsApplicationsChallenges
Multispectral ImageryNDVI, LAI, vegetation healthMonitoring canopy density and stressCoarse resolution for fine structures
LiDAR Scanning3D structure, height, porosityPrecision structural mappingHigh cost and processing complexity
GIS Spatial AnalysisSuitability maps, connectivity corridorsRegional planning and gap analysisData integration inconsistencies
Table 6. Summary of best-practice recommendations for windbreak design in arid agricultural regions.
Table 6. Summary of best-practice recommendations for windbreak design in arid agricultural regions.
ObjectiveRecommended PracticeNotes
Maximize wind reduction30–50% porosity; 15–30 H height; perpendicular orientationAvoid too dense belts to prevent turbulence
Conserve waterUse native xerophytes; space rows 5–10 m apartMinimize irrigation after establishment
Enhance biodiversityMulti-species, multi-strata designInclude shrubs and ground cover
Reduce costsUse locally available species; staggered plantingCombine timber/NTFP species for income
Improve durabilityPlan for pruning, thinning, and species renewalMonitor mortality and replace
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Aili, A.; Bakayisire, F.; Xu, H.; Waheed, A. Enhancing Agroecological Resilience in Arid Regions: A Review of Shelterbelt Structure and Function. Agriculture 2025, 15, 2004. https://doi.org/10.3390/agriculture15192004

AMA Style

Aili A, Bakayisire F, Xu H, Waheed A. Enhancing Agroecological Resilience in Arid Regions: A Review of Shelterbelt Structure and Function. Agriculture. 2025; 15(19):2004. https://doi.org/10.3390/agriculture15192004

Chicago/Turabian Style

Aili, Aishajiang, Fabiola Bakayisire, Hailiang Xu, and Abdul Waheed. 2025. "Enhancing Agroecological Resilience in Arid Regions: A Review of Shelterbelt Structure and Function" Agriculture 15, no. 19: 2004. https://doi.org/10.3390/agriculture15192004

APA Style

Aili, A., Bakayisire, F., Xu, H., & Waheed, A. (2025). Enhancing Agroecological Resilience in Arid Regions: A Review of Shelterbelt Structure and Function. Agriculture, 15(19), 2004. https://doi.org/10.3390/agriculture15192004

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