Fact-checked by Grok 2 weeks ago

Terrestrial ecosystem

A terrestrial ecosystem is a land-based comprising interdependent communities of organisms—such as , animals, fungi, and microbes—interacting with abiotic components like , , , and atmospheric conditions. These ecosystems differ from ones by relying on and rather than submerged water bodies for primary habitat structure, leading to adaptations shaped by factors including temperature extremes, , and nutrient availability in terrestrial soils. Terrestrial ecosystems are classified into major biomes defined by prevailing climate, vegetation dominance, and ecological processes, including , boreal forests (), temperate forests, temperate grasslands, deserts, (Mediterranean shrublands), savannas, and tropical rainforests. Each biome exhibits distinct characteristics: for instance, tropical rainforests feature high rainfall (>2000 mm annually) and multilayered canopies supporting exceptional , while deserts endure low (<250 mm annually) with sparse, drought-resistant flora and fauna. These biomes collectively form the terrestrial biosphere, driving key global functions such as primary production through photosynthesis and decomposition cycles that recycle nutrients. Terrestrial ecosystems underpin Earth's habitability by regulating atmospheric composition, with vegetation acting as a major that sequesters approximately 3.6 Pg of carbon per year, mitigating atmospheric CO₂ accumulation through biomass growth and soil storage. They support over 80% of terrestrial species diversity, influencing evolutionary patterns via habitat specialization, and provide essential services like pollination, soil formation, and freshwater filtration, though their dynamics reflect natural variability in climate and disturbance regimes such as fire and herbivory. Empirical observations indicate that while human activities have altered biome extents—e.g., through conversion to agriculture—these systems demonstrate resilience via successional recovery and adaptation to environmental gradients.

Definition and Fundamentals

Definition

A terrestrial ecosystem comprises a community of interacting biotic components—such as plants, animals, fungi, and microorganisms—and abiotic factors, including soil, climate, topography, and atmospheric conditions, occurring exclusively on landmasses rather than in water bodies. These systems form self-sustaining units where energy flows primarily from solar radiation captured by photosynthetic autotrophs, supporting heterotrophic consumers and decomposers through trophic interactions. Unlike aquatic ecosystems, terrestrial ones rely on precipitation and soil moisture for water cycling, with gravity-driven drainage and evaporation dominating hydrological processes over buoyant dispersion. Terrestrial ecosystems encompass diverse biomes shaped by latitudinal gradients, elevation, and edaphic properties, covering approximately 149 million square kilometers or 29% of Earth's surface as of satellite mapping data from the early 21st century. They function as major regulators of global biogeochemical cycles, sequestering carbon via vegetation biomass—estimated at 2,500 gigatons in soils and biota—and facilitating nutrient turnover through litter decomposition rates that vary from 0.1 to 5 years depending on temperature and moisture. Empirical studies emphasize their resilience to perturbations via biodiversity, where species richness correlates with functional stability, as evidenced by long-term monitoring plots showing higher recovery rates in diverse plots post-disturbance.

Core Characteristics

Terrestrial ecosystems consist of land-based communities of organisms interacting with abiotic components, including soil, air, temperature, precipitation, and topography, distinguishing them from aquatic systems where water serves as the primary medium. These ecosystems span approximately 144 million square kilometers, or 28% of Earth's surface, and are characterized by the dominance of reinforced by lignin, which enables structural rigidity and upright growth absent in most aquatic vegetation. Fungal decomposers play a central role in processing lignocellulose from plant litter, supporting soil humus formation and nutrient retention, with detrital food chains often prevailing over direct grazing pathways. Climate variables, particularly temperature ranges and annual precipitation, primarily dictate vegetation composition, productivity, and biome classification, with higher evaporation rates and desiccation risks necessitating specialized adaptations in flora and fauna, such as waxy cuticles, deep roots, and respiratory systems suited to gaseous exchange in air. Soil serves as the foundational substrate for anchorage, water filtration, and nutrient cycling, facilitating processes like nitrogen fixation by microbes and organic matter decomposition, which sustain primary productivity driven by solar-powered photosynthesis. Unlike aquatic environments, terrestrial systems exhibit pronounced vertical stratification, from canopy layers capturing light to subterranean root zones accessing resources, influenced by gravity and light gradients. Energy flow and biogeochemical cycles in terrestrial ecosystems emphasize carbon sequestration via plant biomass and soil organic matter, with fungi and bacteria mediating decomposition rates that vary by moisture and temperature; for instance, lignin decomposition is slower in drier conditions, enhancing long-term carbon storage. Biodiversity patterns reflect these constraints, with hotspots in mesic regions supporting complex trophic webs, while aridity limits species richness in deserts. These characteristics underscore terrestrial ecosystems' sensitivity to perturbations like altered precipitation regimes, as evidenced by empirical studies linking climate variability to shifts in productivity axes.

Evolutionary History

Initial Land Colonization

The colonization of terrestrial environments by living organisms marked a pivotal transition from aquatic origins, beginning with microbial communities rather than macroscopic plants or animals. Prokaryotic microbes, including , established the first land-based populations around 2.75 billion years ago, as evidenced by carbon isotope ratios in ancient South African paleosols indicating extensive bacterial activity on exposed land surfaces. These early colonizers likely formed protective biofilms and mats in intertidal or periodically desiccated zones, facilitating soil formation through biogenic weathering and nitrogen fixation, which laid the groundwork for later ecosystems. By 2.6 billion years ago, such prokaryotic terrestrialization had become widespread, predating eukaryotic dominance and altering early atmospheric chemistry via oxygen production. Eukaryotic contributions emerged later, with non-vascular plants—precursors resembling modern liverworts and mosses—appearing during the middle , approximately 470 million years ago. Fossilized spores from Saudi Arabian and other deposits confirm these bryophyte-like organisms formed sparse cryptogamic covers alongside fungi, algae, and bacteria, occupying damp, coastal habitats before the evolution of . This phase, termed cryptogamic terrestrialization, involved symbiotic associations, particularly between early land plants and fungi (e.g., arbuscular mycorrhizae), which enhanced nutrient uptake from mineral substrates and mitigated desiccation stress, enabling persistence in low-water environments. Molecular clock analyses suggest the common ancestor of extant land plants diverged from aquatic around 500 million years ago in the , with genetic adaptations for terrestrial stress (e.g., UV resistance and drought tolerance) evolving concurrently. Vascular plants, featuring specialized xylem and phloem for water and nutrient transport, colonized land during the late Silurian, around 428–423 million years ago, initiating more structured ecosystems. Early tracheophytes, such as Cooksonia-like rhyniophytes, grew as small, leafless axes up to 10 cm tall in Rhynie Chert fossils from Scotland, dependent on mycorrhizal partnerships for phosphorus acquisition in nutrient-poor soils. This vascular innovation correlated with increased atmospheric oxygen levels (from ~10% to 15–20%) and CO2 drawdown, driving glaciation events like the Late Ordovician ice age, as pioneer plants accelerated silicate weathering and organic burial. Arthropods, including primitive myriapods (millipedes and centipedes), followed as the first terrestrial animals around 420 million years ago, scavenging detritus in these nascent plant-fungal communities and further promoting soil development through bioturbation. These interdependent microbial, fungal, and plant colonizers transformed barren continents into biologically productive landscapes, setting the stage for Devonian diversification.

Key Evolutionary Milestones

The emergence of vascular plants in the Silurian Period, around 430 million years ago, represented a critical advancement, enabling the internal transport of water and minerals via xylem and phloem tissues, which supported upright growth and resistance to desiccation beyond the capabilities of non-vascular bryophytes. This physiological innovation, exemplified by early rhyniophytes like Cooksonia, allowed plants to achieve heights of up to 50 cm and form rudimentary stands, fostering initial soil stabilization and organic matter accumulation that underpinned nascent terrestrial food chains. By the Devonian Period (approximately 419–359 million years ago), the evolution of lignified woody tissues and true roots further propelled ecosystem development, culminating in the first forests dominated by progymnosperms and early ferns around 385 million years ago. These structures, reaching heights of 10–30 meters in some lycopsid trees, enhanced carbon sequestration, atmospheric oxygen levels (rising to 30% by the late Devonian), and habitat complexity, enabling stratified canopies that supported expanding arthropod communities and early detritivores. The origin of seed plants in the late Devonian, circa 360 million years ago with fossils like Elkinsia polymorpha, decoupled reproduction from moist environments by protecting embryos in durable seeds, facilitating colonization of arid interiors and seasonal variability. This reproductive strategy diversified during the Carboniferous and Permian, promoting resilient vegetation that sustained megaherbivores and intensified global coal formation through vast peatlands. Arthropod diversification paralleled plant advances, with the earliest unequivocal terrestrial forms—myriapods—appearing in the late Silurian (around 420 million years ago), functioning primarily as decomposers and initiating soil food webs through litter breakdown. Winged insects evolved by the Devonian or early Carboniferous (approximately 400–350 million years ago), revolutionizing dispersal, predation, and plant-animal interactions via enhanced mobility and the first instances of flight-mediated pollination precursors. Vertebrate colonization accelerated trophic complexity, as stem-tetrapods transitioned from aquatic origins to semi-terrestrial habits in the late Devonian (around 375 million years ago), with forms like Ichthyostega exploiting vegetated shorelines. Full terrestrialization by amniotes in the Carboniferous (post-359 million years ago) introduced efficient lungs, watertight eggs, and diverse guilds of herbivores and carnivores, stabilizing predator-prey dynamics amid coal swamp ecosystems. The Cretaceous radiation of angiosperms, originating around 140 million years ago, marked a transformative endpoint, with flowers and fruits enabling precise insect pollination and vertebrate seed dispersal, which catalyzed biodiversity surges—angiosperms comprising over 90% of modern plant species and reshaping herbivory, nutrient cycling, and forest structures. This co-evolutionary surge, evidenced by fossil pollen records from 132 million years ago, amplified ecosystem productivity and resilience, setting the stage for Cenozoic dominance.

Structural Components

Biotic Elements

Biotic elements encompass all living organisms within terrestrial ecosystems, including plants, animals, fungi, bacteria, and other microorganisms that interact through trophic relationships and ecological processes. These components are classified primarily into producers, consumers, and decomposers based on their roles in energy flow and nutrient cycling. Producers, predominantly autotrophic organisms such as vascular plants (e.g., trees and grasses) and non-vascular plants like mosses, harness solar energy via photosynthesis to synthesize organic compounds from carbon dioxide, water, and minerals, forming the base of the food web. Consumers are heterotrophic organisms that obtain energy by consuming other organisms, subdivided into primary consumers (herbivores such as insects, rodents, and ungulates that feed directly on producers), secondary consumers (carnivores like birds of prey and small mammals that prey on herbivores), and tertiary consumers (apex predators such as large carnivores including lions or eagles that regulate lower trophic levels). Decomposers, mainly fungi and soil bacteria, break down dead organic matter and waste products through extracellular enzymatic digestion, releasing essential nutrients like nitrogen and phosphorus back into the soil for reuse by producers, thereby preventing nutrient depletion and maintaining ecosystem productivity. In terrestrial settings, decomposer activity is particularly vital in soils, where it influences organic matter decomposition rates, which can vary from months in temperate forests to years in arid deserts due to moisture and temperature constraints. Biotic interactions among these elements drive community dynamics, including competition for limited resources like light, water, and nutrients, which can limit species distributions and promote niche partitioning; predation and herbivory, which control population sizes and prevent overgrazing; and mutualistic relationships, such as pollination by insects or between plant roots and fungi that enhance nutrient uptake. These interactions contribute to biodiversity patterns, with terrestrial ecosystems exhibiting varying species richness— for instance, tropical forests support over 50% of global terrestrial species despite covering only 6% of land area, fostering resilience against disturbances through functional redundancy. High biotic diversity correlates with enhanced ecosystem multifunctionality, including sustained productivity and resistance to invasions, as evidenced by studies linking plant species richness to soil processes and . Disruptions to biotic elements, such as through species loss, can cascade through food webs, altering energy transfer efficiency, which typically declines from 10-20% at each trophic level due to metabolic losses.

Abiotic Factors

Abiotic factors in terrestrial ecosystems comprise the non-living physical and chemical components that regulate biological processes, species distributions, and community structures. Climatic variables, particularly temperature and precipitation, exert primary control over these ecosystems by delineating physiological tolerances and resource availability; for instance, mean annual temperatures ranging from below -10°C in polar regions to over 25°C in tropical zones, coupled with annual precipitation gradients from under 250 mm in deserts to exceeding 2000 mm in rainforests, define biome boundaries and limit plant growth rates and metabolic activities. Solar radiation, varying latitudinally due to Earth's axial tilt and atmospheric attenuation, drives photosynthesis and seasonal cycles, with photosynthetically active radiation (PAR) levels influencing primary productivity; ecosystems at higher latitudes receive less annual insolation, constraining net primary production to approximately 100-500 g C/m²/year compared to 1000-2000 g C/m²/year in equatorial zones. Wind, as a dynamic force, affects transpiration rates, seed dispersal, and erosion, with sustained speeds above 10 m/s capable of mechanically damaging vegetation and altering microclimates in exposed terrains. Edaphic factors, rooted in soil characteristics, profoundly shape nutrient cycling and microbial activity, often overriding biotic influences in determining ecosystem multifunctionality. texture—comprising sand, silt, and clay proportions—governs water retention and aeration; clay-rich soils (e.g., >40% clay) enhance but impede drainage, fostering conditions that suppress aerobic decomposers, whereas sandy soils (>70% sand) promote rapid infiltration yet limit nutrient retention, reducing plant by up to 50% in nutrient-poor settings. , typically ranging from 4.0 in acidic coniferous s to 8.0 in calcareous grasslands, modulates metal and activities; acidic conditions (pH <5.5) mobilize aluminum toxicity, inhibiting root growth and microbial carbon use efficiency, while neutral to alkaline supports higher organic matter decomposition rates. Organic carbon and nitrogen content, varying from <1% in arid soils to >10% in humid layers, directly correlate with soil biota diversity and ecosystem services like . Topographic features, including , , and , modify local abiotic conditions through altitudinal gradients and effects, thereby creating heterogeneity within ecosystems. induces a of approximately 0.6-1.0°C decrease per 100 m rise, compressing regimes and increasing frequency above 2000 m, which limits tree lines and shifts vegetation zonation; for example, exhibit compressed transitions over 1000-3000 m vertical spans. and influence insolation and moisture; south-facing slopes in the receive 20-50% more , accelerating and favoring xerophytic species, while north-facing slopes retain higher , supporting mesic communities. These physiographic elements interact with to amplify or buffer extremes, as evidenced by channeling in valleys that exacerbates disturbance regimes. Overall, abiotic factors exhibit hierarchical dominance, with macro-scale setting broad constraints and micro-scale edaphic and topographic variations enabling fine-scale adaptations.

Major Biomes

Forest Ecosystems

Forest ecosystems are terrestrial biomes dominated by trees forming dense canopies, supporting complex communities of , animals, fungi, and microorganisms that interact through intricate trophic and symbiotic relationships. These ecosystems span approximately 4.14 billion hectares, covering 32% of Earth's land surface as of recent assessments, equivalent to about 0.5 hectares per globally. They exhibit high primary due to efficient capture by multi-layered , with trees typically exceeding 5 meters in and crown cover surpassing 10% over areas larger than 0.5 hectares, per standardized definitions used in global inventories. Major forest types are classified primarily by climate, latitude, and dominant vegetation: tropical rainforests, temperate forests, and boreal forests (). Tropical forests, concentrated in equatorial regions like the and basins, feature evergreen broadleaf trees with minimal seasonal leaf drop, annual rainfall exceeding 2000 mm, and temperatures averaging 25–27°C, fostering year-round growth. Temperate forests, found in mid-latitudes such as eastern and , include species that shed leaves in winter, with around 750–1500 mm annually and temperature ranges from -30°C to 30°C, leading to pronounced seasonal cycles. Boreal forests, spanning high northern latitudes in , , and , are conifer-dominated with species like and adapted to short growing seasons, cold winters below -40°C, and low of 300–850 mm, much of it as snow. Structurally, forests display vertical : emergent trees piercing the canopy, a dense overstory layer, shrubs, herbaceous plants, and , which together sustain diverse microhabitats and facilitate nutrient retention. Abiotic drivers like —often nutrient-poor in due to rapid but enriched by in temperate zones—interact with elements to regulate rates and microbial activity. Forests harbor exceptional , hosting over 80% of terrestrial , including 60,000 , 80% of amphibians, 75% of , and 68% of mammals, with tropical variants alone supporting more than half of global diversity. Functionally, forest ecosystems excel in biogeochemical cycling, where fallen leaves and woody debris decompose to recycle nutrients like and via mycorrhizal fungi and , maintaining despite heavy uptake by . They serve as major carbon sinks, sequestering atmospheric CO2 through and storing it in long-lived wood, with global s holding about 45% of terrestrial carbon despite occupying 31% of land. Energy flows through detritus-based webs in litter-rich floors and grazer chains in canopies, supporting like decomposers that prevent nutrient loss. Succession progresses from post-disturbance (e.g., in zones) to climax communities, enhancing via redundancy. These processes underscore forests' causal role in stabilizing , regulating through (contributing 40% of continental rainfall), and buffering against , though human activities have reduced primary forest extent by altering these dynamics.

Grassland and Shrubland Ecosystems

Grasslands constitute expansive terrestrial ecosystems primarily dominated by vegetation, including from the family and grass-like plants such as sedges and rushes, with tree cover typically limited to less than 10% of the landscape. These thrive in climates featuring seasonal that supports herbaceous growth but restricts taller woody , often maintained by recurrent disturbances like and herbivory. Temperate , such as North American prairies and Eurasian steppes, exhibit some of the most fertile soils globally due to deep root systems that cycle nutrients efficiently and resist . Tropical variants, including savannas, incorporate scattered and accommodate migratory herds, reflecting adaptations to wet-dry cycles. Shrublands, distinct yet sometimes transitional with grasslands, are characterized by dominance of low woody shrubs under 5 meters tall, forming a single canopy layer often interspersed with grasses and forbs, in environments stressed by , nutrient-poor soils, or frequent fires. These ecosystems predominate in Mediterranean-type climates with hot, dry summers and mild, wet winters, receiving 200 to 1000 mm of annual rainfall concentrated in cooler months. Examples include in , in , and kwongan in , where shrub species exhibit sclerophyllous leaves and resprouting abilities post-fire, enhancing resilience to periodic . Collectively, grasslands and shrublands span approximately 40-50% of Earth's ice-free land surface, encompassing regions from the of to the Patagonian , African , and Central Asian steppes for grasslands, and coastal zones around the , southwestern , and for shrublands. gradients—typically 250-1000 mm annually for most grasslands and lower for shrublands—dictate their boundaries, with aridity and temperature extremes (0-25°C means) preventing forest encroachment while averting full . Vegetation in these biomes emphasizes drought- and fire-adapted species; grasslands feature or grasses with extensive root networks penetrating up to 2-3 meters, fostering stability and carbon storage, while shrublands host or shrubs with thick bark and serotinous seed release triggered by heat. Fauna diversity is pronounced, supporting herbivores like (Bison bison) in North American tallgrass prairies and in African savannas, alongside predators such as wolves and ; burrowing mammals and further enhance nutrient turnover. These ecosystems maintain openness through ecological processes: lightning-ignited fires every 1-5 years in many grasslands recycle nutrients and suppress trees, while intense grazing by native ungulates prunes competitors, as evidenced in Serengeti-Mara dynamics where migrations correlate with grass regrowth rates exceeding 10 cm per week post-rain. In shrublands, fire intervals of 10-50 years promote obligate-seeder shrubs, with post-burn biomass recovery reaching 50-70% within two years in systems. Biodiversity hotspots within these biomes rival some forests, with grasslands hosting up to 100+ species per square meter in undisturbed patches and supporting endemic ; however, richness peaks in mesic areas, declining toward arid margins. Shrublands contribute unique assemblages, including pollination-dependent and specialist reptiles, underscoring their role in global beta-diversity gradients. Despite lower aboveground than forests (averaging 1-5 kg/m² dry weight), these biomes excel in belowground productivity, storing 30-50% of terrestrial .

Desert Ecosystems

Desert ecosystems encompass arid terrestrial biomes where annual precipitation is typically less than 250 mm, and potential evapotranspiration greatly exceeds rainfall, leading to pervasive water scarcity. Abiotic conditions include extreme temperature diurnal ranges—often exceeding 30°C, with daytime maxima up to 54°C and nocturnal minima near 4°C in hot deserts—and coarse, nutrient-poor soils such as sandy or rocky substrates with minimal organic matter and high drainage rates. These factors, driven by atmospheric circulation patterns like subtropical high-pressure zones or rain shadows from mountain ranges, constrain primary productivity and foster specialized landforms including dunes up to 180 m high and ephemeral salt flats. Deserts are classified into four principal types based on climate and location: hot and dry (e.g., , covering 9 million km², with rainfall under 50 cm/year and evaporation rates far surpassing it); semiarid (e.g., , with 2-4 cm monthly summer rain); coastal (e.g., Atacama, receiving as little as 1.5 cm/year due to cold ocean ); and cold (e.g., Gobi or interiors, with 15-26 cm precipitation mostly as snow and summer highs of 21-26°C). Each type exhibits low and sparse cloud cover, amplifying solar radiation and wind erosion, which further depletes . Biotic communities feature low plant density with adaptations for , such as () in cacti, where stomata open nocturnally to curb ; reduced or absent leaves replaced by photosynthetic green stems; and extensive root systems, including taproots exceeding 30 m in or shallow lateral spreads in . , termed xerocoles, predominantly display , burrowing to evade heat, and physiological efficiencies like camels tolerating 30% body water loss or rats producing concentrated urine without drinking. Reptiles, , and small mammals dominate, with examples including the and lizard, which channels via skin grooves for hydration. Functional processes in deserts hinge on episodic rainfall events that trigger ephemeral blooms and pulses, but chronic slows decomposition and biogeochemical cycling. Burrowing macro-detritivores, such as isopods (e.g., ), play a pivotal role by fragmenting litter—accounting for up to 89% of removal—and excreting nutrient-rich fecal pellets, elevating soil ammonium by 1.5-fold and nitrate by 2-fold near burrows, thus fostering microbial mineralization in otherwise oligotrophic conditions. aquifers and flash floods episodically recharge systems, carving arroyos and distributing sediments, though overall energy flow remains low due to limited producer . Biodiversity is generally low, with few species adapted to extremes, yet is high among xerocoles; for instance, hot deserts support diverse reptiles and annual plants that complete life cycles post-rain. These ecosystems prove fragile, as minor perturbations in abiotic factors can disrupt balances, exemplified by accelerated rates of 6 million km² annually from or climatic shifts.

Tundra and Alpine Ecosystems

Tundra ecosystems, including both and variants, represent some of the harshest terrestrial environments, defined by prolonged cold temperatures, low precipitation, and the absence of trees due to physiological constraints on growth. tundra occurs in high-latitude regions north of the treeline, such as northern , , , and , where —a permanently frozen soil layer—underlies much of the landscape, restricting root penetration and drainage. , by contrast, forms at high elevations above the treeline on mountains worldwide, including the , , and , lacking widespread but featuring well-drained, rocky soils that limit water retention. Both types experience short growing seasons of 6 to 10 weeks, with annual precipitation typically below 25 cm, often as snow, rendering them functionally similar to deserts despite the icy appearance. Climatic conditions in ecosystems are extreme, with average annual temperatures around -12°C to -28°C; winter lows can plummet to -50°C or below, while brief summers rarely exceed 10°C monthly averages. In arctic , persistent low temperatures maintain , which covers up to 25% of the Northern Hemisphere's land surface and cycles seasonally, thawing only the active layer (10-100 cm deep) in summer. exhibits greater diurnal and seasonal temperature fluctuations due to , with less extreme annual cold but frequent freezing nights even in summer; precipitation patterns vary by but remain low, often 15-30 cm annually, supplemented by or in some areas. These abiotic factors—intense solar radiation at high latitudes or altitudes combined with short photoperiods—impose severe limitations on and nutrient cycling, favoring slow-growing, stress-tolerant organisms. Vegetation in tundra ecosystems is dominated by low-stature perennials adapted to nutrient-poor soils and mechanical stress from wind and ice, including sedges, grasses, mosses, lichens, and dwarf shrubs like and , which rarely exceed 30 cm in height. tundra plant communities form polygonal patterns from , with tussock tundra featuring cotton grasses and lichens covering vast areas; alpine variants show similar assemblages but with more forbs and cushion plants that trap heat and reduce wind exposure. These plants employ strategies such as , shallow roots exploiting the active layer, and physiological to endure scarcity, where decomposition rates are minimal due to cold, leading to thin organic horizons. Faunal diversity is low compared to temperate biomes, with fewer than 50 species per 1000 m² in many areas, though arctic hosts over 21,000 known cold-adapted across taxa when including microbes and . Herbivores like caribou (Rangifer tarandus), lemmings (Lemmus spp.), and hares (Lepus arcticus) dominate food webs, migrating seasonally or exhibiting population cycles that influence vegetation dynamics. Predators such as foxes (Vulpes lagopus) and , including snowy owls (Bubo scandiacus), rely on these cycles, while and migratory birds exploit the brief summer ; alpine fauna overlap with arctic in like pikas (Ochotona spp.) and ptarmigan but include more montane specialists adapted to steeper terrains. Overall, ecosystems sustain unique endemics, contributing disproportionately to despite harsh filters that select for generalist, resilient traits.

Functional Processes

Energy Transfer and Food Webs

Energy enters terrestrial ecosystems predominantly through , where organisms such as vascular capture solar radiation to convert and into organic compounds, forming the base of energy flow. This process yields gross primary production (GPP), with net (NPP) representing the available after autotroph , typically ranging from 0.5 to 2.5 kg/m²/year in temperate forests and grasslands. Heterotrophs then access this energy via consumption, with decomposers recycling to sustain soil nutrient cycles integral to long-term productivity. Trophic levels organize this transfer into hierarchical categories: producers at the base, followed by primary consumers (herbivores like and ungulates), secondary consumers (carnivores preying on herbivores), consumers, and apex predators. In terrestrial settings, such as savannas, browsers like giraffes occupy primary levels while lions serve as apex predators, illustrating predator-prey dynamics that regulate population sizes. Decomposers, including and fungi, operate across levels by breaking down , channeling back into producers via mineralization rather than direct trophic ascent. Food webs depict these interactions as interconnected networks rather than isolated linear food chains, capturing multiple pathways and contingencies like alternative prey or omnivory. For instance, in forest ecosystems, a web might link foliage consumed by caterpillars, which are eaten by birds or spiders, with uneaten litter supporting detritivores that indirectly feed invertebrates preyed upon by the same birds. This complexity enhances , as perturbations in one link—such as outbreaks—can be buffered by compensatory predation or foraging shifts. Energy transfer efficiency between trophic levels averages approximately 10%, as formalized in Lindeman's 1942 trophic-dynamic model, due to losses from metabolic (converting ~60-70% of ingested to ), incomplete consumption, and indigestible . Empirical studies confirm this "10% rule" approximation holds across terrestrial systems, with herbivore assimilation efficiencies around 15-20% and carnivore levels lower at 10-15%, limiting higher trophic levels to sparse . Consequently, energy pyramids depict upright structures with exponentially declining energy flux per level, constraining most terrestrial ecosystems to 3-4 trophic levels beyond producers. This inefficiency underscores the dependence of top predators on vast bases, as seen in African savannas where populations correlate with herbivore densities sustained by grass NPP exceeding 1,000 g/m²/year.

Biogeochemical Cycles

Biogeochemical cycles regulate the flow of essential elements through terrestrial ecosystems, linking biological activity with geological and atmospheric processes to sustain and . These cycles—primarily involving , , , and —operate via microbial transformations, plant uptake, , and abiotic transport, with rates influenced by , type, and properties. In terrestrial systems, biotic components drive rapid internal recycling, while external inputs from or atmospheric deposition provide long-term replenishment; imbalances, such as excess deposition, can accelerate losses through or gaseous emissions, altering . Empirical measurements, including tracing and towers, reveal tight coupling among cycles, where carbon fixation enhances demand, and limitation constrains productivity in weathered soils. The centers on photosynthetic fixation by , which captures atmospheric CO₂ into , with global terrestrial gross estimated at approximately 120 Pg C yr⁻¹. Autotrophic and heterotrophic , along with fires and , return roughly equivalent amounts to the atmosphere, yielding a net terrestrial sink of about 3 Pg C yr⁻¹ that absorbs 25-30% of emissions. serves as the largest terrestrial reservoir, holding 1500-2500 Pg C, where microbial rates vary with temperature and moisture; for instance, fluxes exhibit thermal thresholds around -5°C to -2°C in cold regions and 17°C in warmer ones, beyond which accelerates disproportionately to . Human perturbations, including , reduce this sink capacity, as evidenced by data showing decreased net exchange in disturbed forests. In the , atmospheric N₂ fixation by free-living microbes and symbioses supplies ~62 Tg N yr⁻¹ to global terrestrial ecosystems, primarily supporting plant growth via mineralization of into , followed by to . and remove fixed , with gross transformation rates—such as mineralization at 50-200 mg N kg⁻¹ d⁻¹ in temperate forests—exceeding net uptake and leading to losses in N-saturated systems. These processes are microbially mediated and sensitive to and oxygen levels; for example, conditions in wetlands promote , converting to N₂ and N₂O, contributing to ~10-20% of global N₂O emissions from land. Anthropogenic fertilizers have doubled reactive inputs since pre-industrial times, elevating cycling rates but fostering feedbacks. The in terrestrial ecosystems relies on rock as the primary external input, with internal biotic recycling dominating fluxes at 70-1000 Pg P yr⁻¹ through uptake, litterfall, and microbial mineralization in . Unlike , lacks a significant gaseous phase, resulting in slow mobility and accumulation in organic forms; pools range from 200-1500 mg kg⁻¹, but is limited by to minerals like iron oxides, particularly in tropical, highly weathered where P constrains net . rates, estimated at 1-10 kg P ha⁻¹ yr⁻¹ in unglaciated landscapes, decline over millennia, leading to progressive P depletion; empirical studies in old-growth forests show mycorrhizal fungi enhance acquisition, recycling up to 90% of annual uptake. Climate-driven can export P via runoff, reducing long-term fertility in vulnerable ecosystems. The , integral to terrestrial , facilitates element transport via , which accounts for 60-90% of return to the atmosphere in vegetated lands, with global terrestrial rates around 65,000 km³ yr⁻¹. inputs vary regionally—e.g., 500-2000 mm yr⁻¹ in forests—while by plants, driven by , links to carbon and cycles, as catalyzes ~40% of continental recycling. Infiltration and storage buffer dry periods, but runoff dominates in sparse-cover biomes like deserts; microbial activity peaks at optimal (20-60% water-filled pore space), influencing and mineralization rates. Alterations from land-use change, such as increasing by 10-20% in semi-arid zones, demonstrate causal feedbacks on local and cycle coupling.

Community Dynamics and Succession

Community dynamics in terrestrial ecosystems encompass the temporal and spatial variations in species composition, population sizes, and interspecific interactions within biological communities, shaped primarily by processes such as for resources, predation, herbivory, , facilitation, and dispersal. These dynamics are influenced by both biotic feedbacks, like trophic interactions across levels from producers to apex predators, and abiotic drivers including variability and disturbances such as or , which can alter stability and . Empirical studies indicate that indirect interactions, such as apparent mediated through shared predators or herbivores, often propagate effects across and animal guilds, contributing to non-linear responses in structure under changing environmental conditions. For instance, in communities, outbreaks can suppress dominant tree regeneration while favoring shrubs, demonstrating how pressure modulates competitive hierarchies. Ecological succession represents a directed sequence of assembly and replacement following disturbance or primary , driven by causal mechanisms including facilitation of modification, inhibition by early dominants, and eventual tolerance of late-successional to local conditions. In terrestrial systems, primary succession initiates on substrates devoid of and propagule banks, such as glacial moraines or volcanic lava flows, where initial colonizers like lichens and mosses weather rock to form , enabling establishment over centuries; evidence from Alaskan glacial forelands shows lichen-moss stages persisting 50-100 years before invasion. , conversely, occurs on disturbed but soil-retaining sites like post-fire forests or abandoned fields, accelerating due to residual seed banks and root sprouts; rates vary by , with recovery in temperate zones reaching canopy closure in 20-50 years versus over 200 years for primary in tropical regions. In forest ecosystems, secondary succession post-logging or typically progresses from herbaceous pioneers (e.g., fireweed in stands) to thickets, then shade-tolerant trees, with facilitation via nitrogen-fixing species enhancing ; a study of Yellowstone post-1988 fires documented lodgepole dominance returning within 15-20 years, though outbreaks introduced alternative trajectories. succession, often arrested by or , shifts from annuals to perennial bunchgrasses in early stages, potentially to woody encroachment if disturbances cease; in North American shortgrass steppes, seeded restorations showed reducing dominance and favoring native perennials within 5-10 years compared to ungrazed controls. and successions proceed slowly due to harsh abiotic constraints, with cryptobiotic crusts in arid zones preceding vascular plants over decades, while sites post-glaciation exhibit herb-dominated communities stabilizing after 100+ years. Factors governing these dynamics and succession rates include disturbance frequency, climate (e.g., precipitation driving grassland-to-shrub transitions), soil legacy effects, and propagule availability, with recent analyses revealing environmental variables explaining up to 90% of biome-specific stability variations. Threshold dynamics can lead to hysteresis, where communities fail to revert post-disturbance, as seen in overgrazed savannas shifting to shrublands resistant to grass recovery. Global change amplifies variability, with elevated CO2 and warming accelerating early-successional productivity but disrupting late-stage dominance in some biomes, underscoring the role of empirical monitoring in predicting outcomes.

Biodiversity and Patterns

Spatial Distribution of Diversity

The spatial distribution of in terrestrial ecosystems is characterized primarily by a latitudinal (LDG), in which declines from tropical latitudes toward the poles. This pattern holds across major terrestrial taxa, including , vertebrates, , and prokaryotes, with empirical data from vascular showing peak richness in equatorial bands between 23.5°N and 23.5°S. records indicate the LDG has persisted through geological epochs, including the Permo-Triassic transition for tetrapods, though with varying intensity under different climatic regimes. In mosses, a non-vascular terrestrial group, exhibits a near-perfect negative (r = -0.99) with increasing latitude, underscoring the gradient's pervasiveness even in phylogenetically distinct lineages. Tropical and subtropical regions dominate global terrestrial diversity, with biodiversity hotspots—regions harboring at least 1,500 endemic species and having lost over 70% of primary —concentrating much of this variation. Currently, 36 such hotspots are recognized, spanning approximately 16.7% of Earth's land surface but containing over 50% of species and 42% of terrestrial vertebrates, predominantly in lowland rainforests and montane areas. These hotspots exhibit elevated (regional species pools) due to historical isolation and climatic stability, contrasting with polar and arid zones where (local richness) remains low, often below 100 species per site. At intermediate spatial scales, patterns include elevational diversity peaks at mid-altitudes (typically 1,000–2,500 m) in mountain ecosystems, driven by compressed climatic zones and transitions, as seen in Andean and Himalayan gradients. Vertical stratification within forests amplifies local diversity, with canopy, , and soil layers hosting distinct assemblages; for example, arboreal arthropods in temperate forests show layered turnover exceeding 50% between strata. , reflecting species turnover across landscapes, increases with topographic heterogeneity, as evidenced by higher compositional variation in rugged terrains compared to flat biomes like savannas. Islands and archipelagos display idiosyncratic patterns, with diversity scaling positively with area but modulated by isolation, yielding endemism hotspots like .

Factors Influencing Biodiversity

Biodiversity in terrestrial ecosystems, encompassing , evenness, and functional , is shaped by a combination of abiotic environmental factors, resource availability, and interactions. Empirical studies consistently identify —particularly , , and their —as primary drivers, with species richness gradients often peaking in tropical regions due to higher input and water availability supporting greater niche partitioning. For instance, analyses of global reveal that annual and diurnal range explain significant variance in species richness, with equatorial zones exhibiting up to 10-20 times higher than polar regions owing to reduced seasonal extremes and extended growing periods. Habitat heterogeneity further amplifies by creating varied microhabitats that facilitate species coexistence through niche . Topographic features like gradients and variability increase available niche space, as evidenced by meta-analyses showing that ecosystems with high environmental heterogeneity, such as montane forests, sustain 30-50% greater compared to uniform lowlands by buffering against competitive exclusion. , often proxied by net primary production (NPP), correlates positively with ; higher NPP from nutrient-rich soils and favorable climates supports more trophic levels and specialist , with grassland studies indicating that a 1 Mg/ha increase in NPP can associate with 10-15% higher plant via enhanced resource heterogeneity. Disturbance regimes, including fire, herbivory, and , influence via the , where moderate frequencies prevent dominance by superior competitors, maintaining diversity peaks; for example, ecosystems with biennial fires exhibit 20-40% higher grass and herb richness than fire-suppressed or annually burned sites. Evolutionary history and also play causal roles, as older, stable landmasses like Gondwanan fragments host relict species assemblages, with isolation metrics explaining up to 25% of variance in endemic richness across archipelago-like habitats. Biotic factors, such as predation and , modulate these drivers but are secondary to abiotic controls in broad-scale patterns, as demonstrated by experiments where gradients override strengths in structuring communities.

Human Influences

Direct Modifications and Exploitation

constitutes the dominant form of direct human modification to terrestrial ecosystems, occupying roughly half of the world's habitable land surface and converting vast areas of natural habitats into croplands and pastures. Over the past 50 years, approximately 65% of agricultural land-use change has been driven by rising demand for animal products, intensifying pressure on grasslands and forests. This conversion disrupts native vegetation, soil structures, and wildlife habitats, with cropland expansion alone accounting for a significant portion of impacts from land-use shifts between 1995 and 2022. Deforestation represents a primary mechanism, primarily for agricultural clearance, with global rates averaging 10 million hectares annually according to UN FAO estimates. The FAO reports a slowdown to 10.9 million hectares per year from 2015 to 2025, down from 17.6 million in 1990–2000, though tropical regions bear the brunt, with 6.37 million hectares lost in alone. Commercial and commodity-driven clearing, such as for soy and , amplify these losses, fragmenting ecosystems and releasing stored carbon. Livestock grazing, often exceeding sustainable levels, affects rangelands covering about 40% of Earth's land, leading to that reduces by 26%, by 18%, and by 19%. In and grasslands, excessive stocking densities erode soils, diminish forage quality, and alter community dynamics, with studies showing disrupted microbial associations and heightened vulnerability to . Urbanization, while covering a smaller global footprint (less than 3% of land), drives localized habitat loss through expansion, with projections indicating continued growth; for instance, urban built-up areas in select regions expanded from 3.39 km² in 1998 to an estimated 11.01 km² by 2023. This conversion fragments ecosystems, increases , and facilitates ingress, compounding pressures from adjacent agricultural intensification. Resource extraction, including , directly scars landscapes, with activities linked to vegetation loss, , and decline; mining-related occurs predominantly in tropical rainforests, with over 80% concentrated in 10 countries and concessions encroaching on 20% of Indigenous lands as of 2020. These operations fragment habitats and pollute soils and waterways, with long-term ecological recovery often protracted due to heavy metal residues and altered . Wildlife harvesting through and exploits terrestrial , contributing to defaunation and trophic imbalances; unsustainable offtake, alongside habitat loss, drives 73% average declines in monitored populations since 1970, targeting large-bodied and disrupting and predation dynamics. In biodiversity hotspots, indiscriminate reduces functional , with recent studies documenting extraction rates equivalent to thousands of kilograms of annually in protected areas.

Indirect Effects and Feedbacks

Human-induced represents a primary indirect effect on terrestrial ecosystems, manifesting through altered regimes, patterns, and events that propagate globally beyond localized emissions sources. In regions, observed Arctic amplification has increased near-surface air by approximately 3°C since the 1970s, driving phenological shifts such as earlier and extended growing seasons, which disrupt plant-pollinator synchrony and alter carbon uptake dynamics. Similarly, ecosystems experience upward of treelines, with like Pinus uncinata advancing elevations by 1-2 meters per decade in the European , compressing availability for high-elevation endemics and inducing biotic homogenization. Atmospheric deposition of nitrogen and sulfur compounds from industrial and agricultural emissions constitutes another indirect pathway, enriching remote terrestrial soils and favoring nitrophilous species over oligotrophic natives. In alpine tundra, excess nitrogen inputs exceeding 10 kg N ha⁻¹ yr⁻¹ have been linked to reduced plant diversity and increased graminoid dominance, as documented in long-term monitoring at Niwot Ridge, Colorado, where deposition correlates with a 20-30% decline in forb cover since the 1980s. These depositions, transported via prevailing winds, amplify eutrophication effects far from source regions, altering microbial communities and nutrient cycling in ways that diminish ecosystem resilience to further stressors. Feedback mechanisms in human-influenced terrestrial ecosystems often amplify initial perturbations, particularly in -dominated where thawing soils release stored , estimated at 1,300-1,600 Gt, through enhanced microbial . This carbon-climate could contribute up to 0.1-0.2°C of additional by 2100 under moderate emissions scenarios, as decomposes into CO₂ and CH₄, with abrupt thaw features like lakes accelerating CH₄ emissions by factors of 2-5 compared to intact . In alpine settings, warming-induced drying of soils may conversely enhance CH₄ oxidation sinks in some aerobic microsites, partially offsetting emissions, though net release remains positive due to CO₂ dominance. Vegetation structural changes induce albedo feedbacks, where shrub expansion in tundra—observed to cover 10-20% more area since 1980—lowers surface reflectivity from 0.8 (-dominated) to 0.2 (vegetated), absorbing additional solar radiation and exacerbating local warming by 1-2°C. This shrubification, indirectly promoted by warming, insulates via deeper , promoting further degradation and creating self-reinforcing loops that hinder reversion to states even under stabilization scenarios. In ecosystems, analogous feedbacks arise from reduced persistence, shortening the high- period and amplifying heat uptake, with models projecting 20-50% declines in seasonal cover by mid-century. These processes underscore how indirect human forcings, via greenhouse gas accumulation, trigger nonlinear responses that intensify climate-ecosystem interactions at high latitudes and elevations.

Management and Conservation

Strategies for Sustainability

Sustainable land management (SLM) practices, such as terracing, contour farming, and , have been shown to reduce rates by up to 50-90% in various terrestrial ecosystems compared to conventional , thereby preserving and ecosystem productivity over decades. These methods maintain hydrological cycles and retention, with empirical from assessments indicating that SLM on 1.5 billion hectares could avoid further affecting 20% of . However, success depends on local adaptation, as intensive SLM under projected climate warming may reduce ecological multifunctionality in grasslands by 10-20% due to altered patterns. Agroforestry systems, integrating trees with crops or , enhance by 0.2-3.0 tons per annually while improving and water retention in degraded lands. Systematic reviews of trials in low- and middle-income countries demonstrate boosts agricultural yields by 20-40% in rainfed systems through regulation and nutrient cycling, though benefits diminish in water-scarce regions without support. These practices also support by providing corridors, with meta-analyses reporting 15-30% higher in agroforestry plots versus monocultures. Reforestation and ecological restoration target degraded terrestrial areas, yielding average biodiversity gains of 20% and reduced environmental variability across restored sites globally. In forested landscapes, restoring 350 million hectares—equivalent to the size of current global pledges—could sequester 205 gigatons of carbon over 50 years, but only if paired with emissions reductions, as regrown forests absorb CO2 at rates insufficient to offset ongoing combustion. on marginal agricultural lands further mitigates and enhances , with studies showing 10-25% improvements in services like and . Market-based incentives, including payments for services, have expanded on 100 million hectares since 2000, particularly in carbon markets that reward verified . These approaches outperform regulatory mandates in voluntary adoption rates, with evidence from Latin American programs indicating sustained land stewardship where payments exceed 50% of farm income. Nonetheless, critiques highlight risks of leakage, where protected areas displace elsewhere, necessitating integrated monitoring to ensure net gains.

Empirical Outcomes and Critiques

Empirical assessments of protected areas in terrestrial ecosystems indicate variable success in conserving and reducing loss. A global analysis of over 1,700 protected areas found that they reduced rates by an average of 28% compared to unprotected lands between 2000 and 2012, though effectiveness declined in areas with high governance challenges or external pressures like expansion. Systematic reviews confirm that well-managed reserves can mitigate threats such as and , with meta-analyses showing positive but context-dependent impacts on occupancy rates, particularly in forests where declines were slowed by 10-20% in networked systems. However, coverage expansions under targets like the 17% terrestrial protection goal have included many low-quality sites, yielding minimal gains where enforcement is weak. Restoration initiatives, including and natural regeneration, demonstrate moderate enhancements but face high failure rates. Meta-analyses of global projects report an average 20% increase in and reduced variability in degraded terrestrial sites, with natural regeneration outperforming active planting by 34-56% in metrics. In efforts, empirical data from Asia-Pacific projects reveal 44% average tree mortality within five years, attributed to poor , , and herbivory, though survival improves with species-matched planting and preparation. Long-term studies of large-scale programs, such as China's Grain for Green, show gross primary productivity gains in restored forests but uneven outcomes across climatic zones, with arid regions lagging due to limitations. Critiques highlight systemic shortcomings in conservation outcomes, including overreliance on simplistic metrics like extent without verifying ecological integrity or socioeconomic viability. Peer-reviewed syntheses note that while interventions like control and sustainable logging reduce short-term threats, long-term persistence requires addressing underlying drivers such as land-use intensification, with many projects failing to halt or population declines. Unintended consequences, including displacement of communities from reserves, have led to conflicts and reduced local , as documented in cases where exclusionary policies eroded traditional practices without livelihoods, exacerbating in 20-30% of affected areas. Furthermore, carbon-focused strategies in forests often prioritize sequestration over multifunctional ecosystems, critiqued for undervaluing non-carbon services like habitat connectivity and risking plantations that diminish native diversity. These gaps underscore the need for adaptive, evidence-based approaches integrating local incentives, as current frameworks frequently overlook causal feedbacks like rebound effects from displaced pressures.

Debates and Controversies

Climate Change Attribution

Attribution studies in terrestrial ecosystems seek to distinguish changes driven by anthropogenic climate forcings—primarily —from those arising from natural variability, such as solar cycles, volcanic activity, and ocean-atmosphere oscillations like the El Niño-Southern Oscillation or . These analyses typically employ statistical detection methods or process-based models to compare observed trends against counterfactual scenarios without human influence, revealing fingerprints in metrics like greenness, , and distributions. For instance, satellite-derived (NDVI) data indicate shifts in vegetation activity, with enhanced productivity in high-latitude boreal forests attributable to warming-induced thaw and longer growing seasons. A dominant observed change is global greening, evidenced by a 10% increase in vegetation cover between 2000 and 2020, as measured by MODIS . This phenomenon, spanning forests, croplands, and grasslands, is predominantly linked to CO2 fertilization, where elevated atmospheric concentrations enhance and water-use efficiency in plants, explaining roughly 70% of the effect from 1982 to 2015 according to analyses of AVHRR and MODIS data. While temperature and deposition contribute marginally, variability and explain the rest, underscoring that anthropogenic CO2—a key forcing—has yielded empirically verifiable benefits to primary , countering expectations of uniform degradation. Evidence for biome shifts includes poleward or upslope migrations in some taxa, with field and data documenting altered treelines and encroachment in regions, consistent with modeled responses to 1-2°C warming since pre-industrial times. However, comprehensive reviews of over 10,000 species records find that only 46.6% of range shifts align with climate-driven predictions, such as toward higher latitudes or elevations, while others reflect dispersal limitations, , or interactions rather than temperature alone. NDVI trends further show significant increases (0.05 units on average) across most s from 1982 to 2020, suggesting or rather than wholesale reorganization. Uncertainties in attribution arise from confounding factors, including direct land-use changes like and , which often exceed climate signals in magnitude for regional . Model-dependent approaches, reliant on system models with documented biases in simulating feedbacks and regional , may overestimate climate attribution by underweighting internal variability; for example, multidecadal droughts in the American Southwest correlate more strongly with Pacific patterns than . Critiques emphasize that rapid-attribution frameworks, such as those from , frequently require ensembles of imperfect models and can understate natural forcings, leading to overstated influence in event-specific stresses like wildfires or outbreaks. Empirical hindcasts reveal that pre-20th century biome fluctuations, driven by minima and volcanic aerosols, mirror modern variability, cautioning against exclusive blame on recent trends.

Biodiversity Decline Narratives

The dominant narrative on terrestrial biodiversity decline posits a catastrophic loss, often framed as the onset of a sixth mass , with assessments claiming that up to one million face extinction risk primarily due to , , and . This perspective, advanced by reports like the 2019 IPBES Global Assessment, emphasizes empirical indicators such as population declines in vertebrates (e.g., a 68% average drop in monitored populations from 1970 to 2016 per data integrated into such syntheses) and habitat conversion, projecting irreversible tipping points without immediate policy interventions. However, these narratives frequently rely on modeled projections rather than comprehensive empirical tallies of actual , which remain low: the documents fewer than 1,000 verified species extinctions since 1500 across all taxa, far below predictions of rates 1,000 times the background. Critiques highlight that such narratives overstate global trends by extrapolating local or taxonomic-specific declines (e.g., in or amphibians) to the entire , ignoring countervailing evidence of stability or recovery in terrestrial systems. Peer-reviewed analyses indicate that documented rates have slowed across many and animal groups since the mid-20th century, with weak correlations between historical extinctions and current threats in diverse taxa, challenging assumptions of uniform acceleration. For instance, global forest cover trends show net losses decelerating to 4.12 million hectares annually from 2015 to 2025, down from 10.7 million in the , partly due to agricultural intensification sparing natural lands and in temperate zones, which supports rebound in some regions despite tropical pressures. Overall 20th-century terrestrial decline, as measured by multi-indicator models, ranges from 2% to 11%, nuanced by increases in certain metrics like contributions to local richness. These narratives also face scrutiny for methodological biases, including reliance on non-random sampling and failure to account for under-detection of persistence, leading to inflated threat assessments that serve institutional incentives in and NGOs amid funding dependencies. Empirical demands distinguishing abundance reductions (verifiable in farmlands or fragmented habitats) from species-level extinctions, as the former often reflect adaptive shifts rather than holistic collapse; for example, less than 0.1% of known terrestrial have demonstrably gone extinct in the last 500 years, with projections of future losses limited by successes and . While localized terrestrial declines—such as in primary tropical forests—remain causally linked to land-use changes, global narratives risk causal overreach by downplaying human-induced gains, like expansions covering 17% of terrestrial lands by 2020, which have stabilized populations in targeted taxa. This meta-awareness underscores the need for first-principles validation against over synthesized from potentially biased syntheses.

Land Use Trade-offs

Land use trade-offs in terrestrial ecosystems arise from competing demands for , , , and , often pitting provisioning services like food production against regulating services such as maintenance and integrity. Globally, spanned 4,800 million hectares in 2023, comprising over one-third of terrestrial surface area and driving the majority of habitat conversion. Over 90% of losses linked to land-use change originate from , with cropland responsible for 72% and pastures for 21% of these impacts between 1995 and 2022. These conversions frequently yield short-term gains in productivity but long-term declines in resilience, as intensified uses reduce habitat heterogeneity and . The agriculture-biodiversity nexus exemplifies these tensions, framed in the land sparing versus land sharing debate. Land sparing advocates high-yield farming on minimal footprints to allocate remaining areas to protected s, while land sharing integrates low-intensity, diversified practices to foster on-farm . Empirical assessments, however, reveal no dominant strategy: a 2025 of available data concluded that evidence remains insufficient to generalize either approach for resolving trade-offs, with outcomes varying by type, taxa, and regional context. For instance, intensification may preserve total habitat area but erode local through homogenized landscapes and chemical inputs, whereas extensive sharing sustains fewer overall yields, potentially necessitating broader conversion elsewhere. Contextual modeling underscores these dynamics. In China's , sustainable intensification projections for 2020–2040 forecast 15% higher crop alongside 8% drops in , water , , and , contrasting with scenarios that cut and outputs by 15–20% but boost those services by 15–50%. Similarly, analyses across global ecoregions indicate intensification harms more than farmland expansion in up to 71% of cases where natural habitat is scarce, as gains fail to offset on-site ecological degradation from monocultures and inputs. Trade-offs intensify with non-food demands; cropland growth, for example, has amplified without commensurate safeguards, amplifying net losses. Urbanization introduces further frictions, converting prime arable or forested land into impervious surfaces that fragment ecosystems and elevate , with studies showing disproportionate per-hectare declines compared to agricultural shifts. Forestry-extraction conflicts similarly pit timber yields against carbon storage and wildlife corridors, where selective logging may sustain some diversity but often cascades into full conversion under market pressures. Empirical data emphasize that severity hinges on baseline , availability, and management precision—high-diversity regions benefit more from sparing, while marginal lands favor sharing—rejecting one-size-fits-all prescriptions in favor of spatially explicit assessments. Policies ignoring these nuances risk amplifying unintended harms, as evidenced by historical intensification waves that spared gross land but intensified local extinctions.

References

  1. [1]
    Terrestrial Biomes | Learn Science at Scitable - Nature
    Tropical Forest Biomes · Savanna Biomes · Desert Biomes · Grassland Biomes · Temperate Deciduous Forest Biome · Mediterranean Climate Biomes · Northern Coniferous ...
  2. [2]
    Advancing the EcoVeg approach as a terrestrial ecosystem typology ...
    May 12, 2025 · Vegetation is a critical component of terrestrial ecosystems, given its role in energy capture, biomass production, nutrient and water cycling, ...
  3. [3]
    [PDF] The Terrestrial Carbon Sink - Harvard Forest
    Sep 26, 2018 · The terrestrial carbon sink is the accumulation of carbon on the land surface, removing 3.61 Pg C year−1 from the atmosphere.<|control11|><|separator|>
  4. [4]
    Biodiversity loss reduces global terrestrial carbon storage - Nature
    May 22, 2024 · Natural ecosystems store large amounts of carbon globally, as organisms absorb carbon from the atmosphere to build large, long-lasting, ...
  5. [5]
    The three major axes of terrestrial ecosystem function - PMC
    Sep 22, 2021 · Here, we identify and quantity the major axes of terrestrial ecosystem functions and sources of variation along these axes. ... definition ...
  6. [6]
    Terrestrial ecology - Soil Ecology Wiki
    May 2, 2025 · Terrestrial ecosystems can be defined as a community of interaction between many living organisms and nonliving things on land.Terrestrial Ecosystems vs... · Factors of Terrestrial Ecosystems · Biotic Factors
  7. [7]
    Terrestrial Ecosystem - an overview | ScienceDirect Topics
    The terrestrial ecosystem is a key component in the global carbon cycle, as the repository of a significant carbon stock, as the source and sink of major fluxes ...
  8. [8]
    Terrestrial Ecology - Environmental System Science Program
    Terrestrial ecology research supported by the Environmental System Science (ESS) program seeks to improve the representation of terrestrial ecosystem ...
  9. [9]
    Terrestrial Ecosystem - an overview | ScienceDirect Topics
    Terrestrial ecosystems have some characteristics different from aquatic ecosystems—for example, the dominance of the detritus chain among grazing, detritus, ...
  10. [10]
    Terrestrial Ecosystem - National Geographic Education
    A terrestrial ecosystem is a land-based community of organisms and the interactions of biotic and abiotic components in a given area.
  11. [11]
    Large bacterial population colonized land 2.75 billion years ago
    Sep 23, 2012 · There is evidence that some microbial life had migrated from the Earth's oceans to land by 2.75 billion years ago, though many scientists ...
  12. [12]
    The colonization of land by animals: molecular phylogeny and ...
    Jan 19, 2004 · Prokaryotes were probably the first organisms to colonize land, and this occurred as early as 2.6 billion years ago [1–3].
  13. [13]
    First plants cooled the Ordovician | Nature Geoscience
    Feb 1, 2012 · Fossilized plant spores found in ∼470 Ma sediments indicate that non-vascular land plants were growing in damp environments on continental ...
  14. [14]
    The timescale of early land plant evolution - PNAS
    Feb 20, 2018 · Theories on the process of terrestrialization have long argued for a close temporal relationship between the emergence of land plants and ...
  15. [15]
    Study: First Land Plants Appeared 500 Million Years Ago | Sci.News
    Feb 19, 2018 · “Our results show the ancestor of land plants was alive in the middle Cambrian period, which was similar to the age for the first known ...
  16. [16]
    A timeline for terrestrialization: consequences for the carbon cycle in ...
    Land plants originated in the latter part of the Silurian Period (428–423 Ma), which was followed by rapid diversification during the Devonian Period (416–359 ...
  17. [17]
    Evolution: Out of the Ocean - ScienceDirect.com
    Mar 18, 2013 · The first animals to arrive on land were the myriapods, the centipedes and millipedes. Direct fossil evidence for key events is indicated with ...<|separator|>
  18. [18]
    Seedless Plants – Introductory Biology: Evolutionary and Ecological ...
    The first fossils that show the presence of vascular tissue are dated to the Silurian period, about 430 million years ago. The simplest arrangement of ...Missing: date | Show results with:date
  19. [19]
    The origin and early evolution of vascular plant shoots and leaves
    Dec 18, 2017 · This review discusses fossil, developmental and genetic evidence relating to the evolution of vascular plant shoots and leaves in a phylogenetic framework.
  20. [20]
    Plants evolved complexity in two bursts – with a 250-million-year ...
    Sep 18, 2021 · When land plants first diversified in the early Devonian about 420 million to 360 million years ago, Earth was a warmer world devoid of trees or ...
  21. [21]
    Seed plants: Fossil Record
    The earliest seeds appear in the Late Devonian. The oldest known seed plant is Elkinsia polymorpha, a "seed fern" from Late Devonian (Famennian) of West ...<|separator|>
  22. [22]
    Evolution of Seed Plants - OERTX
    Seed plants evolved from bryophytes, with early seed plants around 350 million years ago. Seeds and pollen enabled non-water reproduction. Gymnosperms ...
  23. [23]
    Early terrestrial arthropods: a fragmentary record - Journals
    The earliest unequivocal terrestrial fossils are uppermost Silurian (Přídolí) myriapods, presumed to be pioneer decomposers.
  24. [24]
    Insect Flight: State of the Field and Future Directions - PMC
    Approximately 400 million years ago, an evolutionary innovation arose that fundamentally altered the history of life. A simple set of protowings evolved in a ...
  25. [25]
    The origin of tetrapods - Understanding Evolution - UC Berkeley
    When we get past coelacanths and lungfishes on the evogram, we find a series of fossil forms that lived between about 390 and 360 million years ago during the ...Missing: colonization | Show results with:colonization
  26. [26]
    Rise of the Earliest Tetrapods: An Early Devonian Origin from ... - NIH
    Tetrapods are now generally considered to have colonized land during the Carboniferous (i.e., after 359 MYA), which is considered to be one of the major events ...
  27. [27]
    The ancestral flower of angiosperms and its early diversification - NIH
    Aug 1, 2017 · The most recent common ancestor of all living angiosperms likely existed ∼140–250 million years ago,,. In contrast, the most recent common ...
  28. [28]
    Uncertainty in the timing of diversification of flowering plants rests ...
    May 28, 2025 · Fossils of unequivocal crown-angiosperms are not known from before the Cretaceous, and yet molecular estimates range from the Late Jurassic to the Permian.
  29. [29]
    Ecosystem Ecology
    Producers. have the ability to make new, complex, organic material from the atoms in their environment · Consumers. Primary consumers - herbivores (Consumers ...
  30. [30]
    Producers - National Geographic Education
    Oct 31, 2023 · Producers convert water, carbon dioxide, minerals, and sunlight into the organic molecules that are the foundation of all life on Earth.
  31. [31]
    Decomposers - National Geographic Education
    Oct 19, 2023 · Decomposers play a critical role in the flow of energy through an ecosystem. They break apart dead organisms into simpler inorganic materials.
  32. [32]
    Food chains & food webs (article) | Ecology - Khan Academy
    Fungi and bacteria are the key decomposers in many ecosystems; they use the chemical energy in dead matter and wastes to fuel their metabolic processes.
  33. [33]
    Global change and species interactions in terrestrial ecosystems
    Global change drivers alter species interactions, including competition, food webs, pathogen infection, mutualisms, herbivory, and predation, with variable ...
  34. [34]
    The links between ecosystem multifunctionality and above - Nature
    Sep 2, 2015 · Plant biodiversity is often correlated with ecosystem functioning in terrestrial ecosystems. However, we know little about the relative and ...
  35. [35]
    Biotic and abiotic factors predicting the global distribution ... - Nature
    Mar 9, 2017 · Abiotic factors, such as temperature and precipitation, are consistently found to be primary determinants of species distributions at broad ...
  36. [36]
    Abiotic Factors - National Geographic Education
    An abiotic factor is a non-living part of an ecosystem that shapes its environment. In a terrestrial ecosystem, examples might include temperature, light, and ...
  37. [37]
    35.2 Biogeography | Texas Gateway
    Abiotic factors such as temperature and rainfall vary based mainly on latitude and elevation. As these abiotic factors change, the composition of plant and ...
  38. [38]
    Relative importance of soil properties and microbial community for ...
    Apr 25, 2016 · Random forest analyses showed that both soil properties (i.e. total C and pH) and microbial abundance determined broad functioning (i.e. soil ...
  39. [39]
    Biotic and abiotic factors affecting soil microbial carbon use efficiency
    Nov 15, 2024 · We reviewed the effects of various abiotic factors, such as temperature, soil moisture, pH, nutrient addition, and substrate type, and biotic factors.
  40. [40]
    Biotic and abiotic factors determine species diversity–productivity ...
    Several abiotic factors relating to climate (air temperature and precipitation) and soil chemistry (pH, organic carbon concentration, total nitrogen ...<|separator|>
  41. [41]
    Effects of topography on tropical forest structure depend on climate ...
    Jul 27, 2019 · Topography affects abiotic conditions which can influence the structure, function and dynamics of ecological communities.<|control11|><|separator|>
  42. [42]
    Biotic and abiotic influences on wind disturbance in forests of NW ...
    Jun 30, 2007 · Abiotic features affecting the degree of wind damage include wind speed (Elie and Ruel, 2005), topography (Gardiner and Quine, 2000), and soils ...Missing: terrestrial | Show results with:terrestrial
  43. [43]
    Environmental conditions are the dominant factor influencing ...
    May 31, 2023 · Abiotic factors such as climate and soil conditions were found to play a more important role compared to biotic factors, directly influencing ...
  44. [44]
    Forest Ecosystem Management: An environmental necessity, but is it ...
    Forest ecosystems are areas of the landscape that are dominated by trees and consist of biologically integrated communities of plants, animals and microbes, ...
  45. [45]
  46. [46]
    [PDF] The State of the World's Forests 2024
    * FAO defines “forest” as land spanning more than 0.5 ha with trees higher ... Almost 75 percent of the world's total land area, particularly forests ...
  47. [47]
    The State of the World's Forests
    The State of the World's Forests 2024. Forest-sector innovations towards a more sustainable future. This edition provides highlights on the state of the ...
  48. [48]
    Four Elements of a Healthy Forest | U.S. Fish & Wildlife Service
    Oct 11, 2022 · Different animals and plants need different forest features to thrive. Forest features change depending on elevation, climate and access to ...
  49. [49]
    Home | Global Forest Resources Assessments
    FAO Global Forest Resources Assessment (FRA) provides essential information for understanding the extent of forest resources, their condition, ...FRA 2015 · Mangrove protection · Past Assessments · About
  50. [50]
    Forest biodiversity, relationships to structural and functional ...
    Mar 7, 2018 · We conclude that forest structure (height) and resources (biomass) are more likely foundational characteristics supporting biodiversity.
  51. [51]
    Nutrient cycling in forests - PubMed
    Studies of nutrient cycling in forests span more than 100 yr. In earlier years, most attention was given to the measurement of the pools of nutrients in plants ...Missing: sequestration | Show results with:sequestration
  52. [52]
    Earth's biodiversity depends on the world's forests - UNEP-WCMC
    Forests contain 60,000 different tree species, 80 percent of amphibian species, 75 percent of bird species, and 68 percent of the world's mammal species. The ...
  53. [53]
    Goal 15: Forests, desertification and biodiversity - UN.org.
    Forests cover nearly 31 per cent of the world and are home to more than 80 per cent of all terrestrial species of animals, plants and insects. However, ...
  54. [54]
    Tropical forests are home to over half of the world's vertebrate species
    Oct 7, 2021 · We determined that tropical forests harbor 62% of global terrestrial vertebrate species, more than twice the number found in any other terrestrial biome on ...
  55. [55]
    Improving forest ecosystem functions by optimizing tree species ...
    Jul 9, 2025 · Promoting carbon sequestration in forests has a strong potential to mitigate increasing atmospheric carbon concentration and thereby climate ...
  56. [56]
    State of the World's Forests 2020
    Forests cover 31 percent of the global land area. Approximately half the forest area is relatively intact, and more than one-third is primary forest.
  57. [57]
    [PDF] Forest Ecology and Management
    Aug 19, 2023 · Mature and old-growth forests are valued for biodiversity, carbon sequestration, habitat, hydrologic function,.
  58. [58]
    Ecological Condition | US EPA
    Jul 25, 2025 · Many ecosystems are defined based on their predominant species (e.g., forested ecosystem) or physical characteristics (e.g., stream ecosystem) ...
  59. [59]
    [PDF] Grassland Ecology - Kansas State University
    One defining feature of grasslands is that they are dominated or codominated by graminoid vegetation, including the true grasses (family. Poaceae) and other ...
  60. [60]
    World Biomes: Grassland - KDE Santa Barbara
    Temperate grasslands have some of the darkest, richest soils in the world (not in wealth, but in nutrients). People who live in grassland regions often use ...
  61. [61]
    Grassland: Mission: Biomes
    Grasslands are generally open and continuous, fairly flat areas of grass. They are often located between temperate forests at high latitudes and deserts at ...
  62. [62]
    Shrubland - an overview | ScienceDirect Topics
    Shrubland is land cover dominated by shrubs, generally under 5m tall, with a single canopy layer, and often has grass cover.
  63. [63]
    Shrubland: Mission: Biomes - NASA Earth Observatory
    Shrublands are in west coastal regions, with hot, dry summers and cool, moist winters. They have 200-1000mm of rain, and are made up of shrubs and short trees.
  64. [64]
    Shrubland ecosystems: Importance, distinguishing characteristics ...
    Shrubland ecosystems are in stressed habitats, often with drought, poor soils, and fire. Examples include sagebrush and chaparral. They have human and wildlife ...
  65. [65]
    Grasslands, Shrublands and Savannahs
    Grasslands, shrublands and savannahs cover approximately half of the world's terrestrial surface. Distributed from Eurasia and Patagonia to Africa and ...
  66. [66]
    [PDF] Grassland Ecosystems | Sala Lab - Arizona State University
    Grassland ecosystems occur in areas of the world that have an annual precipitation between 150 and 1200 mm and mean annual temperature between 0 and 25 1C ( ...
  67. [67]
    The grassland biome
    Grasslands are characterized as lands dominated by grasses rather than large shrubs or trees. In the Miocene and Pliocene Epochs, which spanned a period of ...
  68. [68]
    Grasslands - Natural Resources Institute
    Grasslands have deep roots that enrich soil health and improve water quality. Very little remains of the native prairies, savannas, and grassland birds of the ...
  69. [69]
    Grassland biodiversity - ScienceDirect.com
    Oct 11, 2021 · Grasslands are among the most species-rich habitats on Earth and may even dwarf tropical rainforests in terms of species diversity. Especially ...
  70. [70]
    Recent advances in understanding grasslands - PMC - NIH
    Aug 30, 2018 · Grasslands are typically dominated by grasses ( Poaceae) and other grass-like plants. Some grasslands occur naturally, whilst others must be ...
  71. [71]
    Deserts - Tulane University
    Nov 17, 2015 · Deserts are areas with less than 250 mm of rainfall/year, or where evaporation exceeds precipitation, and are arid. They can be hot or cold.
  72. [72]
    Desert - National Geographic Education
    Aug 5, 2025 · Plants and animals adapt to desert habitats in many ways. Desert plants grow far apart, allowing them to obtain as much water around them as ...
  73. [73]
    The desert biome - University of California Museum of Paleontology
    There are four major types of deserts: Hot and dry; Semiarid; Coastal; Cold. Hot ... Evaporation rates regularly exceed rainfall rates. Sometimes rain ...
  74. [74]
    Plant Adaptations - Teachers (U.S. National Park Service)
    Nov 16, 2022 · Desert plants are adapted to their arid environment in many ways. Small leaves on desert plants help reduce moisture loss during photosynthesis.
  75. [75]
    Burrowing detritivores regulate nutrient cycling in a desert ecosystem
    Oct 30, 2019 · We conclude that burrowing macro-detritivores are important regulators of litter cycling in this arid ecosystem, providing a plausible general mechanism.Missing: impacts | Show results with:impacts
  76. [76]
    The tundra biome - University of California Museum of Paleontology
    Characteristics of tundra include: Extremely cold climate; Low biotic diversity; Simple vegetation structure; Limitation of drainage; Short season of growth and ...
  77. [77]
    Tundra: Mission: Biomes
    The tundra is the coldest of the biomes. It also receives low amounts of precipitation, making the tundra similar to a desert.
  78. [78]
    Alpine Tundra Ecosystem - Rocky Mountain National Park (U.S. ...
    Jul 22, 2020 · Tundra is a biome, or type of environment, which is characterized as treeless, cold, and relatively dry. Across the globe, there are two types ...
  79. [79]
    Tundra Biome - National Geographic Education
    They have long, cold winters with high winds and average temperatures below freezing for six to ten months of the year. On average, only six to ten weeks of the ...
  80. [80]
    13.6: Tundra Biome - Geosciences LibreTexts
    May 24, 2024 · Short grasses, flowers, and grass-like sedges, along with covers of mosses and lichens are the dominate forms of vegetation in the tundra.
  81. [81]
    Tundra Description - Alaska Department of Fish and Game
    Though treeless and often bitter cold, Arctic tundra is an ecosystem of great beauty and abundance, shaped by the dramatic seasons of the far north.
  82. [82]
    Safeguarding Arctic biodiversity - Arctic Council
    The Arctic is home to more than 21000 known species of highly cold-adapted mammals, birds, fish, invertebrates, plants and fungi and microbes.
  83. [83]
    Tundra Ecology - Alaska Department of Fish and Game
    Many tundra animals grow more slowly, and reproduce less frequently, than do their non-tundra relatives. Tundra-dwelling lake trout may take ten years to reach ...Missing: biodiversity | Show results with:biodiversity
  84. [84]
    Tundra - Arctic, Flora, Fauna - Britannica
    Aug 29, 2025 · Many tundra species cannot be found elsewhere, and thus the biome is an important contributor to global biodiversity despite its low species ...
  85. [85]
    Energy Transfer in Ecosystems - National Geographic Education
    Jan 22, 2024 · On average, only about 10 percent of energy stored as biomass in a trophic level is passed from one level to the next. This is known as “the 10 ...
  86. [86]
    The Flow of Energy from Primary Production to Higher Tropic Levels
    Only a fraction of the energy available at one trophic level is transferred to the next trophic level. The rule of thumb is 10%, but this is very approximate.
  87. [87]
    Terrestrial Food Webs | Smithsonian Environmental Research Center
    The terrestrial food web links creatures on land, from the tiniest microbes in the soil to the large mammals of the forests.Missing: structure | Show results with:structure
  88. [88]
    Plant communities and food webs - Frontiers
    Oct 29, 2023 · In savanna ecosystems, vertebrate herbivores can quickly be divided into browsers and grazers, these will in turn divide the habitat into open ...
  89. [89]
    19.1: Introduction to and Components of Food Webs
    Nov 21, 2022 · The three basic ways in which organisms get food are as producers, consumers, and decomposers. Producers (autotrophs) are typically plants or ...
  90. [90]
    Food Web - National Geographic Education
    Oct 19, 2023 · A food web consists of all the food chains in a single ecosystem. Grades 3 - 12+ Subjects Biology, Ecology
  91. [91]
    Food Web: Concept and Applications | Learn Science at Scitable
    Food web offers an important tool for investigating the ecological interactions that define energy flows and predator-prey relationship (Cain et al. 2008).
  92. [92]
    [PDF] The Trophic-Dynamic Aspect of Ecology Raymond L. Lindeman ...
    Sep 24, 2007 · The basic process in trophic dynamics is the transfer of energy from one part of the ecosystem to another. All function, and indeed all life, ...Missing: name= query">
  93. [93]
    Energy Flow and the 10 Percent Rule
    Oct 19, 2023 · On average only 10 percent of energy available at one trophic level is passed on to the next. This is known as the 10 percent rule.
  94. [94]
    Energy flow & primary productivity (article) - Khan Academy
    Energy transfer between trophic levels is not very efficient. Only ‍ of the net productivity of one level ends up as net productivity at the next level. ...
  95. [95]
    Linking human impacts to community processes in terrestrial and ...
    Focusing on the community processes of dispersal, speciation, selection and drift can provide mechanistic insights into how humans alter the dynamics of ...
  96. [96]
    Community structure and trophic level interactions in the terrestrial ...
    Dec 29, 2017 · This paper aims to review and discuss how community structures are organized and trophic level interactions function and shape the biological communities in ...
  97. [97]
    Indirect interactions in terrestrial plant communities: emerging ...
    Jun 25, 2015 · Here, we review the literature on indirect interactions among plants published since 1990, using a novel synthetic framework that accounts for ...
  98. [98]
    (PDF) Global change and terrestrial plant community dynamics
    Aug 6, 2025 · PDF | Significance Global terrestrial vegetation plays a critical role in biogeochemical cycles and provides important ecosystem services.
  99. [99]
    Succession: A Closer Look | Learn Science at Scitable - Nature
    Succession refers to a directional, predictable change in community structure over time (Grime 1979, Huston & Smith 1987).
  100. [100]
    Differences between primary and secondary plant succession ...
    Sep 25, 2018 · Success of both types of succession (a return to target vegetation) differed significantly among biomes, with more likely success in cold than ...
  101. [101]
    Ecological succession in a changing world - Wiley Online Library
    Feb 13, 2019 · By explicitly comparing primary and secondary succession across a broad range of ecosystems in the published literature, Prach and Walker (2019 ...
  102. [102]
    [PDF] Forests, Competition and Succession' - Oregon State University
    For example, in the black spruce forests common to the interior of Alaska, burned sites are quickly occupied by sprouting grasses, shrubs, and small trees ( ...<|separator|>
  103. [103]
    [PDF] and mid-seral seeded grassland compared to shortgrass steppe
    Grazing effects on plant community succession of early- and mid-seral seeded grassland compared to shortgrass steppe. Daniel G. Milchunas & Mark W. Vandever.
  104. [104]
    Climate vulnerability of Earth's terrestrial biomes | Scientific Reports
    Sep 26, 2025 · We show that soil and climate variables can account for 90% of current BCE distributions, providing an accurate classification of Earth's ...
  105. [105]
    Integrating succession and community assembly perspectives - NIH
    Sep 12, 2016 · Classic examples of threshold dynamics in succession are found in the transition from grasslands to shrub-/woodland-dominated ecosystems driven ...
  106. [106]
    Explanations for latitudinal diversity gradients must invoke rate ...
    Aug 3, 2023 · The latitudinal diversity gradient (LDG) describes the pattern of increasing numbers of species from the poles to the equator.
  107. [107]
    The latitudinal diversity gradient of tetrapods across the Permo ...
    Jun 17, 2020 · Reconstructed terrestrial LDGs contrast strongly with the generally unimodal gradients of today, potentially reflecting high global temperatures ...
  108. [108]
    Strong evidence for latitudinal diversity gradient in mosses across ...
    Moss species richness decreases strongly with increasing latitude, showing a latitudinal diversity gradient, with a correlation of -0.99.
  109. [109]
    Comprehensive update to the world's biodiversity hotspots project ...
    Oct 11, 2025 · There are currently 36 terrestrial hotspots, which cover 16.7% of Earth's land surface. Wallacea, which includes more than 1,680 diverse ...
  110. [110]
    Biodiversity Hotspots - Conservation International
    Around the world, 36 areas qualify as hotspots. Their intact habitats represent just 2.5% of Earth's land surface, but they support more than half of the ...
  111. [111]
    Spatial Patterns of Species Biodiversity in Terrestrial Environments
    Jun 27, 2017 · Spatial patterns of biodiversity provide core knowledge for conservation, and understanding the mechanisms for the assembly and maintenance of species co- ...
  112. [112]
    Towards an Understanding of Large-Scale Biodiversity Patterns on ...
    Feb 21, 2023 · This review presents a recent theory called the 'macroecological theory on the arrangement of life' (METAL). METAL proposes that biodiversity is strongly ...
  113. [113]
    Ecological patterns and processes in the vertical dimension of ...
    Feb 3, 2023 · In this review, we explore the ecological evidence that the vertical gradient is an influential engine driving the ecology and evolution of forest species.
  114. [114]
    Common species contribute little to spatial patterns of functional ...
    Feb 14, 2022 · Spatial patterns of functional diversity are important in understanding community assembly as well as spatial variation in ecosystem ...
  115. [115]
    The spatial patterns of diversity and their relationships with ...
    This study analyzed the spatial patterns of diversity and their relationships with environments in rhizosphere microorganisms and host plants along elevations ...
  116. [116]
    Global patterns and drivers of species and genera richness of ...
    Jul 20, 2025 · Species richness was primarily influenced by elevation range (ELR), temperature diurnal range (Tdur), precipitation seasonality (Pseas), annual ...
  117. [117]
    The landscape ecological view of vertebrate species richness in ...
    Oct 3, 2023 · The environmental factors in explaining biodiversity patterns mainly fall into three categories: energy availability, water availability, and ...
  118. [118]
    Environmental heterogeneity as a universal driver of ... - PubMed
    Environmental heterogeneity is regarded as one of the most important factors governing species richness gradients. An increase in available niche space, ...Missing: scientific review
  119. [119]
    Plant species richness and ecosystem multifunctionality in global ...
    Experiments suggest that biodiversity enhances the ability of ecosystems to maintain multiple functions, such as carbon storage, productivity, ...
  120. [120]
    Effects of biodiversity on ecosystem functioning: a consensus of ...
    Species effects act in concert with the effects of climate, resource availability, and disturbance regimes in influencing ecosystem properties. Human activities ...
  121. [121]
    Insights on biodiversity drivers to predict species richness in tropical ...
    The importance of different biodiversity drivers, such as climate, historical factors, or human impact, is another important challenge for biodiversity ...
  122. [122]
    Land Use - Our World in Data
    Half of the world's habitable land is used for agriculture. The extensive land use of agriculture has a major impact on the Earth's environment as it reduces ...Missing: urbanization | Show results with:urbanization
  123. [123]
    The emerging global crisis of land use | 02 The state of the world's ...
    Nov 22, 2023 · Over the past 50 years, some 65 per cent of agricultural land-use change has been driven by increased demand for animal products.
  124. [124]
    Biodiversity impacts of recent land-use change driven by increases ...
    Sep 20, 2024 · This study examines the link between biodiversity impacts from land-use change and shifts in global supply chains from 1995 to 2022
  125. [125]
    Deforestation and Forest Loss - Our World in Data
    The UN FAO estimates that 10 million hectares of forest are cut down each year. This interactive map shows deforestation rates across the world. Read more about ...
  126. [126]
  127. [127]
    [PDF] 2024ForestDeclarationAssessme...
    Oct 6, 2024 · We present core indicators – such as estimates of deforestation,c forest degradation, and area under restoration – that have corresponding 2030.
  128. [128]
    (PDF) The global deforestation footprint of agriculture and forestry
    Aug 8, 2025 · Deforestation footprinting attributes forest loss to commodity production and consumption, identifying global trends, drivers and hot spots to ...
  129. [129]
    Overgrazing is threatening global drylands, study finds
    Nov 25, 2022 · Grazing is a trillion-dollar industry, and is particularly important in drylands, which cover about 40 percent of Earth's land surface and ...
  130. [130]
    Global effects of livestock grazing on ecosystem functions vary with ...
    Feb 1, 2025 · We found that grazing substantially reduced plant productivity (-26 %), followed by water conservation (-18 %) and carbon sequestration (-19 %).
  131. [131]
    Livestock overgrazing disrupts the positive associations between ...
    Apr 19, 2020 · Livestock overgrazing influences both microbial communities and nutrient cycling in terrestrial ecosystems. However, the role of overgrazing ...
  132. [132]
    Prediction of urban expansion by using land cover change detection ...
    However, the extent of urban build-up land will grow from 3.39 km2 in 1998 to 11.01 km2 in 2023 and 12.44 km2 in 2028. Moreover, the future land cover map ...<|separator|>
  133. [133]
    The impact of urbanization on land use land cover change ... - Nature
    Apr 8, 2025 · The aim of this study was to analyze the effect of urbanization on Land Use Land Cover Change (LULCC) at Mizan Aman city, southwest Ethiopia from 1992 to 2022
  134. [134]
    Mining impacts affect up to 1/3 of global forest ecosystems, and ...
    Apr 18, 2023 · More than 80% of direct mining-related deforestation takes place in just 10 countries, with tropical rainforests suffering the most damage – ...
  135. [135]
    Mining Is Increasingly Pushing into Critical Rainforests and ...
    Oct 23, 2024 · A WRI report found that as of 2020, mining concessions and illegal mining covered more than 20% of Indigenous lands in the Amazon, endangering ...
  136. [136]
    Evaluating the environmental and economic impact of mining for ...
    Some of the negative impacts of mining are loss of vegetation cover, mass destruction of water bodies, loss of biodiversity, land-use changes and food ...
  137. [137]
    WWF LPR: Wildlife Populations Down 73% Since 1970
    Oct 9, 2024 · Habitat loss and degradation and overharvesting, driven primarily by our global food system are the dominant threats to wildlife populations ...
  138. [138]
    Subsistence hunting impacts wildlife assemblages and functional ...
    Jan 24, 2025 · Hunting pressure can disrupt trophic webs, resulting in the decline of large-bodied mammals and birds (especially large carnivores and seed ...
  139. [139]
    Exposing illegal hunting and wildlife depletion in the world's largest ...
    Sep 9, 2024 · Of 157 native species targeted by hunters, 19 are currently threatened with extinction. We estimated that 1414 hunters extracted 3251 kg/million ...
  140. [140]
    Habitat degradation and indiscriminate hunting differentially impact ...
    Oct 30, 2019 · Habitat degradation and hunting have caused the widespread loss of larger vertebrate species (defaunation) from tropical biodiversity hotspots. ...
  141. [141]
    Permafrost and Climate Change: Carbon Cycle Feedbacks From the ...
    Oct 17, 2022 · Temperature, organic carbon, and ground ice are key regulators for determining the impact of permafrost ecosystems on the global carbon cycle.
  142. [142]
    Moving up and over: redistribution of plants in alpine, Arctic, and ...
    The climate is changing rapidly, and human influence is intensifying in cold environments. Plant species are responding by moving up in elevation and latitude, ...<|separator|>
  143. [143]
    Contemporary human impacts on alpine ecosystems: the direct and ...
    I review literature regarding current and projected human impacts on alpine ecosystems, including the direct and indirect impacts of human-induced climate ...
  144. [144]
    Permafrost carbon-climate feedbacks accelerate global warming
    Aug 18, 2011 · Permafrost soils contain enormous amounts of organic carbon, which could act as a positive feedback to global climate change due to enhanced respiration rates ...
  145. [145]
    Climate-carbon feedback tradeoff between Arctic and alpine ...
    Sep 17, 2025 · When warming caused drying of alpine permafrost soils, the CO2 sink weakened but the CH4 sink increased. In contrast, warming of relatively wet ...Ghg Sink Strengthened In... · Ghg Responses To Warming... · Materials And Methods
  146. [146]
    Highlighting the role of biota in feedback loops from tundra ...
    We focus specifically on three key feedback loops between tundra and atmosphere (carbon dynamics, albedo and permafrost thaw) and the influences of three key ...
  147. [147]
    Feedbacks of Terrestrial Ecosystems to Climate Change
    Nov 21, 2007 · With modest warming, net feedbacks of terrestrial ecosystems to warming are likely to be negative in the tropics and positive at high latitudes.
  148. [148]
    Soil Erosion Control and Sustainable Land Management - SciTechnol
    Sustainable land management practices play a crucial role in preventing soil erosion and maintaining soil health. The impacts of soil erosion, the factors ...
  149. [149]
    Progress and challenges in sustainable land management initiatives
    Feb 1, 2023 · Sustainable land management (SLM) is widely recognized as the key to reducing rates of land degradation, and preventing desertification.
  150. [150]
    Sustainable land management enhances ecological and economic ...
    Jun 10, 2024 · Results show that intensive management and future climate decrease ecological multifunctionality for most scenarios in both grassland and ...
  151. [151]
    Evidence for the impacts of agroforestry on ecosystem services and ...
    Mar 17, 2022 · Agroforestry is promoted for its potential for carbon sequestration, soil erosion and runoff control, and improved nutrient and water cycling, ...
  152. [152]
    The impacts of agroforestry on agricultural productivity, ecosystem ...
    This EGM collates existing evidence on the impacts of agroforestry on agricultural productivity, ecosystem services, and human well‐being in L&MICs.
  153. [153]
    Evidence for the impacts of agroforestry on agricultural productivity ...
    Oct 29, 2018 · Existing research suggests that integrating trees on farms can prevent environmental degradation, improve agricultural productivity, increase ...
  154. [154]
    Terrestrial ecosystem restoration increases biodiversity and reduces ...
    May 12, 2022 · We found that, relative to unrestored (degraded) sites, restoration actions increased biodiversity by an average of 20%, while decreasing the variability of ...
  155. [155]
    A New Study Indicates Forest Regeneration Provides Climate ...
    Oct 17, 2025 · Forest regeneration works if it is part of a comprehensive climate strategy that aggressively cuts fossil fuel emissions while protecting and ...<|separator|>
  156. [156]
    Balancing the environmental benefits of reforestation in agricultural ...
    Reforestation of agricultural land can improve biodiversity, which can result in increased primary production, reduced susceptibility to invasion by exotic ...
  157. [157]
    Study shows market-based strategies for ecosystem conservation ...
    Mar 12, 2018 · The researchers found strong interest in markets for forestry and land-use practices that store carbon, such as forest conservation and ...
  158. [158]
    [PDF] Sustainable Land Management Sourcebook - World Bank Document
    This sourcebook covers sustainable land management, its definition, challenges, opportunities, and the need for it, as agriculture is fundamental to economic ...<|separator|>
  159. [159]
    The role of sustainable land management practices in alleviating ...
    Sustainable land management practices (SLMPs) are important for ensuring environmental protection, food security, poverty alleviation, and economic growth in a ...
  160. [160]
    A global-level assessment of the effectiveness of protected areas at ...
    Oct 28, 2019 · Our results show that while many PAs are effective, the large focus on increasing terrestrial coverage toward 17% of the earth surface has led ...
  161. [161]
    How effective are protected areas for reducing threats to biodiversity ...
    Sep 8, 2023 · This systematic review aims to identify peer-reviewed and grey literature studies investigating how effective PAs are for reducing threats to biodiversity.
  162. [162]
    Mixed effects of a national protected area network on terrestrial and ...
    Sep 13, 2023 · Our results suggest that the current protected area network can partly contribute to slow down declines in occupancy rates, but alone will not ...
  163. [163]
    Ecological restoration success is higher for natural regeneration ...
    Nov 8, 2017 · Restoration success for biodiversity and vegetation structure was 34 to 56% and 19 to 56% higher in natural regeneration than in active ...Missing: reforestation terrestrial
  164. [164]
    Nearly half of replanted trees die, but careful site selection can help
    Dec 14, 2022 · A recent study tracking forest restoration projects in South and Southeast Asia found that, on average, 44% of trees planted as part of such efforts died ...
  165. [165]
    Long-term and large-scale ecological restoration projects shaped ...
    Dec 15, 2023 · The GPP of new forests has significantly increased trends and the growth rates of new forests vary significantly in different climatic zones.Missing: success empirical
  166. [166]
    Global meta-analysis shows action is needed to halt genetic ...
    Jan 29, 2025 · Conservation strategies designed to improve environmental conditions, increase population growth rates and introduce new individuals (for ...
  167. [167]
    Reviewing the science on 50 years of conservation - PubMed Central
    Jul 18, 2024 · We present a systematic review of published empirical studies about site-level biodiversity conservation initiated between 1970 and 2019.
  168. [168]
    Violence and conservation: Beyond unintended consequences and ...
    Violence and conservation: Beyond unintended consequences and unfortunate coincidences · Conservation as dispossession, violence as erasure · Peasants as ...
  169. [169]
    Communities and conservation: A history of disenfranchisement
    Unintended consequences of wildlife policies and legislation. It has been argued that protected areas have become a lens through which many people see and ...
  170. [170]
    Towards a synthesized critique of forest‐based 'carbon‐fix' strategies
    Feb 21, 2023 · This article synthesizes critiques of 'carbon-fix' strategies in the forestry sector to clarify key concerns about reductionist treatments of forests and ...<|separator|>
  171. [171]
    Challenges and opportunities of area-based conservation in ...
    Dec 1, 2021 · In this review, I identify and relate pressing challenges to promising opportunities for effective and efficient protected area governance and management.
  172. [172]
    Shifts in vegetation activity of terrestrial ecosystems attributable to ...
    Feb 6, 2023 · The most convincing evidence to date for attributing changes in ecosystems to climate change comes from the high latitudes where temperatures ...
  173. [173]
    Global Greening - Climate at a Glance
    NASA shows significant global plant growth over 35 years, due to rising CO2, with a 10% increase between 2000 and 2020.
  174. [174]
    Carbon Dioxide Fertilization Greening Earth, Study Finds - NASA
    Apr 26, 2016 · Results showed that carbon dioxide fertilization explains 70 percent of the greening effect, said co-author Ranga Myneni, a professor in the ...
  175. [175]
    Climate change and the global redistribution of biodiversity
    Apr 11, 2023 · Here, we evaluate the impact of anthropogenic climate change (specifically, changes in temperature and precipitation) on species' ranges ...Missing: peer- | Show results with:peer-
  176. [176]
    Proof of evidence of changes in global terrestrial biomes using ...
    Jul 28, 2023 · Climate change affects plant dynamics and functioning of terrestrial ecosystems. This study aims to investigate temporal changes in global ...
  177. [177]
    Overstating the effects of anthropogenic climate change? A critical ...
    Feb 8, 2023 · The world weather attribution (WWA) initiative demands 'at least two and preferably more models to be good enough for the attribution analysis.
  178. [178]
    Climate variability and vulnerability to climate change: a review - PMC
    We present new analysis that tentatively links increases in climate variability with increasing food insecurity in the future.Climate Change, Climate... · Impacts Of Climate... · Biological Systems<|separator|>
  179. [179]
    [PDF] BIODIVERSITY AND ECOSYSTEM SERVICES - IPBES
    OVERALL REVIEW EDITORS. Manuela Carneiro da Cunha, Georgina M. Mace, Harold Mooney. This report in the form of a PDF can be viewed and downloaded at www.ipbes.Missing: critiques | Show results with:critiques
  180. [180]
    Global Extinction Rates: Why Do Estimates Vary So Wildly?
    Aug 17, 2015 · Only about 800 extinctions have been documented in the past 400 years, according to data held by the International Union for the Conservation of ...
  181. [181]
  182. [182]
    Global trends and scenarios for terrestrial biodiversity and ... - Science
    Apr 25, 2024 · During the 20th century, biodiversity declined globally by 2 to 11%, as estimated by a range of indicators.
  183. [183]
    Opinion Questioning the sixth mass extinction - ScienceDirect.com
    Oct 1, 2024 · But <0.1% of Earth's known species have gone extinct in the last 500 years, and projections of undiscovered species loss are also limited. Other ...Missing: actual | Show results with:actual
  184. [184]
    Land statistics 2001–2023. Global, regional and country trends
    Jun 19, 2025 · Main findings: In 2023, world total agricultural land was 4 800 million hectares (ha), more than one-third of the global land area.Missing: biodiversity | Show results with:biodiversity
  185. [185]
    Empirical evidence supports neither land sparing nor land sharing ...
    Sep 2, 2025 · Empirical evidence supports neither land sparing nor land sharing as the main strategy to manage agriculture–biodiversity tradeoffs | PNAS ...
  186. [186]
    Land sharing versus land sparing—What outcomes are compared ...
    Sep 4, 2021 · Our review showed that both land sharing and land sparing have been represented as high-yielding commodity production combined with preserved ...
  187. [187]
    Trade-offs between agricultural production and ecosystem services ...
    Jul 1, 2025 · This study evaluated the trade-offs between provisioning ecosystem services (crop yields) and other key ecosystem services including regulating services.
  188. [188]
    Intensifying farmland could degrade biodiversity more than expansion
    May 7, 2025 · The intensification of existing farmland can sometimes be more harmful to local biodiversity than expanding the area covered by agricultural land.
  189. [189]
    Global impacts of future cropland expansion and intensification on ...
    Jun 28, 2019 · Identifying trade-offs between ecosystem services, land use, and biodiversity: a plea for combining scenario analysis and optimization on ...
  190. [190]
    Trade‐offs between biodiversity and agriculture are moving targets ...
    Jul 21, 2020 · In summary, our main finding that agriculture–biodiversity trade–offs vary considerably with landscape context, as measured by the share of ...
  191. [191]
    How Agricultural Intensification Affects Biodiversity and Ecosystem ...
    The loss of biodiversity driven by agricultural intensification (AI) is judged to be similar in scale to that expected from climate change (Tilman et al., 2001) ...