Fact-checked by Grok 2 weeks ago

Ecosystem


An ecosystem consists of all living organisms in a defined area interacting with each other and their nonliving physical and chemical , forming a functional where energy flows and materials cycle. Biotic components include autotrophic producers, such as plants and algae, that convert into through ; heterotrophic consumers, encompassing herbivores, carnivores, and omnivores; and decomposers, like and fungi, that mineralize organic matter. Abiotic factors comprise , , water availability, composition, and inorganic nutrients essential for sustaining these interactions.
Ecosystems maintain dynamic balance through unidirectional energy flow, primarily driven by at the base of food chains, with efficiency decreasing across trophic levels due to and heat loss, and closed-loop nutrient cycling involving processes like and that recycle elements such as carbon, , and . These processes underpin ecosystem and , with empirical studies showing that nutrient availability strongly influences and overall stability. Ecosystems vary by scale, from small ponds to vast biomes, and are broadly classified into terrestrial types—such as forests, grasslands, deserts, and tundras—and types, including freshwater rivers and lakes alongside environments like oceans and reefs. Human activities, including and , disrupt these natural dynamics, altering and service provision like and , as evidenced by long-term ecological monitoring data.

Definition and Conceptual Foundations

Core Definition and Principles

An ecosystem is defined as the integrated system comprising a biological of interacting organisms and their surrounding physical within a defined spatial unit, encompassing both biotic (living) and abiotic (non-living) components along with their mutual influences. This concept emphasizes the ecosystem as a holistic entity akin to physical systems, where the organismal complex cannot be isolated from environmental factors such as , , , and inorganic substances. The term "ecosystem" was introduced by British ecologist in 1935 to resolve debates in vegetation studies, building on earlier ideas of organism-environment unity while rejecting overly organism-centric views like those in the holistic "" theories of Frederic Clements; Tansley argued for a mechanistic, dissociative perspective that treats ecosystems as complexes amenable to empirical dissection. Although suggested privately by botanist A.R. Clapham around 1930, Tansley's publication formalized and popularized the concept. At its core, biotic components include autotrophs (primary producers like plants and algae that convert solar energy via ), heterotrophs (consumers such as herbivores, carnivores, and omnivores), and decomposers (microorganisms and detritivores that break down ), all of which interact through predation, , , and to shape structure and dynamics. Abiotic components encompass physical and chemical elements like , , , , nutrient availability, and geological features, which impose constraints and enable processes; for instance, solar radiation drives , while water availability limits species distributions and metabolic rates. Interactions between biotic and abiotic factors are bidirectional: organisms modify their environment (e.g., beavers altering via dams, or plant roots influencing ), while abiotic conditions select for adaptive traits, fostering evolutionary responses over time. Fundamental principles governing ecosystems derive from thermodynamic and material balance laws applied to biological systems. Energy enters primarily via solar input to autotrophs and flows unidirectionally through trophic levels with progressive inefficiency (typically 10% transfer efficiency per level, per Lindeman's trophic-dynamic concept), ultimately dissipating as heat and necessitating continuous external influx to sustain the system. Matter, in contrast, cycles internally through biogeochemical pathways (e.g., carbon, ), with decomposers recycling nutrients from dead material back to producers, preventing indefinite accumulation or depletion absent external perturbations. Ecosystems exhibit self-regulation through negative feedbacks, such as density-dependent predation stabilizing numbers, though they remain open systems vulnerable to external disturbances like climatic shifts; this balance underscores causal dependencies on gradients and availability rather than isolated organismal . Empirical quantification, via metrics like net primary productivity (e.g., global terrestrial average of ~560 g C/m²/year), reveals scale-dependence, with nested hierarchies from microbial patches to biomes, but principles hold across contexts due to universal biophysical constraints.

Historical Origins and Evolution of the Concept

The term "ecosystem" was coined by British ecologist Arthur George Tansley in 1935, in response to prevailing holistic interpretations of vegetation that treated communities as s analogous to biological entities. Tansley introduced the concept in his paper "The Use and Abuse of Vegetational Concepts and Terms," published in , to denote the integrated complex of organisms and their abiotic environment functioning as a physicochemical driven by causal interactions rather than organismic unity. This formulation countered the organismal views advanced by American ecologist Frederic Clements, whose 1916 theory of the climax formation portrayed vegetation as a developing shaped by toward equilibrium. Tansley's emphasis on the ecosystem as a bounded unit of study, encompassing both and abiotic factors, provided a framework for analyzing ecological processes without anthropomorphic or teleological assumptions, drawing implicitly from earlier notions like Karl Möbius's 1877 "" for communities and physical habitats. The ecosystem concept gained functional depth in the 1940s through quantitative modeling of internal dynamics. Raymond Lindeman's 1942 posthumous paper, "The Trophic-Dynamic Aspect of ," applied to energy transfer across trophic levels within ecosystems, conceptualizing them as steady-state systems where efficiency of flow (typically 10-20% between levels) governs and distribution. This marked a shift from descriptive to process-oriented , influencing subsequent work on material and budgets. By the 1950s, Eugene Odum's textbook Fundamentals of (1953) formalized ecosystems as networks of autotrophic producers, heterotrophic consumers, and decomposers, incorporating cycling and self-regulation akin to thermodynamic systems. These advancements aligned with emerging and general , enabling empirical quantification via techniques like radioecology during the Commission's studies in the late 1940s and 1950s. By the 1960s, the ecosystem paradigm had become central to global ecological research, exemplified by the International Biological Programme (1964–1974), which coordinated studies on ecosystem productivity across biomes, revealing patterns such as net primary production varying from 0.1 g/m²/year in deserts to over 2000 g/m²/year in tropical forests. This evolution reflected a progression from Tansley's qualitative integration to measurable fluxes, though debates persisted over scale—whether ecosystems represent discrete units or scalable processes—and the relative roles of biotic feedbacks versus abiotic drivers in maintaining stability. The concept's adoption facilitated interdisciplinary applications, including environmental management, while underscoring ecology's departure from purely biotic-focused paradigms toward holistic yet mechanistic understandings of natural systems.

Structural Components

Biotic Components and Interactions

components encompass all living organisms within an ecosystem, including , animals, fungi, and microorganisms, which interact dynamically to sustain ecological processes. These organisms are classified into three primary functional groups based on their nutritional strategies and roles in trophic dynamics: producers, consumers, and decomposers. Producers, or autotrophs, such as vascular , , and , harness via to fix carbon and generate , forming the base of food webs. In s like reefs, symbiotic () within s contribute significantly to , supporting diverse consumer populations. Consumers, heterotrophic organisms, derive energy by ingesting other organisms and are subdivided into primary consumers (herbivores feeding on producers), secondary consumers (carnivores preying on herbivores), and higher-level predators. Herbivores, such as impalas in savannas, convert plant biomass into animal tissue, transferring upward through trophic levels with efficiency typically around 10%. Apex predators, like cheetahs, occupy top trophic positions, exerting top-down control on prey populations and preventing . Decomposers, including and saprotrophic fungi, mineralize dead organic matter, releasing nutrients like and back into the or for reuse by producers. Interactions among biotic components drive ecosystem structure and function, encompassing predation, , and . Predation involves a predator consuming prey, often stabilizing populations through cycles, as observed in the Canadian lynx-snowshoe hare system where hare densities fluctuate cyclically every 8-11 years due to predator pressure and prey reproduction rates. occurs when organisms vie for limited resources, such as intraspecific density-dependent effects limiting or interspecific exclusion, as in Gause's where similar species cannot coexist indefinitely on the same resources. Symbiotic interactions include , where both parties benefit, exemplified by mycorrhizal fungi enhancing plant nutrient uptake in exchange for carbohydrates, boosting forest productivity by up to 20% in phosphorus-limited soils. and represent asymmetric benefits, with parasites like ticks deriving at expense, potentially influencing and dynamics. These interactions collectively regulate , energy transfer, and , with food webs illustrating complex, non-linear connections rather than simple chains.

Abiotic Components and Environmental Drivers

Abiotic components comprise the non-living physical, chemical, and geological elements of an ecosystem that provide the foundational environment for interactions. These include climatic variables such as , , humidity, wind, and solar radiation; edaphic factors encompassing , , nutrient content, and organic matter; hydrological features like water availability, flow rates, and ; and atmospheric gases including oxygen and concentrations. These abiotic elements function as primary environmental drivers, exerting causal influences on species distributions, , and overall ecosystem through direct physiological constraints and indirect structuring. , for instance, modulates enzymatic reactions and metabolic processes, with optimal ranges varying by ; ectotherms exhibit activity limits around 0–40°C, while extremes beyond these thresholds induce or mortality. regulates water balance, a frequent limiting resource in terrestrial systems, where annual inputs below 250 mm sustain deserts with low , contrasting with >2,000 mm in tropical rainforests supporting high . properties drive nutrient availability and root anchorage, with levels (typically 4–8 in most ecosystems) affecting cation and microbial activity; acidic soils ( <5.5) often limit phosphorus uptake, constraining plant growth in podzols. Light intensity and photoperiod influence photosynthetic rates and seasonal cycles, with photosynthetically active radiation (PAR) averaging 400–700 nm wavelength and intensities up to 2,000 µmol m⁻² s⁻¹ in open habitats declining exponentially under dense canopies. In aquatic environments, dissolved oxygen (5–10 mg/L in well-oxygenated waters) and turbidity control respiration and light penetration, respectively, shaping vertical stratification in lakes and oceans. Interactions among drivers amplify effects; elevated temperatures under reduced intensify evapotranspiration, depleting moisture and shifting compositions toward drought-tolerant species, as observed in semi-arid grasslands. Geological factors like topography influence microclimates and erosion rates, with slopes >30% promoting runoff and reducing depth, thereby limiting vegetation in mountainous ecosystems. Chemical gradients, such as gradients in estuaries (0–35 ppt), impose osmotic challenges that filter tolerant , maintaining distinct zonation patterns. Empirical studies confirm abiotic dominance in structuring basal ecosystem functions, with climatic and edaphic variables explaining up to 60% of variance in multifunctionality metrics across biomes.

Core Processes

Energy Flow and Trophic Dynamics

In ecosystems, energy enters primarily as solar radiation captured by autotrophic producers through , converting approximately 1-2% of incident sunlight into in terrestrial systems and up to 10% in some environments. This initiates a unidirectional flow through trophic levels, where heterotrophic consumers and decomposers assimilate portions of that , with the remainder lost as via and entropy increase, adhering to the second law of . Unlike matter, which cycles, energy dissipates without , limiting ecosystem productivity to continuous external inputs. Trophic levels classify organisms by their position in this energy transfer: primary producers at the base fix ; primary consumers (herbivores) derive nutrition from producers; secondary and tertiary consumers (carnivores) feed on lower levels; and decomposers break down , facilitating partial release back to the . Raymond Lindeman formalized this trophic-dynamic in 1942, analyzing a temperate to demonstrate how budgets integrate across levels, with "trophic" referring to feeding relations and "dynamic" to flux rates and efficiencies. His model quantified inputs like gross (e.g., 8,500 kcal/m²/year in the studied lake) and outputs, revealing cascading transformations from to . Energy transfer efficiency between trophic levels averages around 10%, as only a fraction of ingested is assimilated into after accounting for egestion, , and —empirical from diverse ecosystems support this "10% rule," though values range from 5-20% depending on taxa and . For instance, assimilation from material often yields 10-15% efficiency, while transfers drop lower due to higher metabolic demands. This inefficiency manifests in energy pyramids, where available halves or more per level, constraining most ecosystems to 3-5 trophic levels and explaining sparse (e.g., lions comprising <1% of savanna ). Food chains depict linear sequences of energy passage (e.g., grass → zebra → lion), simplifying dynamics but overlooking redundancy, whereas food webs interconnect multiple chains to model realistic omnivory, alternative pathways, and stability against perturbations. In webs, energy flux distributes across nodes, with keystone species influencing overall throughput; modeling studies show webs sustain higher total transfer than isolated chains by buffering losses. Trophic dynamics thus govern ecosystem function, where disruptions like overfishing cascade to reduce lower-level biomasses, as observed in marine systems with 20-50% efficiency drops post-predator removal.

Nutrient Cycling and Matter Transformation

Nutrient cycling encompasses the continuous movement and chemical transformation of essential elements, including , , and , through biotic and abiotic compartments of an ecosystem. These biogeochemical processes, driven by microbial activity, plant uptake, animal consumption, and abiotic factors like weathering and atmospheric deposition, recycle finite nutrient pools to sustain primary production and overall ecosystem function. Soil biogeochemical cycles form the foundation for nutrient and energy flows that regulate productivity, with inefficiencies leading to losses via leaching or gaseous emissions. The nitrogen cycle exemplifies nutrient transformation, beginning with fixation of atmospheric N₂ into ammonia by symbiotic bacteria in plant roots or free-living diazotrophs, followed by nitrification to nitrate and denitrification back to N₂, maintaining long-term oceanic and terrestrial balances over geological timescales. In contrast, the carbon cycle involves photosynthetic fixation into organic compounds, respiration releasing CO₂, and decomposition mineralizing detritus, with rates varying by ecosystem type; phosphorus cycling relies more on rock weathering for inputs and microbial mineralization for recycling, lacking a gaseous phase. These cycles interconnect, as disruptions in one—such as excess anthropogenic nitrogen inputs—can accelerate carbon sequestration but also eutrophication. Matter transformation primarily occurs via decomposition, where heterotrophic microbes and soil fauna break down dead organic matter through enzymatic hydrolysis and oxidation, converting complex polymers into simpler inorganic ions and gases. This process releases 90-95% of net primary production back into usable forms annually in most terrestrial ecosystems, with activation energies for microbial decomposition typically ranging 40-60 kJ/mol, influenced by temperature and substrate quality. Fauna like earthworms enhance rates by fragmentation, while roots stimulate both decomposition and stabilization of residues into soil organic matter. Variations in cycling efficiency arise from climatic gradients, with tropical ecosystems exhibiting faster turnover due to high temperatures and moisture, releasing nutrients via herbivory and decomposition at rates exceeding temperate zones by factors of 2-5 for key elements. Losses, such as denitrification contributing 10-20% of N₂ emissions globally, underscore incomplete closure, necessitating external inputs in agriculture but risking overload in natural systems.

Primary Production and Decomposition

Primary production constitutes the initial synthesis of organic matter from inorganic precursors by autotrophs, predominantly via photosynthesis in plants, algae, and cyanobacteria, or chemosynthesis in certain prokaryotes, thereby introducing fixed carbon and energy into ecosystems. Gross primary production (GPP) quantifies the total rate of this fixation, typically measured in grams of carbon per square meter per year (g C m^{-2} yr^{-1}), while net primary production (NPP) subtracts autotrophic respiration costs, yielding the biomass available for heterotrophs and storage. NPP rates differ markedly by ecosystem type and environmental conditions; tropical rainforests achieve up to approximately 2000 g C m^{-2} yr^{-1}, temperate forests range from 600 to 1200 g C m^{-2} yr^{-1}, while arid deserts and open oceans often fall below 100 g C m^{-2} yr^{-1}. Key limiting factors include solar irradiance, temperature optima for enzymatic reactions, water availability, and macronutrient concentrations such as nitrogen and phosphorus, with CO_2 saturation influencing photosynthetic efficiency. Decomposition involves the heterotrophic catabolism of detritus—dead plant, animal, and microbial remains—by bacteria, fungi, protozoa, and macroinvertebrates, progressively mineralizing organic compounds into CO_2, water, and inorganic nutrients like ammonium and phosphate. This breakdown proceeds in stages, from initial fragmentation and leaching to microbial colonization and final humification, with rates accelerating under higher temperatures and moisture that enhance decomposer metabolism. Litter quality, characterized by low C/N ratios and minimal recalcitrant compounds like lignin, further promotes faster decomposition by alleviating microbial nutrient constraints. The linkage between primary production and decomposition sustains ecosystem function through nutrient cycling; detrital inputs from high NPP supply substrate for decomposers, whose mineralization releases bioavailable ions critical for autotrophic growth, preventing oligotrophication in closed systems. Inefficient decomposition, as in cold or anaerobic conditions, accumulates refractory organic pools, reducing nutrient turnover and constraining subsequent production, whereas balanced rates in temperate or tropical settings optimize matter recycling and trophic support. Empirical studies confirm that decomposer diversity enhances both decomposition efficiency and feedbacks to primary productivity via improved soil fertility.

Dynamics and Stability

Internal Feedbacks and Resilience

Internal feedbacks in ecosystems consist of regulatory loops where biotic and abiotic components influence one another, either dampening or amplifying perturbations to maintain or alter system states. Negative feedbacks promote stability by counteracting deviations from equilibrium; for example, in classic predator-prey interactions modeled by , rising herbivore densities increase predator reproduction, which subsequently curbs herbivore numbers, fostering oscillatory balance around carrying capacities. Positive feedbacks, conversely, intensify initial changes, potentially leading to rapid shifts; in semi-arid grasslands, overgrazing reduces plant cover, exposing soil to erosion that further diminishes vegetation and perpetuates degradation. These feedbacks underpin ecosystem resilience, defined by C.S. Holling in 1973 as the magnitude of disturbance a system can absorb while retaining its core structure, processes, and alternative stable states, distinct from engineering resilience's emphasis on rapid return to a single equilibrium. Negative feedbacks enhance this capacity by buffering shocks—such as density-dependent mortality in fish populations limiting overexploitation—whereas entrenched positive feedbacks, like nutrient enrichment fueling algal blooms that suppress macrophytes in eutrophic lakes, can lock systems into less productive regimes. Empirical quantification often involves metrics like recovery trajectories post-disturbance, with satellite-derived vegetation indices from 1982 to 2018 revealing global declines in terrestrial resilience, particularly in tropical forests and boreal zones, where feedbacks fail against compounded stressors like drought and land-use change. Resilience varies by ecosystem type and is modulated by attributes such as functional redundancy, where multiple species perform similar roles to sustain processes amid losses. Rewilding initiatives, reintroducing large herbivores to European grasslands, have demonstrated resilience gains in approximately 70% of cases by restoring trophic feedbacks that mitigate flood and fire extremes, though 20% report neutral or adverse effects from altered grazing patterns. However, causal links to biodiversity remain contested; microcosm experiments with lake communities show diverse assemblages resist invasions better but recover more slowly from pulses due to competitive interference overriding stabilizing feedbacks. Connectivity via dispersal corridors further bolsters resilience by replenishing depleted patches, as evidenced in fragmented coral reefs where larval exchange sustains from bleaching events exceeding 50% cover loss. Thresholds where feedbacks tip toward instability—known as tipping points—mark resilience limits, often quantified as the perturbation size precipitating irreversible shifts, such as permafrost thaw releasing 1.5-2.5 GtC annually by 2100, amplifying warming via methane feedbacks. Long-term monitoring underscores that intact feedbacks confer persistence; for instance, Yellowstone's wolf reintroduction in 1995 stabilized aspen regeneration through apex predation cascades, reversing 70 years of decline. Yet, anthropogenic overrides, including nutrient loading and habitat fragmentation, erode these mechanisms, with studies indicating 30-50% of global ecosystems approaching critical thresholds as of 2022.

Disturbance Regimes and Natural Variability

Disturbance regimes refer to the characteristic patterns of disturbance events within an ecosystem, defined by their frequency, intensity, spatial scale, and duration, which collectively shape community structure and dynamics over time. These events, such as wildfires, floods, windstorms, insect outbreaks, and herbivory, disrupt biotic and abiotic components by causing mortality, altering resource availability, or modifying habitat patches. Unlike single events, regimes encompass the cumulative, recurrent nature of disturbances, often operating across landscapes to produce a mosaic of successional stages. In forest ecosystems, fire exemplifies a dominant disturbance regime, with historical patterns varying by vegetation type; for instance, ponderosa pine forests in the western United States experienced low-intensity surface fires every 5 to 25 years, which cleared understory fuels and promoted fire-resilient species without widespread canopy replacement. High-intensity crown fires, occurring less frequently (e.g., every 100–300 years in boreal forests), contrast by killing mature trees and initiating stand-replacing succession, thereby influencing long-term carbon storage and species composition. Such variability in regime attributes—measured via return intervals, burn severity, and patch size—prevents uniform succession toward climax communities, fostering heterogeneity essential for biodiversity. Natural variability arises from the stochastic interplay of these regimes with environmental drivers like climate fluctuations, generating non-equilibrium dynamics where ecosystems rarely reach a stable state. Patch dynamics, a core mechanism, describe how disturbances create discrete habitat patches of varying age and condition, enabling coexistence of early- and late-seral species; for example, post-fire patches in coniferous forests support serotinous seeders like while allowing colonization by shade-intolerant pioneers. This variability enhances resilience by preserving ecological memory—residual structures or propagules from prior states that facilitate recovery—though shifts in regime attributes, such as increased fire intensity from drought, can exceed adaptive thresholds. Empirical studies across biomes indicate that moderate disturbance frequencies maximize species richness, as extreme variability reduces occupancy while infrequency permits competitive exclusion. Disturbances contribute to ecosystem stability not through constancy but via recurrent resets that buffer against chronic stressors, with diverse communities exhibiting lower temporal variability in function amid climatic fluctuations. In lotic systems, flood regimes scour channels and redistribute sediments, maintaining riffle-pool heterogeneity and invertebrate diversity; return intervals of 1–10 years in temperate streams sustain metapopulation dynamics without eradicating populations. Overall, these regimes embody causal realism in ecology, where variability drives adaptation and function rather than deviation from an idealized equilibrium, supported by paleoecological records showing persistent patch turnover over millennia.

Ecological Succession and Long-Term Change

Ecological succession describes the sequential replacement of species assemblages in an ecosystem following a disturbance or initial substrate exposure, driven by biotic interactions and abiotic modifications that alter habitat suitability. Primary succession begins on sterile substrates devoid of soil and propagules, such as retreating glacial till or volcanic lava, where initial colonizers like lichens and cyanobacteria initiate pedogenesis through biochemical weathering and nitrogen fixation, gradually enabling vascular plant establishment over decades to millennia. In contrast, secondary succession follows partial biotic removal—such as post-fire or post-agricultural abandonment—on preexisting soil, accelerating community redevelopment due to residual seed banks and root structures, often reaching mid-seral stages within years to decades. Empirical chronosequences, like those in , document primary succession progressing from herbaceous pioneers to coniferous forests over approximately 200 years, with soil nitrogen accumulation correlating directly with species turnover rates. Mechanisms underlying succession include facilitation, where early species ameliorate harsh conditions (e.g., nitrogen-fixing algae enhancing soil fertility for grasses); inhibition, via competitive exclusion delaying successors until dominant pioneers senesce; and tolerance, where later-arriving species persist under unmodified pioneer conditions without active suppression. These processes, formalized by Connell and Slatyer in 1977, explain variability across sites: facilitation dominates in nutrient-poor primary sequences, while inhibition prevails in productive secondary ones, as evidenced by old-field experiments in Michigan showing grass dominance inhibiting forbs for 10-20 years before woody encroachment. Early 20th-century theories diverged sharply: Frederic Clements (1916) viewed succession as organismic development toward a climatically determined, self-perpetuating "climax" community, implying holistic predictability. Henry Gleason (1926) advocated an individualistic continuum hypothesis, attributing patterns to independent species responses to edaphic and microclimatic gradients rather than community-level teleology, a perspective bolstered by modern phytosociological data revealing fuzzy boundaries over sharp Clementsian zones. Empirical tests, including resurveys of Piedmont old fields from the 1930s-1940s, confirm Gleasonian stochasticity in dispersal-limited assemblages but Clementsian facilitation in soil-building phases, integrating both via initial environmental filtering followed by niche partitioning. Long-term ecosystem change extends beyond seral dynamics, incorporating allogenic forcings like climatic shifts or megadisturbances that reset trajectories or induce regime shifts to alternative stable states, as opposed to autogenic convergence on a singular climax. Paleoecological reconstructions from sediment cores in boreal forests reveal Holocene successional cycles punctuated by millennial-scale aridification events around 6,000-4,000 years ago, shifting dominance from deciduous to coniferous taxa via altered hydrology rather than internal maturation alone. In tropical systems, such as Amazonian rainforests, avian and floral inventories spanning 1979-2016 indicate succession toward higher biomass but disrupted by selective logging, with functional diversity stabilizing only after 40+ years absent recurrent perturbations. Disturbance regimes—fires every 50-200 years in savannas or hurricanes in coastal dunes—prevent equilibrium, fostering mosaics of seral stages; exclusion experiments in ponderosa pine stands since 1920 demonstrate fuel accumulation leading to high-severity fires, altering long-term composition toward shade-tolerant hardwoods. Contemporary syntheses emphasize priority effects and dispersal constraints as causal determinants of variability, with meta-analyses of 100+ chronosequences showing convergence in productivity but divergence in species identity due to propagule arrival timing. Thus, while succession imposes directional structure, long-term persistence hinges on external variability, challenging deterministic models and underscoring empirical observation over theoretical absolutism.

Biodiversity and Ecosystem Function

Measures and Patterns of Biodiversity

Biodiversity measures in ecosystems encompass metrics that quantify variation at genetic, species, and ecosystem levels, though most ecological assessments focus on species-level diversity as a proxy for overall complexity. Species richness, the count of distinct species within a defined area or community (often denoted as S), represents the simplest and most intuitive measure, capturing the raw number of biological entities without regard to abundance. This metric scales with sampling effort and area, necessitating standardized protocols like rarefaction to enable comparisons across studies. Limitations arise from its sensitivity to undersampling rare species, potentially underestimating true diversity in heterogeneous ecosystems. To account for relative abundances, diversity indices integrate both richness and evenness—the uniformity of species distribution. The Shannon index (H), formulated as H = -∑(p_i \ln p_i) where p_i is the proportional abundance of the i-th species, weights rarer species more heavily and approaches higher values with greater evenness. Values typically range from 0 (no diversity) to around 4-5 in highly diverse communities like tropical forests. The Simpson index (D), conversely, emphasizes dominance via the probability that two randomly selected individuals belong to the same species (D = ∑(n_i(n_i-1))/(N(N-1))), with 1-D yielding diversity; it is less sensitive to rare species but more robust to small sample sizes. Evenness (J) normalizes these indices by maximum possible diversity (J = H / \ln S), quantifying deviation from perfect equity. These alpha-diversity metrics describe local variation, while beta-diversity quantifies turnover between sites, often via Sørensen dissimilarity, and gamma-diversity aggregates regional totals. Beyond taxonomic counts, functional diversity assesses trait variation supporting ecosystem roles, such as via convex hull volumes in trait space, revealing how species occupy niche axes like body size or foraging strategy. Phylogenetic diversity measures evolutionary divergence, using branch lengths in trees to prioritize ancient lineages. Empirical data indicate that combining multiple metrics better approximates total biodiversity, as no single index fully captures proxies for underlying genetic or functional variation. Observed patterns of biodiversity exhibit spatial and temporal structure driven by abiotic gradients and biotic interactions. The latitudinal diversity gradient (LDG) manifests as a monotonic increase in species richness from poles to equator, documented across terrestrial, freshwater, and marine realms for taxa including tetrapods, mosses, and insects. For instance, tropical regions harbor over 50% of global species despite covering less than 10% of land area, with evidence from fossil records confirming the pattern's persistence over deep time, albeit with varying steepness. Causal mechanisms include elevated tropical speciation rates from climatic stability and energy availability, exceeding poleward extinction, though productivity alone does not fully explain the gradient's universality. The species-area relationship, empirically fitted as S = cA^z where A is area and z ≈ 0.25-0.35 for islands or habitat patches, underscores how larger extents support more species via immigration and reduced extinction risk. This power-law holds in fragmented ecosystems, informing conservation by predicting diversity loss from habitat reduction; a 90% area decrease correlates with 10-50% species loss depending on z. Altitudinal patterns mirror LDGs on mountains, with richness peaking mid-elevation due to area compression and climate moderation. Temporal patterns reveal hotspots in ancient, stable biomes like coral reefs (up to 1,500 fish species per system) and rainforests, contrasting low-diversity polar tundra. Recent syntheses confirm these gradients persist amid anthropogenic pressures, with human impacts compressing diversity toward novel low-richness states. Experimental manipulations of plant species diversity in grasslands have demonstrated causal positive effects on ecosystem productivity. In the Cedar Creek Biodiversity Experiment, initiated in 1994 at the Cedar Creek Ecosystem Science Reserve in Minnesota, plots were randomly assembled with 1 to 16 native prairie plant species, revealing that aboveground biomass production increased logarithmically with species richness, with 16-species plots yielding approximately twice the biomass of monocultures after 15 years. This effect persisted over two decades, with higher-diversity plots maintaining greater productivity even as initial sampling biases diminished. Similarly, the BIODEPTH experiment, conducted across eight European sites from 1996 to 1999, manipulated species richness from 1 to potentially 32 species per plot and found that primary productivity and nutrient retention improved with diversity, though the strength varied by site due to local environmental conditions. These causal links operate through specific mechanisms, including complementarity effects, where diverse species partitions resources more efficiently—such as differing root depths accessing varied soil nutrients—and selection effects, where higher diversity increases the chance of including highly productive species. Complementarity enhances over time as species interactions stabilize, contributing to sustained multifunctionality, while selection can dominate initial productivity gains but wanes in long-term assemblages. Meta-analyses of over 100 such experiments confirm these patterns, showing biodiversity explains 20-50% of variation in functions like biomass production and decomposition rates across terrestrial and aquatic systems. The insurance hypothesis further links diversity causally to temporal and spatial stability: diverse communities buffer against environmental fluctuations by compensating for species losses via functional redundancy or response diversity, reducing variability in ecosystem processes. For instance, in fluctuating environments, higher-diversity plots exhibit lower temporal variation in productivity compared to low-diversity ones, as evidenced in Cedar Creek data where diversity mitigated drought impacts. This insurance extends to multifunctionality, where diverse assemblages maintain multiple services (e.g., pollination, pest control) more reliably than simplified ones, even under stressors like nutrient enrichment. Scaling experiments indicate these causal relationships hold from plot to landscape levels, though effect sizes diminish at larger scales due to dispersal limitations. Causal evidence for negative or neutral links is rarer but occurs in over-saturated diversity scenarios or when keystone species dominate, yet meta-analytic consensus supports net positive effects on core functions like carbon sequestration and resilience. These findings derive primarily from controlled field trials minimizing confounds like environmental gradients, establishing diversity as a driver rather than mere correlate of function.

Controversies in Biodiversity-Ecosystem Debates

One central controversy in biodiversity-ecosystem debates revolves around the causal direction and strength of the biodiversity-ecosystem functioning (BEF) relationship, particularly whether species richness directly enhances ecosystem productivity and stability or if higher productivity primarily drives diversity. Experimental manipulations, such as the Cedar Creek Biodiversity Experiment initiated in 1994, have demonstrated positive, often saturating effects of plant species richness on productivity in grasslands, with multi-year data showing yields up to 1.7 times higher in 16-species plots compared to monocultures. However, observational studies across larger scales and biomes, including forests and marine systems, frequently reveal the reverse: productivity gradients explain more variance in species richness than richness explains in productivity, suggesting environmental factors like resource availability mediate the link rather than diversity per se. Critics argue that small-scale experiments overestimate BEF effects due to artificial conditions, such as uniform soil and no large herbivores, failing to capture natural variability where species identity or functional traits outweigh raw richness. The diversity-stability hypothesis, positing that higher biodiversity buffers ecosystems against perturbations, remains contested despite meta-analyses affirming temporal stability increases with species richness in controlled settings. Early theoretical work highlighted a complexity-stability paradox, where more species interactions could amplify instability through indirect effects, a view supported by some network models showing destabilizing trophic cascades in diverse food webs. Empirical critiques point to real-world examples, such as productive but low-diversity agricultural systems or coral reefs where keystone species like certain algae or fish dominate function more than overall richness; losses here often stem from specific removals rather than generalized diversity decline. Longitudinal data from experiments indicate BEF relationships may strengthen over decades due to accumulating soil legacy effects or pathogen buildup in low-diversity plots, yet this temporal pattern is inconsistent across ecosystems and challenged by evidence that stability arises more from asynchronous species responses than diversity alone. Debates also encompass the scalability and generalizability of BEF findings, with proponents of the "insurance hypothesis" claiming diversity provides redundancy against environmental change, while skeptics highlight idiosyncratic effects where rare or functional specialists, not total richness, drive resilience. For instance, in fluctuating environments, models predict diversity reduces productivity variance by 20-50% through portfolio effects, but field observations in variable climates show climate mediates BEF, weakening links in high-productivity or extreme sites. This has fueled disputes over conservation priorities: advocates link BEF to justifying broad species protection, yet analyses reveal that ecosystem services like pollination or carbon sequestration often depend on a subset of species, questioning the necessity of preserving all taxa amid trade-offs with human land use. Such intersections with policy underscore biases in research funding toward alarmist narratives, though empirical synthesis cautions against assuming uniform collapse from diversity loss without site-specific validation.

Methods of Investigation

Field-Based and Experimental Approaches

Field-based approaches in ecosystem ecology rely on direct, in-situ observations and long-term monitoring to capture natural variability, processes, and interactions without artificial intervention. These methods include systematic sampling of biotic and abiotic components, such as species inventories, biomass measurements, and flux assessments, often standardized across networks like the U.S. National Science Foundation's Long-Term Ecological Research (LTER) program, established in 1980 with core data collection on primary production, population dynamics, trophic structure, organic matter accumulation, and disturbance. By 2023, the LTER network encompassed 28 sites spanning diverse biomes, enabling detection of slow-changing phenomena like succession and climate-driven shifts that short-term studies miss. Experimental approaches introduce controlled manipulations in natural settings to test causal hypotheses, balancing realism with replicability through treatments like nutrient additions, species removals, or disturbance simulations. The , operational since 1955 and designated an LTER site in 1988, exemplifies whole-ecosystem manipulations; a 1965-1966 watershed deforestation experiment across 15.7 hectares increased streamflow by 40% and nutrient leaching, revealing forest hydrology and biogeochemical feedbacks. Similarly, calcium silicate additions in a 2000s watershed treatment neutralized soil acidity, boosting tree growth and invertebrate populations while altering stream chemistry. Long-running grassland experiments provide insights into community responses over centuries. The , started in 1856 at , UK, applies varied fertilizer regimes to 2.8-hectare plots, demonstrating that liming and nitrogen inputs reduce species richness from over 50 to fewer than 10 taxa while favoring competitive grasses and altering soil pH from 5.5 to 7.5. In prairie systems, the , initiated in 1994 across 168 plots, manipulates plant species richness from 1 to 16 levels, showing that higher diversity enhances aboveground productivity by up to 50% and resistance to drought via complementarity effects. These methods complement each other, with field observations providing baselines for experimental design, though challenges persist: field experiments often struggle with replication due to site heterogeneity and logistical constraints, potentially underestimating responses like biomass declines under drought compared to observational data from natural events. Manipulations like rainout shelters, used in global drought simulations since the 2000s, replicate precipitation reductions of 30-50% to quantify effects on carbon cycling, but short durations (typically 1-5 years) may miss lagged or threshold responses. Integration across scales, as advocated in recent syntheses, enhances inference by combining plot-level treatments with landscape monitoring.

Modeling and Theoretical Frameworks

Modeling in ecosystem ecology employs conceptual, analytical, and simulation approaches to represent interactions among biotic and abiotic components, predict responses to perturbations, and test hypotheses about system behavior. Conceptual models outline ecosystem structure through diagrams of compartments like producers, consumers, and decomposers, facilitating qualitative understanding of energy flows and nutrient cycles. Analytical models, often based on differential equations, quantify dynamics such as population growth and trophic interactions, while simulation models integrate stochastic elements and spatial heterogeneity to forecast outcomes under varying scenarios. Key mathematical models include the , which describe predator-prey oscillations via coupled differential equations: for prey \frac{dN}{dt} = rN - aNP and predators \frac{dP}{dt} = eaNP - dP, where N and P are population sizes, r is prey growth rate, a attack rate, e conversion efficiency, and d predator death rate; these reveal cycles driven by resource limitation but assume constant parameters, limiting realism in diverse ecosystems. Logistic growth models extend single-species dynamics to \frac{dN}{dt} = rN(1 - \frac{N}{K}), incorporating carrying capacity K to account for density-dependent regulation observed in empirical data from microbial and plant communities. Food web models, such as those using , analyze stability through metrics like connectance and interaction strength, with simulations showing that moderate complexity enhances resilience against species loss, as validated in lake and grassland experiments. Theoretical frameworks underpin these models by providing causal structures for interpretation. Systems ecology, pioneered by Eugene Odum in the 1950s, views ecosystems as self-regulating networks of energy throughput and material cycling, emphasizing steady-state fluxes measurable via isotopes and biomass accounting; this contrasts with reductionist population-focused approaches by prioritizing holistic feedbacks, though critics note overemphasis on equilibrium ignores transient disturbances evident in paleoecological records. Hierarchy theory posits ecosystems as nested scales from genes to biomes, where cross-scale interactions drive emergent properties like regime shifts, supported by analyses of coral reefs and forests showing abrupt changes from local stressors propagating upward. Metacommunity theory integrates dispersal and local dynamics via Levins' metapopulation models, predicting biodiversity maintenance through source-sink dynamics, with empirical confirmation in pond networks where patch connectivity correlates with species persistence rates exceeding 70% in high-dispersal scenarios. Validation of models relies on parameter estimation from field data and sensitivity analyses, revealing limitations such as parameter uncertainty amplifying errors in long-term predictions; for instance, ecosystem models for carbon cycling in boreal forests overestimate sequestration by 20-30% when ignoring microbial feedbacks, as shown in intercomparison projects. Bayesian frameworks enhance inference by incorporating prior empirical distributions, improving forecasts for invasive species impacts, though computational demands restrict application to simpler systems. These tools, while powerful for hypothesis generation, require empirical calibration to avoid artifacts from unverified assumptions, as demonstrated by discrepancies between simulated and observed tipping points in Arctic tundra transitions.

Recent Technological Advances in Study

Advances in remote sensing technologies have enhanced large-scale ecosystem monitoring, with satellites providing high-resolution data for tracking vegetation dynamics and biodiversity patterns. For instance, the Plankton, Aerosol, Cloud, ocean Ecosystem (PACE) mission, launched in 2024 by NASA, employs hyperspectral imaging to monitor phytoplankton communities, enabling finer detection of oceanic ecosystem changes compared to previous sensors. Similarly, constellations like those from Planet Labs offer daily imaging at 3-5 meter resolution, facilitating near-real-time assessment of deforestation and habitat fragmentation across terrestrial ecosystems. Unmanned aerial vehicles (UAVs or drones) have enabled high-resolution, fine-scale surveys inaccessible to satellites, particularly for biodiversity inventories. A 2025 review highlights their evolution in plant and animal ecology, including multispectral imaging for species identification and LiDAR for structural analysis of forest canopies, with applications detecting up to 90% of understory vegetation in tropical ecosystems. Drone-based remote sensing has advanced functional diversity monitoring in grasslands, integrating AI for automated classification of plant traits from aerial imagery. Environmental DNA (eDNA) metabarcoding has revolutionized noninvasive biodiversity assessment by detecting genetic material shed into water, soil, or air. Recent standardization efforts, including a 2024 U.S. national strategy, emphasize eDNA's role in aquatic monitoring, where it identifies species presence with sensitivity surpassing traditional surveys; for example, studies have detected rare fish species at concentrations as low as 1 individual per cubic meter of water. Advances in eDNA-RNA kits, detailed in a 2025 PeerJ special issue, extend applications to active biomass estimation in dynamic ecosystems. Artificial intelligence and machine learning algorithms process vast datasets from these sensors, improving predictive models of ecosystem responses. In 2024, AI integration with sensor networks enabled automated detection of species distributions and air quality impacts on ecosystems, achieving prediction accuracies exceeding 85% in landscape-scale analyses. Machine learning has also enhanced eDNA data interpretation, reducing false positives in metabarcoding pipelines through convolutional neural networks trained on reference genomes. These tools collectively address data overload, though challenges persist in model generalizability across ecosystems due to training data biases.

Human Influences

Ecosystem Services Framework

The ecosystem services framework provides a structured approach to identifying, categorizing, and valuing the contributions of ecosystems to human welfare, emphasizing the dependence of societies on natural processes. Developed within ecological economics, the framework traces its modern origins to early conceptualizations in the 1970s and 1980s, but it was advanced significantly by a 1997 peer-reviewed estimate placing the annual global value of these services at $16–54 trillion (in 1997 USD), far exceeding contemporary gross world product. This valuation drew on biophysical data from over 100 studies, highlighting services like pollination and water purification, though it faced methodological critiques for aggregating heterogeneous benefits without accounting for spatial variability or thresholds. The framework was further standardized in the 2005 Millennium Ecosystem Assessment, a synthesis involving over 1,300 experts that classified services into four main categories while documenting degradation in 60% of assessed services due to land-use changes and overexploitation. The four categories delineate distinct types of benefits:
  • Provisioning services, which supply material outputs such as food (e.g., fisheries yielding 100 million tons annually as of 2020), freshwater, timber, and fiber, directly supporting basic human needs but vulnerable to depletion, as seen in global fish stocks where 35% are overfished per 2022 FAO data.
  • Regulating services, including climate regulation (e.g., forests sequestering 2.6 billion tons of CO2 yearly), flood control, and pollination (valued at $235–577 billion annually for crops), which maintain environmental stability through biophysical processes like carbon cycling and habitat moderation.
  • Cultural services, encompassing non-material benefits like recreation (e.g., national parks generating $40 billion in U.S. visitor spending in 2023), aesthetic value, and spiritual fulfillment, often harder to quantify due to subjective preferences.
  • Supporting services, foundational processes such as nutrient cycling, soil formation, and primary production that underpin the other categories but are not directly consumed, with disruptions like soil erosion affecting 33% of global land as of 2015.
Valuation methods under the framework include market-based approaches for provisioning services (e.g., timber prices), revealed preference techniques like hedonic pricing for regulating services (e.g., property value premiums near green spaces, estimated at 5–20% in urban studies), and stated preference methods such as contingent valuation surveys, which elicit willingness-to-pay for cultural services. Benefit transfer extrapolates values from existing studies to new contexts, as in global syntheses assigning $125–145 trillion annually to terrestrial and marine biomes combined, though challenges persist: non-market services resist monetization without double-counting (e.g., valuing soil formation separately from crop yields), irreversibility of losses (e.g., species extinctions) defies discounting, and cultural biases in surveys undervalue services in non-Western contexts. Over 20 monetary techniques have been identified, but peer-reviewed meta-analyses reveal inconsistencies, with regulating services often undervalued relative to provisioning due to data scarcity. Despite its policy influence—evident in frameworks like the U.S. EPA's National Ecosystem Services Classification System, which aids impact assessments for regulations—the framework draws criticisms for conflating ecological functions (e.g., biodiversity maintenance) with human-derived services, potentially overlooking intrinsic ecosystem values independent of utility. It inadequately addresses ecosystem disservices (e.g., disease vectors from wetlands costing $10–50 billion yearly in health burdens) and trade-offs, such as agricultural intensification boosting provisioning at the expense of regulating services, with empirical studies showing inconsistent links to improved conservation outcomes. Academic assessments note incomplete scientific foundations, inconsistent application across scales, and risks of commodification that prioritize short-term economic gains over resilience, as critiqued in SWOT analyses of the approach. Sources advancing the framework, often from UN-affiliated assessments, may exhibit optimism biases toward market solutions, yet causal evidence remains limited, with policies invoking it sometimes failing to halt degradation due to enforcement gaps.

Disservices and Trade-Offs

Ecosystem disservices refer to the outputs of ecological processes that detrimentally affect human well-being, including economic losses, health risks, and physical hazards, often arising concurrently with beneficial services. Unlike services, disservices are not merely anthropogenic externalities but inherent ecosystem functions perceived as negative under specific human values, such as habitat provision for disease vectors or crop pests. A systematic review of 85 studies identified disservices primarily in agroecosystems (49%) and urban areas (35%), with only 13 studies rigorously linking them to measurable ecological interactions rather than subjective perceptions. Common examples include agricultural pest pressures, where increased biodiversity supports natural enemies of crops but also amplifies herbivore populations causing yield reductions; for instance, in southeastern South African arable systems, wildlife crop raiding led to documented economic losses. In urban settings, tree pollen and volatile organic compounds contribute to allergies and air quality degradation, with 91.3% of quantitative studies on urban trees confirming such effects alongside cooling benefits. Zoonotic disease risks exemplify regulating disservices, as diverse wildlife reservoirs heighten pathogen spillover, evidenced by higher Lyme disease incidence in fragmented forests versus homogeneous landscapes. Infrastructure damage from root growth or erosion, and cultural disservices like fear induced by large predators, further illustrate categories under frameworks like , which enumerates eight disservice types. Trade-offs between services and disservices arise from coupled ecological dynamics, where optimizing one often diminishes another; for example, wetland restoration enhances flood regulation and water purification but fosters mosquito breeding grounds, increasing vector-borne disease incidence. In agroecosystems, intensifying provisioning services through monocultures reduces pest-related disservices compared to diverse polycultures, though at the cost of soil regulation decline. Reforestation for carbon sequestration can elevate biogenic emissions of pollutants, as seen in urban forests where volatile organic compounds from trees offset air quality gains. Such synergies and antagonisms, quantified in 43% of reviewed studies acknowledging services, underscore the need for spatially explicit assessments, as small-scale enhancements may yield net benefits while landscape-scale interventions amplify disservices. Management failures, like the Macquarie Island rodent eradication in 2011 that inadvertently boosted vegetation damage via cascading effects, highlight risks of ignoring these interdependencies.
CategoryExample DisserviceAssociated Trade-OffSource
ProvisioningCrop damage by herbivoresBiodiversity for pest control vs. yield loss
RegulatingDisease vector habitatsFlood mitigation vs. pathogen spillover
CulturalPredator-induced fearEcotourism revenue vs. livestock predation
Conceptual ambiguities in defining disservices—often conflating ecological costs with personal nuisances—hinder quantification, with most research relying on proxies like stakeholder surveys rather than direct measurements. Empirical approaches, such as in situ surveys, reveal that disservices are context-dependent, varying by cultural values and management intensity, necessitating integrated social-ecological models for balanced policy. From 2009 to 2022, 524 peer-reviewed papers addressed disservices, predominantly from the U.S. (146 studies), indicating growing recognition but persistent gaps in causal attribution.

Positive Human Modifications

Human interventions aimed at restoring degraded ecosystems, such as reforestation, species reintroduction, and invasive species removal, have yielded measurable improvements in biodiversity and ecological function in multiple cases. A global meta-analysis of 92 restoration projects across terrestrial biomes reported an average 20% increase in biodiversity metrics, including species richness and abundance, compared to unrestored degraded sites, with effects consistent across forests, grasslands, and shrublands. These gains stem from causal mechanisms like enhanced habitat connectivity and reduced erosion, though long-term success depends on ongoing management to counter external pressures such as climate variability. In China's Loess Plateau, the Grain for Green Project, launched in 1999, transformed approximately 11,500 square miles of eroded cropland into forests and grasslands by 2016, achieving a 25% rise in vegetation cover and substantially lowering soil erosion rates from over 10,000 tons per square kilometer annually in untreated areas to under 1,000 tons in restored zones. This initiative boosted local biodiversity by fostering native plant recolonization and supporting faunal recovery, including increased bird and small mammal populations, while improving water retention in a region historically prone to flash floods and desertification. Peer-reviewed assessments confirm these outcomes, attributing them to terracing and afforestation that reversed decades of agricultural overuse, though some studies note trade-offs like temporary reductions in groundwater recharge from denser vegetation. The 1995 reintroduction of 31 gray wolves (Canis lupus) to initiated a trophic cascade by curbing elk (Cervus canadensis) numbers and altering their foraging patterns, which allowed riparian willow (Salix spp.) and aspen (Populus tremuloides) stands to regenerate in valleys like the Northern Range, with beaver (Castor canadensis) colonies expanding due to improved forage availability. Elk calf survival declined by up to 20% in wolf-occupied areas, indirectly benefiting understory vegetation and species like songbirds that rely on denser shrub cover. A 20-year analysis affirmed broader ecosystem benefits, including stabilized soil and enhanced carbon sequestration, though attribution debates highlight confounding factors like concurrent drought relief and reduced human hunting pressure post-park management changes. Rewilding initiatives elsewhere, such as beaver reintroduction in the UK's Rivers Trust projects since 2010, have created wetland mosaics that increased invertebrate diversity by 30-50% and supported fish and amphibian populations through dam-induced flooding that mimics natural hydrology. Similarly, bison (Bison bison) restoration on the American Great Plains has promoted grassland heterogeneity, elevating plant species richness by facilitating grazing patterns that prevent dominance by single grasses. These cases illustrate how targeted modifications can leverage keystone species to amplify positive feedbacks, but empirical monitoring is essential to verify persistence against invasive pressures or land-use intensification.

Negative Impacts and Attribution Debates

Human activities have inflicted substantial damage on ecosystems worldwide, primarily through habitat conversion, fragmentation, and degradation, which are the leading direct drivers of biodiversity loss. Land and sea use changes, such as deforestation, urbanization, and agricultural expansion, account for the majority of recent anthropogenic biodiversity declines, with empirical analyses indicating they dominate over other factors in global assessments spanning 1960 to 2019. For instance, these changes have contributed to an estimated 75% reduction in global terrestrial wildlife populations since 1970, alongside widespread ecosystem restructuring that diminishes resilience to perturbations. Other significant negative impacts include overexploitation of resources, pollution, and the introduction of invasive species, which exacerbate degradation across terrestrial, freshwater, and marine systems. Overfishing has depleted 33% of global fish stocks to unsustainable levels as of 2020, disrupting marine food webs and reducing ecosystem productivity. Pollution from plastics, nutrients, and chemicals has led to dead zones in over 400 coastal systems, covering more than 245,000 square kilometers by 2019, while invasive species, facilitated by trade and transport, have colonized 37% of assessed ecosystems, altering native community structures. These impacts often interact synergistically; for example, nutrient runoff from agriculture not only eutrophies waters but also amplifies hypoxia in conjunction with warming temperatures. Attribution debates focus on quantifying the relative contributions of these drivers, particularly the longstanding primacy of land-use change versus the rising influence of climate change. Observational data consistently attribute most historical biodiversity loss to direct habitat alteration rather than climatic shifts, with land/sea use change identified as the dominant factor in 80% of assessed terrestrial and 50% of marine cases. Projections from integrated models, however, suggest climate change could surpass land use as the primary driver of biodiversity decline by mid-century under moderate emissions scenarios, potentially exacerbating range shifts and phenological mismatches in 15-37% of species. Challenges in attribution arise from confounding factors and methodological limitations, including the difficulty of disentangling synergistic effects and natural variability from anthropogenic signals. For example, while event-attribution studies link specific ecosystem disruptions, such as coral bleaching events in 2014-2017, to anthropogenic warming with high confidence, broader degradation patterns often reflect multiple baselines, with models sometimes overemphasizing climate due to assumptions about future land-use stabilization. Critics argue that institutional biases in academic and intergovernmental assessments, such as those from , may underweight direct exploitation in favor of climate narratives to align with policy priorities, despite empirical primacy of habitat loss in peer-reviewed syntheses. Definitive causal parsing requires longitudinal data and controlled experiments, which remain sparse for complex systems.

Management and Intervention

Conservation and Policy Strategies

Conservation strategies for ecosystems primarily involve establishing protected areas, implementing international agreements, and deploying incentive-based mechanisms to align human activities with ecological preservation. The , ratified by 196 parties since 1992, sets objectives for conserving biological diversity, promoting sustainable use of its components, and ensuring fair benefit-sharing from genetic resources. Its , adopted in 2022, targets protecting 30% of terrestrial and marine areas by 2030, alongside restoring degraded ecosystems and reducing biodiversity loss drivers. National policies often operationalize these through legislation designating reserves and regulating land use, with empirical analyses showing public policies globally reduce tree cover loss risk by approximately 4 percentage points, though effects vary by region and enforcement. Protected areas constitute a core policy tool, covering 17.5% of Earth's land and inland waters and 8.5% of oceans as of August 2024, per assessments aligned with CBD targets. Effectiveness studies indicate these areas curb habitat loss 33% more than comparable unprotected lands, yet mitigation of adjacent human pressures like agriculture remains limited, with only 33% of protected areas achieving high conservation outcomes in integrated evaluations. Rigorous evidence highlights mixed results, as coverage alone does not guarantee biodiversity persistence without addressing poaching, encroachment, or climate shifts, and some designations fail due to inadequate management or local opposition. Incentive programs, such as payments for ecosystem services (PES), compensate landowners for forgoing deforestation or adopting sustainable practices, with randomized trials in Uganda and Ecuador demonstrating 41% deforestation reductions under optimized designs compared to standard contracts. Global reviews of PES schemes reveal predominant focus on outcomes like carbon sequestration, but persistent challenges include funding shortfalls—e.g., Costa Rica's program saw declines post-2015—and uncertain long-term behavioral changes after payments lapse. Empirical debates persist on land management paradigms, with neither "land sparing" (intensified agriculture freeing habitat) nor "land sharing" (wildlife-friendly farming) consistently outperforming the other in biodiversity-agriculture trade-offs. Policy implementation faces evidentiary gaps and causal attribution issues, as many initiatives lack randomized controls or longitudinal data to isolate effects from confounding factors like economic shifts. Conservation failures often stem from overlooking local socio-ecological contexts, insufficient monitoring, or overreliance on unverified assumptions, underscoring the need for adaptive, evidence-driven adjustments rather than static targets. Despite progress in formal designations, systemic underfunding and enforcement inconsistencies—exacerbated by biases in academic reporting favoring positive outcomes—hinder verifiable causal impacts on ecosystem resilience.

Restoration Techniques and Outcomes

Ecological restoration techniques encompass active interventions, such as planting native vegetation, soil amendment, and species reintroduction, alongside passive strategies like fencing to exclude herbivores or halting disturbances to permit natural regeneration. These methods aim to reinstate ecosystem structure, function, and composition degraded by human activities including , agriculture, and urbanization. Success is typically measured by metrics such as biodiversity recovery, carbon sequestration rates, soil stability, and hydrological function restoration, though outcomes depend on site-specific factors like climate, degradation extent, and underlying causal drivers such as altered hydrology or nutrient imbalances. Meta-analyses of terrestrial restoration projects reveal average biodiversity gains of 20% relative to unrestored degraded sites, with reduced variability in ecological indicators across interventions. Forest restoration enhances methane uptake by 90% and grassland by 31%, primarily through improved soil properties like organic matter content and microbial activity. In drylands, active techniques like seeding and mulching outperform passive exclusion in vegetation cover recovery, achieving up to 50% higher biomass accumulation within five years, though cost-effectiveness diminishes in arid conditions due to high seedling mortality rates exceeding 70% without irrigation. Grassland restoration via direct seeding of native species increases plant richness by 15-25% more than grazing cessation alone, as evidenced by systematic reviews of 100+ studies spanning 1980-2020. Passive restoration often yields superior long-term outcomes in forests, with natural regeneration restoring biodiversity and vegetation structure 34-56% more effectively than active planting, based on data from 190 tropical sites monitored over 5-20 years. This disparity arises because active methods frequently fail to replicate successional dynamics, leading to lower survival rates (e.g., 20-40% for planted trees in nutrient-poor soils) and homogenized communities vulnerable to pests. Species reintroduction, as in the 1995 Yellowstone wolf program, has demonstrated cascading benefits, increasing beaver populations by 150% and riparian willow growth by 30% through trophic cascades, though such successes require intact prey bases and minimal human interference. Mangrove restoration projects, analyzed across 50+ global sites, deliver 1.5-2 times higher ecosystem services (e.g., fish production and coastal protection) than unvegetated mudflats, with survival rates averaging 60-80% when hydrological connectivity is restored. Marine ecosystem restoration exhibits an average success rate of 64%, with coral reefs achieving 70-80% recovery in structural complexity via larval propagation and substrate stabilization, per meta-analysis of 200+ interventions from 2000-2023. However, methodological heterogeneity in monitoring—such as short-term (1-5 year) versus long-term assessments—can inflate reported successes by 10-20%, as initial gains often erode without addressing or . Wetland restorations, like those in the U.S. since 2000, have restored 40% of pre-drainage hydrology and bird populations increased by 25%, but persistent invasive species and phosphorus loading limit full functional recovery to under 50% of reference states after two decades. Overall, while restorations mitigate degradation, full equivalence to undisturbed ecosystems remains rare, with only 10-30% of projects attaining >80% functional recovery, underscoring the need for causal diagnosis of degradation drivers prior to intervention.

Economic and Ethical Considerations

Ecosystems underpin economic activities through services such as , water regulation, and , with global estimates placing their annual value at approximately $125–145 trillion in equivalent , though much of this remains unmonetized due to non-market nature. These services support over half of global GDP, roughly $44 trillion in 2023, by enabling sectors like and fisheries. Degradation, including and habitat conversion, incurs costs estimated at $2.7 trillion annually if key services like wild collapse, per analysis, with broader nature-related risks amplifying financial vulnerabilities up to $5 trillion in exposed assets. Payments for ecosystem services (PES) represent a market-based approach to internalize these values, compensating landowners for maintaining habitats that yield benefits like watershed protection. In , the Grain-to-Green Program (GTGP) from 1999 onward converted cropland to forests, reducing and boosting forest cover by over 20 million hectares, though it sometimes displaced food production. Evaluations of PES schemes globally show mixed outcomes: short-term implementations achieve about 46% success in goals, with positive effects on livelihoods and land-use stability when is high, but failures arise from weak or misaligned incentives. Such mechanisms highlight trade-offs, as prioritizing one service (e.g., carbon storage) may undermine others (e.g., ), necessitating empirical assessment over idealized models. Ethically, ecosystem management grapples with anthropocentric views, which assign value based on human utility—such as resource provision—versus biocentric or ecocentric perspectives that posit intrinsic worth in living organisms or systems independent of human benefit. Anthropocentric frameworks dominate policy, justifying interventions like habitat restoration for sustained yields, but critics argue they undervalue non-utilitarian elements, potentially leading to . Biocentric ethics, emphasizing equal moral consideration for all life forms, underpin arguments for limiting human expansion, yet face practical challenges in implementation, as prioritizing individual organisms can conflict with ecosystem-level dynamics like predation. Emerging "" doctrines grant ecosystems legal standing to sue for self-preservation, as in Ecuador's 2008 constitution and subsequent cases enforcing river protections, or New Zealand's 2017 Te Awa Tupua Act recognizing a river's juridical . These precedents aim to counter anthropocentric biases by treating nature as a rights-holder, but outcomes reveal limitations: enforcement often falters due to anthropocentric judicial interpretations, and successes are sporadic, with no broad evidence of reversing degradation trends. Such approaches risk symbolic overreach without causal mechanisms for accountability, underscoring the need for evidence-based policies over declarative ethics.

References

  1. [1]
    Ecosystems | U.S. Geological Survey - USGS.gov
    An ecosystem is made up of all of the living and non-living things in a geographic area. Plants, animals, soil, climate, and other features of the environment ...
  2. [2]
    The Ecosystem and How It Relates To Sustainability
    What is an Ecosystem? An ecosystem consists of the biological community that occurs in some locale, and the physical and chemical factors that make up its non- ...
  3. [3]
    The creation of “Ecosystem Core” hypothesis to explain ... - NIH
    Sep 6, 2019 · An ecosystem is a community of living organisms in conjunction with the nonliving components of their environment [12]. The abiotic components ...
  4. [4]
    Ecosystem - National Geographic Education
    Apr 23, 2025 · Ecosystems contain biotic (living) factors, as well as abiotic (nonliving) factors. Biotic factors include plants, animals and other organisms.
  5. [5]
    Ecosystems - Science Learning Hub
    Jul 19, 2007 · An ecosystem must contain producers, consumers, decomposers, and dead and inorganic matter. All ecosystems require energy from an external ...
  6. [6]
    [PDF] Energy Flow, Nutrient Cycling, and Ecosystem Resilience Author(s)
    Studies of nutrient cycling models have shown that resilience increases as the mean number of cycles that nutrient (or other mineral) atoms make before leaving ...Missing: peer | Show results with:peer
  7. [7]
    Moving toward a new era of ecosystem science - ScienceDirect
    Ecosystem is at the center of the ecological structure: individual, population, community, ecosystem, landscape, region, and biosphere and is the core component ...Moving Toward A New Era Of... · 3. Current Research Progress... · 4. Ecosystem Science In The...
  8. [8]
    Strong nutrient-plant interactions enhance the stability of ecosystems
    Oct 20, 2021 · Their study concludes that strong nutrient-plant interactions increase ecosystem stability, contrasting the notion that weak consumer-resource interactions ...
  9. [9]
    Main Groups of Ecosystems: Terrestrial vs Aquatic Explained Simply
    Jun 21, 2025 · The whole planet's living spaces fall into just two main groups: terrestrial (on land) and aquatic (in water).
  10. [10]
    Major Ecosystem Types | Understanding Earth's Diverse Habitats
    Oct 27, 2024 · Terrestrial ecosystems are found on land and include forests, deserts, grasslands, and tundras. Aquatic ecosystems are found in water and ...
  11. [11]
    The Origin of Ecosystems by Means of Subjective Selection
    “Ecosystem” was defined in 1935 by Sir Arthur Tansley as: the whole system (in the sense of physics) including not only the organism-complex, but also the ...
  12. [12]
    Tansley, A.G. 1935: The use and abuse of vegetational concepts ...
    Aug 9, 2025 · Arthur Tansley first defined the term ecosystem in his seminal work “Use and Abuse of Vegetational Concepts,” as an improved way of viewing the ...
  13. [13]
    The Ecosystem: An Evolving Concept Viewed Historically - jstor
    The word 'ecosystem' was first used in print by A. G. Tansley (1935) in his well-known paper on vegeta- tional concepts and terms. Tansley considered that.
  14. [14]
    [PDF] Ecosystems
    The term "ecosystem" was first coined by Roy Clapham in 1930, but it was ecol- ogist Arthur Tansley who fully defined the ecosystem concept. In a classic ...
  15. [15]
    Biotic/Abiotic - MSU College of Agriculture and Natural Resources
    Biotic factors are living things within an ecosystem; such as plants, animals, and bacteria, while abiotic are non-living components; such as water, soil and ...
  16. [16]
    [PDF] Reading - What is an Ecosystem?
    We define an ecosystem as a community of living things that interact with each other and with the physical environment. So, the saltmarshes along the coastline ...<|control11|><|separator|>
  17. [17]
    Chapter 4 Ecological Principles and Concepts - ScienceDirect.com
    In an ecosystem two major processes are in operation: an unidirectional flow of energy and chemical cycling of important elements. Ecosystems are not static, ...<|separator|>
  18. [18]
    Principles of Ecology – Environmental Biology
    Ecology is the study of how living things interact with each other and with their environment. It is a major branch of biology, but has areas of overlap with ...
  19. [19]
    The Use and Abuse of Vegetational Concepts and Terms - Tansley
    Volume 16, Issue 3 pp. 284-307 Ecology Article Full Access The Use and Abuse of Vegetational Concepts and Terms AG Tansley, AG TansleyMissing: Arthur | Show results with:Arthur
  20. [20]
    [PDF] The ecosystem: an evolving concept viewed historically Early ...
    Tansley considered that organisms, when thinking is fundamental, cannot be separated from 'the environment of the biome – the habitat factors in the widest ...
  21. [21]
    Tansley - an overview | ScienceDirect Topics
    Tansley (1935) coined the term “ecosystem” to recognize the integration of the biotic community and its physical environment as a fundamental unit of ecology.
  22. [22]
    A History of the Ecosystem Concept in Ecology - Yale University Press
    Feb 21, 1996 · In this book an eminent ecologist explains the ecosystem concept, tracing its evolution, describing how numerous American and European researchers contributed ...<|separator|>
  23. [23]
    Units of nature or processes across scales? The ecosystem concept ...
    Feb 5, 2011 · The ecosystem has served as a central organizational concept in ecology for nearly a half century and continues to evolve.
  24. [24]
    Biotic and abiotic factors | Research Starters - EBSCO
    Biotic factors refer to all living organisms within an ecosystem, including plants, animals, and microorganisms. These organisms fulfill critical roles as ...
  25. [25]
    simple DEB-based ecosystem model | Conservation Physiology
    Aug 6, 2022 · The ecosystem contains nutrients, producers, consumers, decomposers and detritus. All three living groups consist of somatic structure and ...Missing: peer- | Show results with:peer-
  26. [26]
    Ecological relationships | Ask A Biologist - Arizona State University
    Sep 30, 2025 · Predation is where one organism (the predator) kills and eats another organism (called the prey). The cheetah is the predator of the impala.
  27. [27]
    Ecological Communities: Networks of Interacting Species
    Species interactions within ecological webs include four main types of two-way interactions: mutualism, commensalism, competition, and predation (which includes ...Missing: symbiosis | Show results with:symbiosis
  28. [28]
    Population & Community Ecology
    Predation involves one species (the predator) consuming the other (the prey), which has the prey die in the exchange. A well-studied example is the lynx-hare ...
  29. [29]
    Ecosystem Dynamics: Exploring Types, Components and the Forces ...
    Aug 8, 2025 · Key biotic (producers, consumers, decomposers) and abiotic (climate, soil, water) factors regulate ecosystem dynamics. Human-induced changes ...
  30. [30]
    19. Symbiosis - SUNY Cortland
    There are further types of symbiosis as well, such as predation, amensalism, antagonism and competition. The symbiosis detailed here is between common milkweed ...Missing: examples | Show results with:examples
  31. [31]
    Symbiosis and Other Relationships - Genetic Science Learning Center
    Through examples, this lesson helps students learn about different types of ecological relationships, including competition, predation, parasitism, commensalism ...Missing: interactions | Show results with:interactions
  32. [32]
    Species Interactions – An Interactive Introduction to Organismal and ...
    Food webs allude to predation interactions and if species are eating the same species, then they are likely competing for the food source.
  33. [33]
    Abiotic/Biotic - Kids Do Ecology
    An ecosystem includes all the abiotic (non-living) factors in addition to the community of biotic (living) things that exists in a certain area. Some ideas ...
  34. [34]
    [PDF] Abiotic Components of a Freshwater System
    Water temperature, pH, phosphate and nitrogen levels, dissolved oxygen, and substrate composition are some of the abiotic factors to consider and measure. ...Missing: definition | Show results with:definition
  35. [35]
    Precipitation and temperature regulate species diversity, plant ...
    Nov 30, 2022 · Our results highlight that precipitation and temperature are two key climatic drivers of species richness, evenness, plant coverage and AGB through complex ...
  36. [36]
    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 ...
  37. [37]
    Impact of climate change on wetland ecosystems: A critical review of ...
    May 15, 2021 · In addition, the impact of key climatic factors such as temperature and water availability on wetlands has been reviewed.
  38. [38]
    Determining the effects of biotic and abiotic factors on the ecosystem ...
    The main factors influencing ecosystem multifunctionality at the community level were soil moisture and functional diversity, whereas those influencing the ...
  39. [39]
    Environmental drivers of increased ecosystem respiration in ... - Nature
    Apr 17, 2024 · These drivers included (1) changes in abiotic and biotic conditions induced by the warming treatment (that is, climatic, soil, vegetation and ...
  40. [40]
    Influences of temperature and moisture on abiotic and biotic soil CO ...
    May 25, 2021 · Results showed that abiotic processes accounted for a considerable proportion (15.6−60.0%) of CO 2 emission in such a biologically active soil under different ...
  41. [41]
    Terrestrial ecology - Soil Ecology Wiki
    May 2, 2025 · Abiotic factors are the unique aspects of the climate, the weather, the type of soil in the region, and the water within the ecosystem. An ...
  42. [42]
    Global patterns and abiotic drivers of ecosystem multifunctionality in ...
    We find that climatic and soil factors jointly drive the EMF in low precipitation areas, and climatic factors dominate the EMF in high precipitation regions.
  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 ...<|separator|>
  44. [44]
    Biotic and abiotic drivers of ecosystem multifunctionality - PubMed
    May 13, 2023 · We conclude that soil texture and plant diversity, representing crucial abiotic and biotic factors, respectively, are important determinants of the EMF of ...Missing: influence | Show results with:influence
  45. [45]
    Energy Flow - an overview | ScienceDirect Topics
    The flow of energy in ecosystems is from energy source, to autotroph, to heterotroph. For most systems, the energy source is the sun and the autotrophs are ...
  46. [46]
    Energy Flow in Ecosystems: A Historical Review - Oxford Academic
    Aug 1, 2015 · Abstract. A generalized model of energy flow applicable both to individual populations and food chains is discussed.
  47. [47]
    The Trophic‐Dynamic Aspect of Ecology - Lindeman - 1942
    Volume 23, Issue 4 pp. 399-417 Ecology Article Full Access The Trophic-Dynamic Aspect of Ecology Raymond L. Lindeman, Raymond L. Lindeman
  48. [48]
    The Trophic-Dynamic Aspect of Ecology - jstor
    October, 1942 TROPHIC-DYNAMIC ASPECT OF ECOLOGY 401. Solar Radiation Solar Radiation. A. External. Dissolved Nutrients N j. Internal. A, PhytoplanKters ...
  49. [49]
    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.
  50. [50]
    Trophic levels and energy loss (article) | Khan Academy
    In general, only 10% of energy is transferred from one trophic level to the next. In this diagram, energy is represented by joules, which are units of energy.
  51. [51]
    Food Web: Concept and Applications | Learn Science at Scitable
    Food webs illustrate energy flow from primary producers to primary consumers (herbivores), and from primary consumers to secondary consumers (carnivores).
  52. [52]
    Energy flow across soil food webs of different ecosystems
    The results indicate that mature and complex soil communities support high energy flux across soil food webs.
  53. [53]
    Soil Biogeochemical Cycle Couplings Inferred from a Function ...
    Mar 10, 2021 · Soil biogeochemical cycles are the foundation of ecosystem function and affect nutrient and energy flows that regulate productivity within both ...
  54. [54]
    Stability of the marine nitrogen cycle over the past 165 million years
    Oct 9, 2025 · The operation of the N cycle is clearly dependent upon global oceanic and orogenic processes, even as biological processes mediate between them.
  55. [55]
    Competing and accelerating effects of anthropogenic nutrient inputs ...
    Jul 1, 2022 · Anthropogenic nutrient inputs increase oceanic productivity and carbon uptake, offsetting climate-induced decrease and accelerating climate-driven ...
  56. [56]
    Activation Energy of Organic Matter Decomposition in Soil and ... - NIH
    Sep 4, 2025 · We reviewed the activation energies of SOM decomposition and mineralization as well as of enzymatic reactions and thermal oxidation published in ...
  57. [57]
    Plant roots increase both decomposition and stable organic matter ...
    Sep 4, 2019 · We provide a framework on the role of plant roots in SOM decomposition and stabilization based on a field experiment in boreal forest soils (Fig ...
  58. [58]
    Linking effect traits of soil fauna to processes of organic matter ...
    Dec 20, 2024 · Soil organic matter (SOM) transformation processes are regulated by the activities of plants, microbes, and fauna. Compared with plants and ...
  59. [59]
    The impact of insect herbivory on biogeochemical cycling in ... - Nature
    Jul 17, 2024 · We show more nutrient release by insect herbivores at non-outbreak levels in tropical forests than temperate and boreal forests, that these fluxes increase ...<|separator|>
  60. [60]
    The Biological Productivity of the Ocean | Learn Science at Scitable
    "Net primary production" (NPP) is GPP minus the autotrophs' own rate of respiration; it is thus the rate at which the full metabolism of phytoplankton produces ...Missing: types | Show results with:types
  61. [61]
    A global database of net primary production of terrestrial ecosystems
    Sep 2, 2025 · The net primary production (NPP) of an ecosystem is the remainder of the photosynthetic gain, or gross primary production (GPP), after ...Methods · Npp Data · Biome And Management...
  62. [62]
    Biomes - Synthesis and Discussion of NPP
    At the highest ranges, tropical rainforests, marshes, and agricultural fields (sugar cane and rice) productivities may be 3,000 g m-2 yr-1 . As temperatures ...Missing: type | Show results with:type
  63. [63]
    [PDF] a constant fraction of gross primary production?
    In the survey, total NPP ranged from 120 to 1660 g C rr1 -2 year -1 , whereas the calculated fraction allocated to roots varied from 0.22 to 0.63. Comparative ...
  64. [64]
    A function-based typology for Earth's ecosystems | Nature
    Oct 12, 2022 · Ecosystem functioning not only underpins biomass production, but also depends on and regulates the stocks and fluxes of resources, energy and ...
  65. [65]
    Joint control of terrestrial gross primary productivity by plant ...
    Terrestrial gross primary productivity (GPP), the total photosynthetic CO2 fixation at ecosystem level, fuels all life on land.
  66. [66]
    The Ecology of Carrion Decomposition | Learn Science at Scitable
    Decomposition of animal carrion is achieved primarily through the activities of invertebrates, such as flies and beetles, and large scavengers.
  67. [67]
    Decomposer diversity increases biomass production and shifts ...
    Dec 17, 2018 · Decomposers play distinct roles in the soil by facilitating different steps of decomposition processes, ranging from litter fragmentation to ...Missing: recycling | Show results with:recycling
  68. [68]
    A meta-analysis on decomposition quantifies afterlife effects of plant ...
    Sep 11, 2020 · Note that litter decomposition is primarily controlled by three factors; climate (e.g., temperature and moisture), decomposers, and litter ...Missing: recycling | Show results with:recycling
  69. [69]
    Role of litter production and its decomposition, and factors affecting ...
    May 7, 2020 · Litter decomposition provides readily available nutrients to plants because it incorporates organic carbon into soil through nutrient cycling ...Anthropogenic And Natural... · Litter Quality · Environment, Climate Factors...
  70. [70]
    [PDF] Decomposition processes: modelling approaches and applications
    Decomposition is a fundamental ecosystem process, comparable to primary production, that influences dynamics by releasing nutrients. It is a complex process.
  71. [71]
    Litter decomposition and nutrient release are faster under secondary ...
    Oct 5, 2023 · Litter decomposition plays a crucial role in regulating nutrient cycling and organic matter turnover within forest ecosystems. At a local scale, ...Missing: recycling | Show results with:recycling
  72. [72]
    1.6 Types of Feedbacks and Their Effects | EME 807
    Negative feedbacks. Consider this example. The population of deer in the area leads to a higher rate of road collisions. The collisions kill a certain number of ...
  73. [73]
    Feedback loops: Regulating natural processes - Eco-intelligent
    Dec 4, 2016 · Negative feedback loop. Positive feedback, on the other hand, increase the initial forcing. Take the example of heavy rains causing soil erosion ...
  74. [74]
    Feedbacks in ecology and evolution - ScienceDirect.com
    Organisms can modify their environment and these modifications can feed back to the organism, generating emergent properties with evolutionary consequences.Missing: internal | Show results with:internal<|separator|>
  75. [75]
    Empirical evidence for recent global shifts in vegetation resilience
    Apr 28, 2022 · Here, we quantify vegetation resilience globally with complementary metrics based on two independent long-term satellite records.
  76. [76]
  77. [77]
    No positive effects of biodiversity on ecological resilience of lake ...
    Apr 2, 2024 · For instance, a meta-analysis study and microcosm experiment showed that biodiversity increases ecosystem resistance but decreases resilience in ...
  78. [78]
    Special feature: measuring components of ecological resilience in ...
    Jan 27, 2021 · Studies that investigate the patterns and underlying drivers behind ecosystem resilience have become common across all areas of ecology ...
  79. [79]
    Reconciling the ecological and engineering definitions of resilience
    Feb 23, 2021 · In this definition resilience is the property of the system and persistence or probability of extinction is the result” (Holling 1973). Holling ...
  80. [80]
    Resilience, Adaptability and Transformability in Social–ecological ...
    Sep 16, 2004 · Three related attributes of social–ecological systems (SESs) determine their future trajectories: resilience, adaptability, and transformability.
  81. [81]
    [PDF] Disturbance regimes and the historical range and variation in ...
    It is far more illustrative to present the concept of disturbance regimes with an example from one of the world's most ubiquitous disturbances – wildland fire ( ...
  82. [82]
    Disturbance theory for ecosystem ecologists: A primer - PMC
    May 30, 2024 · Disturbances affect every scale and level of biological organization. However, disturbance studies are generally guided by discipline‐specific ...
  83. [83]
    Disturbance regimes and the historical range and variation in ...
    Disturbances are major drivers of ecological dynamics and it is the cumulative effects of disturbances across space and time that define a disturbance regime ...
  84. [84]
    Fire Regime - an overview | ScienceDirect Topics
    Fire regimes refer to the spatiotemporal patterns of multiple fires in an ecosystem, characterized by factors such as fire type, frequency, intensity, ...
  85. [85]
    [PDF] Disturbance and Temporal Dynamics
    The incidence of fire is one of many interacting sources of change in this ecosystem. 3.2 Disturbance and disturbance regime descriptors. Not all disturbances ...<|separator|>
  86. [86]
    Disequilibrium and complexity across scales: a patch-dynamics ...
    May 6, 2023 · Importantly, patch-dynamics frameworks can show how natural disturbances produce different types of patches in natural ecosystems, they take ...<|control11|><|separator|>
  87. [87]
    The Ecology of Natural Disturbance and Patch Dynamics
    The Ecology of Natural Disturbance and Patch Dynamics brings together the findings and ideas of those studying varied systems.
  88. [88]
    Changing disturbance regimes, ecological memory, and forest ...
    ... disturbance and is maintained by two types of legacies – information and material ... ecosystem resilience framework with examples of forest disturbances,...
  89. [89]
    How disturbance history alters invasion success: biotic legacies and ...
    Disturbance type. The physical nature of a disturbance event. Different disturbances such as fires and floods may have similarities in their effects, but each ...
  90. [90]
    Biodiversity mediates ecosystem sensitivity to climate variability - PMC
    Jun 27, 2022 · We show that, across multiple biomes, regions of greater plant diversity exhibit lower sensitivity (more stable over time) to temperature variability.
  91. [91]
    Patch dynamics and environmental heterogeneity in lotic ecosystems
    We reviewed concepts of patch dynamics and environmental heterogeneity and their applications to the study of fluvial ecosystems, with emphasis on research ...
  92. [92]
    A theory of pulse dynamics and disturbance in ecology - ESA Journals
    Apr 24, 2019 · The phrase “patch dynamics” has been used to describe ecosystem pattern and process at multiple scales in the disturbance ecology literature ( ...
  93. [93]
    Ecological succession, explained - UChicago News
    Ecological succession is the process by which the mix of species and habitat in an area changes over time. Gradually, these communities replace one another ...
  94. [94]
    Primary and Secondary Succession Mediate the Accumulation of ...
    Aug 18, 2020 · Primary succession is a type of biological succession in native bare systems in which organisms have never grown or have grown but were wiped ...
  95. [95]
    [PDF] Factors Influencing Succession: Lessons from Large, Infrequent ...
    Abundance and spatial arrangement of survivors and arrival patterns of propagules may be the pivotal factors determining how succession differs between intense.
  96. [96]
    History of Ecological Sciences, Part 54: Succession, Community ...
    Jul 1, 2015 · Clements' early publications were empirical studies, in which theory was inconspicuous. He more than kept abreast of Cowles' publications ...
  97. [97]
    [PDF] Revisiting Clements and Gleason - University of Colorado Boulder
    Do the vegetation patterns seen by Clements match his views on communities? Can we empirically distinguish between the two paradigms? We explore these questions.
  98. [98]
    Ecological non-equilibrium and biological conservation
    This is the opposite that happens, for example, in the more gradual and predictable changes that take place in ecological succession. ... Long-term change ...
  99. [99]
    Contrasting patterns of plant and microbial diversity during long ...
    Dec 24, 2018 · ECOLOGICAL SUCCESSION IN A CHANGING WORLD ... Despite the consistent response of above- and below-ground communities to long-term change ...
  100. [100]
    Long-term changes in avian biomass and functional diversity within ...
    Aug 17, 2022 · Long-term change in the avifauna of undisturbed Amazonian rainforest ... Avian ecological succession in the Amazon: a long-term case ...
  101. [101]
    [PDF] Fire exclusion and climate change interact to affect long-term ...
    May 1, 2017 · These communities thus appear to be undergo- ing ecological succession, favouring plant adaptions to better harvest light and carbon in ...<|separator|>
  102. [102]
    Integrating succession and community assembly perspectives - NIH
    Sep 12, 2016 · Classic succession studies involved studying primary and secondary succession after a disturbance using experiments, conceptual and ...
  103. [103]
    Disturbance theory for ecosystem ecologists: A primer - Gough - 2024
    May 30, 2024 · ... long-term change; and third, which disturbance regimes and sources ... Indeed, a Google search (10-19-22) of “ecological succession ...
  104. [104]
    Reconciling the concepts and measures of diversity, rarity and ...
    A biodiversity measure is a calculation method expressed by a mathematical formula that allows specific values for the amount of variety in a biological system ...
  105. [105]
    A conceptual guide to measuring species diversity - Roswell - 2021
    Feb 9, 2021 · Three metrics of species diversity – species richness, the Shannon index and the Simpson index – are still widely used in ecology, ...Diversity metrics · Which Hill diversity to use? · Standardizing samples, then...
  106. [106]
    Perceived biodiversity: Is what we measure also what we see and ...
    Jul 7, 2025 · Thus, there is no single measure of biodiversity and each measure is only a proxy for the 'true' biodiversity it is supposed to describe. The ...
  107. [107]
    Shannon Diversity Index: Definition & Example - Statology
    – The Shannon index incorporates both species richness (the number of different species) and species evenness (the relative abundance of each species).
  108. [108]
    22.2: Diversity Indices - Biology LibreTexts
    Oct 28, 2024 · Shannon-Weiner Index. Another widely used index of diversity that also considers both species richness and evenness is the Shannon-Weiner ...Diversity Indices · Simpson's Index · Shannon-Weiner Index · Evenness Index
  109. [109]
    Characterizing Communities | Learn Science at Scitable - Nature
    Species evenness is highest when all species in a sample have the same abundance. Evenness approaches zero as relative abundances vary.
  110. [110]
    Why it matters how biodiversity is measured in environmental ...
    We find a high degree of overlap in the broad biodiversity metrics used in environmental valuation compared to conservation and ecology.
  111. [111]
    A deep-time perspective on the latitudinal diversity gradient - PNAS
    Jul 15, 2020 · Today, species richness is highest in the tropics and declines toward the poles. Although there are exceptions, this pattern is pervasive ...
  112. [112]
    Strong evidence for latitudinal diversity gradient in mosses across ...
    Our results show that moss species richness decreases strongly with increasing latitude, regardless of whether the globe is considered as a whole or different ...
  113. [113]
    Biodiversity: Concepts, Patterns, Trends, and Perspectives
    We synthesize aspects of this literature, focusing on several key concepts, debates, patterns, trends, and drivers. We review the history of the term and the ...
  114. [114]
    The latitudinal biodiversity gradient through deep time - ScienceDirect
    The present-day decrease in biological diversity from equatorial to polar regions (Figure 1) is a pervasive pattern governing the distribution of life on Earth.
  115. [115]
    Explanations for latitudinal diversity gradients must invoke rate ... - NIH
    Growing evidence suggests that low latitude speciation may have contributed to the formation of LDGs today, with perhaps even larger effects on LDG ...
  116. [116]
    Effect of Productivity on Community Size Explains the Latitudinal ...
    AbstractAlthough many studies have shown that species richness increases from high to low latitudes (the latitudinal diversity gradient), the mechanisms ...<|control11|><|separator|>
  117. [117]
    Process-explicit models reveal the structure and dynamics ... - Science
    Aug 5, 2022 · The patterns of biodiversity we observe at different temporal and spatial scales result from the key evolutionary and ecological processes of ...
  118. [118]
    The geography of climate and the global patterns of species diversity
    Sep 27, 2023 · Our findings suggest that larger and more isolated climate conditions tend to harbour higher diversity and species turnover among terrestrial tetrapods.
  119. [119]
    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 ...
  120. [120]
    The global human impact on biodiversity - Nature
    Mar 26, 2025 · We show that human pressures distinctly shift community composition and decrease local diversity across terrestrial, freshwater and marine ecosystems.
  121. [121]
    Effects of Plant Biodiversity on Population and Ecosystem Processes
    This experiment, often called the "Big" Biodiversity Experiment, determines effects of plant species numbers and functional traits on community and ecosystem ...
  122. [122]
    Sustaining multiple ecosystem functions in grassland communities ...
    We analyzed data from the longest-running biodiversity-ecosystem functioning manipulation to date at the Cedar Creek Ecosystem Science Reserve in Minnesota (15) ...Missing: BIODEPTH | Show results with:BIODEPTH
  123. [123]
    Biodiversity, ecosystem function and plant traits in mature and ...
    May 5, 2005 · Here we compare the relationship between diversity and biomass in BIODEPTH with that at the long-term monitoring site at Bibury in ...
  124. [124]
    Mechanistic links between biodiversity effects on ecosystem ...
    Jun 8, 2021 · Specifically, biodiversity effects on ecosystem functioning were partitioned into complementarity effects (CE) and selection effects (SE), and ...
  125. [125]
    The multiple-mechanisms hypothesis of biodiversity–stability ...
    We hypothesize that multifunctional stability is highest in high-diversity plant communities and that biodiversity–stability relationships increase over time.
  126. [126]
    Scaling‐up biodiversity‐ecosystem functioning research - PMC
    Meta‐analyses of hundreds of BEF experiments have shown consistent relationships between biodiversity and ecosystem functioning across different ecosystem ...
  127. [127]
    Biodiversity as insurance: from concept to measurement and ... - NIH
    Jun 2, 2021 · Biological insurance theory predicts that, in a variable environment, aggregate ecosystem properties will vary less in more diverse ...
  128. [128]
    Quantifying effects of biodiversity on ecosystem functioning across ...
    Feb 28, 2018 · Theory predicts that additional insurance effects of biodiversity on ecosystem functioning could emerge across time and space if species respond ...
  129. [129]
    Biodiversity promotes ecosystem functioning despite environmental ...
    Dec 7, 2021 · Independent of the level of environmental stress, biodiversity helps to maintain ecosystem functioning at relatively stable levels, whereas ...
  130. [130]
    Scaling up biodiversity–ecosystem function relationships across ...
    Aug 27, 2020 · Here we develop quantitative scaling relationships linking 374 experiments that tested plant diversity effects on biomass production across a range of scales.
  131. [131]
    Effects of biodiversity on ecosystem functioning: a consensus of ...
    1) Certain combinations of species are complementary in their patterns of resource use and can increase average rates of productivity and nutrient retention. At ...
  132. [132]
  133. [133]
    The productivity-biodiversity relationship varies across diversity ...
    Dec 12, 2019 · Understanding the processes that drive the dramatic changes in biodiversity along the productivity gradient remains a major challenge.
  134. [134]
    Consistently positive effect of species diversity on ecosystem, but not ...
    May 17, 2021 · We show that ecosystem temporal stability tended to increase with species diversity, regardless of study systems.
  135. [135]
    Complexity and stability of ecological networks: a review of the theory
    Jul 6, 2018 · The use of ecological-network models to study the relationship between complexity and stability of natural ecosystems is the focus of this review.
  136. [136]
    Biodiversity and Ecosystem Function: The Debate Deepens - Science
    It could be argued that the tide is turning against the notion of high biodiversity as a controller of ecosystem function and insurance against ecological ...
  137. [137]
    Biodiversity–stability relationships strengthen over time in a long ...
    Dec 14, 2022 · Numerous studies have demonstrated that biodiversity drives ecosystem functioning, yet how biodiversity loss alters ecosystems functioning ...<|separator|>
  138. [138]
    Why are biodiversity-ecosystem functioning relationships so elusive ...
    Sep 11, 2022 · Why are biodiversity-ecosystem functioning relationships so elusive? Trophic interactions may amplify ecosystem function variability · Authors.
  139. [139]
    Climate mediates the biodiversity–ecosystem stability relationship ...
    Jul 30, 2018 · Our results indicate that spatial, climatic, soil, and plant diversity variables can explain up to 73% of the variation in ecosystem stability ...
  140. [140]
    Biodiversity and ecosystem productivity in a fluctuating environment
    This work shows two major effects of species richness on ecosystem productivity in a fluctuating environment: (i) a reduction in the temporal variance of ...Missing: debates | Show results with:debates
  141. [141]
    Science and values in the biodiversity-ecosystem function debate
    Feb 20, 2022 · This paper explores interactions between ecological science and conservation values in the biodiversity-ecosystem function (BEF) debate of the 1990–2000s.
  142. [142]
    Biodiversity–ecosystem functioning research: Brief history, major ...
    Ecologists have intensively debated the Biodiversity–Ecosystem Functioning (BEF) relationships for more than 50 years (1973 – the present).
  143. [143]
    NSF 24-520: Long-Term Ecological Research (LTER)
    Dec 21, 2023 · Core data collection at LTER sites will continue to center on five areas: 1) primary production, 2) population dynamics and trophic structure, ...
  144. [144]
    Hubbard Brook LTER
    Oct 13, 2025 · The Hubbard Brook LTER is a 3,160 hectare reserve for long-term study of forest and aquatic ecosystems, designated as an LTER site in 1988.
  145. [145]
    About the Hubbard Brook Ecosystem Study
    The Hubbard Brook Ecosystem Study is a long-term research program studying northern forest ecosystems, including biological composition, productivity, and ...
  146. [146]
    The Hubbard Brook Ecosystem Study: Celebrating 50 Years
    Oct 1, 2013 · Six major scientific findings are especially noteworthy: (1) discovery of acid rain in North America and its impact on aquatic and terrestrial ...
  147. [147]
    Hubbard Brook Streamflow Response to Deforestation (Overview)
    The Hubbard Brook experiment tested if deforestation would increase streamflow, with the hypothesis that cutting down trees would lead to increased streamflow.<|separator|>
  148. [148]
    The Park Grass Experiment 1856–2006: its contribution to ecology
    May 5, 2006 · The Park Grass Experiment, begun in 1856, is the oldest ecological experiment in existence. Its value to science has changed and grown since it ...Introduction · The Park Grass Experiment · Plant dynamics · Fauna
  149. [149]
    Replication in field ecology: Identifying challenges and proposing ...
    Jun 8, 2021 · Field ecologists are therefore facing a significant challenge in assessing the replicability of their research with implications for overall ...
  150. [150]
    Field experiments underestimate aboveground biomass response to ...
    Mar 10, 2022 · To assess how climatic changes will affect ecosystems, field researchers commonly use one of two approaches: in situ observations or ...
  151. [151]
  152. [152]
    46.1D: Modeling Ecosystem Dynamics - Biology LibreTexts
    Nov 22, 2024 · Conceptual models describe ecosystem structure, while analytical and simulation models use algorithms to predict ecosystem dynamics.
  153. [153]
    An introduction to mathematical models in ecology and evolution
    This book introduces mathematical models in ecology and evolution, covering exponential growth, density dependence, and spatial models, and is suitable for ...
  154. [154]
    Mathematical models in ecology and evolutionary biology
    Learning basic model mechanics through classic models in EEB · Exponential growth · Logistic growth · When species are in competition – the Lotka-Volterra model.
  155. [155]
    Stability of ecological systems: A theoretical review - ScienceDirect
    Oct 17, 2024 · In this article, we provide a systematic and comprehensive review on the theoretical frameworks for analyzing the stability of ecological systems.
  156. [156]
    A Conceptual Framework to Integrate Biodiversity, Ecosystem ...
    Sep 1, 2022 · We review models used in recent global model intercomparison projects and develop a novel model integration framework to more fully account for ...
  157. [157]
    A theoretical framework for the ecological role of three‐dimensional ...
    Feb 1, 2023 · We present a framework for conceptualizing structural diversity and suggest how to facilitate its broader incorporation into ecological theory ...
  158. [158]
    [PDF] A review of ecosystem modelling frameworks to support the ... - UKRI
    Aim. The aim of this review was to assess how high-level modelling design decisions map to the effectiveness of models that are designed for ecosystem ...
  159. [159]
    Beyond ecosystem modeling: A roadmap to community ...
    Model averaging in ecology: A review of Bayesian, information‐theoretic, and tactical approaches for predictive inference. Ecological Monographs, 88, 485–504.<|separator|>
  160. [160]
    How Earth observation satellites aid climate change research
    May 16, 2024 · 6 ways satellites are helping to monitor our changing planet from space · 1. PACE in search of phytoplankton · 2. Monitoring Earth from the ...
  161. [161]
    Four ways that space is changing ecosystem monitoring - ESA
    Jun 4, 2021 · We look at four projects using satellite data to improve the health of ecosystems around the globe.
  162. [162]
    Drones in ecology: ten years back and forth | BioScience
    Jun 19, 2025 · In the present article, we explore the key developments in ecological drone science since 2013, considering plant and animal ecology, imaging ...
  163. [163]
    Systematic review of remote sensing technology for grassland ...
    On the one hand, UAV remote sensing technology helps grassland biodiversity monitoring to develop from species diversity to functional diversity. For example, ...
  164. [164]
    New national strategy promotes advancements in eDNA technologies
    Jun 12, 2024 · Advancements in biomolecular technology now allow scientists to detect DNA shed by marine life into the water, a technique called environmental ...
  165. [165]
    Advances in environmental DNA monitoring: standardization ...
    Environmental DNA (eDNA) monitoring, a rapidly advancing technique for assessing biodiversity and ecosystem health, offers a noninvasive approach for ...
  166. [166]
    Advancing the environmental DNA and RNA toolkit for aquatic ...
    Mar 18, 2025 · This special issue of PeerJ Life and Environment brings together 20 innovative studies that collectively advance the eDNA toolkit.
  167. [167]
    Artificial intelligence helps drive new frontiers in ecology | BioScience
    May 8, 2024 · AI is helping ecologists monitor air quality, measure the changing footprints of ecosystems, and track changes to species distribution.
  168. [168]
    Integrating AI models into ecological research workflows: The case ...
    Aug 8, 2025 · Artificial intelligence (AI) methods are critical for advancing sensor-based ecology, as they can extract information from autonomously sensed ...
  169. [169]
    A synergistic future for AI and ecology - PMC - NIH
    Sep 11, 2023 · Research in both ecology and AI strives for predictive understanding of complex systems, where nonlinearities arise from multidimensional interactions and ...
  170. [170]
    [PDF] Twenty years of ecosystem services - Robert Costanza
    The chapters covered definitions, history, economic valuation, overarching services like climate and biodiversity, ser- vices from specific biomes including ...
  171. [171]
    [PDF] Ecosystem services: Key concepts and applications
    The Millennium Ecosystem Assessment framework (see Box 2) identified ecosystem services within four categories: □ provisioning services, such as food and water.
  172. [172]
    Ecosystem Services | National Wildlife Federation
    Four Types of Ecosystem Services · 1. Provisioning Services · 2. Regulating Services · 3. Cultural Services · 4. Supporting Services
  173. [173]
    Economic values for ecosystem services: A global synthesis and ...
    This paper presents a global synthesis of economic values for ecosystem services provided by 15 terrestrial and marine biomes.
  174. [174]
    [PDF] Methods for monetary valuation of ecosystem services: A scoping ...
    This review identified over 20 monetary valuation techniques for ecosystem services, including methods like willingness to sell and Delphi method.
  175. [175]
    National Ecosystem Services Classification System - EPA
    Nov 8, 2024 · The main objective of NESCS is to provide a framework that will aid in analyzing the human welfare impacts of policy-induced changes to ecosystems.Missing: categories | Show results with:categories<|separator|>
  176. [176]
    Ecosystem Services: Origins, Contributions, Pitfalls, and Alternatives
    We then outline three sets of weaknesses in the ES framework: confusion over ecosystem functions and biodiversity, omission of dis-services, trade-offs and ...
  177. [177]
    A SWOT analysis of the ecosystem services framework
    Dec 22, 2015 · Weaknesses include an incomplete scientific basis, frameworks being inconsistently applied, and accounting for nature's intrinsic value.Missing: criticisms | Show results with:criticisms
  178. [178]
    Progress and Prospects of Ecosystem Disservices - MDPI
    The interaction and connection between ecosystem services and disservices are mainly reflected in trade-off and synergy. Ecosystem services and disservices are ...
  179. [179]
    Conceptual ambiguity hinders measurement and management of ...
    May 14, 2020 · Disservices are not isolated outcomes of ecosystem function; they are supplied concurrently with services via interactions within the system.
  180. [180]
    A theoretical rethinking of ecosystem services from the perspective ...
    Ecosystem service (ES) research has grown rapidly, but emergent concepts such as disservices, supply-demand, relationships, and flows remain fragmented.
  181. [181]
    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 ...
  182. [182]
    'All the birds returned': How a Chinese project led the way in water ...
    Mar 15, 2025 · By 2016, China had converted more than 11,500 sq miles of rain-fed cropland to forest or grassland – a 25% increase in vegetative cover in a ...Missing: biodiversity | Show results with:biodiversity
  183. [183]
    China restored the world's most eroded land—but not without ...
    By 2016, China had converted over 11,500 square miles of cropland into forest or grassland, improving soil stability and biodiversity but also raising concerns ...
  184. [184]
    Ecological restoration success on the Loess Plateau of China
    Ecological restoration programs such as the "Grain for Green Project" (GFGP) have significantly reduced soil erosion and increased vegetation cover on the ...<|control11|><|separator|>
  185. [185]
    Wolves of Yellowstone - National Geographic Education
    May 10, 2024 · Gray wolves were reintroduced into Yellowstone National Park in 1995, resulting in a trophic cascade through the entire ecosystem.
  186. [186]
    History of Wolf Management - Yellowstone National Park (U.S. ...
    Sep 9, 2025 · Wolves have contributed to decreased survival rates of elk calves and older female adults while also influencing elk habitat use. These effects, ...Early Wolf Management · Restoration Efforts · Legal Challenges
  187. [187]
    Reintroducing wolves to Yellowstone helped entire ecosystem thrive ...
    Feb 25, 2025 · The reintroduction of wolves to Yellowstone National Park in the 1990s had a cascading effect that benefited the entire ecosystem, a new study finds.
  188. [188]
    Do wolves fix ecosystems? CSU study debunks ... - The Coloradoan
    Feb 9, 2024 · In Yellowstone's case, it was claimed in a 2004 study that reintroduced wolves preying on elk created a "landscape of fear" and, either through ...
  189. [189]
    Six global success stories on how rewilding key species ... - One Earth
    Oct 6, 2025 · From beavers in the UK to bison in the Great Plains, here are six stories that demonstrate the power of rewilding in restoring the health of ...Missing: human | Show results with:human
  190. [190]
    Back from the brink: Six species saved by ecosystem restoration
    Oct 4, 2022 · They include the regeneration of forest on abandoned land and the creation of wildlife corridors between protected areas, strategies that are ...
  191. [191]
    The direct drivers of recent global anthropogenic biodiversity loss
    Nov 9, 2022 · We show that land/sea use change has been the dominant direct driver of recent biodiversity loss worldwide.
  192. [192]
    Five drivers of the nature crisis - UNEP
    Sep 5, 2023 · According to scientists, halting and reversing the degradation of lands and oceans can prevent the loss of one million endangered species. In ...
  193. [193]
    Human health impacts of ecosystem alteration - PubMed Central - NIH
    We discuss what is known about the human health implications of changes in the structure and function of natural systems.
  194. [194]
    Climate change could become the main driver of biodiversity decline ...
    Apr 25, 2024 · Land-use change is considered the largest driver of biodiversity change, according to the Intergovernmental Platform on Biodiversity and ...
  195. [195]
    BGC - Yale Center for Biodiversity and Global Change
    Apr 25, 2024 · Climate change could become the main driver of biodiversity decline by mid-century, with combined climate and land-use change leading to loss ...
  196. [196]
    The Latest in Climate Change Attribution and the Law
    Feb 7, 2020 · ... ecosystem degradation. There is a growing body of extreme event and impact attribution studies finding a causal connection between impacts ...Missing: debates | Show results with:debates<|separator|>
  197. [197]
    [PDF] BIODIVERSITY AND ECOSYSTEM SERVICES - IPBES
    However, biodiversity is still being lost, ecosystems are still being degraded and many of nature's contributions to people are being compromised. The ...
  198. [198]
    [PDF] cbd-en.pdf - Convention on Biological Diversity
    The objectives of this Convention, to be pursued in accordance with its relevant provisions, are the conservation of biological diversity, the sustainable use ...
  199. [199]
    2030 Targets (with Guidance Notes)
    The 2030 targets include reducing biodiversity loss, restoring 30% of degraded ecosystems, conserving 30% of land, waters and seas, and halting species ...Implementation and support... · Target 1 · Target 3 · Target 2
  200. [200]
    Public policies and global forest conservation: Empirical evidence ...
    We estimate that public policies reduce the risk of tree cover loss by almost 4 percentage points globally, but there is large variation around this.
  201. [201]
    How much progress have we made towards global ambitions on ...
    As of August 2024, 17.5 per cent of Earth's terrestrial and inland waters are protected or conserved, alongside 8.5 per cent of the ocean.
  202. [202]
    Mixed effectiveness of global protected areas in resisting habitat loss
    Sep 27, 2024 · Protected areas were 33% more effective in reducing habitat loss compared to unprotected areas, though their ability to mitigate nearby human pressures was ...
  203. [203]
    Assessing the integrated conservation effectiveness of protected ...
    The integrated assessment revealed mixed results: 33.03% of the area within PAs achieved high conservation effectiveness, while 60.73% showed medium ...<|separator|>
  204. [204]
    Money for Nothing? A Call for Empirical Evaluation of Biodiversity ...
    Our understanding of the ecological aspects of ecosystem conservation rests, in part, on well-designed empirical studies. In contrast, our understanding of the ...
  205. [205]
    Redesigning Payments for Ecosystem Services to Increase Cost ...
    Jul 11, 2024 · We find that the full-enrollment treatment reduces deforestation by 41% compared to the traditional contract. This extra conservation occurs ...
  206. [206]
    Testing the Effectiveness of Payments for Ecosystem Services to ...
    This is the first randomized evaluation of a deforestation PES program, designed to measure its effectiveness and cost-effectiveness.
  207. [207]
    Redesigning payments for ecosystem services to increase cost ...
    Oct 26, 2024 · Many but not all studies find that PSA has been effective at reducing deforestation. However, PSA's funding has declined. From 2015-2019, ...
  208. [208]
    Payments for ecosystem services programs: A global review of ...
    Jan 15, 2024 · Our global review explores twelve key characteristics of PESPs at three different phases (inputs – implementation – outputs and outcomes)
  209. [209]
    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 ...
  210. [210]
    5 reasons why many conservation efforts fail - Mongabay
    Mar 30, 2016 · They fail to understand the past and current ecology of the place, its wildlife, politics and people. And this can lead to failure. Studies have ...Share This Article · Lack Of Local Buy-In · Lack Of Funding<|control11|><|separator|>
  211. [211]
    Embracing Change in Conservation to Protect Biodiversity and ...
    Aug 22, 2025 · This lack of empirical evidence on the efficacy of novel conservation strategies hampers our understanding of the risks and benefits ...
  212. [212]
    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 ...<|separator|>
  213. [213]
    Meta-analysis shows the impacts of ecological restoration on ...
    Mar 26, 2024 · Our findings reveal that forest and grassland restoration increase CH 4 uptake by 90.0% and 30.8%, respectively, mainly due to changes in soil properties.
  214. [214]
    A meta-analysis contrasting active versus passive restoration ...
    Nov 23, 2020 · We used meta-analysis to contrast different classes of dryland restoration practices. All interventions were categorized as active and passive.
  215. [215]
    a synthesis of outcomes from ecosystem restoration in tropical and ...
    Nov 14, 2022 · Based on community-level data from 11 landscapes, active restoration resulted in faster accumulation of tree basal area and structural ...<|control11|><|separator|>
  216. [216]
    A meta-analysis of the ecological and economic outcomes ... - Nature
    Aug 19, 2021 · The meta-analysis reveals that mangrove restoration provides higher ecosystem benefits over unvegetated tidal flats, while generally lower than ...
  217. [217]
    Assessing the success of marine ecosystem restoration using meta ...
    Mar 29, 2025 · Using this approach, we obtained an average success of ca. 64% for all restoration interventions (Fig. 2), with higher values for coral reefs, ...
  218. [218]
    Conceptual and methodological issues in estimating the success of ...
    We demonstrate here that, due to the heterogeneity of collected data, the success of restoration actions can be overestimated in meta-analyses.
  219. [219]
    The cost and feasibility of marine coastal restoration - ESA Journals
    Nov 4, 2015 · Success in restoration of marine coastal ecosystems was assessed as short-term survival rates, while success in terrestrial ecosystems was ...
  220. [220]
    [PDF] The Future of Changes in Global Ecosystem Services
    Oct 27, 2021 · Some of the $125 trillion ES value is included in GDP, but most ES are non-marketed regulating services such as storm and flood protection, ...<|separator|>
  221. [221]
    It's time to recognise the economic value of an ecosystem
    Feb 27, 2023 · More than half of global GDP, around $44 trillion of economic value, is dependent on nature, we must acknowledge the economic value of an ...
  222. [222]
    Protecting Nature Could Avert Global Economy Losses of $2.7 ...
    Jul 1, 2021 · A new World Bank report estimates that the collapse of select ecosystem services provided by nature – such as wild pollination, provision of food from marine ...
  223. [223]
    $$5 trillion of nature-related economic risks will amplify climate ...
    Dec 14, 2023 · The ECI study concludes, the erosion of natural capital linked with biodiversity loss and environmental degradation generates significant and ...
  224. [224]
    Effects of Payment for Ecosystem Services and Tourism on ...
    The results revealed that the NFCP effectively reduced forest loss and promoted forest gain, while the GTGP significantly fostered forest gain but concurrently ...2. Materials And Methods · B. Forest Data And... · 4. Discussion
  225. [225]
    Evaluating the outcomes of payments for ecosystem services ...
    We found that PES schemes can provide positive conservation and development outcomes with respect to livelihoods, land-use change, household and community ...
  226. [226]
    Critical Analysis of Payments for Ecosystem Services: Case Studies ...
    Jun 11, 2023 · Short-term-implemented PES schemes had success rates of 46% and represented 52% of the analyzed PES schemes. The success was moderate because ...
  227. [227]
    Environmental Ethics - Stanford Encyclopedia of Philosophy
    Jun 3, 2002 · Environmental ethics is the philosophy discipline studying the moral relationship of humans to the environment and its non-human contents.
  228. [228]
    [PDF] Anthropocentric, Biocentric and Ecocentric Ethics - PhilArchive
    Schweitzer's ethic of reverence for life expresses an extreme form of biocentric view that is quite different from the anthropocentric perspectives discussed ...
  229. [229]
    The Rights of Nature — Can an Ecosystem Bear Legal Rights?
    Apr 22, 2021 · The Rights of Nature law recognizes that an ecosystem has the right to exist, flourish, regenerate its vital cycles, and naturally evolve ...
  230. [230]
    Is granting legal rights to nature a promising approach to ...
    Aug 29, 2023 · Legal rights can be complex and difficult to implement, and they have not always been successful in preventing harm to nature.
  231. [231]
    Where Nature's Rights Go Wrong - Virginia Law Review
    Nov 29, 2021 · The shortcomings of nature's rights, however, do not mean that constitutional reform cannot be used to promote environmental goals. Recent work ...Missing: precedents | Show results with:precedents