Fact-checked by Grok 2 weeks ago

Ecosystem ecology

Ecosystem ecology is a subdiscipline of ecology that examines ecosystems as dynamic complexes of biotic and abiotic interactions, emphasizing the exchange of energy and matter through processes like primary production, decomposition, and nutrient cycling. It integrates living organisms with their physical environment to analyze system-level functions, distinguishing itself from organismal, population, or community ecology by prioritizing holistic material and energy budgets over individual species dynamics. The field gained prominence in the mid-20th century, largely through the work of Eugene P. Odum, who popularized the ecosystem concept in his 1953 textbook Fundamentals of Ecology, framing ecosystems as self-regulating systems capable of maintaining internal via feedback loops. Odum and his brother advanced quantitative approaches, including energy flow diagrams and systems modeling, exemplified by the Silver Springs study, which quantified trophic transfers in a spring ecosystem to reveal inefficiencies in propagation across levels. These methods enabled empirical assessments of and limiting factors, grounded in measurable fluxes rather than qualitative descriptions. Key achievements include elucidating biogeochemical cycles and trajectories, providing causal insights into how disturbances disrupt or restore states, as validated through long-term and experimental manipulations. However, controversies persist regarding the field's assumptions of steady-state conditions, which empirical data often challenge amid stochastic events and nonlinear responses, highlighting limitations in predictive modeling for perturbed systems. Despite such debates, ecosystem ecology's emphasis on verifiable processes has informed by revealing bottom-up controls on , such as limitations in terrestrial and systems.

Definition and Scope

Core Principles

Ecosystem ecology examines the and of ecosystems as integrated units where interact with their abiotic through the unidirectional flow of and the cyclic movement of matter. enters ecosystems primarily via by autotrophs, which convert solar radiation into chemical , supporting rates that vary globally from about 1-3 g/m²/year in deserts to over 2,000 g/m²/year in tropical rainforests. This supports heterotrophic across trophic levels, with transfer averaging around 10% between levels due to metabolic losses and incomplete consumption, resulting in most energy being lost as heat. Nutrient cycling constitutes another foundational principle, wherein essential elements like carbon, , and are recycled through biological uptake, transfer via consumption, and return to the abiotic pool via by microbes and detritivores. For instance, in terrestrial ecosystems, by bacteria converts atmospheric N₂ into bioavailable forms at rates of 10-200 kg/ha/year, sustaining ; disruptions, such as excess nitrogen deposition from human activities exceeding 50 kg/ha/year in some regions, can alter cycling dynamics and favor . These cycles maintain ecosystem , the balance of elements in organisms and their , influencing processes from growth to rates. Ecosystems demonstrate emergent properties such as and , arising from interactions among components rather than individual parts. Feedback loops, including density-dependent regulation and nutrient retention, help stabilize systems against perturbations; for example, or nutrient limitation can prevent overexploitation. However, empirical studies reveal that enhances functioning, with diverse assemblages showing 1.5-2 times higher and in experimental grasslands compared to monocultures, though involves complex trade-offs rather than simple diversity-function correlations. These principles emphasize causal links between biophysical processes, rejecting oversimplified models in favor of dynamic, context-dependent behaviors observed in long-term data. Ecosystem ecology distinguishes itself from other ecological subdisciplines by prioritizing the integrated analysis of energy flows, nutrient cycling, and material budgets across biotic and abiotic components within defined spatial units, treating ecosystems as functional systems rather than aggregates of lower-level entities. This contrasts with organismal or physiological ecology, which focuses on individual-level adaptations, metabolic processes, and responses to environmental variables, such as how temperature affects in a single , without scaling to system-wide fluxes. Population ecology, in turn, examines dynamics like birth rates, mortality, and within one , often using models such as the Lotka-Volterra equations for predator-prey interactions limited to two populations, whereas ecosystem ecology incorporates these as subsystems within broader trophic networks and abiotic pools. Community ecology overlaps in studying multi-species interactions like competition, mutualism, and predation but typically emphasizes biotic assembly rules and diversity patterns, abstracting from abiotic processes such as soil nutrient retention or detrital decomposition that dominate ecosystem-level analyses. For instance, while community ecologists might quantify niche overlap via resource partitioning metrics, ecosystem ecologists quantify how such interactions influence whole-system primary production or carbon sequestration rates, revealing emergent properties not predictable from species pairwise data alone. Landscape ecology extends beyond by addressing spatial heterogeneity, fragmentation, and connectivity across multiple ecosystems, using metrics like edge effects or dispersal corridors, in contrast to ecosystem ecology's focus on internal process homogeneity within boundaries defined by hydrological or biogeochemical flows, such as a lake catchment. Biogeochemistry shares ecosystem ecology's interest in elemental cycles but often operates at global or regional scales with less emphasis on biotic mediation, prioritizing abiotic drivers like atmospheric deposition over trophic controls on, for example, mineralization rates. These boundaries are not absolute, as ecosystem ecology draws on data from allied fields—evident in hybrid approaches like energetics that integrate population models—but maintains methodological primacy on empirical measurement of system throughput, such as via tracing of carbon pathways, to validate causal links between structure and function. Overlaps have intensified since the , with interdisciplinary syntheses addressing scale transitions, yet ecosystem ecology retains its core identity through insistence on closed-system approximations for process quantification.

Historical Development

Precursors and Early Concepts (19th–Early 20th Century)

The term oecology (later anglicized to ecology) was coined in 1866 by German zoologist Ernst Haeckel in his work Generelle Morphologie der Organismen, defining it as "the whole science of the relations of the organism to the environment including, in the broad sense, all the 'conditions of existence.'" Haeckel's formulation encompassed both inorganic factors, such as physical and chemical habitat conditions, and organic interactions among organisms, drawing from Charles Darwin's 1859 concepts of natural selection and the struggle for existence while emphasizing a balanced "economy of nature" where organism populations remained relatively constant through shifting species dynamics. This holistic view positioned ecology as a branch of physiology focused on organism-environment interrelations, laying foundational ideas for later ecosystem thinking by integrating biotic and abiotic elements without yet formalizing systemic processes like energy flow. A pivotal precursor emerged in 1887 with American entomologist Stephen A. Forbes's essay "The Lake as a Microcosm," which portrayed freshwater lakes—drawing from observations in Illinois and Wisconsin—as self-contained, primitive systems where organisms formed an interdependent community akin to a single organism. Forbes highlighted how physical conditions, such as water chemistry and seasonal fluctuations, intertwined with biotic chains of predation and competition, arguing that alterations in one species rippled across the entire assemblage, necessitating study of the whole to understand any part. This microcosm analogy anticipated ecosystem holism by stressing equilibrium through mutual dependencies and environmental regulation, influencing subsequent aquatic and community ecology while underscoring the lake's isolation as a "chapter out of the history of a primeval time." In the early 20th century, these ideas evolved through botanical and zoological studies emphasizing larger-scale integrations. Danish botanist Eugenius Warming's 1895 textbook Planzesamfund classified plant formations based on ecological adaptations, bridging individual to community-level patterns in response to gradients. American ecologist Frederic E. Clements advanced this in 1916 with his "biome" concept in Plant Succession, defining as comprehensive regional complexes of and associated animals interacting with climatic and edaphic factors, thus extending organism-centric views toward supra-community units that prefigured boundaries. These developments, building on Haeckel and , shifted focus from isolated relations to dynamic, spatially defined systems, though they retained an organismal emphasis over explicit process quantification.

Formalization and Key Milestones (1930s–1960s)

The concept of the ecosystem was formalized in 1935 by British ecologist Arthur G. Tansley in his paper "The Use and Abuse of Vegetational Concepts and Terms," published in Ecology (16:284–307), where he defined it as "the whole complex of physical factors forming what we call the environment" integrated with the "living organism" complex, emphasizing biotic-abiotic interactions as the basic ecological unit rather than viewing communities as superorganisms. Tansley's formulation, building on earlier suggestions by colleague Roy Clapham, rejected overly holistic interpretations of vegetation and promoted a systems perspective amenable to empirical analysis. This marked a shift toward mechanistic, process-oriented , influencing subsequent quantitative approaches. A pivotal theoretical advancement occurred in 1942 with Raymond L. Lindeman's posthumously published paper "The Trophic-Dynamic Aspect of " in Ecology (23:399–418), which conceptualized ecosystems as dynamic -processing systems organized by trophic levels, where represents progressive stabilization of flow through transformation and efficiency losses (typically 10–20% transfer between levels). Lindeman applied this to lake ecosystems, deriving biomass pyramids from data and highlighting increases as a driving force, thus providing a foundational model for budgeting over descriptive community studies. His work, completed amid personal illness and rejected initially by some reviewers for its mathematical abstraction, bridged with , enabling predictive modeling of material and balances. The 1950s saw empirical consolidation through the Odum brothers' research, with Eugene P. Odum and Howard T. Odum's Fundamentals of Ecology (1953) synthesizing Tansley and Lindeman into a comprehensive framework, defining ecosystems by self-regulating energy circuits and nutrient feedbacks. Howard T. Odum's 1956–1957 study of Silver Springs, Florida—a clear-water river ecosystem—quantified whole-system energy flows (e.g., gross primary production of 6.5 g m⁻² day⁻¹ oxygen equivalent, with 18% exported as detritus), validating trophic-dynamic principles via diel oxygen curves and biomass measurements across 20+ trophic levels. This fieldwork, supported by Atomic Energy Commission grants, demonstrated ecosystems' high efficiency in subtropical settings and spurred large-scale budgeting techniques, though later critiques noted underestimation of subterranean flows. By the late 1960s, these milestones had elevated ecosystem ecology from conceptual abstraction to a quantitative discipline, informing radiation ecology studies at sites like Eniwetok Atoll.

Expansion and Institutionalization (1970s–Present)

The International Biological Program (IBP), spanning 1964 to 1974, catalyzed the expansion of ecosystem ecology by prioritizing large-scale studies of biological productivity, resource management, and human-induced environmental changes, which drew unprecedented funding—exceeding $1 billion globally—and fostered multi-site, interdisciplinary research on ecosystem processes. This effort built on earlier systems approaches, incorporating mathematical modeling to quantify material and energy fluxes, as exemplified by Howard T. Odum's (1971), which formalized energy circuit language for depicting trophic dynamics and in ecosystems. Concurrently, heightened public and policy attention to and resource depletion, spurred by events like the 1969 fire, integrated ecosystem-level analyses into environmental assessments, such as Gene Likens' Hubbard Brook studies documenting acid rain's biogeochemical impacts starting in the early 1970s. Institutionalization accelerated in the with the U.S. National Science Foundation's establishment of the Long-Term Ecological Research (LTER) Network in 1980, initially funding six sites representing major biomes to collect decadal-scale data on core variables like , disturbance regimes, and population patterns through integrated observation, experimentation, and modeling. By 2023, the network expanded to 28 U.S. sites and inspired the International LTER (ILTER) framework, enabling cross-site syntheses that revealed patterns such as nitrogen saturation thresholds and climate-driven shifts in . This infrastructure professionalized the field, training thousands of researchers and generating datasets underpinning meta-analyses, while academic programs proliferated, with ecosystem ecology departments or centers emerging at institutions like the and . Subsequent decades saw methodological advances, including process-based models like the CENTURY soil organic matter model developed in the 1980s, which simulated nutrient cycling and plant-soil interactions using site-specific and management inputs to predict long-term responses to perturbations. The marked a pivot toward applied valuation, with Robert Costanza et al.'s 1997 analysis estimating global services at $16–54 trillion annually (in 1997 USD), aggregating 17 services across 16 biomes based on 131 prior studies to highlight economic dependencies on biophysical processes like and . This quantification influenced policy frameworks, though critiques noted uncertainties in discounting future values and non-market services. From the 2000s onward, ecosystem ecology institutionalized further through integration with global challenges, including biodiversity-ecosystem functioning (BEF) experiments like the Jena Experiment (initiated 2002), which demonstrated linear increases in biomass production with plant via niche complementarity and selection effects. Advances in and enabled scaling from plots to biomes, as in NASA's MODIS data for monitoring net primary productivity since 2000, revealing anthropogenic drivers of terrestrial carbon sinks absorbing ~30% of annual emissions. Contemporary emphases include metrics in social-ecological systems, informed by loop analysis and qualitative modeling frameworks refined since the , to assess tipping points in disturbed ecosystems like coral reefs under warming. These developments underscore ecosystem ecology's maturation into a predictive , with over 2,000 LTER-affiliated researchers contributing to evidence-based amid climate variability.

Fundamental Processes

Energy Flow and Trophic Interactions

Energy flow in ecosystems begins with the capture of by autotrophic producers through , forming the basis of all subsequent trophic transfers. This process yields gross primary production, which represents the total energy fixed before respiratory losses, with available for consumption or storage. Raymond Lindeman's 1942 analysis of a freshwater ecosystem introduced the trophic-dynamic framework, quantifying energy budgets across levels and emphasizing the sequential transfer from producers to herbivores, carnivores, and decomposers. In this model, energy diminishes progressively due to the second law of thermodynamics, as organisms expend most ingested energy on and export, leaving limited amounts for growth and reproduction. Trophic transfer efficiency, the fraction of energy from one level assimilated and passed to the next, averages approximately 10%, though empirical measurements range from 5% to 20% depending on ecosystem type and taxa. This "10% rule" arises from Lindeman's observations, where predation and assimilation inefficiencies, combined with respiratory heat loss, constrain upward flow; for instance, in aquatic systems, herbivore assimilation of plant material often falls below 20% due to indigestible structural compounds like cellulose. Howard T. Odum's 1956 study of the Silver Springs, Florida, ecosystem provided empirical quantification, revealing annual net primary production of 8,430 kcal/m², with only about 11% transferred to herbivores and further declines to top carnivores, underscoring the pyramid-like structure of energy distribution. Decomposers play a parallel role, processing detritus and facilitating nutrient return, but their energy pathways similarly exhibit low transfer efficiencies. Trophic interactions extend beyond linear chains to complex food webs, encompassing predation, herbivory, omnivory, and that dictate energy partitioning and ecosystem stability. Food webs represent interconnected feeding links, where or strong interactions can amplify or dampen energy flows; for example, top predators regulate populations, preventing and sustaining biomass. In Silver Springs, Odum's model highlighted the dominance of and detrital paths, with and mediating 384 and 152 kcal/m²/yr to higher levels, respectively, illustrating how interaction influences overall throughput. These dynamics reveal causal mechanisms where inefficient transfers enforce biomass pyramids, limiting higher-level abundance and shaping against perturbations. Empirical validations, such as stable analyses, confirm that not all links contribute equally, with basal resources driving most variance in trophic structure.

Nutrient Cycling and Decomposition

Nutrient cycling encompasses the biological, geological, and chemical processes that transfer essential elements such as carbon, , and between living organisms, , and the abiotic environment within ecosystems, thereby sustaining and overall functioning. These cycles prevent nutrient depletion by recycling materials rather than relying solely on external inputs, with rates determined by factors including organismal demand, environmental conditions, and stoichiometric balances. Empirical studies in ecosystems, for instance, demonstrate that nitrogen limitation dominates in early , shifting to phosphorus limitation later, highlighting how cycling efficiency influences long-term community structure. Decomposition, a core component of nutrient cycling, involves the microbial and faunal breakdown of dead , mineralizing nutrients into inorganic forms available for plant uptake and preventing accumulation of undecomposed . Rates of decomposition vary globally, with faster turnover in warmer, moist environments; for example, decomposition in tropical forests can exceed that in boreal systems by factors of 2–5 due to higher microbial activity and enzymatic efficiency. Key drivers include quality, such as carbon-to-nitrogen ratios below 25:1 promoting rapid breakdown, and extrinsic factors like and , which together explain up to 70% of variance in mass loss across biomes. Microorganisms dominate decomposition, with and fungi hydrolyzing complex polymers into monomers via extracellular enzymes, while detritivores like and fragment material to enhance surface area for microbial access. redox cycling, for instance, regulates long-term litter decay by oxidizing recalcitrant compounds, as shown in controlled experiments where Mn amendments increased by 20–50% in organic-rich s. Nutrient release during follows models, with initial rapid mineralization of labile fractions (e.g., sugars, ) followed by slower processing of lignins and , influencing soil nutrient pools and feedback to primary producers. Disruptions to , such as enrichment, can invert traditional positive relationships with decay rates, leading to and reduced carbon turnover, as observed in long-term fertilization experiments where addition slowed by 10–30% over decades. In ecosystems, human alterations like and flow regulation shape patterns, with global riverine rates correlating strongly with dissolved oxygen and organic inputs, underscoring the connectivity between terrestrial and . These processes underpin , as efficient cycling buffers against losses via or , though climate warming projections indicate accelerated in cold biomes, potentially releasing 10–20% more by 2100.

Disturbance, Succession, and Resilience

Disturbances in ecosystems are defined as relatively discrete events that disrupt community, population, or structure and alter resource availability, often leading to mortality or relocation of organisms. Common examples include wildfires, floods, hurricanes, droughts, insect outbreaks, and volcanic eruptions, which vary in spatial extent, frequency, severity, and duration. These events create heterogeneous patches within landscapes, influencing by removing dominant and allowing opportunistic , as evidenced in studies of large, infrequent disturbances like the 1980 eruption, which affected over 200,000 hectares and initiated long-term monitoring of recovery dynamics. Unlike chronic stressors, disturbances are typically pulsed and can enhance function by recycling nutrients and preventing monopolization by late- , though excessive frequency or intensity may exceed recovery thresholds. Ecological succession follows disturbance as a predictable sequence of community changes driven by species interactions, resource gradients, and abiotic facilitation or inhibition. Primary succession initiates on barren substrates devoid of and biotic legacies, such as glacial retreats or lava flows, where pioneer autotrophs like and mosses initiate pedogenesis through and organic accumulation, progressing over centuries to complex forests; for instance, the 1963 volcanic island has documented colonization within years and vascular plants by decades. occurs on disturbed sites retaining and banks, accelerating recovery—often 10-50 times faster than primary—via rapid herb and shrub dominance before tree re-establishment, as observed in post-fire systems where nutrient pulses from ash decomposition favor early seral . Both types culminate in a self-sustaining "climax" adapted to local conditions, though Clementsian holistic views of deterministic progression have been critiqued in favor of , multiple-endpoint models incorporating dispersal limitations and priority effects. Resilience quantifies an ecosystem's capacity to absorb disturbances while maintaining core processes and structures, avoiding flips to alternative stable states. C.S. Holling formalized in 1973 as the magnitude of perturbation a system can tolerate before losing , contrasting with resilience focused on return speed to ; for example, coral reefs demonstrate low resilience to repeated bleaching events exceeding thermal thresholds, shifting to algal dominance. This property emerges from , functional redundancy, and connectivity, with empirical data from long-term experiments like the Hubbard Brook forest showing nitrogen retention resilience post-logging due to microbial and vegetative feedbacks. Succession contributes to resilience by rebuilding complexity, but chronic disturbances like climate-driven droughts can erode it, as in Amazonian forests where fragments reduce recovery from fires by 30-50%. Interactions among these elements underscore that moderate disturbances promote resilient, diverse states via the , where peak diversity occurs at optimal frequencies preventing competitive exclusion.

Biodiversity and Functioning

Diversity-Productivity Relationships

In ecosystem ecology, the diversity-productivity relationship examines how the number and variety of influence ecosystem productivity, typically measured as net primary productivity (NPP) or biomass accumulation. Experimental studies, such as those from long-term grassland manipulations like the Cedar Creek Ecosystem Science Reserve, have demonstrated positive effects where increasing plant enhances productivity through mechanisms like niche complementarity and selection for high-performing . A meta-analysis of 11 manipulation experiments across found that reductions in plant diversity decreased productivity comparably to resource limitations, underscoring 's role akin to abiotic drivers. However, observational data from natural ecosystems often reveal more complex patterns, including unimodal or negative relationships. In a 2023 of plot-level data from diverse biomes, a 10% increase in was causally linked to declining , attributed to competitive exclusion and environmental filtering rather than facilitative effects dominant in controlled settings. Forest inventories spanning temperate to tropical regions similarly show no consistent positive link, with relationships varying by scale—positive at local plots but neutral or inverse at levels—and influenced by dominant species traits over raw richness. Invasive species further complicate dynamics; a 2023 study of NPP gradients indicated that native-richness positives erode when exotics dominate, as invaders often monopolize resources and suppress diverse assemblages without boosting overall output. Critiques highlight methodological biases in experiments, such as exaggerated effects from non-random species selection or failure to partition interaction versus compositional changes, suggesting observational reversals better reflect causal realities in unconstrained systems. Across facets like phylogenetic or functional diversity, productivity correlations weaken or shift unimodal in high-diversity forests, challenging universal claims of biodiversity as a productivity driver. These discrepancies emphasize that while diversity can stabilize or enhance function under controlled conditions, natural gradients prioritize abiotic controls like soil fertility and climate, rendering positive relationships context-dependent rather than generalizable.

Empirical Patterns and Causal Mechanisms

Empirical studies, primarily from biodiversity-ecosystem functioning (BEF) experiments, reveal positive relationships between plant and key ecosystem processes such as primary and biomass production. Meta-analyses of over 300 experiments indicate that higher enhances aboveground biomass by 20-50% compared to monocultures, with effect sizes strengthening in more diverse assemblages due to reduced temporal variability and improved resource capture. In long-term grassland experiments spanning 15-17 years, initially boosts through dominant species, but over time, the relationship intensifies via sustained increases in community-level complementarity, leading to up to 30% higher yields in diverse plots. These patterns hold across scales, though observational data from natural forests and grasslands sometimes show weaker or context-dependent links, highlighting challenges in extrapolating manipulated designs to unmanaged systems. For and , demonstrates that diverse communities exhibit lower temporal variability in , with buffering against environmental fluctuations like or limitation. A global of sites found that reduces year-to-year fluctuations by enhancing asynchrony in species responses, effectively insuring function against losses. Similarly, higher correlates with greater carbon storage, as diverse plant assemblages increase belowground allocation and litter decomposition efficiency, amplifying accumulation by 10-25% in experimental settings. Patterns weaken in high-disturbance or nutrient-poor environments, where environmental filters override effects, underscoring the role of site-specific conditions. Causal mechanisms underlying these patterns include selection effects and complementarity effects, quantifiable through additive partitioning in experiments. Selection effects arise when diverse mixtures are disproportionately influenced by highly productive species that would dominate monocultures, accounting for 40-60% of the net biodiversity effect in early successional stages. Complementarity effects, however, drive long-term gains through niche partitioning, where co-occurring species exploit complementary resources—such as differing root depths or light requirements—leading to overyielding beyond single-species expectations. Evidence from trait-based analyses shows that functional trait diversity, rather than taxonomic richness alone, mediates these gains, with phylogenetic diversity enhancing complementarity in forests by reducing competitive exclusion. Additional mechanisms involve facilitation and reduced consumer pressure; diverse canopies in experiments foster mutualistic interactions that boost and dilution, indirectly elevating by 15-20%. For stability, the portfolio effect—statistical averaging of species-specific variabilities—causally links to lower ecosystem variance, as asynchronous prevent synchronous crashes. Critically, in observational contexts requires accounting for confounders like environmental heterogeneity; modern approaches using confirm that endogenous changes, not just exogenous drivers, propagate to functioning via these pathways, though reverse ( selecting for ) persists in some systems.

Methods and Analysis

Observational and Experimental Techniques

in ecosystem ecology primarily involve non-invasive monitoring to capture spatial and temporal patterns in biotic and abiotic components. Long-term ecological research (LTER) sites, supported by the since 1980, exemplify this approach, with 28 sites as of 2023 conducting standardized observations on , , trophic structure, accumulation, and spatial-temporal distributions of ecosystems. These sites employ plot-based sampling, sensor networks, and repeated censuses to track variables like accumulation and composition over decades, enabling detection of slow processes such as succession and climate-driven shifts. techniques, including and , complement ground-based observations by providing large-scale data on vegetation structure, productivity via indices like NDVI, and changes, particularly useful for inaccessible or heterogeneous landscapes. For instance, active has quantified forest canopy height and at resolutions down to meters, aiding assessments of carbon stocks and disturbance legacies. Challenges in observational methods include in data processing, of causal drivers from correlations, and issues across spatial extents, as outlined in frameworks like LIES (Latency, Identifiability, Effort, Scale). Despite these, integrating observations with models has improved insights into functioning, such as terrestrial carbon fluxes, by assimilating diverse data streams like measurements of net ecosystem exchange. Experimental techniques emphasize manipulative interventions to infer , contrasting with observational correlatives. Field manipulations alter specific factors—such as additions, exclusions via , or simulated disturbances like prescribed burns—within replicated plots or enclosures to isolate effects on processes like or . Whole-ecosystem experiments, such as the Hubbard Brook clearcuts initiated in 1965-1966, have demonstrated impacts of on export, with concentrations rising over 100-fold in treated streams compared to controls. In aquatic systems, lake whole-ecosystem manipulations, like whole-lake fertilization or acidification reversals at the Experimental Lakes Area since the 1960s, have quantified recovery trajectories and trophic cascades. Large-scale networks of manipulative experiments, including those simulating via warming chambers or elevated CO2, address and response thresholds across biomes, with over 100 such sites globally by 2021 revealing variable effects on turnover. Ethical considerations limit manipulations in sensitive ecosystems, such as tropical forests, where small-scale canopy manipulations have tested light-nutrient interactions without broad harm. Replication and controls mitigate risks, ensuring treatments exceed background variability, as emphasized in designs sampling pre-treatment baselines before interventions. These methods, while powerful for , face constraints and unintended artifacts from enclosures, necessitating validation against unmanipulated references.

Modeling Approaches

Modeling approaches in ecosystem ecology encompass conceptual, analytical, and frameworks designed to quantify biotic-abiotic interactions, energy flows, nutrient cycles, and responses to perturbations. These methods integrate empirical to test hypotheses about causal mechanisms driving ecosystem , such as trophic transfers and biogeochemical processes. Conceptual models provide qualitative overviews via diagrams, while analytical and simulation models employ mathematics and computation for quantitative predictions. Conceptual models represent ecosystems as interconnected compartments, illustrating flows of or without numerical simulation. A foundational example is Howard T. Odum's 1957 analysis of , which depicted trophic structure through an energy , estimating gross primary at 20,810 kcal/m²/year and highlighting detritus-based pathways dominating secondary . Such models aid in visualizing steady-state assumptions and identifying key pathways but lack for transient . Analytical models use mathematical equations to describe rates of change, often differential equations for continuous processes like or nutrient uptake. For instance, extensions of Lotka-Volterra predator-prey equations to multi-species food webs model trophic interactions by balancing birth, death, and consumption rates, enabling analysis of stability under varying parameters. These approaches excel in deriving analytical solutions for simple systems but struggle with nonlinearities and in complex ecosystems. Simulation models, typically computational, replicate behaviors over time through algorithms handling stochasticity and feedbacks. Ecopath with Ecosim (Ecopath, developed in the 1980s for mass-balanced snapshots; Ecosim for temporal ) simulates impacts on webs by parameterizing flows and predation mortalities, as applied in global assessments since the 1990s. Individual-based models (IBMs) track agent-level behaviors, such as or dispersal, aggregating to emergent properties like patterns; examples include simulations of predator reintroductions forecasting population recoveries. approaches combine multiple models to quantify uncertainty, as in marine projections integrating hydrodynamic and trophic components. These methods support scenario testing for disturbances but require extensive parameterization and validation against field data to avoid . Process-based models emphasize causal mechanisms, such as biogeochemical cycles via coupled differential equations for carbon, , and transformations, contrasting empirical-statistical fits reliant on correlations. Hybrid models increasingly incorporate for inputs like net primary productivity, enhancing scalability from plots to landscapes. Validation remains critical, with techniques like hindcasting against historical datasets ensuring models reflect empirical realities rather than artifacts of untested assumptions.

Data Challenges and Validation

Ecosystem ecology faces significant data challenges due to the inherent complexity and variability of natural systems, including spatial heterogeneity, temporal dynamics over decades or longer, and the difficulty of measuring elusive processes like belowground nutrient fluxes or microbial decomposition rates. For instance, sampling biases arise from uneven coverage of ecosystems, with remote or extreme environments (e.g., deep oceans or high altitudes) underrepresented, leading to incomplete representations of global patterns. Uncertainty in data collection is compounded by measurement errors, such as those in remote sensing estimates of biomass, which can vary by 20-50% due to atmospheric interference or sensor resolution limits, and by missing data from short-term studies that fail to capture rare disturbances like wildfires or floods. Integration of disparate data sources exacerbates these issues, as ecosystem studies often combine field observations, , and experimental manipulations, resulting in scale mismatches—e.g., plot-level data extrapolated to scales without accounting for patchiness—and unbalanced datasets where certain variables (e.g., ) are overrepresented relative to others like predator-prey interactions. Noisy or incomplete data further hinders analysis, with institutional barriers like constraints limiting long-term networks, as evidenced by the decline in sustained ecological observatories post-2010 due to cuts in agencies like the U.S. . Open data initiatives reveal additional hurdles, including lack of standardization in and dispersed repositories, which impede reusability and in macrosystems ecology. Validation of ecosystem data and models requires rigorous assessment to ensure reliability, typically involving comparison against independent empirical datasets, such as hindcasting model outputs to historical records or cross-validating predictions with field measurements from networks like the Long-Term Ecological Research (LTER) sites established since 1980. Techniques include to quantify parameter uncertainty, Bayesian methods for propagating input errors (e.g., in carbon flux estimates), and face validation where expert judgment evaluates mechanistic plausibility alongside quantitative metrics like root-mean-square error against observed trophic dynamics. distinguishes from validation by adjusting parameters to fit data, but over-reliance risks ; thus, hold-out validation sets, drawn from diverse biomes, are essential, as demonstrated in continental-scale tests where models matched 1675 data points with R² values exceeding 0.7 for provisioning services but underperformed for regulating ones due to unmodeled feedbacks. Despite advances, validation remains challenged by epistemic uncertainties, such as equifinality where multiple parameter sets yield similar outputs, necessitating ensemble modeling to bracket plausible ranges—e.g., in nutrient cycling simulations where decomposition rates vary by and , with uncertainties up to 30% in global models. Peer-reviewed critiques emphasize that incomplete validation, often due to data scarcity, can propagate errors in policy applications, underscoring the need for standardized protocols like those proposed for (, , , ) data to enhance credibility.

Applications and Outcomes

Resource Management and Sustainability

Ecosystem ecology informs by quantifying fluxes of energy, nutrients, and biomass to determine harvest rates that preserve system productivity and structure. , the extraction level allowing resource regeneration, underpins practices in and ; for instance, (MSY) in fisheries is modeled as the peak catch from a under constant environmental conditions, often estimated via logistic growth equations where equals minus natural mortality at half . However, MSY calculations assume stable single-species dynamics, frequently underestimating trophic cascades and environmental variability, leading to in 33% of global exceeding limits as of 2020 assessments. Ecosystem-based management (EBM) extends these principles to multi-trophic interactions, integrating human activities with ecological processes for resilience. In marine systems, EBM case studies from NOAA's Northeast U.S. shelf demonstrate reduced and habitat degradation through spatial zoning and predator-prey modeling, sustaining yields while maintaining ; for example, implementing ecosystem indicators like and adjusted quotas in the , preventing serial depletions observed in single-species approaches. In forestry, sustained yield protocols limit annual cuts to growth increments, preserving soil nutrient retention and ; a U.S. Forest Service analysis in the applied ecosystem modeling to balance timber harvest with fire regime restoration, yielding 1.2 billion board feet annually without net loss in old-growth equivalents since reforms. Sustainability requires adaptive monitoring of empirical indicators, such as biomass thresholds and recovery rates post-disturbance, to counter model uncertainties from variability and illegal harvesting. Failures, like the 1992 collapse of Newfoundland stocks despite MSY —reducing biomass by 99% from 1960s peaks due to unmodeled shifts—underscore the need for precautionary buffers and real-time data validation over static quotas. Successes in integrated , including loading controls to avert , have restored productivity in systems like the , where reductions since 1987 halved algal blooms while supporting fisheries valued at €2.5 billion yearly. reveals that prioritizing ecological baselines over short-term economic maximization enhances long-term yields, as evidenced by meta-analyses showing 20-50% higher stable outputs in EBM versus conventional regimes.

Policy Integration and Economic Valuation

Ecosystem ecology informs policy by quantifying ecosystem processes such as nutrient cycling and energy flows, which underpin services like and , enabling integration into frameworks like the U.S. (NEPA) and that require assessments of ecological impacts on . Payments for ecosystem services (PES) programs, grounded in ecological data on service provision, compensate landowners for maintaining habitats that deliver verifiable benefits, such as watershed protection in Costa Rica's program established in 1997, which has enrolled over 1.2 million hectares by 2020 through hydrological monitoring. In the , ecosystem-based adaptation policies under the EU Biodiversity Strategy incorporate ecological modeling to enhance against climate impacts, prioritizing restoration of services like flood regulation valued at €9-45 billion annually in avoided damages. Economic valuation of ecosystem services translates ecological functions into monetary terms to support policy decisions, using methods such as techniques like , which estimate values from market behaviors (e.g., premiums near forests reflecting air quality benefits), and stated preference approaches like surveys that elicit willingness-to-pay for non-market services. For instance, global meta-analyses have quantified regulating services like at $235-577 billion annually, informing agricultural policies that subsidize . In policy applications, these valuations underpin PES schemes, such as the U.S. Conservation Reserve Program, which pays farmers $1.7 billion yearly as of 2022 to retire cropland for soil retention and , with ecological validating service delivery via metrics like reduction. Despite methodological rigor, economic valuations face critiques for potential overestimation due to assumptions in future benefits or commensurability of services with goods, as noted in analyses arguing against aggregating total values which may exceed $33 trillion globally per outdated estimates prone to double-counting biophysical flows. Empirical challenges include data gaps in long-term ecological dynamics, leading some policies to rely on conservative valuations; for example, the UK's National Assessment adjusted service values downward based on causal from field experiments showing variable productivity responses to . Integration successes, however, demonstrate causal links: PES in Mexico's forests reduced by 7.5% from 2003-2012, per satellite and ground data, yielding net economic gains through sustained timber and water services. Such underscores the value of ecosystem ecology in prioritizing policies that align incentives with measurable biophysical outcomes over unsubstantiated projections.

Human-Dominated Systems

Human-dominated systems encompass landscapes where anthropogenic activities, such as , , and , exert primary control over ecological processes, overriding many natural drivers of energy flow, nutrient cycling, and community assembly. These systems, including agroecosystems and urban matrices, cover substantial portions of Earth's terrestrial surface; for example, human modification affects approximately 50% of global land with low human influence remaining, while broader alterations extend to 75% or more of ice-free land through land-use change and resource extraction. In these environments, ecosystem ecology examines altered trophic dynamics, such as simplified food webs with dominant primary producers like crops and reduced predator-prey interactions, alongside intensified biogeochemical cycles driven by fertilizers and impervious surfaces. Empirical studies reveal consistent patterns of —often 20-50% lower than in unmodified habitats—and heightened vulnerability to perturbations like or variability, though some novel assemblages emerge with functional redundancies. Agroecosystems exemplify managed human-dominated systems where ecological principles inform sustainable intensification; for instance, diversification practices enhance synergies between crops, pollinators, and soil microbes, reducing reliance on synthetic inputs by up to 30-50% in systems while maintaining yields equivalent to conventional monocultures. Key agroecological tenets, such as input reduction, promotion, and preservation, derive from observations of natural stability, where of and pest regulation via natural enemies sustain productivity without external subsidies. Causal analyses, including long-term experiments, demonstrate that these approaches mitigate nutrient runoff—responsible for 70% of global river —and bolster against droughts, as diversified systems recover 2-3 times faster from disturbances than uniform ones. However, implementation challenges persist, with economic pressures often favoring high-input models despite evidence of long-term yield declines in over-fertilized s. Urban ecosystems, another prevalent human-dominated type, feature disrupted hydrological and thermal regimes; impervious cover reduces by 50-90% and elevates local temperatures by 1-3°C, altering rates and in adjacent streams and soils. Ecosystem ecology applications here focus on restoring functions through , such as vegetated roofs and riparian buffers, which enhance retention—capturing up to 70% of rainfall events—and support urban food webs by providing corridors for mobile species like and . Studies in tropical and temperate cities quantify how these interventions boost ecosystem services, including valued at $10-20 billion annually in U.S. , though structural homogenization limits native diversity recovery without active management. In managed forests, selective and plantation monocultures simplify dynamics, with net primary productivity often 20-40% below natural stands due to reduced complexity, prompting ecological modeling to optimize harvest schedules for sustained timber yields and retention. Overall, outcomes in these systems hinge on integrating empirical trophic and flux data into policy, yielding measurable gains in where human interventions align with underlying causal mechanisms rather than ignoring biophysical limits.

Controversies and Critiques

Stability vs. Resilience Debates

In ecosystem ecology, traditionally denotes the capacity of a to maintain its and in the face of perturbations, often measured by (minimal change during disturbance) and engineering (rapid return to a pre-disturbance state). This perspective, rooted in early models like those of emphasizing and , posited that greater and complexity enhance through compensatory interactions and negative feedbacks. However, Robert May's 1973 analysis of models demonstrated that increasing connectance and in such idealized networks typically reduces local asymptotic , suggesting complex ecosystems are prone to chaotic fluctuations rather than inherent constancy. Empirical observations, such as oscillatory in predator-prey s, align with May's theoretical predictions, challenging the assumption of universal in diverse communities. C.S. Holling's seminal 1973 paper reframed the discussion by distinguishing —the ability to absorb disturbances, undergo reorganization, and retain essential functions and controls—from traditional focused on efficiency and minimal variance. Holling argued that real ecosystems operate in non-equilibrium conditions with multiple basins of attraction, where persistence arises not from rigidity but from and regime shifts, as evidenced by cyclic disturbances like fires in boreal forests or floods in wetlands that prevent collapse. This resilience paradigm critiques equilibrium-centric models for underestimating variability in heterogeneous environments, advocating instead for management strategies that buffer against thresholds, such as in fisheries where can trigger irreversible shifts to low-productivity states. Debates persist over whether stability and resilience are complementary or conflicting properties. Proponents of unification, as in a 2021 synthesis, propose integrating metrics like return time () with domain of attraction size () to capture both local robustness and global persistence, supported by simulations showing trade-offs where high resistance may reduce adaptability to novel stressors like . Critics, however, maintain the distinction is causal: stability assumes predictable equilibria verifiable through linear approximations, while resilience accommodates nonlinear dynamics and , as observed in coral reefs transitioning to algal dominance post-bleaching events in the 1990s and 2010s. Empirical meta-analyses indicate bolsters resilience in experimental grasslands but not always stability, with structured interactions (e.g., strong trophic controls) mitigating May-style instability more than randomness alone. These tensions influence , where stability-focused policies risk in dynamic systems, whereas resilience-oriented approaches prioritize variability tolerance, though both require validation against long-term data amid modeling assumptions that often overlook .

Model Limitations and Empirical Shortfalls

Ecosystem models in often rely on simplifying assumptions about interactions, cycling, and environmental drivers, which can lead to oversimplification of . For instance, many models assume linear or equilibrium-based relationships that fail to account for nonlinear feedbacks, , or emergent properties arising from multi-trophic interactions. This limitation is exacerbated in transferability efforts, where models calibrated to one perform poorly in others due to unmodeled factors like interactions and environmental nonstationarity. sensitivity further undermines reliability, as small changes in input values—often derived from sparse data—can yield divergent outcomes, highlighting the gap between model elegance and real-world . Predictive accuracy remains a core shortfall, with ecosystem models frequently failing to forecast population responses to perturbations or management interventions. A 2024 analysis of multispecies models demonstrated poor out-of-sample predictions for conservation scenarios, attributing failures to calibration processes that cannot distinguish between plausible interaction structures within ecosystems. In fisheries applications, similar models exhibit limited foresight for stock dynamics under exploitation, often overestimating stability or underestimating collapse risks due to unvalidated assumptions about predator-prey couplings. These issues persist despite computational advances, as models struggle with rare events, regime shifts, and long-term transients that empirical observations rarely capture comprehensively. Empirical validation faces profound challenges from data scarcity and methodological constraints. Long-term, high-resolution datasets are rare for most ecosystems, limiting tests of model projections against independent observations and fostering overfitting to historical patterns rather than causal mechanisms. Short time series, common in ecological monitoring, hinder detection of underlying dynamics in empirical dynamic modeling approaches, reducing confidence in extrapolations to future states. Experimental manipulations at ecosystem scales are logistically infeasible or ethically restricted, leaving models reliant on correlative data that conflate causation with coincidence, particularly under multiple stressors like climate variability. Consequently, many models prioritize explanatory power over verifiable foresight, with validation often confined to hindcasting rather than prospective testing.

Policy Overreach and Ideological Biases

Policies derived from ecosystem ecology have occasionally extended beyond empirical support, particularly when relying on equilibrium-based models that assume stable states amenable to precise human control. Such approaches, embedded in frameworks like the U.S. Endangered Species Act since the , presuppose ecosystems revert to fixed balances post-disturbance, justifying interventions like exclusion zones or harvest quotas aimed at predictability. However, empirical analyses reveal that suppressing natural variability—such as through levees for or rigid shrub management on rangelands—amplifies risks of , as variability buffers against extremes in nutrient-polluted lakes, overharvested fisheries, and grazed grasslands. This overreach manifests in strategies that prioritize short-term over long-term adaptability, contracting the "safe operating space" for exploited systems by up to 50% in simulated scenarios. The persistence of equilibrium paradigms in policy, despite their partial rejection in ecological theory since the 1990s, stems from institutional inertia and nonequilibrium critiques highlighting transient dynamics and stochasticity. For example, wildlife management policies enforcing "balance of nature" concepts often misguide decisions by underestimating regime shifts, leading to ineffective protections that fail to account for ecological flux. Overreach intensifies when these models underpin expansive regulations, such as biodiversity offsets or restoration mandates, where equivalence assumptions ignore context-specific contingencies, resulting in averted losses overstated as gains without verifiable net benefits. Ideological biases compound these issues, with environmental research exhibiting personal, institutional, and socio-cultural distortions that skew policy toward control-oriented interventions. Personal biases arise from scientists' self-interests or ideological commitments, while institutional pressures—prevalent in academia and funding bodies—favor narratives aligning with preservationist agendas over utilitarian trade-offs. Socio-cultural emphases on mechanistic dominance of nature, rooted in Western paradigms, further bias ecology toward policies undervaluing human adaptation, as seen in critiques of ecosystem services frameworks that imprecisely conflate biophysical processes with anthropocentric valuations, marginalizing non-economic elements and justifying regulatory expansions without causal rigor. These biases, amplified by shared disciplinary values among ecologists favoring anti-anthropocentric views, often yield policies that prioritize intrinsic nature values at the expense of empirical cost-benefit analysis, particularly in left-leaning institutions where alarmist extrapolations from models eclipse data on human-dominated resilience.

Recent Advances

Technological Integrations

Remote sensing technologies, including satellite imagery and LiDAR, have become integral to ecosystem ecology by enabling large-scale monitoring of vegetation structure, land cover changes, and biodiversity patterns. For instance, hyperspectral remote sensing data from platforms like NASA's Earth Observing System allow ecologists to quantify ecosystem productivity and detect invasive species across vast areas, with resolutions improving to sub-meter scales in recent missions such as the European Space Agency's Sentinel-2 launched in 2015 and updated through 2020s reorbits. Geographic Information Systems (GIS) integrate these datasets with ground-based observations to model spatial dynamics, such as nutrient cycling in watersheds, supporting causal analyses of disturbance effects like deforestation, which reduced global forest cover by 420 million hectares between 1990 and 2020. Machine learning algorithms enhance the processing of remote sensing data in ecosystem ecology, automating the classification of habitat types and prediction of species distributions with accuracies exceeding 90% in some forest inventory applications. Convolutional neural networks, applied to multispectral imagery, have identified ecosystem shifts due to climate variability, as demonstrated in studies forecasting mangrove extent changes with inputs from Landsat archives spanning 1984 to 2023. These tools address data volume challenges by extracting patterns from petabyte-scale repositories, though validation against empirical field data remains essential to mitigate overfitting risks inherent in black-box models. Wireless sensor networks and (IoT) devices provide high-frequency, in-situ measurements for validating remote and modeled data, capturing variables like and atmospheric CO2 fluxes in across heterogeneous terrains. Deployments in ecosystems, for example, have utilized low-power IoT nodes to monitor hydrological regimes, revealing diurnal carbon exchange variations with precision comparable to traditional towers but at scales up to 100 times larger since implementations scaled post-2020. Integration with AI-driven analytics further refines these networks for , such as early warning of drought-induced die-offs, enhancing in ecosystem responses to perturbations.

Global Change Responses and Predictions

Ecosystems respond to global change drivers—including rising atmospheric CO₂ concentrations, warming temperatures, altered precipitation patterns, and nitrogen deposition—through shifts in primary productivity, nutrient cycling, and trophic dynamics. Long-term empirical observations from sites like the U.S. Long-Term Ecological Research Network reveal enhanced net primary production in some temperate forests under elevated CO₂, attributed to fertilization effects, but these gains are often transient due to nutrient limitations and increased respiration rates. Warming has accelerated decomposition and carbon release from soils, as evidenced by increased heterotrophic respiration in boreal ecosystems, potentially reducing soil carbon stocks by 10-20% in vulnerable regions over decades. These responses vary by biome: tropical systems show heightened sensitivity to drought-induced tree mortality, while Arctic tundra experiences shrub encroachment and permafrost thaw, releasing methane and altering energy flows. Biodiversity loss interacts with climatic drivers to amplify ecosystem vulnerabilities, reducing functional redundancy and impairing recovery from disturbances. Empirical data indicate that species richness declines correlate with diminished complementarity effects in resource use, leading to less efficient ; for instance, grasslands with halved diversity under CO₂ enrichment and nitrogen addition exhibit 15-30% lower productivity stability. In marine and terrestrial systems, warming disrupts predator-prey interactions, as seen in phenological mismatches between pollinators and plants, which can cascade to lower trophic levels and reduce overall . These patterns underscore causal links where biodiversity erosion—driven partly by exceeding climatic range shifts—exacerbates sensitivity to warming, with meta-analyses showing 20-50% greater impacts on low-diversity assemblages. Predictive models in ecosystem ecology, integrating empirical data with dynamic global vegetation frameworks, forecast widespread disruptions under high-emission scenarios like SSP5-8.5. Projections estimate a 20.8% global decline in belowground exogenous mycorrhizal fungi by 2100, impairing uptake and growth in temperate and zones, with cascading effects on . Terrestrial carbon storage faces losses of 7.4-103 PgC from combined and land-use driven declines, as diverse systems buffer against warming-induced spikes. ecosystem may persist structurally until mid-century but risks tipping points from compounded stressors, with models predicting 10-25% reductions in resistance to perturbations. Uncertainties persist in scaling local experiments to global predictions, particularly regarding acclimation feedbacks like microbial adaptations that could moderate CO₂ fertilization. These forecasts emphasize the need for integrated assessments accounting for multiple drivers, as isolated projections often overestimate by underweighting synergies.

References

  1. [1]
    [PDF] Ecosystem Ecology - NSF Public Access Repository
    Ecosystem ecology is a branch of study and thinking within the ecological sciences that focuses on the ecosystem – a dynamic complex of interactions of ...
  2. [2]
    Eugene Odum - New Georgia Encyclopedia
    Eugene Pleasants Odum, called “the father of modern ecology,” brought the word ecosystem into common parlance by making it the organizing concept in his 1953 ...
  3. [3]
    Origins of Ecosystem Ecology - Part3 - Appalachian State University
    On campus, Odum started the Institute of Ecology, which actually directs the work at Savannah River, and attracted a large group of ecologists with interests ...
  4. [4]
    [PDF] ECOSYSTEM ECOLOGY - Yiqi Luo
    Ecosystem ecology is a subdiscipline of ecology that focuses on exchange of energy ... Global change. An original equilibrium can be defined at a reference.
  5. [5]
    10 Limitations on Progress in Ecology - UBC Zoology
    Dec 23, 2013 · 1. A shortage of ecologists or more properly jobs for ecologists. In particular a scarcity of government agencies employing ecologists in secure jobs.Missing: achievements controversies
  6. [6]
    Introduction to Ecosystem Ecology I: Energy Flow and Nutrient Cycles
    Ecosystem ecology is an extension of organismal, population, and community ecology. The ecosystem comprises all the biotic components (living things) and ...
  7. [7]
    The Ten Principles of Ecology | The Kaspari Lab
    Jul 17, 2017 · 1. Evolution organizes ecological systems into hierarchies. Individual organisms combine into populations, populations combine into species, species combine ...
  8. [8]
    Principles of Ecology Revisited: Integrating Information ... - Frontiers
    Principle 1: Information is a fundamental feature of living systems, and therefore also of all ecological systems.Abstract · Principles of Ecology, Revisited · Advancing the Role of... · Conclusion
  9. [9]
    [PDF] Ecosystem Ecology: In Pursuit of Principles - SUNY ESF
    The first challenge in teaching ecosystem ecology is to figure out where to start, given that everything depends on everything else. This book manages to convey ...
  10. [10]
    Principles of Ecosystem Sustainability | The American Naturalist
    Principles of Ecosystem Sustainability. F. Stuart Chapin III,; Margaret S ... ecosystem ecology, Biodiversity and Conservation 23, no.77 (Mar 2014): 1619 ...
  11. [11]
    [PDF] Eleven Basic Ecological Concepts for Natural Resource Management:
    Ecosystem ecology considers biotic and abiotic components as an integrated system and, as such ... Principles of terrestrial ecosystem ecology. Springer.
  12. [12]
    35.1 The Scope of Ecology - Biology for AP® Courses | OpenStax
    Mar 8, 2018 · Ecology integrates multiple scientific disciplines ... Ecosystem ecology is an extension of organismal, population, and community ecology.
  13. [13]
    Linking biodiversity and ecosystems: towards a unifying ecological ...
    Community ecology and ecosystem ecology provide two perspectives on complex ecological systems that have largely complementary strengths and weaknesses.<|separator|>
  14. [14]
    Moving toward a new era of ecosystem science - ScienceDirect
    Ecosystem science, also known as ecosystem ecology, focuses on the patterns, processes, and services of ecosystems, and explores how biological communities and ...
  15. [15]
    History of Ecological Sciences, Part 47: Ernst Haeckel's Ecology
    Jul 1, 2013 · Ernst Haeckel, one of the world's best-known and most-read zoologists, named and defined a new science, “Oecologie” (1866, II:286).
  16. [16]
    History of Ecological Sciences, Part 59: Niches, Biomes, Ecosystems ...
    Sep 29, 2017 · Two American ecologists who led the way in developing systems ecology were EP Odum and Stanley Auerbach (1922–2004), at the Oak Ridge National ...Pre-biome · Ecosystem · Post-war investigations · Systems ecology
  17. [17]
    [PDF] The Lake as a Microcosm - Illinois Open Publishing Network
    The Lake as a Microcosm*. By Stephen A. Forbes. A lake is to the naturalist a chapter out of the history of a primeval.
  18. [18]
    Tansley, A.G. 1935: The use and abuse of vegetational concepts ...
    The concept of the ecosystem was introduced in 1935 by Arthur Tansley and reflects the ecological connection between the environment and organisms, creating ...
  19. [19]
    A History of the Ecosystem Concept in Ecology - Yale University Press
    Feb 21, 1996 · Odum helped to popularize the ecosystem concept and how numerous other scientists investigated its principles and published their results.
  20. [20]
    The Trophic‐Dynamic Aspect of Ecology - Lindeman - 1942
    The Trophic-Dynamic Aspect of Ecology. Raymond L. Lindeman,. Raymond L ... First published: 01 October 1942. https://doi.org/10.2307/1930126. Citations ...
  21. [21]
    [PDF] LINDEMAN'S TROPHIC- DYNAMIC ASPECT OF ECOLOGY
    In the fall of 1941, only six years after the word “ecosystem” was coined, Raymond Lindeman submitted “The trophic- dynamic aspect of ecology” (the final ...
  22. [22]
    [PDF] Raymond Laurel Lindeman and the Trophic Dynamic Viewpoint
    Raymond Lindeman was an intellectually daring, determined young man whose insightful paper, “The Trophic. Dynamic Aspect of Ecology” (Lindeman 1942) has ...
  23. [23]
    Eugene Pleasants Odum | Biographical Memoirs: Volume 87
    Eugene P. Odum was recognized nationally and internationally as a pioneer of ecosystem ecology. It is rare that an individual makes major contributions.<|control11|><|separator|>
  24. [24]
    Through the MACROSCOPE: the legacy of HT Odum
    His first ecosystem studies were conducted on the Silver Springs in the early 1950s. ... Eugene Odum, Ecosystem Ecologist and Environmentalist. University of ...
  25. [25]
    Eugene and Howard Odum - Ecology - Oxford Bibliographies
    Oct 25, 2017 · During the early 1950s, they conducted an important study of ecosystem function on Eniwetok Atoll, the site of US nuclear weapons testing. Like ...
  26. [26]
    [PDF] The International Biological Program (IBP) - National Academies
    Aug 24, 2004 · The committee recommended that the IBP focus on the effect on biological communities of changes in the natural environment, and on the ...
  27. [27]
    [PDF] The rise of ecosystem ecology and its applications to environmental ...
    Nov 13, 2015 · Also, in the late 1960s and early 1970s period, a new public and institutional awareness of compelling envi- ronmental problems was emerging ...
  28. [28]
    About the Network - LTER
    Jan 28, 2025 · To promote training, teaching, and learning about long-term ecological research and the Earth's ecosystems, and to educate a new generation of ...
  29. [29]
    The International Long Term Ecological Research Network: a ...
    Feb 21, 2017 · The ILTER Network differs from many other global networks in that members engage in long-term, site-based ecological research. ILTER researchers ...
  30. [30]
    LTER Network
    We are over 2000 researchers at 27 sites who apply long-term observation, experiments, and modeling to understand how ecological systems function over decades.LTER Site Profiles · Opportunities · About the Network · Using LTER Data
  31. [31]
    History of Ecosystem Model Development at Colorado State ...
    Jan 15, 2019 · During the 1980s CENTURY was used for site-level simulations, while regional application was limited due to lack of climate, soils, and plant ...
  32. [32]
    The value of the world's ecosystem services and natural capital
    May 15, 1997 · We have estimated the current economic value of 17 ecosystem services for 16 biomes, based on published studies and a few original calculations.
  33. [33]
    The history of ecosystem services in economic theory and practice
    Apr 1, 2010 · The origins of the modern history of ecosystem services are to be found in the late 1970s. It starts with the utilitarian framing of beneficial ...
  34. [34]
    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).
  35. [35]
    On the History of Ecosystem Dynamical Modeling - PubMed Central
    Nov 8, 2023 · Since the 1980s, several qualitative frameworks have been developed, such as loop analysis [9], qualitative reasoning models [12], fuzzy models ...
  36. [36]
    [PDF] Ecosystem trophic foundations: Lindeman exonerata
    Thus, one is led to imagine a trophic pyramid of energy flow, where the amount of energy transferred during successive feeding events (as represented by the ...
  37. [37]
    [PDF] Productivity, trophic structure, and energy flow in the steady-state ...
    We briefly describe the ambitious data col- lection program Odum developed to quantify trophic energy flows in this ecosystem, as well as his clever methods ...
  38. [38]
    Food Web: Concept and Applications | Learn Science at Scitable
    Some trophic relationships are more important than others in dictating how energy flows through ecosystems.
  39. [39]
    (PDF) Productivity, trophic structure, and energy flow in the steady ...
    Aug 6, 2025 · (LTER, Callahan, 1984). Odum's characterization of Silver Springs as a. steady-state ecosystem with inputs of constant tem-. perature, flow and ...
  40. [40]
    Food webs speak of human impact: Using stable isotope-based ...
    In their essence, food webs are adaptive networks of nodes (trophic entities) interacting over fluxes of energy and matter within and across ecosystems (Fig. 1a) ...
  41. [41]
    The Nitrogen Cycle: Processes, Players, and Human Impact - Nature
    The nitrogen cycle involves nitrogen fixation, nitrification, denitrification, anammox, and ammonification, where nitrogen changes forms and is critical for ...
  42. [42]
    Nutrient Cycling and Limitation: Hawaii as a Model System
    Aug 10, 2025 · One such ecosystem process is the cycling of nutrients within ecosystems ( Vitousek, 2004 ). The importance of nutrient cycling for ...
  43. [43]
  44. [44]
    General reversal of N-decomposition relationship during long-term ...
    May 10, 2024 · Decomposition of dead organic matter is fundamental to carbon (C) and nutrient cycling in terrestrial ecosystems, influencing C fluxes from the ...
  45. [45]
    Human activities shape global patterns of decomposition rates in rivers
    May 30, 2024 · ... empirical measurements with fine-scale geospatial and environmental data ... Global Biogeochemical Cycles 36, e2021GB007163 (2022). Crossref · Web ...
  46. [46]
    A meta-analysis on decomposition quantifies afterlife effects of plant ...
    Sep 11, 2020 · Here, we provide a global-scale meta-analysis to quantify how changes in the diversity of organic matter derived from plants (i.e. litter) ...
  47. [47]
    Long-term litter decomposition controlled by manganese redox cycling
    Sep 8, 2015 · Here, we show how the ability of the integrated plant–soil system to promote active redox cycling of manganese (Mn) regulates litter decomposition.<|separator|>
  48. [48]
    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, ...
  49. [49]
  50. [50]
    [PDF] Disturbance processes and ecosystem management
    Disturbance--White and Pickett (1985) define disturbance as "any relatively discrete event in time that disrupts ecosystem, community, or population structure.
  51. [51]
    Disturbance and change in biodiversity - PMC - NIH
    The word disturbance is used in ecology to refer to a great variety of phenomena. Examples of disturbance include fires, storms, diseases, volcanic eruptions, ...
  52. [52]
    [PDF] Comparing Large, Infrequent Disturbances: What Have We Learned?
    Large, infrequent disturbances (LIDs) are large in spatial extent and infrequent, like volcanic eruptions, big fires, and extreme floods or storms. Examples ...
  53. [53]
    Disturbance theory for ecosystem ecologists: A primer - PMC
    May 30, 2024 · For example, disturbance has non‐random impacts that are dependent upon frequency, severity, source, and duration, resulting in the retention of ...
  54. [54]
    Difference Between Primary And Secondary Succession - BYJU'S
    When natural communities are developed in a barren habitat with no soil or extremely less soil, it is called primary succession. Abiotic factors such as water, ...
  55. [55]
    What Is the Difference Between Primary and Secondary Ecological ...
    Sep 26, 2025 · Secondary succession differs from primary succession in that it begins after a major disturbance—such as a devastating flood, wildfire, ...
  56. [56]
    Differences between primary and secondary plant succession ...
    Sep 25, 2018 · Primary and secondary seres differed significantly in trends of species richness and in importance of aliens in late successional stages, but ...
  57. [57]
    Ecological Resilience–In Theory and Application - ResearchGate
    Aug 6, 2025 · Ecological resilience was defined as the amount of disturbance that an ecosystem could withstand without changing self-organised processes and structures.
  58. [58]
    Biodiversity promotes primary productivity and growing season ...
    Sep 5, 2017 · Research of the past decades has shown that biodiversity promotes ecosystem functions including primary productivity.
  59. [59]
    Biodiversity impacts ecosystem productivity as much as resources ...
    Results of 11 experiments show that ecologically relevant decreases in grassland plant diversity influenced productivity at least as much as ecologically ...
  60. [60]
    Clarifying the effect of biodiversity on productivity in natural ...
    May 5, 2023 · Contrary to many prior studies, we estimate that increases in plot-level species richness caused productivity to decline: a 10% increase in ...
  61. [61]
    Tree species richness and its effects on productivity: neither global ...
    Jan 22, 2019 · The relationship between species richness and ecosystem function is a field of ecology that has always puzzled me.
  62. [62]
    Review of biodiversity–productivity relationships in forests published ...
    Mar 29, 2023 · Fourthly, the relationship varies markedly across spatial scales and may not be strongly positive at scales relevant to conservation. Lastly, ...
  63. [63]
    [PDF] Net primary productivity relationships are eliminated by invasive ...
    Nov 6, 2023 · These results indicate that relationships between NPP and richness depend on whether native or exotic species are dominant, and that exotic ...<|separator|>
  64. [64]
    Partitioning changes in ecosystem productivity by effects of species ...
    Jun 28, 2024 · Species interactions affect ecosystem productivity. Positive interactions (resource partitioning and facilitation) increase productivity while negative ...
  65. [65]
    The productivity-biodiversity relationship varies across diversity ...
    Dec 12, 2019 · We demonstrate that unimodal productivity–biodiversity relationships only dominate for taxonomic diversity. In forests, trait and phylogenetic diversity ...
  66. [66]
    We should not necessarily expect positive relationships between ...
    Sep 16, 2021 · Our review indicates that we should not necessarily expect any specific relationship between local biodiversity and ecosystem function in observational field ...
  67. [67]
    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.
  68. [68]
    Biodiversity promotes ecosystem functioning despite environmental ...
    Dec 7, 2021 · Moreover, several continental‐scale studies showed that plant diversity increased ecosystem productivity or biomass in dry regions, but such ...
  69. [69]
    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 and
  70. [70]
    Biodiversity effects on ecosystem functioning in a 15-year grassland ...
    Biodiversity effects on ecosystem functioning in a 15-year grassland experiment: Patterns, mechanisms, and open questions
  71. [71]
    Global evidence of positive biodiversity effects on spatial ecosystem ...
    Jul 19, 2019 · We show that biodiversity stabilizes rather than increases plant productivity in natural grasslands at the global scale.Results · Biodiversity And... · Biodiversity And Spatial...
  72. [72]
    Species richness promotes ecosystem carbon storage - Journals
    Nov 25, 2020 · This meta-analysis provides strong empirical evidence that plant species richness significantly enhances ecosystem C storage, with stronger ...Introduction · Methods · Results · Discussion<|separator|>
  73. [73]
    Unravelling biodiversity–productivity relationships across a large ...
    Sep 22, 2021 · We found that the spatial extent of the tree community could change the importance of the multiple facets of biodiversity on productivity.
  74. [74]
    Scaling‐up biodiversity‐ecosystem functioning research - PMC
    Biodiversity‐ecosystem functioning (BEF) has focused on isolating the causal pathways by which biodiversity change alters the magnitude and stability of ...
  75. [75]
    Modern causal inference approaches to investigate biodiversity ...
    Apr 6, 2023 · A typical observational causal study would start with building a statistical model of species richness and ecosystem productivity, as the naive ...
  76. [76]
    A graphical causal model for resolving species identity effects and ...
    Apr 17, 2020 · Formal causal analysis showed that biodiversity–ecosystem function correlations are noncausal associations. Using simulations, we showed how ...
  77. [77]
    The Use and Potential of Remote Sensing for Studying Ecosystem ...
    Sep 7, 2022 · Remote sensing techniques are increasingly used for studying ecosystem dynamics, delivering spatially explicit information on the properties of Earth.
  78. [78]
    Active Remote Sensing for Ecology and Ecosystem Conservation
    Active remote sensing data and methods have been mainly leveraged for answering ecological questions via providing information on forest structure, function ...
  79. [79]
    LIES of omission: complex observation processes in ecology
    We propose the LIES framework (defining observation processes in terms of issues of Latency, Identifiability, Effort and Scale)
  80. [80]
    Combining Observations and Models: A Review of the CARDAMOM ...
    Aug 26, 2025 · The rapid increase in the volume and variety of terrestrial biosphere observations offers a unique opportunity to derive ecological insights, ...
  81. [81]
    Experimental ecology and the balance between realism ... - Nature
    Jun 3, 2025 · We argue that many of the challenges we currently face can be overcome by embracing multidimensional ecological experiments, moving beyond classical model ...
  82. [82]
    Long Term Ecological Research (LTER)
    Established as an LTER in 1988, the focus of research has consistently been on the impact of disturbance on the Northern Forest.<|separator|>
  83. [83]
    Manipulative experiments networks on response and adaptation of ...
    Manipulative experiments with changing biological or environmental factors and large-scale field experiment networks were regarded as effective approaches.Missing: techniques | Show results with:techniques
  84. [84]
    Whole-ecosystem experimental manipulations of tropical forests
    Mar 6, 2015 · Ethics of manipulating tropical forest ecosystems. Experimental manipulation can harm tropical forests, but is also one of the most efficient ...
  85. [85]
    Sampling design of ecological experiments - David Zelený
    Oct 14, 2022 · In manipulative experiments, we first sample baseline data (before the treatment is applied), then apply the treatment, and then (after allowing ...Introduction · Design of manipulative... · Beware of pseudoreplication...Missing: techniques | Show results with:techniques
  86. [86]
    A review of landscape ecology experiments to understand ...
    Sep 9, 2022 · This review explores the intersection between experimentation, ecological processes and landscape ecology.
  87. [87]
    A guide to ecosystem models and their environmental applications
    Sep 14, 2020 · We provide an overview of the main types of ecosystem models and their uses, and discuss challenges related to modelling complex ecological systems.
  88. [88]
    6.5: Ecosystem models and remote sensing - Geosciences LibreTexts
    Dec 26, 2023 · Ecosystem models include conceptual models (flow charts), analytical models (mathematical formulas), and simulation models (computer algorithms ...
  89. [89]
    Trophic Structure and Productivity of Silver Springs, Florida - Odum
    Trophic Structure and Productivity of Silver Springs, Florida. Howard T. Odum, Howard T. Odum. Search for more papers by this author.
  90. [90]
    Addressing data integration challenges to link ecological processes ...
    Feb 1, 2021 · We highlight four common analytical challenges to data integration in macrosystems ecology research: data scale mismatches, unbalanced data, sampling biases, ...
  91. [91]
    [PDF] Addressing data integration challenges to link ecological ... - Directory
    Data integration challenges include data scale mismatches, unbalanced data, sampling biases, and model development and assessment.
  92. [92]
    Recommendations for quantitative uncertainty consideration in ...
    Uncertainty in input data is common in ecology and evolution, and arises from imperfect mapping between the variable of interest and the observed data [16,20,23] ...
  93. [93]
    The current state of uncertainty reporting in ecosystem studies: a ...
    Jun 8, 2021 · Sampling error—the uncertainty in replicate measurements—was the source most commonly reported in studies of vegetation (84%), soil mass (56%), ...
  94. [94]
    Overcoming the challenges of data integration in ecosystem studies ...
    Apr 10, 2024 · Challenges include defining conceptual models, data quantity differences, missing/noisy data, model optimization, accurate predictions, and ...
  95. [95]
    Too much and not enough data: Challenges and solutions for ...
    Mar 27, 2025 · Each type of data challenge can have multiple underlying causes, resulting from ecosystem attributes as well as from financial, institutional, ...
  96. [96]
    [PDF] Challenges and Opportunities of Open Data in Ecology - eScholarship
    Open data in ecology faces challenges like dispersed data, lack of standardization, and inadequate rewards for sharing, but benefits from open access and the ...<|separator|>
  97. [97]
    A Continental-Scale Validation of Ecosystem Service Models
    Apr 22, 2019 · We describe a study of multiple models of five ES, which we validate at an unprecedented scale against 1675 data points across sub-Saharan Africa.
  98. [98]
    Is my model fit for purpose? Validating a population ... - ESA Journals
    Sep 17, 2023 · Model validation is critically important to evaluate and reduce model uncertainty, to determine whether models are fit for purpose, and to ...<|separator|>
  99. [99]
    A practical guide to understanding and validating complex models ...
    Nov 18, 2022 · ' Thus, validation includes everything from whether code is correct, to whether parameters are identifiable and estimates unbiased, to whether ...
  100. [100]
    Testing ecological models: the meaning of validation - ScienceDirect
    Validation means that a model is acceptable for its intended use because it meets specified performance requirements.
  101. [101]
    WEC216/UW256: Validation, Verification, and Calibration
    Verification checks if the model's formalism is correct. Validation checks the model's accuracy. Calibration adjusts model parameters to improve output.
  102. [102]
    Rigorous (in)validation of ecological models | bioRxiv
    Sep 23, 2024 · Common approaches in ecological model validation and limitations. Panels (A) and (B) depict a classic example of predator-prey dynamics: the ...
  103. [103]
    Guidance framework to apply best practices in ecological data ...
    Feb 12, 2025 · We propose a practical and operational framework for researchers and experts in ecology to achieve best practices for building analytical procedures.
  104. [104]
    Challenges for Marine Ecological Assessments: Completeness of ...
    Jan 26, 2022 · We review existing findability, accessibility, interoperability, and reusability (FAIR) biodiversity data coming from one of the most reliable online ...
  105. [105]
    Maximum Sustainable Yield - an overview | ScienceDirect Topics
    Maximum sustainable yield (MSY) is the highest average catch that can be continuously taken from a population, especially in fisheries, over an indefinite ...
  106. [106]
    Maintaining ecological stability for sustainable economic yields of ...
    Sep 25, 2025 · Real-world fisheries are mostly managed using maximum sustainable yields to maximize the value of net revenues, which requires foreseeing ...
  107. [107]
    Ecosystem-Based Fisheries Management Case Studies
    Apr 16, 2024 · Three case studies highlighting ecosystem-based fisheries management implementation.
  108. [108]
    Sustained Yield Use In Forestry And Natural Resource Management
    Sustained yield use allows people to obtain benefits from forests and other reserves of natural resources without undermining their ecological integrity.
  109. [109]
    [PDF] USGS DDS-43, Case studies in ecosystem management
    The Mammoth-June case study examines how a single na- tional forest is attempting to implement the new Forest. Service policy for ecosystem analysis.
  110. [110]
    Ecologically Sustainable Yield | American Scientist
    Marine conservation requires a new ecosystem-based concept for fisheries management that looks beyond sustainable yield for individual fish species.
  111. [111]
    The emergence of ecosystem management as a tool for meeting ...
    Ecosystem management is an approach that attempts to involve all stakeholders in defining sustainable alternatives for the interactions of people and the ...
  112. [112]
    Ecosystem-Based Management
    A fundamental principle of EBM is that individual ecosystem components (biological, physical, chemical, social, cultural, economic) are intrinsically linked to ...<|separator|>
  113. [113]
    [PDF] Best Practices for integrating ecosystem services into federal ...
    These laws include the National Environmental Policy Act, the Clean Water Act, the Endangered Species Act, the Forest. Land Policy and Management Act, and the ...
  114. [114]
    Designing payments for ecosystem services: Lessons from previous ...
    Payments for ecosystem services (PES) policies compensate individuals or communities for undertaking actions that increase the provision of ecosystem ...<|separator|>
  115. [115]
    Ecosystem-based approaches | Adaptation in EU policy sectors
    Ecosystem-based approaches focus on ecosystem restoration and enhancement of ecosystem services to protect society against negative impacts of climate change.
  116. [116]
    Economic Valuation of Ecosystem Services - ScienceDirect.com
    Economic valuation of ecosystem services (ES) refers to the process of assessing the social-economic benefits and costs associated with natural ecosystems, ...
  117. [117]
    Economic value of regulating ecosystem services - PubMed
    Sep 6, 2019 · This study is the first meta-regression analysis of the economic value of regulating ecosystem services at the global level.Missing: critiques | Show results with:critiques
  118. [118]
    Payments for Ecosystem Services - Ecosystem Marketplace
    Payments for Ecosystem Services (PES) are economic arrangements rewarding conservation, often through direct public payments to landowners.
  119. [119]
    Why not to calculate the value of the world's ecosystem services and ...
    Ecosystem Services economic Valuation (ESV) is often seen as a tool that can potentially enhance our collective choices regarding ecosystem services as it ...
  120. [120]
    The First Economic Valuation of Global Ecosystem Services - NCEAS
    The working group found that, every year, the world's diverse ecosystems combined produce an average of $33 trillion USD worth of ecosystem services.
  121. [121]
    [PDF] Guidance on the economic valuation of ecosystem services and ...
    Oct 26, 2023 · Economic valuation of ecosystem services involves identifying and quantifying the contribution of environmental resources to human well-being; ...
  122. [122]
    Payments for ecosystem services programs: A global review of ...
    Jan 15, 2024 · Payments for ecosystem services programs (PESPs) are increasingly being adopted globally to enhance sustainability outcomes.
  123. [123]
    50% of Earth's Land Surface Remains Relatively Untouched By ...
    Aug 25, 2020 · They conclude that just over 50% of Earth's land surface can be classified as having low human influence or being untouched completely, with a ...
  124. [124]
    (PDF) Human Domination of Earth's Ecosystems - ResearchGate
    Aug 10, 2025 · Human alteration of Earth is substantial and growing. Between one-third and one-half of the land surface has been transformed by human action.<|separator|>
  125. [125]
    Human-modified ecosystems and future evolution - PMC
    Human-modified ecosystems are shaped by our activities and their side effects. They share a common set of traits including simplified food webs, landscape ...
  126. [126]
    Human-Dominated Ecosystems - Science
    In this Special Issue, Science takes a look at the progress made to date in the study of human influence on the planet.
  127. [127]
    Agroecological principles and elements and their implications for ...
    Oct 27, 2020 · 13 consolidated agroecological principles: recycling; input reduction; soil health; animal health; biodiversity; synergy; economic diversification.
  128. [128]
    Agroecology: Principles and Practices for Diverse, Resilient, and ...
    Oct 30, 2019 · Agroecology is a science that applies ecological concepts and principles to the design and management of sustainable agricultural ecosystems.
  129. [129]
    The 13 principles of Agroecology
    The 13 principles of Agroecology · INput reduction · Soil Health · animal health and welfare · BIODIVERSITY · Synergy · Economic diversification · SOCIAL VALUES AND ...
  130. [130]
    Overstory #95 - Agroecology: principles and strategies
    Nov 13, 2001 · Agroecology goes beyond the use of alternative practices to develop agroecosystems with minimal dependence on high agrochemical and energy inputs.
  131. [131]
    Urban ecosystem | Human Impact, Biodiversity & Pollution | Britannica
    Oct 11, 2025 · Urban ecosystems are often warmer than other ecosystems that surround them, have less infiltration of rainwater into the local soil, and show ...
  132. [132]
    Urbanisation affects ecosystem functioning more than structure in ...
    We found that urbanisation affected function 32% more than structure in the studied tropical streams, likely driven by greater algal growth in urban streams.
  133. [133]
    Ecosystem Services in Urban Landscapes: Practical Applications ...
    Analyzing how urban ecosystems function, provide goods and services for urban dwellers; and how they change, and what allows and limits their performance can ...
  134. [134]
    [PDF] Forests as Human-Dominated Ecosystems - Dirzo Lab
    Forests are human-dominated ecosystems, often used for agroforestry, hunting, and gathering. Human activity, including clearing, has shrunk the forest by about ...
  135. [135]
    How ecological research on human-dominated ecosystems ...
    Nov 16, 2021 · We conducted an analysis of ecology literature dealing with agricultural and forestry practices (AF practices), scanning a total of 27 556 references.
  136. [136]
    Chapter: Engineering Resilience versus Ecological Resilience
    The two contrasting aspects of stability—essentially one that focuses on maintaining efficiency of function (engineering resilience) and one that focuses on ...
  137. [137]
    Conservation Ecology: Sustainability, Stability, and Resilience
    The purpose of this essay is to define and refine the concepts of stability and resilience and to demonstrate their value in understanding the behavior of ...Missing: debates | Show results with:debates
  138. [138]
    Qualitative Stability in Model Ecosystems - May - 1973 - ESA Journals
    May 1, 1973 · We discuss deductions that can be made as to the stability of multispecies communities, knowing only the structure of the food web.Missing: critique | Show results with:critique<|control11|><|separator|>
  139. [139]
    Resilience and Stability of Ecological Systems - jstor
    These features have all been shown to be essential properties of the predation process (Holling 12, 13) and the effect of adding some of them has been analyzed.
  140. [140]
    Unifying the concepts of stability and resilience in ecology
    Mar 19, 2021 · With rapidly changing environmental conditions, the stability of ecosystems has become a major focus of ecological research.THE BABYLONIAN... · DEFINING ECOLOGICAL... · MEASURING ECOLOGICAL...
  141. [141]
    Ecosystem stability and resilience: a review of their relevance for the ...
    This paper provides a selective review of ecosystem stability from a conservation perspective. Existing models of ecosystem resilience are discussed.
  142. [142]
    Does functional redundancy affect ecological stability and resilience ...
    Jul 8, 2020 · The results of this meta-analysis indicate that functional redundancy may positively affect community stability and resilience to disturbance.Missing: debates | Show results with:debates
  143. [143]
    No complexity–stability relationship in empirical ecosystems - Nature
    Aug 24, 2016 · This prediction was put forth by Robert May, who studied the relationship between complexity and stability in random ecosystems. Ecosystem ...Missing: critique | Show results with:critique
  144. [144]
    Outstanding Challenges in the Transferability of Ecological Models
    Predictions from transferred ecological models are affected by species' traits, sampling biases, biotic interactions, nonstationarity, and the degree of ...
  145. [145]
    [PDF] The Limits to Models in Ecology
    Limited data availability implies that site or subject-specific models may be possible only for a limited number of subjects (ecosystems, communities, or ...
  146. [146]
    [PDF] Full title Ecosystem models cannot predict the consequences ... - arXiv
    Ecosystem models have poor predictive accuracy when forecasting future dynamics and offer ambiguous predictions about how species will respond to management ...
  147. [147]
    [PDF] Fisheries and Ecological Models in Fisheries Management - AWS
    is a key point in ecosystem models and much of the profuse literature ... social models relatively poor predictive power. MODELS AND DECISION-MAKING.
  148. [148]
    Harmonizing nature's timescales in ecosystem models - Cell Press
    Apr 25, 2025 · Existing ecosystem models either collapse or omit timescales, which can compromise biological accuracy, or include all timesteps, which reduces ...
  149. [149]
    [PDF] Recent developments in empirical dynamic modelling
    Many ecological time series are short relative to the time- scale of the system, which presents a major obstacle to the successful ap- plication of EDM: if the ...<|separator|>
  150. [150]
    Empirical ecology to support mechanistic modelling: Different ...
    Jun 18, 2025 · These problems are characterized by large scales, rapid change in climate and other environmental conditions, multiple stressors on multiple ...
  151. [151]
  152. [152]
    Why too stringent ecosystem management can cause serious damage
    Oct 7, 2015 · A new study concludes that managing ecosystems for predictable outcomes is risky and can seriously backfire.
  153. [153]
    [PDF] The Nonequilibrium Paradigm in Ecology and the Partial Unraveling ...
    Apr 1, 1994 · In the twenty-five years since it has been enshrined in environmental law, the equilibrium paradigm has been re- jected in ecology and replaced ...
  154. [154]
    Ecological non-equilibrium and biological conservation
    We argue that the concepts of ecological equilibrium and balance still hinder wildlife managers from optimizing proper decision-making and correctly ...
  155. [155]
    [PDF] KEY ISSUES IN BIODIVERSITY OFFSET LAW AND POLICY
    Key issues include mitigation hierarchy, goals of offsetting, limits to offsetting, establishing equivalence, duration, proximity, use of averted losses, and ...
  156. [156]
    The inherent biases in environmental research and their effects on ...
    Three types of bias: personal, institutional, and socio-cultural, will be examined together with their effects on public policy.
  157. [157]
    The use and abuse of ecosystem service concepts and terms
    The term 'ecosystem services' is widely used inappropriately and imprecisely. The term obscures context and conflates ecological processes with human services.Missing: achievements | Show results with:achievements
  158. [158]
    Subtle biases with important consequences - Dynamic Ecology
    07-Jan-2021 · As a collective, ecologists tend to share a set of values that might create an important source of bias in our science.
  159. [159]
    Remote Sensing of Ecology, Biodiversity and Conservation
    Remote sensing, the science of obtaining information via noncontact recording, has swept the fields of ecology, biodiversity and conservation (EBC).
  160. [160]
    Harnessing Remote Sensing Data for Biological and Ecological ...
    Nov 4, 2020 · Remotely sensed data can complement traditional field data and help researchers directly measure the distribution of plants and animals.
  161. [161]
    Application of GIS and Remote-Sensing Technology in Ecosystem ...
    Geographic information systems (GISs) and remote-sensing technologies have become indispensable tools in the fields of ecosystem services assessment and ...
  162. [162]
    Algorithms going wild – A review of machine learning techniques for ...
    General ML algorithms are helpful in ecological forecasting, ecosystem monitoring, and conservation planning, enabling ecologists to make more informed ...
  163. [163]
    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.
  164. [164]
    A synergistic future for AI and ecology - PMC - NIH
    Sep 11, 2023 · Applying AI tools to ecological domains has vastly expanded our abilities to quantify phenomena that were previously unquantifiable or difficult ...Fig. 1 · Fig. 2 · 2. Ecological Systems...
  165. [165]
    Internet of Things to network smart devices for ecosystem monitoring
    Sep 15, 2019 · The breakthroughs in the ecosystem monitoring IoT include new data loggers and long-distance wireless sensor network technology that supports ...
  166. [166]
    Assessing and Improving IoT Sensor Data Quality in Environmental ...
    Sep 4, 2024 · This article provides a long-term evaluation of IoT sensors data quality in environmental monitoring networks, particularly focusing on peatlands.
  167. [167]
    A low‐cost IoT network to monitor microclimate variables in ...
    Feb 1, 2023 · This low-cost Internet of Things system can capture the microclimate in contrasting environments with accuracy comparable to commercial ...
  168. [168]
    Long-Term Ecological Research on Ecosystem Responses to ...
    Aug 16, 2022 · We describe how a long-term ecological research perspective facilitates insights into an ecosystem's response to climate change.abstract · Climate change at LTER sites · Ecosystem responses to...
  169. [169]
    Evaluating ecosystem responses to rising atmospheric CO2 and ...
    Mar 22, 2004 · Ecosystem models demonstrate the importance of examining the combined effects of the gradually rising concentration of atmospheric CO 2 and the climatic change ...
  170. [170]
    Review Climate change effects on biodiversity, ecosystems ...
    Sep 1, 2020 · We present the most up-to-date assessment of climate change impacts on biodiversity, ecosystems, and ecosystem services in the US and implications for natural ...
  171. [171]
    Nitrogen and CO2 enrichment interact to decrease biodiversity ... - NIH
    Aug 12, 2025 · This study reveals that nitrogen addition and elevated CO₂ diminish the effects of biodiversity on complementarity and selection effects, but do ...
  172. [172]
    Biodiversity loss reduces global terrestrial carbon storage - Nature
    May 22, 2024 · We find that biodiversity declines from climate and land use change could lead to a global loss of between 7.44-103.14 PgC (global sustainability scenario) and ...
  173. [173]
    Interconnecting global threats: climate change, biodiversity loss, and ...
    In this Review, we define and exemplify the causal pathways that link the three global pressures of climate change, biodiversity loss, and infectious disease.
  174. [174]
    Changes in forest ecosystem stability under climate ... - Frontiers
    Jan 9, 2025 · We hypothesize that even with the expected changes in climate, the structural stability of the species assembly will remain unchanged until the ...
  175. [175]
    Ecological acclimation: A framework to integrate fast and slow ...
    Jun 23, 2025 · A framework for assessing ecosystem dynamics in response to chronic resource alterations induced by global change. Ecology, 90, 3279–3289.1 Introduction · 4.2 Coral Reef Acclimation... · 4.3 Soil Microbiome Response...