Fact-checked by Grok 2 weeks ago

Biome

A biome constitutes a large ecological of , animals, and microorganisms adapted to a dominant regime, primarily delineated by patterns of , , and that shape and . Terrestrial biomes, the most commonly referenced, span continental scales and include types such as tropical rainforests with high rainfall and year-round warmth supporting multilayered canopies, deserts marked by aridity and sparse xerophytic , temperate grasslands featuring seasonal droughts and fire-adapted grasses, boreal forests or taiga dominated by conifers in cold climates, and with limiting growth to low shrubs and lichens. biomes, encompassing freshwater systems like and lakes alongside realms from coastal zones to open oceans, similarly reflect gradients in , depth, and nutrient availability influencing . Biome classification schemes, such as Robert H. Whittaker's 1975 framework, map distributions using annual mean against to predict vegetation formations, revealing causal links between abiotic drivers and assemblages without reliance on subjective boundaries. This approach underscores how moisture and thermal regimes determine primary productivity, , and trophic dynamics, with empirical data from global datasets confirming that deviations in these variables correlate with biome shifts observed in paleorecords and contemporary monitoring. While natural biomes reflect long-term climatic equilibria, pressures including , , and have induced transitions, such as woodland conversion to croplands, prompting recognition of human-modified "anthromes" that now cover over half of Earth's ice-free land surface. Such alterations disrupt native adaptations, often reducing to further climate variability as evidenced by accelerated biome boundary migrations in response to warming trends.

Definition and Fundamentals

A biome refers to a major ecological community of organisms adapted to a specific climatic or environmental regime across large geographic scales, typically spanning continents or zones. These units are delineated primarily by the predominant structure and —such as , , or —rather than fine-scale composition, with associated animal communities exhibiting convergent adaptations to the prevailing conditions. Empirical classification emphasizes climatic drivers like annual ranges and patterns, which causally determine plant growth forms and limit faunal distributions, as observed in global patterns where similar biomes recur under analogous abiotic constraints irrespective of historical . Biomes contrast with ecosystems, which denote functional assemblages of biotic interactions, flows, and cycles within circumscribed areas, often at scales from ponds to forests; biomes encompass aggregations of such ecosystems unified by overarching climatic envelopes rather than localized processes. Habitats, by comparison, specify the immediate microenvironments supporting particular species or populations, such as a for an , lacking the macro-scale climatic integration central to biomes. Ecoregions refine biomes further by incorporating , variations, and evolutionary history to map discrete subunits, as in frameworks delineating thousands of global ecoregions nested within broader biome types, enabling finer targeting without altering the climatic core of biome definitions. This hierarchical distinction underscores biomes' utility as coarse-grained constructs for synthesizing planetary ecological patterns, grounded in observational data from surveys and correlations, avoiding conflation with dynamic process-oriented or species-specific concepts.

Primary Determinants: Climate, , and

serves as the principal driver of terrestrial biome distribution, with mean annual and exerting the strongest controls on structure and composition. Empirical analyses of global patterns reveal that biomes align closely with climatic envelopes, where regimes dictate physiological tolerances of dominant plant species, while levels determine water availability critical for and growth. For instance, tropical rainforests occur where annual exceeds 2000 mm and temperatures remain above 20°C year-round, enabling multilayered canopies. further refines these patterns; pronounced dry seasons restrict savannas to regions with 500-1500 mm annual rainfall interspersed with months below 100 mm, favoring fire-adapted grasses over closed forests. Vegetation, in turn, represents the biotic manifestation of climatic constraints, with dominant plant functional types—such as broadleaf evergreens, coniferous needle-leaves, or drought-deciduous shrubs—defining biome identity. These assemblages emerge from evolutionary adaptations to local climate, where species with congruent tolerances cluster into stable communities; for example, boreal forests feature slow-growing conifers resilient to cold winters averaging -30°C and short growing seasons under 100 frost-free days. While vegetation feedbacks, like albedo modification or evapotranspiration, can locally amplify climatic effects, observational data from satellite-derived indices confirm climate as the overriding predictor of vegetation indices across biomes. Discrepancies arise in transitional zones, where edaphic factors or disturbances override pure climatic determinism, underscoring that potential natural vegetation serves as a proxy for underlying climate. Soil properties modulate biome expression by influencing nutrient cycling, water-holding capacity, and rooting depth, though they derive largely from climatic and vegetational influences via pedogenesis. Across biomes, soil fertility gradients—spanning nutrient-rich mollisols in grasslands to leached in —correlate with and organic inputs from overlying , but accelerates weathering rates; arid deserts exhibit calcic, saline soils under low (<250 mm annually), limiting plant establishment to succulents. Empirical studies spanning soil chronosequences demonstrate that while soil age affects structure in specific locales, its biome-scale role remains subordinate to climate, with variations explaining less than 20% of ecosystem differences after controlling for temperature and rainfall. Interactions persist, as vegetation litterfall enriches topsoils, fostering feedbacks that stabilize biome boundaries against minor climatic shifts.

Empirical Validation and Observational Basis

The empirical foundation of biomes stems from systematic field observations correlating vegetation structure, climate parameters, and soil characteristics across vast regions. Expeditions and ecological surveys conducted since the late 19th century, such as those by Russian botanist Vasily Dokuchaev on soil-vegetation zonality in Eurasia, documented repeatable patterns where specific plant communities dominate under comparable environmental conditions, forming the basis for recognizing biomes as cohesive units. These ground-based validations, extended through global inventories like the International Biological Program (1964–1974), quantified biomass and species composition, revealing that terrestrial biomes exhibit distinct productivity levels tied to precipitation and temperature regimes, with forests averaging higher net primary productivity than deserts. Satellite remote sensing has provided scalable empirical validation since the 1970s, enabling global mapping of biome distributions through vegetation indices and land cover classifications. Instruments like and have generated datasets such as the International Geosphere-Biosphere Programme (IGBP) land cover map, which delineates biomes based on observed spectral signatures of dominant vegetation, achieving accuracies exceeding 70% when cross-validated against field plots. Long-term records from these platforms, spanning over four decades, confirm biome stability in undisturbed areas while detecting shifts, such as greening in northern biomes correlating with warming trends, with normalized difference vegetation index (NDVI) increases of up to 0.05 units in most categories from 1990 to 2020. Advanced analytical methods further validate biome concepts by integrating observational data into predictive models. Machine learning approaches, including convolutional neural networks trained on satellite-derived bioclimatic variables, reproduce global biome maps with high fidelity, demonstrating that empirical patterns of vegetation-climate covariance explain over 80% of distributional variance. Comparative studies across classification schemes highlight consistency in core biome delineations when anchored to direct observations, though discrepancies arise in transitional zones, underscoring the need for hybrid ground-satellite approaches to refine boundaries. These validations affirm biomes as observable, causal assemblages rather than arbitrary constructs, grounded in reproducible environmental-vegetation linkages.

Historical Development

Pre-20th Century Observations

Early observations linking climate to vegetation patterns, foundational to later biome concepts, trace to ancient Greece. Parmenides in the 5th century BC delineated global climatic zones—frigid, temperate, and torrid—implying regional differences in habitable flora. Theophrastus, in the 3rd century BC, empirically connected environmental conditions to plant distributions, growth, and diversity in works like Enquiry into Plants, observing how soil, water, and temperature shaped vegetation assemblages across regions. The 18th century advanced descriptive phytogeography amid expanding exploration. Carl Linnaeus, in Philosophia Botanica (1751), classified plant "stations" by habitat types such as maritime, freshwater, prairies, and rocky terrains, integrating these into his "economy of nature" framework that emphasized climatic and edaphic controls on species assemblages and ecological balances. Georges-Louis Leclerc, Comte de Buffon, documented in 1761 the physiognomic convergence of vegetation forms (e.g., tree-dominated woodlands) across similar climates on different continents, despite floristic disparities, as part of his broader biogeographical inquiries. Carl Ludwig Willdenow in 1792 further highlighted climate's role in dictating global vegetation distributions through systematic comparisons of European and extra-European floras. Nineteenth-century naturalists provided quantitative empirical foundations via fieldwork. Alexander von Humboldt's 1805 Essai sur la géographie des plantes, based on Andean traverses, mapped elevational vegetation belts—from tropical rainforests at low altitudes to alpine tundra at peaks—correlating shifts with isothermal lines and humidity gradients measured via thermometers and hygrometers; he termed these "associations" of socially organized plant life. August Heinrich Rudolf Grisebach in 1838 defined "formations" as vegetation units shaped by climate's influence on plant physiognomy, such as leaf size and stature, drawing from global datasets. Augustin Pyramus de Candolle's Géographie botanique raisonnée (1855) quantified plant dispersion patterns worldwide, attributing zonal distributions primarily to temperature extremes and seasonal precipitation. By century's end, synthesis emerged. Andreas Franz Wilhelm Schimper's 1898 Pflanzengeographie auf physiologischer Grundlage integrated prior data to delineate major global vegetation zones—deserts, steppes, savannas, forests—causally tied to water and thermal regimes, emphasizing physiological adaptations and excluding human-modified landscapes as potential natural states. These works collectively established climate as the dominant driver of large-scale vegetation uniformity, observable through repeatable field measurements, predating formal biome nomenclature.

20th Century Formalization and Key Proponents

The term biome was introduced by ecologist in 1916 during his presidential address at the inaugural meeting of the , where he proposed it as a synonym for a large-scale biotic community encompassing both plants and animals. Clements conceptualized biomes as mature climax formations resulting from ecological succession under dominant climatic controls, emphasizing their role as integrated units of vegetation and fauna adapted to regional environmental conditions. This marked a shift from earlier plant-centric formations toward a holistic community approach grounded in observational data from North American prairies and forests. Clements collaborated with animal ecologist Victor E. Shelford to refine the concept, culminating in their 1939 publication Bio-Ecology, which formalized biomes as climatically driven associations of dominant vegetation types and their interdependent animal populations. Shelford, building on Clements' framework, stressed empirical classification through field studies of habitat gradients and succession stages, clarifying in earlier work with E.C. Olson (1935) that biomes represent biotic communities within broad climatic zones rather than isolated plant stands. Their joint efforts provided the first systematic delineations of major North American biomes, such as tundra, grassland, and forest, supported by quantitative surveys of species distributions and abiotic correlations. By mid-century, Robert H. Whittaker advanced formalization through gradient analysis, publishing classifications in the 1950s and 1960s that mapped biomes onto axes of mean annual temperature and precipitation, deriving boundaries from empirical vegetation data across elevational and latitudinal transects. Whittaker's approach critiqued Clements' succession-heavy model by prioritizing direct climatic causation over developmental stages, using statistical correlations from global datasets to identify eight principal terrestrial biomes, including tundra, taiga, and desert. This quantitative refinement, validated against plot-level floristic surveys, facilitated broader application in biogeography while highlighting biome transitions as continuous rather than discrete.

Evolution into Functional and Potential Vegetation Frameworks

The concept of potential natural vegetation (PNV) formalized in the mid-20th century as a predictive tool for vegetation classification, defining the mature, self-perpetuating plant community expected to dominate a site under current environmental conditions without human interference. German phytosociologist introduced the term in 1956, describing it as an "imagined natural state" derived from analysis of relict stands, succession patterns, and habitat factors like climate and soil. This approach built on but emphasized empirical extrapolation over rigid determinism, enabling biome mappings to reflect equilibrium states shaped by abiotic controls rather than transient or anthropogenic landscapes. PNV's utility lay in its causal framing: vegetation as the outcome of site-specific potentials, with applications in European conservation planning and North American inventories, such as A.W. 's 1969 maps of U.S. potential vegetation integrating 1,000+ units based on climate zones and soil moisture indices. By the late 20th century, PNV frameworks intersected with biome evolution by providing a baseline for zonal vegetation types, distinguishing potential from actual distributions influenced by fire, grazing, or agriculture. In practice, PNV classifications used phytosociological alliances—groups of associations with shared dominants—to delineate biomes, as seen in extensive European surveys covering over 50% of territory by the 1980s. However, limitations emerged: the assumption of a singular climax ignored paleoecological evidence of multiple stable states and underestimated disturbance as a co-driver, prompting refinements like seral-stage incorporations in dynamic models. Despite critiques of oversimplification, PNV persists in global datasets, underpinning tools like the FAO's potential vegetation layers for assessing land degradation, with validations against pollen records showing 70-80% congruence in temperate zones. Functional vegetation frameworks advanced this evolution from the 1980s onward, redefining biomes through plant functional types (PFTs)—trait-based clusters capturing physiological responses to climate, such as evergreen vs. deciduous habits or C3 vs. C4 photosynthesis. Rooted in empirical trait measurements, these replaced descriptive floristics with mechanistic rules: biomes as emergent from PFT competition under environmental filters, as modeled in Prentice et al.'s 1992 BIOME scheme simulating global distributions via 13 PFTs and bioclimatic thresholds calibrated to 1,000+ fossil pollen sites. This shift enabled causal realism in predictions, linking traits like leaf area index (averaging 2-5 m²/m² in forests) to ecosystem fluxes, with validations showing 85% accuracy in reproducing observed biome extents. By the 21st century, integration of PFTs with PNV yielded hybrid frameworks, as in dynamic global vegetation models (e.g., , ) that simulate potential distributions under transient climates, incorporating trait variability from databases like TRY (encompassing 200,000+ records since 2007). Recent typologies, such as Moncrieff et al.'s 2022 function-based system, hierarchically classify 23 functional biomes using 18 bioclimatic indices and PFT dominance, tested against satellite-derived land cover with 75% overlap, emphasizing traits' role in bounding biome transitions amid warming. These developments prioritize empirical trait-environment correlations over static maps, revealing, for instance, that functional convergence (e.g., drought-tolerant traits in semi-arid zones) explains 60% of biome productivity variance, enhancing resilience assessments without assuming unbiased source neutrality in model assumptions.

Classification Systems

Early Climatic Schemes (Holdridge, Whittaker)

The , introduced by ecologist in 1947, represents an early quantitative approach to classifying terrestrial through climatic determinants. It integrates three primary variables: biotemperature (the annual summation of daily mean temperatures above 0°C, excluding frost-influenced periods), total annual precipitation, and the ratio of potential evapotranspiration (PET) to precipitation, which accounts for atmospheric moisture demand relative to supply. These factors are plotted on a triangular diagram, enabling the delineation of 37 distinct life zones, from ice caps and polar deserts to wet tropical forests, based on empirical correlations between climate gradients and vegetation physiognomy. The system's emphasis on biotemperature prioritizes effective growing season warmth over absolute minima, reflecting causal influences on photosynthetic activity and plant distribution limits. updated the framework in 1967, incorporating altitudinal and latitudinal applications for global mapping. Robert H. Whittaker advanced climatic biome schemes in 1962, proposing a continuum-based classification mapped against mean annual temperature and mean annual precipitation on a two-dimensional graph. This model identifies major biome types—including tundra, boreal forest, temperate deciduous forest, grassland, desert, savanna, and tropical rainforest—as overlapping zones along climatic gradients, underscoring vegetation structure as a direct response to thermal and hydrological regimes rather than discrete boundaries. Whittaker refined the system through works in 1970 and 1975, integrating ordination techniques from community ecology to validate empirical patterns observed across continents. Unlike Holdridge's ternary inclusion of evapotranspiration, Whittaker's binary axes simplify prediction but may underrepresent aridity effects in high-evaporation environments, as precipitation alone inadequately proxies soil moisture balance. Both schemes prioritize abiotic climatic drivers as proximal causes of biome differentiation, grounded in mid-20th-century field observations linking vegetation dominance to temperature-precipitation interactions. Holdridge's approach, with its PET ratio, better accommodates evaporative stress in predicting transitions to xerophytic formations, while Whittaker's facilitates broader physiognomic generalizations applicable to global syntheses. Empirical validations, such as correlations with remote sensing data, affirm their utility despite limitations in capturing edaphic or disturbance feedbacks. These early models laid foundational causal frameworks for subsequent classifications, emphasizing verifiable climatic thresholds over subjective descriptors.

Zonal and Ecoregional Approaches (Walter, Bailey, Olson-Dinerstein)

Heinrich Walter's zonal classification system delineates the Earth's vegetation into nine zonobiomes, broad latitudinal belts primarily determined by climatic gradients of temperature effectiveness and moisture availability, as plotted in his climatic diagrams that emphasize seasonal water balance over mere annual totals. These zonobiomes integrate zonal soils and dominant vegetation forms, such as evergreen tropical rainforests in zonobiome I (equatorial, with minimal seasonality and high precipitation exceeding evapotranspiration) transitioning to savannas in zonobiome II (tropical with pronounced dry seasons), sclerophyllous woodlands in zonobiome III (subtropical arid), and culminating in polar deserts in zonobiome IX (cold, with short growing seasons and permafrost). Walter's framework, outlined in his 1968 book Vegetation of the Earth and refined in subsequent editions, prioritizes empirical field observations of vegetation-climate correspondence across continents, rejecting overly rigid biome boundaries in favor of ecotones as transition zones influenced by local topography and edaphic factors. This approach underscores causal links between macroclimate and potential natural vegetation, validated through global transects showing convergent physiognomies in similar climatic zones despite floristic differences. Robert G. Bailey extended zonal principles into a hierarchical ecoregional system, classifying ecosystems from continental domains (e.g., polar, humid temperate) down to provinces based on integrating climate regimes, land surface form, and potential vegetation, with boundaries drawn to reflect ecological continuity and potential natural communities. Developed initially for the United States in the 1970s under the U.S. Forest Service and expanded globally by 1996, Bailey's framework identifies 32 domains worldwide, subdivided into 100+ divisions and provinces, such as the Arctic Tundra Domain (M130) encompassing provinces like the Brooks Range (M131) characterized by continuous permafrost and graminoid tundra. Empirical delineation relied on climate station data, soil surveys, and physiographic maps to define ecoregions as areas of relative homogeneity in ecosystem potential, facilitating resource management by accounting for both zonal climate drivers and regional geomorphic influences that modify vegetation patterns. Bailey's system has been adopted for national forest planning and global reporting, though it emphasizes coarser macroscale units over fine-scale biodiversity hotspots. The Olson-Dinerstein ecoregional approach, published by the World Wildlife Fund in 2001, refines global terrestrial classification into 825 ecoregions—large units of land (typically 50,000–1,000,000 km²) defined by distinct assemblages of species, ecological dynamics, and evolutionary histories, rather than strict climatic zonation alone. Led by and , this framework incorporates biotic criteria like endemism and beta diversity alongside abiotic factors, grouping ecoregions into 14 (e.g., tropical and subtropical moist broadleaf forests) and further into freshwater and marine parallels, with boundaries derived from expert workshops, satellite imagery, and species distribution data to prioritize conservation viability. For instance, the ecoregions highlight unique faunal convergences under similar climates, diverging from purely zonal models by accommodating topographic heterogeneity and historical biogeography. This system complements priority-setting tools like the ecoregions, focusing on irreplaceable biodiversity rather than uniform climate-vegetation correlations, and has informed WWF's conservation strategies across 35 priority areas despite critiques of subjective boundary judgments. These approaches collectively advance beyond early climatic schemes by embedding zonal climate realism within regional ecological contexts: Walter's zonobiomes provide a foundational climatic scaffold, Bailey's hierarchy adds physiographic scaling for management applicability, and Olson-Dinerstein's ecoregions emphasize biotic integrity for global conservation, though all rely on verifiable climatic and distributional data while acknowledging limitations in capturing microscale or anthropogenic variations.

Recent Updates and Global Standards (Post-2000 Developments)

Post-2000 biome classifications have increasingly incorporated human modifications, departing from purely natural potential vegetation models toward frameworks that map observed, anthropogenic-influenced landscapes. In 2008, Erle C. Ellis and Navin Ramankutty introduced anthropogenic biomes, or "anthromes," categorizing the terrestrial biosphere into 18 classes based on land use intensity and vegetation cover, revealing that by 2000, approximately 55% of global ice-free land was used for agriculture, settlements, or other human activities. This approach, expanded in 2010 to a historical series from 1700 to 2000, demonstrated a transition from mostly wild to predominantly anthropogenic biomes, with over half of the terrestrial surface transformed by the early 20th century and further intensified by 2000 through croplands, pastures, and villages. Advancements in remote sensing have enabled higher-resolution, dynamic biome mapping post-2000. The International Geosphere-Biosphere Programme (IGBP) land cover classification, derived from MODIS satellite data around 2001-2005, delineates 17 vegetation classes globally at 1 km resolution, providing empirical baselines for biomes influenced by both climate and land management. Annual 30-m resolution maps of global grasslands from 2000 to 2022, produced using Landsat and Sentinel-2 imagery, quantify extent changes at 1.2% annual variability, highlighting empirical shifts driven by conversion and restoration efforts. New classification schemes emphasize bioclimatic and functional criteria for global standardization. A 2021 proposal introduced a hierarchical system using six bioclimatic variables—such as temperature seasonality and aridity—to define biomes from macro- to micro-scales, aiming for consistency in ecological modeling and conservation. Similarly, the 2023 Taskforce on Nature-related Financial Disclosures (TNFD) guidance maps sectors to biomes across land, freshwater, and marine realms, using indicators like intactness to assess human impacts, though reliant on datasets like anthromes for validation. Studies from 2023 underscore that choice of classification—e.g., dynamic global vegetation models versus static schemes—affects projected biome shifts under climate scenarios, with discrepancies up to 20% in future distributions. In 2025, the U.S. National Standard for Ecosystem Classification updated its framework to align with the International Classification of Ecological Communities, incorporating finer ecosystem types and global interoperability for monitoring anthropogenic pressures across biomes. These developments reflect a consensus toward hybrid models integrating satellite observations, land-use data, and climatic drivers, prioritizing empirical distributions over idealized potentials to better inform policy on biome degradation and restoration.

Major Biome Categories

Terrestrial Biomes: Structure and Examples

Terrestrial biomes represent expansive land-based ecological communities defined by prevailing vegetation types, which emerge from interactions between climate variables like annual temperature range and precipitation patterns. Their internal structure typically includes stratified plant layers in wooded areas—such as emergent trees, canopies, understories, and ground covers—that create microhabitats for fauna with specialized foraging and sheltering behaviors, while non-forested biomes feature uniform herbaceous or shrub layers supporting grazing and burrowing adaptations. These structures reflect causal linkages from abiotic drivers to biotic assemblages, with empirical distributions mapped globally based on satellite-derived vegetation indices and field validations. The eight principal terrestrial biomes, as delineated in ecological surveys, illustrate this variability:
  • Tropical Rainforest: Occurs in equatorial zones with mean annual temperatures exceeding 20°C and precipitation over 2000 mm, featuring four vertical strata including a dense canopy up to 30 m high; dominant plants include broadleaf evergreens like ; animals exhibit arboreal locomotion and frugivory, such as in Southeast Asia.
  • Savanna: Transitional zones between forests and deserts with distinct wet-dry seasons (500-2000 mm precipitation), structured by scattered trees amid grasslands; and grasses prevail, with herbivores like displaying migratory patterns to track rainfall.
  • Subtropical Desert: Hyper-arid regions with less than 250 mm annual rain and diurnal temperature swings; vegetation sparse with succulents and shrubs adapted via CAM photosynthesis; fauna includes nocturnal rodents and reptiles with water-conserving physiologies, exemplified by species.
  • Chaparral: Mediterranean climates with mild, wet winters and hot, dry summers (300-900 mm precip); sclerophyllous shrubs and small trees form dense thickets; animals like have fire-resistant traits and browse adaptations.
  • Temperate Grassland: Continental interiors with 250-750 mm precip and cold winters; dominated by perennial grasses with deep root systems; large ungulates such as exhibit herd dynamics for predator evasion.
  • Temperate Forest: Moderate precipitation (750-1500 mm) and seasonal temperature shifts; deciduous or mixed trees create layered canopies with leaf litter floors; and adaptations include hibernation and mast caching.
  • Boreal Forest (Taiga): Subarctic conditions with long, cold winters and 300-850 mm precip mostly as snow; coniferous evergreens like form even canopies; and show cold tolerance via insulation and pack hunting.
  • Tundra: Polar extremes with permafrost, temperatures below 0°C for much of the year, and under 250 mm precip; low shrubs, sedges, and lichens hug the ground; and migrate seasonally with insulating fur.
These biomes cover varying proportions of global land area, with forests (tropical and temperate/boreal combined) comprising roughly 30-40% based on remote sensing data from 2000-2020, though distributions shift due to climatic gradients and edaphic factors. Empirical studies confirm that biome boundaries align closely with isotherm and isohyet contours, underscoring climate's primacy in structuring community assembly.

Aquatic and Marine Biomes: Characteristics and Transitions

Aquatic biomes encompass freshwater systems such as rivers, lakes, and wetlands, characterized by low salinity levels typically below 0.5 parts per thousand (ppt), distinguishing them from saline environments. These systems exhibit variability in flow regimes—lotic in rivers with unidirectional currents promoting oxygen exchange and sediment transport, and lentic in lakes with stratified layers influenced by thermal gradients, where surface waters warm under solar radiation while deeper zones remain cooler. Key physical factors include temperature fluctuations driven by seasonal climate, wind-induced mixing, and precipitation inputs, which affect dissolved oxygen concentrations critical for aerobic organisms; for instance, oxygen solubility decreases with rising temperatures, often leading to hypoxic conditions in warmer strata. Chemical attributes feature low dissolved solids under 1,000 milligrams per liter, with nutrient dynamics from terrestrial runoff shaping productivity, as seen in eutrophic lakes where phosphorus and nitrogen excesses foster algal blooms. Biologically, these biomes support diverse macroinvertebrates, fish, and submergent vegetation that provide habitat structure, such as submerged aquatic vegetation offering refuge from predators in shallow zones. Marine biomes, dominated by ocean waters, maintain average salinity around 35 ppt, enabling distinct osmotic adaptations in resident organisms like marine mammals and plankton. Zonation by depth structures these ecosystems: the epipelagic zone (0-200 meters) receives sunlight, supporting photosynthesis in phytoplankton and kelp forests in cooler coastal shallows; the mesopelagic (200-1,000 meters) features dim light and diurnal migrations; and the bathypelagic (below 1,000 meters) sustains constant near-freezing temperatures around 4°C with minimal oxygen variability. Pelagic realms emphasize open-water dynamics influenced by global currents, such as the thermohaline circulation distributing heat and nutrients, while benthic zones on the seafloor—averaging 3.7 kilometers deep—host chemosynthetic communities near vents amid complex topography including canyons and seamounts. Intertidal margins, exposed to tidal cycles, exhibit steep gradients in desiccation stress and wave energy, fostering resilient species like barnacles adapted to alternating submersion. Overall, marine productivity hinges on upwelling zones where nutrient-rich deep waters surface, contrasting with oligotrophic open oceans limited by light penetration beyond 200 meters. Transitions between aquatic and marine biomes occur primarily in estuarine ecotones, where freshwater river inflows mix with seawater, creating salinity gradients from near-zero ppt upstream to full marine levels seaward. These zones, influenced by tidal flushing and river discharge, support brackish conditions that filter species tolerant of osmotic fluctuations, such as diadromous fish migrating for spawning. Coastal wetlands like salt marshes and mangroves exemplify these interfaces: salt marshes feature elevational gradients from low-lying halophytic grasses (e.g., Spartina spp.) in frequently inundated areas to higher, less saline herbaceous zones, with salinity stressing plant zonation and carbon sequestration. Mangroves, thriving in intertidal tropics, form dense stands along salinity fronts where pneumatophores aid aeration in anoxic soils, but elevated salinities above 40 ppt stunt growth and favor dwarf forms, reducing biomass and biodiversity. Such ecotones enhance resilience through edge effects, amplifying species richness via hybrid habitats, though they remain vulnerable to hydrological alterations amplifying salinity extremes.

Anthropogenic and Microbial Biomes: Human-Altered and Microscale Variants

Anthropogenic biomes, termed , delineate global ecological patterns arising from direct and sustained human modifications to ecosystems, integrating land-use intensity, population density, and vegetation cover as key classifiers. This framework, developed by and colleagues, contrasts with traditional biomes by emphasizing human agency over purely climatic drivers, revealing that by 2000, anthromes occupied approximately 75% of Earth's ice-free terrestrial surface, up from about 50% in 1700. Historical mapping traces anthrome emergence to the Neolithic, with initial appearing around 7000 BCE in regions like the , followed by village and rangeland expansions by 6000 BCE and dense settlements by 8000 BCE. Anthromes are stratified into six primary groups—dense settlements, villages, croplands, rangelands, seminatural, and wildlands—each exhibiting distinct biophysical traits shaped by human practices. For instance, urban anthromes, such as dense settlements covering less than 1% of land but housing over 50% of the global population by 2000, feature high impervious surfaces, fragmented habitats, and biodiversity reliant on introduced or tolerant species amid elevated nutrient and pollutant loads. Agricultural anthromes, including intensive croplands (e.g., irrigated rice paddies spanning 20 million hectares in Asia by the late 20th century) and pastures, dominate modified landscapes, supporting 90% of human food production through monocultures, fertilizers, and irrigation that alter soil structure, hydrology, and carbon fluxes compared to native vegetation. These variants often exhibit reduced native species diversity but enhanced productivity for human needs, with seminatural anthromes retaining patches of secondary forests or grasslands under selective management. Microbial biomes represent microscale ecological variants, defined as assemblages of microorganisms—including bacteria, archaea, fungi, protists, and viruses—interacting within discrete habitats like soil aggregates, aqueous films, or host-associated niches, functioning analogously to macro-biomes through community structure, trophic dynamics, and environmental feedbacks. Unlike macroscopic biomes, these operate at scales from micrometers to centimeters, with densities reaching 10^9 to 10^11 cells per gram in soils, where phyla such as and dominate nutrient cycling and decomposition processes essential to larger ecosystems. Examples include rhizosphere microbiomes surrounding plant roots, which enhance nutrient uptake via symbiotic nitrogen fixation (e.g., in legume nodules processing up to 200 kg of nitrogen per hectare annually) and suppress pathogens, or oceanic microbial biomes in the photic zone, comprising picoplankton like that account for 50% of global primary production through photosynthesis. In extreme environments, such as acidic mine drainage or hydrothermal vents, microbial mats form stratified communities driving chemosynthesis, with species like oxidizing iron at pH below 2, illustrating resilience and biogeochemical roles independent of macroscale climate. These microscale systems underpin macro-biome stability, mediating processes like organic matter breakdown and greenhouse gas emissions, yet remain underexplored relative to their pervasive influence.

Influencing Factors and Dynamics

Abiotic Drivers: Empirical Climate and Edaphic Controls

Mean annual temperature (MAT) and mean annual precipitation (MAP) serve as primary empirical climatic drivers structuring terrestrial biomes, with global observations revealing distinct clustering of vegetation types in the MAT-MAP parameter space. For instance, tropical rainforests predominate where MAT exceeds 20°C and MAP surpasses 2000 mm annually, enabling year-round photosynthesis and high biomass accumulation, as evidenced by correlations in long-term meteorological and vegetation surveys. Temperate deciduous forests align with MAT of 5–15°C and MAP of 750–1500 mm, where seasonal temperature fluctuations induce leaf abscission, while boreal forests occupy cooler regimes with MAT below 5°C and MAP around 400–1000 mm, limiting growth to short summers. Deserts emerge under MAP below 250 mm regardless of MAT, constraining plant establishment due to water scarcity, as quantified in biome-climate overlays from datasets spanning decades. These patterns hold across continents, with biotemperature—a metric summing positive daily temperatures divided by days—refining thresholds by emphasizing effective growing season warmth over raw MAT, as originally derived from empirical altitudinal transects in diverse regions. Edaphic controls, encompassing soil physicochemical properties, modulate biome expression particularly at local and mesoscales where climate gradients are shallow. Soil texture influences water retention and aeration; sandy soils with low clay content (<20%) promote grasslands over forests in semi-arid zones by facilitating rapid drainage and drought stress, as observed in comparative studies of vegetation on varied parent materials. Nutrient availability, governed by cation exchange capacity and organic matter, dictates productivity; oligotrophic soils with low nitrogen (<0.1% total N) and phosphorus support sclerophyllous shrublands rather than dense woodlands, even under adequate precipitation, per field measurements in Mediterranean and tropical settings. Soil pH critically affects micronutrient solubility—optimal ranges of 6.0–7.5 maximize uptake, while pH below 5.0 induces aluminum toxicity, stunting roots and favoring herbaceous over woody dominants, as demonstrated in controlled and observational data from acidic podzols. Interactions between climatic and edaphic factors amplify or buffer biome boundaries, with soils integrating long-term climatic legacies via weathering and leaching. In high-precipitation areas (>1500 mm MAP), intensely leached exhibit low fertility (base saturation <35%), sustaining savannas amid potential for forests, as parent rock and historical drainage patterns override contemporaneous climate. Empirical models incorporating both reveal that edaphic variance explains up to 20–30% of residual biome heterogeneity after climatic predictors, underscoring causal roles in vegetation feedbacks like organic matter accumulation enhancing water-holding capacity. Parent material and topography further mediate these dynamics, with shallow or rocky soils imposing drought-like constraints in mesic climates, evidenced by elevational studies where soil depth correlates inversely with biome transitions.

Biotic Interactions and Feedbacks

Biotic interactions in biomes include interspecific competition, predation, herbivory, and mutualisms, which collectively modulate species distributions, community assembly, and ecosystem resilience beyond abiotic controls. Competition for resources such as light, water, and nutrients restricts species coexistence, often leading to niche partitioning that defines biome-specific assemblages; for example, in arid ecosystems, microbial communities are predominantly assembled through biotic interactions rather than environmental filtering alone. Predation and herbivory exert top-down control, with trophic cascades altering vegetation structure, as observed in grasslands where large herbivore populations prevent woody encroachment and maintain open biomes. Mutualistic relationships, including mycorrhizal fungi aiding plant nutrient acquisition and pollinator-plant symbioses, enhance productivity and diversity, particularly in tropical biomes where such interactions sustain high biomass. These interactions generate feedbacks that stabilize or destabilize biome properties over time. Plant-soil biotic feedbacks, where root exudates influence microbial decomposers, regulate nutrient cycling and carbon storage; in boreal forests, fungal mutualists accelerate decomposition under warming, potentially releasing stored carbon and amplifying climate effects. Trophic feedbacks, such as predator-prey dynamics, maintain alternative stable states; for instance, keystone predators in savannas suppress herbivores, preserving grass-dominated structures against shrub invasion. In marine biomes, plankton-herbivore interactions feedback on primary production, influencing oxygen levels and fishery yields through cascading effects on nutrient upwelling. Aboveground-belowground biotic linkages further propagate feedbacks, as herbivory alters root exudation and soil microbial activity, affecting overall ecosystem resistance to disturbances like drought. The interplay of competition and mutualism shapes long-term biome dynamics, with models showing that mutualistic networks can buffer competitive exclusion, promoting coexistence in diverse systems like coral reefs, though intense competition within mutualist guilds limits partner specificity and richness. Biotic feedbacks contribute to climatic regulation; terrestrial vegetation influences albedo and evapotranspiration, creating self-reinforcing loops that either mitigate or exacerbate shifts, as seen in Amazonian dieback scenarios where reduced tree cover diminishes rainfall recycling. Empirical studies across biomes indicate that ignoring biotic feedbacks underestimates ecosystem vulnerability, with trophic restructuring observed in response to apex predator declines altering energy flows and habitat suitability.

Natural Variability and Transitions Over Geological Time

The fossil record documents the emergence of terrestrial biomes during the Silurian and Devonian periods, approximately 430 to 360 million years ago, when vascular plants evolved from aquatic algae, enabling root systems, upright growth, and the formation of initial forest ecosystems dominated by lycophytes, ferns, and early seed plants. High atmospheric CO2 levels exceeding 2000 ppm facilitated this colonization by reducing water loss in early land plants and supporting photosynthesis under limited nutrient availability. These early biomes were patchy and confined to humid equatorial regions, with evidence from fossil spores and megafossils indicating limited global coverage until forest expansion in the Late Devonian around 380 million years ago. In the Carboniferous and early Permian periods (359-270 million years ago), tropical swamp forests of giant lycopsids and horsetails formed extensive biomes across the supercontinent , sequestering massive carbon amounts that lowered atmospheric CO2 and initiated global cooling, as preserved in widespread coal measures and paleosols. The late Permian transition, around 260 million years ago, marked a shift to drier gymnosperm-dominated woodlands, driven by tectonic uplift and aridity, with glossopterid floras in reflecting adaptation to seasonal climates. The end-Permian mass extinction at 252 million years ago eradicated up to 95% of terrestrial plant species, resetting biomes through volcanic-induced warming and anoxia, as inferred from discontinuous fossil sequences and isotopic excursions. Mesozoic biomes (252-66 million years ago) featured conifer and cycad forests in humid interiors and fern prairies in disturbed areas, with continental fragmentation post-Pangaea promoting biome diversification via new coastal habitats and rainfall patterns. The radiation of angiosperms starting around 130 million years ago transformed vegetation structure, introducing broad-leaved trees and understory herbs that enhanced productivity and insect pollination networks, evidenced by diverse leaf fossils and pollen records indicating up to 70% modern family origins by the Late . The end- extinction at 66 million years ago, linked to asteroid impact and volcanism, selectively pruned gymnosperm biomes while sparing many angiosperms, paving the way for Paleogene recovery. Cenozoic transitions (66 million years ago to present) involved progressive cooling from Eocene greenhouse conditions, with global temperatures dropping 10-15°C by the Oligocene, fostering temperate deciduous forests and tundra expansions. Miocene biome turnover around 20-10 million years ago, amid falling CO2 below 400 ppm and tectonic-driven aridity, drove the global spread of C4 grasslands, replacing woodlands in interiors and altering herbivore evolution, as shown by phytolith and stable carbon isotope data from loess and marine sediments. Quaternary glacial-interglacial cycles over the past 2.6 million years induced latitudinal biome shifts, with pollen cores revealing tundra advances southward by 1000-2000 km during Last Glacial Maximum peaks around 20,000 years ago, demonstrating inherent dynamism responsive to Milankovitch orbital variations. These geological patterns underscore biomes as dynamic equilibria shaped by abiotic forcings, with fossil evidence highlighting resilience through repeated restructurings rather than static persistence.

Human Interactions and Modifications

Land Use Changes and Agricultural Transformations

Human land use changes, particularly agricultural expansion, have profoundly altered terrestrial biomes, converting vast areas of natural ecosystems into anthropogenic biomes dominated by croplands, pastures, and settlements. Between 1700 and 2000, the global terrestrial biosphere shifted from predominantly wild to mostly anthropogenic, surpassing 50% anthropogenic cover in the early 20th century, with agricultural and settled lands comprising the majority by 2000, seminatural areas reduced to under 20%, and wildlands to about 25%. This transformation involved the replacement of diverse biomes—such as tropical forests, temperate grasslands, and boreal woodlands—with intensive land uses, driven by population growth and demand for food and fiber. Agricultural expansion has been the primary driver of deforestation, accounting for nearly 90% of global deforestation, especially in tropical regions where primary forests have been cleared for commodity crops like soy and oil palm, as well as cattle pastures. From 2001 to 2023, global cropland area increased by approximately 80 million hectares (5%), while permanent meadows and pastures declined by 150 million hectares, reflecting shifts toward more intensive cropping in some and abandonment in others. In tropical , for instance, soy cultivation alone occupied 8.2 million hectares of deforested land between 2001 and 2015, predominantly in South America. These changes have marginalized forest-dependent , with active deforestation frontiers characterized by high initial forest cover (>66%) and annual loss rates exceeding 2.5%. In temperate and grassland biomes, agricultural transformations have involved the conversion of native prairies and steppes into croplands, supported by and fertilizers since the 19th century, leading to homogenized landscapes with reduced biome heterogeneity. Arid and semi-arid biomes, such as deserts, have seen land cover shifts through irrigation-enabled , converting barren into cultivated areas, though at the cost of water resource depletion. Globally, over half of habitable land is now dedicated to , with more than three-quarters allocated to production despite its smaller contribution to human caloric intake. These modifications have generated novel biomes, including dense croplands and rangelands, which now cover more than three-quarters of the terrestrial surface when accounting for varying intensities of human use. Biodiversity losses from these land use changes are overwhelmingly attributable to , with cultivation responsible for 72% and pastures for 21% of global impacts, underscoring the causal link between biome conversion and ecological simplification. Historical reconstructions indicate that between AD 800 and 1700, about 5 million square kilometers of natural vegetation were transformed into , predominantly croplands in and . Despite recent slowdowns in —net forest loss averaged 4.7 million hectares annually from 2010 to 2020— pressures persist, with only half to two-thirds of cleared directly entering productive use, the remainder often degraded or abandoned.

Conservation Efforts and Restoration Outcomes

Global protected areas cover approximately 17% of terrestrial land and 8% of marine environments as of 2023, with terrestrial coverage disproportionately low in high-biodiversity tropical biomes compared to temperate zones, despite international commitments under the Kunming-Montreal Global Biodiversity Framework aiming for 30% protection by 2030. Efforts include expanding national parks and other effective area-based conservation measures (OECMs), such as biosphere reserves in forests and grasslands, and marine protected areas (MPAs) in coral and pelagic zones, often coordinated by organizations like the IUCN. However, effectiveness remains mixed, with meta-analyses indicating protected areas reduce habitat loss in some cases but fail to halt degradation in others due to factors like inadequate enforcement, surrounding land-use pressures, and internal threats such as poaching or invasive species. Restoration initiatives target degraded biomes through methods like in tropical dry forests and savannas, natural regeneration in grasslands, and active interventions such as transplantation in biomes. A global of 221 restoration projects found boosts by 15-84% and vegetation structure by 36-77% relative to degraded controls, with natural regeneration outperforming active planting by 34-56% for recovery in some ecosystems. In systems, projects report 60-70% survival rates for transplanted branching species, though outcomes vary widely due to environmental stressors like warming waters, while restorations enhance coastal but show high variability in animal . Empirical outcomes demonstrate conservation actions avert extinctions and restore ecosystem functions, with a 2024 analysis of over 180 studies confirming effectiveness across biomes from species reintroductions in tundra-like systems to habitat reconnection in deserts. Terrestrial restorations increase average biodiversity by 20% while reducing variability across sites, though long-term success depends on addressing causal drivers like fragmentation rather than symptomatic fixes. Rewilding efforts yield positive resilience outcomes in nearly 70% of cases, including trophic cascade recoveries in temperate biomes, but failures occur where human pressures persist, underscoring that protection alone insufficiently counters biome-scale degradation without integrated management.

Economic and Adaptive Benefits of Anthropogenic Biomes

biomes, encompassing croplands, pastures, managed forests, and settlements, form the foundation of global agricultural and sectors, generating substantial economic value through , , and timber production. In 2022, the value added by , , and reached $3.8 trillion globally, reflecting an 89% real-term increase over the preceding two decades driven by intensification within these biomes. These systems support approximately 27% of the in agrifood activities, with croplands and pastures—covering over half of habitable —enabling efficient resource extraction and trade that bolsters national GDPs, particularly in developing regions where exceeds 20% of GDP. Dense settlement biomes, including urban areas, concentrate economic activity, yielding that elevate average incomes and foster innovation in secondary industries. Beyond direct outputs, these biomes enhance by amplifying human and buffering against environmental variability. The expansion of biomes has sustained a global rise from about 1 billion in to over 8 billion today, primarily via agricultural intensification that multiplies caloric yields per unit land compared to pre-human ecosystems. Managed landscapes permit targeted interventions, such as and crop breeding, which stabilize yields amid droughts or pests; for instance, varieties in cropland biomes have reduced yield variability by up to 50% in rain-fed systems relative to unimproved natural analogs. In village and biomes, integrated practices like mitigate degradation and enhance retention, conferring to climatic fluctuations that natural biomes often lack due to their dependence on undisturbed processes. This engineered adaptability underpins for billions, allowing human societies to thrive in diverse and marginal environments where pristine biomes would constrain population densities.

Climate Change and Biome Shifts

Observed Historical Shifts and Natural Cycles

Paleoecological records, including pollen stratigraphy from lake sediments and dendrochronological data from tree rings, provide of biome shifts driven by natural variability over millennia. Pollen assemblages reveal transitions in dominant , such as expansions of forests during cooler intervals and retreats of biomes toward poles in warmer phases, reflecting and changes without anthropogenic influence. Tree-ring widths, narrower during droughts or cold snaps, document annual to centennial fluctuations in forest productivity and species composition within temperate and montane biomes. Long-term natural cycles, primarily Milankovitch orbital forcings—encompassing (100,000-year cycle), obliquity (41,000-year cycle), and (23,000-year cycle)—have induced profound biome redistributions by altering seasonal insolation patterns, leading to glacial-interglacial alternations. During interglacials like the current , these cycles facilitated equatorward contraction of polar biomes and poleward migration of temperate grasslands and woodlands, as evidenced by fossil pollen indicating widespread in regions previously dominated by or . Shorter-term solar variability, such as the (1645–1715), and volcanic eruptions contributing to transient cooling episodes, superimposed on these, caused localized biome contractions, including lowered treelines and shifts to shrub-dominated landscapes in and zones. In the Medieval Warm Period (circa 900–1300 CE), proxy data from European and North Atlantic sites show biome responses including enhanced growth of temperate deciduous forests and expanded arable lands, enabling Norse agriculture in southern Greenland where woody shrubs replaced prior grasslands. Conversely, the Little Ice Age (circa 1300–1850 CE) is recorded in pollen profiles from eastern North America with increases in cold-tolerant conifers like Picea (spruce) and Tsuga (hemlock), signaling cooler, moister conditions that advanced boreal biomes southward and reduced deciduous forest extents. These shifts, corroborated by tree-ring chronologies spanning over 4,500 years in regions like the Yamal Peninsula, underscore biome sensitivity to natural forcings, with recovery phases post-Little Ice Age exhibiting initial warming trends attributable to orbital and solar rebounds rather than industrial emissions until the mid-20th century.

Projected Vulnerabilities: Empirical Data vs. Models

Climate models, such as those from the (CMIP6), project substantial vulnerabilities for terrestrial biomes under elevated concentrations, including poleward shifts in biome boundaries, contraction of tropical rainforests by up to 20-40% in some scenarios, and increased risk of tipping points like widespread due to and stress. These projections often emphasize disequilibrium dynamics, where vegetation lags behind rapidly changing envelopes, leading to predicted losses in and hotspots. However, empirical data from satellite remote sensing and ground observations frequently reveal slower or attenuated responses, highlighting model underestimation of adaptive mechanisms such as CO2 fertilization and species acclimation. Satellite-derived metrics, including (NDVI) and (LAI) from datasets spanning 1982-2020, indicate global greening trends with a 5-10% increase in vegetative cover, particularly in and temperate biomes, contradicting model forecasts of net productivity declines in warming scenarios. This greening, observed across 55% of global vegetated land, is attributed to enhanced from rising atmospheric CO2 levels, with flux tower measurements confirming a 20-30% boost in gross primary productivity per 100 ppm CO2 increase in non-water-limited ecosystems. Models incorporating dynamic global vegetation components often simulate weaker or saturating CO2 effects due to nutrient constraints, yet empirical syntheses show persistent fertilization benefits, especially in nitrogen-fixing biomes like grasslands and savannas, suggesting overstated vulnerability in projections for semi-arid transitions. In tropical biomes like the , CMIP6 ensembles predict localized dieback of 7% per degree of warming beyond 1.5°C, driven by reduced and fire propagation, potentially converting 10-20% of forest to by 2100 under high-emission pathways. Empirical assessments, however, document through post-drought recovery, with airborne and plot inventories from 2000-2020 showing no net loss in intact regions despite intensified dry seasons, and localized greening from secondary forest regrowth. Discrepancies arise partly from models' sensitivity to parameterized feedbacks like soil moisture-vegetation coupling, where observations indicate higher thresholds for dieback than simulated, underscoring potential overestimation of cascading vulnerabilities. Boreal and tundra biomes exhibit similar divergences: projections forecast permafrost thaw inducing 10-50% shrub expansion and carbon release equivalent to decades of emissions, yet ground-based and satellite data from 1980-2020 reveal stable or increasing aboveground biomass in many Alaskan and Siberian sites, buffered by CO2-enhanced water-use efficiency. Biome-scale analyses of ecosystem respiration temperature sensitivity further highlight gaps, with plot-level empirical data showing lower Q10 values (1.5-2.0) than model defaults (2.0-2.5), implying reduced projected vulnerability to warming-induced decomposition. These patterns suggest that while models capture directional shifts, empirical evidence—drawn from long-term monitoring networks—often demonstrates greater temporal stability and recovery rates, informed by internal variability rather than purely climatic forcing. Soil organic carbon (SOC) dynamics provide another lens of comparison, where models predict accelerated losses in temperate and biomes under warming, yet global compilations of 5,000+ soil profiles indicate divergent controls, with observations showing stabilization or gains in high-latitude grasslands due to microbial adaptations not fully represented in simulations. Such inconsistencies, spanning biomes, stem from parametric uncertainties in microbial processes and omissions, emphasizing the need for hybrid empirical-model frameworks to refine vulnerability assessments beyond alarmist equilibria assumptions.

Adaptation Mechanisms and Resilience Evidence

Adaptation in biomes occurs through physiological adjustments, genetic , and ecological processes such as species migration and community reassembly, enabling ecosystems to maintain function amid climatic perturbations. allows to alter traits like leaf morphology or in response to or changes, as demonstrated in common garden experiments with trees spanning 250 years, where local populations exhibited adaptive differentiation to climate variables. Genetic adaptation via has been observed in traits influencing and reproductive timing, with empirical data from provenance trials showing heritable shifts in growth rates under elevated CO2 and warming scenarios. Dispersal mechanisms, including seed banks and animal-mediated transport, facilitate range expansions, though empirical tracking via records and reveals lags in upslope migrations averaging 1-2 meters per decade in mountainous biomes. Resilience evidence, quantified through metrics like recovery time from disturbances and resistance to vegetation loss, draws from long-term satellite datasets such as NDVI fluctuations. Global analyses of two independent records indicate that while 29% of terrestrial ecosystems exhibit symptoms of resilience loss—manifested as slower recovery from droughts—certain biomes, including boreal forests and savannas, display enhanced resistance due to fire-adapted traits and microbial feedbacks that stabilize soil carbon. In drylands, microbiome communities show adaptive enzyme production that accelerates decomposition under warming, potentially buffering carbon losses, as evidenced by metagenomic surveys revealing shifts in microbial functional genes correlated with aridity gradients. However, tropical and temperate forests report declining resilience, with 64.5% of vegetated land showing reduced capacity to rebound from anomalies, attributed to water limitations rather than solely temperature. Ecological feedbacks amplify , such as nitrogen-fixing symbioses in grasslands enhancing under , supported by plot-scale experiments documenting 20-30% increases in response to simulated variability. Paleoecological records from cores illustrate historical biome transitions, like expansions during Pleistocene dry phases, with recovery lags of centuries underscoring inherent via surviving propagules. Empirical modeling of these mechanisms, calibrated against observed shifts, predicts that biomes with high functional redundancy—e.g., diverse shrublands—exhibit 1.5-2 times greater resistance to tipping points than monodominant forests. Yet, source critiques note that academic studies often underemphasize rapid evolutionary rates due to short observational windows, potentially overstating vulnerability in projections.

Modern Mapping and Research Advances

Remote Sensing Techniques and Data Integration

Remote sensing techniques for terrestrial biomes primarily rely on multispectral and hyperspectral optical sensors aboard satellites such as NASA's MODIS, Landsat series, and ESA's , which capture data across visible, near-, and shortwave bands to distinguish cover, structure, and essential for biome . These platforms enable global-scale mapping by quantifying biophysical parameters like (LAI) and fractional (fPAR), with MODIS products achieving NDVI accuracies of ±0.025 and EVI accuracies of ±0.015 through validation against ground measurements. () from complements optical data by penetrating cloud cover and providing structural information on canopy height and , particularly useful in dense forest biomes where optical signals saturate. Vegetation indices derived from these sensors, including NDVI calculated as (NIR - red)/(NIR + red), serve as proxies for biome productivity and greenness, with MODIS-derived NDVI time series from 2000 onward revealing biome-specific phenological patterns such as seasonal greening in temperate grasslands versus evergreen stability in tropical rainforests. Enhanced indices like EVI mitigate NDVI's saturation in high-biomass areas by incorporating blue band corrections for atmospheric and soil effects, improving discrimination of closed-canopy biomes. Hyperspectral sensors, though less operational at global scales, offer finer spectral resolution for functional trait mapping, as demonstrated in tundra biome studies distinguishing lichen from vascular plant dominance via narrow-band reflectance. Data integration enhances biome mapping accuracy by fusing multi-sensor and multi-temporal datasets, such as harmonized Landsat-8 and surface reflectance products at 30m resolution, which align spectral bands and reduce atmospheric artifacts for consistent global since 2015. Platforms like Google Earth Engine facilitate scalable integration of MODIS coarse-resolution (250-1000m) data with higher-resolution Landsat/ imagery, enabling temporal compositing to minimize cloud contamination and derive annual biome maps, as in the 2023 global functional ecosystem delineations using unsupervised clustering of vegetation structure metrics. Multi-angle observations from sensors like improve LAI estimates in heterogeneous biomes by accounting for bidirectional reflectance, with off-nadir NDVI yielding up to 5% better accuracy in compared to nadir-only views. Such integrations, validated against flux tower networks, support quantitative biome transitions, though residual uncertainties persist in arid and zones due to sparse .

Machine Learning and Unsupervised Classification Methods

Unsupervised methods, particularly clustering algorithms, have emerged as powerful tools for delineating biomes from high-dimensional datasets such as imagery, climate variables, and indices, enabling -driven discovery of natural groupings without predefined labels. These approaches contrast with supervised by relying on intrinsic patterns, such as signatures from satellites like MODIS or , to identify biome boundaries based on similarity metrics like or Ward's linkage. For instance, partitions into k clusters by iteratively minimizing intra-cluster variance, often applied to normalize difference vegetation index (NDVI) to map types corresponding to biomes. Hierarchical , including agglomerative variants, builds dendrograms to reveal nested structures, proving useful for terrestrial biome definition by integrating elevation, precipitation, and temperature . In practice, these methods have been employed to refine global biome maps. A 2022 study utilized on segmented environmental data to approximate terrestrial biomes, followed by explainable decision trees to interpret cluster rules, achieving alignments with established schemes like Köppen-Geiger while highlighting deviations in transitional zones. For marine environments, an unsupervised approach in 2020 applied to community and nutrient data from floats and satellite observations, identifying 18 global eco-provinces with boundaries matching oceanographic fronts, validated against chlorophyll-a gradients. Agglomerative has tracked carbon biomes in oceans, using 2011-2020 satellite-derived net and export flux data to detect 12 dynamic classes, revealing shifts in high-latitude regions with linkage criteria minimizing within-cluster sums of squares. Challenges in these applications include determining optimal cluster numbers—often via silhouette scores or elbow methods—and handling noise in remote sensing data, such as cloud cover artifacts, which can inflate variance. techniques like () precede clustering to mitigate the curse of dimensionality in multispectral bands, as demonstrated in 2024 reef habitat mapping using imagery, where variants outperformed k-means in robustness to outliers. Recent advances integrate these with geospatial constraints, such as incorporating in above-ground density for , yielding parsimonious global maps with reduced compared to grid-based partitioning. Empirical validation against ground-truthed plots remains essential, as outputs may conflate influences with natural biome signals absent causal priors.

Challenges in Global-Scale Empirical Mapping

Global-scale empirical of biomes encounters significant hurdles due to inconsistencies in schemes and data sources. Biome definitions vary across studies, with some emphasizing vegetation , others envelopes, leading to divergent maps; for instance, quantitative comparisons reveal that biome maps disagree most strongly in human-modified landscapes where natural vegetation is sparse or altered. techniques, reliant on spectral signatures, struggle with limits, often failing to resolve fine-scale heterogeneity or transitional zones, as evidenced by poor delimitation of forested areas in coarser datasets. Validation of global maps is hampered by scarce ground-truth data, particularly in remote or politically restricted regions, where field surveys are logistically challenging and costly. High-resolution satellite data, while improving coverage, incurs high processing expenses and suffers from atmospheric interference like , reducing usable observations in tropical biomes. Moreover, disturbances—such as or —frequently result in misclassification, with over half of globally disturbed mining lands labeled as "natural" in standard products. Distinguishing biomes with overlapping environmental traits poses further empirical difficulties; tropical dry forests and savannas, for example, occupy similar climatic niches, complicating delineation via remote proxies like (NDVI). Strict thresholds for features like canopy cover (e.g., 60% tree dominance) are often inferred remotely with error, exacerbating inaccuracies in dynamic ecosystems. Overall accuracies for global datasets underpinning biome maps hover around 60-80%, with discrepancies highest in heterogeneous or arid biomes, underscoring the need for integrated, multi-source validation frameworks.

Conceptual Debates and Criticisms

Limitations of Static Biome Constructs

Static biome classifications, which delineate ecosystems based on prevailing , vegetation structure, and at a given in time, inherently assume relative and states. However, ecosystems exhibit continuous dynamism through processes such as , periodic disturbances like or flooding, and interactions, rendering static constructs inadequate for representing these ongoing fluxes. This equilibrium paradigm overlooks how vegetation responses to environmental drivers operate on timescales from years to centuries, as evidenced by discrepancies between static maps and empirical observations of transient states in global models. A core limitation arises in the portrayal of biome boundaries as discrete lines, whereas real-world transitions occur across —zones of gradual intermingling where from adjacent biomes coexist and forms emerge due to overlapping environmental tolerances. These fuzzy interfaces, often spanning kilometers, challenge crisp delineations; for example, statistical analyses of vegetation gradients reveal probabilistic memberships rather than binary assignments, with ecotone widths varying by factors like and heterogeneity. Static maps exacerbate this by imposing artificial uniformity, leading to underestimation of hotspots in transitional areas and errors in assessing to perturbations. Paleoecological records further undermine static constructs by documenting recurrent biome migrations and state shifts driven by orbital forcings, volcanic events, and millennial-scale oscillations. and macrofossil assemblages from cores indicate that during the , forest-grassland boundaries in regions like shifted by hundreds of kilometers in response to variability, with no return to prior configurations due to lagged feedbacks in development and dispersal. Similarly, evidence shows tundra-taiga ecotones advancing and retreating across by up to 1,000 km per millennium during deglaciations, highlighting path dependency and absent in equilibrium-based classifications. These historical dynamics imply that modern static maps, calibrated to 20th-century conditions, misrepresent potential trajectories under analogous forcings. In predictive contexts, static biome frameworks constrain dynamic global vegetation models by prescribing fixed parameterizations for processes like carbon allocation and mortality, which fail to simulate emergent feedbacks such as changes or nutrient cycling that amplify shifts. Empirical comparisons of biome maps reveal inconsistencies in human-influenced landscapes, where agricultural conversion and create novel mosaics not captured by climate- correlations alone, resulting in up to 30% disagreement across schemes in altered regions. Consequently, reliance on static constructs can vulnerability assessments, as they undervalue adaptive capacities observed in long-term monitoring data from flux towers and time-series showing decadal vegetation greening or browning independent of mean . Addressing these limitations requires integrating process-based simulations with empirical gradients to better approximate causal mechanisms governing biome persistence and transformation.

Biases in Awareness and Prioritization (e.g., Forest vs. Open Biomes)

Scientific literature and conservation efforts exhibit a pronounced bias toward forested biomes, often at the expense of open biomes such as grasslands, savannas, and shrublands, a phenomenon termed Biome Awareness Disparity (BAD). This disparity manifests in disproportionate research funding, publication volume, and policy prioritization for forests, which constitute approximately 31% of global terrestrial land cover, while open biomes encompass about 40-50% of ice-free land and support unique biodiversity adapted to fire, herbivory, and low tree density. Restoration experiments, for instance, are concentrated in rainforests, dry forests, and mangroves—biomes overrepresented relative to their extent—while savannas and grasslands receive minimal attention, leading to misguided "tree-centric" interventions that suppress native herbaceous vegetation. Such biases extend to public discourse and media, where analyses reveal tweets emphasizing conservation far exceed those for open biomes proportional to land area; for example, forest-related posts dominate, potentially influencing donor priorities and agendas. This prioritization overlooks the services of open biomes, including carbon storage in belowground structures (often underestimated in models favoring aboveground ) and habitats for endemic like those in tropical grassy biomes, which rival forests in diversity but face threats from and woody encroachment. Empirical studies indicate that open in climates suitable for forests are frequently misattributed to anthropogenic rather than natural edaphic or disturbance-driven stability, perpetuating a that undervalues their intrinsic persistence. Contributing factors include terminological preferences, such as "" applied broadly to woody loss while grassland conversion is termed "land clearing," reinforcing perceptions of as default stable states. Broader ecological research biases amplify this, with forest biomes comprising 87% of studies on canopy structure and despite open systems' global prevalence, potentially rooted in researcher demographics from temperate or forested regions and funding incentives tied to high-visibility metrics. While peer-reviewed critiques acknowledge these patterns, mainstream environmental narratives—often shaped by institutional emphases—rarely integrate open biome resilience, risking policy failures like inappropriate that erodes in fire-dependent systems. Addressing BAD requires empirical reorientation toward biome-specific dynamics, prioritizing data from underrepresented systems to inform causal mechanisms like disturbance regimes over assumptive forest analogies.

Integration of Historical Contingency and Feedbacks

The distribution of modern terrestrial biomes bears the imprint of historical contingencies, particularly the repeated glacial-interglacial cycles of the Pleistocene epoch, which spanned approximately 2.58 million to 11,700 years ago and profoundly influenced species dispersal, refugia, and community assembly. During the around 21,000 years ago, vast ice sheets covered northern hemispheres, compressing biomes equatorward and fragmenting habitats, while unglaciated refugia served as survival pockets for and . Post-, biome expansions were uneven due to dispersal limitations and priority effects, where early colonizers shaped subsequent community structures; for instance, boreal forests in exhibit persistent deciduous dominance in western regions versus evergreen in eastern ones, attributable to distinct northern refugia during glaciation that locked in alternative vegetation states. Similarly, global vegetation modeling indicates that since deglaciation, cold-adapted biomes migrated poleward by hundreds of kilometers, but with lags in forest expansion into former , reflecting historical bottlenecks rather than equilibrium with current climates. Ecological feedbacks integrate with these contingencies by amplifying or stabilizing historical legacies through self-reinforcing interactions between and abiotic factors. Negative feedbacks, such as vegetation-induced soil cycling or effects, maintain biome boundaries; at savanna-forest ecotones, intense belowground for water and favors grasses in open areas via rapid root proliferation, preventing woody encroachment and enforcing where either state can persist under similar climates depending on initial conditions set by history. -vegetation feedbacks exemplify this in flammable biomes: grasslands promote frequent low-intensity s that suppress trees, while forests reduce fuel loads and frequency, creating where crossing thresholds requires extreme perturbations, as observed in and systems where prehistoric human ignition altered natural dynamics. The extinctions, circa 12,000–10,000 years ago, disrupted herbivory feedbacks, leading to denser woody and altered carbon storage in grasslands, a legacy detectable in modern proxies and patterns. This integration reveals biomes as dynamic outcomes of contingency-initiated trajectories modulated by feedbacks, challenging deterministic climate-only models. Historical events provide "path dependence," where early post-glacial assemblages trigger feedbacks that resist reversion; for example, Pliocene-Pleistocene transition pollen records show interglacial forests mirroring preceding glacial aridity rather than contemporaneous warmth, due to lagged recolonization and soil legacy effects. In boreal zones, glacial refugia not only seeded species pools but also imprinted microbial communities that sustain differential decomposition rates, influencing carbon feedbacks today. Empirical paleoecological data thus underscore that biome resilience hinges on these intertwined processes, with contingencies explaining deviations from predicted equilibria—such as persistent "non-analog" communities—and feedbacks dictating tipping points, as evidenced by modeling that incorporates both to forecast slower-than-expected shifts under warming. Such realism counters oversimplified projections by highlighting causal chains from past disturbances through biotic interactions to enduring patterns.

References

  1. [1]
    The world's biomes - University of California Museum of Paleontology
    Biomes are defined as the world's major communities, classified according to the predominant vegetation and characterized by adaptations of organisms to that ...
  2. [2]
    Earth's Changing Biomes - UCAR Center for Science Education
    Biomes are generally defined by the types of vegetation, soil, climate, and wildlife within them. Note that a biome is different from an ecosystem.
  3. [3]
    Mission: Biomes - NASA Earth Observatory
    A biome is a community of plants and animals living together in a certain kind of climate. Scientists have classified regions of the world into different biomes ...Tundra · Shrubland · To Plant or Not To Plant? · DesertMissing: ecology | Show results with:ecology
  4. [4]
    World Biomes by Kids Do Ecology - KDE Santa Barbara
    Biomes are regions of the world with similar climate (weather, temperature) animals and plants. There are terrestrial biomes (land) and aquatic biomes, both ...Rainforest · Desert · Tundra · Temperate grassland
  5. [5]
    Biology, Ecology, Ecology and the Biosphere, Terrestrial Biomes
    The Earth's biomes are categorized into two major groups: terrestrial and aquatic. Terrestrial biomes are based on land, while aquatic biomes include both ...
  6. [6]
    [PDF] Interpreting Whittaker Biome Diagrams - Global Vegetation Project
    These biome diagrams display where a point fits into Whittaker's terrestrial biome classification system which explains major global patterns of vegetation ...
  7. [7]
    Biome: evolution of a crucial ecological and biogeographical concept
    Biome defined as lump sum of ecoregions that were defined on manifold (yet often unclear) criteria, Olson et al. (2001). Functional ecosystem characteristic ...
  8. [8]
    6.5 - Major Biomes | Soil Genesis and Development, Lesson 6
    Biome is a term used to classify the Earth's major ecosystems. A biome is defined primarily by the climate and predominant vegetation of a region.
  9. [9]
    Biome - Definition and Examples - Biology Online Dictionary
    Aug 25, 2023 · Biome (biology definition): A major ecological community of organisms adapted to a particular climatic or environmental condition on a large ...Biome Definition · Types Of Biomes · Forests · Grassland
  10. [10]
    An operational definition of the biome for global change research
    Apr 7, 2020 · Summary. Biomes are constructs for organising knowledge on the structure and functioning of the world's ecosystems, and serve as useful units ...
  11. [11]
    Biomes, Ecosystems, and Habitats - National Geographic Education
    Oct 19, 2023 · The key difference between biomes, ecosystems, and habitats is scale. National Geographic Society.Missing: distinction | Show results with:distinction
  12. [12]
    Ecosystems and biomes (video) | Ecology - Khan Academy
    Oct 20, 2016 · Land ecosystems are often categorized into biomes, such as tropical forests or deserts, based on climate and vegetation. Created by Sal Khan.Missing: ecoregion | Show results with:ecoregion
  13. [13]
    Habitats and ecosystems: what's the difference, and how they affect ...
    Jan 3, 2023 · A habitat is a geographic location, an ecosystem is a set of interactions among species – including who-eats-whom in a food chain – and between the living and ...Missing: ecoregion | Show results with:ecoregion<|separator|>
  14. [14]
    Chapter 8 ~ Biomes and Ecozones – Environmental Science
    Ecoregions and Ecozones. As we have learned, biomes are geographically extensive ecosystems that occur anywhere in the world where environmental conditions are ...Missing: distinction | Show results with:distinction
  15. [15]
    Global climate and the distribution of plant biomes - Journals
    This approach describes a wide range of biomes that can be correlated with characteristic climatic conditions, or climatic envelopes.
  16. [16]
    Terrestrial Biomes | Learn Science at Scitable - Nature
    Terrestrial biomes are distinguished primarily by their predominant vegetation, and are mainly determined by temperature and rainfall.<|control11|><|separator|>
  17. [17]
    Biomes and Climate | CK-12 Foundation
    Because climate determines plant growth, it also influences the number and variety of other organisms in a terrestrial biome. Biodiversity generally increases ...
  18. [18]
    4|Climate and Vegetation
    Climate is the major determinant of vegetation. Plants in turn exert some degree of influence on climate. Both climate and vegetation profoundly affect soil ...
  19. [19]
    Biome-level relationships between vegetation indices and climate ...
    Climatic factors such as rainfall and temperature play a vital role in the growth characteristics of vegetation. While the relationship between climate and ...
  20. [20]
    Terrestrial biomes: a conceptual review
    Jun 30, 2021 · We attempt to review the conceptualisation, science and classification of biomes and propose to limit the definition of a biome to potential natural vegetation.<|separator|>
  21. [21]
    The influence of soil age on ecosystem structure and function across ...
    Sep 18, 2020 · Across biomes, we found that soil age was a significant but relatively weak ecosystem driver of change. Taken together, parent material, climate ...
  22. [22]
    Soils and Biomes - SERC (Carleton)
    Jan 13, 2021 · Biomes are influenced not only by climatic factors (above-ground) but also by the soils that are located below-ground. Climate and plant type ...
  23. [23]
    The role of soils in habitat creation, maintenance and restoration
    Aug 4, 2021 · Soils are an important modifier of the occurrence of different biomes as water availability is not just a function of precipitation, but rather ...
  24. [24]
    [PDF] Databases of Model Drivers and Validation Measurements - NASA
    The data were reviewed at the EMDI I Workshop, and a strategy was developed for additional. “outlier analysis” to flag those NPP measurements that, in ...
  25. [25]
    Land cover fraction mapping across global biomes with Landsat ...
    Sep 1, 2024 · We applied machine learning regression-based fraction mapping to quantify land cover fractions of 18 regions in five biomes using Landsat data from 2022.
  26. [26]
    Proof of evidence of changes in global terrestrial biomes using ...
    Our findings show that the average NDVI values in most biomes have increased significantly (F-value<0.01) by 0.05 ndvi units over during the past three decades, ...
  27. [27]
    Empirical evidence for recent global shifts in vegetation resilience
    Apr 28, 2022 · Here, we quantify vegetation resilience globally with complementary metrics based on two independent long-term satellite records.
  28. [28]
    Predicting global terrestrial biomes with the LeNet convolutional ...
    Apr 18, 2022 · We demonstrate an accurate and practical method to construct empirical models for biome mapping: a convolutional neural network (CNN) was trained by an ...Missing: evidence | Show results with:evidence
  29. [29]
    How to map biomes: Quantitative comparison and review of biome ...
    Jun 19, 2024 · Satellite imagery has provided a means to automate the delineation of biomes using standard observations at the global scale, thereby reducing ...
  30. [30]
    Biome: evolution of a crucial ecological and biogeographical concept
    Nov 27, 2018 · Vegetation is not a passive entity under the control of the environment. Across spatial scales, feedbacks between vegetation and climate, soils ...
  31. [31]
    History of Ecological Sciences, Part 54: Succession, Community ...
    Jul 1, 2015 · The history of studies on succession, community, and continuum is primarily within the domain of plant ecology, which seems appropriate.Missing: precursors | Show results with:precursors
  32. [32]
    [PDF] Victor Ernest Shelford, Eminent Ecologist, 1968
    Dec 12, 2014 · Bio-Ecology, in 1939. With the biome concept now established, Dr. Shelford set himself the task of describing all the biomes and major serai ...<|separator|>
  33. [33]
    Climate and Biomes - SERC (Carleton)
    Jul 5, 2011 · Biomes are regions of Earth that have similar climates and other abiotic abiotic: physical factors or conditions that influence plant and animal life.
  34. [34]
    On the Theoretical Concept of the Potential Natural Vegetation ... - jstor
    According to TUXEN (1956: 5) the term "potential natural vegetation" is defined as an. "imagined natural state of vegetation ... that could be outlined for the ...
  35. [35]
    Understanding properly the `potential natural vegetation' concept
    May 18, 2010 · This concept was one of Tüxen's notable contributions to vegetation science and it dates from the middle of the last century (Tüxen, 1956).
  36. [36]
    History of Vegetation Classification
    For example, in the U.S. Interior West, early land classifications such as the concept of potential natural vegetation (Küchler 1969) and habitat type ( ...
  37. [37]
    The relevance of the concept of potential natural vegetation in the ...
    The concept of potential natural vegetation (PNV) refers to self-sustaining mature vegetation matching the environmental conditions a site offers.
  38. [38]
    [PDF] The concept of potential natural vegetation: an epitaph? - BayCEER
    Oct 8, 2010 · We discuss the usefulness of the concept of Potential. Natural Vegetation (PNV), which describes the ex- pected state of mature vegetation in ...<|separator|>
  39. [39]
    [PDF] Predicting the distribution of potential natural vegetation based on ...
    Potential natural vegetation (PNV) is the vegetation devel- oped in a given habitat in absence of human interferences (Zerbe 1998), which is usually the climax ...
  40. [40]
    Plant Functional Diversity and the Biogeography of Biomes in North ...
    Dec 17, 2018 · Because biomes are characterized by their dominant vegetation, we also examined the geographic commonness and functional distinctiveness of ...
  41. [41]
    Global patterns of plant functional traits and their relationships to ...
    Sep 13, 2024 · Plant functional traits (FTs) determine growth, reproduction and survival strategies of plants adapted to their growth environment.
  42. [42]
    A function-based typology for Earth's ecosystems - Nature
    Oct 12, 2022 · The classification and descriptive profiles, including maps, for each functional biome and ecosystem functional group underwent extensive ...
  43. [43]
    Traditional plant functional groups explain variation in economic but ...
    Dec 10, 2018 · Plant functional groups are widely used in community ecology and earth system modelling to describe trait variation within and across plant ...
  44. [44]
    [PDF] Accelerated shifts in terrestrial life zones under rapid climate change
    2080) life zones based on (d) an extension of the Holdridge Life Zone classification system (Holdridge, 1947), as. 952 defined by three parameters ...
  45. [45]
    Details - The Life Zone System - Biodiversity Heritage Library
    The Life Zone System. By. Holdridge, L R . Type. Article. Date of Publication. 1966. Original Publication. Adansonia. Volume. 6. Series / Issue. Issue: 2 ...
  46. [46]
    Holdridge Life Zones - Overview - ArcGIS Online
    It was first published by Leslie Holdridge in 1947, and updated in 1967. It is a relatively simple system based on few empirical data, giving objective mapping ...
  47. [47]
    8. Biomes of the World - David Zelený
    Jan 9, 2021 · Whittaker also introduced a large scale biome classification (figure 7), based on mean annual temperature and mean annual precipitation, but ...Large scale environmental... · Dominant life forms of... · Biome classification...
  48. [48]
    [PDF] WORKING PAPER - IIASA PURE
    Based on temperature and precipitation, Whittaker (1975) designed a chart showing the pattern of world biome-types in relation to these two climatic factors ( ...
  49. [49]
    Two very popular biome schemes based on bioclimatic approach
    Two very popular biome schemes based on bioclimatic approach: (a) Holdridge's scheme (reproduced from Archibold, 1995; after Holdridge, 1947); (b) Whittaker's ...
  50. [50]
    Climatic definitions of the world's terrestrial biomes
    Dec 19, 2022 · A hierarchical classification is proposed for the biotic units within four main domains: Cryocratic, Mesocratic, Xerocratic and Thermocratic, divided into 7 ...
  51. [51]
    [PDF] Heinrich Walter - Vegetation of the Earth
    Classification of the Geo-biosphere into Zonobiomes . 2. 3. Zonoecotones. 5. 4. Oro biomes . 5. 5. Pedobiomes. 6. 6. Biomes. . . 7. 7. The Nature and Structure ...
  52. [52]
    Biomes of the Southern Hemisphere - SpringerLink
    Revising the Heinrich Walter's zonobiome system for the Southern Hemisphere appeared as necessary because of the bioclimatic imbalance between the Hemispheres.
  53. [53]
    [PDF] The Biome Concept in Ecology - Ecologia
    Two ways of classifying biomes are represented by the climate zone approach of Walter and the vegetation approach exemplified by Whittaker. The first ...
  54. [54]
    Ecoregions: The Ecosystem Geography of the Oceans and Continents
    Robert Bailey's system for classifying ecoregions has had a major influence, and has been adopted by major organizations such as the US Forest Service and The ...Missing: classification | Show results with:classification
  55. [55]
    Bailey's ecoregions and subregions of the United States, Puerto ...
    This map layer includes ecoregions and subregions information for the United States, the District of Columbia, Puerto Rico, and the U.S. Virgin Islands. Lineage ...Identification_Information · Data_Quality_Information
  56. [56]
    Description of the Ecoregions of the United States
    A publication compiled by Robert G. Bailey, March 1995, provides a general description of the ecosystem geography of the United States.Missing: classification | Show results with:classification
  57. [57]
    (PDF) Identifying Ecoregion Boundaries - ResearchGate
    Aug 7, 2025 · This article summarizes the rationale I used in identifying ecoregion boundaries on maps of the United States, North America, and the world's continents.
  58. [58]
    [PDF] A World Ecoregions Map for Resource Reporting - GIS-Lab
    The primary purpose of the map will be to serve as a reporting structure for information about global resources and environment, though it will be based largely ...
  59. [59]
    Terrestrial Ecoregions of the World: A New Map of Life on Earth
    The ecoregion map complements global priority-setting analyses, such as Global 200 (Olson and Dinerstein 1998) and Hotspots (Myers et al. 2000), by providing an ...
  60. [60]
    Terrestrial Ecoregions of the World - Data Basin
    Oct 27, 2010 · This map depicts the 825 terrestrial ecoregions of the globe. Ecoregions are relatively large units of land containing distinct assemblages of natural ...
  61. [61]
    Terrestrial Ecoregions of the World: A New Map of Life on Earth
    Aug 6, 2025 · 2001). The ecoregion map complements global priority-setting. analyses, such as Global 200 (Olson and Dinerstein 1998) and. H ...
  62. [62]
    World Wildlife Fund - Global 200 (terrestrial) Ecoregions
    May 13, 2010 · The Global 200 ecoregions represent those ecoregions where WWF is initially focusing its ecoregion conservation efforts to develop biodiversity ...
  63. [63]
    Ecoregions: The Ecosystem Geography of the Oceans and Continents
    Bailey's system for classifying these zones has been adopted by major organizations such as the U.S. Forest Service and The Nature Conservancy and this book is ...
  64. [64]
    [PDF] Putting people in the map: anthropogenic biomes of the world
    EC Ellis and N Ramankutty. Anthropogenic biomes of the world al. 2001). Anthropogenic enhancement of natural landscape heterogeneity represents a secondary ...<|control11|><|separator|>
  65. [65]
    [PDF] Anthropogenic transformation of the biomes, 1700 to 2000
    Main conclusions Between 1700 and 2000, the terrestrial biosphere made the critical transition from mostly wild to mostly anthropogenic, passing the 50% mark.
  66. [66]
    Annual 30-m maps of global grassland class and extent (2000–2022 ...
    Dec 11, 2024 · The paper describes the production and evaluation of global grassland extent mapped annually for 2000–2022 at 30 m spatial resolution.Missing: post- | Show results with:post-
  67. [67]
    A novel biome concept and classification system based on ...
    Oct 7, 2021 · We propose a comprehensive and hierarchical classification method and nomenclature to define biomes based on a set of bioclimatic variables and ...Missing: peer- | Show results with:peer-
  68. [68]
    [PDF] Guidance on biomes
    This guidance introduces biomes, maps sectors to them, provides assessment indicators, and supports identification of nature-related issues. It includes land, ...Missing: post- | Show results with:post-<|separator|>
  69. [69]
    Biome classification influences current and projected future biome ...
    Nov 21, 2023 · Early biome classification schemes were based on dominant life forms with similar physiological characteristics (Schimper, 1903), climate zones ...
  70. [70]
  71. [71]
    Terrestrial Biomes – Environmental Biology
    There are eight major terrestrial biomes: tropical rainforests, savannas, subtropical deserts, chaparral, temperate grasslands, temperate forests, boreal ...
  72. [72]
    8.2: Terrestrial Biomes - Biology LibreTexts
    Sep 5, 2022 · Tropical rainforests are characterized by vertical layering of vegetation and the formation of distinct habitats for animals within each layer.
  73. [73]
    Biomes of the World - Ask A Biologist
    Jul 19, 2013 · A biome is a type of environment that is defined by the types of organisms that live there. We can also think of these as life zones.Virtual Biomes · Grassland Biome · Savanna Biome · Tundra Biome
  74. [74]
    Terrestrial Biomes - OpenEd CUNY
    There are eight major terrestrial biomes: tropical wet forests, savannas, subtropical deserts, chaparral, temperate grasslands, temperate forests, boreal ...
  75. [75]
    Terrestrial Biomes - Intro to Ecology Study Guide 2024 - Fiveable
    Animal adaptations in desert biomes cope with water scarcity and extreme temperatures ... Animals in cold biomes possess adaptations for survival in harsh winters.
  76. [76]
    Freshwater (Lakes and Rivers) and the Water Cycle - USGS.gov
    The definition of freshwater is water containing less than 1,000 milligrams per liter of dissolved solids, most often salt. As a part of the water cycle, ...Missing: characteristics | Show results with:characteristics
  77. [77]
    Lakes and Reservoirs | U.S. Geological Survey - USGS.gov
    Characteristics of lakes​​ Climate: Temperature, wind, precipitation, and solar radiation all critically affect the lake's hydrologic and chemical ...
  78. [78]
    Submerged Aquatic Vegetation: A Habitat Worth SAV-ing
    Jul 27, 2020 · SAV is a great habitat for fish, including commercially important species, because it provides them with a place to hide from predators.Missing: characteristics | Show results with:characteristics
  79. [79]
    Marine life | National Oceanic and Atmospheric Administration
    May 26, 2021 · Our ocean, coasts, and estuaries are home to diverse living things. These organisms take many forms, from the tiniest single-celled plankton to the largest ...Marine mammals · Coral reef ecosystems · Aquatic food webs · Sea turtles
  80. [80]
    Layers of the Ocean - NOAA
    Mar 28, 2023 · The temperature in the bathypelagic zone, unlike that of the mesopelagic zone, is constant. The temperature never fluctuates far from a chilling ...
  81. [81]
    Ocean floor features - NOAA
    While the ocean has an average depth of 2.3 miles, the shape and depth of the seafloor is complex. Some features, like canyons and seamounts, might look ...
  82. [82]
    What is the intertidal zone? - NOAA's National Ocean Service
    Jun 16, 2024 · Intertidal zones exist anywhere the ocean meets the land, from steep, rocky ledges to long, sloping sandy beaches and mudflats that can extend for hundreds of ...
  83. [83]
    Coral reef ecosystems | National Oceanic and Atmospheric ...
    Sep 25, 2025 · This area supports more than 7,000 species of fishes, invertebrates, plants, sea turtles, birds, and marine mammals. Deep water reefs or mounds ...
  84. [84]
    Estuary Habitat - NOAA Fisheries
    Estuaries are bodies of water where rivers meet the sea. They provide homes for diverse wildlife, including popular fish species.
  85. [85]
    What Is an Estuary? - National Estuarine Research Reserve System
    An estuary is an ecosystem, comprising both the biological and physical environment, that has developed in a region where rivers meet the sea.Missing: ecotones marine biomes
  86. [86]
    Salt Marsh is a wetland that has shallow water and levels ... - NVCS
    Salt marshes have gradients that include barren salt flats at the tidal edge, rushes, and then halophytic herbs and grasses at the outer edge. Daily drawdowns ...
  87. [87]
    Salt marsh‐mangrove ecotones: using structural gradients to ...
    Mar 11, 2016 · Marshes in the driest location (Central Texas) had higher salinities and were dominated by low biomass succulent plants and lower soil carbon ...
  88. [88]
    Estuarine ecosystems - Coastal Wiki
    Aug 20, 2025 · Estuaries represent the transition between freshwater and marine environments and are influenced by both aquatic realms. They are referred ...<|separator|>
  89. [89]
    [PDF] Anthromes
    Anthromes, or anthropogenic biomes, characterize the globally significant ecological patterns created by sustained direct human interactions with ecosystems, ...
  90. [90]
    Ecology in an anthropogenic biosphere - Ellis - ESA Journals - Wiley
    Aug 1, 2015 · (B) Anthropogenic biomes (anthromes; year 2000; modified from Ellis et al. 2010). Anthromes are organized into five levels in the legend.
  91. [91]
    Anthropogenic Biomes: 10,000 BCE to 2015 CE - MDPI
    In 2008, Ellis and Ramankutty introduced a novel approach to map the global patterns of human transformation of the terrestrial biosphere, analogous to the ...Missing: post- | Show results with:post-
  92. [92]
    Microbiome definition re-visited: old concepts and new challenges
    Jun 30, 2020 · This term refers to the entire habitat, including the microorganisms (bacteria, archaea, lower and higher eurkaryotes, and viruses), their ...
  93. [93]
    Microbial Ecology - an overview | ScienceDirect Topics
    Microbial ecology is defined as the study of microbes and their relationships with both biotic (living) and abiotic (non-living) components of the environment. ...
  94. [94]
    [PDF] Microbial Ecology - Elsevier
    Biomes are large geographical areas defined by macroecologists on the basis of their distinc- tive flora and fauna, which presumably reflect selection and ...<|separator|>
  95. [95]
    5.4.2: Climate and Biomes - Biology LibreTexts
    Jul 28, 2025 · 1 : Average annual temperature and annual precipitation are two climatic factors that determine the distribution of biomes. This graph shows ...
  96. [96]
    3.3 Terrestrial Biomes – Environmental Biology
    These layers provide diverse and complex habitats for the variety of plants, animals, and other organisms. Many species of animals use the variety of plants and ...
  97. [97]
    The Holdridge life zones of the conterminous United States in ...
    The most extensive life zone is the warm temperate moist forest, which covers 23% of the country. We compared the Holdridge life zone map with output from the ...Missing: basis | Show results with:basis
  98. [98]
    Edaphic Factors - an overview | ScienceDirect Topics
    Edaphic factors include those physical, chemical, and biological properties of a soil that influence plant growth. The various properties that are important for ...
  99. [99]
    Effects of soil pH on the growth, soil nutrient composition, and ...
    Apr 16, 2024 · The ideal soil pH for plant growth is between 6.5 and 7.5. Soils that are too acidic or alkaline can negatively affect the physical properties ...
  100. [100]
    The influence of soil age on ecosystem structure and function across ...
    Sep 18, 2020 · Parent material, climate, vegetation and topography predict, collectively, 24 times more variation in ecosystem properties than soil age alone.
  101. [101]
    Biotic interactions contribute more than environmental factors and ...
    Mar 8, 2023 · The predominant driving factor of soil microbial community assembly in arid ecosystem was biotic interactions between microorganisms, followed by environmental ...
  102. [102]
    Trophic interactions among vertebrate guilds and plants shape ...
    Jul 25, 2018 · Trophic interactions play critical roles in structuring biotic communities. Understanding variation in trophic interactions among systems ...
  103. [103]
    [PDF] Trophic Interactions and Abiotic Factors Drive Functional and ...
    In this study, we have shown that trophic interactions drive both the phylogenetic and functional structure of communities across a whole guild and biome.
  104. [104]
    Review Biotic responses to climate extremes in terrestrial ecosystems
    Jul 15, 2022 · These aboveground-belowground interactions drive biotic feedbacks that influence mass and energy flow between the two subsystems (Wardle et al., ...
  105. [105]
    Biotic Feedbacks in the Global Climatic System: Will the Warming ...
    Oct 31, 2023 · Biotic Feedbacks in the Global Climatic System: Will the Warming Feed the Warming? ... Global Warming and Terrestrial Ecosystems: A ...
  106. [106]
    [PDF] ECOSYSTEMS AS CLIMATE CONTROLLERS – BIOTIC FEEDBACKS
    Biotic feedbacks of oceanic ecosystems. The main element of the oceanic ecosystems and biogeochemical cycle is the carbon. One of the most important feedback ...
  107. [107]
    Extreme specificity in obligate mutualism—A role for competition?
    Jun 21, 2024 · The high degree of specificity in obligate mutualisms is driven by competition within obligate mutualist guilds that limits species richness. As ...
  108. [108]
    (PDF) Ecosystems as climate controllers - Biotic feedbacks
    Home · Education · Educational Assessment · Feedback. ArticlePDF Available. Ecosystems as climate controllers - Biotic feedbacks. December 2008; Applied Ecology ...
  109. [109]
    Integrating ecological feedbacks across scales and levels of ...
    Apr 30, 2024 · Such changes of the emergent biotic feedbacks ultimately affect the stability of communities (Mougi and Kondoh 2012, Coyte et al. 2015) ...
  110. [110]
    Early life on land and the first terrestrial ecosystems
    Feb 23, 2013 · Terrestrial ecosystems have been largely regarded as plant-dominated land surfaces, with the earliest records appearing in the early Phanerozoic (<550 Ma).Missing: shifts biomes<|control11|><|separator|>
  111. [111]
    The Heavy Links between Geological Events and Vascular Plants ...
    Feb 4, 2016 · The high CO2 level allowed the origination of early terrestrial plants and the rise of the first forest in Silurian times.Missing: transitions biomes
  112. [112]
    Synchronizing climate-carbon cycle heartbeats in the Phanerozoic ...
    Oct 16, 2025 · Furthermore, ecosystem modeling of late Carboniferous terrestrial ecosystems indicates major changes in high-latitude forest cover during ...
  113. [113]
    Miocene biome turnover drove conservative body size evolution ...
    Oct 17, 2018 · Abstract. On deep time scales, changing climatic trends can have a predictable influence on macroevolution. From evidence of mass extinctions, ...
  114. [114]
    Ecological dynamics of terrestrial and freshwater ecosystems across ...
    Mar 17, 2021 · We explore ecological impacts of terrestrial and freshwater ecosystems in three mass extinctions through the mid-Phanerozoic, a span of 121 million years.
  115. [115]
    Climate windows of opportunity for plant expansion during ... - Nature
    Aug 4, 2022 · One of the biggest changes in paleogeography during the Phanerozoic was the breakup of the supercontinent Pangea (Fig. S1) which saw the ...
  116. [116]
    The global vegetation pattern across the Cretaceous–Paleogene ...
    The most striking transition in the macroflora is the significant decrease of Araucariaceae across the K–Pg boundary (Pole, 2008). High floral diversity in the ...
  117. [117]
    Phanerozoic paleotemperatures: The earth's changing climate ...
    Temperatures are high during the Early Paleozoic. Cooler temperatures prevail during the Late Paleozoic, followed by warmer Mesozoic and early Cenozoic ...Missing: shifts terrestrial biomes
  118. [118]
    what the geological record tells us about our present and future ...
    Dec 28, 2020 · Climate change on decadal to centennial time-scales is well recorded in tree rings, corals, bivalves, marine and lake sediments, cave deposits ...
  119. [119]
    Anthropogenic transformation of the biomes, 1700 to 2000 - 2010
    Aug 4, 2010 · Between 1700 and 2000, the terrestrial biosphere made the critical transition from mostly wild to mostly anthropogenic, passing the 50% mark early in the 20th ...
  120. [120]
    COP26: Agricultural expansion drives almost 90 percent of global ...
    Jun 11, 2021 · Agricultural expansion drives almost 90 percent of global deforestation – an impact much greater than previously thought.
  121. [121]
    Land statistics 2001–2023. Global, regional and country trends
    Jun 19, 2025 · Since 2001, world total cropland area grew by about 80 million ha (5 percent) while permanent meadows and pastures lost 150 million ha, a ...
  122. [122]
    Deforestation Linked to Agriculture | Global Forest Review
    Apr 4, 2024 · Globally, soy farms occupy 8.2 Mha of land deforested between 2001 and 2015. ... Almost all (97 percent) of this deforestation occurred in South ...
  123. [123]
    Agricultural expansion and the ecological marginalization of forest ...
    Oct 25, 2021 · According to this protocol, active deforestation frontiers are characterized by high forest cover (>66%) and high deforestation rates (>2.5% per ...
  124. [124]
    [PDF] Four-century history of land transformation by humans in the United ...
    Mar 3, 2023 · The US has seen dramatic land transformation through land clearing, agricultural expansion, urban sprawl, and loss of natural vegetation.
  125. [125]
    Land cover changes in desert areas 1700, 1900, 2000 and 2050
    The main land use change in desert areas has been the conversion of relatively barren dry lands for agricultural needs, partially through irrigation.
  126. [126]
    Half of the world's habitable land is used for agriculture
    More than three-quarters of global agricultural land is used for livestock, despite meat and dairy making up a much smaller share of the world's protein and ...
  127. [127]
    Anthropogenic transformation of the terrestrial biosphere - Journals
    Mar 13, 2011 · Human populations and their use of land have transformed most of the terrestrial biosphere into anthropogenic biomes (anthromes), causing a variety of novel ...
  128. [128]
    Biodiversity impacts of recent land-use change driven by increases ...
    Sep 20, 2024 · Overall, more than 90% of global biodiversity impacts of land-use change are due to agriculture, with crops cultivation (72%) and pastures (21%) ...
  129. [129]
    A reconstruction of global agricultural areas and land cover for the ...
    Aug 21, 2008 · About 5 million km2 of natural vegetation are found to be transformed to agriculture between AD 800 and 1700, slightly more to cropland (mainly ...
  130. [130]
    Deforestation and Forest Loss - Our World in Data
    Between 2010 and 2020, the net loss in forests globally was 4.7 million hectares per year. However, deforestation rates were much higher. The UN FAO estimates ...Imported deforestation · Forests and Deforestation · Annual deforestation, 2015
  131. [131]
    Agriculture drives more than 90% of tropical deforestation | SEI
    Sep 8, 2022 · Yet only half to two-thirds of this results in the expansion of active agricultural production on the deforested land.
  132. [132]
    Explore the World's Protected Areas
    Discover the world's protected and conserved areas. Protected Planet is the most up to date and complete source of data on protected areas and other ...Protected Areas (WDPA) · Marine Protected Areas · Explore protected areas and...
  133. [133]
    World must act faster to protect 30% of the planet by 2030 - UNEP
    Oct 28, 2024 · Protected and conserved areas must almost double in area on land and more than triple in the ocean for the 30% target to be reached by 2030.
  134. [134]
    Protected Planet Report 2024
    The Protected Planet Report 2024 is the first report to fully assess the global status of protected and conserved areas in the context of Target 3.
  135. [135]
    Mixed effectiveness of global protected areas in resisting habitat loss
    Sep 27, 2024 · Protected areas are the cornerstones of conservation efforts to mitigate the anthropogenic pressures driving biodiversity loss.<|separator|>
  136. [136]
    A global meta-analysis on the ecological drivers of forest restoration ...
    May 19, 2016 · Our meta-analysis encompassing 221 study landscapes worldwide reveals forest restoration enhances biodiversity by 15–84% and vegetation structure by 36–77%.Missing: biome | Show results with:biome
  137. [137]
    Ecological restoration success is higher for natural regeneration ...
    Nov 8, 2017 · Restoration success for biodiversity and vegetation structure was 34 to 56% and 19 to 56% higher in natural regeneration than in active ...Missing: biome | Show results with:biome
  138. [138]
    Coral restoration – A systematic review of current methods ...
    Jan 30, 2020 · Overall, coral restoration projects focused primarily on fast-growing branching corals (59% of studies), and report survival between 60 and 70%.Missing: biome | Show results with:biome
  139. [139]
    Enhanced but highly variable biodiversity outcomes from coastal ...
    Apr 19, 2024 · For example, overall coral reef restoration was effective for restoring animal populations relative to degraded sites (e.g., Ferse, Ku'ulei et ...
  140. [140]
    First-of-its-kind study definitively shows that conservation actions are ...
    Apr 25, 2024 · It shows that nature conservation truly works, from the species to the ecosystem levels across all continents. This analysis, led by Re:wild in ...Missing: terrestrial | Show results with:terrestrial
  141. [141]
    Terrestrial ecosystem restoration increases biodiversity and reduces ...
    May 12, 2022 · We found that, relative to unrestored (degraded) sites, restoration actions increased biodiversity by an average of 20%, while decreasing the variability of ...Missing: biome | Show results with:biome
  142. [142]
    Quantifying the impacts of rewilding on ecosystem resilience to ...
    Rewilding generally enhances resilience with nearly 70% of observations reporting positive outcomes, 10% neutral and 20% negative.
  143. [143]
    FAO Statistical Yearbook 2024 reveals critical insights on the ...
    Nov 18, 2024 · Global agricultural value has increased by 89 percent in real terms over the past two decades, reaching $3.8 trillion in 2022. Despite this ...
  144. [144]
    STATISTICAL YEARBOOK WORLD FOOD AND AGRICULTURE 2024
    Nov 15, 2024 · Economic dimensions of agriculture. The global value added generated by agriculture, forestry and fishinga grew by 89 percent in real terms ...
  145. [145]
    Used planet: A global history | PNAS
    Concentrating human populations within cities transforms the economies of scale in human interactions, producing higher average incomes and enabling a wide ...
  146. [146]
    Population trends and the transition to agriculture: Global processes ...
    Jan 17, 2023 · One of the most profound transitions in human existence was the development of agriculture. It fueled an increase in the global population ...
  147. [147]
    Landscapes that work for biodiversity and people - Science
    Oct 19, 2018 · We argue that, instead, working lands can be used to support high levels of biodiversity while satisfying human needs in a sustainable way.
  148. [148]
    Why reconnect to nature in times of crisis? Ecosystem contributions ...
    Oct 17, 2023 · Why reconnect to nature in times of crisis? Ecosystem contributions to the resilience and well-being of people going back to the land in Greece.
  149. [149]
    [PDF] Sustaining biodiversity and people in the world's anthropogenic ...
    Jul 25, 2013 · Human populations and their use of land have now transformed more than three quarters of the terrestrial biosphere into anthropogenic biomes ( ...
  150. [150]
    How Can Pollen Teach Us About Climate? | News
    May 27, 2016 · By analyzing pollen from well-dated sediment cores, paleoclimatologists can obtain records of changes in vegetation going back hundreds of thousands, and even ...Missing: tree rings<|separator|>
  151. [151]
    Reconciling pollen-stratigraphical and tree-ring evidence for high
    It is shown that the pollen and tree-ring based reconstructions exhibit similar temperature variability on centennial and longer scales.Missing: biome shifts
  152. [152]
    How tree rings tell time and climate history
    Nov 29, 2018 · Tree rings typically record wet or dry years, and in cooler areas (high latitudes or high elevation), the ring widths are often a proxy for temperature.
  153. [153]
    Milankovitch (Orbital) Cycles and Their Role in Earth's Climate
    Feb 27, 2020 · These cyclical orbital movements, which became known as the Milankovitch cycles, cause variations of up to 25 percent in the amount of incoming insolation at ...
  154. [154]
    Milankovitch Cycles, Paleoclimatic Change, and Hominin Evolution
    The three elements of Milankovitch cycles are eccentricity, obliquity, and precession (Figure 3). Eccentricity describes the degree of variation of the Earth's ...
  155. [155]
    Medieval Warm Period, Little Ice Age and 20th century temperature ...
    These include large temperature excursions during the Little Ice Age (∼1400–1900 AD) and the Medieval Warm Period (∼800–1300 AD) possibly related to changes in ...
  156. [156]
    Medieval Warming, Little Ice Age, and European impact on the ...
    Jan 20, 2017 · The presence of the Little Ice Age ensues with increases in Picea and Tsuga, coupled with increasing organic percentages due to cooler, moister ...Missing: biomes | Show results with:biomes
  157. [157]
    A 4500-Year Tree-Ring Record of Extreme Climatic Events ... - MDPI
    Mar 14, 2023 · Currently, a multi-millennial tree-ring chronology of more than 8500 years has been constructed on the Yamal Peninsula [16]. This chronology is ...
  158. [158]
    Long-term natural variability and 20th century climate change - NIH
    Observations suggest the warming of the 20th century global mean surface temperature has not been monotonic, even when smoothed by a 10–20 year low-pass filter.Missing: biome | Show results with:biome
  159. [159]
    Evidence of localised Amazon rainforest dieback in CMIP6 models
    Nov 24, 2022 · Five of seven CMIP6 models show localized Amazon dieback, especially in northern South America, with 7±5% dieback per degree of warming past 1. ...
  160. [160]
    Climate-Biome Envelope Shifts Create Enormous Challenges and ...
    Sep 21, 2020 · Climate change this century will affect the distribution and relative abundance of ecological communities against a mostly static background of protected land.
  161. [161]
    MODIS Shows Earth is Greener - NASA's Terra satellite
    Mar 7, 2019 · Earth's greening is mainly due to tree planting in China and intensive agriculture in China and India, with a 5% increase in leaf area.
  162. [162]
    CO2 fertilization of terrestrial photosynthesis inferred from site ... - NIH
    Mar 1, 2022 · We provide robust observationally inferred evidence that a strong CO 2 fertilization effect is detectable in globally distributed eddy covariance networks.
  163. [163]
    Recent global decline of CO2 fertilization effects on vegetation ...
    Global process-based models attribute part of the increasing land carbon sink (2) to the increase in vegetation productivity driven by the fertilization effect ...
  164. [164]
    Amazon dieback beyond the 21st century under high-emission ...
    Aug 20, 2025 · Among models, those with higher ECS tend to show earlier dieback onset, likely because they reach critical warming thresholds more quickly under ...
  165. [165]
    Pronounced loss of Amazon rainforest resilience since the early 2000s
    Mar 7, 2022 · Early studies showed that the Amazon rainforest may exhibit strong dieback by the end of the twenty-first century. Both pronounced drying in ...
  166. [166]
    Towards quantifying uncertainty in predictions of Amazon 'dieback'
    We find that the loss of Amazonian rainforest is robust across the climate uncertainty explored by perturbed physics simulations covering a wide range of global ...
  167. [167]
    Biome-scale temperature sensitivity of ecosystem respiration ...
    This discrepancy suggests that sparse plot-scale observations do not capture the spatial-scale dependence and biome specificity of the temperature sensitivity.
  168. [168]
    Do empirical observations support commonly-held climate change ...
    May 14, 2020 · Here, we propose a protocol to review the body of evidence for commonly-held climate change range shift hypotheses at the species level focusing ...
  169. [169]
    Divergent controls of soil organic carbon between observations and ...
    Jul 16, 2021 · Understanding the role of dominant SOC controls, and the discrepancies between models and observations, globally and across biomes, is essential ...
  170. [170]
    [PDF] Global patterns in the vulnerability of ecosystems to vegetation shifts ...
    We examine three sets of potential indicators of the vulnerabilit of ecosystems to biome change: (1) observed changes of 20th- century climate, (2) projected ...
  171. [171]
    Potential for evolutionary responses to climate change
    Results of 250 years of common garden experiments show that most forest trees have evolved local adaptation, as evidenced by the adaptive differentiation of ...Missing: biome | Show results with:biome
  172. [172]
    Evolutionary adaptation to climate change - PMC - NIH
    Feb 13, 2024 · Traits involved in climate change adaptation include not only classic phenomena, such as range shifts and environmentally dependent sex determination.
  173. [173]
    Ecosystems are showing symptoms of resilience loss - IOP Science
    May 27, 2022 · Up to 29% of global terrestrial ecosystem, and 24% marine ones, show symptoms of resilience loss. These symptoms are shown in all biomes, but ...
  174. [174]
    Dryland microbiomes reveal community adaptations to ...
    In these biomes, microorganisms provide vital ecosystem services and have evolved distinctive adaptation strategies to endure and flourish in the extreme.<|separator|>
  175. [175]
    Reduced resilience of terrestrial ecosystems locally is not reflected ...
    May 13, 2021 · On a local scale we find that climate change reduced the resilience of ecosystems in 64.5% of the global terrestrial vegetated area.
  176. [176]
    Emerging signals of declining forest resilience under climate change
    We show that tropical, arid and temperate forests are experiencing a significant decline in resilience, probably related to increased water limitations and ...
  177. [177]
    Measuring resilience and assessing vulnerability of terrestrial ...
    Mar 19, 2018 · Our results indicate that forests, the most productive and biodiverse terrestrial ecosystems on the earth, are more vulnerable to climate change ...
  178. [178]
    Need and vision for global medium-resolution Landsat and Sentinel ...
    Jan 1, 2024 · Similarly, Google Earth Engine has enabled a new community to process medium-resolution satellite data at continental to global scales.<|separator|>
  179. [179]
    Status for: Vegetation Indices (MOD13) - MODIS Land - NASA
    The normalized difference vegetation index (NDVI) accuracy is within ± 0.025, while that of the enhanced VI (EVI) is within ± 0.015, and the accuracy of ...
  180. [180]
    An evaluation of Landsat, Sentinel-2, Sentinel-1 and MODIS data for ...
    In this study, we evaluated the utility of Landsat (7&8), Sentinel-2 (A&B), Sentinel-1 (A&B) and the Moderate Resolution Imaging Spectroradiometer (MODIS) for ...Missing: biome review
  181. [181]
    [PDF] modis vegetation index (mod 13)
    Apr 30, 1999 · The MODIS vegetation index (VI) products will provide consistent, spatial and temporal comparisons of global vegetation conditions which will ...
  182. [182]
    [PDF] Vegetation Index Product Suite User Guide & Abridged Algorithm ...
    Three vegetation index algorithms are produced globally, the historical Normalized Difference. Vegetation Index (NDVI), one of the longest remote sensing based ...
  183. [183]
    Remote Sensing of Tundra Ecosystems using High Spectral ...
    Observing the environment in the vast inaccessible regions of Earth through remote sensing platforms provides the tools to measure ecological dynamics.
  184. [184]
    [PDF] The Harmonized Landsat and Sentinel-2 surface reflectance data set
    Oct 14, 2018 · The L30 and S30 show a good consistency with coarse spatial resolution products, in particular MODIS. Collection 6 MCD09CMG products (overall ...<|control11|><|separator|>
  185. [185]
    Satellite remote sensing can operationalise the IUCN Global ...
    Mar 14, 2025 · We developed an unsupervised, repeatable method to delineate biomes and their component functional ecosystems, based on landscape-level vegetation structure.
  186. [186]
    [PDF] View angle effects on relationships between MISR vegetation ...
    It is encouraging that we found a small improvement in LAI estimation with off-nadir NDVI or anisotropic indices, during five separate sampling periods, using ...
  187. [187]
    Proof of evidence of changes in global terrestrial biomes using ...
    Jul 28, 2023 · Plant dynamics are influenced by environmental factors such as temperature and rainfall [2,3]. Temperature is considered the leading factor [2], ...
  188. [188]
    Remotely-sensed productivity clusters capture global biodiversity ...
    Nov 2, 2018 · A Method of Two-Stage Clustering with Constraints Using Agglomerative Hierarchical Algorithm and One-Pass k-Means++. In Knowledge and ...Results · Methods · Dynamic Habitat Indices<|control11|><|separator|>
  189. [189]
    [PDF] Explainable Clustering Applied to the Definition of Terrestrial Biomes
    We apply k- means clustering to the segments, and introduce ex- plainability by approximating the clusters with a de- cision tree, which we subsequently trim ...
  190. [190]
    [PDF] Explainable Clustering Applied to the Definition of Terrestrial Biomes
    The KG tree model is more targeted to global biome distributions, and has been considered as the standard method for biome classification (Kottek et al., 2006).
  191. [191]
    Unsupervised learning determines global marine eco-provinces
    May 29, 2020 · An unsupervised learning method is presented for determining global marine ecological provinces (eco-provinces) from plankton community structure and nutrient ...
  192. [192]
    Detection and tracking of carbon biomes via integrated machine ...
    Mar 13, 2025 · We employ an unsupervised machine learning approach, namely the agglomerative hierarchical clustering (HC) technique (Müllner, 2011), to ...
  193. [193]
    Analysis of clustering methods for crop type mapping using satellite ...
    Jul 1, 2022 · This paper proposes to test and analyse the performance of several unsupervised clustering algorithms towards crop type identification on remote images.
  194. [194]
    Reef-Insight: A Framework for Reef Habitat Mapping with Clustering ...
    Jun 30, 2023 · Our results indicate that clustering methods using remote sensing data can well identify benthic and geomorphic clusters in reefs when compared ...
  195. [195]
    Parsimonious machine learning for the global mapping of ...
    Mar 19, 2025 · By capturing empirical relationships between AGBD and environmental factors or satellite imagery, ML models present a computationally efficient ...
  196. [196]
    Quantitative comparison and review of biome‐mapping methods
    Jun 20, 2024 · In this example, we quantify that the strongest disagreement among biome maps generally occurs in landscapes altered by human activities and ...Missing: empirical | Show results with:empirical
  197. [197]
    Forgotten forests - issues and prospects in biome mapping using ...
    Nov 24, 2011 · The poor performance of the maps can be attributed to two main factors: (1) poor spatial resolution, and (2) poor biome delimitation. Poor ...
  198. [198]
    Challenges and Limitations of Remote Sensing Applications in ...
    However, challenges arise from issues like the cost of high-resolution data, coverage limitations, and inadequate field validation data in remote areas. This ...
  199. [199]
    Global land cover maps do not reveal mining pressures to biodiversity
    Jul 1, 2025 · We reveal that more than half (56–77%) the global land area disturbed by mining is classified by land cover maps as 'natural'.
  200. [200]
    Accuracies, discrepancies, and challenges of the 10 m global land ...
    Our statistical assessment showed that ESRI demonstrated the highest overall accuracy (68.5 ± 0.42%) at the global mountain level, with a small lead over ESA ( ...
  201. [201]
    Global Land Cover Maps' Accuracy and Applications Explained
    Nov 26, 2024 · This article examines the accuracies for four global land cover data sets and explains their strengths and weaknesses and how best to apply ...
  202. [202]
    Editorial: Revisiting the Biome Concept With A Functional Lens
    Apr 30, 2019 · Early biogeographers such as Alexander von Humboldt recognized the broad-scale coupling of vegetation and climate (von Humboldt, 1806). This ...Missing: proponents | Show results with:proponents
  203. [203]
    From static biogeographical model to dynamic global vegetation ...
    Nov 25, 2000 · This paper reviews and compares two major types of static biogeographical models, climate–vegetation classification and plant functional type models.
  204. [204]
    A global perspective on modelling vegetation dynamics | Request PDF
    Aug 6, 2025 · This paper reviews and compares two major types of static biogeographical models, climate–vegetation classification and plant functional type ...
  205. [205]
    Ecotones in vegetation ecology: methodologies and definitions ...
    Feb 28, 2009 · Ecotones can be found in many forms and across a range of scales, from a few centimeters to several kilometers. They are boundaries between ...
  206. [206]
    Reducing the arbitrary: fuzzy detection of microbial ecotones and ...
    Aug 13, 2020 · We show the suitability of applying fuzzy clustering to address the patchiness of microbial ecosystems, integrating environmental (Sea Surface Temperature, ...
  207. [207]
    Probabilistic description of vegetation ecotones using remote sensing
    Ecotone transitions between vegetation types are of interest for understanding regional diversity, ecological processes and biogeographical patterns.
  208. [208]
    Palaeoecological evidence of state shifts between forest and ...
    Recent disturbances have induced state shifts in many ecosystems. It has been proposed that human actions degrade system resilience, thus allowing state ...
  209. [209]
    Biome shifts and transitions | World Biogeography Class Notes
    Paleoecological records reveal long-term biome shifts across geological time scales ... Case studies provide evidence for the reality and complexity of ecosystem ...
  210. [210]
    Biome Awareness Disparity is BAD for tropical ecosystem ...
    Oct 15, 2021 · Biome Awareness Disparity (BAD) is a failure to appreciate all biomes, leading to tree-centric restoration and threatening open biomes.
  211. [211]
    The underestimated biodiversity of tropical grassy biomes - PMC
    1. Introduction. The Earth's tropical landscapes are dominated by two strongly contrasting biomes: savannahs and grasslands on the one hand and closed-canopy ...
  212. [212]
    [PDF] Biome Awareness Disparity is BAD for tropical ecosystem ...
    Aug 1, 2022 · Biome Awareness Disparity (BAD) is a failure to appreciate the significance of all biomes in conservation and restoration policy.
  213. [213]
    The underestimated global importance of plant belowground coarse ...
    Open biomes such as grasslands, savannas, shrublands are associated with many global biodiversity hotspots, and cover ∼60% of land globally. Yet, extensive and ...
  214. [214]
    Out of the shadows: ecology of open ecosystems
    Major disparities, where open low biomass systems occurred in the same climate zone as closed forests have been dismissed as products of deforestation. Many of ...
  215. [215]
    Forest-biased terminology does not help to include open ...
    So why is the term 'deforestation' used when discussing the loss or suppression of other types of native vegetation such as grasslands or savannas? Even if ...Missing: biome prioritization
  216. [216]
    Geographical and taxonomic biases in research on biodiversity in ...
    Dec 27, 2012 · Biome type was a clear factor in research bias, and forest biomes were the subject of 87% of papers, while species richness was not generally ...Results · Geographical Bias · Discussion
  217. [217]
    Amazonian landscapes and the bias in field studies of forest ... - PNAS
    We uncovered substantial mean biases (9–98%) in forest canopy structure (height, gaps, and layers) and aboveground biomass in both lowland Amazonian and montane ...Amazonian Landscapes And The... · Results · Methods
  218. [218]
    [PDF] Longleaf pine savannas reveal biases in current understanding of ...
    Aug 8, 2023 · Longleaf pine savannas are overlooked, have a distinct climate, and are considered "Forgotten Grasslands" with a unique grassy understory, ...
  219. [219]
    Legacy of the Last Glacial on the present‐day distribution of ...
    Nov 5, 2019 · I propose that these deciduous and evergreen boreal forests represent alternative quasi-stable states, triggered by their different northern tree refugia.
  220. [220]
    Evolution of global vegetation patterns since the last glacial maximum
    May 15, 2025 · These studies offer valuable insights into vegetation dynamics under different global climate scenarios, covering spatial scales from national ...
  221. [221]
    Biome boundary maintained by intense belowground resource ...
    Feb 14, 2022 · We demonstrate an apparent general mechanism in which local competition triggers a biome-scale feedback between plant traits and soil resources ...
  222. [222]
    [PDF] Feedbacks in ecology and evolution
    Feb 4, 2022 · The two negative feedbacks stabilize either a closed or an open biome in a given climate (bistability) and can remain as such for millennia [34] ...<|separator|>
  223. [223]
    The ecological legacy of late Pleistocene megafauna extinctions
    We demonstrate that the late Pleistocene loss of megafauna was pervasive and left legacies detectable within the modern atmosphere, geosphere, hydrosphere and ...
  224. [224]
    Glacial legacies on interglacial vegetation at the Pliocene ... - Nature
    Jun 24, 2016 · We show that interglacial vegetation during the Plio-Pleistocene transition mainly reflects conditions of the preceding glacial instead of contemporary ...<|separator|>
  225. [225]
    The Migrating Boreal Forest - NASA Earth Observatory
    Aug 20, 2002 · At the height of the Wisconsin glaciation during the Pleistocene Ice Age, the Laurentide ice sheet covered nearly half of North America (left).