Fact-checked by Grok 2 weeks ago

Biological interaction

Biological interactions, also known as biotic interactions or species interactions, refer to the relationships among organisms of different species that coexist in the same habitat or ecosystem, influencing each other's survival, reproduction, and distribution through direct or indirect effects that can be positive, negative, or neutral. These interactions are a core focus of community ecology, shaping the structure and dynamics of biological communities by determining how species coexist, compete, or cooperate. The primary types of biological interactions are classified based on their effects on the interacting species, often denoted using a sign convention where "+" indicates benefit, "−" indicates harm, and "0" indicates no effect. occurs when two species vie for limited resources, resulting in negative impacts on both (−/−), as seen in cases where one species outcompetes another for food or space, potentially leading to the where only one species can occupy a specific niche. Predation and involve one species consuming another for sustenance, benefiting the predator or herbivore (+) while harming the prey or plant (−), such as sea otters preying on sea urchins to control their populations. benefits one species (+) while harming the other (−), such as ticks feeding on the blood of mammals. Mutualism provides mutual benefits to both species (+/+), exemplified by pollinators like hummingbirds aiding while gaining . benefits one species (+) without affecting the other (0), as in attaching to whales for without impacting the host. Less commonly emphasized but notable is amensalism, where one species is harmed (−) by another that remains unaffected (0), such as through the release of allelopathic chemicals by inhibiting nearby growth. Beyond pairwise classifications, biological interactions often occur in , including indirect effects through food webs where changes in one interaction cascade across multiple , influencing stability and resilience. For instance, the reintroduction of wolves in demonstrated how predator-prey dynamics can indirectly promote vegetation recovery by reducing herbivore overgrazing, thereby benefiting diverse community members. These interactions drive evolutionary processes, such as in mutualistic pairs, and play critical roles in maintaining , regulating population sizes, and modulating functions like nutrient cycling and primary productivity. In the context of global environmental change, understanding biological interactions is essential for predicting how assemblages respond to stressors like habitat loss or climate shifts.

Introduction

Definition

Biological interactions encompass any process in which one biological entity influences the state, activity, function, or of another distinct biological entity. These processes span multiple scales of , from molecular-level events such as the of enzymes to substrates, which alter molecular conformations and catalyze reactions, to organismal-level associations in ecosystems where affect each other's distribution and abundance. At their core, such interactions are dynamic and energy-dependent, driving the complexity of by enabling coordination, , and across hierarchical levels from cells to communities. A primary distinction lies between direct and indirect interactions. Direct interactions occur through immediate physical, chemical, or physiological contact between entities, resulting in an unmediated effect on the recipient's , , or . Indirect interactions, by contrast, are mediated by one or more entities or environmental factors, propagating effects through chains of without direct contact. This dichotomy applies universally across scales, from regulatory networks where transcription factors indirectly modulate distant via signaling cascades, to ecological dynamics like apparent between prey mediated by a shared predator. Interactions can further be classified as obligatory or facultative based on their for . Obligatory interactions require the involvement of both (or at least one) entities for , , or normal function, as seen in certain symbiotic molecular complexes where leads to functional failure. Facultative interactions, however, confer benefits such as enhanced or but allow entities to function independently under suitable conditions. This classification highlights the spectrum of dependency in biological systems, excluding intra-entity processes like within a single or , which do not involve distinct external influencers. For example, in , predator-prey dynamics illustrate a direct interaction affecting population levels.

Importance

Biological interactions play a pivotal role in by driving and through mechanisms such as , predation, and , which impose selective pressures that shape and over time. For instance, interactions can alter evolutionary responses to environmental changes, facilitating the of lineages and the formation of new even in isolated populations. These processes highlight how interactions mediate effects, influencing and across generations. In ecological contexts, biological interactions are essential for maintaining by structuring communities and stabilizing populations through interdependent relationships that prevent dominance by any single species. They underpin key services, including , where interspecific exchanges—such as by microbes and uptake by —recycle essential elements like and , supporting productivity and . Overall, diverse interactions enhance stability, enabling services like and water regulation that sustain global . The practical applications of understanding biological interactions span multiple fields, informing strategies in , , and . In , targeting molecular and cellular interactions within protein networks has revolutionized , allowing precise modulation of disease pathways through network-based approaches. In , leveraging predator-prey interactions enables , where natural enemies suppress pest populations and account for 50–90% of pest regulation in crop fields, reducing reliance on chemical pesticides. For , recognizing interactions guides efforts to protect trophic networks, ensuring the persistence of and associated services amid environmental threats.

Sub-organismal Interactions

Molecular Interactions

Molecular interactions form the foundational level of biological associations, where biomolecules such as proteins, nucleic acids, and small molecules engage through non-covalent forces including bonds, van der Waals interactions, electrostatic forces, and hydrophobic effects. These interactions enable precise and functional within cells, underpinning processes like and enzymatic catalysis. At this scale, interactions are typically transient and reversible, governed by thermodynamic principles that determine stability and specificity. Key types of molecular interactions include -receptor binding, enzyme-substrate interactions, and protein-protein interactions (PPIs). In -receptor binding, a molecule such as a or binds to a specific receptor protein, often initiating conformational changes that propagate signals. Enzyme-substrate interactions involve the precise of a substrate into an enzyme's , facilitating chemical transformation through stabilization of the . PPIs, meanwhile, allow proteins to form complexes that coordinate multi-step reactions or structural assemblies, with high specificity arising from complementary surface topologies. Central to these interactions are concepts like binding affinity, specificity, and allostery. Binding affinity quantifies the strength of association, commonly expressed by the K_d, defined as K_d = \frac{[A][B]}{[AB]}, where [A] and [B] are the equilibrium concentrations of the free binding partners and [AB] is the complex; lower values indicate higher binding affinity under conditions. Specificity ensures selective recognition of particular partners, driven by structural complementarity and energetic discrimination against non-cognate s. Allostery refers to regulation where binding of a at one site modulates affinity at a distant site, as described in the concerted model where proteins exist in between tense (T) and relaxed () states. Representative examples illustrate these principles. In transcription, DNA-protein interactions occur when transcription factors bind specific promoter sequences via or zinc-finger motifs, recruiting to initiate . Antibody-antigen exemplifies immune , where the variable regions of antibodies form complementary interfaces with epitopes on pathogens, achieving affinities often in the nanomolar range to facilitate neutralization. Techniques for detecting molecular interactions include the yeast two-hybrid system and co-immunoprecipitation. The yeast two-hybrid system, introduced in , fuses one protein to a and another to a transcription activation domain; interaction reconstitutes transcriptional activity, enabling of PPIs in yeast cells. Co-immunoprecipitation isolates protein complexes from cell lysates using an antibody against one partner, pulling down associated molecules for identification via or , confirming interactions in native contexts. These methods have revealed extensive interactomes, such as those involving signaling proteins.

Cellular Interactions

Cellular interactions encompass the dynamic processes through which cells adhere, communicate, and respond to one another, emerging from molecular foundations such as receptor-ligand engagements to orchestrate collective behaviors in s and microbial communities. These interactions are pivotal for maintaining cellular organization and enabling responses to environmental cues, with disruptions often leading to pathological states. At the core, cell-cell adhesion molecules like cadherins facilitate direct physical connections between cells, promoting tissue stability and through calcium-dependent homophilic binding. Key types of cellular interactions include adhesion mechanisms, signaling pathways, and density-dependent communication systems. Cadherins, for instance, form adherens junctions that not only anchor cells but also initiate intracellular signaling to regulate cytoskeletal dynamics and . The (MAPK) cascade exemplifies signaling pathways, where extracellular stimuli activate a sequential relay—from receptor tyrosine to MAP kinase kinases (MAP2Ks) and MAPKs—culminating in nuclear modulation for changes. In , enables population-level coordination via autoinducer molecules like acyl-homoserine lactones, which accumulate to threshold levels and trigger communal for processes such as production. in these interactions typically proceeds through three phases: reception by surface receptors, amplification via second messengers or cascades, and response through effector activation, ensuring precise and amplified signal propagation. Emergent cellular behaviors from these interactions include and fusion events critical for development. can be induced by intercellular signals, such as binding to death receptors on target cells, activating cascades that dismantle the cell in a controlled manner to prevent . Cell fusion, observed in processes like myoblast merger during skeletal muscle formation, relies on fusogenic proteins that destabilize membranes and promote hemifusion intermediates, integrating cytoplasms for multinucleated syncytia. In the , T-cell activation exemplifies cooperative interactions, where antigen-presenting cells engage T-cell receptors via major histocompatibility complex-peptide complexes, co-stimulated by CD28-B7 ligation to initiate IL-2 production and . Similarly, microbial biofilms arise from quorum sensing-driven signaling, where bacterial cells aggregate via adhesins and production, enhancing resistance to antibiotics and host defenses. Dysregulation of cellular interactions underlies diseases like cancer, where aberrant signaling perpetuates uncontrolled growth. For example, oncogenic mutations in the MAPK pathway, such as BRAF V600E, lead to constitutive activation, evading and promoting through enhanced and defects in function. These insights highlight the therapeutic potential of targeting interaction interfaces, such as inhibitors to disrupt biofilms in infections.

History of Organismal Interactions

Early Concepts

Early observations of biological interactions trace back to ancient naturalists, who documented predator-prey dynamics and emerging mutualisms through descriptive accounts rather than formal theories. , in his Historia Animalium (circa 350 BCE), noted various animal predation patterns, such as like eagles and hawks hunting smaller animals for sustenance, and herbivores like sheep on specific plants while avoiding toxic ones, illustrating early recognition of trophic dependencies. His pupil extended these in Historia Plantarum (circa 300 BCE), describing plant-animal mutualisms, including the role of in fig via caprification—where wasps transfer between male and female fig trees—and manual pollination of date palms, highlighting interdependent reproduction. These 4th-century BCE records, preserved in herbalist and philosophical texts, laid anecdotal foundations for understanding organismal interrelations without mechanistic explanations. In the 18th century, naturalists shifted toward systematic documentation of interactions within broader natural economies. , in his 1749 essay "The Oeconomy of Nature," described symbiotic associations such as birds dispersing plant seeds by consuming fruits—like thrushes aiding propagation—framing these as balanced contributions to nature's harmony, though he did not coin the term "." , during his 1799–1804 South American expeditions, advanced perspectives by observing interconnected competitions in diverse flora-fauna networks, emphasizing how interactions influence environmental balance and human alterations disrupt it. Gilbert White's 1789 The Natural History and Antiquities of Selborne provided detailed local accounts, including birds like and flycatchers preying on , and seasonal insect swarms affecting avian foraging, portraying interactions as integral to parish . The saw further integration of interactions into evolutionary theory. Charles Darwin's 1859 described how competition for resources, predation, and mutualistic relationships drive , with examples like orchids and their pollinators illustrating coevolutionary dependencies. Darwin's work built on earlier observations, emphasizing interactions as mechanisms shaping and . This era also marked conceptual transitions from teleological interpretations—viewing interactions as divinely purposed—to more mechanistic ones grounded in empirical limits. Thomas Malthus's 1798 An Essay on the Principle of Population exemplified this by arguing that populations grow geometrically while resources increase arithmetically, leading to natural checks like and among organisms for sustenance, as seen in animal herds limited by food scarcity. Influenced by Enlightenment , figures like in his 1790 critiqued overt , suggesting apparent purposes in biology arise from organized complexity rather than final causes, paving the way for later formalized models. These pre-1900 insights, rooted in observation, established interactions as dynamic processes shaped by environmental constraints.

Modern Developments

In the early , mathematical modeling advanced the quantitative understanding of organismal interactions, particularly through the Lotka-Volterra equations developed independently by in 1925 and in 1926, which described oscillatory predator-prey dynamics based on differential equations capturing population growth and decline. This framework shifted ecological studies from descriptive accounts to predictive models, enabling simulations of interaction stability and cycles. Complementing these efforts, introduced the concept in 1935, defining it as a system of and abiotic components where organismal interactions, such as nutrient cycling and energy transfer, maintain holistic function. Mid-20th-century developments integrated energy dynamics and evolutionary perspectives into interaction studies. Eugene P. Odum's 1953 textbook Fundamentals of Ecology formalized energy flow models for ecosystems, emphasizing how mutualistic interactions, like and , facilitate unidirectional energy transfer from producers to consumers while recycling matter. Building on this, and Peter H. Raven's 1964 paper on butterfly-plant relationships proposed as a driver of reciprocal adaptations in mutualistic and antagonistic interactions, illustrating how selective pressures from one species shape another's traits over generations. From the late into the 21st, network ecology emerged as a key approach, with analyses of webs revealing structural patterns like low connectance that underpin interaction stability, followed by early 2000s studies identifying scale-free topologies in ecological networks. Concurrently, post-2000 microbiome research, spearheaded by the Human Microbiome Project launched in 2007, uncovered extensive hidden mutualisms between human-associated microbes and host cells, such as gut aiding and immune modulation, transforming views of from pairwise to community-level phenomena. Recent advancements in the and have leveraged genomic and computational tools to dissect and forecast interaction dynamics. CRISPR-Cas9 editing, widely adopted since 2012, has enabled targeted disruption of genes involved in symbiotic interactions, such as those mediating legume-rhizobium , providing causal insights into mutualistic specificity. In parallel, 2020s climate models incorporating species interaction networks predict widespread shifts, including disrupted mutualisms like pollinator-plant mismatches and intensified competitions due to altered phenologies and range overlaps under warming scenarios.

Classifications of Organismal Interactions

Duration-based Classification

Biological interactions can be classified based on their into short-term and long-term categories, providing a framework to understand their temporal persistence and ecological implications. Short-term interactions, also known as ephemeral or transient interactions, are characterized by brief s, typically spanning hours, days, or a single event, without ongoing association between the organisms involved. These interactions often involve minimal or no physical contact beyond the immediate exchange, such as a predator capturing and consuming prey in one encounter. In contrast, long-term interactions, frequently referred to as symbioses, persist over extended periods, including the lifespan of individuals or multiple generations, fostering prolonged physical or physiological intimacy between partners. Examples include vertically transmitted endosymbionts in , where are inherited across generations and provide essential nutrients, maintaining association for millions of years. The primary criteria for this classification revolve around the time frame of the interaction, which can range from minutes to evolutionary timescales spanning generations, and the duration of intimacy, assessed by the extent of sustained physical contact or metabolic integration. Transmission mode further refines this: horizontal transmission often aligns with short-term interactions reformed each generation through environmental acquisition, while vertical transmission supports long-term persistence via direct inheritance from parent to offspring. This duration-based approach offers advantages in predicting interaction stability and evolutionary trajectories, as long-term associations typically promote and genome streamlining in symbionts due to genetic bottlenecks, enhancing mutual dependency and . However, it has limitations in hybrid or facultative cases, such as that can shift from short-term opportunistic encounters to prolonged infections based on host availability, blurring categorical boundaries and requiring contextual evaluation. The origins of duration-based classification trace back to 1970s research in symbiosis literature, where studies on endophytic mutualisms began emphasizing temporal persistence to distinguish casual from obligatory relationships. For instance, predation exemplifies a predominantly short-term interaction.

Fitness-based Classification

Biological interactions are often classified based on their effects on the fitness of the interacting organisms, using a simple sign convention where "+" indicates a positive effect (increase in fitness), "−" indicates a negative effect (decrease in fitness), and "0" indicates no effect (neutral). This framework, originally proposed to standardize the categorization of pairwise interactions, distinguishes six main types by combining the effects on each participant. The classification is summarized in the following table:
Interaction TypeEffect on Species 1Effect on Species 2
++
+0
Predation/Parasitism+
Amensalism0
Neutralism00
This table represents direct, pairwise effects under typical conditions, though real-world outcomes can vary. Key concepts in this classification include the distinction between direct and indirect fitness effects. Direct effects stem from immediate interactions between two , such as resource sharing or harm, while indirect effects arise through chains involving other species, like apparent via a shared predator. Additionally, interaction outcomes exhibit strong context-dependency, where the net effect can shift based on environmental conditions; for instance, between may become mutualistic under as neighbors facilitate resource access. Fitness effects are typically measured through changes in demographic rates, such as , , or , which collectively determine an organism's lifetime reproductive output. Experimental designs like removal studies are commonly used to isolate effects; by removing one and observing fitness changes in the other, researchers quantify the sign and magnitude of the . For example, removal experiments with coexisting wood warbler demonstrated negative fitness effects from interspecific competition on nesting success and . Despite its utility, the +/- has faced criticisms for oversimplifying complex ecological dynamics. It primarily addresses pairwise interactions, neglecting diffuse effects where multiple collectively influence , as in Pianka's concept of diffuse competition among desert lizards sharing resources. Post-2000 critiques highlight its limitations in capturing network-level phenomena, such as higher-order interactions involving three or more , which can alter and coexistence in ways not predicted by binary signs alone.

Trophic-based Classification

Trophic interactions in refer to relationships between organisms that involve the of , facilitating transfer along food chains or webs, such as predation, herbivory, or . In contrast, non-trophic interactions encompass direct effects between species that do not involve feeding, including for resources or without , territorial disputes, or facilitation without . These distinctions highlight how trophic links center on nutritional dependencies, while non-trophic ones focus on abiotic or behavioral influences that shape coexistence. The primary criterion for classifying interactions as trophic is the direct involvement of , where one ingests another, altering through energy flow. Chain length in food webs, defined by the number of sequential trophic levels from producers to top predators, further characterizes these interactions, with longer chains indicating more complex energy pathways. Non-trophic interactions, lacking this , are evaluated based on their indirect mediation through environmental modifications or non-lethal . Trophic interactions are significant for driving flow and in , often propagating effects across multiple levels as seen in trophic cascades. Non-trophic interactions, meanwhile, primarily influence community structure by modulating and without altering budgets. In evolutionary contexts, trophic interactions frequently promote through co-evolutionary pressures, leading to refined predator-prey adaptations. Studies from the 1980s, such as Paine's analysis of linkages, demonstrated how varying interaction strengths in trophic networks foster such evolutionary refinements and stability.

Types of Organismal Interactions

Predation

Predation is a biological interaction in which one , the predator, kills and consumes another , the prey, typically resulting in immediate death of the prey and a net benefit to the predator through energy acquisition. This process is directional and antagonistic, with the predator gaining advantages such as increased and , while the prey experiences a decrement due to mortality. Unlike prolonged exploitative interactions, predation is generally short-term and lethal, often involving active where the predator locates, captures, and consumes the prey in a single encounter. Predators often employ strategies guided by , which posits that they select prey types and habitats to maximize net energy intake relative to the costs of searching, pursuing, and handling. This theory, originally developed to explain resource use in patchy environments, suggests that predators prioritize high-profit prey when abundant and broaden their diet as profitability declines, thereby optimizing efficiency in variable ecological conditions. For instance, a predator might ignore low-energy prey if more rewarding options are available nearby, balancing risks like injury or time expenditure against nutritional rewards. Classic examples of predation include gray wolves (Canis lupus) hunting (Odocoileus virginianus) in North American forests, where packs coordinate to chase and subdue prey; African lions (Panthera leo) ambushing zebras (Equus quagga) on savannas through stealth and group tactics; and great white sharks ( carcharias) striking or fish in marine environments using speed and surprise. These interactions highlight diverse predatory tactics, from pursuit in open habitats to ambush in concealed settings, all aimed at overcoming prey defenses. Predator-prey dynamics frequently produce cyclical fluctuations in population sizes, where increases in prey abundance support predator growth, followed by prey declines that eventually reduce predator numbers, allowing prey recovery. Such cycles, observed in systems like snowshoe hares and Canadian lynx, arise from the time-lagged responses of predators to prey density changes, preventing either population from reaching . In response, prey evolve anti-predator adaptations, including morphological traits like to blend with backgrounds and reduce detection, or behavioral and physiological enhancements such as burst speed for evasion during chases. These adaptations, shaped by coevolutionary pressures, can significantly lower predation risk, with camouflaged prey often experiencing up to 50% fewer attacks in visual predator systems. In human contexts, predation principles underpin biological control efforts, where predators are introduced to manage pest populations. For example, ladybird beetles (Coccinellidae), such as the seven-spot ladybird (Coccinella septempunctata), are deployed against aphid infestations in agriculture; a single adult can consume over 50 aphids per day, leading to reductions exceeding 50% in aphid densities within greenhouses when released at rates of 5-10 individuals per square meter. This approach minimizes chemical pesticide use, promoting sustainable pest management while leveraging natural predatory efficiency.

Mutualism

Mutualism refers to a symbiotic interaction between two or more in which each participant derives a net benefit, often through exchanges that enhance , , or . These benefits can include access to nutrients, from predators or environmental stressors, or improved dispersal mechanisms, fostering interdependence that contributes to and . Unlike other interactions, mutualism requires ongoing reciprocity to persist, as exploitation by one partner can destabilize the relationship. Mutualisms are categorized as , where at least one species is entirely dependent on the partner for —such as certain plants unable to acquire essential nutrients independently—or facultative, where benefits accrue but independent survival remains possible for both. Trophic mutualisms involve the direct exchange of nutritional resources or energy between partners, optimizing resource acquisition in complementary ways. A representative example is pollination syndromes, where pollinators like bees obtain nectar and pollen as food rewards while inadvertently transferring pollen to stigmas, enabling plant fertilization and seed production; this interaction supports over 80% of flowering plants worldwide. Defensive mutualisms, by contrast, center on protection services, with one partner deterring threats to the other. In the Acacia-ant system, species of Acacia trees provide specialized domatia (hollow thorns for nesting) and extrafloral nectar or protein-rich Beltian bodies as food, while Pseudomyrmex ants aggressively patrol and remove herbivores, competing vegetation, and even encroaching ant colonies, reducing herbivory damage by up to 90% in some savanna ecosystems. Prominent examples illustrate the ubiquity and specificity of mutualisms across taxa. Mycorrhizal associations between and Glomeromycota fungi exemplify trophic mutualism, with fungi forming extensive hyphal networks that enhance plant uptake of and —up to 90% of a 's needs in nutrient-poor environments—in exchange for 20-30% of the plant's photosynthetically fixed carbon; this occurs in over 80% of vascular and is essential for productivity. In the human gut , facultative mutualism arises between the host and diverse bacterial communities, where microbes ferment indigestible fibers to produce short-chain fatty acids and vitamins (e.g., and ), bolstering host energy harvest, immune regulation, and pathogen resistance, while the host supplies a niche and undigested substrates. Mutualistic stability hinges on mitigating , where exploiters (e.g., non-pollinating nectar thieves or ineffective symbionts) gain benefits without reciprocating, potentially leading to interaction collapse. Enforcement mechanisms promote by favoring cooperative partners: partner choice allows hosts to preferentially associate with high-quality mutualists, as seen in yucca plants selecting pollinators via floral traits, while sanctions impose fitness costs on cheaters, such as reduced resource allocation to underperforming rhizobial bacteria in legume roots. These mechanisms, evolved through , ensure reciprocity and prevent , though their efficacy varies with partner density and environmental conditions. Mutualisms can span short-term encounters, like single events, to lifelong associations.

Commensalism

Commensalism represents a symbiotic in which one derives a benefit while the other experiences no net effect, often denoted as +/0 in fitness-based classifications of organismal interactions. This relationship is typically opportunistic, emerging from incidental associations where the benefiting exploits resources or structures provided by without imposing costs. Unlike more reciprocal symbioses, is generally non-obligatory and can persist long-term if environmental conditions remain stable, though it may dissolve if the association becomes disadvantageous for either party.30044-1) Classic examples illustrate these dynamics in marine and terrestrial ecosystems. attaching to the skin of exemplify phoretic commensalism, where the gain mobility, dispersal to nutrient-rich feeding grounds, and protection from predators through the whale's movement, while the whale incurs no detectable harm as the barnacles do not feed on its tissues or impede locomotion. Similarly, epiphytic plants such as orchids and bromeliads grow on branches in tropical forests, benefiting from elevated access to , , and air circulation without extracting nutrients from the host tree or altering its growth. These interactions highlight how commensals often utilize the host's physical for support or transport. Detecting true commensalism poses significant challenges due to the difficulty in empirically verifying neutrality, as subtle or context-dependent effects on the host may go undetected in field studies. Relationships initially classified as commensal can shift toward mutualism under changing conditions, such as when the commensal provides incidental protection against herbivores, complicating long-term assessments. Advanced experimental designs, including controlled manipulations of interaction intensity, are often required to distinguish neutrality from minimal benefits or costs. Some researchers even question the existence of purely neutral interactions, arguing that close associations invariably involve some degree of influence.00536-5.pdf) In ecological communities, plays a key role by enhancing habitat complexity, as commensals like epiphytes create microhabitats that support additional , including invertebrates, birds, and microbes, without disrupting the primary host dynamics. This structural augmentation fosters niche diversification and contributes to overall resilience, particularly in diverse environments like rainforests and coral reefs where layered interactions amplify resource partitioning.

Parasitism

Parasitism is a biological interaction in which one organism, the parasite, exploits another, the host, by deriving nutrients or resources from it, typically resulting in harm to the host's fitness while allowing the host to survive, albeit weakened. Unlike short-term encounters, parasitism often involves prolonged associations where the parasite resides on or within the host, extracting sustenance without immediately causing death. This exploitation can lead to reduced host growth, reproduction, or survival, and in some cases, chronic disease. Parasites are classified into several types based on their location and nature. Ectoparasites live on the external surface of , feeding on , , or secretions, as seen in fleas and ticks that attach to mammals. , in contrast, inhabit the internal tissues or organs of the host, such as tapeworms in the intestines of vertebrates, where they absorb digested food. Microparasites, including viruses, , and , are typically microscopic and multiply rapidly within host cells or fluids, often eliciting strong immune reactions. Representative examples illustrate the diversity of parasitic strategies. The parasite Plasmodium infects human red blood cells and liver cells, using mosquitoes as vectors for transmission, causing fever and in hosts while completing its . In brood parasitism, avian species like the (Cuculus canorus) lay eggs in the nests of other birds, such as reed warblers, tricking hosts into incubating and feeding the parasitic young at the expense of their own offspring. Hosts have evolved defenses to counter , primarily through immune responses that aim to detect and eliminate invaders. Innate immunity involves rapid, non-specific mechanisms like by macrophages, while adaptive immunity deploys antibodies and T-cells to target specific parasites, such as IgE-mediated responses against helminths. , or the degree of harm inflicted, evolves under the trade-off hypothesis, where higher virulence may enhance transmission but reduces host longevity, favoring balanced strategies over time, as observed in the attenuation of in rabbits.

Competition

Competition in biology refers to a negative between organisms of the same or different that arises from their simultaneous demand for a limited environmental , resulting in reduced for both parties. This is symmetric in its harm, as both competitors experience decreased , , or due to resource scarcity or direct . Competition plays a crucial role in structuring communities by influencing distributions and abundances, often leading to evolutionary adaptations that minimize overlap in resource use. Competition manifests in two primary types: exploitative and interference. Exploitative competition occurs indirectly when organisms deplete shared resources, such as nutrients or space, thereby limiting availability for others; for instance, forest plants compete for , where taller individuals shade shorter ones, reducing the latter's photosynthetic capacity. In contrast, interference competition involves direct behavioral interactions, such as or territorial defense, that prevent access to resources; examples include engaging in fights to claim nesting sites or mammals marking territories to exclude rivals. These types can coexist within the same system, with exploitative effects dominating at low densities and interference becoming prominent as populations grow. Competition is frequently trophic, involving resources like food within food webs. Illustrative examples highlight competition's ecological impacts. On the , , such as Geospiza fortis and Geospiza scandens, compete intensely for seeds during droughts, where medium ground finches with larger beaks survive better by cracking harder seeds, leading to shifts in beak size and to reduce overlap. Similarly, invasive species often outcompete natives through superior resource acquisition; for example, invasive plants like cheatgrass () in North American grasslands rapidly deplete soil moisture and nutrients, displacing native bunchgrasses and altering community composition. These cases demonstrate how competition drives selection and invasion success. Key outcomes of prolonged competition include the and niche partitioning. The , articulated by Gause in 1934, posits that two species occupying identical niches cannot coexist indefinitely; one will eventually dominate and exclude the other due to superior resource use efficiency, as demonstrated in laboratory experiments with paramecia. To avoid exclusion, species often undergo niche partitioning, evolving differences in resource utilization—such as times or habitats—to coexist stably. Competition intensity is quantified using resource overlap indices, such as Pianka's niche overlap index, which measures similarity in resource use between species on a scale from 0 (no overlap) to 1 (complete overlap), calculated as O_{jk} = \frac{\sum p_{ij} p_{ik}}{\sqrt{\sum p_{ij}^2 \sum p_{ik}^2}}, where p_{ij} is the proportion of resource i used by species j. High overlap values indicate strong potential for competition.

Amensalism

Amensalism is an interspecific biological interaction in which one harms or inhibits another without experiencing any cost or benefit to itself, denoted in -based as a (−, 0) relationship where the affected suffers reduced while the actor remains neutral. This asymmetry often arises from incidental byproducts of the actor's activities, such as the release of chemical compounds or physical disturbances, rather than deliberate targeting. Unlike , amensalism imposes no reciprocal cost on the actor, distinguishing it as a one-sided negative interaction that can occur across diverse taxa including microbes, , and animals. Classic examples illustrate these characteristics through chemical mechanisms known as or . In microbial systems, the fungus secretes penicillin, a that inhibits bacterial growth, such as in , while the fungus itself remains unaffected by the antibiotic's presence. Among plants, () exemplifies by releasing volatile compounds like into the soil and air, which suppress seed germination and growth of neighboring , such as grasses or forbs, without impacting the sagebrush's own establishment. These interactions highlight how amensalism can manifest as an unintended consequence of resource acquisition or defense strategies in the actor . Ecologically, amensalism plays a key role in regulating community structure and dynamics, particularly by controlling the distribution and abundance of sensitive species to maintain or facilitate transitions in ecosystems. In , may employ amensalistic tactics, such as allelopathic inhibition, to suppress competitors and create space for later seral stages, thereby directing the trajectory of habitat development from bare substrates to climax communities. For instance, allelopathic plants like can inhibit early-successional herbs, promoting a shift toward shrub-dominated landscapes in arid environments. At larger scales, amensalistic interactions enhance network stability in complex food webs by reducing overall connectance and mitigating the spread of perturbations, outperforming symmetric interactions like in diverse assemblages. Despite its prevalence in natural systems, amensalism is considered relatively rare or understudied compared to other interactions, often debated as an unintentional form of where the "neutral" effect on the actor is difficult to verify empirically. Field observations suggest it occurs frequently in multispecies communities, such as assemblages where it outnumbers traditional by a 2:1 , yet its incidental nature leads to challenges in distinguishing it from broader antagonistic processes. This ambiguity underscores the need for more targeted research to clarify its evolutionary and functional significance.

Neutralism

Neutralism refers to a type of biological interaction in which two or more coexist within the same or without exerting any detectable influence on each other's , , , or , denoted as a (0,0) outcome in fitness-based classifications. This lack of effect typically arises from spatial or temporal separation, where species occupy distinct niches, utilize non-overlapping resources, or avoid contact, rendering their coexistence incidental rather than interdependent. In the of the interaction , neutralism occupies the central "null" position, representing the baseline absence of positive or negative fitness impacts that contrasts with directional effects seen in (+,+), (-,-), or other asymmetric interactions. Examples of neutralism are prevalent among the vast majority of species pairs in complex, diverse ecosystems, where direct interactions are negligible due to the sheer scale of biodiversity. For instance, in microbial communities, neutralism dominates observed associations, accounting for approximately 65.6% of positive and 35.7% of negative network links in experimental bacterial networks, illustrating how unrelated microbes often coexist without mutual influence. Similarly, unrelated insect species in a forest ecosystem, such as a butterfly and a wood-boring beetle that exploit different plant parts at different times, exemplify neutralism, as their presence does not alter each other's population dynamics or resource availability. Proving true neutralism poses significant challenges in ecological research, as subtle, diffuse, or indirect effects—such as through shared environmental changes or higher-order interactions—may go undetected, complicating the distinction from weak or context-dependent influences. This difficulty underscores neutralism's theoretical role in community assembly models, where it serves as the default state for non-interacting pairs, facilitating the assembly of diverse communities by assuming minimal interference among most taxa unless evidence of effects emerges.

Non-trophic Interactions

Characteristics

Non-trophic biological interactions encompass ecological relationships among that lack direct energy or biomass transfer, in contrast to trophic interactions involving such as predation or mutualistic feeding. These interactions center on access to non-consumable resources like , mates, or behavioral opportunities, often mediated by physical structures, chemical signals, or environmental modifications provided by one organism to another. For example, may create habitats that shelter associates without nutritional exchange, thereby shaping community structure. Key characteristics of non-trophic interactions include their frequent reliance on interference mechanisms, where organisms directly influence each other through exclusion, attraction, or alteration of the shared , bypassing feeding dynamics. These interactions exhibit variability in temporal scale, manifesting as transient events like brief agonistic encounters between competitors or enduring effects, such as seasonal habitat provisioning that persists across generations and stabilizes populations. Unlike purely diffusive processes, they often demand proximity, amplifying their role in dense or structured habitats like coral reefs or soil microbiomes. Regarding fitness consequences, non-trophic interactions can impose positive (+/+), negative (-/-), or asymmetric (+/-) effects on participant ' survival, , and , paralleling trophic outcomes but via indirect pathways like resource partitioning or stress reduction rather than caloric gain or loss. Facilitative non-trophic exchanges, for instance, boost by ameliorating harsh conditions, while competitive interference diminishes it through denied access to breeding sites or foraging areas, ultimately influencing and maintenance. Detection of non-trophic interactions relies on behavioral assays, which involve controlled or field-based of responses such as territorial displays or aggregation patterns to quantify interaction strength. Complementary spatial modeling approaches, including statistical analyses of distributions and networks, enable of these links from observational data, accounting for environmental covariates to distinguish causal effects from spatial . Long-term experimental manipulations further validate these methods by isolating interaction impacts on metrics.

Examples

Non-trophic interactions manifest in various forms beyond direct energy or nutrient exchanges, influencing organismal behavior, space utilization, and chemical environments. One prominent behavioral example occurs in lek-breeding birds, where males engage in disruption tactics during displays to interfere with rivals' mating efforts. For instance, in species like the wire-tailed manakin, a rival male may supplant another at a display perch, temporarily halting the ongoing and potentially redirecting female attention, thereby reducing the displaced male's without consuming resources from the competitor. Spatial non-trophic interactions are evident in plant root systems, where physical crowding limits access to volume independent of depletion. In mixed stands of grasses and trees, such as those involving early-successional like , neighboring impose non-resource by mechanically occupying space, restricting the expansion of fine roots and altering their growth patterns through direct rather than resource theft. This physical barrier effect can reduce seedling in crowded conditions, highlighting how spatial constraints shape community structure. Chemical non-trophic interactions, such as , involve the release of inhibitory compounds that affect neighboring plants without physical contact or resource competition. Black walnut trees () exemplify this through , a natural exuded from , leaves, and hulls, which inhibits the growth of understory vegetation like tomatoes and potatoes by disrupting activity and elongation. Low concentrations of , such as 10 ppm (approximately 57 μM), can reduce growth of sensitive plants like tomatoes by 50%. In modern urban ecology, anthropogenic noise serves as a non-trophic disruptor of animal signaling. Elevated traffic and industrial sounds in cities mask bird songs, compelling species like European robins to shift vocalizations to higher frequencies or sing at dawn when noise levels drop, thereby altering mating and territorial interactions. A meta-analysis of 32 studies across wildlife (primarily avian) shows that noise pollution leads animals to increase minimum call frequency (effect size β = 0.86 ± 0.29), potentially affecting signal transmission and reception. While these spatial and behavioral cases may overlap with competitive dynamics, they primarily emphasize interference mechanisms (detailed in Competition).

Evolutionary Dynamics

Coevolution

Coevolution refers to the reciprocal evolutionary changes that occur between interacting , where adaptations in one species exert selective pressures that drive evolutionary responses in the other. This was first formally described in the context of plant-herbivore interactions, highlighting how and their host have influenced each other's diversification through specialized chemical defenses and counteradaptations. In a strict sense, coevolution involves specific, paired evolutionary changes in traits of two populations resulting directly from their , such as enhanced defenses in prey prompting improved strategies in predators. This can manifest as an "," where ongoing adaptations in one species, like a predator's more efficient capture mechanisms, continually select for countermeasures in the prey, such as faster escape behaviors or toxic secretions. Beyond pairwise interactions, diffuse coevolution occurs when multiple within a community exert collective selective pressures, leading to broader evolutionary shifts across guilds, as seen in assemblages of pollinators and flowering . A classic example of strict is the obligate mutualism between yucca plants ( and Hesperoyucca) and their pollinating (Tegeticula and Parategeticula), where female moths actively pollinate flowers using specialized mouthparts while laying eggs, ensuring offspring survival but limited to a fraction of seeds to avoid overexploitation. This system has driven parallel diversification, with moth tentacle-like appendages evolving to handle yucca stigmas and plants developing mechanisms to detect and abort fruits with excessive eggs. In antagonistic interactions, predator-prey dynamics illustrate paired evolution, such as (Lophiiformes) developing bioluminescent lures that mimic prey or conspecifics to attract victims, potentially co-evolving with prey sensory adaptations like low-light in deep-sea crustaceans, though direct reciprocity remains inferred from lure complexity matching prey detection thresholds. Key mechanisms underlying include gene-for-gene dynamics in host-parasite systems, where specific resistance genes in hosts correspond to avirulence genes in parasites, leading to rapid cycles of and counteradaptation as parasites evolve to overcome host defenses. This is exemplified in plant-pathogen interactions, where matching allelic variants determine success, promoting polymorphism maintenance through negative . The posits that species must continuously evolve to maintain relative fitness against evolving antagonists, akin to running to stay in place, as biotic interactions like impose perpetual selective pressure independent of abiotic changes. Proposed to explain constant risks across taxa, this framework underscores how coevolutionary arms races sustain diversity by favoring rare genotypes. Evidence for coevolution is drawn from phylogenetic congruence, where the branching patterns of interacting lineages mirror each other, indicating cospeciation driven by reciprocal selection, as observed in yucca-yucca moth phylogenies showing matched divergence times. Experimental evolution studies provide direct proof, such as laboratory coevolution of bacteria () and viruses (T7 phage), where hosts rapidly evolved resistance while parasites countered with increased infectivity over dozens of generations, demonstrating measurable trait shifts under controlled conditions. These approaches confirm that reciprocal selection can occur on contemporary timescales, altering interaction outcomes without external variables.

Interaction Networks

Biological interaction networks represent complex webs of relationships among multiple species, extending beyond pairwise interactions to encompass emergent properties that influence community structure and dynamics. Food webs, which map trophic interactions such as predation and herbivory, form a foundational type of these networks, illustrating energy flow and species dependencies across ecosystems. Mutualistic networks, in contrast, highlight cooperative relationships like pollination or seed dispersal, where species benefit reciprocally without direct energy transfer, often exhibiting patterns that enhance ecosystem services. Key structural metrics in these networks include modularity, which quantifies the division into densely connected subgroups (modules) of interacting species, and nestedness, which measures the extent to which the interaction partners of less-connected species are subsets of those for more-connected species. These properties, analyzed through comparative studies of over 50 mutualistic networks, reveal that modularity and nestedness often trade off, shaping network organization independently in many cases. Network stability emerges from the interplay between structural features like connectance—the proportion of realized possible interactions—and robustness to perturbations such as species loss. Higher connectance in food webs correlates with increased robustness to random species extinctions, as denser links distribute impacts and prevent cascading failures. In mutualistic networks, high buffers against disturbances by promoting redundancy, where diverse interactions maintain function despite losses; for instance, simulations show that mutualisms enhance persistence and temporal in multiplex networks combining trophic and non-trophic links. Weighted connectance, accounting for interaction strengths, further predicts better than unweighted measures, with positive correlations observed in flux-based models of and terrestrial webs. Overall, these dynamics underscore how , including elevated , mitigates vulnerability to environmental changes. Illustrative examples highlight these principles in natural systems. In coral reef networks, a balance of predation and —such as fish-algae and coral-zooxanthellae —structures trophic pathways, with global analyses of 6 reefs showing congruent predator-prey modules that sustain through modular compartmentalization. Perturbations like disrupt this balance, reducing modularity and . networks, conversely, demonstrate vulnerability to loss; studies in fragmented landscapes reveal that such degradation increases modularity while decreasing nestedness, leading to network collapse as specialist pollinators decline and interactions homogenize, with empirical data from 15 sites. These cases emphasize how structural shifts under pressure can precipitate systemic failures. Modeling interaction networks relies on graph theory to simulate and predict behaviors, representing species as nodes and interactions as edges to quantify properties like centrality and clustering. Applications post-2010 have integrated spatial and dynamic elements, enabling assessments of invasion resilience; for example, network analyses show that high modularity in mutualistic webs resists non-native species integration by limiting invasion pathways, as seen in analyses of 58 empirical networks where native generalists buffered against exotic generalists. Graph-based models, drawn from reviews of over 100 ecological graphs since 2010, inform conservation by identifying critical links for maintaining stability against biological invasions.

References

  1. [1]
    Species interactions - Understanding Global Change
    Species interact and thus affect one another. Species interactions can be direct or indirect, and have positive or negative effects.
  2. [2]
    Species Interactions – An Interactive Introduction to Organismal and ...
    A list of ways that two species can interact with each other, including their definitions and how the species are impacted by the interaction.
  3. [3]
    [PDF] Biological Systems from Molecular Interactions to Ecological Dynamics
    Sep 25, 2024 · A biological system is typically defined as a collection of interacting components that work together to maintain life processes. These systems ...
  4. [4]
    Biological Interaction - an overview | ScienceDirect Topics
    Biological interactions are defined as ongoing, energy-consuming processes that evolve in space and time, characterized by their complexity and dynamic ...
  5. [5]
    Direct and Indirect Interactions | Learn Science at Scitable - Nature
    The second major class of interactions, indirect effects, can be defined as the impact of one organism or species on another that is mediated or transmitted by ...
  6. [6]
    Species Interactions and Competition | Learn Science at Scitable
    Commensalism is an interaction in which one individual benefits while the other is neither helped nor harmed. For example, orchids (examples of epiphytes) ...Missing: entity | Show results with:entity
  7. [7]
    Mutualism - an overview | ScienceDirect Topics
    Mutualistic interactions can be facultative or obligatory. Facultative interactions are those in which neither species requires the other to persist but ...
  8. [8]
    14: Introduction to Species Interactions
    ### Definition of Species Interactions or Biological Interactions in Ecology
  9. [9]
    The importance of species interactions in eco-evolutionary ... - Nature
    Aug 6, 2021 · For example, species interactions can affect a species' evolutionary response to altered environmental conditions; and dispersal may release a ...
  10. [10]
    A species interaction kick-starts ecological speciation in allopatry
    Adaptation to different environments is thought to play a key role in speciation. However, speciation typically begins in allopatry, where reproductive ...
  11. [11]
    How Do Species Interactions Affect Evolutionary Dynamics Across ...
    Dec 4, 2015 · Stronger species interactions, however, can alter evolutionary outcomes and either dampen or promote evolution of constituent species depending ...<|separator|>
  12. [12]
    Biodiversity: The role of interaction diversity - ScienceDirect.com
    May 9, 2022 · Links between the diversity of ecological interactions and community structure have long been recognised but remain broadly understudied.
  13. [13]
    Soil-based ecosystem services: a synthesis of nutrient cycling and ...
    Aug 12, 2014 · Nutrient cycling influences nutrient dynamics in ecosystem, inter-specific interaction and ecosystem stability, thereby affecting ES provision ( ...
  14. [14]
    Biodiversity - World Health Organization (WHO)
    Feb 18, 2025 · Biodiversity supports key ecosystem services like soil fertility, natural pest control, pollination and water regulation. Preserving ...Key Facts · Impact · Sustainable, Healthy Food...
  15. [15]
    Molecular interaction networks and drug development: Novel ...
    Dec 5, 2022 · Network medicine‐based approaches to drug target identification and drug development offer avenues for advancing therapeutics in the current ...
  16. [16]
    Natural enemy interactions constrain pest control in complex ... - PNAS
    The natural enemies of insect pests are responsible for an estimated 50–90% of the biological pest control occurring in crop fields (4). In landscapes ...
  17. [17]
    Biodiversity conservation requires integration of species-centric and ...
    Species interactions are fundamental to ecological processes and associated ecosystem services (90). In particular, trophic interactions regulate the flow of ...
  18. [18]
    Insights into Protein–Ligand Interactions: Mechanisms, Models, and ...
    Essentially, proteins realize their biological functions through their direct physical interaction with other molecules, including proteins and peptides, ...
  19. [19]
    Overview of Protein–Protein Interaction Analysis - US
    This article reviews the nature of protein–protein interactions, the types of interactions, and describes common methods of and assays for detection.
  20. [20]
    Molecular Interaction - an overview | ScienceDirect Topics
    Molecular interactions involving proteins regulate virtually all aspects of cellular function. Here, the term “interaction” may be defined as the formation of a ...
  21. [21]
    Simple methods to determine the dissociation constant, Kd - PMC
    Sep 16, 2024 · The dissociation constant (Kd) is an equilibrium constant used universally in biochemistry and pharmacology to represent the binding affinity of ...
  22. [22]
    Allostery: a revolution in molecular biology, in 1965 | - Institut Pasteur
    Oct 15, 2025 · Allostery describes the phenomenon whereby small molecules known as ligands bind to certain regulatory proteins at a site that is topologically ...
  23. [23]
    On the nature of allosteric transitions: A plausible model
    Journal of Molecular Biology, Volume 12, Issue 1, May 1965, Pages 88-118, On the nature of allosteric transitions: A plausible model.Missing: seminal | Show results with:seminal
  24. [24]
    Following the tracks: How transcription factor binding dynamics ...
    The DBD recognizes and binds to a specific motif sequence (purple box) in regulatory promoter or enhancer regions of target genes. Upon binding, the TF recruits ...
  25. [25]
    The interaction of the antibody molecule with specific antigen - NCBI
    As a general principle, antibodies bind ligands whose surfaces are complementary to that of the antibody. A small antigen, such as a hapten or a short peptide, ...3-7. Antibodies bind antigens... · 3-8. Antibodies bind to...
  26. [26]
    A novel genetic system to detect protein–protein interactions - Nature
    Jul 20, 1989 · We have tested this system using two yeast proteins that are known to interact—SNF1 and SNF4. High transcriptional activity is obtained only ...
  27. [27]
    Protein-Protein Interactions: Co-Immunoprecipitation - PubMed
    Co-IP is considered to be one of the standard methods of identifying or confirming the occurrence of protein-protein interaction events in vivo.
  28. [28]
    Regulation of cadherin-mediated adhesion in morphogenesis - Nature
    Jul 15, 2005 · This article considers how the dynamic regulation of cadherin-mediated adhesion at the cell surface controls tissue morphogenesis.Missing: paper | Show results with:paper
  29. [29]
    MAPK signal pathways in the regulation of cell proliferation ... - Nature
    Mar 1, 2002 · In this paper we discussed their functions and cooperation with other signal pathways in regulation of cell proliferation. Similar content being ...
  30. [30]
    Progress in and promise of bacterial quorum sensing research
    Nov 16, 2017 · This Review highlights how we can build upon the relatively new and rapidly developing field of research into bacterial quorum sensing (QS).Missing: review paper
  31. [31]
    Cell–cell communication: new insights and clinical implications
    Aug 7, 2024 · The steps of cell signal transduction. Signal transduction and ... The role of the Hedgehog signaling pathway in cancer: a comprehensive review.
  32. [32]
    Apoptosis in Cellular Society: Communication between Apoptotic ...
    Dec 20, 2016 · In this review, we focus on the mutual interactions between apoptotic cells and their neighbors in cellular society and discuss issues relevant to future ...
  33. [33]
    T cells in health and disease | Signal Transduction and Targeted ...
    Jun 19, 2023 · In this review, we comprehensively summarize the current understandings of T cell biology and functions in both physiological and pathological ...
  34. [34]
    Communication is the key: biofilms, quorum sensing, formation and ...
    Quorum sensing regulates the metabolic activity of planktonic cells, and it can induce microbial biofilm formation and increased virulence. In this review we ...2. Biofilm-Producing... · Figure 1. Biofilm... · Quorum Sensing
  35. [35]
    Targeting the RAS/RAF/MAPK pathway for cancer therapy - Nature
    Dec 18, 2023 · In this review, we delve deeper into the mechanistic insights surrounding RAF kinase signaling in tumorigenesis and RAFi-resistance.
  36. [36]
    (PDF) Aristotle and Theophrastus on plant-animal interactions
    The observation of ecological interactions began with early natural historians including Aristotle, Hippocrates and Theophrastus who made several observations ...
  37. [37]
    History of Ecological Sciences, Part 52: Symbiosis Studies - Egerton
    Jan 1, 2015 · 450 BC–AD 1858. The most fundamental symbiotic relationship is animals eating plant material and animal physiological wastes becoming ...
  38. [38]
    Alexander von Humboldt: the first environmentalist
    May 20, 2019 · This insight was on the brink of Darwin's idea that competition among species is what drives evolutionary change. Nature was no Garden of Eden ...
  39. [39]
    The Natural History of Selborne by Gilbert White
    ### Summary of Bird-Insect Interactions in *The Natural History of Selborne*
  40. [40]
    Teleological Notions in Biology - Stanford Encyclopedia of Philosophy
    Mar 20, 1996 · In the eighteenth and nineteenth centuries, the status of teleology in biology was also contested as part of the vitalist-mechanist debate in ...
  41. [41]
    Alfred J. Lotka and the origins of theoretical population ecology - PMC
    Aug 4, 2015 · The equations describing the predator–prey interaction eventually became known as the “Lotka–Volterra equations,” which served as the starting ...
  42. [42]
    The Use and Abuse of Vegetational Concepts and Terms - Tansley
    Volume 16, Issue 3 pp. 284-307 Ecology Article Full Access The Use and Abuse of Vegetational Concepts and Terms AG Tansley, AG Tansley
  43. [43]
    Fundamentals of Ecology - Eugene Pleasants Odum, Gary W. Barrett
    ECOLOGY was first published in 1953 and was the vehicle Odum used to educate ... energy flow environment environmental example experimental factors ...
  44. [44]
    BUTTERFLIES AND PLANTS: A STUDY IN COEVOLUTION - 1964
    BUTTERFLIES AND PLANTS: A STUDY IN COEVOLUTION ... This work has been supported in part by National Science Foundation Grants GB-123 (Ehrlich) and GB-141 (Raven).
  45. [45]
    [PDF] Community Food Webs: Data and Theory. - The Rockefeller University
    This book describes food webs, which show which organisms feed on which others, and presents data and theory about their regularities.
  46. [46]
    The Human Microbiome Project - Nature
    Oct 17, 2007 · A strategy to understand the microbial components of the human genetic and metabolic landscape and how they contribute to normal physiology and predisposition ...Missing: mutualisms | Show results with:mutualisms
  47. [47]
    Understanding the Plant-microbe Interactions in CRISPR/CAS9 Era
    CRISPR/CAS9 for mechanistic insights in symbiotic nitrogen fixation. Symbiotic nitrogen fixation is one of the major beneficial legume-Rhizobium interactions ...Missing: mutualisms | Show results with:mutualisms
  48. [48]
    Implications of exceeding the Paris Agreement for mammalian ...
    Feb 6, 2023 · We evaluated the potential future impact of climate change on mammalian biodiversity across biogeographic assemblages, accounting for the level ...2 Methods · 3.1 The Bioclimatic Space... · 4 Discussion
  49. [49]
  50. [50]
  51. [51]
    On classifying interactions between populations | Oecologia
    On classifying interactions between populations. Original Papers; Published: September 1987. Volume 73, pages 272–281, (1987); Cite this article. Download PDF.
  52. [52]
    How context dependent are species interactions? - Chamberlain
    Apr 16, 2014 · Interaction outcomes are context-dependent when the sign and/or magnitude change as a function of the biotic or abiotic context.
  53. [53]
    Fitness Estimation for Ecological Studies: An Evaluation ... - Frontiers
    The topics of evolution by natural selection and ecological interactions are closely intertwined. Thus, measurements of evolutionary fitness are ubiquitous ...
  54. [54]
    Niche Overlap and Diffuse Competition - PNAS
    Although the average amount of overlap between pairs of species decreases with the intensity of diffuse competition, the overall degree of competitive ...Missing: criticism | Show results with:criticism
  55. [55]
    When can higher-order interactions produce stable coexistence?
    A pitfall of the common pairwise approach is that it misses the higher-order interactions potentially responsible for maintaining natural diversity.
  56. [56]
    Species with a Large Impact on Community Structure - Nature
    eating or being eaten — are called trophic interactions. In addition, some organisms, called foundation species, exert influence on a ...
  57. [57]
    [PDF] than a meal... integrating non-feeding interactions into food webs
    We thus define a non-trophic interaction as: (1) a direct non-feeding effect of a species on another or (2) a direct non-feeding effect of a species on an ...
  58. [58]
    Food Web: Concept and Applications | Learn Science at Scitable
    Food web offers an important tool for investigating the ecological interactions that define energy flows and predator-prey relationship (Cain et al. 2008).Missing: specialization | Show results with:specialization
  59. [59]
    Trophic Cascades Across Diverse Plant Ecosystems - Nature
    Trophic cascades are powerful indirect interactions that can control entire ecosystems. Trophic cascades occur when predators limit the density and/or behavior ...The History Of The Concept · When And Where Do... · Trophic Cascades Across...
  60. [60]
    Non-trophic interactions strengthen the diversity—functioning ...
    Aug 29, 2019 · Ecological communities are not only diverse in terms of the species that compose them but also in terms of the way they interact with each other ...
  61. [61]
  62. [62]
    15 |Predation and Parasitism - Utexas
    It is directional in the sense that one member of the pair (the predator) benefits from the association while the other (the prey) is affected adversely.
  63. [63]
    [PDF] The Role of Predation in Wildlife Population Dynamics
    Predation has been defined as individuals of one species eating living individuals of another species (Taylor. 1984). The role of predators in the population ...
  64. [64]
    Predator-Prey Relationships
    A predator is an organism that eats another organism. The prey is the organism which the predator eats. Some examples of predator and prey are lion and zebra, ...
  65. [65]
    [PDF] Dynamics of Predation - UNL Digital Commons
    Jun 4, 2019 · Predator and prey populations cycle through time, as predators decrease numbers of prey. Lack of food resources in turn decrease predator ...<|separator|>
  66. [66]
    Raptor Perception of Mismatch in Seasonally Polyphenic Prey
    Camouflage in particular has long been appreciated as an effective defense against predation and as a visual metaphor for antipredator adaptation (Thayer et al.
  67. [67]
    Identification of Conditions for Successful Aphid Control by ... - NIH
    Mar 28, 2017 · Ladybirds are effective aphid predators in greenhouses. Aphid population reduction exceeded 50% in most studies and ladybird release rates usually did not ...
  68. [68]
    Lady Beetles - Biological Control
    Lady beetles are effective predators if aphids are abundant (high pest density) but are thought to be less effective at low pest densities. There may also ...
  69. [69]
    Ecological theory of mutualism: Robust patterns of stability and ...
    Dec 15, 2021 · We review and synthesize this work, finding that mutualisms tend to exhibit stable coexistence at high density and destabilizing thresholds at low density.
  70. [70]
    Partner abundance controls mutualism stability and the pace ... - PNAS
    Mar 24, 2017 · Our study reveals that mutualistic strategy profoundly affects the pace of morphological change in traits involved in the interaction.Results And Discussion · Recurrent Mutualism... · Loss Of Mutualism With Ants...
  71. [71]
    Mighty Mutualisms: The Nature of Plant-pollinator Interactions
    Fig wasps and fig plants (Cook & Rasplus 2003), and yucca moths and yucca plants (Pellmyr 2003), are good examples of this extremely rare relationship.
  72. [72]
    Divergent investment strategies of Acacia myrmecophytes ... - PNAS
    We used a defensive ant-plant mutualism to investigate whether reward production by congeneric host plant species affects their level of exploitation. Such ...
  73. [73]
    The Mutualistic Interaction between Plants and Arbuscular ...
    Mycorrhizal fungi belong to several taxa and develop mutualistic symbiotic associations with over 90% of all plant species, from liverworts to angiosperms.
  74. [74]
    Host-Bacterial Mutualism in the Human Intestine - Science
    Mar 25, 2005 · New studies are revealing how the gut microbiota has coevolved with us and how it manipulates and complements our biology in ways that are mutually beneficial.Missing: microbiome | Show results with:microbiome
  75. [75]
    Cheating can stabilize cooperation in mutualisms - PMC
    Mechanisms of enforcement can be required to maintain cooperation but, when costly, these mechanisms will only be maintained when there is some way for cheaters ...2. The Model · (d) The Simulation · 3. Results
  76. [76]
    The stability of mutualism | Nature Communications
    May 27, 2020 · In this paper we explore the use of random matrix theory for studying these issues. The analysis shows how mutualism allows species to build up ...
  77. [77]
    Commensalism - an overview | ScienceDirect Topics
    Commensalism is defined as the interaction between two species in which one gains a fitness advantage while the other neither benefits nor is harmed.
  78. [78]
    (PDF) Barnacles - ResearchGate
    Barnacles attached to marine mammals are often referred to as ectoparasites; however, in actuality, they do not feed on their hosts but use them as a moving ...
  79. [79]
    Evaluating the structure of commensalistic epiphyte–phorophyte ...
    Mar 7, 2019 · This study showed that commensalistic epiphyte–phorophyte networks are more nested, but less modular, than the mutualistic ones. Overall, these ...
  80. [80]
    (PDF) Commensalism, Amensalism, and Synnecrosis - ResearchGate
    Commensalism in the narrow sense can be understood as an interaction strictly neutral for one organism and positive for the other. Neutral interaction is the ...
  81. [81]
    Commensalism - an overview | ScienceDirect Topics
    Commensalism is a relationship in which only one partner benefits unlike mutualism. However, there is a very thin line that separates mutualism and commensalism ...
  82. [82]
    [PDF] An overview of the current research on epiphyte ecology
    Outside of their role in storage, vascular epiphytes are also a source of food and habitat for many animals and microorganisms as well as useful ecological ...<|separator|>
  83. [83]
    Habitat Complexity Affects the Structure but Not the Diversity of ...
    These results indicate that complexity, whether physical or biological, is only one of many influences on biodiversity on coastal infrastructure.
  84. [84]
    Habitat complexity: approaches and future directions | Hydrobiologia
    Dec 21, 2011 · Areas of high habitat complexity can play a critical role in nutrient processing within an ecosystem due to the rescaling of turbulence ( ...Habitat Complexity... · Role In Conservation · Abiotic Complexity
  85. [85]
    Predation, Herbivory, and Parasitism | Learn Science at Scitable
    In parasitism, an individual organism, the parasite, consumes nutrients from another organism, its host, resulting in a decrease in fitness to the host. In ...
  86. [86]
    Principles of Parasitism: Host–Parasite Interactions - PMC
    The relationship between two living organisms can be classified as parasitic, symbiotic, or commensal.
  87. [87]
    The Ecology of Avian Brood Parasitism | Learn Science at Scitable
    Figure 1: Common Cuckoos (Cuculus canorus) parasitizing Common Redstarts (Phoenicurus phoenicurus) in Europe lay eggs whose colors mimic closely host egg colors ...Brood Parasitism As A... · The Coevolutionary... · Egg Mimicry
  88. [88]
    Evolution of parasite virulence when host responses cause disease
    The trade-off hypothesis of virulence evolution rests on the assumption that infection-induced mortality is a consequence of host exploitation by parasites.Missing: defenses | Show results with:defenses
  89. [89]
    A shift from exploitation to interference competition with increasing ...
    Exploitative competition is an indirect effect that occurs through use of a shared resource and depends on resource availability. Interference competition ...<|separator|>
  90. [90]
    Intense Natural Selection in a Population of Darwin's Finches ...
    Large birds, especially males with large beaks, survived best because they were able to crack the large and hard seeds that predominated in the drought.
  91. [91]
    Alien species affect the abundance and richness of native species in ...
    Dec 4, 2022 · Additional evidence shows that invasive plants generally outcompete native and other alien species due to functional traits related to growth ...
  92. [92]
    The struggle for existence : Gauze, G. F., 1910 - Internet Archive
    Apr 15, 2008 · The struggle for existence ; Publication date: 1934 ; Topics: Natural selection, Biomathematics ; Publisher: Baltimore, The Williams & Wilkins ...
  93. [93]
    The roles of amensalistic and commensalistic interactions in large ...
    Jul 13, 2016 · Amensalism can be defined as an interaction in which one organism inflicts harm to another organism without receiving any costs or benefits. ...
  94. [94]
    Amensalism - an overview | ScienceDirect Topics
    Amensalism is defined as an ecological interaction between two species in which one species is destroyed or inhibited while the other remains unaffected. ... How ...
  95. [95]
    Amensalism (Antagonism) Interaction: Types, Examples
    Aug 3, 2023 · Amensalism is a type of negative ecological interaction where one of the species is harmed or destroyed while the other either benefits or remains unaffected.<|control11|><|separator|>
  96. [96]
    Methyl Jasmonate as an Allelopathic Agent: Sagebrush Inhibits ...
    We conclude that MeJA release from sagebrush plays an allelopathic role for N. attenuata seed banks, but other unidentified compounds are also involved.Missing: amensalism | Show results with:amensalism
  97. [97]
  98. [98]
  99. [99]
    [PDF] 11|Interactions Between Populations - Utexas
    Neutralism (0, 0) occurs when the two populations do not interact and neither affects the other in any way whatsoever; it is thus of little ecological inter-.
  100. [100]
    Orienting the Interaction Compass: Resource Availability as a Major ...
    Oct 12, 2016 · This variation, known as context dependence, affects biotic and abiotic interactions alike and frustrates predictive efforts. For example, to ...
  101. [101]
    Enhancing microbial predator–prey detection with network and trait ...
    Feb 4, 2025 · Moreover, Jiang et al. [13] observed that neutralism predominated in 65.6% and 35.7% of positive and negative tested network associations, ...
  102. [102]
    Disruption, Dispersion, and Dominance in Lek-Breeding Birds
    Disruption usually is performed by a rival male who, as a result of his actions, may be able to mate with the female disrupted.
  103. [103]
    Resource and non‐resource root competition effects of grasses on ...
    Apr 8, 2009 · This study assessed the effects of resource (i.e. nutrients) and non-resource (i.e. interference for space) competition from fine roots of ...Missing: nutritional | Show results with:nutritional
  104. [104]
    Effect of Juglone and Other Allelochemicals in Walnut Leaves ... - NIH
    Jan 12, 2023 · The data suggest that juglone itself inhibits secondary metabolism in the plant, making it more susceptible to stress and pathogen attacks.
  105. [105]
    A meta‐analysis of the influence of anthropogenic noise on ...
    Apr 1, 2021 · Our meta-analysis shows that wildlife exposed to noise pollution modify communications to be higher pitched, although this result is primarily ...
  106. [106]
    Review: Ecological networks – beyond food webs - Ings - 2009
    Dec 11, 2008 · Mutualistic webs define the nexus of ecosystem services such as pollination and seed dispersal, rather than population dynamics or energy fluxes ...
  107. [107]
    Nestedness versus modularity in ecological networks: two sides of ...
    Jun 7, 2010 · We found that the correlation between nestedness and modularity for a population of random matrices generated from the real communities decreases significantly.
  108. [108]
    The modularity of pollination networks - PNAS
    This lack of correlation between nestedness and modularity suggests that modularity dictates the basic building blocks of networks. These building blocks or ...
  109. [109]
    Food-web structure and network theory: The role of connectance ...
    We analyze a broad range of 16 high-quality food webs, with 25–172 nodes, from a variety of aquatic and terrestrial ecosystems.
  110. [110]
    Mutualism increases diversity, stability, and function of multiplex ...
    May 1, 2020 · We find that consumer-resource mechanisms underlying plant-pollinator mutualisms can increase persistence, productivity, abundance, and temporal stability.
  111. [111]
    Food web stability and weighted connectance: the complexity ...
    Jan 12, 2016 · We found no relation between unweighted connectance and food web stability, but weighted connectance was positively correlated with stability.Food Web Stability And... · Methods · Food Web Data And Fluxes<|control11|><|separator|>
  112. [112]
    Complexity and stability of ecological networks: a review of the theory
    Jul 6, 2018 · The use of ecological-network models to study the relationship between complexity and stability of natural ecosystems is the focus of this review.Complexity--Stability Debate · Food Webs · Mutualistic Communities
  113. [113]
    Congruent trophic pathways underpin global coral reef food webs
    Sep 20, 2021 · All networks are characterized by species with narrow dietary preferences, arranged into distinct groups of predator–prey interactions. These ...Sign Up For Pnas Alerts · Results And Discussion · Materials And Methods
  114. [114]
    Habitat loss alters the architecture of plant–pollinator interaction ...
    Dec 1, 2013 · We found that interaction networks become more modular and connected with habitat loss in communities of plants and their pollinators.
  115. [115]
    Implications of non-native species for mutualistic network resistance ...
    Jun 11, 2019 · In real-world terms, invaded networks may withstand disturbance because their generalists are well-buffered, but are because non-native species are unlikely to ...
  116. [116]
    Graph theory in ecological network analysis: A systematic review for ...
    Sep 25, 2024 · Graph theory (GT) has become an indispensable tool for exploring environmental interconnections and elucidating ecological relationships.
  117. [117]
    Using network analysis to study and manage human-mediated ...
    Jul 15, 2023 · ... network theory to understand and predict biological invasions. Trends ... Hui C, Richardson DM (2019) How to invade an ecological network. Trends Ecol ...<|control11|><|separator|>