Fact-checked by Grok 2 weeks ago

Microevolution

Microevolution refers to evolutionary changes occurring within a or over relatively short timescales, characterized by shifts in frequencies due to processes such as , , , and . These mechanisms operate on already present or newly arising, leading to adaptations that enhance survival and reproduction in specific environments. Unlike , which encompasses larger-scale patterns like and the emergence of higher taxa, microevolution is directly observable and empirically verifiable through field studies and laboratory experiments. Key empirical evidence for microevolution includes the rapid development of antibiotic resistance in bacterial populations, where selective pressure from drugs favors pre-existing resistant mutants, altering within generations. Similarly, variations in beak morphology among on the demonstrate adaptive responses to food availability, with measurable changes in trait distributions correlating to environmental shifts. In wild populations, climate-driven selection has been documented to shift breeding , illustrating microevolutionary responses to ongoing environmental pressures. While microevolution is widely accepted as a foundational process supported by genetic and observational data, debates persist regarding its sufficiency to explain macroevolutionary patterns without additional causal factors, though empirical studies emphasize continuity in underlying mechanisms. These small-scale changes underscore the dynamic nature of populations, informing fields from to by revealing how interacts with selective forces in real time.

Definition and Conceptual Foundations

Core Definition and Distinctions

Microevolution is defined as the change in frequencies within a over time, encompassing observable shifts in genetic composition driven by mechanisms such as , , , and . This process occurs on a small scale, typically within a single interbreeding or closely connected groups, and can manifest as adaptations to local environmental pressures or random fluctuations in . Unlike broader patterns, microevolutionary changes are directly measurable through techniques like tracking in experiments or studies, often over spans of years to decades. The primary distinction from macroevolution lies in scope and taxonomic level: microevolution pertains to variations below the species boundary, such as shifts in trait distributions within a population (e.g., beak size variations in ), while involves larger-scale divergences, including events or the emergence of novel body plans across lineages. Proponents of a strict separation, often from creationist perspectives, argue that microevolutionary changes represent limited adaptations without the capacity to generate fundamentally new forms, citing empirical limits in observed transitions. In contrast, mainstream posits that arises as an accumulation of microevolutionary processes over geological timescales, supported by shared underlying genetic mechanisms, though direct real-time observation of macroevents remains infeasible due to their extended durations. This demarcation is not always rigid, as some transitional cases—such as incipient in laboratory settings—blur the lines, highlighting that the terms primarily serve purposes for categorizing evolutionary phenomena rather than denoting mechanistically distinct processes. Empirical validation of microevolution relies on quantifiable data, such as shifts in bacterial populations under selective antibiotics, underscoring its testability compared to macroevolutionary inferences drawn from records and phylogenetic analyses.

Observability and Timescales

Microevolution manifests as measurable shifts in allele frequencies within populations, observable over timescales ranging from single generations in rapidly reproducing organisms to decades in longer-lived species. In laboratory settings with bacteria, such changes occur within hours to days; for instance, Escherichia coli populations exposed to antibiotics develop resistance through selection on pre-existing variants or new mutations, with full adaptation evident after 10-20 generations. Similarly, experimental evolution in yeast or fruit flies demonstrates allele frequency alterations under controlled selection pressures over weeks, confirming microevolutionary dynamics in real time. In natural populations, microevolution is documented over years to centuries via field studies tracking phenotypic and genotypic shifts. The peppered moth (Biston betularia) exemplifies this: the melanic form rose from rarity in 1848 to comprising 98% of populations by 1895 amid industrial pollution darkening tree bark, favoring against predation, before declining post-1970s clean air acts. on the provide another case, with genomic analyses revealing beak morphology adaptations over 30 years on Daphne Major, driven by environmental droughts and interspecies competition altering selection pressures on loci controlling beak size. These observations span dozens of generations, contrasting with macroevolutionary requiring longer durations. Human populations also exhibit observable microevolution on contemporary timescales, such as changes linked to or resistance over a few generations, detectable through genomic surveys. Overall, microevolutionary rates vary by organismal and selection intensity, enabling direct empirical verification unlike processes inferred over geological epochs.

Mechanisms Driving Microevolution

Mutation as a Source of Variation

Mutations are heritable changes in the DNA sequence that serve as the ultimate source of genetic variation in populations, providing the raw material upon which natural selection, genetic drift, and other evolutionary processes act. Without mutations, populations would lack novel alleles, preventing adaptation to changing environments or the emergence of new traits within species. In microevolution, which encompasses changes in allele frequencies over short timescales, mutations introduce diversity at the molecular level, often occurring in germline cells to ensure transmission to offspring. Common types of mutations include point mutations, such as substitutions where one is replaced by another, which can result in silent changes that do not alter the sequence, missense mutations that substitute one for another, or mutations that prematurely terminate protein synthesis. Insertions and deletions of can cause frameshift mutations, shifting the of the and typically leading to nonfunctional proteins, while larger structural variants like duplications create redundant copies that may evolve new functions over time. Most mutations are neutral or deleterious, reducing , but rare beneficial mutations, such as those conferring antibiotic resistance in , can spread rapidly under selective pressure. Mutation rates vary across organisms but are generally low per generation, on the order of 10^{-9} to 10^{-10} per in eukaryotes, ensuring stability while allowing gradual variation accumulation. In , rates are higher per , around 10^{-3} to 10^{-10} per , facilitating observable evolutionary changes in laboratory settings, as demonstrated in long-term E. coli experiments where s enabled citrate utilization. These rates are influenced by factors like efficiency and environmental mutagens, but intrinsic genomic properties set baseline levels. In population contexts, even low rates generate substantial variation due to large population sizes and numerous replication events. Empirical studies confirm ' role in microevolutionary variation, with examples including the evolution of in via single substitutions that alter target proteins, allowing survival in treated populations. Similarly, in microbes, in efflux pump genes have driven multidrug resistance, illustrating how point provide adaptive alleles within generations. While beneficial are infrequent—estimated at less than 1% of total —their fixation under selection underscores ' primacy as variation generators in microevolutionary dynamics.

Natural Selection and Adaptation

Natural selection operates as a differential process wherein individuals with phenotypes conferring higher —defined as greater survival and in a specific —contribute disproportionately to subsequent generations, thereby altering frequencies within populations. This mechanism, central to microevolution, favors heritable variations that enhance to prevailing conditions, such as availability or predation pressures, without requiring novel genetic information beyond existing variation. Adaptations emerge as populations shift toward traits optimizing , over generations in response to environmental changes. In field studies of on the , documented acting on beak morphology during a 1977 drought, where birds with larger, thicker beaks survived better due to access to harder seeds, leading to heritable increases in beak size in the next generation. Over 30 years of observation from 1973 to 2003, selection fluctuated—unidirectional, oscillating, or episodic—resulting in measurable shifts in traits like beak depth and wing length, with heritability estimates around 0.65–0.87, demonstrating adaptation to varying food supplies and hybridization events. These changes occurred within species, illustrating microevolutionary adaptation without speciation in the studied lineages. Laboratory and industrial examples further substantiate selection's role. Bernard Kettlewell's 1950s experiments on peppered moths (Biston betularia) in polluted English woodlands showed melanic forms, rising from 5% to 95% frequency by 1895 amid soot-darkened trees, experienced 50% higher predation survival by birds compared to light forms, confirming camouflage-driven selection; post-pollution decline reversed this trend. In bacteria, exposure to antibiotics selects for rare resistant mutants; for instance, populations evolve resistance via mutations in target genes like gyrA, with fitness costs offset by compensatory changes, increasing resistant frequencies under selective pressure. Such instances highlight selection's efficacy in generating adaptive responses within microbial generations, often spanning hours to days. Critiques of specific experiments, such as moth resting postures not matching natural behaviors, underscore methodological challenges but do not negate the broader empirical pattern of selection-driven shifts in allele frequencies. Overall, natural selection's causal role in adaptation relies on pre-existing , with directional pressures yielding population-level changes verifiable through and metrics, distinguishing it from random processes in microevolutionary dynamics.

Genetic Drift and Random Changes

Genetic drift denotes the stochastic variation in frequencies within a arising from random sampling of gametes during , independent of selective pressures. This process operates in all finite populations but exerts a proportionally greater influence in smaller ones, where chance events can substantially alter genetic composition over generations. Unlike , lacks directionality toward , potentially leading to the fixation of deleterious alleles or the loss of beneficial ones solely by probability. The Wright-Fisher model formalizes genetic drift by assuming a diploid population of size N, where the next generation's allele count follows a binomial distribution based on the current frequency p, with 2N trials each having success probability p. Under this model, the expected change in allele frequency is zero, but the variance per generation equals p(1-p)/(2N), quantifying the random fluctuation's magnitude. Over time, this drift drives alleles toward fixation (frequency 1) or extinction (frequency 0), with the probability of fixation for a neutral allele equaling its initial frequency. Simulations and analytical solutions from this model demonstrate that, in the absence of other forces, genetic diversity erodes, as measured by heterozygosity declining by a factor of 1 - 1/(2N) each generation. Extreme manifestations of genetic drift include the bottleneck effect, where a sharp population reduction—such as from environmental catastrophes—amplifies random sampling, drastically curtailing . For instance, northern elephant seals underwent a in the , reducing their numbers to about 20 individuals, resulting in near-complete at many loci today. Similarly, the founder effect occurs when a small of individuals establishes a new population, inheriting only a fraction of the original ; human examples include elevated frequencies of certain alleles in isolated groups like the Finnish population due to historical migrations. Empirical studies, including genomic sequencing, confirm these effects reduce and increase drift's impact, often measurable via heterozygosity excess or patterns. In microevolutionary contexts, genetic drift interacts with mutation and selection but dominates in small or fragmented populations, contributing to local adaptations or maladaptations without deterministic fitness benefits. Observations in laboratory populations of Drosophila and bacteria reveal drift-induced allele frequency shifts aligning with model predictions, underscoring its role in generating variation amenable to subsequent selection. While mainstream population genetics emphasizes drift's neutrality, critiques from information-theoretic perspectives highlight its potential to degrade functional genetic complexity in isolated lineages, though direct causal links remain debated in empirical data.

Gene Flow and Population Connectivity

Gene flow refers to the movement of genetic material, specifically , between distinct populations through the of individuals or the dispersal of gametes, such as in or in aquatic . This process alters allele frequencies within recipient populations, serving as a key mechanism of alongside , selection, and drift. Unlike , which generates novel variation , or drift, which randomly fixes alleles in isolated groups, directly imports existing from donor populations, potentially introducing adaptive alleles or diluting locally fixed ones. In terms of connectivity, maintains genetic cohesion across spatially separated groups by counteracting divergence driven by local selection or drift. High rates of —quantified as the product of migration rate (m) and (N_e), or —reduce genetic differentiation, as measured by metrics like F_ST, where values approach zero under sufficient exchange. For instance, in the model of , values exceeding 1 typically prevent substantial divergence, fostering even over large distances. Empirical studies confirm this: in a of , from neighboring demes offset , preserving diversity despite small s on the order of dozens of individuals. Conversely, barriers to dispersal, such as geographic isolation or , elevate F_ST and promote differentiation; for example, in strongly philopatric seabirds like the , limited correlates with structured genetic clusters across breeding colonies. Gene flow's evolutionary impacts hinge on its rate and the compatibility of exchanged alleles. Low-to-moderate levels can enhance local by spreading beneficial variants, as seen in experimental crosses of small, inbred populations where immigrant alleles boosted by 20-50% without introducing deleterious loads. In marginal habitats, unidirectional flow from core populations of similar environments has been shown to increase recipient , mitigating . However, excessive gene flow may swamp adaptive divergence, homogenizing populations and constraining local specialization; simulations and genomic data from species like indicate that larval dispersal connectivity scales inversely with F_ST, with high Nm eroding environmental clines. Asymmetric flow, influenced by factors like wind patterns in or currents in taxa, can further bias differentiation, as documented in global analyses where directional dispersal yields imbalanced sharing. Overall, thus modulates microevolutionary trajectories by balancing connectivity against isolation, with observable effects in shifts over timescales of generations to centuries.

Empirical Evidence from Observations

Laboratory and Experimental Examples

One prominent laboratory example is the long-term evolution experiment (LTEE) with initiated by Richard Lenski in 1988 at , involving 12 replicate populations propagated daily in a glucose-limited medium, reaching over 75,000 generations by 2023. These populations exhibited parallel increases in fitness relative to the ancestor, measured by competitive assays showing up to fivefold gains driven by mutations in metabolic and regulatory genes. A key innovation occurred in one population around generation 31,500, where aerobic citrate utilization evolved via a tandem duplication enabling and rearrangement, conferring a growth advantage in citrate-rich conditions absent in the ancestral strain. Genomic sequencing revealed contingent evolution, with potentiating mutations preceding the citrate-enabling event, underscoring the role of historical contingency in adaptive trajectories. In , artificial selection experiments on bristle number traits, such as abdominal or sternopleural bristles, have demonstrated microevolutionary responses over tens of generations. For instance, selection for high or low sternopleural bristle number in replicate lines yielded realized heritabilities of approximately 0.25, with divergent lines differing by 2-3 bristles after 20-50 generations, linked to polygenic variation across multiple quantitative trait loci (QTLs). Quantitative genetic analyses identified 30-50 QTLs influencing bristle number, with additive effects explaining much of the response, though non-additive interactions and also contributed to correlated responses in components. These experiments, conducted since the mid-20th century and refined with modern mapping, illustrate how shifts allele frequencies for quantitative traits under controlled laboratory conditions. Laboratory evolution of antibiotic resistance in bacteria provides further evidence of rapid microevolutionary adaptation. In or serial transfer setups, E. coli and other species exposed to sublethal concentrations of antibiotics like trimethoprim evolve resistance within 10-100 generations through target site mutations (e.g., in folA for pathway inhibition) or amplified efflux pumps, increasing minimum inhibitory concentrations by 100- to 1000-fold. High-throughput platforms, such as mega-plates with antibiotic gradients, visualize spatial evolution where populations sequentially acquire mutations, forming rings of increasing resistance, with parallel genetic paths across replicates but occasional novel solutions like . Such experiments quantify costs, often 5-20% rate reductions in drug-free , which can be mitigated by compensatory mutations, highlighting trade-offs in adaptation.

Natural Population Studies

Studies of natural populations have documented microevolutionary changes through shifts in morphological traits and genetic frequencies in response to environmental pressures. In the , long-term observations of , particularly the (Geospiza fortis), reveal rapid adaptations in beak size driven by . Following a severe drought in 1977, finches with deeper beaks survived at higher rates due to better access to larger, harder seeds, resulting in an average increase in beak depth of approximately 4-5% in the subsequent generation. This shift was heritable, with genetic analysis later identifying the HMGA2 gene as a key regulator of beak morphology, where regulatory changes contributed to smaller beaks in response to a later hybridization event and food scarcity in 2004-2005. Peter and Rosemary Grant's multi-decade fieldwork demonstrated these changes occurring over timescales of years to decades, illustrating selection's role in altering population means without requiring new mutations. Industrial melanism in the (Biston betularia) provides another well-documented case, where the frequency of the dark melanic form rose from less than 5% in early 19th-century to over 90% in polluted areas by the mid-20th century. This increase correlated with sooty tree trunks favoring against bird predation, conferring a survival advantage estimated at 50% or more for melanics in polluted habitats. Post-1950s clean air regulations reduced , leading to a decline in melanic frequency to under 10% by the 1990s, with field experiments confirming predation as the primary selective agent through recapture rates showing 2:1 advantages for the better-camouflaged form. Genetic studies pinpoint a single locus () responsible for the melanistic , underscoring how rare variants can sweep through populations under strong selection. Threespine stickleback fish (Gasterosteus aculeatus) exhibit microevolutionary divergence in natural lakes, where marine ancestors colonized freshwater post-glaciation, evolving reduced armor plating within thousands of years. In , populations on islands uplifted by the 1964 earthquake showed genetic differentiation and trait shifts, such as changes in length, within 50 years, driven by to prey abundance. Genomic analyses reveal at loci like EDA for armor loss, with effective sizes influencing the rate of local . These observations highlight and drift's interactions with selection in shaping contemporary variation.

Human-Specific Instances

One prominent example of microevolution in humans is the evolution of , where genetic variants enable continued production into adulthood, allowing digestion of in beyond infancy. This trait emerged independently in multiple populations following the of dairy animals around 10,000 years ago, with strong positive selection driving increases; for instance, the -13910T variant rose from rarity to over 90% in northern populations within 5,000-7,000 years due to nutritional advantages in pastoralist societies. Similar mutations, such as the G13915 variant, show frequencies up to 30-50% in herding groups, reflecting localized selection pressures from as a famine-resistant source amid outbreaks. The sickle cell (HbS) exemplifies balancing selection in human populations exposed to , where heterozygotes (AS ) gain resistance to infection, maintaining the at intermediate frequencies despite homozygous (SS) individuals suffering . In regions like , HbS frequencies reach 10-20%, correlating directly with historical endemicity, with protective effects reducing severe risk by up to 90% in heterozygotes via mechanisms like altered sickling that inhibits parasite growth. Ongoing selection persists, as evidenced by higher AS survival rates in malarial areas, though frequencies decline in low- migrant populations, such as at ~8% versus 15-20% in African ancestral regions. High-altitude adaptation among populations demonstrates rapid microevolutionary change through selection on introgressed variants, particularly in the , which regulates response by modulating levels and oxygen transport. Derived from human admixture around 40,000 years ago, the reached frequencies of 80-90% in within the last 3,000-5,000 years, enabling efficient adaptation to plateau (above 4,000 meters) without the excessive seen in lowlanders, thus reducing risks like . This selection signature exceeds neutral expectations, with downregulating production while preserving oxygen delivery, contrasting with Andean adaptations relying more on elevated .

Relation to Macroevolution and Ongoing Debates

Proposed Mechanisms Linking Micro to Macro

Evolutionary biologists propose that macroevolutionary patterns, such as and diversification, emerge from the prolonged action of microevolutionary processes including , , , and across populations and geological timescales. This extrapolation assumes continuity between short-term adaptations within populations and long-term lineage divergence, though empirical translation remains incomplete in many cases. A key mechanism is protracted speciation, where microevolutionary dynamics at the level—such as splitting rates (λ') and rates (μ')—generate macroevolutionary outcomes like net diversification rates. Simulations using data from Weir and Schluter (2007) illustrate this: in temperate regions, higher splitting (λ' = 1.16) and rates yield rates of 0.58 per million years and rates of 0.45, producing observed ; tropical scenarios with lower rates (λ' = 1.13) result in slower (0.17) but lower (0.04), explaining higher diversity (60.81 on average). These processes thus account for patterns like latitudinal gradients without invoking distinct macro-level forces. Another proposed bridge involves demographic and ecological factors balancing geographic expansion against , captured by the conceptual index Φ = t_exp / TTBS, where t_exp is expansion time (area / dispersal rate) and TTBS is time to biological isolation (N_diff / (α + β)), with α as resource-partitioning rate and β as geographic opportunity rate. When Φ < 1 (rapid expansion precedes ), bursts occur, as modeled in adaptive radiations; examples include the Junco's quick colonization and , and lautus's local amid expansion. Conversely, resource-driven (high α) facilitates in heterogeneous environments, as seen in . Gene duplication provides a genetic , enabling one copy to retain original function while the other acquires novelty through and selection, potentially yielding macroevolutionary innovations like new morphological traits over time. This , observed in sequences, is theorized to scale up micro-variations into complex adaptations, though functional shifts often require coordinated regulatory changes. Challenges in predictability arise from factors like varying genetic correlations and adaptation speeds; for instance, short-term microevolutionary responses may not forecast macro patterns due to shifting environmental contexts or drift dominance in small populations, highlighting variable success in empirical links. Despite this, models suggest reduced fosters , connecting intra-population changes to inter-lineage splits.

Creationist Critiques and Limits of Extrapolation

Creationists and advocates accept as a mechanism for variation and within biological kinds but maintain that extrapolating it to —producing fundamentally new forms, organs, or body plans—exceeds the demonstrated capabilities of , selection, and other microevolutionary processes. for Creation Research (ICR) distinguishes , which generates varieties within a type (e.g., breeds from wolves), from , which purportedly creates new types; the latter, ICR argues, remains unobserved in real time and relies on unverified historical inference rather than repeatable . A core critique centers on the rarity of beneficial mutations sufficient for complex innovations. In The Edge of Evolution (2007), biochemist Michael Behe examines empirical data from microbial evolution, such as Plasmodium falciparum's resistance to chloroquine, which requires at least two specific amino acid substitutions in a transporter protein, occurring at a probability of approximately 1 in 10^{20} parasite replications under intense selection pressure. Behe contends this illustrates the "edge" of Darwinian evolution: while single-mutation changes enable microevolutionary tweaks, coordinated multi-mutation events needed for novel protein-protein interactions or cellular systems are probabilistically prohibitive over geological timescales, limiting evolution to minor modifications rather than the origin of irreducibly complex apparatuses like the blood-clotting cascade or cilial transport. Irreducible complexity further delimits extrapolation, as certain systems lose function entirely if any component is absent, precluding gradual assembly via microevolutionary steps without non-functional intermediates. Behe identifies the bacterial flagellum as exemplifying this, where its ~40 protein components form a rotary motor that defies stepwise Darwinian co-option due to interdependent parts, challenging claims that microevolutionary tinkering scales to macroevolutionary novelty. Critics from creationist perspectives, including those at ICR, note that documented microevolutionary cases—such as lens loss in cavefish or beak variations in Darwin's finches—typically involve regulatory shifts or deletions yielding specialized but reduced fitness in ancestral environments, not information-gaining innovations requisite for upward complexity. These limits, proponents argue, align with empirical in the fossil record, where transitional forms bridging major phyla remain scarce despite extensive sampling, suggesting microevolutionary processes operate within bounded parameters rather than unboundedly toward . Organizations like ICR and the , grounded in literal interpretations of , prioritize such data over uniformitarian assumptions in mainstream , which they view as influenced by naturalistic presuppositions that undervalue design inferences from probability and complexity barriers.

Empirical Gaps and Information Theory Considerations

While microevolutionary changes, such as shifts in allele frequencies due to selection or drift, are well-documented in laboratory settings and natural populations, empirical evidence for their extrapolation to macroevolutionary innovations remains limited. For instance, bacterial long-term evolution experiments, like Richard Lenski's E. coli study spanning over 75,000 generations since 1988, have produced adaptations such as citrate utilization under aerobic conditions via tandem gene duplications and point mutations, yet these involve regulatory tweaks to existing metabolic pathways rather than the origin of novel protein folds or irreducible complexes. Similarly, observations of speciation in plants and insects, such as polyploidy-induced reproductive isolation in Tragopogon species documented in the 1920s-1940s, demonstrate barriers to gene flow but do not exhibit the creation of fundamentally new genetic architectures required for transitions to higher taxonomic levels. Critics, including biochemist Michael Behe, argue that such examples fail to bridge to macroevolution because they lack evidence of coordinated mutations building irreducible complexity, as no transitional intermediates for systems like the bacterial flagellum have been observed forming de novo through microevolutionary processes. These gaps persist despite extensive genomic sequencing, with mainstream evolutionary models relying on inference from fossils and comparative anatomy rather than direct process observation, highlighting a reliance on untested assumptions about scalability. From an information theory standpoint, the functional specificity of DNA sequences poses a causal challenge to microevolution accounting for macroevolutionary novelty. Biological information, akin to specified complexity in algorithmic terms, requires not just raw sequence length (Shannon entropy) but improbable functional arrangements improbable under random variation; Douglas Axe's experimental surveys of protein folds, involving randomization of 10^74 possible sequences for a 150-amino-acid domain, found functional proteins in fewer than 1 in 10^77 configurations, underscoring the rarity of viable innovations via mutation alone. Beneficial mutations observed in microevolution, such as those conferring pesticide resistance in insects (e.g., knockdown resistance in mosquitoes via voltage-gated sodium channel alterations since the 1950s), typically entail loss-of-function or regulatory adjustments that reduce overall genomic viability, consistent with genetic entropy models positing net informational decline over generations due to mutation accumulation rates exceeding repair fidelity. John Sanford's simulations, drawing on human mutation rates of approximately 100-200 new variants per genome per generation, predict a fitness decay threshold within thousands of generations absent purifying selection strong enough to counter near-neutral deleterious effects, which empirical pedigree studies in isolated populations like the Finnish disease heritage confirm through rising genetic load. Mainstream rebuttals invoke gene duplication and co-option, yet these mechanisms recycle pre-existing information without empirically verified net gains in specified complexity for complex traits, as neo-Darwinian models struggle to quantify the probabilistic barriers without invoking unobservable deep time or convergence. This informational asymmetry suggests microevolutionary dynamics may be constrained to variation within bounded potential, limiting causal realism in extrapolations to macroevolution.

Historical Development of the Concept

Coinage and Early Usage

The term microevolution was introduced by Yuri Filipchenko in his 1927 German-language book Variabilität und Variation, where he defined it as intraspecific evolutionary change driven by processes such as selection and variation within populations or species, in contrast to , which he viewed as requiring distinct, potentially non-gradual mechanisms for the emergence of new taxa. Filipchenko, an early and orthogeneticist who emphasized directed variation, used the term to highlight observable small-scale adaptations while questioning whether they sufficed to explain larger phylogenetic transitions without additional saltatory elements. This coinage reflected the era's tensions between Mendelian and Darwinian , as Filipchenko sought to integrate emerging data on with evolutionary theory. Early adoption of microevolution occurred primarily in Russian and German scientific literature, with Filipchenko's student Theodosius Dobzhansky popularizing the concept in English through his 1937 book Genetics and the Origin of Species, where he reframed it within the emerging modern synthesis as changes in gene frequencies observable over short timescales. Dobzhansky treated microevolutionary processes—mutation, selection, drift, and —as empirically verifiable and foundational, arguing they provided the mechanistic basis for without invoking non-Darwinian jumps, though he acknowledged gaps in extrapolating to macro scales. By the late 1930s, the term appeared in Western discussions of , such as in works by and , who modeled shifts mathematically to quantify microevolutionary dynamics. This usage solidified microevolution as a core concept in , distinct from earlier vague notions of minor variation predating genetic frameworks.

Integration into Modern Evolutionary Biology

The modern evolutionary synthesis, emerging in the 1930s and 1940s, reconciled Charles Darwin's theory of with Gregor Mendel's principles of by formalizing microevolution as changes in frequencies within populations, modeled through . This integration demonstrated that small-scale genetic variations, driven by mechanisms such as , , , and , could account for adaptive shifts observable in natural and experimental populations, without invoking Lamarckian or saltational changes. Key mathematical foundations were laid by Ronald A. Fisher in his 1922 paper on the correlation between relatives and , J.B.S. Haldane's 1924 work on selection intensities, and Sewall Wright's 1931 shifting balance theory, which quantified how frequencies deviate from Hardy-Weinberg equilibrium under evolutionary forces. Theodosius Dobzhansky's 1937 book and the Origin of Species provided empirical genetic evidence from experiments, showing how microevolutionary processes like chromosomal inversions and selection on polygenic traits generate population-level adaptations, thus bridging Mendelian genetics with Darwinian . Ernst Mayr's 1942 Systematics and the Origin of Species emphasized the role of geographic isolation in microevolutionary divergence, integrating with population-level changes to explain speciation precursors. Julian Huxley's 1942 Evolution: The Modern Synthesis synthesized these contributions, explicitly terming the framework and highlighting microevolution's role in unifying disparate biological fields including and . In contemporary terms, this integration underpins and genomic studies, where microevolutionary models predict trajectories using tools like the Wright-Fisher model, validated by sequencing data from evolving populations such as under selection. These processes remain central to , as shifts—directly measurable via markers like SNPs—provide the mechanistic basis for phenotypic evolution, though debates persist on their sufficiency for larger-scale patterns addressed elsewhere.

Contemporary Applications and Research

Resistance Phenomena in Pests and Pathogens

Resistance to antibiotics in bacterial pathogens exemplifies microevolutionary change through natural selection on genetic variants, where exposure to antimicrobial agents increases the frequency of pre-existing or newly arisen resistance alleles within populations. Mechanisms include enzymatic inactivation of drugs, efflux pumps expelling antibiotics from cells, modification of drug targets such as ribosomal proteins or cell wall synthesis enzymes, and reduced permeability of bacterial membranes. For instance, penicillin resistance in Staphylococcus aureus emerged shortly after the antibiotic's introduction in the 1940s, driven by mutations in penicillin-binding proteins and beta-lactamase production, leading to strains like methicillin-resistant S. aureus (MRSA) that now cause over 80,000 invasive infections annually in the United States alone. Horizontal gene transfer via plasmids and integrons further accelerates resistance dissemination, as seen in multidrug-resistant Enterobacteriaceae carrying extended-spectrum beta-lactamases (ESBLs). In insect pests, insecticide resistance arises similarly from selection favoring alleles that confer physiological tolerance, with over 600 species documented as resistant to one or more compounds by 2020. Key mechanisms encompass enhanced metabolic detoxification via enzymes, glutathione S-transferases, and esterases that conjugate or oxidize toxins; target-site insensitivity, such as altered in organophosphate-resistant mosquitoes; and behavioral avoidance, though less common. A classic case is resistance in houseflies (Musca domestica), first reported in 1946 after widespread agricultural use, where resistant populations evolved elevated oxidase activity to break down the insecticide, spreading globally within years. In agricultural pests like the (Leptinotarsa decemlineata), multiple resistance to neonicotinoids and pyrethroids has evolved through and point mutations, imposing fitness costs like reduced in absence of selection but persisting under continuous pressure. Herbicide resistance in weeds demonstrates rapid allele frequency shifts in plant populations under monoculture farming, with 267 unique cases across 93 weed species confirmed by 2023, primarily to glyphosate and ALS-inhibiting herbicides. Evolutionary pathways involve non-target-site resistance through accelerated herbicide metabolism by cytochrome P450s or GSTs, and target-site mutations like proline-to-serine substitutions in EPSPS enzymes for glyphosate tolerance in species such as Palmer amaranth (Amaranthus palmeri). The first resistant weed, Echinochloa spp. to propanil, appeared in 1957 in Arkansas, but proliferation accelerated post-1996 with glyphosate-resistant crops, leading to resistant Amaranthus populations doubling in size and yield losses exceeding 50% in untreated fields. Gene duplication events amplify detoxification genes, enabling polygenic resistance without complete loss of fitness. Fungal pathogens exhibit antifungal resistance via analogous genetic adaptations, particularly in Candida and Aspergillus species, where azole resistance has surged due to agricultural fungicide use selecting environmental reservoirs transmissible to humans. Mechanisms include efflux pump overexpression, mutations in ergosterol biosynthesis genes like CYP51A in Aspergillus fumigatus, and biofilm formation enhancing tolerance. For example, triazole-resistant A. fumigatus isolates, carrying G54 or L98 mutations, increased from rare in the 1990s to over 15% of clinical strains by 2020 in , linked to demethylation inhibitor fungicides in crop protection. Candida auris, an emerging multidrug-resistant first identified in 2009, shows intrinsic resistance to via efflux and acquired echinocandin resistance through fks1 gene alterations, contributing to hospital outbreaks with mortality rates up to 60%.

Predictive Models and Recent Genetic Findings

Genomic prediction models have emerged as powerful tools for forecasting microevolutionary changes by estimating individual breeding values from genome-wide markers, enabling detection of shifts in quantitative traits under selection. In a longitudinal study of Soay sheep (Ovis aries) spanning 35 years (1985–2020), researchers applied genomic best linear unbiased prediction (GBLUP) to adult weight data, revealing a significant increase in mean breeding values consistent with ongoing natural selection, thereby validating the model's ability to quantify microevolutionary responses over decadal timescales. Similarly, in wild populations of great tits (Parus major), population genetic models calibrated with empirically derived fitness estimates accurately predicted allele frequency changes at causal loci into the subsequent generation, with prediction errors below 5% for selected variants, demonstrating the feasibility of short-term evolutionary forecasting in natural settings. These models extend classical frameworks like the , which assumes polygenic from numerous small-effect loci, to incorporate genomic data for enhanced precision in predicting trait evolution amid and selection. For instance, multi-population genomic prediction using algorithms outperformed traditional multitrait GBLUP in forecasting breeding values across breeds, achieving up to 15% higher accuracy by accounting for and population structure—principles directly applicable to wild microevolutionary dynamics. However, predictability remains constrained by factors such as mutational supply and , as highlighted in theoretical analyses showing that while bias can bias evolutionary trajectories, empirical validation requires dense genomic sampling to resolve cryptic variation. Recent genetic findings underscore the empirical grounding of these predictions, with whole-genome sequencing revealing rapid allele frequency shifts in response to environmental pressures. A 2024 analysis of big datasets from living organisms demonstrated that microevolutionary rates observed over years to decades—such as adaptive shifts in Drosophila populations—extrapolate reliably to predict macroevolutionary patterns over millennia, bridging short- and long-term scales through consistent selection gradients. In parasite-host systems, genomic surveys of Trichinella species confirmed microevolutionary divergence via restricted gene flow and selection on immune-related loci, aligning with model predictions of localized adaptation without requiring novel mutations. These advances, supported by high-throughput sequencing since 2020, affirm microevolution's predictability while emphasizing the need for integrated models that parse neutral drift from adaptive signals to avoid overestimation of evolutionary rates.

Implications for Conservation and Human Health

Microevolution plays a pivotal role in conservation biology by enabling populations to adapt to rapid environmental changes, such as those driven by climate change and habitat alteration, thereby potentially averting extinction in endangered species. Genetic variation within populations allows for selection of traits conferring resilience, a process termed evolutionary rescue, which has been documented in species like salmonids facing warming waters and altered flow regimes. However, low genetic diversity in fragmented or bottlenecked populations—common in endangered taxa—constrains this adaptive potential, increasing vulnerability; for instance, studies on island birds reveal slower microevolutionary rates in isolated habitats, limiting responses to anthropogenic pressures. Conservation strategies increasingly incorporate genomic assessments to forecast adaptive capacity, emphasizing the maintenance of gene flow and diversity to harness microevolutionary processes against threats like invasive species and pollution. In human health contexts, microevolution underlies the rapid emergence of in bacterial pathogens, where selective pressures from use drive shifts in allele frequencies toward resistant genotypes, complicating treatments for infections like and . studies confirm that can acquire via and within days to weeks under drug exposure, as seen in Escherichia coli and Staphylococcus aureus lineages. Similarly, pathogen microevolution post-vaccination leads to serotype replacement and immune escape, exemplified by Streptococcus pneumoniae, where introduction of the 7-valent (PCV7) in 2000 shifted dominant strains, reducing targeted serotypes but elevating non-vaccine types in carriage and disease. This within-host and population-level evolution necessitates ongoing surveillance and adaptive updates, as rapid genetic changes—often on timescales of months—outpace static interventions, contributing to over 1.27 million annual deaths from resistant infections as of 2019 estimates. Understanding these dynamics informs stewardship programs to curb spread, such as cycling and development.

References

  1. [1]
    What is microevolution? - Understanding Evolution - UC Berkeley
    Microevolution is simply a change in gene frequency within a population. Evolution at this scale can be observed over short periods of time.
  2. [2]
    9.4: Microevolution - Biology LibreTexts
    Sep 4, 2021 · Evolutionary change that occurs over relatively short periods of time within populations is called microevolution.
  3. [3]
    Mechanisms of microevolution - Understanding Evolution
    There are a few basic ways in which microevolutionary change happens. Mutation, migration, genetic drift, and natural selection are all processes that can ...
  4. [4]
    Evolution at different scales: micro to macro
    Microevolution happens on a small scale (within a single population), while macroevolution happens on a scale that transcends the boundaries of a single species ...Examples of microevolution · What is microevolution? · What is macroevolution?
  5. [5]
    Examples of microevolution - Understanding Evolution - UC Berkeley
    Pesticide resistance, herbicide resistance, and antibiotic resistance are all examples of microevolution by natural selection.
  6. [6]
    Climate change drives microevolution in a wild bird - Nature
    Feb 22, 2011 · We show the first evidence that recent climate change alters natural selection in a wild population leading to a microevolutionary response.
  7. [7]
    Evidence for evolution in response to natural selection in a ...
    Our results show that microevolution can be detectable over relatively few generations in humans and underscore the need for studies of human demography and ...
  8. [8]
    Defining microevolution - Understanding Evolution - UC Berkeley
    within a single population. That means narrowing our focus to one branch of the tree of life.
  9. [9]
    A Perspective on Micro-Evo-Devo: Progress and Potential - PMC - NIH
    The term “micro-evo-devo” refers to the combined study of the genetic and developmental bases of natural variation in populations and the evolutionary ...
  10. [10]
    the relevance of microevolution for human health and disease - PMC
    Apr 29, 2013 · We propose a different approach by addressing more empirical and health-oriented research concerning past, current and future microevolutionary changes.
  11. [11]
    Microevolution vs Macroevolution - BYJU'S
    Large-scale and visible changes that occur above the level of species are known as macroevolution. The change occurs at an intraspecific level.
  12. [12]
    What Is The Difference Between Macroevolution And Microevolution?
    Oct 1, 1996 · Macroevolution refers to major evolutionary changes over time, the origin of new types of organisms from previously existing, but different, ...
  13. [13]
    [PDF] Microevolution and Macroevolution Are Governed by the Same ...
    Nov 1, 2009 · Most scientists would accept that there are distinct phenomena that can be categorized as microevolutionary and macroevolutionary.
  14. [14]
    12.2 Microevolution and Macroevolution - General Biology II - Fiveable
    Microevolution can be directly observed, while macroevolution is inferred from fossils and genetic data. Together, they explain the incredible diversity of life ...
  15. [15]
    Microevolution and Macroevolution | CK-12 Foundation
    Dec 17, 2015 · Microevolution is the process by which organisms change in small ways over time. · Macroevolution refers to larger evolutionary changes that ...
  16. [16]
    Bacterial Evolution: The road to resistance - eLife
    Oct 25, 2019 · Resistance to antimicrobial treatments, also known as AMR, evolves rapidly, often over the course of a single infection. It occurs either by ...
  17. [17]
    Genomic evolution of antibiotic resistance is contingent on genetic ...
    Jan 13, 2021 · Populations of bacteria, yeast, and viruses reproduce quickly and grow to large numbers, enabling one to observe evolution over timescales of ...<|separator|>
  18. [18]
    The Peppered Moth - Age of Revolution
    The first dark Peppered Moth was recorded in Manchester in 1848 and, by 1895, this variety accounted for 98% of these Moths in the city. Around this time, ...
  19. [19]
    Selective bird predation on the peppered moth: the last experiment ...
    Melanics became common during the industrial revolution, but since 1970 there has been a rapid reversal, assumed to have been caused by predators selecting ...<|control11|><|separator|>
  20. [20]
    Community-wide genome sequencing reveals 30 years of Darwin's ...
    Sep 29, 2023 · This study takes advantage of 30 years of study of a classic system to elucidate the role of genetic architecture and introgression in adaptation.
  21. [21]
    Evidence for evolution in response to natural selection in a ... - PNAS
    Oct 3, 2011 · Our results show that microevolution can be detectable over relatively few generations in humans and underscore the need for studies of human demography and ...Missing: observability | Show results with:observability
  22. [22]
    New research shows microevolution can be used to predict how ...
    May 10, 2024 · New research shows microevolution can be used to predict how evolution works on much longer timescales · Big datasets from living creatures and ...
  23. [23]
    Mutation and the evolution of recombination - PMC - NIH
    Mutation is the ultimate source of all genetic variation, and is essential for evolution by natural selection: indeed, most of our genome has been shaped ...
  24. [24]
    DNA and Mutations - Understanding Evolution - UC Berkeley
    Mutations are essential to evolution; they are the raw material of genetic variation. Without mutation, evolution could not occur. In this tutorial, we'll ...
  25. [25]
    The effects of mutations - Understanding Evolution - UC Berkeley
    Not all mutations matter for evolution. Somatic mutations occur in non-reproductive cells and so won't be passed on to offspring.
  26. [26]
    Types of mutations - Understanding Evolution - UC Berkeley
    Types of mutations include substitution (base exchange), insertion (extra base pairs), deletion (lost DNA), and frameshift (altered codon parsing).
  27. [27]
    Mutation and Genetic Variation – Molecular Ecology & Evolution
    Duplications can provide raw material for evolution, as the extra copy of a gene can accumulate mutations and potentially gain a new function. Deletions ...
  28. [28]
    The effects of mutations - Understanding Evolution - UC Berkeley
    Beneficial effect​​ Other mutations are helpful to the organisms that carry them. For example, DDT resistance in insects is sometimes caused by a single mutation ...
  29. [29]
    Experimental estimates of germline mutation rate in eukaryotes
    Mutation rates vary greatly among eukaryotes; plants and mammals have higher rates than arthropods, and unicellular organisms have the lowest. Rates increase ...Lay Summary · Introduction · Results and discussion · Concluding remarks
  30. [30]
    Studying Mutation and Its Role in the Evolution of Bacteria - PMC - NIH
    Mutation is the engine of evolution in that it generates the genetic variation on which the evolutionary process depends.
  31. [31]
    A new era of mutation rate analyses: Concepts and methods - PMC
    This paper aims to provide a comprehensive overview of the key concepts and methodologies frequently employed in the study of mutation rates.
  32. [32]
    Understanding Natural Selection: Essential Concepts and Common ...
    Apr 9, 2009 · This paper provides an overview of the basic process of natural selection, discusses the extent and possible causes of misunderstandings of the ...
  33. [33]
    Unpredictable evolution in a 30-year study of Darwin's finches
    Natural selection occurred frequently in both species and varied from unidirectional to oscillating, episodic to gradual. Hybridization occurred repeatedly ...
  34. [34]
    The peppered moth and industrial melanism: evolution of a natural ...
    Dec 5, 2012 · In Kettlewell's experiments, melanic and typical moths were at relatively high densities. ... Selection experiments on industrial melanism ...
  35. [35]
    The roles of history, chance, and natural selection in the evolution of ...
    Antibiotics can impose strong selection pressure on microbial populations, leading to highly repeatable evolutionary outcomes (Vogwill et al., 2014; ...
  36. [36]
    Mutations and selection - ReAct - Action on Antibiotic Resistance
    Mutations can result in antibiotic resistance in bacteria. Resistant bacteria survive antibiotic treatment and can increase in numbers by natural selection.
  37. [37]
    Genetic drift - Understanding Evolution - UC Berkeley
    Genetic drift is one of the basic mechanisms of evolution. In each generation, some individuals may, just by chance, leave behind a few more descendants.
  38. [38]
    Genetic Drift - National Human Genome Research Institute
    Genetic drift is a mechanism of evolution characterized by random fluctuations in the frequency of a particular version of a gene (allele) in a population.
  39. [39]
    Genetic Drift and Diversity – Molecular Ecology & Evolution
    Genetic drift is an evolutionary process that results in random changes in allele frequencies within a population from one generation to the next.
  40. [40]
    Learning mitigates genetic drift - PMC - NIH
    Nov 27, 2022 · Genetic drift is a major evolutionary process characterized by random fluctuations of allele frequencies in a population from one generation to ...
  41. [41]
    Introduction to the Wright-Fisher Model
    Mar 31, 2019 · The Wright-Fisher model is a discrete-time Markov chain that describes the evolution of the count of one of these alleles over time.
  42. [42]
    Wright-Fisher Model - an overview | ScienceDirect Topics
    The Wright–Fisher model is defined as a population genetics model that describes the genetic composition of a population with discrete, ...
  43. [43]
    Bottlenecks and founder effects - Understanding Evolution
    Because genetic drift acts more quickly to reduce genetic variation in small populations, undergoing a bottleneck can reduce a population's genetic variation ...Missing: evidence | Show results with:evidence
  44. [44]
    20.9.2: Genetic Drift - Biology LibreTexts
    Dec 16, 2021 · The bottleneck effect occurs when only a few individuals survive and reduces variation in the gene pool of a population. The genetic structure ...Genetic Drift vs. Natural... · The Bottleneck Effect · The Founder EffectMissing: evidence | Show results with:evidence
  45. [45]
    A Population-Genetic Test of Founder Effects and Implications for ...
    A founder effect can account for the presence of an allele at an unusually high frequency in an isolated population if the allele is selectively neutral.
  46. [46]
    Genetic Drift and Founder Effects: Implications for Population ...
    Mar 17, 2024 · Genetic factors, such as founder effects and population bottlenecks, may contribute to the increased prevalence of certain genetic disorders ...Missing: evidence | Show results with:evidence
  47. [47]
    Serial founder effects and genetic differentiation during worldwide ...
    These findings demonstrate that genetic drift played a major role in shaping allele frequencies and created genetic differentiation among newly formed ...
  48. [48]
    Gene flow - Understanding Evolution - UC Berkeley
    Gene flow includes lots of different kinds of events, such as pollen being blown to a new destination or people moving to new cities or countries. If genetic ...
  49. [49]
    Natural Selection, Genetic Drift, and Gene Flow Do Not Act in ...
    Natural selection, genetic drift, and gene flow are the mechanisms that cause changes in allele frequencies over time.
  50. [50]
    Gene Flow and Population Differentiation | Science
    Gene Flow and Population Differentiation: Studies of clines suggest that differentiation along environmental gradients may be independent of gene flow. John A.
  51. [51]
    Gene flow counteracts the effect of drift in a Swiss population of ...
    Using simulations and empirical data, we show that effective population size is small, and that genetic drift would lead to a marked decline in genetic ...
  52. [52]
    Genetic differentiation in an endangered and strongly philopatric ...
    Jun 19, 2021 · For such populations, gene flow and connectivity to other populations are crucial for maintaining genetic variation [9, 10]. The effects of ...
  53. [53]
    [PDF] Experimental Evidence for Beneficial Fitness Effects of Gene Flow in ...
    A small amount of gene flow increases fitness in small, inbred populations. Higher rates of gene flow may increase fitness but decrease phenotypic divergence.
  54. [54]
    Gene flow effects on populations inhabiting marginal areas: Origin ...
    Jun 24, 2020 · Our results suggest that gene flow between marginal populations from similar environmental conditions increases the fitness of the recipient ...
  55. [55]
    Patterns and effects of gene flow on adaptation across spatial scales
    Jun 18, 2024 · Gene flow can have rapid effects on adaptation and is an important evolutionary tool available when undertaking biological conservation and restoration.
  56. [56]
    Global wind patterns shape genetic differentiation, asymmetric gene ...
    Apr 19, 2021 · The “asymmetry” hypothesis: Population pairs linked by more directionally asymmetrical winds have more imbalanced gene flow asymmetry ratios. 4.
  57. [57]
    How species evolve collectively: implications of gene flow and ... - NIH
    Species evolve collectively through the spread of advantageous alleles, even with low gene flow, while also differentiating at other loci.
  58. [58]
    Experimental evolution and the dynamics of adaptation and genome ...
    May 16, 2017 · The long-term evolution experiment, or LTEE, is simple both conceptually and practically. Twelve populations were started the same ancestral ...
  59. [59]
    Long-Term Experimental Evolution in Escherichia coli. XII. DNA ...
    The longest-running evolution experiment involves 12 populations of Escherichia coli founded from the same ancestral strain, which have been serially propagated ...
  60. [60]
    Innovation in an E. coli evolution experiment is contingent on ...
    Overall, both models give very similar results that reflect the well-known deceleration in fitness gains during the Lenski long-term experiment [3,4,14,16].
  61. [61]
    Modes of Selection and Recombination Response in Drosophila ...
    Selection for low bristle number initially resulted in an increase; for example, at generation 2, L1 had a mean of 19.6 ± 0.54 and L2 had a mean number of 18.2 ...
  62. [62]
    Drosophila bristles and the nature of quantitative genetic variation
    The number of QTLs detected that affect bristle number increased to 53, with 33 affecting sternopleural bristle number, 31 affecting abdominal bristle number, ...
  63. [63]
    Selection for increased abdominal bristle number in Drosophila ...
    Selection for increased abdominal bristle number in Drosophila melanogaster with concurrent irradiation. II. Populations derived from an outbred cage population.
  64. [64]
    Laboratory Evolution of Antimicrobial Resistance in Bacteria to ...
    Jan 18, 2024 · This review article introduces the methodologies employed in the laboratory evolution of AMR in bacteria and presents recent discoveries concerning AMR ...
  65. [65]
    High-throughput laboratory evolution reveals evolutionary ... - Nature
    Nov 24, 2020 · Understanding the constraints that shape the evolution of antibiotic resistance is critical for predicting and controlling drug resistance.
  66. [66]
    [PDF] Effects of Natural Selection on Finch Beak Size - HHMI BioInteractive
    The mean is just above 8.8 mm and the mode is 8.8 mm. The data appears to be normally distributed.<|separator|>
  67. [67]
    Gene behind 'evolution in action' in Darwin's finches identified
    Apr 21, 2016 · Environmental change coupled with the gene HMGA2 drove the rapid evolution of a smaller overall beak size in the medium ground finch (Geospiza ...
  68. [68]
    Evolution of Darwin's finches caused by a rare climatic event
    We show that Darwin's finches on a Galapagos island underwent two evolutionary changes after a severe El Nino event caused changes in their food supply.
  69. [69]
    The peppered moth and industrial melanism: evolution of a ... - Nature
    Dec 5, 2012 · The peppered moth was the most diagrammatic example of the phenomenon of industrial melanism that came to be recognised in industrial and smoke- ...
  70. [70]
    Evolution of stickleback in 50 years on earthquake-uplifted islands
    Our deep population genomic data clearly support a recent oceanic origin and subsequent adaptive differentiation of resident freshwater stickleback in post-1964 ...
  71. [71]
    The genomic basis of adaptive evolution in threespine sticklebacks
    Apr 4, 2012 · After the retreat of Pleistocene glaciers, marine sticklebacks colonized and adapted to many newly formed freshwater habitats, evolving repeated ...<|control11|><|separator|>
  72. [72]
    Population Structure Limits Parallel Evolution in Sticklebacks
    May 6, 2021 · Population genetic theory predicts that small effective population sizes (N e ) and restricted gene flow limit the potential for local adaptation.
  73. [73]
    Evolution of lactase persistence: an example of human niche ...
    However, some humans continue to express lactase throughout adult life, and are thus able to digest the lactose found in fresh milk. This trait is called LP.
  74. [74]
    The evolutionary tale of lactase persistence in humans - Nature
    Sep 25, 2023 · The historical understanding of lactase persistence in humans is closely tied to the advent of dairy farming.
  75. [75]
    How Humans' Ability to Digest Milk Evolved from Famine and Disease
    Jul 27, 2022 · Researchers think they know why: lactose tolerance was beneficial enough to influence evolution only during occasional episodes of famine and disease.
  76. [76]
    Global distribution of the sickle cell gene and geographical ...
    Nov 2, 2010 · Sixty years ago it was suggested that the sickle cell disease mutation survives because the heterozygous genotype confers resistance to malaria ...
  77. [77]
    Malaria protection due to sickle haemoglobin depends on parasite ...
    Dec 9, 2021 · The alleles are in strong linkage disequilibrium and have frequencies that covary with the frequency of HbS across populations, in particular ...
  78. [78]
    Malaria continues to select for sickle cell trait in Central Africa - PMC
    falciparum malaria continues to exert strong selective pressure in favor of the sickle cell allele.
  79. [79]
    Altitude adaptation in Tibetans caused by introgression of ... - Nature
    Jul 2, 2014 · A hypoxia pathway gene, EPAS1, was previously identified as having the most extreme signature of positive selection in Tibetans4,5,6,7,8,9,10, ...
  80. [80]
    Genetic signatures of high-altitude adaptation in Tibetans - PNAS
    Apr 3, 2017 · The EPAS1 gene, which encodes the hypoxia inducible factor-2α (HIF-2α) subunit of HIF complex, is a transcription factor involved in body ...
  81. [81]
    Metabolic insight into mechanisms of high-altitude adaptation ... - NIH
    Recent studies have identified genes involved in high-altitude adaptation in Tibetans. Genetic ... Hypoxia: Adapting to High Altitude by Mutating EPAS-1, the Gene ...
  82. [82]
    Microevolutionary processes impact macroevolutionary patterns
    Aug 10, 2018 · Macroevolutionary modeling of species diversification plays important roles in inferring large-scale biodiversity patterns.
  83. [83]
    Conceptual and empirical bridges between micro- and macroevolution
    Jul 10, 2023 · The paper argues that much work remains to identify links between micro and macroevolution, and how mechanisms at one scale translate to ...
  84. [84]
    Linking population‐level and microevolutionary processes to ...
    Our intention here is to highlight some interesting ideas and mechanisms that could help us to better link microevolutionary mechanisms to macroevolutionary ...
  85. [85]
    Bridging micro and macroevolution: insights from chromosomal ...
    This review focuses on chromosomal evolution in plants, especially angiosperms, from both micro- and macroevolution perspectives. It spans processes at the ...
  86. [86]
    Variable success in linking micro- and macroevolution | Oxford
    Abstract. Attempts to predict macroevolution from microevolution, and microevolution from macroevolution, when natural selection is the main cause have met.
  87. [87]
    Linking micro and macroevolution in the presence of migration
    Feb 7, 2020 · Decreasing migration between subpopulations captures divergence between lineages, linking micro with macroevolution. This is modeled using an ...
  88. [88]
    The Scientific Case Against Evolution
    The fact that macroevolution (as distinct from microevolution) has never been observed would seem to exclude it from the domain of true science.
  89. [89]
    The Edge of Evolution | Michael J. Behe
    The “edge” of evolution, a line that defines the border between random and nonrandom mutation, lies very far from where Darwin pointed. Behe argues convincingly ...
  90. [90]
    Review Michael Behe Edge of Evolution
    May 22, 2009 · Behe comes to the conclusion that the edge of Darwinian evolution for a vertebrate lies somewhere between the level of species and class. That ...Missing: summary | Show results with:summary
  91. [91]
    Introduction and Responses to Criticism of Irreducible Complexity
    Feb 20, 2006 · In Darwin's Black Box (1996), Lehigh University biochemist Michael Behe proposed that many of these molecular machines exhibit irreducible complexity.<|control11|><|separator|>
  92. [92]
    The Limits of Evolution | Intelligent Design
    Apr 15, 2020 · Professor Michael Behe invites us to review the sweep of his argument for intelligent design, as he has presented it in his books and other publications.
  93. [93]
    Information theory, evolutionary innovations and evolvability - Journals
    Oct 23, 2017 · Evolutionary biologists have a long-standing interest in information theory, because it is ultimately information encoded in DNA that renders ...Abstract · Introduction · Results · Discussion
  94. [94]
    An introduction to microevolution: rate, pattern, process | Genetica
    Filipchenko, I.A., 1927. Variabilität und variation (Variability and Variation). Gebrüder Borntraeger, Berlin. Google Scholar. Filipchenko, I.A., 1929.
  95. [95]
    Macroevolution FAQ
    The terms macroevolution and microevolution were first coined in 1927 by the Russian entomologist Iurii Filipchenko (or Philipchenko, depending on the ...<|separator|>
  96. [96]
    [PDF] Macro and Micro Evolution
    Both of these terms were coined by Yuri Filipchenko, a Russian entomolo- gist and mentor to the renowned geneticist and evolutionary biologist Theodosius ...
  97. [97]
    5. Macroevolution - Digital Atlas of Ancient Life
    ... Yuri Filipchenko (1927) in “Variabilität und Variation.” Photograph of Filipchenko and the front cover of his work Variabilitat und Variation. Left ...
  98. [98]
    The Modern Synthesis - Darwin
    The modern synthesis united geneticists, naturalists and palaeontologists for the first time and laid the foundations for the explosion of research into ...
  99. [99]
    Darwin and Genetics - PMC - PubMed Central - NIH
    We review the interaction between evolution and genetics, showing how, unlike Mendel, Darwin's lack of a model of the mechanism of inheritance left him unable ...
  100. [100]
    Population genetics - ScienceDirect.com
    Darwin's theory of natural selection succeeded because it was the first to propose an evolutionary mechanism that invokes population genetics: genotypes that ...
  101. [101]
    The Modern Evolutionary Synthesis · 150 Years of On the Origin of ...
    Some of the principal scientists who contributed significantly to the Modern Evolutionary Synthesis were Theodosius Dobzhansky, Ernst Mayr, George Gaylord ...
  102. [102]
    Modern Synthesis - an overview | ScienceDirect Topics
    Dobzhansky's (1937) seminal Genetics and the Origin of Species was followed by important synthetic works by the zoologists Mayr (1942; 1904–2004) and Rensch ( ...
  103. [103]
    Timeline of evolutionary theory
    Julian Huxley (1887-1975) publishes Evolution: The Modern Synthesis. 1942. Ernst Mayr. Ernst Mayr (1904-2005) publishes Systematics and the Origin of Species ...
  104. [104]
    Origins and Evolution of Antibiotic Resistance - PMC - PubMed Central
    This review presents the salient aspects of antibiotic resistance development over the past half-century, with the oft-restated conclusion that it is time to ...
  105. [105]
    Mechanisms of Antibiotic Resistance | Microbiology Spectrum
    From an evolutionary perspective, bacteria use two major genetic strategies to adapt to the antibiotic “attack”: (i) mutations in gene(s) often associated with ...
  106. [106]
    A Brief History of Antimicrobial Resistance - AMA Journal of Ethics
    Antimicrobial resistance has dogged infectious disease treatment processes since the first modern antimicrobials were discovered.Abstract · Antimicrobials In The Modern... · Mycobacteria, Retroviruses...
  107. [107]
    Insecticide Resistance: Challenge to Pest Management and Basic ...
    The most important resistance mechanisms are enhancement of the capacity to metabolically detoxify insecticides and alterations in target sites.
  108. [108]
    Insecticide Resistance: Overview and Management | Land-Grant Press
    May 19, 2020 · Some of the most common mechanisms of insecticide resistance include enhanced metabolic processing of toxins, target site alterations, and ...
  109. [109]
    Insights into insecticide-resistance mechanisms in invasive species
    There is various mechanism mediating insecticide resistance development in insects. The major factors are behavioral resistance, fitness cost, reduced ...
  110. [110]
    Mechanisms of evolved herbicide resistance - PMC - NIH
    For example, enzymes involved in herbicide metabolism–based resistances include cytochromes P450, GSH S-transferases, glucosyl and other transferases, aryl ...
  111. [111]
    [PDF] Deciphering the evolution of herbicide resistance in weeds
    This golden age of herbicides was quickly cut short, however, by the detection of the first herbicide-resistant weeds in 1957 [5]. Today, herbicide resistance ...
  112. [112]
    Evolutionary and ecological insights from herbicide‐resistant weeds ...
    Feb 1, 2019 · The evolution of herbicide resistance in crop weeds presents one of the greatest challenges to agriculture and the production of food.
  113. [113]
    Tackling the emerging threat of antifungal resistance to human health
    Mar 29, 2022 · For example, echinocandin resistance is more common in individuals previously treated with echinocandins, and azole-resistant genotypes of ...
  114. [114]
    Emerging Antifungal Resistance in Fungal Pathogens - PMC
    Mar 18, 2024 · Novel resistant variants of fungal pathogens that were previously susceptible are evolving (such as Aspergillus fumigatus) as well as newly ...
  115. [115]
    Combating Antifungal Resistance - American Society for Microbiology
    Nov 30, 2022 · Resistant fungi include species of Candida. Azoles, such as fluconazole, block metabolism by preventing ergosterol synthesis. Resistant fungi ...
  116. [116]
    Using genomic prediction to detect microevolutionary change ... - NIH
    May 11, 2022 · One way to detect microevolution in a population is to estimate each individual's 'genetic merit' or breeding value, and test whether ...
  117. [117]
    Predicting evolutionary change at the DNA level in a natural ...
    We use fitness estimates to calibrate population genetic models that effectively predict allele frequency changes into the next generation.
  118. [118]
    Predictive ability of multi-population genomic prediction methods of ...
    Jun 26, 2024 · Our study demonstrates that ML algorithms can achieve better prediction performance than multitrait GBLUP models in multi-population genomic ...
  119. [119]
    The utility of genomic prediction models in evolutionary genetics
    Aug 4, 2021 · Based on this principle, genomic prediction methods are used to make predictions of breeding value for an individual using genome-wide molecular ...
  120. [120]
    Towards evolutionary predictions: Current promises and challenges
    Dec 9, 2022 · Many genetic factors can influence the predictability of evolution, and here, we discuss three of them, namely mutation bias, mutational supply ...
  121. [121]
    [PDF] MICROEVOLUTION AND THE GENETIC STRUCTURE OF ...
    theoretical population genetic models are useful because of their predictive value regarding specific features of microevolution, such as the potentially ...<|separator|>
  122. [122]
    The paradox of predictability provides a bridge between micro
    Abstract. The relationship between the evolutionary dynamics observed in contemporary populations (microevolution) and evolution on timescales of millions.
  123. [123]
    Conservation - Understanding Evolution - UC Berkeley
    Efforts to conserve threatened trout populations are helped by considering both the past and future evolution of the populations.Missing: implications | Show results with:implications
  124. [124]
    Slower tempo of microevolution in island birds - PubMed
    A more rapid evolutionary tempo in larger areas has important ramifications for biodiversity conservation.
  125. [125]
    [PDF] Managing Microevolution: Restoration in the Face of Global Change ...
    A more complete understanding of the role of evo- lution in shaping populations and species will help conser- vation biologists and restoration ecologists make.
  126. [126]
    Connecting research and practice to enhance the evolutionary ...
    Jan 3, 2023 · Common-garden, animal-model, or experimental-evolution investigations can indicate whether a species may be able to adapt to climate change.
  127. [127]
    Evolutionary Pathways and Trajectories in Antibiotic Resistance
    Jun 30, 2021 · Evolutionary trajectories of antibiotic resistance find their way in these changing landscapes subjected to random variations, becoming highly entropic and ...
  128. [128]
    Adaptive Laboratory Evolution of Antibiotic Resistance Using ...
    Antibiotic resistance is a global threat to human health, wherefore it is crucial to study the mechanisms of antibiotic resistance as well as its emergence ...Missing: microevolution | Show results with:microevolution
  129. [129]
    The post-vaccine microevolution of invasive Streptococcus ... - Nature
    Oct 23, 2015 · The 7-valent pneumococcal conjugated vaccine (PCV7) has affected the genetic population of Streptococcus pneumoniae in pediatric carriage.
  130. [130]
    Within-host microevolution of Streptococcus pneumoniae is rapid ...
    Jul 10, 2020 · Our findings suggest that within-host microevolution is rapid and adaptive during natural colonisation.Results · Frequently Mutated Genes And... · Genetic Similarity Between...
  131. [131]
    Pre- and postantibiotic epoch: The historical spread of antimicrobial ...
    Sep 25, 2025 · Modern plasmids have evolved through complex microevolution and fusion events into a distinct group of highly recombinogenic, multi-replicon, ...Abstract · Discussion · Materials And Methods...