Fact-checked by Grok 2 weeks ago

Insect flight

Insect flight is the powered aerial locomotion achieved by winged in the superorder , representing a pivotal evolutionary that originated approximately 400 million years ago in the period and facilitated the diversification of over one million extant across diverse ecosystems. This capability relies on specialized chitinous wings, asynchronous flight muscles, and intricate neural control systems that enable maneuvers ranging from hovering to rapid evasion, distinguishing as the most aerially dominant group of animals. Unlike vertebrate flight, insect wings operate through high-frequency flapping—often exceeding 100 Hz in smaller —generating lift and thrust via unsteady aerodynamics rather than fixed-wing gliding. The evolution of insect flight traces back to the , with estimates placing its emergence between 419 and 393 million years ago, predating the oldest known winged s from the Early around 325 million years ago. Key theories on wing origins include the paranotal hypothesis, positing that wings developed as extensions of thoracic terga (dorsal plates) for before powered flight, supported by evidence from primitive Palaeoptera like dragonflies; and the gill hypothesis, suggesting derivation from aquatic gill-like structures in ancestors, evidenced by genetic similarities in development. The split into Palaeoptera (non-folding wings) and (foldable wings with for elasticity) estimated during the late to , around 350-400 million years ago, spurred explosive diversification, with flight enabling habitat expansion, predator escape, and resource exploitation, though secondary flightlessness has evolved repeatedly in stable environments. Insect wings consist of thin, veined membranes that provide structural integrity while allowing passive deformation for enhanced aerodynamics, with veins forming a hierarchical network that boosts fracture toughness and flexibility. Power is supplied by two primary muscle types: synchronous muscles in larger, basal insects like dragonflies, which contract once per neural impulse for precise control; and asynchronous (myogenic) muscles in advanced orders, which oscillate at frequencies up to 200 Hz with minimal neural input, achieving energetic efficiency through stretch-activation mechanisms. Wing kinematics involve bidirectional strokes—upstroke and downstroke—modulated by small steering muscles at the wing base, with body size influencing frequency: smaller insects compensate for low Reynolds numbers by increasing flap rates, while larger ones rely on broader wings. Passive elements, such as elastic resilin in hinges and bristles for wing coupling in moths and bees, further optimize stability and power storage during flight. Aerodynamically, insect flight exploits unsteady flow phenomena to generate forces at low speeds and small scales, where traditional fixed-wing principles falter due to viscous effects. A primary mechanism is the leading-edge vortex (LEV), a stable, low-pressure swirl that forms on the wing during translation and remains attached to augment by up to 50% at high angles of attack, as observed in fruit flies and hawkmoths. Additional contributions come from the clap-and-fling motion, where wings clap together at stroke reversal to fling apart, creating circulatory flow for enhanced circulation; rotational forces during pronation/supination; and wake capture from prior strokes. Wing flexibility passively induces and , improving efficiency by 10–20% and aiding disturbance recovery within 1–2 wingbeats, as seen in hoverflies and . Flight control integrates sensory inputs from compound eyes, antennae, and halteres (gyroscopic organs in flies) via the brain's central complex, enabling real-time adjustments for navigation, obstacle avoidance, and stability amid gusts. Ecologically, insect flight underpins pollination, predation, and nutrient cycling, with its loss in groups like ants and termites reflecting trade-offs for ground-based lifestyles, yet its innovation remains central to insects comprising over half of all known biodiversity.

Anatomy of Insect Wings and Flight Muscles

Wing Structure and Venation

Insect wings are composed of a thin, flexible supported by a framework of tubular . The membrane primarily consists of an epicuticle layer, with varying presence of exocuticle and mesocuticle in species such as those in , while is notably absent from the membrane in locusts but present in vein cuticles across many taxa. form multilayered cuticles (up to six layers) containing and , serving as hollow tubes that transport for nutrient delivery, oxygenation, and waste removal to the living wing tissues. These also house tracheae for oxygen supply and nerves for sensory functions, with flow patterns including looped circulation in crane flies or oscillatory flow in butterflies, driven by thoracic pulsatile organs and vein elasticity. Wing attachment occurs via regions, often reinforced by flexion lines that act as one-way hinges promoting controlled and twisting during deployment. The typically features denser venation for enhanced rigidity, while the trailing edge is thinner and more flexible, allowing dynamic changes of up to 12% in species like hoverflies. Wing venation follows a primitive pattern of six symmetrical pairs of veins—costa (C), subcosta (Sc), (R), media (M), cubitus (Cu), and anal (A)—each with a anterior and concave posterior arising from blood sinuses. In archaic Paleoptera (e.g., ), venation is retained in a straight series with wings held extended, whereas exhibit a derived V-shaped enabling wing folding over the body. This venation network functions as a , with tubular struts optimizing -to-weight ratios by resisting cracks and forming corrugation vertices that enhance structural integrity without excessive mass. Spanwise flexural scales with the cube of wing due to leading-edge veins, achieving 1–2 orders of magnitude greater rigidity than chordwise bending across diverse . Venation patterns vary markedly across insect orders to suit ecological roles. In Coleoptera (), forewings are thickened into rigid elytra up to 1 mm thick for protection, while hindwings remain membranous and folded with aids at hinges. (dragonflies and damselflies) possess flexible, reticulate-veined wings with extensive secondary cross-veins forming net-like patterns over areas of 20–725 mm². In Diptera (flies), hindwings are modified into clubbed —stiff basal structures serving as mechanosensory organs rather than flight surfaces. Surface features further adapt wings for airflow interaction. Scales on wings create rough textures that generate micro-turbulence, reducing and enhancing through vortex formation, as seen in the peacock butterfly (Inachis io). Hairs on wings produce micro-vortices to improve and by altering local flow. Corrugations, prominent in leading edges, increase spanwise stiffness while modifying airflow by trapping vortices in profile valleys, thereby boosting production without compromising flexibility.

Direct and Indirect Flight Muscles

Insect flight muscles are classified into direct and indirect types based on their anatomical attachments and functional roles in wing movement. Direct flight muscles attach directly to the bases of the wings, typically via sclerites or ligaments, enabling precise control over individual wings. These muscles are prevalent in more primitive insect orders, such as (dragonflies and damselflies) and certain small insects like (), where they allow for independent manipulation of each wing for maneuvers like hovering or rapid directional changes. In these insects, key direct muscles include tergosternal (dorsoventral) and pleurosternal muscles that connect to wing base elements, such as the axillary sclerites, to drive downstrokes and upstrokes. Direct muscles operate synchronously, with each contraction triggered by a single from motor neurons, limiting wingbeat frequencies to typically less than 100 Hz. In contrast, indirect flight muscles do not attach to the wings but instead connect to the of the , deforming its structure to indirectly transmit motion to the wings through notal and pleural sclerites. These muscles are dominant in more advanced insect orders, such as Diptera (flies), (bees), and Coleoptera (beetles), facilitating high-power output for sustained flight. The primary indirect muscles consist of dorsal-ventral (tergosternal) muscles, which extend from the dorsal notum to the ventral and elevate the wings during , and dorso-longitudinal muscles, which run along the length of the between anterior and posterior apodemes and depress the wings by arching the notum. This antagonistic pairing causes oscillatory deformation of the thorax, with up to 85% of the energy for wing downstrokes stored as elastic potential in the sclerites. A key distinction between direct and indirect muscles lies in their contraction mechanisms: synchronous versus asynchronous. Synchronous contraction, characteristic of direct muscles in basal insects like Odonata, involves one muscle twitch per neural , with calcium release and tightly coupled to each cycle for controlled but lower-frequency motion. Asynchronous contraction, typical of indirect muscles in higher , allows multiple wingbeat cycles per through stretch , where mechanical stretch of the muscle during the opposing phase triggers delayed force enhancement without repeated neural input. In asynchronous systems, impulses occur at low frequencies (e.g., 3–5.5 Hz in fruit flies), yet wingbeat frequencies reach 185–195 Hz due to resonant thoracic oscillations and sustained calcium levels in the . This mechanism enables wingbeat rates exceeding 100 Hz, up to 1,000 Hz in small like midges, far surpassing synchronous limits. Asynchronous indirect muscles feature specialized fibrillar fiber types, where myofibrils are loosely packed with abundant mitochondria (e.g., 43% in flight muscles) and a highly ordered, near-crystalline arrangement of sarcomeres to support rapid oscillations. These fibrillar muscles contain C-filaments linking filaments to Z-lines, enhancing stretch sensitivity and minimizing energy loss from calcium cycling, which contrasts with the more compact, cross-striated fibers in synchronous muscles. This structural adaptation allows for high-frequency performance while maintaining efficiency, with myofibrillar volume comprising about 53% of the muscle in asynchronous types.

Aerodynamic Principles

Direct Flight Mechanics

Direct flight mechanics in involve the direct attachment of flight muscles to the wing bases, allowing for synchronous contraction that pulls the wings through their strokes. This mechanism is characteristic of more primitive insect orders, where muscles contract once per neural impulse, enabling precise control over wing motion. In dragonflies (), for instance, these direct muscles facilitate independent movement of the fore and hind wings, which are not coupled as in more derived . The of typically produce complex trajectories, such as figure-eight paths traced by the wing tips during . In dragonflies, wing motion involves a downstroke and upstroke in a near-vertical plane, with the figure-eight pattern arising from the and feathering of the wings at stroke reversal, optimizing generation. This path allows for effective hovering and rapid directional changes, as the wings can adjust amplitude, phase, and independently on each of the four wings. Force generation in direct flight relies on synchronous muscle activation, where each wingbeat corresponds to a single neural signal to the muscles, matching the wingbeat frequency. In , this frequency typically ranges from 20 to 40 Hz, varying with body size and temperature, lower than in asynchronous systems due to the direct neural pacing. Neural control occurs via motor neurons that innervate the muscles for each , providing fine-tuned adjustments for maneuvers. This synchronous direct control offers advantages in precise maneuvering and hovering, as seen in the agile flight of dragonflies, which can execute sharp turns and stationary positions with independent wing actions. However, it limits efficiency and maximum frequency compared to indirect mechanisms, as the slower contraction rates constrain power output for sustained high-speed flight. In Ephemeroptera (mayflies), direct flight similarly enables short, erratic flights with uncoupled fore and hind wings, though at frequencies around 25 Hz, suited to their brief adult lifespans. The wing structure, with muscles attaching directly to the base via sclerites, supports this mechanics by allowing pivotal motion without thoracic deformation.

Indirect Flight Mechanics

Indirect flight mechanics in involve specialized muscles that do not attach directly to the wings but instead deform the thoracic , causing the wings to pivot on hinge-like articulations at their bases. This system, prevalent in advanced flying , relies on antagonistic pairs of indirect flight muscles: the dorsal longitudinal muscles (DLM), which run along the length of the , and the dorsoventral muscles (DVM), which span vertically. When the DLM contract, they compress the longitudinally, bowing the notum downward; conversely, DVM elevates the notum, expanding the vertically. These deformations are amplified through the rigid yet elastic thoracic structure, transmitting forces to the wing hinges without direct muscular control, enabling rapid oscillations. A hallmark of indirect flight is the asynchronous operation of these muscles, where wingbeat frequency decouples from neural firing rates due to stretch activation. In this , sustained low-frequency neural impulses maintain calcium levels in the muscle, but actual contractions are triggered by mechanical stretch from the antagonistic muscle pair, leading to delayed force enhancement and self-sustained oscillations. This allows one neural impulse to drive multiple contraction cycles—often 3 to 100 or more—facilitating wingbeat frequencies far exceeding neural limits, such as 185–195 Hz in fruit flies () and up to 1,000 Hz in midges. In contrast, primitive muscles operate synchronously, with contractions tied 1:1 to neural impulses at lower frequencies. The 's architecture is crucial for force transmission in indirect flight, featuring the notum as the dorsal plate that bows under and the pleura as lateral walls that flex to accommodate deformation. elements, including the rubber-like protein integrated into myofibrillar structures and cuticular regions, provide resilience and energy return, preventing fatigue during high-frequency cycles; for instance, pads at pleural hinges enhance deformability. In Diptera, this setup produces a characteristic "" during rapid thorax buckling, observable in tethered flight preparations. This indirect asynchronous system powers flight in most Endopterygota, including Hymenoptera (e.g., bees with wingbeats around 200–250 Hz) and Lepidoptera (e.g., moths and butterflies achieving 10–50 Hz but with efficient thorax leverage for sustained hovering). These insects, comprising over 90% of flying species, evolved this mechanism to support agile maneuvers and high power output, distinct from the synchronous direct muscles in more basal orders like Odonata.

Leading Edge Vortex

The leading edge vortex (LEV) is a prominent unsteady aerodynamic feature in insect flight, characterized by a stable, low-pressure vortex that forms and remains attached to the dorsal surface of the wing during the translational phase of the flapping stroke, particularly at high angles of attack exceeding 45 degrees. This vortex generates substantial lift by creating a region of low pressure above the wing, enabling insects to achieve forces far beyond those predicted by steady-state aerodynamics. The LEV was first visualized experimentally in the hawkmoth Manduca sexta using smoke-wire flow visualization on a dynamically scaled mechanical flapper, revealing an intense vortex on the downstroke sufficient to account for observed high-lift production. Formation of the LEV occurs primarily during the downstroke and upstroke translations, where the wing's rapid motion separates at the , initiating dynamic —a process where the vortex detaches from the but stabilizes due to the wing's three-dimensional and spanwise . Unlike two-dimensional airfoils, where such vortices shed periodically and cause , the insect LEV grows to a constant size without bursting, sustained by a spanwise component comparable to the speed that transports outward toward the wingtip, mimicking the conical vortex on delta wings. This stability is further reinforced by the spanwise and the absence of additional generation at the trailing edge, adhering to a modified that prevents premature shedding. The LEV contributes substantially to the total during translation in model experiments simulating () kinematics, with peak circulatory forces enhanced by up to 70% over quasi-steady estimates. In hovering flight, this mechanism allows to maintain equilibrium by balancing weight against mean , as demonstrated in robotic wings mimicking strokes where LEV attachment persisted across Reynolds numbers typical of small (100–10,000). The vortex's persistence is conserved across diverse taxa, from hawkmoths to hoverflies, despite variations in wing planform, underscoring its generality in flapping at low Reynolds numbers. Experimental validation through () on tethered insects and computational models confirms that LEV strength scales with wing velocity and , but its attachment is modulated by wing flexibility, which can delay vortex bursting and enhance endurance. For instance, in with flexible wings, the LEV remains stable longer than on rigid models, contributing to efficient transitions. Overall, the LEV exemplifies how unsteady effects dominate insect flight, integrating with rotational and wake-capture mechanisms to enable agile locomotion.

Clap and Fling

The clap and fling mechanism enables small to generate impulsive circulatory during the reversal between upstroke and downstroke phases of the wingbeat cycle. First identified through high-speed in the parasitic wasp , the process begins with the wings clapping together at the dorsal end of the upstroke, expelling air from between them and aligning their leading edges. As the downstroke initiates, the wings fling apart while remaining in close proximity near the trailing edges, forming a narrow slot that creates a low-pressure region and promotes rapid circulation around each wing, enhancing through delayed stall of the vortex. In with flexible wings, the fling employs a peeling mode, where separation propagates from the trailing edge toward the , rather than an abrupt rigid separation. This peeling delays the shedding of trailing edge vortices, preventing their interaction and cancellation with the beneficial leading edge vortices, thereby sustaining higher over a longer duration and reducing by up to 50% compared to rigid-wing flings. Computational simulations at low Reynolds numbers (Re ≈ 10) demonstrate that this mode maintains vortical asymmetry, crucial for efficient force production in tiny like . High-speed imaging of free-flying reveals the clap and fling occurring over approximately 15% of the wingbeat cycle, providing a significant enhancement to at stroke reversal through amplified peak force. The of the fling can be modeled as the separation of two closely spaced airfoils, yielding enhanced circulation relative to isolated wing translation and doubling in theoretical models.

Flight Dynamics and Performance

Hovering

Hovering represents a specialized form of insect flight characterized by zero net forward velocity, enabling stationary positioning in the air for tasks such as foraging or mating. In many insects, such as bees, this is achieved through symmetric wing kinematics involving near-horizontal strokes in the stroke plane, with the wings undergoing pronation—rotation such that the leading edge moves downward—at the beginning of the downstroke and supination—leading edge upward—at the start of the upstroke. These rotational adjustments allow the wings to maintain a positive angle of attack throughout both the downstroke and upstroke, generating approximately equal lift during each half-stroke to balance body weight without translational thrust. For instance, honeybees employ relatively low stroke amplitudes of about 90° combined with high wingbeat frequencies around 230 Hz to sustain this balanced lift production. Stability during hovering requires precise active control to counteract perturbations like body s from asymmetric forces or wind gusts. Insects achieve this by dynamically modulating stroke amplitude, which primarily influences vertical force, and feathering angles (), which adjust both and . These rapid adjustments, often on the scale of wingbeat cycles, enable corrections for , roll, and yaw instabilities; for example, increasing stroke amplitude on one side can generate counter-s to stabilize . Such control mechanisms ensure long-duration hovering, with the leading-edge vortex playing a key role in augmenting efficiency across strokes. The metabolic demands of hovering are exceptionally high due to the continuous power output required for static lift, often exceeding resting rates by 50–100 times in many species. In hummingbird hawkmoths (Macroglossum stellatarum), hovering elevates metabolic rates up to 170 times the resting level, reflecting the intense aerobic demands fueled primarily by oxidation. Similarly, hoverflies (Syrphidae family, e.g., ) exhibit flight metabolic rates around 100 times resting, underscoring the energy-intensive nature of sustained hovering for feeding or prey inspection. These costs highlight the evolutionary premium on efficient power management in hovering specialists. Hovering variations include normal and inverted orientations, adapting to ecological needs like obstacle navigation. In normal hovering, typical of bees, hawkmoths, and dragonflies, wings beat in a horizontal plane with the body horizontal to produce downward lift. Dragonflies (Odonata) employ normal hovering, with wings positioned laterally and strokes in a near-horizontal plane to generate downward lift, allowing enhanced maneuverability. The induced power required for hovering, which represents the minimum aerodynamic power to sustain lift, is given by P_\text{ind} = \frac{L^{3/2}}{\sqrt{2 \rho A}} where L is the lift (equal to body weight), \rho is air density, and A is the effective disc area swept by the wings; this formulation illustrates how smaller insects incur disproportionately higher specific power costs due to reduced A.

Forward and Maneuvering Flight

In forward flight, insects transition from the near-horizontal stroke plane typical of hovering to a more forward-inclined orientation, which redirects aerodynamic forces to produce net thrust while sustaining lift. This adjustment enhances translational velocity by aligning the wing motion with the direction of travel, allowing insects to achieve efficient progression through the air. For instance, dragonflies (order Odonata) can reach forward speeds of up to 10 m/s, facilitated by this stroke plane reconfiguration and powerful wing beats that exploit unsteady aerodynamics like the leading-edge vortex. To minimize energy expenditure during forward motion, insects pitch their bodies downward into a streamlined, more horizontal posture, reducing the projected frontal area and thereby lowering body . This body alignment optimizes the balance between , , and , particularly at moderate to high speeds where induced drag from hovering diminishes but parasite drag becomes prominent. In flies such as bluebottles (Calliphora spp.), forward speed correlates directly with this angle, enabling sustained cruising while countering gravitational and aerodynamic loads. Maneuvering in forward flight relies on asymmetric wing kinematics to generate torques for yaw, pitch, and roll, allowing rapid directional changes without halting translation. Insects achieve yaw and roll by differentially altering stroke amplitude or deviation between contralateral wings, creating unbalanced forces that induce rotation around the body's vertical or longitudinal axes. For pitch control, symmetric shifts in stroke position relative to the body's center of mass adjust the vertical force vector. In fruit flies (), these asymmetries enable saccadic turns with angular velocities exceeding 1000°/s, demonstrating the precision of such mechanisms. Recent whole-body physics simulations (as of 2025) confirm the role of these angular velocities in turns and stability. Flies further stabilize maneuvers using , modified hindwings that function as gyroscopes to detect and compensate for rotational perturbations during turns. During sharp turns, insects counteract centrifugal forces through dynamic wing twisting and feathering, which modulate local angles of attack to produce corrective lateral forces and enable accelerations up to 5 . This twisting allows rapid adjustments in aerodynamic , preventing and maintaining control at high angular rates. For efficiency in prolonged forward and maneuvering flight, such as during , locusts (Locusta migratoria) align their bodies to minimize drag, achieving lift-to-drag ratios of approximately 1.7:1 in phases interspersed with , which extends range over hundreds of kilometers.

Governing Equations

The governing equation for aerodynamic in insect flight follows the conventional form adapted to the flapping kinematics: L = \frac{1}{2} \rho U^2 S C_L where \rho denotes air density, U the between and surrounding air (typically the vector sum of flapping and induced velocities), S the effective area, and C_L the lift coefficient, which varies with , (typically 10^3–10^4 for ), and unsteady effects. In unsteady flow regimes prevalent during rapid oscillations, this equation is augmented by contributions from and rotational mechanisms to account for inertial and circulatory forces. The term arises from the virtual mass of air accelerated normal to the surface, yielding a force L_{am} \propto \rho c^2 l \dot{v}_n, where c is , l spanwise element , and \dot{v}_n the normal ; this provides peak during stroke initiation. Rotational , prominent at stroke reversals, stems from bound circulation generated by pitching, approximated as L_{rot} = \rho \Gamma U c, with \Gamma the rotational circulation scaling with \dot{\theta} and ; together, these unsteady terms enable C_L values up to 2 or higher, far exceeding steady-state limits. Drag in insect flight is quantified analogously as D = \frac{1}{2} \rho U^2 S C_D with C_D the drag coefficient encompassing multiple components: profile drag from viscous shear and pressure differences on the wing surface, induced drag from trailing vortices associated with lift generation, and parasitic drag from bluff body effects on wings and fuselage. Profile drag predominates during slow, high-angle-of-attack translations, scaling with C_D \approx 2 \sin^2 \alpha + C_{f}, where \alpha is angle of attack and C_f skin friction; induced drag follows C_{D_i} = C_L^2 / (\pi AR e), with AR aspect ratio and e span efficiency, becoming critical in low-speed regimes like hovering. Parasitic contributions are minor but increase with body size and speed. These components collectively determine energy dissipation, with total C_D often 0.5–1.5 in flapping cycles. The mechanical power output required for sustained flight integrates these forces over the wing motion, expressed as the sum of aerodynamic power (to counter drag during translation), induced power (to sustain vortical lift), and profile power (inertial costs of flapping). A simplified form is P = P_{aero} + P_{ind} + P_{pro}, where P_{aero} = D \cdot U, P_{ind} \approx (L^{3/2} / (2 \rho A)^{1/2}) / \eta with A the swept area and \eta efficiency (typically 0.7–0.9), and P_{pro} \propto \rho f^2 S^{3/2} b^4 / m, with f stroke frequency, b mean radius, and m mass. Profile power reaches a minimum at an optimal advance ratio (forward speed over flapping speed ≈ 0.3–0.5), balancing reduced induced costs against increased translational drag; in hovering, induced power can comprise 60–80% of total, yielding specific powers of 20–100 W/kg across insect taxa. Blade element theory facilitates computation of these forces by segmenting the wing into independent annular elements along the , assuming local without spanwise interactions. For an element at radius r with width dr, local velocity U_{loc} = \sqrt{( \omega r )^2 + V_f^2 + 2 \omega r V_f \sin \phi} (combining rotational \omega, forward V_f, and feathering angle \phi) yields incremental dL = \frac{1}{2} \rho U_{loc}^2 c(r) dr C_L(\alpha_{loc}) and dD = \frac{1}{2} \rho U_{loc}^2 c(r) dr C_D(\alpha_{loc}), where \alpha_{loc} incorporates geometric and induced angles. Quasi-steady approximations neglect time delays in vortex formation, integrating dL and dD over the and cycle for total forces; this approach, calibrated with empirical C_L(\alpha) and C_D(\alpha) from dynamic data, accurately predicts mean within 10–20% for hover and forward flight in model insects.

Physiological Mechanisms

Power Generation and Output

Most advanced rely on asynchronous (indirect) flight muscles to achieve high wingbeat frequencies through delayed stretch , contrasting with synchronous muscles in basal like dragonflies that contract once per neural impulse for precise control. Asynchronous muscles feature specialized - interactions where cross-bridge cycling is accelerated, with dissociation rates from reaching up to 3,700 s⁻¹ in indirect flight muscles, far exceeding those in . This rapid cycling allows a single to sustain multiple contractions per cycle, powering oscillations at frequencies over 100 Hz in small . In synchronous muscles, power is generated through direct 1:1 of neural impulses to contractions, enabling lower frequencies (e.g., 20–40 Hz in dragonflies) but greater , with cross-bridge adapted for sustained force rather than . Asynchronous systems, however, dominate in derived orders, where power input to the wings is generated via work loops in these muscles, with sinusoidal mimicking wing motion. During the stretch , delayed calcium triggers increased cross-bridge attachment, producing positive net mechanical work that exceeds energy input during shortening. For example, in basalar muscle, work loops under imposed sinusoidal motion yield positive work loops with peak power outputs aligned to the stretch-activated , optimizing energy transfer for flight. Mechanical power output (P_mech) from the muscles to the wings is defined as the product of and , P_mech = F × v, where arises from cross-bridge attachments and from wing kinematics. In flying , this output achieves efficiencies of 20–40%, reflecting the conversion of to useful aerodynamic work, with higher values in larger species due to optimized muscle . Power generation and output are measured using respirometry to estimate metabolic input combined with high-speed kinematic analysis of wing motion to compute mechanical work. For instance, in hovering , respirometry yields a metabolic rate supporting a mechanical power output of approximately 80 W kg⁻¹ muscle mass, sufficient for sustained flight at body weights up to 150% of normal. Wingbeat , a key factor influencing power output, scales allometrically with body mass as f ∝ M^{-1/3}, reflecting geometric similarity and inertial demands in asynchronous fliers. This relation ensures smaller achieve higher frequencies (e.g., >200 Hz in ) to generate adequate , while larger ones operate at lower rates (e.g., ~20–40 Hz in moths) with greater stroke amplitudes. Allometric patterns also link muscle mass (typically 20–40% of body mass) and , constraining maximum body sizes in flying .

Elasticity and Energy Storage

Insect flight relies on elastic structures to store and release efficiently, minimizing the power demands on flight muscles. A key component is , a rubber-like protein discovered in cuticles, particularly concentrated in wing hinges and the thoracic of many flying insects. enables passive energy recycling by deforming during muscle contraction and snapping back to propel wing movements, with nearly perfect elastic recovery that returns approximately 97% of stored . Its material properties include a low of about 1 , allowing high extensibility up to 300% , combined with exceptional fatigue resistance capable of enduring millions of cycles without significant degradation. The primary function of in flight is to facilitate and during reversals, particularly through like the "click" in dipteran , where deformation stores from one and recaptures it for the next. This elastic recapture reduces the net mechanical power required from muscles by 20-30%, enhancing overall flight efficiency in small with high wingbeat frequencies. In with indirect flight muscles, such as flies, integrates with the compliant to amplify these savings, though the core benefit stems from the protein's ability to buffer inertial loads without active muscular input. Representative examples illustrate resilin's role across taxa. In cicadas, which employ muscles, exocuticle regions in the act as composite springs reinforced by , storing to augment wing oscillations and sustain prolonged flight despite their large body size. Similarly, in locusts, elastic tendons containing connect flight muscles to the , enabling rapid transfer and storage that supports powerful takeoffs and sustained forward flight. These structures highlight how optimizes use by combining with stiffer cuticular elements to form hybrid springs tailored to species-specific flight demands.

Wing Coupling

Wing coupling refers to the physical and kinematic mechanisms that synchronize the motion of forewings and hindwings in many , enabling them to function as a single aerodynamic surface during flight to enhance , , and . These linkages minimize aerodynamic between the wings, allowing for coordinated that reduces expenditure while improving performance. In with coupled wings, such as those in the orders and , the fore- and hindwings overlap or interlock mechanically, contrasting with decoupled systems where independent control provides maneuverability. Various physical mechanisms facilitate this coupling. In Lepidoptera, such as butterflies and moths, the jugal lobe—a lobe-like extension at the base of the forewing—overlaps with the humeral angle of the hindwing, creating a stable linkage that synchronizes wing motion. For instance, in the small white butterfly Pieris rapae, the forewing basal area contacts the hindwing humeral area dorsally, allowing synchronous flapping without additional hooks. In Hymenoptera, like bees and wasps, hamuli—rows of hook-like setae along the anterior margin of the hindwing—interlock with a folded posterior margin (plexus) of the forewing, forming a robust, reversible coupling that permits a wide range of wing angles during flight. This structure in honeybees (Apis mellifera), composed of sclerotized cuticle, ensures the wings act as a unified pair, improving aerodynamic performance. In Diptera, such as flies, where hindwings are modified into halteres, spanwise coupling occurs through thoracic mechanical linkages and wing-halter interactions, treating the system as coupled oscillators that maintain phase-locked motion for balance and stability. Kinematically, wing coupling often involves controlled phase differences between fore- and hindwing strokes, which can generate clap-like effects to boost lift without requiring a full wing clap-fling motion. In coupled systems, a small phase lag (e.g., 0–90°) aligns the wings to trap and expel air efficiently, increasing circulatory force and lift by recovering wake energy. For example, in butterflies, synchronous fore- and hindwing motion during the upstroke creates a partial clap, where the wings cup and collide to form a low-pressure air pocket that enhances propulsion by 28% relative to rigid-wing models. Similarly, in bees, hamuli-mediated synchronization maintains near-zero phase difference, allowing the coupled wings to produce clap-like interactions at stroke reversal, which amplifies lift through vortex interactions while minimizing drag. Adaptations in wing coupling reflect diverse flight needs, with decoupling in primitive groups like enabling independent fore- and hindwing control for agile maneuvers, such as rapid turns or hovering, at the cost of some . In contrast, advanced pterygote insects, including those with tight , achieve higher by reducing slip between wings, as seen in where the mechanism supports sustained forward flight and load-carrying. Elastic hinges at the wing base can briefly aid this inter-wing coordination by storing and releasing energy to fine-tune phase alignment.

Biochemical Processes

Insect flight muscles exhibit specialized biochemical pathways to sustain the extraordinarily high rates of required for oscillatory contractions during flight. The primary energy sources are , primarily from and stores within the muscle, which undergo aerobic to produce pyruvate for mitochondrial oxidation via the tricarboxylic acid cycle and . In many , such as honeybees and flies, flight is fueled almost exclusively by carbohydrate oxidation, with enzymes like , , and exhibiting high activities to support rapid flux. oxidation plays a secondary role in short-burst fliers but becomes prominent in long-distance migrants like locusts, where beta-oxidation enzymes in flight muscle mitochondria enable sustained provision from lipid reserves mobilized from the . Although activity is minimal in most flight muscles, preventing significant accumulation, a lactate shuttle mechanism has been observed in certain contexts, such as under hypoxic stress in , where produced in the is transported to mitochondria or other tissues for oxidation, facilitating metabolic flexibility. Myosin isoforms in insect flight muscles are adapted for fast-twitch, high-frequency performance, featuring heavy chain variants with altered lever arm structures that enhance shortening velocity and power output during stretch-activated contractions. In , alternative splicing of the myosin heavy chain generates at least 15 isoforms, including embryonic types that promote rapid cross-bridge and delayed tension development essential for asynchronous flight. Calcium ion (Ca^{2+}) of these interactions occurs via the troponin-tropomyosin complex on thin filaments, where Ca^{2+} binding to induces a conformational change that exposes myosin-binding sites on , enabling contraction at low Ca^{2+} concentrations typical of sustained flight. Specialized troponin isoforms, such as those with enhanced mechanical sensitivity, further optimize this for oscillatory demands. Mitochondria, often termed sarcosomes in insect flight muscle, achieve exceptional densities to match the aerobic demands, comprising up to 40% of the myofibrillar volume in mature asynchronous muscles of species like blowflies and bees. This high packing density supports mass-specific ATP production rates exceeding those in vertebrate muscles by orders of magnitude, with cristae-rich membranes housing dense electron transport chains for efficient oxidative phosphorylation. Hormonal regulation mobilizes these biochemical resources, with serving as the principal "fight-or-flight" in , rapidly elevating cyclic AMP levels in flight muscles to activate and trehalase for and breakdown during takeoff. In locusts, injection prior to flight increases carbohydrate oxidation rates by 50-100%, enhancing readiness through allosteric modulation of key enzymes like , which shifts to an active conformation via cascades. Adrenaline-like effects of also promote in the , supplying fatty acids for extended flight, with feedback via allosteric inhibition of glycolytic enzymes to balance substrate use.

Sensory and Neural Feedback

Insect flight control depends on intricate sensory systems that detect environmental and self-motion cues, feeding into neural circuits for instantaneous adjustments. In Dipteran insects like flies, —modified hindwings—serve as specialized mechanoreceptors that sense Coriolis forces generated by rotations during flight, enabling gyroscopic stabilization. These forces deflect the from their beating plane, with campaniform sensilla and chordotonal organs at the base transducing the signals into neural impulses. eyes provide visual feedback through optic flow, where the apparent motion of the across the informs the insect about translational , altitude, and obstacles, facilitating course corrections in cluttered environments. Additionally, wind-sensitive hairs (trichoid sensilla) distributed on the head, antennae, and detect aerodynamic perturbations and , contributing to the perception of wind direction and attitude. Neural processing occurs primarily in the thoracic ganglia, where (CPGs) establish the rhythmic motor output for wingbeats, typically at frequencies of 10–200 Hz depending on the species. These CPGs, composed of interconnected and motoneurons, generate alternating patterns for antagonistic flight muscles even in isolated preparations, but rely on sensory for fine-tuning amplitude and phase. Feedback loops integrate inputs from multiple modalities via reflex arcs, with latencies as short as 5–10 ms, allowing corrections at rates matching wingbeat frequencies, such as approximately 200 Hz in . For instance, proprioceptive stretch receptors in the wing hinges and monitor wing position and deformation, providing closed-loop control to synchronize contralateral wings and prevent desynchronization during perturbations. In without , such as hawkmoths, antennal mechanoreceptors including function as vestibulo-lateral sensors, detecting rotational accelerations and to stabilize posture during hovering and maneuvering. A classic example is the optomotor response in , where visual optic flow elicits compensatory yaw torques by modulating wingbeat amplitude, with the response gain peaking at spatial frequencies matching natural environments. Experimental of in flies results in immediate loss of rotational stability, causing tumbling and inability to maintain controlled flight, underscoring the indispensability of this mechanosensory input for equilibrium. These systems collectively ensure robust stability, with brief integration into maneuvering behaviors like turns, where sensory fusion prioritizes rapid threat avoidance.

Evolutionary Origins and Adaptations

Overview of Evolutionary Hypotheses

The of insect flight represents a pivotal in history, likely originating around 400 million years ago during the period based on analyses of phylogenomic data from diverse lineages. The earliest definitive evidence of winged dates to the early , approximately 325 million years ago, with precursors such as precoxal leg exites in ancestral arthropods serving as the structural foundation for wing development. These winged precursors enabled the transition to powered flight, marking as the first animals to achieve this capability and facilitating their subsequent radiation across terrestrial ecosystems. Multiple lines of evidence support this timeline and origin. The fossil record, particularly from deposits, includes early pterygote groups like the Palaeodictyoptera, which display primitive wing venation and articulations indicative of flapping flight capabilities. highlights homologies between wings and proximal leg segments in relatives, suggesting a shared exite-based origin for these structures. Genetic markers further illuminate this process; for instance, the vestigial gene, a key regulator of wing patterning, is expressed in presumptive wing fields of apterous insects such as (order ), revealing conserved developmental pathways even in wingless descendants of flying ancestors. Central debates surround the environmental context and multiplicity of flight's emergence among pterygotes—the winged insect clade. Proponents of aquatic origins argue for a "water-up" , where flight structures derived from gill-like appendages in amphibious ancestors, while terrestrial "trees-down" models posit from elevated arboreal positions as the precursor to active flight. Another key contention is whether wings evolved once () or independently in separate pterygote lineages (dual or polyphyletic origins), with recent evo-devo and genomic studies favoring a single derivation from thoracic leg bases. These discussions underscore the challenges in reconciling sparsity with molecular and morphological data. Flight conferred profound adaptive benefits, enabling rapid escape from predators, long-distance dispersal to exploit patchy resources, and aerial displays that enhanced across insect orders. Yet, these advantages came at substantial costs, including elevated metabolic rates—up to 100 times resting levels during sustained flight—and the energetic demands of wing maintenance, which have driven secondary flightlessness in over 20% of extant species in stable or insular habitats.

Epicoxal Hypothesis

The epicoxal hypothesis proposes that insect wings originated from gill-like exites attached to the epicoxa, the proximal segment of the ancestral , particularly on abdominal coxae in aquatic environments. According to this model, these exites functioned initially as respiratory and locomotor flaps in water, later adapting for aerial flapping upon the transition to land. Jarmila Kukalová-Peck introduced this idea in her 1983 analysis of arthropod limb morphology and fossil evidence, arguing that the pro-wing formed from an epicoxal exite mobilized by leg-derived musculature, with the epicoxa incorporating into the pleural body wall to form the wing's articular base. Supporting evidence draws from embryological and morphological homologies, especially in Ephemeroptera (mayflies), where abdominal gills exhibit structural and developmental parallels to wings. In mayfly nymphs, these gills evaginate from the pleural region as movable, muscled appendages, mirroring the embryonic formation of wing buds that separate from leg discs rather than dorsal terga. Furthermore, coxal muscle similarities reinforce this link, as the intrinsic pleural muscles actuating mayfly gills are homologous to the direct flight muscles in adult wings, both tracing to epicoxal leg origins in the arthropod ground plan. Fossil nymphal wing pads from Paleozoic Ephemeroptera, such as those in Protereismatidae, show articulated structures consistent with serial exites derived from abdominal segments. Criticisms of the epicoxal hypothesis center on its limited fossil corroboration and overreliance on an ancestral phase. Direct transitional fossils depicting flaps evolving into thoracic wings are absent, with records showing winged appearing abruptly without clear intermediates from abdominal exites. Additionally, the model is challenged for implying complex wing articulations evolved from simple exites, a process genetically and mechanically demanding, and for emphasizing larvae despite many lineages being fully terrestrial. While it aligns with exopterygote , where wings form externally in nymphs, it contrasts with pleural theories by prioritizing exite appendages over body wall extensions. The hypothesis carries significant implications for understanding exopterygote flight origins, suggesting that powered flight arose from serial modifications of larval aquatic gills into adult thoracic wings during the Devonian-Carboniferous transition. This serial homology posits wings on multiple abdominal segments primitively, with thoracic specialization enabling aerial locomotion while abdominal remnants persisted as gills in basal groups like mayflies. Overall, it supports a monophyletic pterygote origin tied to limb evolution, influencing interpretations of diversification in early terrestrial ecosystems.

Paranotal Hypothesis

The paranotal hypothesis posits that insect wings originated as lateral extensions, or paranotal lobes, of the dorsal thoracic tergum in early terrestrial arthropods, initially serving as passive gliding structures before evolving into actively flapping appendages capable of powered flight. This theory, first articulated in the and widely endorsed by mid-20th-century entomologists, suggests these lobes provided an adaptive advantage for controlled descent from heights, such as trees or elevated vegetation, in a terrestrial . Sir Vincent Wigglesworth reinforced this view in 1973, emphasizing the transition from gliding to flapping as a gradual evolutionary process driven by selection for improved aerial maneuverability. Supporting evidence draws from both fossil and extant forms. Fossil arthropods like , a millipede relative, exhibit pronounced tergal expansions interpreted as paranotal lobes that could have facilitated . In modern apterygote hexapods, such as Collembola (springtails), small vestigial dorsal lobes on the are seen as remnants of these ancestral structures, consistent with a tergal origin restricted to thoracic segments. Biomechanical models further bolster the by demonstrating that paranotal lobes could generate sufficient for short-distance flight. Simulations by Kingsolver and Koehl () analyzed aerodynamic performance across size scales, showing that small, flap-like extensions produce positive and during descent, aligning with the proposed gliding-to-flapping progression. Despite its strengths, the paranotal hypothesis faces criticisms for inadequately accounting for key wing features. It struggles to explain the pleural articulation that enables wing movement in modern insects, as tergal extensions would lack the necessary basal joints for rotation and folding. Similarly, the complex venation patterns in fossil and extant wings, which resemble branched limb structures rather than simple dermal outgrowths, remain difficult to derive from dorsal lobes alone.

Endite-Exite Hypothesis

The Endite-Exite Hypothesis proposes that wings evolved through the of an anterior endite lobe and a posterior exite lobe originating from the pleural region of the , rather than from dorsal tergal expansions. This model, developed as a variant by Kukalová-Peck in her 1987 analysis of , interprets the primitive wing as a biflagellate structure derived from proximal leg podomeres that were incorporated into the body wall and flattened for aerodynamic function. These lobes, movable and articulated, preadapted the structure for flapping flight by leveraging existing limb mechanics. Supporting evidence stems from comparative morphology across arthropods, particularly the segmental homologies in biramous limbs, where the inner endopodite aligns with endite components and the outer exopodite with exites. In , pleural sclerites such as the episternum and epimeron represent fused remnants of these subcoxal and epicoxal elements, with ontogenetic studies in primitive orders like Ephemeroptera revealing wing buds emerging from the pleuron. Fossil specimens from pterygotes further corroborate this by showing articulated wing bases integrated with pleural structures, suggesting a lateral origin tied to leg-derived appendages. The hypothesis accounts for wing articulation through the derivation of key basal structures from endite-exite joints: the subcostal sulcus forms from the anterior-posterior boundary between lobes, while axillary sclerites (e.g., 1Ax and 2Ax in neopterans) arise from fused proximal podomere hinges, enabling pivoting and muscle attachment for powered flight. This setup contrasts with simpler fixed articulations in other models and aligns with the multi-sclerite band observed in early pterygote fossils. Criticisms of the Endite-Exite Hypothesis center on its structural complexity, which posits multiple podomere migrations and fusions that are difficult to verify phylogenetically without intermediate fossils or molecular markers confirming homologies. While it integrates elements of epicoxal origins—such as endite-exite lobes potentially stemming from epicoxal podomeres—the model's reliance on detailed groundplan reconstructions has led to debates over its compared to more parsimonious alternatives.

Other Hypotheses and Dual Origin

In addition to the primary hypotheses, alternative theories propose that insect wings originated from gill-like structures or thoracic limb derivatives. The gill-wing hypothesis suggests that wings evolved from movable gill plates on the abdominal segments of ancestors, which initially served respiratory functions before adapting for aerial locomotion. A 2024 study analyzed fossil larvae from deposits, proposing that thoracic and abdominal outgrowths functioned as s and represent serial homologs to wings, supporting an origin. This view is supported by fossil evidence of larvae exhibiting flattened thoracic and abdominal projections resembling s, indicating a shared developmental origin for breathing and flight organs. In contrast, the leg-branchial hypothesis posits that wings derived from proximal segments of ancestral s, where outgrowths or lobes on thoracic limbs were incorporated into the body wall and later modified for flight. Experimental gene knockouts in crustaceans demonstrate that patterning genes for these leg segments correspond to wing-forming regions in , reinforcing this thoracic limb origin. The dual origin model integrates elements from multiple theories by proposing that wings evolved from the fusion of tergal (dorsal body wall) and pleural (lateral, leg-derived) tissues in pterygotes. This is evidenced by distinct expression patterns, where (Ubx) represses wing development in the hindwing imaginal discs of basal insects but promotes it in derived lineages, allowing differential evolution of wing pairs. Activation of an ancient in crustacean epipodites—gill-like leg outgrowths—further supports the fusion of dorsal and lateral tissues into a unified structure. Recent phylogenomic analyses from the 2020s reveal multiple instances of flight loss and regain, challenging strict of wing evolution and highlighting evolutionary flexibility. For example, in stick and leaf insects (), genomic data indicate at least three independent regains of wings from flightless ancestors, driven by shifts in developmental gene networks. Fossil records from the align with aquatic-to-terrestrial transitions. Hybrid models synthesize these insights by combining pleural and paranotal elements, positing that wings arose through the integration of ancestral leg-derived and dorsal body wall structures. This approach reconciles competing hypotheses, with developmental studies showing how gene co-option enabled such fusions to produce functional flight appendages.

References

  1. [1]
    Insect Flight: State of the Field and Future Directions - PMC
    The evolution of flight in an early winged insect ancestral lineage is recognized as a key adaptation explaining the unparalleled success and diversification of ...
  2. [2]
    The aerodynamics of insect flight | Journal of Experimental Biology
    Dec 1, 2003 · This review covers the basic physical principles underlying flapping flight in insects, results of recent experiments concerning the aerodynamics of insect ...
  3. [3]
    [PDF] Evolution of flight in insects
    Jul 31, 2024 · This review article consolidates current understanding of the evolutionary process of insect flight, encompassing subjects such as the fossil ...
  4. [4]
    Passive mechanisms in flying insects and applications in bio ...
    Jul 2, 2025 · This review synthesizes current research progress on passive mechanisms in insect flight systems, aiming to establish foundational knowledge frameworks.
  5. [5]
    Beyond aerodynamics: The critical roles of the circulatory and ...
    This review provides a survey of the various living components in insect wings, as well as the specific contribution of the circulatory and tracheal systems.
  6. [6]
    [PDF] Materials, Structure, and Dynamics of Insect Wings as Bioinspiration ...
    In addition to regional differences in wing venation pattern and the material properties of veins and membranes, many insect wings contain flexion lines – bands ...<|control11|><|separator|>
  7. [7]
    Origin and evolution of insect wings and their relation to ...
    The full primitive wing venation consists of six symmetrical pairs of veins; in each pair, the first branch is always convex and the second always concave; ...
  8. [8]
    A simple developmental model recapitulates complex insect wing ...
    Sep 17, 2018 · Insect wings are typically supported by thickened struts called veins. These veins form diverse geometric patterns across insects.
  9. [9]
    Flexural stiffness in insect wings I. Scaling and the influence of wing ...
    Sep 1, 2003 · In this study, we address the relationship between venation pattern and wing flexibility by measuring the flexural stiffness of wings.
  10. [10]
    Aerodynamics, sensing and control of insect-scale flapping-wing flight
    While conventional aerofoils are smooth and streamlined, insect wings exhibit rough surfaces, e.g. the cross-sectional corrugations of dragonfly wings or scales ...
  11. [11]
    Evolution of Flight Muscle Contractility and Energetic Efficiency - PMC
    Oct 9, 2020 · Larger insects, such as dragonflies and locusts, use direct flight muscle for wing beating during flight. Contraction of elevator muscles pulls ...
  12. [12]
    Locomotion – ENT 425 – General Entomology
    These are called “indirect flight muscles” because they have no direct contact with the wings. They stretch from the notum to the sternum. When they contract, ...
  13. [13]
    The thorax musculature of Anisoptera (Insecta: Odonata) nymphs ...
    The muscles and wing base sclerites in Odonata form a direct flight mechanism. Dorso-ventral muscles are attached directly to elements of the wing base, ...Missing: tergosternal pleurosternal
  14. [14]
    Structure, function and evolution of insect flight muscle - PMC
    In many insects, IFMs do not directly drive the wings, but do so indirectly by deforming the thoracic exoskeleton (indirect flight muscle, Fig. 4).
  15. [15]
    Insect Flight | Smithsonian Institution
    True flies are a large group of insects with only one pair of wings, although they have small balancing organs known as halteres where a second pair of wings ...
  16. [16]
    Flight Muscle - an overview | ScienceDirect Topics
    Direct flight muscles (DFM): Muscles that act directly on the wing base. The power-producing muscles of primitive insects (e.g., Odonata) are directly attached ...Birds, Biodiversity Of · Flight: Constraints And... · The Ecological Diversity Of...
  17. [17]
    Dragonfly (Sympetrum flaveolum) flight: Kinematic measurement ...
    From the kinematic experiments, two flapping patterns of the dragonfly wing are revealed as a simple figure-eight and a double figure-eight flapping pattern.
  18. [18]
    Dragonfly flight: free-flight and tethered flow visualizations reveal a ...
    The direct flight musculature of dragonflies means that stroke frequency,amplitude, phase and angle of attack can be varied independently on each of the four ...
  19. [19]
    Evolution of Flight Muscle Contractility and Energetic Efficiency
    Synchronous direct flight muscles are found in lower species of insects like locusts (Orthoptera) and dragonfly (Odonata), which have wing beats at lower ...
  20. [20]
    Dragonfly Flight : II. Velocities, Accelerations and Kinematics of ...
    Feb 1, 1997 · The wingbeat frequency of dragonflies during free flight has been shown to be correlated with body temperature (Sotavalta, 1954; May, 1981).
  21. [21]
    Aspects of Flight Mechanics in Anisopterous Dragonflies
    The present work will show that in Odonata the twisting of all four wings is controlled by small accessory flight muscles. Whereas in Schistocerca both ...Missing: activation | Show results with:activation
  22. [22]
    On the natural frequencies and mode shapes of dragonfly wings
    Aug 6, 2025 · The average flapping frequency 27 Hz of dragonflies is about 16% of the fundamental natural frequency. At this frequency ratio, the inertial ...
  23. [23]
    Bridging two insect flight modes in evolution, physiology ... - Nature
    Oct 4, 2023 · This equation captures the indirect actuation of synchronous or asynchronous insect flight muscle, which act via the deformation of the thorax ...
  24. [24]
    Mechanical analysis of Drosophila indirect flight and jump muscles
    Instead, the IFM fibers of Drosophila are asynchronous, meaning they do not contract synchronously with each nerve action potential from the flight muscle ...
  25. [25]
    The flying insect thoracic cuticle is heterogenous in structure and in ...
    Jan 15, 2022 · The thorax is a specialized structure central to insect flight. In ... We investigated cuticle structure using histology, resilin ...
  26. [26]
  27. [27]
    Wing Rotation and the Aerodynamic Basis of Insect Flight - Science
    The enhanced aerodynamic performance of insects results from an interaction of three distinct yet interactive mechanisms: delayed stall, rotational circulation ...
  28. [28]
    Flexible clap and fling in tiny insect flight
    Oct 1, 2009 · In this paper, we use the immersed boundary method to simulate clap and fling in rigid and flexible wings.Missing: mode | Show results with:mode
  29. [29]
    Vortex trapping recaptures energy in flying fruit flies | Scientific Reports
    Mar 26, 2021 · The images suggests that clap-and-fling wing motion in fruit flies lasts approximately one millisecond or ~ 15% of the ~ 150 Hz stroke cycle. We ...
  30. [30]
    Short-amplitude high-frequency wing strokes determine the ... - PNAS
    Here we demonstrate that honeybees use an alternative strategy, hovering with relatively low stroke amplitude (≈90°) and high wingbeat frequency (≈230 Hz).
  31. [31]
    [PDF] The role of drag in insect hovering - Z. Jane Wang
    2A), each half-stroke generates almost equal lift in the vertical direction and almost equal drag in the opposite horizontal direction. The averaged.
  32. [32]
    Insect flight dynamics: Stability and control | Rev. Mod. Phys.
    May 16, 2014 · This review begins with an overview of the flapping kinematics and aerodynamics of insect flight. It is followed by a summary of the ...
  33. [33]
    Active and passive stabilization of body pitch in insect flight - PMC
    The magnetic torque perturbation (thin blue stripe) tips the insect downwards, and the insect responds by correcting its orientation.
  34. [34]
    Hawkmoths use nectar sugar to reduce oxidative damage from flight
    Feb 17, 2017 · Hovering flight is the most energetically expensive form of locomotion known (3), with metabolic rates reaching 170 times higher than at rest (4) ...
  35. [35]
    Power and efficiency of insect flight muscle
    Mar 1, 1985 · ... Pringle, 1959 ... In the more primitive synchronous insect flight muscles there is direct nervous stimulation of each muscle contraction.
  36. [36]
    A comprehensive review of long‐distance hover fly migration ...
    Aug 24, 2024 · Tomlinson and Menz (2015) examined metabolic rate (i.e., energy production) and water loss patterns in Ep. balteatus and Er. tenax to better ...
  37. [37]
    The aerodynamics of hovering insect flight. VI. Lift and power ...
    Power estimates are compared with the available measurements of metabolic rate during hovering to investigate the role of elastic energy storage, the ...
  38. [38]
    Dragonfly Flight: Power Requirements at High Speed and Acceleration
    Jul 1, 1991 · Rüppell (1989) reported maximum speeds of 7.5 ms−1 in A. junius and 10 ms−1 in Aeshna cyanea, although average flight velocities were much lower ...
  39. [39]
    Speed control and force-vectoring of bluebottle flies in a ...
    Feb 19, 2019 · Forward flight speed and its dependency on body pitch angle and aerodynamic damping. Using three angle-pin angles and three dampers (D0, D1 ...
  40. [40]
    Aerodynamic force generation and power requirements in forward ...
    Sep 1, 2003 · In forward flight, the body drag is not zero and the stroke plane is tilted forward to produce thrust.Missing: reduction | Show results with:reduction
  41. [41]
    The aerodynamics and control of free flight manoeuvres in Drosophila
    Sep 26, 2016 · Unlike throttle and pitch control, changes in roll and yaw involve asymmetrical alterations in the stroke kinematics of the two wings. Flies ...
  42. [42]
    Manoeuvring flight of a model insect—Saccadic yaw and sideslip
    Saccadic body yaw ranging from 45° to 180° were obtained through small asymmetric adjustments in the twist of the contralateral wing pair, while sideslip flight ...
  43. [43]
    Flies tune the activity of their multifunctional gyroscope - ScienceDirect
    Aug 19, 2024 · Halteres are renowned for acting as biological gyroscopes that rapidly detect rotational perturbations and help flies maintain a stable flight posture.
  44. [44]
    The natural flight of the migratory locust,Locusta migratoria L.
    A lift drag ratio of 1.7∶1 has been calculated, and a minimum sinking rate of just over 2 m/s has been measured. 6. Glides commence with a positive body angle ...
  45. [45]
    [PDF] A Blade Element Approach to Modeling Aerodynamic Flight of an ...
    Blade element theory provides advantages over more complex models by providing a simple analytical model for the aerodynamic forces. Although blade element ...
  46. [46]
    Three-dimensional wing structure attenuates aerodynamic efficiency ...
    Mar 11, 2020 · The computed flight efficiencies of 17–23% are somewhat below experimentally derived estimates that range from 26–32% in various species of ...<|separator|>
  47. [47]
    The Changes in Power Requirements and Muscle Efficiency During ...
    Apr 1, 1997 · The limits of flight performance have been estimated in tethered Drosophila melanogaster by modulating power requirements in a 'virtual ...
  48. [48]
    Mechanism of resilin elasticity | Nature Communications
    Aug 14, 2012 · It could be stretched to over 300% of its original length before breaking with low elastic modulus in the range of 0.1–3 MPa as indicated by ...Missing: Young's | Show results with:Young's
  49. [49]
    Resilin in Insect Flight Systems - The Advanced Portfolio - Wiley
    Aug 18, 2023 · Wing folding involves the intrinsic elasticity (energy storage) of resilin and the active flexion of the wing towards the longitudinal body ...
  50. [50]
    The “click” mechanism in dipteran flight: if it exists, then what effect ...
    With the “click” mechanism, it is possible to change the velocity profile at the root of the wing so that it is non-sinusoidal for a sinusoidal input force. It ...
  51. [51]
  52. [52]
    A Comparative Study of the Flight Mechanism of Diptera
    The click action observed in CCl4-anaesthetized Calliphora was found to be due to an interaction between the radial stop and the pleural wing process at the top ...
  53. [53]
    Regional differences in degree of resilin cross-linking in the desert ...
    ... energy storage in the flight system. In structures where deformations tend to occur more slowly, such as the clypeo-labral springs and tracheae, di- and ...
  54. [54]
    Development and deposition of resilin in energy stores for locust ...
    A second structure that is suggested to store the remaining 43% of the requisite energy is the tendon of the extensor muscle (Bennet-Clark, 1975), the materials ...
  55. [55]
    Resilin and chitinous cuticle form a composite structure for energy ...
    Sep 30, 2008 · In this study, we analyse the nature and action of specialised structures that store and release energy to power the most effective jumping ...
  56. [56]
    Wing Coupling in Bees and Wasps: From the Underlying Science to ...
    Jun 3, 2021 · By synchronizing fore and hind wings, the coupling mechanism enables insects to attain more lift and better gliding performance. The disability ...
  57. [57]
    Structure, properties and functions of the forewing-hindwing ...
    The coupling structure in honeybee wings consists of a posterior rolled margin at the trailing margin of the forewing and a set of V-shaped hamuli at the ...<|control11|><|separator|>
  58. [58]
    Flight of the dragonflies and damselflies - Journals
    Sep 26, 2016 · In §4, we see how the Odonata overcome this trade-off by operating their four wings independently, achieving excellent flight performance.
  59. [59]
    Wing coupling mechanism in the butterfly Pieris rapae (Lepidoptera ...
    In general, a wing coupling mechanism is normally composed of two components situated at the fore- and hindwing, respectively. As an evolutionary consequence, ...
  60. [60]
    Wings and halteres act as coupled dual oscillators in flies - PMC
    In Diptera, the forewings power flight, whereas the hindwings have evolved into specialized structures called halteres, which provide rapid mechanosensory ...Missing: odonata | Show results with:odonata
  61. [61]
    Phasing of dragonfly wings can improve aerodynamic efficiency by ...
    When operating at a phase shift of +25%, the hind wing experiences a peak in lift enhancement (lift compared with the hind wing flapping in isolation) at 35 ...
  62. [62]
    Energy metabolism in orchid bee flight muscles: carbohydrate fuels all
    Sep 15, 2005 · ... in fatty acid oxidation, was not detectable in any species. Thoracic ... fat bodies and flight muscles that are sufficient to support flight.
  63. [63]
    Metabolism - ScienceDirect
    ... fat, and proline oxidation in insect flight muscles. Included are the ... flight is accompanied by activation of fatty acid oxidation and inhibition of ...
  64. [64]
    Flexibility in energy metabolism supports hypoxia tolerance in ...
    Aerobic energy metabolism in insect flight muscle is similar to that of humans in ... Fatty acid metabolism is not important for flight in Diptera, and is rarely ...Missing: oxidation | Show results with:oxidation
  65. [65]
    Insect Flight Muscle Chemomechanics - NCBI - NIH
    The biochemical and mechanical basis of insect flight has captivated the interest of biologists for decades. This chapter presents a brief review of the
  66. [66]
    MITOCHONDRIA IN THE FLIGHT MUSCLES OF INSECTS
    When mature, the sarcosomes account for 32.6 per cent of the total muscle dry weight, or close to 40 per cent on a wet weight basis. 3. It appears probable that ...Missing: density | Show results with:density
  67. [67]
    The role of neurohormonal octopamine during 'fight or ... - PubMed
    Octopamine has been called the 'fight or flight' hormone of insects. We tested this hypothesis by measuring octopamine levels in the haemolymph of field ...Missing: control takeoff mobilization allostery enzyme kinetics
  68. [68]
    [PDF] REVIEW INSECT FLIGHT MUSCLE METABOLISM - DSpace
    Insect flight muscles contain a limited amount of carbohydrate reserves, which can meet the energy requirements of the muscles only during the first few.<|control11|><|separator|>
  69. [69]
    Metabolic Biochemistry of Insect Flight
    Hormonal control of flight will be considered with respect to the activation of flight muscle metabolism and the long-term mobilization of fuels for sustained ...
  70. [70]
  71. [71]
  72. [72]
  73. [73]
    What serial homologs can tell us about the origin of insect wings - NIH
    Mar 14, 2017 · Among the several wing genes identified from Drosophila studies, vestigial ( vg) is often considered one of the most critical wing marker genes ...
  74. [74]
    Dual evolutionary origin of insect wings supported by an ... - PNAS
    Jan 9, 2018 · There are currently two prominent, but contrasting wing origin hypotheses (the tergal origin hypothesis and the pleural origin hypothesis).
  75. [75]
    [PDF] Origin of the insect wing and wing articulation from the arthropodan leg
    The pro-wing originally operated on a row of pivots from the epicoxa and subcoxa (pleuron) and became mobilized by epicoxal leg musculature. KUKALOVA-PECK,. J.
  76. [76]
    Paleozoic Nymphal Wing Pads Support Dual Model of Insect Wing ...
    Jan 12, 2017 · The gradual development of movable wing pads during post- embryonic development was considered primary evidence to support the ''epicoxal'' ...Missing: embryological | Show results with:embryological
  77. [77]
    Evolutionary origin of insect wings from ancestral gills - ResearchGate
    Aug 9, 2025 · Two hypotheses have been proposed for the origin of insect wings. One holds that wings evolved by modification of limb branches that were ...
  78. [78]
    A century and a half of research on the evolution of insect flight
    Aug 7, 2025 · Lacking transition fossils, neither theory could be definitively rejected. Winged insects are abundant in the fossil record from the mid- ...Missing: criticisms | Show results with:criticisms
  79. [79]
    The Origin of Insect Wings | The Backyard Arthropod Project
    Mar 30, 2013 · The main objection to this hypothesis is that it would have been difficult to evolve the joints from scratch, and the genetics of wing joints ...
  80. [80]
    “Only Insects (and Angels) Were Able to Acquire Wings Without ...
    Jul 7, 2025 · Together with her 1978 review, the introduction of the epicoxal model of insect wing evolution became Kukalová-Peck's most widely cited work (> ...
  81. [81]
    Report Abdominal serial homologues of wings in Paleozoic insects
    Aug 8, 2022 · Epicoxal or exite origin hypothesis originally proposed by Kukalová-Peck for the origin of insect wings received further support from a ...
  82. [82]
    Evolution of Insect Wings and Flight | Nature
    Nov 16, 1973 · Sir Vincent Wiggles-worth, however, here gives evidence that the thoracic styli were perhaps the precursors of insect wings, and describes a ...
  83. [83]
    A century and a half of research on the evolution of insect flight
    The gill and paranotal lobe theories of insect wing evolution were both proposed in the 1870s. For most of the 20th century, the paranotal lobe theory was more ...Missing: Arthropleura | Show results with:Arthropleura
  84. [84]
    Hypothesis Testing in Evolutionary Developmental Biology: A Case ...
    Aug 6, 2025 · This hypothesis was based on the presence of tergal expansions (called paranotal lobes) in some extant primitively wingless insects and fossil ...
  85. [85]
    [PDF] aerodynamics, thermoregulation, and the evolution of - insect wings
    We examine several aerodynamic and thermoregulatory hypotheses about pos- sible adaptive factors in the evolution of wings from small winglets in insects. Using ...
  86. [86]
  87. [87]
    A second view on the evolution of flight in stick and leaf insects ...
    May 12, 2022 · Along these lines, the regain of wings in stick and leaf insects (Phasmatodea) was hypothesised to have occurred several times independently ...
  88. [88]
    What serial homologs can tell us about the origin of insect wings.
    Mar 14, 2017 · The tergal origin hypothesis (also known as the paranotal lobe hypothesis) proposes that wings originated from expansions of the dorsal body ...<|control11|><|separator|>