Fact-checked by Grok 2 weeks ago

Radiation pressure

Radiation pressure is the mechanical pressure exerted by on a surface, arising from the transfer of carried by photons or electromagnetic waves upon or . This phenomenon occurs because electromagnetic waves possess linear proportional to their energy divided by the , p = E/c, leading to a force when the wave interacts with matter. The concept of radiation pressure has a rich historical background, with early suspicions dating to Johannes Kepler in 1619, who attributed the orientation of comet tails away from the Sun to the pressure of sunlight. James Clerk Maxwell provided the first theoretical foundation in 1873, predicting it as a consequence of electromagnetic wave momentum in his Treatise on Electricity and Magnetism. Experimental confirmation came in 1901 through Pyotr Lebedev's measurements of light pressure on small mirrors, followed by independent verifications by Ernest Fox Nichols and Gordon Ferrie Hull in 1901–1903. In physical terms, the radiation pressure P on a perfectly absorbing surface is given by P = I/c, where I is the of the and c is the , while for a perfectly reflecting surface, it doubles to P = 2I/c due to the change in direction. Within stellar interiors, pressure takes the form P_\text{rad} = \frac{1}{3} a T^4, where a is the radiation constant and T is the , playing a crucial role in by counteracting in massive stars. This contribution is particularly significant near the Eddington limit, the maximum at which pressure balances gravitational attraction, setting an upper bound on stellar masses during formation. Radiation pressure has diverse applications, notably in space propulsion through , which harness sunlight's for fuel-free acceleration; for instance, NASA's Advanced Composite Solar Sail System (ACS3) demonstrates this with a 9-meter sail providing via solar radiation pressure. In , it influences phenomena such as the repulsion of dust particles in comet tails and the dynamics of interstellar dust, as well as limiting by halting infall in massive protostars. These effects underscore its importance across scales, from laboratory to cosmic structures.

History

Discovery

The earliest observation potentially linked to radiation pressure dates to 1619, when noted that the tails of comets consistently point away from the Sun, suggesting a repulsive force emanating from the body that pushes lightweight particles outward. Kepler interpreted this as a mechanical effect from solar emission, though he attributed it to a stream of thin rather than itself; the connection to radiation pressure was only recognized retrospectively as evidence of light's momentum transfer. Theoretical groundwork for radiation pressure emerged in the 19th century with the development of electromagnetic theory. In 1873, James Clerk Maxwell predicted that electromagnetic waves, including , carry and thus exert on absorbing or reflecting surfaces, deriving this from the stresses within the as outlined in his seminal work. This prediction implied a equal to the of the radiation for perfect absorption, providing a quantitative basis for the phenomenon independent of . Experimental verification proved challenging due to the minuscule forces involved—on the order of the weight of a bacterium for on a small mirror—and initial confusion with thermal effects in devices like the , which rotates due to gas molecule recoil rather than direct radiation pressure. The first reliable confirmation came from Pyotr Lebedev, whose experiments conducted in 1899 and announced in 1900 measured deflection in suspended mirrors under electric lamp illumination, followed closely by Ernest Fox Nichols and Gordon Ferrie Hull in 1901–1903 at . Nichols and Hull employed a sensitive torsion balance with lightweight silvered vanes exposed to , achieving agreement with Maxwell's theory within 1% after evacuating the apparatus to eliminate gas effects. These results overcame widespread skepticism, establishing radiation pressure as a verifiable physical reality despite its subtlety.

Early theoretical developments

The theoretical understanding of radiation pressure advanced significantly in the late 19th and early 20th centuries, building on James Clerk Maxwell's electromagnetic theory, which predicted that carries and exerts pressure on matter. In 1900–1901, Russian physicist Pyotr Lebedev conducted the first laboratory experiments confirming Maxwell's prediction of radiation pressure. Using a bright lamp as the light source and a delicate torsion balance inspired by contemporary designs, Lebedev measured the deflection caused by impinging on small suspended mirrors and blackened surfaces, distinguishing between and effects; his results showed pressures consistent with theoretical expectations within experimental error. Independent confirmation came in 1901–1903 from American physicists Ernest Fox Nichols and Gordon Ferrie , who achieved greater sensitivity using a refined torsion and improved optical isolation to measure radiation pressure from both and visible sources on polished and absorbing surfaces, yielding values agreeing with formula to within a few percent. These experiments solidified the wave-based electromagnetic interpretation but reignited historical debates on whether light pressure stemmed from a corpuscular or undulatory , as earlier corpuscular theories had intuitively explained transfer while wave models required electromagnetic interactions. By the 1910s, applications extended to , with Arthur Schuster exploring radiation pressure's role in stellar atmospheres through models of , where pressure from absorbed and scattered influences atmospheric structure and formation. The transition to quantum interpretations began with Albert Einstein's 1905 light quantum hypothesis, which posited discrete energy packets () carrying p = h/\lambda, providing a particle-like framework that reconciled wave predictions with pressure observations and paved the way for developments in photon models.

Theoretical Foundations

Electromagnetic wave momentum

In classical electromagnetism, light is described as an electromagnetic wave that carries both energy and . The momentum arises from the interaction of electric and magnetic fields, as predicted by James Clerk Maxwell in his 1873 treatise on electromagnetism. This wave underpins the concept of radiation pressure, where the transfer of to a surface results in a mechanical force. The of an electromagnetic wave is given by the \vec{S} = \frac{1}{\mu_0} \vec{E} \times \vec{B}, which represents the power per unit area carried by the fields. The associated is \vec{g} = \frac{\vec{S}}{c^2}, reflecting the relativistic relation between and for electromagnetic fields in . For a propagating in the z- with \vec{E} and \vec{B} perpendicular to each other and to the propagation , |\vec{S}| = \frac{E B}{\mu_0} = \frac{E^2}{Z_0} where Z_0 = \sqrt{\mu_0 / \epsilon_0} is the , and the points in the of propagation./08%3A_Electromagnetic_Fields_and_Energy_Flow/8.02%3A_Poyntings_Theorem) When such a wave impinges normally on an absorbing surface, the momentum flux through the surface equals the rate of momentum transfer per unit area, yielding a radiation pressure P = \frac{1}{c} \int \vec{S} \cdot d\vec{A}./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.05%3A_Momentum_and_Radiation_Pressure) For a plane wave with time-averaged intensity I = \langle |\vec{S}| \rangle, this simplifies to P = \frac{I}{c} on a perfectly absorbing surface, as all incident momentum is deposited./University_Physics_II_-Thermodynamics_Electricity_and_Magnetism(OpenStax)/16%3A_Electromagnetic_Waves/16.05%3A_Momentum_and_Radiation_Pressure) For more general geometries and field configurations, the radiation pressure is derived from the \overset{\leftrightarrow}{T}, whose components are T_{ij} = \epsilon_0 \left( E_i E_j - \frac{1}{2} \delta_{ij} E^2 \right) + \frac{1}{\mu_0} \left( B_i B_j - \frac{1}{2} \delta_{ij} B^2 \right). The force on a surface is the surface of the normal component of this tensor, \vec{F} = \oint \overset{\leftrightarrow}{T} \cdot d\vec{A}, where the normal stress T_{nn} directly gives the pressure as the momentum transfer rate per unit area. This framework accounts for both normal and shear components of flux in arbitrary electromagnetic fields. Relativistically, the total momentum of the electromagnetic field in a volume V is \vec{p} = \epsilon_0 \int_V \vec{E} \times \vec{B} \, dV = \frac{1}{c^2} \int_V \vec{S} \, dV, consistent with the field acting as a relativistic fluid with equal to energy divided by c in the propagation direction for unidirectional . The energy density is u = \frac{1}{2} \left( \epsilon_0 E^2 + \frac{B^2}{\mu_0} \right), but for isotropic radiation—such as where fields propagate equally in all directions—the differs, given by P = \frac{u}{3}. This relation emerges from integrating the flux over all directions, analogous to the equation of state for photon gas.

Absorption and reflection pressures

Radiation pressure arises from the transfer of momentum carried by electromagnetic waves to a surface upon interaction. For a unidirectional beam incident normally on a perfectly absorbing surface, the pressure is given by P = \frac{I}{c}, where I is the intensity (energy flux) of the radiation and c is the speed of light; this follows classically from the momentum flux of the electromagnetic field, which equals the energy flux divided by c. For a perfectly reflecting surface under normal incidence, the pressure doubles to P = \frac{2I}{c}, as the momentum change involves both absorption and re-emission in the opposite direction, reversing the incident momentum. In the general case of partial reflection, the pressure is P = \frac{(1 + r)I}{c}, where r (0 ≤ r ≤ 1) is the reflection coefficient, representing the fraction of incident energy reflected; here, r = 0 recovers the absorbing case and r = 1 the reflecting case. When the beam strikes at an oblique angle \theta (measured from ), the effective on is reduced by the projection factor \cos \theta, and the component of transfer introduces an additional \cos \theta. For , the becomes P = \frac{I \cos^2 \theta}{c}; for perfect , P = \frac{2I \cos^2 \theta}{c}; and generally, P = \frac{(1 + r) I \cos^2 \theta}{c}. The total force on a surface of area A is then F = P A, directed to ; for example, in applications, this force propels spacecraft by reflecting sunlight at non- angles, with the \cos^2 \theta dependence optimizing orientation for maximum thrust. These expressions apply to unidirectional radiation fields, such as a . In contrast, for isotropic radiation fields—like in a —the pressure on a surface is P = \frac{u}{3}, where u is the total of the ; this arises from integrating the momentum flux over all directions in the , yielding one-third of the regardless of the surface's or properties, as the field remains isotropic internally.

Photon-based derivation

In the quantum mechanical description of light, electromagnetic radiation is modeled as a stream of , each carrying discrete energy and . This perspective, introduced by in his 1905 paper on the , posits that the energy E of a single is E = h f, where h is Planck's constant and f is the frequency of the light. The p of a , derived from for a , is then p = E / c = h f / c = h / \lambda, where c is the and \lambda is the . This particle-like transfer provides an alternative derivation of radiation pressure, complementary to the classical electromagnetic wave approach. For a beam of incident on a perfectly absorbing surface, each imparts its full p upon , transferring in the direction of . If N photons are absorbed per unit time over an area A, the total transfer rate is N p = N (E / c), yielding a force F = N E / c. The resulting radiation P is this force divided by the area: P = F / A = (N E) / (c A). Since the I of the beam is the power per unit area, I = (N E) / A, the pressure simplifies to P = I / c. In the case of a perfectly reflecting surface at normal incidence, each reverses its direction upon , resulting in a change in of $2p per (from +p to -p). For oblique incidence at an angle \theta to the , the relevant component normal to the surface changes by $2p \cos \theta. Thus, for normal incidence, the total transfer rate doubles to $2 N p = 2 N (E / c), and the becomes P = 2 (N E) / (c A) = 2 I / c. For monochromatic , the number flux of (photons per unit time per unit area) is dn/dt = I / (h f), confirming that the scales with the arrival rate and individual transfer. The model is particularly advantageous for describing discrete interactions in scenarios where fields are sparse or intensities are low, such as in single-photon experiments, or in high-energy regimes like gamma-ray bursts where quantum effects dominate over classical wave descriptions. Einstein further elaborated on this particle nature in his analysis of radiation pressure fluctuations, demonstrating how photon-like exchanges explain statistical variations in pressure on a mirror immersed in .

Emission contributions

Radiation pressure contributions from emission arise when a source emits , imparting to the surrounding medium or experiencing due to conservation of . Unlike pressure from incident radiation, which pushes inward on absorbing or reflecting surfaces, emission-based pressure acts outward from the source, as the radiated photons carry away in the direction of . This effect is fundamental in scenarios where the emitter itself generates , such as in self-luminous bodies or processes. For an isolated particle or body emitting radiation, the net force due to recoil is given by the negative of the rate at which momentum is carried away by the photons. If the momentum emission rate is \frac{d\vec{p}}{dt}, the recoil force on the emitter is \vec{F} = -\frac{1}{c} \frac{d\vec{p}}{dt}, where c is the speed of light; for emission of power P in a specific direction, this simplifies to F = P/c in magnitude. In the case of isotropic emission from a point source with total luminosity L, symmetry ensures the net recoil force on the source is zero, as momenta cancel in all directions. However, this emitted radiation exerts pressure on an enclosing spherical shell of radius r: the energy flux at the shell is S = L / (4\pi r^2), and for perfect absorption, the outward pressure is P = S / c = L / (4\pi r^2 c). This configuration models the expansive force of radiation from a central source on surrounding material, such as in proto-stellar envelopes. In stellar and planetary contexts, emission recoil becomes significant for hot bodies with non-uniform temperature distributions, leading to anisotropic . The Yarkovsky effect exemplifies this, where a rotating or small body re-emits absorbed unevenly due to thermal lag, resulting in a net from the recoil of photons emitted preferentially from the warmer . This outward-directed force perturbs orbits over long timescales, with magnitude scaling as F \propto L / c but modulated by the body's thermal properties and rotation; for typical near-Earth asteroids, it induces semimajor axis drifts of up to 10^{-4} /Myr. Such effects are precursors to more complex radiative accelerations in non-spherical or rotating emitters. In plasmas, emission contributions to radiation pressure often involve Thomson scattering, where free electrons scatter incident photons elastically and re-emit them in random directions, transferring net to the plasma. The scattered light's recoil imparts a force equivalent to the incident momentum flux divided by c, with the Thomson cross-section \sigma_T = 6.65 \times 10^{-25} cm² determining the opacity; for an electron density n_e, the pressure is P = (\sigma_T n_e S)/c, where S is the incident . This process drives outward acceleration in relativistic plasmas or stellar winds, distinguishing it from pure by the diffuse re-emission that sustains the momentum transfer across the medium. In astrophysical settings like accretion disks, this scattering-enhanced pressure can approach the Eddington limit, balancing .

Uniform field effects

In a uniform, isotropic radiation field, such as that of in , the radiation pressure P is related to the u by the equation P = \frac{1}{3} u, where u = a T^4 and a is the radiation constant. This relation arises because the isotropic distribution of momenta contributes equally to in all directions, analogous to the but adjusted for the relativistic nature of photons. The derivation from photon gas statistics treats the radiation as a collection of massless particles with energy E = pc, where p is and c is the . The pressure is computed as the average flux across a surface, given by P = \frac{1}{3} \int p_x v_x f(\mathbf{p}) d^3p, where f(\mathbf{p}) is the integrated over to account for ; for , this yields P = \frac{1}{3} u. This flux perspective highlights how the random, isotropic collisions of with a surface produce a net compressive force without directional bias. In radiation-dominated regimes, where the energy density is primarily from , the equation of state for the photon gas is P = \frac{1}{3} u, leading to adiabatic compression behaviors distinct from non-relativistic gases. During compression, the scales as u \propto V^{-4/3} for a fixed , implying that pressure resists collapse more effectively at higher densities due to its stiff equation of state. This property is crucial for understanding stability in high-temperature environments where matter interactions are subordinate to radiation. An analogy to ' swindle in illustrates how radiation pressure modifies instability criteria: just as the original Jeans analysis assumes a uniform background to derive a critical for , incorporating radiation pressure extends this by adding a term to the effective speed, stabilizing perturbations when P = \frac{1}{3} u dominates over gas . In such cases, the Jeans length increases, suppressing fragmentation in radiation-supported structures. In opaque media, where radiation diffuses rather than streams freely, the uniform field approximation still holds locally, with pressure arising from the gradient of the radiation energy density under the diffusion regime. This diffusive transport maintains isotropic pressure contributions, enabling radiation to compress over scales where opacity prevents direct momentum transfer.

Solar Radiation Pressure

Intensity and basic calculations

The , defined as the average intensity of solar radiation at Earth's distance from (1 , or AU), is approximately 1366 W/m². This value represents the total perpendicular to the incoming rays just outside Earth's atmosphere. For radiation pressure calculations in the solar system, the pressure exerted by this flux on a perfectly absorbing surface is given by P_{\text{sun, abs}} = \frac{I}{c}, where I is the intensity and c = 3 \times 10^8 m/s is the in vacuum. Substituting the solar constant yields P_{\text{sun, abs}} \approx 4.5 \, \mu\text{N/m}^2. For a perfectly reflecting surface, the pressure doubles due to change upon reflection, so P_{\text{sun, ref}} = \frac{2I}{c} \approx 9 \, \mu\text{N/m}^2. Solar radiation intensity decreases with distance r from according to the , I \propto \frac{1}{r^2}, as the same power output spreads over a larger spherical surface area. Consequently, radiation pressure scales similarly, P \propto \frac{1}{r^2}, making it significantly weaker beyond 1 —for instance, at Jupiter's distance of about 5.2 , the pressure drops to roughly 1/27th of its value at . This radial dependence influences orbital dynamics, particularly for lightweight objects where non-gravitational forces become comparable to solar gravity. For non-normal incidence, the effective on a surface is reduced by the cosine of \theta between the direction and the surface , yielding I_{\text{eff}} = I \cos \theta and thus P \propto \cos \theta for (or $2 \cos \theta for perfect , assuming ). This angular factor is crucial for calculating pressures on tilted or orbiting bodies. In the context of small particles, such as or interplanetary , pressure often dominates over gravitational attraction for sizes below approximately 1 μm. The ratio \beta = P_{\text{rad}} / P_{\text{grav}} > 1 in this regime arises because force scales with cross-sectional area while gravitational force scales with (proportional to ), favoring smaller particles where surface effects prevail.

Absorption versus reflection

In solar contexts, the radiation pressure exerted on a material depends fundamentally on whether the incident sunlight is absorbed or reflected. For perfectly absorbing surfaces, such as black bodies that capture all incoming photons without re-emission in the forward direction, the momentum transfer is complete, resulting in a pressure of P = I / c, where I is the solar intensity and c is the speed of light. This produces a drag-like force opposing the direction of propagation, as the surface gains the full momentum of the absorbed photons. For perfectly reflecting surfaces, like ideal mirrors that reverse the direction of incoming photons, the change in is doubled, yielding a pressure of P = 2I / c. This enhanced force points away from the Sun and can enable applications by providing net without mass expulsion. Real materials exhibit partial and , characterized by the wavelength-dependent \alpha (ranging from 0 for perfect to 1 for perfect ), which modifies the net pressure to P = (1 + \alpha) I / c. In the spectrum, materials like silicates or carbon-rich surfaces have albedos typically between 0.1 and 0.5, leading to intermediate pressures that influence orbital dynamics and surface interactions within the solar system. For and interplanetary dust grains exposed to solar , the efficiency of pressure is quantified by the radiation pressure efficiency Q_{\mathrm{pr}}, which accounts for both and the forward-directed component of and depends on , , and structure. Small grains (effective radius a_{\mathrm{eff}} \lesssim 0.3 \, \mu\mathrm{m}) often experience enhanced Q_{\mathrm{pr}} due to resonances, but reduces it compared to compact spheres. The key dynamical parameter is \beta = P_{\mathrm{rad}} / P_{\mathrm{grav}}, the ratio of to gravitational force; grains with \beta > 1 are blown out of the solar system on trajectories, a threshold typically met by submicron grains with metallic inclusions but not by larger or fluffier ones. Polarization effects in unpolarized solar light have minimal impact on overall radiation pressure for most dust grains and surfaces, as the net momentum transfer averages out over the random orientations.

Spacecraft perturbations and solar sails

Solar radiation pressure induces perturbations on spacecraft in heliocentric orbits by exerting a continuous acceleration a = \frac{P A}{m}, where P is the solar radiation pressure at 1 AU (approximately $4.56 \times 10^{-6} N/m²), A is the spacecraft's projected cross-sectional area normal to the Sun direction, and m is its mass; for reflective surfaces, this is multiplied by a coefficient up to 2. This non-conservative force disrupts Keplerian motion, leading to along-track drifts, variations in semi-major axis, and precession of the orbit's node and perigee, with effects on the order of millimeters per second squared for typical satellites and accumulating to observable changes over months. For geostationary satellites, these perturbations necessitate active station-keeping maneuvers, as the unbalanced force can cause east-west drifts of up to several degrees per year without correction. To harness rather than mitigate this pressure, solar sails employ large, lightweight reflective membranes to generate propellant-free via transfer from photons. The concept originated with Konstantin Tsiolkovsky's 1921 proposal for light-based propulsion, later expanded by Friedrich Tsander, envisioning sails for interplanetary travel. A landmark demonstration was Japan's mission in 2010, which successfully deployed a 196 m² sail—measuring 14 m per side—and verified a of 1.12 mN from radiation pressure, enabling a controlled to while powering onboard electronics via thin-film . This reflective design doubled the pressure compared to absorption, producing an of about 0.1 mm/s² near . The dynamics of solar sails rely on orienting the sail normal to the Sun for maximum thrust along the Sun-spacecraft line, facilitating non-Keplerian trajectories such as spiral escapes from the inner solar system or stationary "hovering" displaced from equilibrium points. Sail performance is quantified by the lightness number \beta, defined as the ratio of the sail's radiation-pressure acceleration to the local solar gravitational acceleration (\beta = \frac{a_\text{SRP}}{GM_\odot / r^2}), where values above 1 enable hyperbolic escapes; for IKAROS, \beta \approx 0.001, but advanced designs target \beta > 0.01 with areal densities below 5 g/m². Modern concepts build on this, as seen in the Planetary Society's LightSail 2 mission (2019), which used a 32 m² sail for controlled orbital maneuvers in Earth orbit, proving scalability for deep-space applications like asteroid reconnaissance. More recently, NASA's Advanced Composite Solar Sail System (ACS3), launched on April 23, 2024, demonstrated the deployment of a 9-meter sail in September 2024 to test lightweight composite booms for future solar sail propulsion systems. Key limitations include the need for precise attitude control to maintain optimal orientation, often via vanes or micro-ers, as misalignment reduces and induces unwanted torques from flexibility. Additionally, prolonged exposure to radiation causes material degradation, such as yellowing and reduced reflectivity in polymers like , potentially halving over years; mitigation involves aluminized coatings and periodic adjustments.

Astrophysical and Cosmic Effects

Interstellar dust and gas dynamics

In interstellar environments, radiation pressure significantly influences the dynamics of grains, particularly those on the order of micron-sized, where the β—the of the radiation pressure force to the force—typically ranges from 0.1 to 1. This value arises primarily from the and of stellar photons by grains composed of silicates or carbonaceous materials, leading to a net outward force that can exceed or balance stellar depending on grain size, composition, and proximity to the radiation source. For β > 0.5, grains become partially or fully decoupled from the , resulting in outward acceleration along trajectories. This outward push creates dust-free zones around stars, notably in H II regions where intense ultraviolet radiation from young, massive stars ionizes surrounding gas and expels smaller grains. In these regions, radiation pressure forms cavities with reduced dust densities, as grains with β ≈ 1 are blown out to distances of several parsecs, altering the local structure and facilitating the propagation of ionizing photons. Observations of such zones, including deficits in scattered light near stellar sources, confirm this depletion mechanism. Complementing the radial outward force, the Poynting-Robertson drag introduces a tangential component due to the aberration of in the grain's , causing a gradual loss of . For dust orbiting stars, this drag effect—arising from the re-emission of absorbed photons in the forward direction of motion—leads to spiral infall toward the central star over timescales of thousands to millions of years, depending on and orbital radius. This process shapes the inner edges of circumstellar dust distributions and contributes to the replenishment of material in accretion disks. Radiation pressure also affects gas dynamics through with , as grains absorb stellar and re-emit it thermally, transferring to the surrounding gas via collisions. In clouds, this indirect force on gas can enhance turbulent motions or stabilize structures against , particularly when dust-to-gas ratios increase due to differential drift under radiation pressure. Such prevents wholesale cloud disruption in many cases, maintaining overall stability while allowing localized outflows. Observational evidence for these dynamics includes infrared excesses around stars, attributed to warm populations maintained or redistributed by pressure-driven flows. These excesses, detected in mid- to far- bands by telescopes like Spitzer, reflect the thermal emission from grains heated by stellar and positioned at distances where pressure balances other forces, such as in disks or envelopes. Analogous to the in the solar system—produced by sunlight scattered off interplanetary under similar pressure effects—interstellar counterparts manifest as diffuse glows tracing streams. Multi-wavelength effects further modulate these interactions, with radiation from young stars exerting strong direct pressure on via high-efficiency (Q_pr ≈ 1-2), driving rapid expulsion of small grains. In contrast, re-emission from heated provides a more isotropic, lower-momentum transfer, sustaining longer-term dynamics in cooler regions and contributing to the overall pressure balance in star-forming environments. This UV-dominated push versus IR-mediated drag highlights the wavelength-dependent role in shaping distributions around evolving stellar sources.

Star formation and clusters

Radiation pressure plays a crucial role in the feedback processes during within molecular clouds, particularly by interacting with grains in protostellar outflows to regulate accretion onto forming stars. In these outflows, radiation from the central exerts pressure on , which can reprocess and redirect the , driving material away and halting further infall when the outward force balances or exceeds gravitational attraction. This feedback mechanism is especially effective in dense environments with surface densities below approximately 10^3 M_⊙ pc^{-2}, where it limits the efficiency of by ejecting significant fractions of the cloud mass. A key consequence of this radiation-driven is the imposition of an upper mass limit on stars, analogous to the Eddington limit, where the -supported radiation pressure prevents accretion beyond roughly 100 M_⊙. For massive protostars, the Eddington barrier arises as the star's increasing accelerates and gas outward, reducing net accretion rates and growth at masses around 30–50 M_⊙ when combined with outflows, though theoretical models suggest radiation alone caps it near 100 M_⊙. This limit ensures that very massive stars form only under specific conditions, such as high from infalling filaments that temporarily overcome the barrier. In young star clusters, collective radiation pressure from multiple forming stars disperses the surrounding molecular envelopes, which can either trigger rapid by compressing nearby gas or quench it by clearing material too efficiently. A prominent example is the Cluster, where intense radiation from the O-star θ¹ Ori C drives photoevaporation of proplyd envelopes at rates of about 0.4 × 10^{-6} M_⊙ yr^{-1}, leading to rapid mass loss and potential truncation of disk evolution within 10^4 years for low-mass proplyds. This dispersal quenches further accretion in the inner cluster regions while possibly inducing triggered formation in the outer envelope through shock-induced compression. Recent studies as of 2024 have explored the effect of radiation pressure on the dispersal of photoevaporating protoplanetary disks, investigating whether radiation-pressure-driven outflows can remove enough dust to align with observational data. Radiative acceleration in protostellar accretion disks further modulates by providing an outward force that balances , allowing disks to persist despite the intense from the central object. In massive , the disk's flattened geometry reduces the effective radiative force in the radial direction compared to spherical accretion, enabling sustained high accretion rates (>10^{-3} M_⊙ yr^{-1}) necessary for building stellar masses above 8 M_⊙. This balance prevents premature disk disruption and supports the transport of outward, facilitating material delivery to the star. Numerical simulations incorporating radiation hydrodynamics demonstrate how pressure influences cloud dynamics, often supporting molecular clouds against and altering the fragmentation into stars. Adaptive mesh refinement () models of massive cloud collapse show that radiation feedback reduces accretion onto protostars by up to 50% in high-luminosity phases, while also stabilizing cloud cores by providing turbulent support equivalent to in some cases. These simulations predict that radiation pressure disperses low-density envelopes early, favoring the formation of clustered massive stars over isolated ones. Observational evidence for these effects comes from polarized light in star-forming regions, where dust grain under pressure gradients reveals the anisotropic fields driving . In protostellar cores, submillimeter polarization maps show elongated dust grains aligned to flux directions, indicating radiative torques that respond to pressure-induced asymmetries and support models of outflow launching. Such observations in regions like confirm that pressure gradients enhance dust , providing indirect tracers of the regulating formation.

Galactic formation and evolution

In active galactic nuclei (AGN), radiation pressure from quasars expels interstellar gas through momentum-driven outflows, where the momentum flux is approximately P \approx L_{\rm AGN}/c with L_{\rm AGN} the AGN luminosity and c the , thereby regulating growth by limiting accretion rates. These outflows also quench in host galaxies by sweeping away molecular gas reservoirs, preventing further collapse into stars and maintaining a balance between black hole accretion and galactic evolution. Observational evidence from fast outflows in quasars supports this feedback mechanism, showing that radiation pressure on grains accelerates material to velocities exceeding 0.1c, influencing gas dynamics on kiloparsec scales. In barred galaxies, radiation pressure acting on dust grains within prominent dust lanes contributes to driving bar instabilities by perturbing gas orbits and enhancing non-axisymmetric perturbations in the . Dust lanes, aligned along leading edges of the , experience compressive shocks where radiation force amplifies gravitational instabilities, promoting bar strengthening and elongation over dynamical timescales of several hundred million years. This process alters the overall disc morphology, as the coupled dust-gas response under pressure facilitates transport and fuels central activity. During galaxy mergers, radiation pressure enhances outflows that clear dense gas clouds, facilitating the coalescence of supermassive s by reducing stalling. In gas-rich mergers, AGN-driven outflows powered by on dust remove intervening material along the inspiral path, allowing binaries to harden and merge via emission without prolonged gas torques. Simulations of major mergers demonstrate that this prevents excessive growth while promoting post-merger morphological transformations, such as the formation of elliptical remnants. Radiation pressure plays a key evolutionary role in high-redshift dusty , aiding morphological evolution by regulating clump migration and disc stability during the peak of cosmic at z ≈ 2–3. In these obscured systems, pressure on grains disperses giant molecular clouds, preventing excessive central concentration and fostering the transition from clumpy, irregular discs to more structured spirals over gigayears. This mechanism suppresses over-quenching, allowing sustained while -reprocessed influences overall galaxy assembly. Numerical models incorporating radiation pressure into N-body + hydrodynamic simulations reveal its importance for bulge formation, as momentum transfer from and AGN compresses central gas, driving bar dissolution and spherical bulge growth. In cosmological contexts, these simulations show that radiation enhances vertical in discs, stabilizing against fragmentation and channeling into dense bulges with masses exceeding 10^{10} M_\odot. Adaptive mesh refinement approaches confirm that omitting underpredicts bulge-to-disc ratios in high-z progenitors by factors of 2–3.

Early universe implications

In the radiation-dominated era of the early , shortly after the and lasting until approximately 47,000 years post-inflation, radiation pressure played a pivotal role in driving the isotropic expansion of the . The pressure exerted by relativistic particles, primarily photons and neutrinos, is given by P = \frac{u}{3}, where u is the of the radiation field. This , with w = \frac{1}{3}, results in a scale factor a(t) \propto t^{1/2} in a flat , contrasting with the later matter-dominated where w = 0 and expansion slows differently. This pressure ensured a smooth, homogeneous expansion before began to dominate around z \approx 3000, influencing the initial conditions for subsequent . Following recombination at around 380,000 years after the , when the cooled sufficiently for electrons and protons to form neutral , radiation pressure continued to affect perturbations through photon-baryon interactions. In the post-recombination epoch, the (CMB) photons exerted pressure that damped (BAO) on small scales, typically below 100 Mpc, by resisting and smoothing out fluctuations. This damping arises from the finite thickness of the recombination surface and the of photons, which random-walked over distances comparable to the sound horizon, suppressing power in the field. Closely related is Silk damping, a viscous effect from the tight coupling between photons and baryons before full , which erases primordial fluctuations on even smaller scales of about 10 Mpc or less through photon in an expanding . These processes, first theoretically described in the context of in opaque media, fundamentally shape the initial spectrum of cosmic perturbations. During the epoch of reionization, beginning around redshift z \approx 10-15 when the first stars and galaxies formed, ultraviolet radiation from these sources not only ionized neutral hydrogen but also generated radiation pressure that pushed surrounding neutral gas outward. This feedback mechanism, driven by momentum transfer from absorbing UV photons, expelled dense gas from star-forming regions, regulating early star formation and contributing to the inhomogeneous reionization of the intergalactic medium. Recent research as of 2024 emphasizes the role of Lyman-α radiation pressure feedback at Cosmic Dawn, which may inject up to 100 times more momentum than other mechanisms, acting as a dominant form of early stellar feedback. The pressure effects helped create ionized bubbles around proto-galaxies, with photoionization timescales shorter than those for full momentum deposition, leading to an initial rapid clearing of neutral hydrogen before sustained outflows. Observationally, these radiation pressure effects are imprinted in the power spectrum, where anisotropies on small angular scales (high multipoles l > 1000) exhibit exponential damping consistent with and processes, constraining cosmological parameters like the density \Omega_b and the fluctuation amplitude. The acoustic peaks in the temperature and power spectra, modulated by photon pressure during the pre-recombination era, provide precise measurements of these imprints, with data from missions like Planck confirming the suppression of power on scales below the diffusion length. signatures appear as a large-scale bump in the E-mode power spectrum at low l \approx 10, further linking early radiation pressure to the \tau \approx 0.054. These constraints highlight radiation pressure's role in bridging physics to cosmic structure.

Planetary systems and comets

In planetary systems around other stars, radiation pressure plays a key role in shaping the distribution of exozodiacal , which is analogous to the observed in our Solar System from scattered off interplanetary particles. This , primarily composed of small grains produced by collisions in outer debris belts or cometary activity, migrates inward through mechanisms like Poynting-Robertson drag before being cleared from the inner regions (<1 AU) by stellar radiation pressure. Grains smaller than the blowout size (typically <1–10 μm, depending on composition and stellar luminosity) experience a net outward force exceeding stellar gravity, leading to their rapid ejection on timescales of years, preventing accumulation near the star and maintaining low dust densities in habitable zones. Observations of hot exozodiacal around stars like Vega confirm this process, with detected emission requiring continuous replenishment rates of ~10^{-9} M_\oplus/yr to sustain the observed flux against expulsion. Radiation pressure also influences planetary migration by exerting subtle torques on protoplanetary dust in evolving disks, altering the gas-dust dynamics that drive orbital changes. In transitional protoplanetary disks, the outward push on small dust grains counters inward radial drift due to gas drag, creating size-sorted populations and clumping that modify the disk's surface density profile. This redistribution generates asymmetric torques on embedded protoplanets, potentially slowing inward type I migration or inducing outward shifts, particularly for low-mass planets interacting with dust-rich regions. Such effects are most pronounced during disk evolution stages where radiation pressure contributes to cavity opening and recession, indirectly shaping the final orbital architecture of planetary systems by trapping planets at pressure maxima. For comets, radiation pressure enhances outgassing by accelerating icy grains away from the nucleus, which increases the effective surface area exposed to solar heating and alters cometary through momentum transfer. Upon sublimation, volatile ices release dust aggregates that, if small enough (β > 0.1, where β is the ratio of radiation force to ), are promptly deflected antisunward, boosting expansion and contributing to non-gravitational observed in trajectory deviations. In the case of Comet 2P/Encke, radiation pressure on its dust trail creates a depleted gap in the particle distribution, as sub-micron grains are blown out while larger ones remain bound, leading to a structured stream that intersects and produces the Taurid meteor complex. The around exemplifies how radiation pressure interacts with inclined dust to produce observed warped structures. Small grains in the inner disk (<50 AU), launched from planetesimal collisions, experience enhanced radiation forces when tilted relative to the midplane, amplifying vertical excursions and bending the disk's geometry over time. This results in the characteristic inner warp detected at ~30–50 AU, where the disk deviates by ~5° from planarity, contrasting with the flatter outer regions and indicating pressure-driven sculpting superimposed on planetary gravitational influences. Simulations including radiation pressure show that such warping intensifies after ~2 Myr, distorting the disk's overall flatness and contributing to the observed asymmetry in scattered light images. In distant reservoirs like the Oort cloud, radiation pressure induces long-term perturbations primarily on dust components released from icy bodies, subtly eroding the reservoir over gigayears. Particles ejected during cometary passages near the Sun (D ≲ 10 μm) are blown out of the Solar System on hyperbolic trajectories, reducing the dust-to-ice ratio in surviving Oort cloud objects and influencing the dynamical stability of the cloud against external galactic tides. For larger grains (D ≳ 1 mm), cumulative effects over multiple orbits lead to gradual orbital expansion, potentially populating the inner Oort cloud with altered inclinations and contributing to the influx of long-period comets with perturbed dust envelopes. These processes ensure that radiation pressure acts as a selective filter, preserving larger aggregates while dispersing fine dust, which shapes the long-term evolution of comet reservoirs in mature planetary systems.

Modern Applications

Optical tweezers

Optical tweezers utilize focused laser beams to exert radiation pressure for the non-contact manipulation of microscopic particles, leveraging both gradient and scattering forces to achieve stable trapping. The technique was pioneered by in 1970, who first demonstrated the acceleration and trapping of micron-sized dielectric particles using the radiation pressure from continuous-wave laser beams, marking the initial observation of optical scattering and gradient forces on such particles. In 1986, Ashkin advanced the method by inventing the single-beam gradient force optical trap, employing a tightly focused laser beam through a high-numerical-aperture objective to stably confine transparent particles against gravitational and other forces. The trapping mechanism arises from two primary components of the radiation pressure force acting on a particle in a Gaussian laser beam. The gradient force, \mathbf{F}_\text{grad} \propto \nabla (\alpha I), where \alpha is the particle's real polarizability and I is the local light intensity, attracts dielectric particles toward the region of highest intensity at the beam focus, enabling three-dimensional confinement for particles with refractive index higher than the surrounding medium. The scattering force, \mathbf{F}_\text{scat} = \frac{n}{c} \sigma I, where n is the refractive index of the medium, c is the speed of light in vacuum, and \sigma is the particle's scattering cross-section, propels particles along the beam propagation direction due to momentum transfer from absorbed or scattered photons; in stable traps, this forward force is balanced by the axial gradient component pulling backward. These forces, typically on the order of piconewtons, allow precise control over particle position with sub-nanometer resolution. Optical tweezers have found extensive applications in biophysics and nanotechnology, including the trapping and manipulation of living cells such as bacteria and sperm, DNA molecules for stretching and unzipping studies, and nanoparticles for assembly into structures. For instance, they enable measurement of forces in molecular motors like kinesin, revealing step sizes of 8 nm during microtubule transport. The development of optical tweezers earned Arthur Ashkin half of the 2018 Nobel Prize in Physics, shared with Gérard Mourou and Donna Strickland, recognizing their transformative impact on biological and physical sciences. Various configurations enhance versatility, with single-beam traps suitable for isolating individual particles and holographic optical tweezers enabling simultaneous creation of multiple, independently addressable traps through computer-generated phase holograms displayed on spatial light modulators. This holographic approach, developed in the late 1990s, facilitates complex manipulations such as rotating arrays of particles or sorting heterogeneous mixtures without mechanical contact. Despite their precision, optical tweezers are limited by potential sample heating from laser absorption and photodamage from high intensities, particularly exceeding $10^6 W/m², which can denature biological molecules or kill cells through thermal or photochemical effects. Wavelengths in the near-infrared (e.g., 1064 nm) minimize such damage by reducing absorption in aqueous media, but power levels must be carefully calibrated to avoid cytotoxicity during extended trapping.

Laser-matter interactions

In intense laser-matter interactions, radiation pressure manifests through nonlinear effects when ultrashort, high-power laser pulses couple to plasmas or solid targets, driving complex dynamics beyond linear photon momentum transfer. At sufficiently high intensities, the oscillating electric field of the laser imparts a time-averaged force on charged particles, leading to expulsion from high-intensity regions and subsequent plasma response. This regime is particularly prominent in petawatt-class lasers, where relativistic effects dominate, enabling applications in particle acceleration and coherent radiation sources. The ponderomotive force is central to these interactions, representing the nonlinear response of electrons to the inhomogeneous laser field. In the non-relativistic limit, it is expressed as \mathbf{F}_\text{pond} = -\frac{e^2}{4 m_e \omega^2} \nabla \langle E^2 \rangle, where e and m_e are the electron charge and mass, \omega is the laser angular frequency, and \langle E^2 \rangle is the time-averaged squared electric field amplitude. This force expels electrons from regions of peak intensity, creating charge separation that can form plasma density channels or cavities. At relativistic intensities, defined by the normalized vector potential a_0 \approx 1 when I \lambda^2 > 10^{18} W cm^{-2} \mum^{2} (with I the intensity and \lambda the wavelength in \mum), the electron quiver velocity approaches the speed of light (v_\text{osc} \approx c), enhancing the force and introducing relativistic corrections that modify plasma transparency and wave propagation. A key application is laser-driven ion acceleration via radiation pressure, where the laser pulse reflects off a thin , exerting a forward pressure gradient that accelerates the entire ion layer coherently. In the radiation pressure acceleration (RPA) regime, typically accessed with circularly polarized to minimize electron heating, ions in ultrathin foils (\sim 10 nm) can reach GeV energies over micrometer scales, as demonstrated in simulations and early experiments with petawatt lasers. For instance, stable GeV proton beams have been predicted for under intensities exceeding $10^{20} W cm^{-2}, with the pressure balancing the target expansion to maintain monoenergetic output. High-harmonic generation (HHG) from plasma surfaces is also indirectly influenced by radiation pressure, which modulates the reflecting plasma mirror to achieve phase matching and coherence. By compressing large-scale plasma surfaces through radiation pressure, the Doppler upshift of reflected harmonics is enhanced, enabling efficient attosecond pulse production. Recent experiments with relativistic lasers have verified this mechanism, showing orders-of-magnitude improvements in harmonic yield when pressure-induced surface modulation balances relativistic oscillation effects. In the 2020s, petawatt laser facilities such as those at ELI-NP and have demonstrated in experiments pressures exceeding $10^{15} Pa (petapascal scale) in overdense plasmas, corresponding to intensities above $10^{20} W cm^{-2} for near-perfect reflection (P \approx 2I/c). These conditions have enabled observations of RPA-driven GeV-scale ions in thin-foil targets, confirming theoretical predictions and advancing compact accelerator concepts.

Emerging technologies

In , radiation pressure enables all-optical switching within metamaterials and optomechanical cavities, where light-induced forces modulate optical properties without electronic intermediaries. For instance, in photonic metamaterials, resonant optomechanical forces drive giant nonlinear responses, allowing control over and at the nanoscale, with forces on the order of femtonewtons (fN) arising from cavity-enhanced interactions. These systems leverage radiation pressure to achieve , switching, and memory functions in optomechanical metamaterials, where mechanical resonances couple to photonic modes for tunable electromagnetic responses. In gravitational wave detectors like the Laser Interferometer Gravitational-Wave Observatory (LIGO), radiation pressure noise from quantum fluctuations in the laser beam limits sensitivity at low frequencies, but advanced mitigation techniques employing frequency-dependent squeezing reduce this noise. Squeezed vacuum states injected into the interferometer suppress quantum radiation pressure noise by reshaping the uncertainty in the light's quadratures, achieving up to 4.0 dB noise reduction near 1 kHz in operational detectors. This approach enhances broadband sensitivity, enabling clearer detection of from events such as mergers. Quantum utilizes radiation pressure to cool mechanical resonators to their , bridging classical and quantum regimes in macroscopic systems. In a seminal experiment, a nanomechanical oscillator at 3.68 GHz was cooled from a bath temperature of 20 K to its using optical radiation pressure in a setup, demonstrating cooling where anti-Stokes removes phonons efficiently. This technique, relying on the dynamical backaction of cavity-enhanced , has paved the way for preparation and coherent control of mechanical modes, with applications in processing. Emerging medical applications harness pure pressure for , particularly through optical trapping of nanoparticles to achieve precise spatiotemporal control. nanoparticles, manipulated by focused beams exerting forces, enable intracellular of therapeutic agents, enhancing efficacy in photothermal and while minimizing off-target effects. These optomechanical interactions allow nanoparticles to be guided to specific cellular sites, facilitating controlled release mechanisms driven by light-induced momentum transfer. Looking to the future, concepts for probes propelled by sails build on radiation pressure principles, aiming for relativistic speeds beyond limitations. Although Starshot initiative is on indefinite hold as of September 2025 due to funding challenges, recent 2025 research has advanced ultra-thin lightsail designs, such as scalable nanotechnology-based membranes and pentagonal mirrors, to better withstand intense fluxes. These developments emphasize scalable photon propulsion for potential deep-space exploration.

References

  1. [1]
  2. [2]
    [PDF] On the History of the Radiation Reaction1 - Kirk T. McDonald
    Jan 9, 2020 · Apparently, Kepler considered the pointing of comets' tails away from the Sun as evidence for radiation pressure of light [2].2 Following ...
  3. [3]
    Stellar Radiation Pressure - HyperPhysics
    If you consider the hot interior of a star, the radiation energy density can be related to the radiation pressure which can act to prevent further gravitational ...
  4. [4]
    Advanced Composite Solar Sail System (ACS3) - NASA
    The unfurled solar sail is approximately 30 feet (about 9 meters) on a side. Since solar radiation pressure is small, the solar sail must be large to ...
  5. [5]
    Nichols' radiometer and discovering that radiation exerts pressure
    Apr 5, 2024 · ... Maxwell in 1873 when he unified electricity and magnetism. Shortly after this, Adolfo Bartoli suggested that radiation must have a pressure.
  6. [6]
    A treatise on electricity and magnetism : Maxwell, James Clerk, 1831 ...
    Mar 16, 2006 · Publication date: 1873. Topics: Electricity, Magnetism, Electricity, Magnetics. Publisher: Oxford : Clarendon Press.
  7. [7]
    Cavity optomechanics | Rev. Mod. Phys.
    Dec 30, 2014 · The first unambiguous experimental demonstrations of the radiation-pressure force predicted by Maxwell were performed using a light mill ...Missing: history | Show results with:history
  8. [8]
  9. [9]
    The Pressure Due to Radiation. (Second Paper.) | Phys. Rev. (Series I)
    The Pressure Due to Radiation. (Second Paper.) EF Nichols and GF Hull PDF Share Phys. Rev. (Series I) 17, 26 – Published 1 July, 1903
  10. [10]
    Radiation Pressure 1 | Nature
    The existence of pressure due to waves, though held by Euler, seems to have dropped out of sight until Maxwell, in 1872, predicted its existence as a ...
  11. [11]
    Radiation Through a Foggy Atmosphere
    6 ARTHUR SCHUSTER As for x = o, B = o, and A = R0, it follows that a = R0; or replacing BbyA-R, 2(A-R0) -sR0x When x = -I, the value of A is equal to S, the ...Missing: 1915 | Show results with:1915
  12. [12]
    Pyotr Lebedev - Physics Today
    Feb 24, 2015 · His discovery in 1899 of radiation pressure constituted the first experimental vindication of a prediction of James Clerk Maxwell's theory of ...<|control11|><|separator|>
  13. [13]
    27 Field Energy and Field Momentum - Feynman Lectures
    ... g is 1/c2 times the Poynting vector: g=1c2S. So the Poynting vector gives not only energy flow but, if you divide by c2, also the momentum density.
  14. [14]
    [PDF] Chapter 5 Energy and Momentum
    To calculate the radiation pressure P we integrate Maxwell's stress tensor on an infinite planar surface A parallel to the interface as shown in Fig. 5.2 ...
  15. [15]
    [PDF] 3-1 3. Radiation - Center for Astrophysics
    2q = 1/3. The radiation pressure of an isotropic radiation field is one third the energy density: P = 1. 3 u . Page 8. 3-8. 3.5 Flux from a sphere of uniform ...
  16. [16]
    Radiation pressure - MIT
    The radiation pressure is defined as the force per area that a wave exerts on a surface. If the wave is fully absorbed: P r a d a b s = F A = I c \mathcal P ...Missing: absorbing | Show results with:absorbing
  17. [17]
    [PDF] 24.6 - momentum and radiation pressure
    5 Finally, the radiation pressure exerted on a perfect reflecting surface for normal incidence of the wave is twice that given by Equation 24.33, or P 2S/c.
  18. [18]
    [PDF] Concerning an Heuristic Point of View Toward the Emission and ...
    Energy quanta penetrate into the surface layer of the body, and their energy is transformed, at least in part, into kinetic energy of electrons. The simplest ...
  19. [19]
    The Radiation Pressure of Light: historical perspectives and the role ...
    Apr 26, 2025 · Kepler's concept of radiation pressure was speculative and rooted in an intuitive grasp of the forces emanating from the Sun, yet it bore a ...
  20. [20]
    Radiation pressure | tec-science
    May 27, 2019 · The energy E in Einstein's equation denotes the energy Wp of a photon and the velocity v in the formula for the momentum, denotes the speed of ...
  21. [21]
    [PDF] 1 Physikalische Zeitschrift Vol. 10. No. 22, pg. 817 (1909) A. Einstein ...
    Those variations of radiation pressure, in order that thermal equilibrium be possible, must now be so constituted that, on average, they compensate for the ...
  22. [22]
    34 Relativistic Effects in Radiation - Feynman Lectures
    When light is emitted from a source there is a recoil effect: the same thing in reverse. If an atom is emitting an energy W in some direction, then there is a ...
  23. [23]
    [PDF] RADIATIVE PROCESSES IN ASTROPHYSICS
    The radiation pressure of an isotropic radiation field is one-third the energy density. This result will be useful in discussing the thermodynamics of ...
  24. [24]
    [PDF] Today in Astronomy 111: radiation forces on Solar system bodies ...
    Nov 12, 2024 · Also related to the Poynting-Robertson effect is the Yarkovsky effect, which can be significant for bodies large enough to have non-uniform ...
  25. [25]
    I. O. Yarkovsky and the Discovery of 'his' Effect - NASA ADS
    ... Yarkovsky effect. Calculating the radiation pressure on small bodies, he noted that if a body exposed to solar radiation “is so small that it is practically ...
  26. [26]
    Radiation force on a relativistic plasma and the Eddington limit - ADS
    Calculations of the radiation force due to Thomson scattering on a relativistic plasma ... Radiation Pressure;; Relativistic Plasmas;; Thomson Scattering ...
  27. [27]
    [PDF] Thermodynamics of Radiation
    The behaviour of such photon “gases” in thermal equilibrium can be analysed experimentally by measuring the thermodynamical properties of the electro-magnetic ...
  28. [28]
    [PDF] Thermodynamics of blackbody radiation - Robert E. Kelly - DF-UBA
    The thermodynamics of homogeneous, isotropic, unpolarized electromagnetic radiation in a cavity with volume and temperature controllable as the independent.
  29. [29]
    [PDF] Thermodynamics of Radiation Pressure and Photon Momentum
    The goal of the present paper is to review the physics of blackbody radiation and to draw a few conclusions concerning radiation pressure from the principles of ...
  30. [30]
    GRAVITATIONAL INSTABILITY IN RADIATION ...
    ABSTRACT. I consider the physics of gravitational instabilities in the presence of dynamically important radiation pressure and gray radiative diffusion.
  31. [31]
    [PDF] Radiation Diffusion: An Overview of Physical and Numerical Concepts
    Jan 18, 2005 · Radiation pressure, changes is spectral shape due to moving ... that diffusion predicts faster than light flow speeds in transparent media.
  32. [32]
    Solar Constant - an overview | ScienceDirect Topics
    The latest value of the solar constant is 1366.1 W m−2 [12]. The spectral distribution of the solar constant at TOA is given in Appendix A. The calculation of ...
  33. [33]
    [PDF] arXiv:1205.4762v1 [astro-ph.IM] 21 May 2012
    May 21, 2012 · At 1 AU, this is only 4.6 to 9.2 μN per m2 of difference in effective areas of the two spacecraft normal to the Sun, depending on the reflection ...Missing: m² | Show results with:m²
  34. [34]
    [PDF] Solar Sail Propulsion – Ready for SmallSat Mission Implementation
    Characteristic Acceleration (Ac): acceleration experienced at 1 AU with the sail normal to the Sun. • With 9.08 μN/m2 radiation pressure at 1 AU.Missing: m² | Show results with:m²
  35. [35]
    [PDF] The Physics of Solar Sails
    ... the Sun experiences a light pressure of. 9.1 µN/m. 2 . It is often wrongly stated that particle pressure from the “solar wind” powers sail spacecraft. Solar ...
  36. [36]
    Dynamics of the charged particles released from a Sun‐grazing ...
    Jun 7, 2021 · In Figure 1, the ratios of radiation pressure to gravity (β) shows an anti-correlation with dust particle diameter, which is greater than 1 μm.
  37. [37]
    16.4 Momentum and Radiation Pressure - UCF Pressbooks
    The radiation pressure p rad applied by an electromagnetic wave on a perfectly absorbing surface turns out to be equal to the energy density of the wave:.<|control11|><|separator|>
  38. [38]
    [PDF] RADIATION PRESSURE ON FLUFFY SUBMICRON-SIZED GRAINS
    Feb 16, 2016 · ABSTRACT. We investigate the claim that the ratio β of radiation pressure force to gravitational force on a dust grain in our solar.
  39. [39]
    [PDF] arXiv:1508.00646v2 [astro-ph.EP] 7 Mar 2016
    Mar 7, 2016 · In this paper we use two different codes to calculate radiation pressure on dust grains. First, we use the Mul- tiple Sphere T Matrix (MSTM) ...
  40. [40]
    The polarimetric properties of cometary dust and a possible effect of ...
    The polarisation is caused mainly by scattering of solar radiation by cometary dust grains. At certain wavelengths, however, the polarisation features, are ...
  41. [41]
    Solar Radiation Pressure - an overview | ScienceDirect Topics
    Solar radiation pressure (SRP) is the force caused by the exchange in momenta between the photons emitted by the Sun and the satellite's surface.
  42. [42]
    [PDF] An Analytical Theory for the Perturbative Effect of Solar Radiation ...
    Solar radiation pressure is the largest non-gravitational perturbation for most satellites, significantly influencing their orbital dynamics.
  43. [43]
    [PDF] RADIATION-PRE SSURE AND AIR-DRAG EFFECTS ON THE ...
    Jan 18, 1974 · Solar radiation pressure on high altitude satellites. Aerospace Res. Lab. Tech. Doc. Rep. ARL 72-0034, February, 9 pp. COOK, G. E.. 1965 ...<|control11|><|separator|>
  44. [44]
    The Physics of Solar Sails - NASA Technical Reports Server (NTRS)
    Apr 1, 2003 · The concept of using photon pressure for propulsion has been considered since Tsiolkovsky in 1921. In fact, Tsiolkovsky and Tsander wrote of ...
  45. [45]
    Small Solar Power Sail Demonstrator 'IKAROS' Confirmation ... - JAXA
    Jul 9, 2010 · The thrust by solar light pressure is 1.12 mili-Newton (*2,) which is the expected value. With this confirmation, the IKAROS was proved to ...
  46. [46]
    [PDF] Solar Sail Propulsion - NASA Technical Reports Server (NTRS)
    Solar Sailing was initially developed at JPL as a measure to save the Mariner 10 mission which had lost a large portion of its propellant margin.
  47. [47]
    [PDF] Solar Sails: Modeling, Estimation, and Trajectory Control - CCAR
    The sail lightness number β is defined as the ratio of the magnitude of the sail acceleration due to the solar radiation pressure and the acceleration ...
  48. [48]
    The future of solar sailing | The Planetary Society
    Mar 6, 2023 · The Planetary Society's LightSail 2 spacecraft sailed into history last year, successfully completing its mission to demonstrate flight by light.The Future Of Solar Sailing · A Useful Technology · Nasa Solar Sails<|separator|>
  49. [49]
    Potential Solar Sail Degradation Effects on Trajectory and Attitude ...
    The standard optical solar sail models that are currently used for trajectory and attitude control design do not take optical degradation into account, hence ...Missing: UV | Show results with:UV
  50. [50]
    Solar sail propulsion limitations due to hydrogen blistering
    May 1, 2021 · Blistering results in degradation of the reflective properties of the solar sail membrane. Specimens were tested at environmental conditions ...
  51. [51]
  52. [52]
  53. [53]
  54. [54]
    effect of radiation pressure on spatial distribution of dust inside H ii ...
    We investigate the impact of radiation pressure on spatial dust distribution inside H ii regions using one-dimensional radiation hydrodynamic simulations.2.4 Hydrodynamics · 2.4. 1 Dust And Gas Dynamics · 4.1 Dust Cavity RadiusMissing: excesses | Show results with:excesses
  55. [55]
    [PDF] Radiation Pressure and the Poynting–Robertson Effect for Fluffy ...
    When a dust particle is assumed to be spherical with a homogeneous structure, the solar radiation exerts a force on the particle in the direction of propagation ...
  56. [56]
    Interstellar Graines - B.T. Draine
    These radiation pressure forces can dominate the dynamics of gas in certain regions, such as dust-forming winds from cool stars. 15.3. H2 Formation on Grains.
  57. [57]
    The capture of interstellar dust: the pure Poynting–Robertson case
    To the usual equations of motion with Poynting–Robertson drag will be added the effect of evaporation of the particle. This comes about because the captured ...
  58. [58]
    A DYNAMICS CODE FOR A DUSTY INTERSTELLAR CLOUD ...
    Inside the cloud, the dust grains drift inward by radiation field and the dust-to-gas ratio at the intermediate radius increases by a factor of 20. Export ...
  59. [59]
    Driving gas shells with radiation pressure on dust ... - Oxford Academic
    Our goals are (1) to investigate the ability of AGN radiation pressure on dust to launch galactic outflows and (2) to constrain the efficiency of infrared (IR) ...
  60. [60]
    Spectral shifting strongly constrains molecular cloud disruption by ...
    In this paper we focus on radiation pressure on dust grains. The energy in the radiation field produced by a typical stellar population exceeds the ...Missing: seminal | Show results with:seminal
  61. [61]
    MID-INFRARED SPECTRA OF DUST DEBRIS AROUND MAIN ...
    ABSTRACT. We report spectra obtained with the Spitzer Space Telescope in the k ¼ 14 35 m range of 19 nearby main- sequence stars with infrared excesses.
  62. [62]
    Distribution of Interplanetary Dust Detected by the Juno Spacecraft ...
    Nov 11, 2020 · This cloud of dust particles can be observed in reflected sunlight (the Zodiacal light), emission of thermal radiation (Hauser et al., 1984; ...
  63. [63]
    Outflows driven by direct and reprocessed radiation pressure in ...
    Radiation pressure is a strong candidate to explain recently observed molecular outflows found in young super-star clusters in nearby starburst galaxies.
  64. [64]
    Outflows Driven by Direct and Reprocessed Radiation Pressure in ...
    Oct 6, 2022 · We quantify the relative importance of UV and IR radiation pressure in different regimes, and deduce that both are equally important for \Sigma ...
  65. [65]
    Investigating ultraviolet and infrared radiation through the turbulent ...
    Our goal is to analyze the UV and IR emission of a MC at different stages of its evolution and relate its absorption and emission properties with its ...
  66. [66]
    PROTOSTELLAR OUTFLOWS AND RADIATIVE FEEDBACK FROM ...
    The star-formation efficiency varies with the ratio of ejection to accretion rates from 50% in the case of very weak outflows to as low as 20% for very strong ...<|separator|>
  67. [67]
    [PDF] The Role of Outflows, Radiation Pressure, and Magnetic Fields in ...
    Stellar feedback in the form of radiation pressure and magnetically driven collimated outflows may limit the maximum mass that a star can achieve and affect ...
  68. [68]
    [PDF] HOW OUTFLOWS AND RADIATIVE FEEDBACK LIMIT ACCRETION ...
    High accretion rates and infalling filaments provide sufficient ram pressure to overcome radiation pressure. (20,000 AU)2. (3,000 AU)2 tff = 42.7 kyr. D.
  69. [69]
    [PDF] The Orion Nebula and its Associated Population - Astronomy in Hawaii
    The restricting force is produced by stellar radiation pressure pushing on the imbedded grains. Because it is well established that there is expansion away from ...
  70. [70]
    [PDF] ORION AND THEORIES OF STAR FORMATION - Yale Astronomy
    Star formation and its various effects are evidently responsible for many of the complex phenomena observed in the Orion Nebula region, and one would ...
  71. [71]
    On the Formation of Massive Stars - NASA ADS
    By enhancing the nonisotropic character of the radiation field, the accretion disk reduces the effects of radiative acceleration in the radial direction-a ...
  72. [72]
    RADIATION-HYDRODYNAMIC SIMULATIONS OF MASSIVE STAR ...
    In this paper we present a series of adaptive mesh refinement (AMR) radiation-hydrodynamic simulations of the collapse of massive star-forming clouds using the ...<|separator|>
  73. [73]
    Formation of Massive and Wide First-star Binaries in Radiation ...
    Nov 30, 2023 · We study the formation of Population III stars by performing radiation hydrodynamic simulations for three different initial clouds extracted ...
  74. [74]
    Recent progress in theory and observational study of dust grain ...
    Here, we review the recent progress in the theoretical development and observational testing of these effects using dust polarization observed toward star- ...
  75. [75]
    energetics of AGN radiation pressure-driven outflows
    We show that AGN radiation pressure on dust can adequately reproduce the observed outflow energetics (mass outflow rate, momentum flux, and kinetic power)
  76. [76]
    Momentum-driven Winds from Radiatively Efficient Black Hole ...
    Abstract. We explore the effect of momentum-driven winds representing radiation-pressure-driven outflows from accretion onto supermassive black holes in a set ...
  77. [77]
    Evidence for quasar fast outflows being accelerated at the scale of ...
    Feb 11, 2022 · Quasar outflows may play a crucial role in regulating the host galaxy, although the spatial scale of quasar outflows remain a major enigma, ...<|separator|>
  78. [78]
    Radiation pressure in galactic discs: stability, turbulence, and winds ...
    Abstract. The radiation force on dust grains may be dynamically important in driving turbulence and outflows in rapidly star-forming galaxies.Missing: lanes barred
  79. [79]
    GASEOUS STRUCTURES IN BARRED GALAXIES - IOP Science
    Dust lanes approximately follow one of the x1-orbits and tend to be straighter under a stronger and more elongated bar, but are insensitive to the presence of ...
  80. [80]
    Radiation-driven efflux and circulation of dust in spiral galaxies.
    Jan 1, 1972 · Mixing and escape also occur by means of instabilities of the galactic disk caused partially by radiation pressure on the grains. These ...Missing: barred | Show results with:barred
  81. [81]
    The growth of massive black holes in galaxy merger simulations with ...
    In this paper, we present simulations of major mergers of spiral galaxies using a model for the growth of BHs that includes (1) a BH accretion rate prescription ...
  82. [82]
    Electromagnetic counterparts to massive black-hole mergers - PMC
    ... radiation pressure inflate the accretion flow and produce outflows. In ... black hole ejected by gravitational-radiation recoil in a galaxy merger.
  83. [83]
    Optical and UV Flares from Binary Black Hole Mergers in Active ...
    Jul 13, 2024 · After the merger, the remnant is gravitationally kicked out, accretes from the unperturbed disc and produces a quasi-spherical outflow which ...
  84. [84]
    birth of a galaxy – II. The role of radiation pressure - Oxford Academic
    Here we present three adaptive mesh refinement radiation hydrodynamics simulations that illustrate the impact of momentum transfer from ionizing radiation to ...
  85. [85]
    Star formation and clumps in cosmological galaxy simulations with ...
    We study the global and morphological effects of radiation pressure (RP) in eight pairs of high-resolution cosmological galaxy formation simulations.
  86. [86]
    Star Formation and Clumps in Cosmological Galaxy Simulations ...
    Oct 21, 2014 · We study the global and morphological effects of radiation pressure (RP) in eight pairs of high-resolution cosmological galaxy formation ...
  87. [87]
    THE PRESSURE OF THE STAR-FORMING INTERSTELLAR ...
    We examine the pressure of the star-forming interstellar medium (ISM) of Milky-Way-sized disk galaxies using fully cosmological SPH+N-body, high-resolution ...
  88. [88]
    [PDF] Lecture VI: The radiation-dominated era
    The non-relativistic plasma (3000 <z< 2 × 109). The radiation is the dominant component of the Universe at these early times, so I will focus on that. We will ...
  89. [89]
    [PDF] Cosmic Microwave Background Anisotropies - Wayne Hu's Tutorials
    ABSTRACT: Cosmic microwave background (CMB) temperature anisotropies have and will continue to revolutionize our understanding of cosmology.
  90. [90]
    Stellar feedback by radiation pressure and photoionization
    Radiation pressure needs a comparatively long time period to deposit the momentum in the gas. The photoionization time-scales are shorter, yielding an early ...
  91. [91]
    Hot exozodiacal dust resolved around Vega with IOTA/IONIC
    Such grains would be expected to be expelled from the inner planetary system by radiation pres- sure within only a few years, which indicates inordinate replen ...
  92. [92]
    Unraveling the Water Sources in Comet 103P/Hartley 2 from Deep ...
    Apr 17, 2025 · 2O-ice-containing grains are sublimating after being pushed by solar radiation pressure in the antisunward direction and contribute to the ...
  93. [93]
    Dynamics of HVECs emitted from comet C/2011 L4 as observed by ...
    Jul 4, 2015 · If the ejected clumps are composed of dust grains, radiation pressure contributes to the initial size sorting of the dust expelled from the ...
  94. [94]
    A proposed alternative dynamical history for 2P/Encke that explains ...
    Observations of the comet's dust trail in the mid-infrared and optical revealed that the core of trail is more than 20 000 km wide and composed of particles in ...
  95. [95]
    DYNAMICS OF DUST PARTICLES RELEASED FROM OORT ...
    We find that OCC particles with diameters D ≲ 10 μm are blown out from the solar system by radiation pressure, while those with D ≳ 1 mm have a very low Earth- ...
  96. [96]
  97. [97]
    [PDF] Optical Tweezers and their Application to Biological Systems
    This axial gradient generates a light pressure force in the backward direction relative to the beam and of a magnitude that can exceed the forward scattering ...
  98. [98]
  99. [99]
    Optical Tweezers: Phototoxicity and Thermal Stress in Cells and ...
    In this review the main negative effects of optical tweezers upon biostructures and living systems will be presented.
  100. [100]
    [PDF] arXiv:2103.16406v2 [physics.optics] 4 Jun 2021
    Jun 4, 2021 · e2. 4mω2 ∇|E|2 ,. (2) where e, m, ω, and E are the elementary charge, mass of the electron, angular frequency of light, and electric field. Page ...
  101. [101]
    Nonlinear dielectric optomechanical metamaterials | Light - Nature
    Aug 30, 2013 · We introduce a dielectric photonic metamaterial presenting a giant nonlinear optical response driven by resonant optomechanical forces.
  102. [102]
    Isolation, switching and memory based on optomechanical ...
    We show that cavity optomechanics can provide a rich platform to obtain a range of electromagnetic properties and functionalities that are otherwise ...
  103. [103]
    Frequency-Dependent Squeezing for Advanced LIGO
    To simultaneously reduce shot noise at high frequencies and quantum radiation pressure noise at low frequencies requires a quantum noise filter cavity with low ...
  104. [104]
    Laser cooling of a nanomechanical oscillator into its quantum ... - arXiv
    Jun 18, 2011 · Starting from a bath temperature of T=20K, the 3.68GHz nanomechanical mode is cooled into its quantum mechanical ground state utilizing optical radiation ...Missing: optomechanics | Show results with:optomechanics
  105. [105]
    Starshot - Breakthrough Initiatives
    Breakthrough Starshot aims to demonstrate proof of concept for ultra-fast light-driven nanocrafts, and lay the foundations for a first launch to Alpha Centauri.Missing: 2025 | Show results with:2025
  106. [106]
    The Pressure to Explore: Caltech Researchers Take First ...
    Jan 31, 2025 · The concept is to use lasers to propel miniature space probes attached to "lightsails" to reach ultrafast speeds and eventually our nearest star ...
  107. [107]
    Space sails for achieving major space exploration goals
    Oct 1, 2024 · Among flown solar sails, only IKAROS has experienced measurable photon-induced acceleration and thrust in real flight: 0.004 mm/s 2 and 1.12 mN ...