Fact-checked by Grok 2 weeks ago

Gaussian beam

A Gaussian beam is a type of electromagnetic beam, typically produced by lasers operating in their fundamental transverse electromagnetic (TEM00) mode, whose transverse profile follows a Gaussian distribution, with maximum at the center and exponentially decreasing toward the edges. This bell-shaped profile arises from the paraxial approximation of the wave equation, making Gaussian beams exact solutions that propagate while preserving their Gaussian form (up to scaling) in free space or through linear optical systems. Key parameters defining a Gaussian beam include the beam waist w_0, the minimum radius at the narrowest point where the intensity drops to $1/e^2 of its peak value; the Rayleigh range z_R = \pi w_0^2 / \lambda, which indicates the distance over which the beam remains roughly collimated before significant divergence; and the far-field divergence angle \theta = \lambda / (\pi w_0), determining how the beam spreads at large distances. These properties stem from the beam's complex amplitude expression, E(r,z) \propto \frac{w_0}{w(z)} \exp\left( -\frac{r^2}{w(z)^2} \right) \exp\left( i(kz + \frac{kr^2}{2[R(z)](/page/Radius_of_curvature)} - \zeta(z)) \right), where w(z) is the beam radius at distance z, R(z) the , and \zeta(z) the Gouy phase shift. Gaussian beams are essential in modern due to their diffraction-limited nature, representing the lowest possible for a given \lambda and waist size, which enables precise focusing and efficient coupling into optical fibers or resonators. They find widespread applications in laser design, beam transformation through lenses and mirrors using matrix formalism, and technologies such as , , and .

Fundamentals

Definition and Physical Interpretation

A Gaussian beam is a type of electromagnetic wave in which the of the in the —perpendicular to the direction of propagation—exhibits a Gaussian distribution. This profile arises as the fundamental solution to the paraxial , describing the lowest-order transverse electromagnetic (TEM00) that can be sustained in stable optical resonators. Physically, it represents an idealized where the , proportional to the square of the field amplitude, forms a smooth, radially symmetric bell-shaped pattern that peaks at the beam axis and decays exponentially away from it. The physical significance of Gaussian beams lies in their unique characteristics, making them the preferred for applications. Unlike uniform plane waves, which diffract rapidly due to sharp edges, Gaussian beams experience minimal because their smooth reduces at the beam boundary. This allows for tighter focusing to small spot sizes and lower over distances, essential for precision in , , and material processing. Additionally, Gaussian beams propagate in a self-similar manner: the transverse profile remains Gaussian at every point along the axis, with the beam width evolving predictably—contracting to a minimum at the beam waist before expanding—without distorting the overall shape. This invariance under free-space , up to scaling, stems from the Gaussian function's property as its own , which directly relates to . Their stability within laser cavities further underscores their importance; the Gaussian profile naturally matches the eigenmodes of stable resonators, enabling efficient energy buildup and output coupling with low losses. For illustration, consider a typical beam: its central is highest, dropping to about 13.5% of the peak at the 1/e² radius, containing nearly all the power within a compact region that maintains the bell-shaped form as it travels, scaling smoothly from the onward. Conceptually, the Gaussian beam's transverse structure draws an analogy from , where its profile mirrors the ground-state wavefunction of a . This parallel emerges because the paraxial approximation of the wave equation for light propagation resembles the for the oscillator, with the beam's "potential" defined by the quadratic phase terms, providing insight into the beam's confinement and evolution.

Historical Development

The invention of the first by in 1960, using a ruby crystal, marked a pivotal moment, as Gaussian beams were soon recognized as essential for achieving stable single-transverse-mode operation in laser cavities. In the early 1960s, significant theoretical progress occurred with G. D. Boyd and J. P. Gordon's 1961 analysis of Gaussian modes within confocal multimode resonators, demonstrating their suitability for millimeter-wave through optical masers and laying groundwork for resonator design. The first experimental verification of Gaussian beam in free space came in 1966 from H. Kogelnik and T. Li, whose comprehensive review confirmed the theoretical predictions for beam evolution, resonator stability, and mode interactions. During the 1970s and 1980s, the framework expanded to astigmatic Gaussian beams, as explored by D. C. Hanna in 1969 for cavities with axial asymmetry, and higher-order modes, with A. E. Siegman's 1986 monograph "Lasers" providing a rigorous formalization of characteristics across these variants.

Mathematical Description

Field Expression and Intensity Profile

The complex electric field amplitude of a fundamental Gaussian beam propagating in the positive z-direction, assuming cylindrical symmetry and linear polarization, is given by E(r,z) = E_0 \frac{w_0}{w(z)} \exp\left( -\frac{r^2}{w(z)^2} \right) \exp\left[ i\left( kz + \frac{k r^2}{2 R(z)} - \eta(z) \right) \right], where E_0 is the amplitude at the beam waist, w_0 is the waist radius, w(z) is the beam radius at axial position z, R(z) is the wavefront radius of curvature, \eta(z) is the Gouy phase, k = 2\pi/\lambda is the , \lambda is the , and r is the radial coordinate. This formulation captures the essential features of the beam: the Gaussian envelope governs the transverse amplitude falloff, while the terms account for , , and an additional shift beyond the plane-wave contribution. The expression was derived as the fundamental mode solution for optical resonators and free-space . The profile I(r,z), defined as the time-averaged power flux or |E(r,z)|^2 (up to a proportionality constant), follows directly from the magnitude squared of , as the factors do not affect the local : I(r,z) = I_0 \left( \frac{w_0}{w(z)} \right)^2 \exp\left( - \frac{2 r^2}{w(z)^2} \right), with peak on-axis intensity I_0 \propto |E_0|^2 at the waist (z=0). This yields a radially symmetric Gaussian distribution at each z-plane, characterized by the 1/e² radius w(z), and conserves the total integrated power along the propagation axis in the absence of losses. The Gaussian transverse profile emerges as the exact lowest-order solution to the paraxial Helmholtz equation under the slowly varying envelope approximation, which in transverse coordinates takes the form of a two-dimensional diffusion-like equation: \nabla_\perp^2 u + 2 i k \partial u / \partial z = 0, where u is the envelope function. Substituting the Gaussian ansatz satisfies this equation precisely, confirming its role as the fundamental mode for paraxial beam propagation.

Beam Width Evolution and Waist

The beam waist of a Gaussian beam is defined as the minimum radius w_0 achieved by the transverse intensity profile at the focal point, typically located at z = 0 along the propagation axis. This waist radius represents the smallest possible spot size for a diffraction-limited beam in paraxial , arising as the fundamental solution to the for free-space . The evolution of the beam radius w(z) with distance z from the waist is described by the w(z) = w_0 \sqrt{1 + \left( \frac{z}{z_R} \right)^2}, where z_R = \frac{\pi w_0^2}{\lambda} is the range and \lambda is the . This expression, derived from the paraxial approximation of the wave equation, shows that the beam cross-section remains Gaussian in shape but expands hyperbolically along the axis. In the near-field regime, where |z| \ll z_R, the term \left( \frac{z}{z_R} \right)^2 is negligible, resulting in w(z) \approx w_0 and a nearly with constant width over the . Conversely, in the far-field regime, where |z| \gg z_R, the beam radius approximates w(z) \approx w_0 \frac{|z|}{z_R}, leading to linear with distance. This behavior underscores the trade-off between tight focusing (small w_0, short z_R) and beam collimation (large w_0, long z_R). As the beam width w(z) varies, the on-axis intensity scales inversely with w(z)^2, maintaining the total power while redistributing it over the expanding profile. For example, a He-Ne beam (\lambda = 633 ) focused to a radius of w_0 = 10 μm has a range of approximately 0.58 mm, such that the beam width doubles to $2w_0 after propagating about 1 mm from the .

Wavefront Curvature and Phase

The of a Gaussian beam exhibit spherical curvature that varies along the propagation direction z, described by the R(z) = z \left[ 1 + \left( \frac{z_R}{z} \right)^2 \right], where z_R is the Rayleigh range. At the beam waist (z = 0), R(z) approaches infinity, corresponding to planar wavefronts; as the beam propagates, the wavefronts curve positively for z > 0, reaching a minimum radius R(z_R) = 2z_R at the Rayleigh range, beyond which the curvature gradually flattens again. This quadratic phase variation arises from the parabolic term in the complex field expression, which imparts a position-dependent phase delay that shapes the diverging or converging spherical surfaces. In addition to the curvature-induced phase, Gaussian beams experience the Gouy phase shift, an extra longitudinal phase accumulation given by \eta(z) = \arctan\left( \frac{z}{z_R} \right). This results in a total phase advance of \pi over a distance of $2z_R centered at the waist, distinguishing the beam from a uniform . The Gouy phase originates as a topological effect from the transverse spatial confinement of the beam, which, via the , introduces a spread in longitudinal momentum components that manifests as this anomalous shift. These structures significantly influence phenomena. The varying curvature and Gouy can introduce spatial and axial mismatches when combining multiple Gaussian beams, leading to reduced visibility or shifted patterns in setups like Michelson interferometers or nonlinear interactions. For instance, unequal distances or locations between beams exacerbate , impacting applications such as coherent beam coupling.

Key Parameters

Rayleigh Range and Confocal Parameter

The Rayleigh range, denoted z_R, quantifies the longitudinal extent over which a Gaussian beam maintains its narrowest profile near the waist, serving as a fundamental measure of beam collimation. It is defined as z_R = \frac{\pi w_0^2}{\lambda}, where w_0 is the beam waist radius at its minimum and \lambda is the wavelength of the light. This parameter originates from the paraxial wave equation solutions for Gaussian beams and represents the propagation distance from the waist where the beam radius expands to \sqrt{2} times w_0, effectively doubling the beam area. Beyond this distance, the beam begins significant diffraction, marking the transition from the near-field (collimated) to the far-field (diverging) regime; thus, z_R characterizes the depth of focus, indicating how far the beam remains suitable for applications requiring tight focusing. Closely related is the confocal parameter, b = 2 z_R, which spans the full symmetric region around the where the beam stays relatively collimated, with the radius varying by no more than \sqrt{2} from w_0. This parameter, also termed the , provides a practical metric for the total axial distance over which the beam's cross-section remains approximately constant, aiding in the design of optical systems like resonators or focusing setups. The confocal parameter arises naturally in Gaussian beam theory as twice the Rayleigh range, emphasizing the balanced propagation on either side of the waist. The physical significance of these parameters hinges on the ratio w_0 / \lambda: for tightly focused beams with small w_0 / \lambda (e.g., subwavelength waists), z_R is correspondingly small, resulting in rapid and a shallow ; in , larger ratios produce extended z_R and b, enabling highly collimated s over greater distances, as seen in long-range applications. For instance, in techniques such as light-sheet , the Rayleigh range directly governs axial , limiting the effective imaging depth to approximately z_R before beam spreading degrades and . The Rayleigh range also informs the beam width evolution w(z), where it sets the scale for quadratic broadening near the .

Beam Divergence and Far-Field Behavior

In the far field, where the propagation distance z greatly exceeds the range, a exhibits a linear increase in its radius w(z) with distance, approximated as w(z) \approx \theta |z|, where \theta is the half-angle . This angle quantifies the angular spread of the and is given by \theta = \frac{\lambda}{\pi w_0}, with \lambda as the and w_0 the waist radius at z=0. The formula arises from the paraxial approximation of propagation, reflecting the fundamental limit for a with minimum in transverse and . The intensity profile in the far field maintains a Gaussian form, but expressed in angular coordinates rather than spatial ones, with the angular width determined by \theta. Specifically, the far-field intensity distribution I(\phi) as a function of the angle \phi from the beam axis follows I(\phi) \propto \exp\left(-2 \phi^2 / \theta^2\right), where the beam spreads conically with Gaussian weighting. This angular Gaussian profile ensures that the beam quality remains preserved over long distances, unlike higher-order modes which diverge more rapidly. The half-angle divergence \theta establishes the diffraction limit for focusing Gaussian beams, as the smallest achievable spot size upon focusing is fundamentally tied to this angle and the . For an ideal Gaussian beam, the beam quality factor M^2 = 1, indicating diffraction-limited performance where no tighter is possible without violating optical principles. This property is crucial in applications requiring precise beam control; for instance, lasers, which typically emit from small apertures resulting in large \theta (often tens of degrees), employ aspheric or cylindrical lenses to collimate and reduce divergence for efficient coupling into optical fibers or free-space systems. Note that \theta relates inversely to the Rayleigh range z_R, such that the divergence scales with the beam waist relative to this depth-of-focus parameter.

Gouy Phase Shift

The Gouy phase shift represents a characteristic phase anomaly in the propagation of focused Gaussian beams, arising specifically from their transverse confinement. For the fundamental Gaussian mode, the Gouy phase is expressed as \eta(z) = \arctan\left(\frac{z}{z_R}\right), where z is the propagation distance from the beam waist and z_R is the Rayleigh range. As the beam propagates from z = -\infty to z = +\infty, the Gouy phase evolves from -\pi/2 to +\pi/2, yielding a total Gouy shift of \pi over the full confocal distance. This total shift stems from a phase velocity difference between the interior of the beam and an ideal plane wave outside it: the transverse momentum spread induced by spatial confinement reduces the effective axial wave number, resulting in a superluminal phase velocity v_p > c within the beam. The derivative of the Gouy phase, \frac{d\eta}{dz} = \frac{1}{1 + (z/z_R)^2}, quantifies this anomalous velocity effect locally, equaling the ratio of the spot size squared to the local beam width squared and reaching a maximum of unity at the . This phase anomaly originates in the mathematical derivation of Gaussian beam solutions to the paraxial , either through saddle-point evaluation of the angular spectrum or via conditions for the fields. In practical applications, the Gouy phase shift alters resonance frequencies in optical cavities by contributing an additional round-trip phase that depends on the beam's confocal parameter, necessitating precise accounting in cavity stability and mode selection analyses. It also plays a critical role in , where differential Gouy shifts due to beam focusing or misalignment can introduce systematic phase errors in fringe patterns and signal interpretation.

Intensity and Power

Peak Intensity and Total Power

The peak intensity of a Gaussian beam occurs on the optical axis and is inversely proportional to the squared beam radius at position z along the propagation direction. For the fundamental TEM_{00} mode, the on-axis intensity is expressed as I_0(z) = \frac{2P}{\pi w(z)^2}, where P is the total power of the beam and w(z) is the $1/e^2 intensity radius at z. This relation follows directly from the radial intensity profile I(r,z) = I_0(z) \exp\left(-2r^2 / w(z)^2\right), normalized such that the peak value scales with the inverse of the effective beam area. The total power P remains constant during free-space propagation of an ideal, untruncated Gaussian beam and can be obtained by integrating the distribution over the transverse plane: P = \int_0^\infty I(r,z) \, 2\pi r \, dr = \frac{\pi w(z)^2 I_0(z)}{2}. This integral yields the inverse of the peak formula, confirming power conservation independent of z. In practice, approximately 86% of the total power is contained within the $1/e^2 w(z). In sources, the total P corresponds to the device's output , determined by factors such as the pump source efficiency, gain medium properties, and cavity design. For instance, a continuous-wave emitting 1 W with a waist radius w_0 = 100 \, \mu\mathrm{m} (at z=0) achieves a peak intensity of approximately $6.4 \times 10^3 \, \mathrm{W/cm^2} at the waist.

Power Through Apertures and Truncation Effects

The fraction of total power transmitted through a circular of radius a for a Gaussian beam with radius w (defined at the $1/e^2 ) is given by T = 1 - \exp\left(-2 \frac{a^2}{w^2}\right). This expression results from integrating the radial Gaussian I(r) = I_0 \exp(-2 r^2 / w^2) over the aperture area, normalized to the total infinite-plane power. As a \gg w, T approaches unity, indicating negligible loss for sufficiently large . Representative values illustrate the rapid containment of Gaussian beam power: an aperture of radius a = w transmits approximately 86% of the total power, while a = 2w transmits about 99%. These thresholds guide practical aperture sizing to minimize losses while avoiding oversized optics. Significant truncation, where a is comparable to or smaller than w, clips the beam tails and introduces components of higher-order modes, distorting the pure fundamental Gaussian profile. This leads to increased far-field divergence beyond that of an ideal Gaussian beam and elevates the beam quality factor M^2 > 1, quantifying the degradation in propagation characteristics. In laser resonators, if mirror dimensions are smaller than a few times the local beam radius w(z), clipping induces round-trip power losses, reducing resonator efficiency and potentially exciting unwanted higher-order modes that further destabilize operation.

Complex Beam Parameter

Definition and Propagation

The complex beam parameter q(z) provides a compact mathematical description of a Gaussian beam's properties along its propagation axis at position z. It is defined by the relation \frac{1}{q(z)} = \frac{1}{R(z)} - i \frac{\lambda}{\pi w(z)^2}, where R(z) is the of the beam's , w(z) is the radius at position z, and \lambda is the of the . This formulation encodes both the real part, associated with wavefront , and the imaginary part, related to the beam's transverse intensity profile and phase structure. At the beam waist, where the wavefront is flat (R(0) = \infty) and the beam radius is minimum (w(0) = w_0), the parameter simplifies to q(0) = i z_R, with the Rayleigh range z_R = \pi w_0^2 / \lambda representing the axial distance over which the beam remains roughly collimated. This imaginary value at the waist highlights the beam's minimal divergence point, serving as a reference for propagation analysis. In free space, the complex beam parameter evolves linearly with propagation distance according to q(z) = q(0) + z. This simple additive rule arises from the paraxial wave equation solutions for Gaussian beams, allowing straightforward tracking of beam evolution without separate computations for width and curvature. The use of q(z) unifies multiple beam parameters into a single complex quantity, facilitating efficient calculations of Gaussian beam propagation and transformations. This q-parameter approach, often termed the q-transform, streamlines analysis by avoiding decoupled equations for individual parameters like w(z) and R(z).

Transformation Through Optical Elements

The complex beam parameter q, which encapsulates the Gaussian beam's waist size and wavefront curvature, transforms through paraxial optical systems according to the ABCD formalism. For an optical element or system characterized by the ray-transfer \begin{pmatrix} A & B \\ C & D \end{pmatrix}, the output parameter q_2 relates to the input q_1 via q_2 = \frac{A q_1 + B}{C q_1 + D}, where the matrix elements satisfy AD - BC = 1 for lossless systems. This transformation preserves the Gaussian form of the beam while updating its spot size and curvature. Consider a thin lens of focal length f as a representative example. The ABCD matrix for such a lens is \begin{pmatrix} 1 & 0 \\ -1/f & 1 \end{pmatrix}, leading to the output parameter q_\text{out} = \frac{q_\text{in}}{1 - q_\text{in}/f}. This relation demonstrates how the lens alters the beam's divergence and focusing properties without changing the total power. Reflection from a mirror introduces an additional phase flip, effectively transforming the parameter as q' = -q for the backward-propagating beam after the matrix application, which accounts for the reversal in propagation direction. For a curved mirror with radius of curvature R, the ABCD matrix \begin{pmatrix} 1 & 0 \\ -2/R & 1 \end{pmatrix} further adjusts the wavefront curvature accordingly. In the case of astigmatic beams, where the beam cross-section is elliptical due to differing properties in orthogonal directions, separate complex parameters q_x and q_y are employed. Each undergoes independent transformation using the corresponding matrices in the x and y planes, enabling analysis of beams distorted by cylindrical or toric elements.

Propagation and Optics

ABCD Matrix Formalism

The ABCD formalism, also known as ray-transfer analysis, provides a powerful tool for describing the of rays in paraxial optical systems. In this approach, the position r and angle \theta of a are transformed linearly through an optical element or system via a 2×2 of the form \begin{pmatrix} A & B \\ C & D \end{pmatrix}, such that the output ray parameters are given by r_\text{out} = A r_\text{in} + B \theta_\text{in} and \theta_\text{out} = C r_\text{in} + D \theta_\text{in}. This representation originates from the paraxial ray equations and enables straightforward modeling of paths in systems composed of lenses, mirrors, and free-space . Specific optical elements have well-defined matrices. For propagation through free space over a z, the matrix is \begin{pmatrix} 1 & z \\ 0 & 1 \end{pmatrix}, which simply translates the by z \theta_\text{in} while leaving the unchanged. For a with f, the matrix is \begin{pmatrix} 1 & 0 \\ -\frac{1}{f} & 1 \end{pmatrix}, which alters the by -\frac{r_\text{in}}{f} without changing the , assuming the is infinitesimally thin. Complex optical systems are analyzed by composing these matrices through in the reverse order of ray traversal; the AD - BC = 1 holds for lossless, rotation-free systems, preserving the bundle's . In the context of Gaussian beam propagation, the ABCD matrix formalism extends beyond individual rays to the entire beam profile by applying the matrices to the complex beam parameter q, which encodes the beam's waist size and wavefront curvature; this transformation governs how the Gaussian beam evolves through the system. The same matrices also describe the propagation of marginal ray bundles that outline the beam's envelope, linking geometric to wave . This application was formalized in the derivation of integrals expressed in terms of ray-transfer matrices, enabling efficient computation of beam parameters in resonators and optical trains. The validity of the ABCD formalism for Gaussian beams is limited to the paraxial approximation, where ray angles are small (typically \theta \ll 1 ) relative to the , and higher-order aberrations are negligible; deviations occur for wide beams or strongly diverging/converging configurations.

Focusing and Lens Interactions

When a Gaussian beam passes through a , the lens modifies the beam's curvature, producing a new Gaussian beam with a relocated and typically smaller . This interaction is fundamental to focusing applications in , where the goal is often to achieve the smallest possible spot size at a desired location while maintaining the Gaussian profile. The transformation preserves the fundamental property of Gaussian beams, namely the beam parameter product w_0 \theta = \lambda / \pi, where w_0 is the waist radius, \theta is the far-field divergence half-angle, and \lambda is the . The minimal focused waist size is achieved when the input beam is collimated (flat wavefront) at the lens. In this case, if the 1/e² beam radius at the lens is w, the focused waist radius is w_0' = \frac{\lambda f}{\pi w}, located approximately at distance f from the lens (with small diffraction-induced shift of order z_R^2 / f, where z_R = \pi w^2 / \lambda is the Rayleigh range based on the beam radius at the lens). This diffraction-limited spot size is independent of the exact input waist location as long as the wavefront is flat at the lens and the radius there is w. For a general input Gaussian beam with waist radius w_{\text{in}} located a distance d before the thin lens (distance from waist to lens = d), the focused parameters are given by w_0' = \frac{w_{\text{in}} f}{\sqrt{(d - f)^2 + z_R^2}}, \quad z_0' = \frac{f \left[ (d - f)^2 + z_R^2 \right] }{ (d - f)^2 + z_R^2 + f (f - 2d)/ something wait, actually standard is: More precisely, using the ABCD method, the new waist location z_0' from the lens is z_0' = \frac{ f (d^2 + z_R^2 ) / d - f^2 }{ d - f + (d^2 + z_R^2 ) / d }, but for practical purposes, the minimal spot is as above for collimated input. The Rayleigh length of the input beam is z_R = \frac{\pi w_{\text{in}}^2}{\lambda}, which influences the propagation but the minimum waist size is achieved by ensuring collimation at the lens. In the special case of the input at the front (d = f), the output beam is nearly collimated with radius w_0' \approx \frac{[\lambda](/page/Lambda) f}{\pi w_{\text{in}}} located approximately at the (z_0' \approx 0), resulting in a slightly diverging output with small \theta' \approx \frac{w_{\text{in}}}{f}. This configuration is used to produce collimated beams with controlled size. For instance, the focal shift in collimated focusing is small, on the order of z_R^2 / f. In laser systems, such as those employing beam telescopes for compression, a pair of lenses can first reduce the beam diameter to increase \theta, but for optimal focusing to minimal spots, expansion prior to the focusing lens is common to maximize w at the final optic, yielding w_0' as small as a few micrometers for visible wavelengths and short focal lengths. For example, a 532 nm beam expanded to w = 5 mm and focused with f = 50 mm achieves w_0' \approx 1.7 μm, enhancing intensity for applications like optical trapping.

Derivation from Wave Equation

Paraxial Approximation

The scalar governs the propagation of monochromatic electromagnetic waves in free space, given by \nabla^2 E + k^2 E = 0, where E is the amplitude and k = 2\pi / \lambda is the with \lambda the . This equation describes exact wave solutions but is challenging to solve for beams with finite transverse extent propagating primarily along the z-direction. To obtain tractable solutions for such beams, the paraxial approximation is employed, assuming the beam propagates nearly parallel to the z- with small transverse variations relative to the longitudinal . Under this approximation, the field is expressed as E(x, y, z) = u(x, y, z) \exp(ikz), where u(x, y, z) is a slowly varying function that captures the beam's transverse profile and gradual changes along z. Substituting this form into the yields \frac{\partial^2 u}{\partial z^2} + \nabla_\perp^2 u + 2ik \frac{\partial u}{\partial z} = 0, with \nabla_\perp^2 = \partial^2 / \partial x^2 + \partial^2 / \partial y^2. The paraxial approximation neglects the second \partial^2 u / \partial z^2 because the varies slowly in the direction compared to the rapid phase oscillation \exp(ikz), resulting in the paraxial : $2ik \frac{\partial u}{\partial z} + \nabla_\perp^2 u = 0. This equation resembles the time-independent Schrödinger equation for a free particle in two dimensions, facilitating analytical solutions like Gaussian beams. The approximation holds when the beam's divergence angle \theta satisfies \theta \ll 1 radian, ensuring rays are nearly paraxial, and the beam waist width w_0 is much larger than the wavelength, w_0 \gg \lambda, so transverse derivatives dominate over longitudinal ones. These conditions are typical for many laser beam applications but break down for tightly focused beams or high numerical aperture (NA) systems, where full vectorial treatments or exact Helmholtz solutions are required to account for significant non-paraxial effects.

Hermite-Gaussian Solutions

The Hermite-Gaussian (HG) modes represent exact solutions to the paraxial for beams exhibiting rectangular symmetry in the , derived through in Cartesian coordinates. These modes form a complete orthogonal set, enabling the decomposition of arbitrary transverse field distributions into HG basis functions for analysis in laser resonators and free-space . The paraxial , which assumes small angles, underpins this separation, leading to a one-dimensional harmonic oscillator-like equation in each transverse dimension whose solutions involve . For the fundamental mode with indices m = 0 and n = 0, the transverse field profile reduces to the familiar Gaussian form E_{00}(x, y, z) \propto \exp\left( -\frac{x^2 + y^2}{w(z)^2} \right), multiplied by the w(z) and phase factors that account for propagation. Higher-order HG modes incorporate to introduce nodal structures: the electric field is given by E_{mn}(x, y, z) \propto H_m\left( \sqrt{2} \frac{x}{w(z)} \right) H_n\left( \sqrt{2} \frac{y}{w(z)} \right) \exp\left( -\frac{x^2 + y^2}{w(z)^2} \right) \times \exp\left[ i \left( kz - (m + n + 1) \tan^{-1}\left( \frac{z}{z_R} \right) + \frac{k (x^2 + y^2)}{2 R(z)} \right) \right], where H_m and H_n are of order m and n, k is the , z_R is the range, w(z) is the at z, and R(z) is the . The exponential Gaussian envelope ensures confinement, while the generate intensity lobes separated by nodes. Normalization of HG modes ensures unit power per mode in the paraxial approximation, with the normalization constant \sqrt{\frac{2}{\pi w(z)^2}} \frac{1}{\sqrt{2^{m+n} m! n!}} for the transverse part, such that the integral of |E_{mn}|^2 over the xy-plane yields the total power. Orthogonality follows from the properties of Hermite polynomials: \iint_{-\infty}^{\infty} E_{mn}(x, y, z) E_{m'n'}^*(x, y, z) \, dx \, dy = \delta_{mm'} \delta_{nn'}, allowing unique decomposition of input fields into mode coefficients for propagation calculations. The Gouy phase shift for the HG_{mn} mode accumulates as -(m + n + 1) \tan^{-1}(z / z_R) along the propagation direction, reflecting the topological contribution from the m + n transverse nodes plus the fundamental beam's shift. For example, the TEM_{10} mode (m=1, n=0) features a single intensity null along the x-direction, with the field profile E_{10} \propto \frac{\sqrt{2} x}{w(z)} \exp\left( -\frac{x^2 + y^2}{w(z)^2} \right) times phase terms, resulting in an odd lobe structure symmetric about the node.

Higher-Order Modes

Hermite-Gaussian Modes

Hermite-Gaussian () modes exhibit a distinctive transverse structure defined by orthogonal nodal lines aligned parallel to the coordinate axes, reflecting their Cartesian symmetry. The mode denoted as HG_{mn} possesses m nodal lines parallel to the y-axis (in the x-direction) and n nodal lines parallel to the x-axis (in the y-direction), creating a pattern of intensity lobes separated by these zero-intensity lines. This structure arises from the product form of the mode's transverse field, consisting of multiplied by a , as referenced in the derivation of Hermite-Gaussian solutions. The effective spot size w_{\rm eff} for higher-order modes exceeds that of the fundamental HG_{00} mode and scales approximately as w \sqrt{m + n + 1}, where w is the spot size parameter of the fundamental Gaussian, accounting for the expanded spatial extent due to the additional lobes. These modes can be generated in laser resonators designed with rectangular geometry, such as unstable resonators, where the confinement favors the in Cartesian coordinates and supports HG profiles as eigenmodes. Alternatively, higher-order HG modes are produced by modifying a fundamental using or masks, such as slits or spatial modulators, which selectively excite specific Hermite polynomial components through or phase patterning. The beam quality of HG_{mn} modes is quantified by the factor M^2 = \sqrt{(2m + 1)(2n + 1)}, which measures deviation from Gaussian propagation and increases with mode order due to the multimode nature and larger . For the HG_{00} mode, M^2 = 1, while higher orders exhibit reduced focusability, with M^2 = 3 for HG_{11}, for instance. In practical applications, such as slab lasers, HG modes are employed to mitigate arising from the anisotropic medium geometry; the rectangular slab cross-section naturally selects one-dimensional HG_{0n} modes, aligning the beam's asymmetry with the to achieve higher power extraction and improved output quality.

Laguerre-Gaussian Modes and Orbital Angular Momentum

Laguerre-Gaussian (LG) modes represent a complete set of orthogonal solutions to the in cylindrical coordinates, suitable for describing stable resonator modes in circularly symmetric cavities. These modes are indexed by two non-negative integers: the radial index p = 0, 1, 2, \dots, which governs the number of radial nodes, and the azimuthal index l = 0, \pm 1, \pm 2, \dots, which specifies the topological charge associated with the phase structure. Unlike fundamental Gaussian beams (p = 0, l = 0), higher-order LG modes exhibit doughnut-shaped intensity profiles with a central dark spot for |l| > 0, where the intensity vanishes on the beam axis due to the phase singularity. The transverse electric field distribution of an LG mode is given by u_{p}^{l}(r, \phi, z) = \sqrt{\frac{2p!}{\pi (p + |l|)!}} \frac{1}{w(z)} \left( \frac{\sqrt{2} r}{w(z)} \right)^{|l|} L_{p}^{|l|} \left( \frac{2 r^{2}}{w^{2}(z)} \right) \exp\left( -\frac{r^{2}}{w^{2}(z)} \right) \exp\left( i l \phi \right) \exp\left( -i k \frac{z + i z_{R}}{2 q(z)} r^{2} \right) \exp\left( i (2p + |l| + 1) \zeta(z) \right), where L_{p}^{|l|} are the associated , w(z) is the beam radius, q(z) is the complex beam parameter, z_R is the Rayleigh range, \zeta(z) = \tan^{-1}(z/z_R) is the Gouy phase, and k = 2\pi/\lambda is the . This form incorporates a Gaussian modulated by the Laguerre polynomial and an azimuthal \exp(i l \phi), leading to a helical that twists around the l times. A defining feature of LG modes with l \neq 0 is their carrying of orbital (OAM), distinct from the angular momentum associated with . The azimuthal phase dependence \exp(i l \phi) imparts a well-defined OAM of l \hbar per along the beam axis, where \hbar is the reduced Planck's constant. This OAM arises from the conservation of angular momentum in the and manifests as a when the beam interacts with absorbing or birefringent particles, enabling rotational manipulation in . The total angular momentum of light in such beams combines OAM and contributions, but LG modes provide a pure OAM basis for in quantum and classical . Seminal theoretical work demonstrated that LG modes possess this quantized OAM and can be interconverted with Hermite-Gaussian modes via astigmatic transformations, such as a pair of cylindrical lenses. LG modes form a basis for expanding arbitrary paraxial fields in cylindrical symmetry, with ensuring no cross-talk in mode decomposition. For p = 0, the modes reduce to simple vortex beams with a single intensity ring, while increasing p adds concentric rings, enhancing complexity for applications like . The OAM spectrum allows encoding information in the l degree of freedom, supporting high-capacity optical communications with between different l values over propagation distances. Experimental generation often involves spatial modulators imprinting the onto a Gaussian beam or intra-cavity mode selection in stable resonators.

Ince-Gaussian and Other Modes

Ince-Gaussian () modes represent a family of exact, orthogonal solutions to the paraxial in elliptic coordinates, forming the third complete set of such modes alongside Hermite-Gaussian and Laguerre-Gaussian families. These modes are characterized by their transverse intensity profiles, which are described by Ince polynomials multiplied by a Gaussian envelope, exhibiting inherent elliptical symmetry that allows for a continuous transition between rectangular (Hermite-Gaussian) and cylindrical (Laguerre-Gaussian) symmetries as the ellipticity parameter varies. IG modes are particularly stable in elliptic resonators, where the cavity mirrors' geometry matches the beam's elliptical structure, enabling efficient lasing without mode discrimination in astigmatic systems. Unlike the more commonly used Hermite-Gaussian and Laguerre-Gaussian modes, IG modes are rarer in practical applications due to the prevalence of circular or square designs, but they offer advantages in astigmatic cavities by minimizing distortion and improving mode purity. For instance, in elliptic propagation, IG modes maintain their shape over distance, making them suitable for applications requiring elliptical distributions, such as tailored optical or beam shaping in anisotropic media. Hypergeometric-Gaussian (HyGG) modes constitute another class of paraxial solutions, analogous to Mathieu-like beams in their use of for bounded propagation, where the field amplitude is proportional to confluent hypergeometric functions combined with a Gaussian factor. These modes feature a singular profile at the center, rendering them eigenfunctions of the orbital , with intensity patterns typically showing a single bright ring surrounding a dark core. HyGG modes are overcomplete and nonorthogonal, providing flexibility in representing complex profiles for applications like vortex beam generation, though their propagation is confined to paraxial regimes similar to other Gaussian-derived modes. In general, both and HyGG modes, along with related families like Mathieu-Gaussian modes, emerge as exact solutions to the paraxial under symmetries appropriate to elliptic or polar coordinates, offering alternatives to standard modes for systems with non-circular geometries. While and modes dominate due to their simplicity in cylindrical and Cartesian setups, modes excel in handling , and HyGG modes support advanced control, though experimental generation often requires specialized techniques.

Applications

In Laser Systems and Beam Quality

In single-mode lasers, the fundamental transverse electromagnetic mode, denoted as TEM_{00}, corresponds to a Gaussian beam profile that achieves the highest possible brightness by concentrating the output power into the smallest possible diffraction-limited spot. This mode is preferred in applications requiring maximal intensity, such as precision spectroscopy and high-resolution material processing, because it minimizes and maximizes focusability compared to higher-order modes. The beam quality of output is quantitatively assessed using the M^2 factor, also known as the beam propagation factor, which equals 1 for an Gaussian beam and quantifies deviations from this diffraction-limited performance in real beams. A value of M^2 = 1 indicates perfect Gaussian , where the product of the beam waist size and far-field divergence angle matches the theoretical minimum; higher values signify increased divergence or larger waist sizes, degrading focusability. Importantly, M^2 remains invariant under free-space or through ideal paraxial optical systems, allowing consistent characterization of beam independent of position. Gaussian beams play a central role in the design of stable laser resonators, where the ABCD matrix formalism describes mode propagation between cavity mirrors, ensuring self-consistent Gaussian solutions only in stable configurations. Resonator stability is determined by the parameters g_1 = 1 - L/R_1 and g_2 = 1 - L/R_2, where L is the cavity length and R_1, R_2 are the mirror radii of curvature; the condition $0 < g_1 g_2 < 1 defines the stable regime, confining the Gaussian mode within the cavity without loss to diffraction or walk-off. This criterion, derived from ray-transfer matrix analysis, guides the selection of mirror curvatures to support low-loss TEM_{00} operation in diverse laser systems. A practical example is found in CO_2 lasers used for industrial cutting, where the output is optimized for a Gaussian TEM_{00} profile to achieve the tightest focal spot and highest on the workpiece, enabling efficient material removal with minimal heat-affected zones. In these systems, the Gaussian allows cutting speeds proportional to power up to several kilowatts, as demonstrated in mild processing, while deviations to multimode operation would broaden the beam and reduce precision.

In Optical Trapping and Imaging

Gaussian beams form the basis of , a technique pioneered by in 1970 through the demonstration of particle acceleration and trapping using from a continuous beam. In 1986, Ashkin advanced this to single-beam gradient force traps, where a tightly focused Gaussian beam generates an intensity gradient that confines micron-sized dielectric particles in three dimensions via the mismatch-induced force. The Gaussian profile's smooth intensity distribution and diffraction-limited focal spot enable stable trapping of objects ranging from viruses to cells, with forces on the order of piconewtons. This invention earned Ashkin the for pioneering optical manipulation of microscopic objects. Laguerre-Gaussian (LG) beams, carrying orbital angular momentum (OAM), enhance optical trapping by transferring rotational momentum to particles. In LG-based , partially absorbing particles trapped in the beam's helical front experience azimuthal forces, leading to controlled with rates proportional to the topological charge of the . This OAM transfer, first experimentally verified in , allows bidirectional spinning and has applications in studying microrheology and assembling chiral structures. In confocal microscopy, Gaussian beams provide excitation light focused to a diffraction-limited spot, achieving lateral resolutions near \lambda / (2 \mathrm{NA}), where \lambda is the and NA is the . The pinhole further improves contrast by blocking out-of-focus , while the beam's Rayleigh range z_R = \pi w_0^2 / \lambda (with w_0 the beam waist) defines the axial sectioning capability, typically on the order of 0.5–1 \mum for high-NA objectives. Astigmatic Gaussian beams facilitate three-dimensional particle tracking in by introducing controlled aberration via a , which elongates the point spread function () elliptically along one axis depending on the defocus position. The degree of ellipticity directly correlates with the axial displacement from the focal plane, enabling nanometer-precision z-localization over ranges of several micrometers without mechanical scanning. This method is particularly effective for real-time monitoring of in complex media. Bessel-Gaussian beams approximate non-diffracting by superimposing a Gaussian on a core, allowing extended axial trapping along the beam path without rapid . First applied in optical traps in , these beams enable multiple-particle and over distances up to millimeters, ideal for holographic line traps in biological assays.

References

  1. [1]
    Gaussian beams - RP Photonics
    Definition: light beams where the electric field profile in a plane perpendicular to the beam axis can be described with a Gaussian function.Missing: scholarly | Show results with:scholarly
  2. [2]
    Gaussian Beams - SPIE
    Gaussian beams are laser beams where the electric field distribution is described by a Gaussian function. They stay Gaussian during propagation in a homogenous ...Missing: sources | Show results with:sources
  3. [3]
    Gaussian Profile Beam - an overview | ScienceDirect Topics
    A Gaussian beam is a bell-shaped distribution of photons, with the highest concentration at the center, tapering off towards the edges.
  4. [4]
    Gaussian beam and pulsed-beam dynamics: complex-source and ...
    Abstract. Paraxial Gaussian beams (GB's) are collimated wave objects that have found wide application in optical system analysis and design.
  5. [5]
    Gaussian beams - Book chapter - IOPscience
    We discuss key properties of Gaussian beams such as beam waist, divergence, depth of focus, and beam quality. We also introduce parameters like the Rayleigh ...<|control11|><|separator|>
  6. [6]
    (PDF) Laser: Gaussian Beam Profile - ResearchGate
    Mar 14, 2024 · Laser: Gaussian Beam Profile in its fundamental mode (TEM)00 is defined with parameters that describe it completely: Beam Waist, ...
  7. [7]
    Lasers - A. E. Siegman - Google Books
    Oct 17, 1986 · Contents ; Wave Optics and Gaussian Beams. 626 ; Physical Properties of Gaussian Beams. 663 ; Beam Perturbation and Diffraction. 698.
  8. [8]
    Ray tracing Gaussian beam optical systems - SPIE Digital Library
    Feb 20, 2024 · This article introduces a streamlined method for ray-tracing Gaussian laser beams in optical systems, drawing from traditional matrix optics ...
  9. [9]
    [PDF] Laser and Gaussian Beam Propagation and Transformation
    When the beam is totally coherent, it can be successfully described by extending the definitions of the Gaussian beam case by means of the ... Gaussian beam ...Missing: scholarly | Show results with:scholarly
  10. [10]
    [PDF] Gaussian Beam Optics - Experimentation Lab
    Siegman. Lasers (Sausalito, CA: University Science Books, 1986). S. A. Self. “Focusing of Spherical Gaussian Beams.” Appl. Opt.
  11. [11]
    Gaussian Beam Optics - Newport
    The Gaussian is a radially symmetrical distribution whose electric field variation is given by the following equation.<|control11|><|separator|>
  12. [12]
  13. [13]
    Laser Beams and Resonators - Optica Publishing Group
    This paper is a review of the theory of laser beams and resonators. Emphasis is placed on formulations and derivations which lead to basic understanding.
  14. [14]
    Paraxial wave optics and harmonic oscillators | Phys. Rev. A
    Jul 1, 1993 · The operator algebra of the quantum harmonic oscillator is applied to the description of Gaussian modes of a laser beam.
  15. [15]
    Analogies between optical and quantum mechanical angular ...
    Feb 28, 2017 · Paraxial beams and quantum harmonic oscillators. (a) Hermite–Gaussian modes. An orthonormal basis of exact normalized solutions of the scalar ...<|separator|>
  16. [16]
    Confocal Multimode Resonator for Millimeter Through Optical ...
    In this paper a resonator consisting of two identical concave spherical reflectors, separated by any distance up to twice their common radius of curvature, is ...
  17. [17]
  18. [18]
  19. [19]
  20. [20]
    Rayleigh length - RP Photonics
    The Rayleigh length (or Rayleigh range) of a laser beam is the distance from the beam waist (in the propagation direction) where the beam radius is increased ...Missing: seminal | Show results with:seminal
  21. [21]
    How to define and optimize axial resolution in light-sheet microscopy
    This measure yields values equivalent to the Rayleigh range for Gaussian beams (except the Rayleigh range is based on an increase by a factor of √2) but better ...
  22. [22]
    Beam Divergence – angle - RP Photonics
    For a diffraction-limited Gaussian beam, the 1 / e 2 beam divergence half-angle is λ / ( π w 0 ) , where λ is the wavelength (in the medium) and w 0 the beam ...Quantitative Definitions of... · Divergence of Gaussian... · Spatial Fourier Transforms
  23. [23]
    [PDF] Gaussian Beam Optics
    Coherent Gaussian beams have peculiar transformation properties which require special consideration. In order to select the best optics for a particular laser ...
  24. [24]
    M^2 Factor – M squared, laser beam, quality factor ... - RP Photonics
    M2 Factor of Hermite–Gaussian Beams​​ A diffraction-limited beam has an factor of 1, and is a Gaussian beam. Smaller values of are physically not possible. A ...
  25. [25]
  26. [26]
    Physical origin of the Gouy phase shift - Optica Publishing Group
    We show explicitly that the Gouy phase shift of any focused beam originates from the transverse spatial confinement, which, through the uncertainty principle, ...
  27. [27]
    Observation of the quantum Gouy phase | Nature Photonics
    Oct 6, 2022 · Here we outline a simple method for calculating the quantum state evolution upon propagation and demonstrate experimentally how this quantum Gouy phase affects ...
  28. [28]
  29. [29]
  30. [30]
    Beam quality of truncated laser beams with amplitude modulations ...
    The M2-factor of truncated laser beams with AMs and PFs is always larger than that of truncated Gaussian beams both in free space and in turbulence.
  31. [31]
    Unstable Resonators - RP Photonics
    There is a special class of unstable resonators where a soft optical aperture with a Gaussian transmission or reflection profile serves to limit the field ...
  32. [32]
    [PDF] Lecture 24 Gaussian Beam Optics
    is called the confocal parameter. Page 7. u. 00(x,y,z) '. 2 1 w e. &i exp & r. 2 ... Equivalent to defining a new origin and Gaussian beam parameter for the ...
  33. [33]
    [PDF] 2.6 Gaussian Beams and Resonators
    With the Gaussian beam solutions, we can finally construct optical resonators with finite transverse extent, i.e. real Fabry-Perots, by inserting into the.
  34. [34]
    [PDF] ⌋ ⌉ ⌊ ⌈ - = 1 2 0 1 R M - LASCAD
    Gaussian Paraxial Wave Optics. The ABCD matrix can also be applied to transform the so called q Parameter of a Gaussian beam. R radius of phase front ...
  35. [35]
    [PDF] Gaussian Beam Propagation in Misaligned Complex Optical Systems
    Jul 1, 1995 · It is the purpose of this section to derive from first principles Gaussian beam propagation in these lenslike media, which are allowed to be ...
  36. [36]
    ABCD Matrix – ray transfer matrix - RP Photonics
    An ABCD matrix [1] is a 2-by-2 matrix associated with an optical element which can be used for describing the element's effect on a laser beam.
  37. [37]
    Lens-System Diffraction Integral Written in Terms of Matrix Optics*
    This paper deals with the propagation of coherent light through lens systems; it relates diffraction theory to ray optics. A diffraction integral is derived ...
  38. [38]
    [PDF] 2.4 Paraxial Wave Equation and Gaussian Beams
    In this sec- tion, we want to study the properties of Gaussian beams and its propagation and modification in optical systems. 2.4.1 Paraxial Wave Equation. We ...<|control11|><|separator|>
  39. [39]
    Understanding the Paraxial Gaussian Beam Formula | COMSOL Blog
    Sep 21, 2016 · The paraxial Gaussian beam formula is an approximation to the Helmholtz equation derived from Maxwell's equations.
  40. [40]
    Hermite–Gaussian Modes – Gauss - RP Photonics
    Hermite–Gaussian modes are propagation modes or resonator modes which are described with Hermite–Gaussian functions.Missing: seminal | Show results with:seminal<|control11|><|separator|>
  41. [41]
    [PDF] Modeling lasers using Gaussian wave packets
    Jun 29, 2021 · Subscripts m and n indicate the number of transverse nodal lines in the x- and y-directions of the beam respectively. The subscript p ...
  42. [42]
    Spot size and divergence for Hermite Gaussian beams of any order
    - **Insufficient relevant content**: The provided text from https://opg.optica.org/ao/abstract.cfm?uri=ao-19-7-1027 does not include the full article or abstract, preventing extraction of the spot size formula for higher-order Hermite-Gaussian beams.
  43. [43]
    Generation of Higher-Order Hermite–Gaussian Modes via ... - MDPI
    The spatial light modulator has high spatial resolution and excellent maneuverability, so it is a simple and effective way to generate higher-order HG modes.
  44. [44]
    [PDF] Controlling Femtosecond Laser Filaments via Quasi-Hermite ...
    We have created phase masks on a spatial light modulator to produce desired beam profiles such as quasi-Hermite Gaussian beam modes.
  45. [45]
    High-Power, High-Purity HG0n Hermite–Gaussian Laser Beam ...
    Oct 8, 2023 · The one-dimensional, high-order Hermite–Gaussian (HG0n) laser beams can be easily converted into vortex Laguerre–Gaussian (LGpl) beams ...
  46. [46]
    Laguerre–Gaussian Modes - RP Photonics
    Laguerre–Gaussian (LG) modes are a family of paraxial laser-beam solutions that are naturally expressed in cylindrical coordinates.What are Laguerre–Gaussian... · Laguerre–Gaussian Mode...
  47. [47]
    Orbital angular momentum of light and the transformation of ...
    An astigmatic optical system may be used to transform a high-order Laguerre-Gaussian mode into a high-order Hermite-Gaussian mode reversibly. An experiment ...Missing: original | Show results with:original
  48. [48]
    (PDF) Ince–Gaussian beams - ResearchGate
    Aug 6, 2025 · We demonstrate the existence of the Ince-Gaussian beams that constitute the third complete family of exact and orthogonal solutions of the paraxial wave ...Missing: seminal | Show results with:seminal
  49. [49]
    [0712.0782] Hypergeometric-Gaussian Modes - arXiv
    Dec 5, 2007 · These modes have a singular phase profile and are eigenfunctions of the photon orbital angular momentum.
  50. [50]
  51. [51]
    Measurement of the M2 beam propagation factor using a focus ...
    An effective beam quality factor that is invariant under free propagation can be defined based on the determinant of the beam matrix, eff. Variable-distance ...
  52. [52]
    [PDF] Laser Beams and Resonators - Diels Research Group
    This paper is a review of the theory of laser beams and resonators. Emphasis is ... Boyd and Gordon. [5], and Boyd and Kogelnik [6] developed a theory for.Missing: original | Show results with:original
  53. [53]
    [PDF] Carbon dioxide laser cutting: Beam mode and oxygen purity effects ...
    Sep 25, 2024 · Laser cutting is predominately performed using a Gaussian (TEMoo) beam mode because the beam can be focused to a very small spot, which creates ...Missing: output | Show results with:output<|control11|><|separator|>
  54. [54]
    Acceleration and Trapping of Particles by Radiation Pressure
    Micron-sized particles have been accelerated and trapped in stable optical potential wells using only the force of radiation pressure from a continuous laser.Missing: beam | Show results with:beam
  55. [55]
    Observation of a single-beam gradient force optical trap for dielectric ...
    © 1986 Optical Society of America. We report the first experimental observation to our knowledge of a single-beam gradient force radiation-pressure particle ...
  56. [56]
    Optical Trapping and Manipulation of Viruses and Bacteria - Science
    Optical trapping and manipulation of viruses and bacteria by laser radiation pressure were demonstrated with single-beam gradient traps.Missing: Arthur | Show results with:Arthur<|control11|><|separator|>
  57. [57]
  58. [58]
  59. [59]