Fact-checked by Grok 2 weeks ago

Resonance

Resonance is a fundamental physical phenomenon in which a system, such as an oscillating object or circuit, exhibits a significantly amplified response or vibration when driven by an external force at a frequency that matches its natural frequency of oscillation. This natural frequency is determined by the system's intrinsic physical properties, such as mass, stiffness, or inductance and capacitance in electrical contexts. The effect arises because energy transfer from the driving force is most efficient at this resonance frequency, leading to maximal amplitude buildup, though damping mechanisms like friction typically limit the response to prevent unbounded growth. In mechanical and acoustic systems, resonance manifests when external vibrations couple with the system's modes, producing effects ranging from constructive reinforcement in musical instruments to destructive amplification in susceptible structures. For instance, a playground swing achieves greater height when pushed periodically at its natural period, illustrating how small inputs can yield large outputs near resonance. In , resonance occurs in LRC circuits where the inductive reactance equals capacitive reactance, enabling efficient energy storage and applications in radio tuning and . The sharpness of the resonance peak is quantified by the quality factor Q, defined as Q = ω₀ / γ, where ω₀ is the natural and γ is the coefficient; high Q values indicate narrow, intense resonances useful in oscillators. Beyond and , resonance extends to and , where it describes transient states or "resonant particles" with definite energies, detectable through sharp peaks in cross-sections, as in reactions or high-energy collisions. In and , controlled resonance underpins technologies like MRI scanners, which exploit to image tissues, and seismic dampers that mitigate building vibrations during earthquakes. Overall, resonance underscores the interplay between driving forces and across disciplines, enabling both innovative applications and caution against unintended amplifications.

Fundamentals

Definition and Basic Principles

Resonance is a fundamental phenomenon in physics where a 's of is significantly amplified when subjected to a periodic driving at or near its , resulting in a maximum response. This occurs in various , from mechanical structures to electrical circuits, where the is the rate at which the would oscillate freely if displaced from . It was formalized by in 1665, who observed that two pendulum clocks suspended from the same beam would synchronize their swings due to mutual , an early recognition of resonant . A simple illustrates this: consider pushing a child on a playground swing. By applying gentle pushes timed precisely with the swing's natural back-and-forth rhythm, the height of the swing increases dramatically without requiring additional force, as each push adds energy constructively. For those unfamiliar with oscillators, these are systems—like a on a or a swinging —that naturally vibrate at a characteristic frequency set by their inherent properties, such as and restoring force. The basic condition for resonance is given by the equation \omega_d = \omega_0 where \omega_d is the of the driving force and \omega_0 is the system's natural . At this match, energy transfer from the driver to the system is maximized per cycle, allowing small inputs to accumulate into large oscillations, as the system's motion aligns perfectly with the applied force.

Harmonic Motion and Natural Frequency

Simple harmonic motion (SHM) describes the oscillatory behavior of a system where the restoring force is directly proportional to the displacement from equilibrium and acts in the opposite direction. This relationship is expressed by Hooke's law, F = -kx, where k is the spring constant and x is the displacement. Applying Newton's second law, F = ma, yields the differential equation m \frac{d^2x}{dt^2} + kx = 0, which has the general solution x(t) = A \cos(\omega_0 t + \phi), where A is the amplitude, \omega_0 is the natural angular frequency, and \phi is the phase constant determined by initial conditions. The natural frequency arises from the system's inherent properties and represents its rate in the absence of external influences. For a mass-spring system, substituting the restoring force into Newton's second gives \frac{d^2x}{dt^2} = -\frac{k}{m} x, leading to the natural \omega_0 = \sqrt{\frac{k}{m}}, where m is the . This frequency depends solely on the stiffness k and m, illustrating how softer springs or heavier masses result in slower oscillations. For a simple pendulum, under the where \sin \theta \approx \theta, the equation simplifies to \frac{d^2 \theta}{dt^2} + \frac{g}{l} \theta = 0, yielding \omega_0 \approx \sqrt{\frac{g}{l}}, with g as and l as the pendulum length. In undriven SHM, mechanical energy is conserved, with the total energy remaining constant as it interchanges between kinetic and potential forms. The potential energy is U = \frac{1}{2} k x^2, and the kinetic energy is K = \frac{1}{2} m v^2, so the total energy E = K + U = \frac{1}{2} k A^2 at maximum displacement, where velocity is zero. This conservation implies that the amplitude A is fixed, and the motion persists indefinitely without energy loss. The period T_0 of SHM, the time for one complete , is given by T_0 = \frac{2\pi}{\omega_0}, independent of for ideal systems. The natural frequency \omega_0 thus defines the system's intrinsic rhythm, crucial for understanding how it responds to perturbations. The \phi shifts the , allowing alignment with initial and ; for instance, \phi = 0 starts at maximum . Graphically, SHM exhibits sinusoidal patterns across key variables. x(t) varies as a cosine wave between -A and A. v(t) = -A \omega_0 \sin(\omega_0 t + \phi) leads by \pi/2 radians, reaching maxima at . Acceleration a(t) = -A \omega_0^2 \cos(\omega_0 t + \phi) is out of phase with , peaking at extremes and zero at , confirming a = -\omega_0^2 x. These plots highlight the synchronized, periodic nature of the motion.

Linear Systems

Driven Damped Harmonic Oscillator

The driven damped models a where an external sinusoidal force excites a damped mass-spring setup, leading to a steady-state response that exhibits resonance when the driving frequency approaches the . The equation of motion is derived from Newton's second law, incorporating inertial, , restoring, and driving forces: m \ddot{y} + b \dot{y} + k y = F_0 \cos(\omega_d t), where m is the mass, b the damping coefficient, k the spring constant, F_0 the driving force amplitude, and \omega_d the driving angular frequency. The natural angular frequency is \omega_0 = \sqrt{k/m}, and the damping ratio is \beta = b/(2m). After transients decay, the steady-state solution is y(t) = A(\omega_d) \cos(\omega_d t - \phi), with amplitude A(\omega_d) = \frac{F_0 / m}{\sqrt{(\omega_0^2 - \omega_d^2)^2 + (2 \beta \omega_d)^2}}. This amplitude peaks at the resonance frequency \omega_r = \omega_0 \sqrt{1 - 2 \beta^2} for \beta < \omega_0 / \sqrt{2}, which shifts below \omega_0 due to damping but approximates \omega_0 for light damping (\beta \ll \omega_0). The phase lag is \phi = \arctan\left( \frac{2 \beta \omega_d}{\omega_0^2 - \omega_d^2} \right), starting at 0 for \omega_d \ll \omega_0 (in phase with the force), reaching \pi/2 near resonance, and approaching \pi for \omega_d \gg \omega_0 (out of phase). The average power delivered by the driving force, \bar{P} = \frac{1}{2} F_0 \omega_d A(\omega_d) \sin \phi, maximizes exactly at \omega_d = \omega_0, independent of damping, as this aligns the force with the velocity for optimal energy transfer. In frequency response plots, the amplitude curve shows a sharp peak near \omega_0 for low \beta, broadening with increased damping; the phase curve transitions smoothly from 0 to \pi; and the power curve peaks precisely at \omega_0 with a Lorentzian shape, dropping to half-maximum at \omega_0 \pm \beta. A practical example is a playground swing, modeled as a driven pendulum where periodic pushes apply the sinusoidal force; resonance occurs when pushes match the swing's natural frequency, building large amplitudes with minimal effort despite air damping, but mistimed pushes reduce the response.

RLC Circuits

In a series RLC circuit, consisting of a resistor R, inductor L, and capacitor C connected in series with a voltage source V(t), the governing equation arises from Kirchhoff's voltage law, balancing the voltage drops across each component: V(t) = I(t) R + L \frac{dI(t)}{dt} + \frac{1}{C} \int I(t) \, dt. This differential equation describes the circuit's response to an applied voltage, analogous in form to the equation for a driven damped mechanical oscillator. For a sinusoidal driving voltage V(t) = V_0 \cos(\omega t), the steady-state current I(t) is also sinusoidal at the same frequency \omega, with amplitude determined by the circuit's impedance Z: Z = R + j\left(\omega L - \frac{1}{\omega C}\right), where j is the imaginary unit. The magnitude of the impedance is |Z| = \sqrt{R^2 + \left(\omega L - \frac{1}{\omega C}\right)^2}, and the current amplitude is I = V_0 / |Z|. Resonance occurs when the imaginary part of Z vanishes, i.e., \omega L = 1/(\omega C), yielding the resonant angular frequency \omega_0 = 1 / \sqrt{LC}. At this frequency, the impedance minimizes to |Z| = R, maximizing the current amplitude to I_{\max} = V_0 / R. The voltages across the individual components reflect the phase shifts inherent in reactive elements. The resistor voltage V_R = I R remains in phase with the current and thus with the source voltage at resonance. The inductor voltage V_L = I \omega L leads the current by 90°, peaking above \omega_0 as its magnitude increases with frequency. Conversely, the capacitor voltage V_C = I / (\omega C) lags the current by 90°, reaching its maximum below \omega_0 due to the inverse frequency dependence. At resonance, V_L = V_C, and these reactive voltages cancel in the phasor diagram, leaving the total source voltage V = V_R aligned with the current phasor along the real axis. In general, the phasor sum satisfies V = \sqrt{V_R^2 + (V_L - V_C)^2}, illustrating how the net voltage vector results from the in-phase V_R and the opposing V_L and V_C components on the imaginary axis. The frequency response of the circuit, plotting current or voltage amplitudes versus \omega, shows a peak at \omega_0 for the series current, with the resistor voltage mirroring this shape. The capacitor voltage curve peaks at a frequency below \omega_0, while the inductor voltage peaks above it, both exhibiting broader resonances due to the fixed current amplitude at \omega_0. The bandwidth \Delta \omega, defined as the full width at half-maximum power (or current amplitude at $1/\sqrt{2} of peak), is \Delta \omega = R / L. The quality factor Q, measuring the sharpness of the resonance, is given by Q = \omega_0 L / R = \omega_0 / \Delta \omega, indicating how selectively the circuit responds near \omega_0. Antiresonance, the condition of minimum current, occurs in parallel RLC configurations at \omega = 1 / \sqrt{LC}, where the impedance maximizes, paralleling antiresonant behavior in mechanical systems.

Wave Phenomena

Standing Waves

Standing waves represent a resonant phenomenon in wave mechanics, arising in bounded media where waves interfere constructively at specific discrete frequencies. These patterns form through the superposition of an incident wave and its reflection from boundaries, resulting in fixed positions of zero displacement known as nodes and positions of maximum displacement called antinodes. Unlike traveling waves, standing waves exhibit no net propagation of energy across the medium, as the forward and backward wave components cancel each other's energy flux, though local energy oscillates between kinetic and potential forms. The boundary conditions imposed by the medium's constraints dictate the allowed wavelengths and frequencies for resonance. For a one-dimensional medium fixed at both ends, such as a string of length L, the wavelengths satisfy \lambda_n = \frac{2L}{n}, where n = 1, 2, 3, \dots is a positive integer, ensuring nodes at the boundaries. This quantization leads to resonant frequencies given by f_n = \frac{n v}{2 L}, where v is the wave speed in the medium; energy accumulates preferentially at these modes when driven externally. The general behavior of such waves is governed by the one-dimensional wave equation, \frac{\partial^2 u}{\partial t^2} = v^2 \frac{\partial^2 u}{\partial x^2}, whose standing wave solutions take the form u(x,t) = [A \sin(k x) + B \cos(k x)] \sin(\omega t + \phi), with k = \frac{2\pi}{\lambda} and \omega = 2\pi f, where coefficients A and B are determined by boundary conditions. In resonant conditions, when an external driving force matches one of these natural frequencies f_n, the amplitude at the antinodes grows significantly due to constructive reinforcement over time, analogous to frequency matching in harmonic oscillator resonance. This buildup distinguishes standing wave resonance from non-resonant cases, where destructive interference limits amplitude, highlighting the role of spatial patterning in energy localization within the bounded system.

Resonance in Strings and Pipes

Resonance in strings occurs through the formation of standing waves, where the string vibrates at specific natural frequencies determined by its length L, tension T, and linear mass density \mu. The fundamental frequency is given by f_1 = \frac{1}{2L} \sqrt{\frac{T}{\mu}}, representing the lowest mode with one antinode in the middle. Higher harmonics follow as integer multiples, f_n = n f_1 for n = 1, 2, 3, \dots, allowing the string to support multiple antinodes. These modes are excited by plucking, which typically emphasizes the fundamental and lower harmonics, or bowing, which sustains higher modes through continuous friction. In air columns, such as pipes, resonance arises from longitudinal standing waves, with frequencies depending on the speed of sound v and pipe length L. For a closed pipe (one end closed), the fundamental frequency is f_1 = \frac{v}{4L}, with odd harmonics f_n = (2n-1) f_1 for n = 1, 3, 5, \dots. An open pipe (both ends open) has f_1 = \frac{v}{2L}, with all integer harmonics f_n = n f_1. Real pipes require end corrections, adding an effective length \Delta L \approx 0.6 r (where r is the radius) to the open end(s) to account for the antinode displacement outside the pipe. Practical examples include guitar strings, which produce transverse standing waves resonating at harmonics to generate musical notes, and organ pipes, which create longitudinal acoustic waves for sustained tones. When two nearby resonators, such as slightly detuned strings or pipes, vibrate together, they produce beats—a periodic amplitude variation at the difference frequency, audible as a pulsating sound. Damping in these systems causes amplitude decay over time due to energy loss from friction in strings or viscosity and thermal conduction in air columns, leading to exponential decrease in vibration intensity. The quality factor Q, defined as Q = 2\pi \times \frac{\text{stored energy}}{\text{energy lost per cycle}}, quantifies the mode lifetime, with higher Q indicating longer resonance duration before significant decay. Experimental observation of string resonance is demonstrated in Melde's experiment, where a tuning fork drives a string either transversely (frequency matching) or longitudinally (half-frequency), verifying the harmonic frequencies by counting loops under varying tension. As a two-dimensional analog, Chladni patterns on vibrating plates reveal nodal lines of standing waves, illustrating resonance modes in extended media.

Advanced Concepts

Resonance in Complex Networks

Resonance in complex networks extends the principles of synchronization observed in simple coupled oscillators to interconnected systems with intricate topologies, where collective behaviors emerge from interactions among multiple components. A classic example is the synchronization of coupled pendulums, as first observed by in 1665, where two clocks suspended from a common beam gradually aligned their swings due to weak mechanical coupling through the support structure, leading to anti-phase or in-phase locking at a common frequency. This phenomenon builds on the resonance in individual oscillators by demonstrating how mutual influence can entrain disparate natural frequencies toward a shared rhythm, a process amplified when the driving or coupling frequency matches the system's inherent modes. In more general settings, the Kuramoto model provides a foundational mathematical description of phase locking in large ensembles of coupled oscillators, where each oscillator's phase evolves according to its natural frequency plus sinusoidal interactions from neighbors, resulting in synchronization when coupling strength exceeds a critical threshold determined by frequency heterogeneity. For networks, resonance manifests through eigenmodes of the graph Laplacian, which capture the system's natural vibrational patterns; driving the network at frequencies aligning with these eigenmodes enhances coherent responses, such as amplified signal propagation or collective oscillations, unlike the single natural frequency in isolated linear systems. Practical examples illustrate this network-scale resonance. In power grids, modeled as Kuramoto-like oscillators representing generators, synchronization maintains a nominal 50 or 60 Hz frequency to prevent blackouts; mismatches in driving frequencies can trigger resonant instabilities, but proper coupling ensures phase locking across the topology. Similarly, in biological networks like neural circuits, Kuramoto dynamics describe synchronized firing patterns, where resonance facilitates information processing, such as in gamma oscillations linking distant brain regions for cognitive tasks. Nonlinear effects introduce richer dynamics in strongly coupled networks, including subharmonic resonances where the system oscillates at fractions of the driving frequency, potentially leading to bifurcations into chaotic states or stable resonant clusters. These transitions arise from nonlinear interactions amplifying small perturbations, contrasting with linear cases by enabling multistable resonant regimes. The underlying framework relies on eigenvalues of the adjacency matrix (or Laplacian) to predict resonant frequencies, though complex topologies preclude simple closed-form solutions, requiring numerical spectral analysis for precise characterization.

Q Factor and Bandwidth

The quality factor, denoted Q, quantifies the sharpness of resonance in oscillatory systems by measuring how selectively the system responds near its natural frequency. It is defined as the ratio of the resonant angular frequency \omega_0 to the full width at half maximum (FWHM) \Delta \omega of the power response curve:
Q = \frac{\omega_0}{\Delta \omega}.
This definition highlights the inverse relationship between Q and the resonance bandwidth, where higher Q values indicate narrower peaks and greater frequency selectivity.
From an energy perspective, Q represents the efficiency of energy storage relative to dissipation, expressed as Q = 2\pi times the ratio of the peak energy stored in the resonator to the energy lost per oscillation cycle. In the context of a damped harmonic oscillator governed by \ddot{x} + 2\beta \dot{x} + \omega_0^2 x = 0, this yields Q = \omega_0 / (2 \beta), where \beta is the damping coefficient that governs the rate of amplitude decay. Low damping (small \beta) results in high Q, allowing sustained oscillations with minimal energy loss per cycle. The bandwidth follows directly as \Delta \omega = \omega_0 / Q, emphasizing that high-Q systems exhibit narrow resonances suited for precise frequency discrimination, while low-Q systems have broader responses indicative of higher damping. In applications such as bandpass filters, Q governs selectivity by determining how effectively the filter passes signals near \omega_0 while attenuating those outside the bandwidth, enabling designs with sharp cutoffs for signal processing in electronics. The amplitude response of a driven damped oscillator, A(\omega), derives from the steady-state solution and approximates a Lorentzian near resonance for low damping:
A(\omega) \approx \frac{F_0 / (2 m \beta \omega_r )}{\sqrt{1 + \left( \frac{\omega - \omega_r}{\beta} \right)^2}},
where F_0 is the driving force amplitude, m is mass, and \omega_r \approx \omega_0 is the frequency of maximum amplitude. The power response, proportional to A^2(\omega), then has an FWHM of $2\beta, aligning with the bandwidth definition and underscoring Q's role in peak sharpness.
Practically, Q is often measured via the ring-down time \tau, the characteristic decay time of free oscillations after excitation ceases, related by \tau = Q / \omega_0. This approach applies to mechanical systems, where vibrations decay slowly in high-Q structures, and electrical circuits, where charge oscillations in inductors and capacitors similarly reveal Q through transient response duration.

Applications Across Disciplines

Mechanical and Structural Resonance

Mechanical and structural resonance occurs when an external periodic force excites a mechanical system at or near its natural frequencies, leading to amplified vibrations that can cause excessive stress or failure in structures and machines. In engineering practice, is employed to identify these natural frequencies and mode shapes, often using to model complex structures and compute eigenfrequencies where resonance arises if the forcing frequency matches these values. This analysis helps predict and avoid conditions where small inputs produce large outputs, ensuring structural integrity under dynamic loads like wind or machinery operation. A prominent historical example is the collapse of the Tacoma Narrows Bridge in 1940, where wind-induced aeroelastic flutter excited torsional vibrations near the bridge's natural frequency of approximately 0.2 Hz, resulting in catastrophic failure after amplitudes reached up to 28 feet. The incident highlighted the dangers of resonance in slender structures, as the self-reinforcing oscillations from aerodynamic forces overwhelmed the bridge's limited damping capacity. In machinery, resonance is managed by tuning engine operating speeds to avoid natural frequencies of components, preventing amplified vibrations that could lead to wear or breakdown. Vibration isolation techniques, such as damped mounts made from rubber or elastomers, are commonly used to decouple machines from their supports, reducing transmission of resonant energy and maintaining operational stability. For space applications, the International Space Station (ISS) employs dampers on its solar arrays to control resonance induced by structural dynamics or thruster firings, mitigating fatigue in the lightweight panels that span over 100 meters. These passive damping systems, often integrated into the array masts, absorb vibrational energy to prevent mode amplification and ensure long-term durability in the microgravity environment. To counteract resonance in tall structures, tuned mass dampers (TMDs) are installed to absorb oscillatory energy; for instance, the 660-ton spherical TMD in Taipei 101 skyscraper is tuned to the building's fundamental frequency, reducing sway amplitudes by up to 40% during wind or seismic events. Design of such dampers incorporates the Q factor to optimize damping levels, balancing energy dissipation without overly restricting motion.

Acoustic and Optical Resonance

Acoustic resonance occurs in structures where sound waves interfere constructively at specific frequencies, leading to enhanced oscillations in enclosed or bounded spaces. A classic example is the , which consists of a rigid cavity connected to the exterior via a narrow neck, behaving like a mass-spring system for air. The resonant frequency is given by f = \frac{v}{2\pi} \sqrt{\frac{A}{V L}}, where v is the speed of sound, A is the neck's cross-sectional area, V is the cavity volume, and L is the effective neck length. This device is widely employed in automotive mufflers to target and attenuate low-frequency exhaust noise by absorbing energy at the design frequency. In rooms and other enclosed volumes, acoustic resonance manifests as room modes, which are standing wave patterns determined by the dimensions of the space. These modes result in peaks and nulls in the frequency response, causing uneven sound distribution, boomy bass buildup, and prolonged echoes at modal frequencies, particularly below 300 Hz. Whispering gallery modes provide another acoustic resonance mechanism, where sound waves propagate along curved boundaries via successive total internal reflections, achieving high quality factors in structures like cathedral domes or stadium galleries for efficient sound confinement near the surface. Optical resonance parallels acoustic phenomena but involves electromagnetic waves in cavities, enabling precise control of light. In a Fabry-Pérot interferometer, formed by two parallel partially reflecting mirrors separated by distance L, resonance occurs when the round-trip phase shift is a multiple of $2\pi, yielding angular frequencies \omega_m = m \pi c / L for integer mode number m, with c the speed of light. Such cavities are essential in lasers, where they sustain coherent light amplification by confining photons; the quality factor Q = 2\pi \nu \tau quantifies the sharpness of resonance, with \nu the frequency and \tau the photon lifetime inside the cavity. High-Q whispering gallery modes in optical microspheres trap light circumferentially through total internal reflection, yielding quality factors exceeding $10^9 for enhanced light-matter interactions and minimal losses. Practical applications include acoustic levitation, where standing waves in resonant cavities create pressure nodes to suspend small objects against gravity. In telecommunications, Fabry-Pérot cavities serve as tunable optical filters, selectively transmitting narrow wavelength bands for wavelength-division multiplexing. Nonlinear optical processes benefit from parametric resonance, as in second-harmonic generation, where efficiency surges when the pump laser frequency aligns with a cavity mode, enabling phase-matched frequency doubling within the resonator.

Electrical and Electronic Resonance

In electrical and electronic systems, resonance manifests through tuned circuits that selectively amplify or filter signals at specific frequencies. A prominent example is the parallel resonant LC circuit augmented with resistance (RLC), where at the resonant frequency, the inductive and capacitive reactances cancel, resulting in maximum circuit impedance dominated by the resistor. This high-impedance state at resonance allows parallel RLC circuits to function as effective bandpass filters in radio tuners, where varying the capacitance tunes the resonant frequency to select desired broadcast signals while rejecting others. Crystal oscillators leverage the piezoelectric properties of quartz crystals to achieve precise frequency control, operating at the crystal's mechanical resonance frequency typically in the range of 1 to 100 MHz. The quartz crystal vibrates mechanically when an electric field is applied, generating an electrical signal at its resonant frequency, which is fed back into an amplifier to sustain oscillation. These devices exhibit exceptionally high quality factors (Q > 10^4, often up to 10^6), enabling superior frequency stability essential for applications like electronic clocks and timekeeping systems. Wireless power transfer utilizes resonant coupling between coils to enable efficient energy transmission without physical connections. Tesla coils exemplify early coupled resonator systems, where primary and secondary coils tuned to the same resonant frequency exchange energy via oscillating , achieving non-radiative transfer over moderate distances. Modern implementations, such as magnetic resonance coupling, employ self-resonant coils operating in the strongly coupled regime to deliver power efficiently up to several times the coil diameter, powering devices like wireless chargers with minimal losses. Electronic filters exploit resonance to shape frequency responses, particularly in bandpass configurations that pass a narrow band around the center frequency \omega_0 while attenuating others. The quality factor Q = \frac{\omega_0}{\Delta \omega} defines the filter's selectivity, where \Delta \omega is the bandwidth between the -3 dB points; higher Q yields narrower bandwidths and sharper resonance peaks. Design choices include Butterworth filters, which provide a maximally flat passband response with no ripple for smooth signal processing, versus Chebyshev filters, which introduce controlled passband ripple (e.g., 0.5 dB) for steeper roll-off and narrower transition bands at the cost of transient ringing. Superconducting resonators achieve ultra-high Q factors (often exceeding $10^6) by eliminating ohmic losses through zero-resistance at cryogenic temperatures, enabling with minimal dissipation. These cavities, typically fabricated from materials like or , serve as critical components in superconducting for , where high Q extends coherence times beyond 1 ms and supports precise control of quantum states. By avoiding resistive heating and losses, such resonators enhance qubit fidelity in scalable quantum processors.

Atomic and Molecular Resonance

In , resonance occurs when matches the energy difference between discrete quantum levels, inducing transitions. For atoms, the exemplifies this, where electrons transition from higher energy levels (n > 2) to the n=2 level, emitting photons at visible wavelengths such as 656 nm for the Hα line. These resonant frequencies arise from the quantized energy levels described by the , refined by , enabling precise spectroscopic identification of atomic species. When atoms are driven by an external oscillating at the resonant between two levels, coherent leads to Rabi oscillations, where the probability of finding the atom in the oscillates sinusoidally with a frequency proportional to the field strength. This quantum effect, first theoretically described for transitions, has been extended to optical driving in two-level systems, allowing controlled manipulation of atomic states in experiments. At the molecular scale, resonance manifests in vibrational and rotational modes excited by radiation matching the spacings of these quantized motions. In (IR) spectroscopy, resonant absorption occurs when the aligns with fundamental vibrational transitions, such as or modes in diatomic or polyatomic molecules, providing fingerprints for molecular identification. Rotational resonances, typically in the range, couple with vibrations to form rovibrational spectra, revealing structural details through selection rules that favor ΔJ = ±1 changes in . Nuclear magnetic resonance (NMR) in molecules exploits the resonant precession of nuclear spins in a , occurring at the Larmor given by \omega = \gamma B, where \gamma is the and B is the applied . This resonance allows radiofrequency pulses to flip spins, enabling high-resolution mapping of molecular environments in liquids and solids. In , describes charged particles orbiting in magnetic fields at the cyclotron \omega_c = \frac{q B}{m}, where q is charge, B is , and m is ; when an applied electromagnetic wave matches \omega_c, absorption enhances particle acceleration, as observed in diagnostics and solid-state cyclotron resonance experiments. Resonances also appear in as peaks in cross-sections, corresponding to short-lived intermediate states or unstable particles, such as the Z boson resonance near 91 GeV in electron-positron collisions, where the Breit-Wigner form quantifies the width and position. Quantum effects like level anticrossing, or avoided crossings, occur when nearly interact via coupling, repelling each other and forming a minimum gap proportional to the interaction strength, as seen in fine-structure perturbations or molecular potential curves. In vibrational spectra, arises from anharmonic coupling between a fundamental mode and an of another mode with similar energy, such as the symmetric stretch and bending in CO₂, splitting and intensifying spectral lines beyond harmonic predictions. Applications of atomic and molecular resonance include (MRI), which uses NMR principles to generate three-dimensional images of tissue water content and structure by detecting proton spin resonances in varying magnetic fields, revolutionizing non-invasive diagnostics since the 1970s. Laser cooling techniques rely on optical resonance detuning, where laser light slightly off-resonance from atomic transitions imparts momentum via photon absorption and , reducing kinetic energy and temperatures to microkelvin levels for Bose-Einstein condensate formation.

Broader Implications

Universal Resonance Curve

The universal resonance curve represents the normalized response of a lightly damped to a driving force, applicable across diverse physical systems. It is typically plotted as the of the |χ(ω)| normalized by its static value |χ(0)| versus the driving ω normalized by ω₀, yielding a characteristic lineshape. For lightly damped systems, this curve exhibits a symmetric peak near ω/ω₀ ≈ 1, with the peak height approaching Q (the factor) for high Q values, and a width inversely proportional to the coefficient γ, reflecting the sharpness of the resonance. The of the system wraps from 0 to π radians across the resonance, with the steepest change occurring near ω = ω₀, where the driving force and are approximately 90 degrees out of phase. This universal form arises from the equation of motion for a driven, damped oscillator, leading to the complex susceptibility χ(ω) = 1 / (ω₀² - ω² - i γ ω), where the real part describes the dispersive response and the imaginary part the absorptive response, both contributing to the envelope. The magnitude |χ(ω)| peaks sharply when ω ≈ ω₀ for small γ, while the (FWHM) of the curve is approximately γ, underscoring the role of in broadening the response. In the limit of low (γ ≪ ω₀), the curve's shape becomes nearly identical regardless of the specific system parameters, confirming its universality. The same normalized Lorentzian curve governs resonance phenomena in mechanical systems (e.g., driven masses on springs), electrical circuits (e.g., RLC tuned circuits), and atomic responses (e.g., electron transitions in dielectrics), with the quality factor Q = ω₀ / γ determining the peak height scaling as Q. This cross-disciplinary applicability stems from the shared underlying dynamics of second-order linear differential equations describing these oscillators. Historically, introduced this model in 1905 to explain anomalous dispersion in dielectrics, treating bound electrons as damped oscillators driven by electromagnetic fields. Modern computational simulations, including finite-difference time-domain methods and , consistently reproduce this universal curve across scales, from nanoscale optomechanical devices to macroscopic structural vibrations.

Disadvantages and Damping Effects

Uncontrolled resonance can lead to catastrophic structural failures when external forces match the natural frequency of a system, causing amplified oscillations that exceed design limits. A historical example is the collapse of the in on April 12, 1831, where soldiers marching in unison induced , resulting in the bridge's disintegration and the death of one soldier. Similarly, during earthquakes, seismic waves can excite a building's natural vibrational modes, amplifying ground motions and potentially causing collapse if the structure's resonant frequency aligns with dominant earthquake frequencies. Resonance also contributes to material in applications, where repeated stress cycles at or near the natural accelerate crack initiation and propagation. In aircraft wings, for instance, vibrational resonance from engine or aerodynamic forces can lead to cracks, compromising structural integrity over time and necessitating rigorous monitoring and protocols. Unintended resonance poses risks in everyday technologies, where frequencies are carefully selected to avoid harmful effects. ovens operate at 2.45 GHz, an band allocated for such devices that provides effective of food with a of about 1–5 cm, unlike higher frequencies near the ∼22 GHz rotational transitions of free water molecules, which would result in much shallower penetration (∼1 mm) and primarily surface heating. In (MRI), radiofrequency pulses are tuned to the Larmor of protons for diagnostic resonance, but safety limits on specific absorption rates () and strengths are enforced to mitigate risks of tissue heating or induced currents in implants. To mitigate these dangers, damping techniques are employed to dissipate and suppress resonant amplifications by increasing the damping coefficient β, which in turn reduces the quality factor and broadens the curve. Viscous damping, where resistive forces are proportional to , is commonly used in absorbers and hydraulic systems to smoothly attenuate oscillations. Frictional damping, involving dry or friction that provides constant opposition regardless of , is effective in bolted joints and systems for absorbing vibrational through sliding interfaces. Active control methods, such as piezoelectric actuators or electromagnetic devices that apply counter-forces in real-time based on , offer adaptive damping for dynamic environments like structures. While damping enhances stability by lowering Q factor peaks and preventing excessive amplitudes, it trades off sensitivity in applications requiring sharp resonance, such as sensors or filters, where broader responses reduce selectivity and efficiency.

References

  1. [1]
    Resonance - HyperPhysics
    A resonant frequency is a natural frequency of vibration determined by the physical parameters of the vibrating object.
  2. [2]
    The Feynman Lectures on Physics Vol. I Ch. 23: Resonance
    When there is a very sharp resonance, it corresponds to a very definite energy, just as though there were a particle of that energy present in nature. When the ...
  3. [3]
    6.3 Formal Charges and Resonance – Chemistry Fundamentals
    Resonance occurs in cases where two or more Lewis structures with identical arrangements of atoms but different distributions of electrons can be written. The ...
  4. [4]
    [PDF] MECHANICAL RESONANCE - Rutgers Physics
    The objectives of this experiment are: •. To study the resonance behavior of a mechanical oscillator. •. To measure the oscillator amplitude and phase as a ...
  5. [5]
    123. 16.8 Forced Oscillations and Resonance - UH Pressbooks
    The phenomenon of driving a system with a frequency equal to its natural frequency is called resonance. A system being driven at its natural frequency is said ...
  6. [6]
    Full article: Huygens' clocks: 'sympathy' and resonance
    1. In February 1665 Huygens made a discovery: two of the pendulum clocks hanging from a common wooden bar supported by two chairs demonstrated an odd ...
  7. [7]
    Resonant Pendulum | Exploratorium
    A very small force, when applied repeatedly at just the right time, can induce a very large motion. This phenomenon is known as resonance.<|control11|><|separator|>
  8. [8]
    Damped Driven Oscillator - Galileo
    We shall be using ω for the driving frequency, and ω0 for the natural frequency of the oscillator (meaning that ignoring damping, so ω0=√k/m. ) The Driven ...
  9. [9]
    [PDF] Physics Demonstrations in Sound & Waves: Part I
    At resonance, there is maximum energy transfer from the driving source to the oscillator, and a relatively small force can be used to obtain a large ...<|control11|><|separator|>
  10. [10]
    [PDF] Chapter 23 Simple Harmonic Motion - MIT OpenCourseWare
    Jul 23, 2022 · force acting on the spring is a linear restoring force, Fx = −k x (Figure 23.3). ... F = −k x − b. (23.5.2) x dt. Figure 23.13 Free-body ...
  11. [11]
    [PDF] Lecture 1: Simple Harmonic Oscillators
    Recall Hooke's law: if your displace a spring a distance x from its equilibrium position, the restoring force will be F = −kx for some constant k. You ...
  12. [12]
    15.1 Simple Harmonic Motion – General Physics Using Calculus I
    The angular frequency ω , period T, and frequency f of a simple harmonic oscillator are given by ω = k m , T = 2 π m k , and f = 1 2 π k m , where m is the mass ...Missing: natural | Show results with:natural
  13. [13]
    [PDF] Oscillations Simple Harmonic Motion - De Anza College
    The spring force is the restoring force. ... Figure 15.20 One example of. In the picture, that would be a fluid resistance force,. #». R = −b#»v. Fnet = −kx − b.
  14. [14]
    Simple Harmonic Motion Graphs - Physics
    Graphs of position, velocity, and acceleration. In SHM, the general equations for position, velocity, and acceleration are: x(t) = A cos(ωt + φ).
  15. [15]
    [PDF] RES.8-009 (Summer 2017), Lecture 5: Driven Oscillations
    2 = (ω + ω0)(ω − ω0)=2ω0Δω + O (Δω)2 . When the driving frequency matches the natural frequency of an oscillator we say that the system is in resonance. One of ...<|control11|><|separator|>
  16. [16]
    [PDF] Lecture 2: Driven oscillators
    Physically, the oscillator can't keep up with the driving force: it experiences phase lag. 2.3 Power and energy. We see from Eq. (25) there is a part of x(t) ...Missing: transfer | Show results with:transfer
  17. [17]
    driven oscillations - University of Hawaii System
    Jul 1, 2002 · It's a fancy word, but you will find that you already understand resonance in every day life. Example: Pushing a child on a swing (a pendulum).
  18. [18]
    [PDF] Lecture 4: RLC Circuits and Resonant Circuits
    For the series RLC circuit the impedance (Z) is: ◇. At resonance (series, parallel etc):. ◇. At the resonant frequency the following are true for a series RLC ...
  19. [19]
    [PDF] Chapter 21: RLC Circuits
    →The figure shows the current and emf of a series RLC circuit. To increase the rate at which power is delivered to the resistive load, which option should ...
  20. [20]
    [PDF] Frequency response: Resonance, Bandwidth, Q factor
    The current flowing in the system is in phase with the source voltage. The power dissipated in the RLC circuit is equal to the power dissipated by the resistor.
  21. [21]
    14.6 RLC Series Circuits – University Physics Volume 2
    Key Equations ; Time constant for a RL circuit, τ L = L / R ; Charge oscillation in LC circuits, q ( t ) = q 0 cos ( ω t + φ ) ; Angular frequency in LC circuits ...
  22. [22]
    [PDF] RLC Series Circuit - Physics
    What is the equation, and does the frequency given by the equation match the frequency you found in step 2? Hint: resonance is when the impedance is minimized.
  23. [23]
    RLC Series Circuit - HyperPhysics
    The component voltages can be obtained by multiplying the current times the component impedances. ... at resonance. Default values will be entered for unspecified ...Missing: across | Show results with:across
  24. [24]
    [PDF] RESONANCE IN SERIES AND PARALLEL RLC NETWORKS
    Series Resonance​​ The resonance of a series RLC circuit occurs when the inductive and capacitive reactance are equal in magnitude but cancel each other because ...Missing: derivation equations
  25. [25]
    16.6 Standing Waves and Resonance – University Physics Volume 1
    In Oscillations, we defined resonance as a phenomenon in which a small-amplitude driving force could produce large-amplitude motion. Think of a child on a swing ...
  26. [26]
    Reflection of Sinusoidal Waves from Boundaries
    Oct 8, 2011 · As the reflected wave superposes with the incident wave, a standing wave is quickly formed. There is no net transfer of energy The fixed end ...
  27. [27]
    16.6: Standing Waves and Resonance - Maricopa Open Digital Press
    The resonance produced on a string instrument can be modeled in a physics lab using the apparatus shown in (Figure). A lab setup for creating standing waves ...
  28. [28]
    General Solution of 1D Wave Equation
    In other words, standing waves are not fundamentally different to traveling waves. The wave equation, (8.33), is linear. This suggests that its most general ...
  29. [29]
    Standing Waves: A string fixed at both ends - Physics
    The lowest resonance frequency (n=1) is known as the fundamental frequency for the string. ... All stringed musical intruments have strings fixed at both ends.
  30. [30]
    Resonances of closed air columns - HyperPhysics
    A closed cylindrical air column will produce resonant standing waves at a fundamental frequency and at odd harmonics. The closed end is constrained to be a ...
  31. [31]
    Resonances of open air columns - HyperPhysics
    The resonant frequencies of air columns depend upon the speed of sound in air as well as the length and geometry of the air column.
  32. [32]
    11.4 Standing Waves
    When a guitar string is plucked or a violin string is bowed or a piano string is struck, standing waves are produced. This restricts the wavelengths-or the ...
  33. [33]
    Acoustic Waves with in an Organ Pipe - UCLA Physics & Astronomy
    The sound waves produced by the pipe are variations in air pressure that propagate in the direction of the pipe's axis. White regions indicate areas in which ...
  34. [34]
    [PDF] Seminar Notes: The Mathematics of Music - Yale University
    Sep 1, 2010 · Such beats can be heard, for instance, when two instruments of a musical ensemble play the same note slightly out of tune from each other.
  35. [35]
    [PDF] 15. Musical Acoustics - UC Davis Math
    resonance indicates the extent of the perturbative coupling. (a) tuning the A-string resonance through a coupled resonance at ≈ 460 Hz; (b) the splitting ...Missing: detuned | Show results with:detuned
  36. [36]
    [PDF] Oscillations Of A String - NYU Physics department
    Sometimes people refer to the ”Q” of a system, which is the ratio of the stored energy in the system / work done by the driving force during each radian. Q = ω ...
  37. [37]
    [PDF] Chapter 11 - Two and Three Dimensions - MIT OpenCourseWare
    We discuss the example of Chladni plates. iv. We give a two-dimensional example of a waveguide, in which the waves are con- strained to propagate only in one ...
  38. [38]
    Understanding the dynamics of biological and neural oscillator ...
    May 27, 2020 · Networks of Kuramoto–Sakaguchi oscillators have been used as models for synchronization phenomena. In neuroscience, individual oscillators ...
  39. [39]
    The Kuramoto model in complex networks - ScienceDirect.com
    This report is dedicated to review main contributions in the field of synchronization in networks of Kuramoto oscillators.
  40. [40]
    [PDF] Q factor
    ∆ω = ω0/Q. Thus, Q factor could also be defined as the ratio of the central frequency ω0 to the bandwidth ∆ω. 3 Summary. The main point of this write-up is to ...
  41. [41]
    [PDF] THE DAMPED HARMONIC OSCILLATOR - Oregon State University
    This number (times π) is the Quality factor or Q of the system. Page 10. L ... ˙ ˙ q +2β˙ q +ω. 0. 2 q = 0. L. R. C. I. Page 11. LCR circuit obeys precisely the.
  42. [42]
    Q Factor and Bandwidth of a Resonant Circuit | Electronics Textbook
    The Q, or quality, factor of a resonant circuit is a measure of the “goodness” or quality of a resonant circuit.Missing: interpretation | Show results with:interpretation
  43. [43]
    [PDF] Oscillators, Resonances, and Lorentzians - Todd Satogata
    Figure 1: Resonant amplitude response (scaled to driving amplitude X0, left) ... This equation is known as a Lorentzian function, related to the Cauchy ...
  44. [44]
    What Is Modal Analysis and Why Is It Necessary? | SimScale Blog
    Mar 11, 2024 · The modal analysis provides an overview of the limits of the response of a system, studying the response amplitude in terms of frequency.
  45. [45]
    [PDF] Chapter 2b - Method of Finite Elements I
    Institute of Structural Engineering. Page 23. Method of Finite Elements I. Modal Analysis. What are the eigenmodes of a given structure ? Global system of ...
  46. [46]
    Eigenfrequency Analysis - COMSOL
    Apr 19, 2018 · Eigenfrequencies or natural frequencies are certain discrete frequencies at which a system is prone to vibrate.
  47. [47]
  48. [48]
  49. [49]
    Resonance and the Factor of 2 Rule - Vibration Research
    To avoid over-amplification (even when underdamped), a component's resonant frequencies must be separated from the system's operational vibration inputs. When ...
  50. [50]
  51. [51]
    [PDF] INTERNATIONAL SPACE STATION MODAL CORRELATION ...
    Having a better understanding of the on orbit array mode frequencies and damping allows for less uncertainty to be placed on array modes during loads analysis ...Missing: resonance dampers
  52. [52]
    Development of a Passive Vibration Damping Structure for Large ...
    The damper was mounted at the bottom of the solar panel masts to dissipate the elastic energy induced by the solar panels and improve the overall damping ...3.2. Free Vibration Test Of... · 5.2. Free Vibration Test · 6.2. Thermal Vacuum Test
  53. [53]
    [PDF] Vibration Control in a 101-Storey Building Using a Tuned Mass ...
    May 23, 2014 · This study investigates the mitigating effects of a TMD on the structural dynamic responses of. Taipei 101 Tower, under the action of winds ...
  54. [54]
    [PDF] THE ANALYSIS OF TUNED MASS DAMPERS OF TAIPEI 101 TOWER
    The natural frequency of the damper is tuned so that under an excitation the damper is out of phase with the structure and goes to resonance itself, thus ...
  55. [55]
    Helmholtz resonators - Hao Tang - People | MIT CSAIL
    Jun 9, 2020 · This note derives the resonance frequency of Helmholtz resonators. The result is useful for building intuitions in acoustic phonetics.Missing: formula | Show results with:formula
  56. [56]
    Resonators for Noise Control - BYU Acoustics Research Group
    Acoustic resonators are used to amplify or absorb sound in very specific frequency ranges. Car mufflers and bass traps absorb unwanted noise.
  57. [57]
    Driving Room Modes - Dr. Russell's Acoustics Animations
    All rooms (a cavity volume enclosed by boundaries) has resonant frequencies at which the acoustic response to a source can be extremely large.Missing: echoes | Show results with:echoes
  58. [58]
    Acoustic whispering gallery modes within the theory of elasticity
    Whispering gallery modes, which are strongly localized near the rim of a resonator owing to total internal reflection, date back to the works of Rayleigh1,2 ...
  59. [59]
    [PDF] The Fabry-Perot Cavity Contents
    The dependence on the frequency argument Ω occurs through k = n(Ω)Ω/c. 1.3 Examples. Transmission for a train of pulses. Fabry-Perot as a frequency filter.
  60. [60]
    Q-factor - RP Photonics
    Definition via resonance bandwidth: the Q-factor is the ratio of the resonance frequency and the full width at half-maximum (FWHM) bandwidth of the resonance:Missing: interpretation | Show results with:interpretation
  61. [61]
    Acoustic Levitation Made Simple - AIP Publishing LLC
    Jan 5, 2015 · The reflected waves interact with newly emitted waves, producing what are known as standing waves, which have minimum acoustic pressure ...
  62. [62]
    High finesse of optical filter by a set Fabry-Perot cavity
    The Fabry–Perot (FP) filter or etalon is the most widely used optical filter for lightwave telecommunication systems. This popularity is due to its ...
  63. [63]
    Cavity-enhanced second-harmonic generation via nonlinear ...
    We apply this approach to obtain novel micropost and grating microcavity designs supporting strongly coupled fundamental and harmonic modes at infrared and ...
  64. [64]
    Parallel Resonant Circuits
    ### Summary of Parallel RLC Resonance from HyperPhysics
  65. [65]
    [PDF] Introduction to Quartz Frequency Standards. Revision - DTIC
    In most crystal oscillator types, a variable-load capacitor is used to adjust the frequency of oscillation to the desired value. Such oscillators operate at the ...
  66. [66]
    Wireless Power Transfer - Stanford University
    Nov 14, 2014 · 1: Tesla's system of resonant coils for wireless power transfer. The coils are "tuned" to resonate using a capacitor, labeled "condenser".<|separator|>
  67. [67]
    Wireless Power Transfer via Strongly Coupled Magnetic Resonances
    Using self-resonant coils in a strongly coupled regime, we experimentally demonstrated efficient nonradiative power transfer over distances up to 8 times ...
  68. [68]
    [PDF] CHAPTER 8 ANALOG FILTERS
    The transfer function of a band-pass filter is then: ωo here is the frequency (F0 = 2 π ω0) at which the gain of the filter peaks. Ho ...
  69. [69]
    [PDF] Advancements in Superconducting Microwave Cavities & Qubits
    Superconducting microwave cavities with ultra-high Q-factors are revolutionizing the field of quantum computing, offering long coherence times exceeding 1 ms, ...
  70. [70]
    Superconducting microwave cavities and qubits for quantum ...
    Jan 17, 2024 · Superconducting microwave cavities featuring ultrahigh Q-factors, which measure the efficiency of energy storage in relation to energy loss ...
  71. [71]
    6.3: Atomic Spectra and Models of the Atom - Chemistry LibreTexts
    Apr 12, 2023 · Like Balmer's equation, Rydberg's simple equation described the wavelengths of the visible lines in the emission spectrum of hydrogen (with n1 = ...Missing: seminal review
  72. [72]
    Balmer Series - an overview | ScienceDirect Topics
    The Balmer series refers to the set of spectral lines corresponding to the transitions of an electron in a hydrogen atom from higher energy levels to the ...Missing: seminal paper
  73. [73]
    [PDF] Pulsed Rabi oscillations in quantum two-level systems
    Dec 15, 2017 · In this article, we show a simple and intuitive model to better understand the temporal excitation dynamics of quantum two-level systems.Missing: explanation | Show results with:explanation
  74. [74]
    [PDF] MITOCW | watch?v=MVOJloovd18
    They had previously observed the Rabi oscillations at early times, but now the experiment had to be adjusted, I think by using slower atoms, to observe the ...
  75. [75]
    Vibrational Spectroscopy as a Tool for Bioanalytical and ... - NIH
    The main part of the review describes the basic principles and concepts of vibration spectroscopy and microspectrophotometry, in particular IR spectroscopy, mid ...Missing: paper | Show results with:paper
  76. [76]
    Vibrational spectroscopy by means of first‐principles molecular ...
    Mar 1, 2022 · This review article summarizes the field of vibrational spectroscopy by means of FPDM and highlights recent advances made such as the ...
  77. [77]
    [2401.01389] A mini review of NMR and MRI - arXiv
    Jan 2, 2024 · In this mini-review, we consider the concepts of NMR and MRI technologies from their fundamental origins to applications in medical science.
  78. [78]
    Cyclotron Resonance - an overview | ScienceDirect Topics
    Cyclotron resonance is defined as the condition when the wave frequency matches the cyclotron frequency, characterizing the gyro-motion of charged particles ...
  79. [79]
    Cyclotron Resonance - National MagLab
    Aug 22, 2023 · Application of a magnetic field gives rise to a Lorentz force that drives quasiparticles (correlated electrons) across the Fermi surface (FS) at ...
  80. [80]
    [PDF] 50. Resonances - Particle Data Group
    Dec 1, 2021 · Typically, a resonance appears as a peak in the total cross section. If the structure is narrow and if there are no relevant thresholds or ...
  81. [81]
    Level crossings and anticrossings - ScienceDirect.com
    Anticrossing refers to the situation where the levels involved are prevented from crossing by the presence of a small interaction which couples the two ...
  82. [82]
    Avoided level crossings with exponentially closing gaps in quantum ...
    Avoided level crossings can create exponentially closing gaps, which can lead to exponentially long running times for optimization problems. In this paper, we ...
  83. [83]
    Fermi Resonance Effects in the Vibrational Spectroscopy of Methyl ...
    Results are compared to the infrared (IR) spectroscopy of four molecules studied under supersonic expansion cooling in gas phase conditions. The molecules ...Missing: explanation | Show results with:explanation
  84. [84]
    Laser cooling at resonance | Phys. Rev. A
    May 3, 2018 · We show experimentally that three-dimensional laser cooling of lithium atoms on the D 2 line is possible when the laser light is tuned exactly to resonance.
  85. [85]
    [PDF] Oscillators, Resonances, and Lorentzians - Todd Satogata
    1 The Driven, Damped Simple Harmonic Oscillator. Consider a driven and damped simple harmonic oscillator with resonance frequency ω0: x(t) + ζω0 ˙x(t) + ω2.
  86. [86]
    [PDF] Where from Lorentzian?
    Oct 11, 2017 · The Lorentzian describes the line shape of transition processes involving atoms and laser fields. The three main processes are stimulated ...
  87. [87]
    Resonance - Richard Fitzpatrick
    We conclude that the height and width of the resonance peak of a weakly damped ( $Q\gg 1$ ) oscillator scale as $Q$ and $Q^{-1}$ , respectively. Thus, the ...
  88. [88]
    [PDF] Lorentz Oscillator Model - BYU Physics and Astronomy
    Jul 7, 2025 · The Lorentz oscillator model models an electron as a driven damped harmonic oscillator, bound to a nucleus with a spring force and a damping ...
  89. [89]
    [PDF] The Lorentz Oscillator and its Applications - MIT OpenCourseWare
    Now we will combine the concepts of complex permittivity, the Lorentz oscillator model, and the wave equation to describe how electromagnetic fields propagate ...
  90. [90]
    The British Army stopped walking in-step on bridges because of an ...
    Sep 28, 2021 · In 1831 a suspension bridge in Greater Manchester, England, collapsed while British troops marched over it. The bridge had a few design flaws.
  91. [91]
    Building Resonance: Structural stability during earthquakes - IRIS
    All buildings have a natural period, or resonance, which is the number of seconds it takes for the building to naturally vibrate back and forth.
  92. [92]
    NON-LINEAR VIBRATION METHOD FOR DETECTION OF FATIGUE ...
    Some results of mathematical simulations of bending vibrations encountered in a cracked aircraft wing under external harmonic excitation are presented.
  93. [93]
    Why are the microwaves in a microwave oven tuned to water?
    Oct 15, 2014 · The microwaves in a microwave oven are not tuned to a resonant frequency of water. In fact, the microwaves generated inside a microwave oven ...Missing: avoidance | Show results with:avoidance
  94. [94]
    Benefits and Risks - FDA
    Dec 9, 2017 · An MR Safe device is nonmagnetic, contains no metal, does not conduct electricity and poses no known hazards in all MR environments.
  95. [95]
    Viscous Damping - an overview | ScienceDirect Topics
    Viscous damping is defined as a damping model where the damping force is proportional to the velocity of a kinematic variable and acts in the opposite ...
  96. [96]
    [PDF] Damping and Vibration Control - NRAO Library
    Oct 30, 1994 · The damping mechanisms existing include material damping, friction damping, viscoelastic damping, viscous damping, dynamic absorber, tuned mass.
  97. [97]
    8.3: Damping and Resonance - Physics LibreTexts
    Nov 8, 2022 · A common damping force to account for is one for which the force is proportional to the velocity of the oscillating mass, and in the opposite direction of its ...