Fact-checked by Grok 2 weeks ago

Standing wave

A standing wave, also known as a stationary wave, is a wave pattern that oscillates in time while maintaining a fixed spatial profile, resulting from the superposition of two waves of identical and traveling in opposite directions. This creates regions of constructive and destructive , leading to points of no (nodes) and maximum (antinodes), giving the appearance of a stationary vibration without net propagation. Standing waves typically form in confined media where a traveling wave reflects off a , such as a fixed end, and the reflected wave interferes with the incoming wave. For the pattern to stabilize, the driving must one of the system's frequencies, or harmonics, allowing the waves to reinforce each other in a periodic manner. This phenomenon is fundamental in mechanical systems like stretched strings and longitudinal waves in air columns, where conditions determine the possible wavelengths and frequencies. In practical applications, standing waves are essential to the production of in musical instruments, such as the fundamental tones on guitar strings or . They also appear in electromagnetic contexts, like microwaves in cavities or resonators, influencing fields from acoustics to . The mathematical description of a one-dimensional standing wave can be expressed as the sum of two counter-propagating sinusoidal waves: y(x,t) = 2A \sin(kx) \cos(\omega t), where k is the wave number and \omega is the .

Formation of Standing Waves

Superposition of Counter-Propagating Waves

A standing wave arises from the , where two waves of identical and interfere in a uniform medium, producing points of constructive interference (antinodes) with maximum displacement and points of destructive interference (nodes) with zero displacement. This phenomenon occurs when two traveling waves propagate in opposite directions toward each other; for instance, one wave moves to the right while the other moves to the left, both maintaining the same \lambda and speed v in the medium. The displacement \psi(x,t) of the resulting standing wave is derived by adding the displacements of these counter-propagating . Assume the rightward wave is \psi_1(x,t) = A \sin(kx - \omega t) and the leftward wave is \psi_2(x,t) = A \sin(kx + \omega t), where A is the , k = 2\pi / \lambda is the wave number, and \omega = 2\pi f is the . Using the trigonometric identity \sin(a) + \sin(b) = 2 \sin\left(\frac{a+b}{2}\right) \cos\left(\frac{a-b}{2}\right), the superposition yields: \psi(x,t) = 2A \sin(kx) \cos(\omega t) In this , the term \sin(kx) governs the , fixing the locations of nodes where \sin(kx) = 0 (i.e., at x = n\pi / k for n) and antinodes where |\sin(kx)| = 1 (i.e., maximum ). The \cos(\omega t) term describes the uniform temporal across the medium, with all points oscillating in but with varying amplitudes determined by position. Phase differences between the counter-propagating waves cause this pattern: at nodes, the waves are always 180 degrees out of phase, canceling each other completely regardless of time; at antinodes, they are in phase, reinforcing to twice the original amplitude. Between nodes and antinodes, partial reinforcement occurs based on the local phase alignment. The discovery of harmonic ratios in vibrating musical strings dates back to the 6th century BCE, when experimented with string lengths to identify pleasing musical intervals, laying groundwork for understanding in acoustics.

Influence of Medium and Boundaries

Standing waves arise from the reflection of traveling waves at boundaries, where the reflected wave interferes with the incident wave to produce counter-propagating components of equal and frequency. In a bounded medium, such as a or , the boundary acts as a reflector, sending back a wave that travels in the opposite direction and superposes with the original wave via the principle of superposition. This process sustains the standing pattern only if the reflections maintain coherence, preventing the wave from dissipating as it propagates indefinitely. For constructive to form stable standing , the length must correspond to an multiple of half-wavelengths, ensuring that the relationship between incident and reflected aligns to reinforce fixed and antinodes. The type of significantly influences this: a fixed , where is zero (e.g., a clamped end), inverts the wave by 180 degrees upon , leading to a at the . In contrast, a free , where or transverse is zero (e.g., a loose end), reflects the wave without inversion, resulting in an antinode at the . These conditions dictate the allowable mode shapes, with fixed favoring odd harmonics in some systems and free allowing even modes as well. The properties of the medium further govern standing wave formation; in a uniform, non-dispersive medium, waves propagate at constant speed regardless of , enabling perfect patterns. Non-uniform media introduce variations in or , causing partial reflections and that disrupt the needed for standing waves, often resulting in quasi-standing or traveling patterns. , where wave speed depends on , prevents ideal standing waves by causing different components to dephase over time, leading to or of the pattern even in bounded systems. Resonance occurs when an external driving matches the natural frequencies set by the medium's length and conditions, amplifying the standing wave as input aligns with the system's modes. In real media, however, from , , or internal losses gradually reduces the amplitude of standing waves over time, as dissipates into , limiting sustained to low-damping environments. This temporal highlights the distinction between ideal theoretical models and practical observations, where standing waves persist only briefly without continuous supply.

Mathematical Description

Standing Wave on an Infinite String

A standing wave on an infinite arises in a non-dispersive medium where propagate at a constant speed, independent of or . For a taut , this wave speed v is determined by the T and the linear density \mu, given by the formula v = \sqrt{T/\mu}, derived from the balance of forces in small string segments leading to the one-dimensional . This assumption holds for small-amplitude transverse vibrations where effects are negligible, allowing sinusoidal solutions without distortion over distance. The mathematical form of such a standing wave emerges from the superposition of two counter-propagating traveling waves of equal amplitude A and , traveling in opposite directions along the string. Consider a right-going wave \psi_R(x,t) = A \sin(kx - \omega t) and a left-going wave \psi_L(x,t) = A \sin(kx + \omega t), where k = 2\pi/\lambda is the and \omega = 2\pi f is the , with the \omega = k v ensuring both waves share the same speed v. The total displacement is then \psi(x,t) = \psi_R(x,t) + \psi_L(x,t). Applying the trigonometric identity for the sum of sines, \sin a + \sin b = 2 \sin\left(\frac{a+b}{2}\right) \cos\left(\frac{a-b}{2}\right), yields: \begin{align} \psi(x,t) &= A \sin(kx - \omega t) + A \sin(kx + \omega t) \\ &= 2A \sin(kx) \cos(\omega t). \end{align} This separates into a spatial part $2A \sin(kx) that oscillates in amplitude along the string and a temporal part \cos(\omega t) that modulates the entire pattern uniformly in time, producing fixed nodes and antinodes. The infinite extent of the string permits this pure standing pattern, as there are no boundaries to cause reflections or interfere with the counter-propagating waves, allowing the superposition to maintain a stationary wave profile indefinitely. In this ideal case, the \lambda = 2\pi / k and f = \omega / 2\pi = v / \lambda are related directly through the constant speed v, without constraints on possible values, unlike bounded systems. Any \lambda and corresponding f can form a standing wave, as the lack of endpoints imposes no quantization. However, real strings are finite in length, so the string model serves as an valid primarily near the center of a long string, far from boundary influences that would otherwise quantize modes and alter the wave pattern.

Standing Wave on a Finite String with Fixed Ends

A standing wave on a finite string fixed at both ends arises when transverse waves reflect off the boundaries and interfere constructively, resulting in a stationary pattern characterized by nodes at the endpoints. The boundary conditions require that the transverse displacement \psi(x, t) satisfies \psi(0, t) = 0 and \psi(L, t) = 0 for all times t, where L is the length of the string; these conditions enforce zero displacement at the fixed ends. Applying these to the general wave solution leads to quantized normal modes, where the allowed wavelengths are \lambda_n = 2L / n for positive integers n = 1, 2, 3, \dots, corresponding to half-wavelength segments fitting exactly between the ends. The general solution for the displacement is a superposition of these normal modes: \psi(x, t) = \sum_{n=1}^{\infty} B_n \sin\left(\frac{n\pi x}{L}\right) \cos\left(\frac{n\pi [v](/page/V.) t}{L} + \phi_n\right), where B_n and \phi_n are the amplitude and of the nth mode, respectively, and v is the wave speed on the string. This form satisfies the wave equation and boundary conditions, with the sine term ensuring nodes at x = 0 and x = L, while the cosine term describes temporal . The frequencies of these modes derive from the \omega_n = 2\pi f_n = n \pi v / L, yielding f_n = n v / (2L); the (n=1) is f_1 = v / (2L), and higher are integer multiples thereof, forming a harmonic series. In these quantized modes, the total of the is the of the energies in each , with the of proportional to B_n^2 \omega_n^2, reflecting the of the modes that prevents between them. Thus, the overall is E = \frac{1}{2} \mu L \sum_{n=1}^{\infty} B_n^2 \omega_n^2 / 2, where \mu is the linear density, though the exact partitioning depends on initial conditions. Specific harmonics are excited by initial disturbances such as plucking or striking the at particular locations; for instance, plucking at the (x = L/2) primarily excites odd-n modes due to , while plucking closer to one end emphasizes higher harmonics.

Standing Wave on a String with One Fixed End

A standing wave on a fixed at one end and at the other arises when transverse reflect and interfere under these asymmetric boundary conditions. The fixed end, typically at position x = 0, enforces a where the is zero for all time, so \psi(0, t) = 0. At the end, located at x = L, there is no transverse , resulting in zero of the , \partial \psi / \partial x (L, t) = 0. These conditions contrast with the symmetric at both ends in the fixed-fixed case and lead to distinct resonant modes. To satisfy the boundary conditions, the allowed standing wave modes correspond to odd quarter-wavelength multiples fitting the string length L. The wavelengths are given by \lambda_n = \frac{4L}{2n-1} for positive integers n = 1, 2, 3, \dots, producing only odd harmonics. The corresponding frequencies are f_n = \frac{(2n-1) v}{4L}, where v is the wave speed on the string, determined by the tension and linear density. For the fundamental mode (n=1), the frequency is f_1 = \frac{v}{4L}, and higher modes follow as odd multiples of this value. The general displacement function for the standing wave is a superposition of these modes: \psi(x, t) = \sum_{n=1}^{\infty} C_n \sin\left( \frac{(2n-1)\pi x}{2L} \right) \cos\left( \frac{(2n-1)\pi v t}{2L} + \phi_n \right), where C_n are amplitudes determined by initial conditions and \phi_n are phase angles. Each term represents a , with the spatial part \sin\left( \frac{(2n-1)\pi x}{2L} \right) ensuring the node at x=0 and antinode (maximum amplitude) at x=L. In terms of mode shapes, the mode (n=1) spans a along the string, with a at the fixed end and an antinode at the free end, differing from the half-wavelength in the fixed-fixed where nodes occur at both boundaries. Higher modes (n=2, 3, \dots) add additional nodes between the ends, but always maintain the antinode at the free end, resulting in asymmetric patterns with odd frequencies. This fixed-free configuration is relevant in physical systems like flagpoles, where wind-induced vibrations create standing waves with the base fixed and the top free, or beams in mechanical structures exhibiting similar transverse resonances.

Standing Wave in a Pipe

Standing waves in pipes typically refer to longitudinal acoustic waves in cylindrical tubes filled with a , such as air, where the wave manifests as variations in and along the pipe's length. Unlike transverse waves on a , these are compression-rarefaction waves; at a closed end, the has a (no longitudinal motion), leading to a antinode (maximum variation), while at an open end, the has a (equal to ), resulting in a displacement antinode. For a pipe closed at one end and open at the other, the boundary conditions yield standing waves only for odd harmonics. The wavelengths are given by \lambda_n = \frac{4L}{2n-1}, where L is the pipe length and n = 1, 3, 5, \dots, corresponding to frequencies f_n = \frac{(2n-1)v}{4L}, with v the in the fluid. The fundamental mode (n=1) has \lambda_1 = 4L and f_1 = v/(4L), placing a displacement node at the closed end and an antinode at the open end. In an open pipe, with both ends open, standing waves support all integer harmonics, with wavelengths \lambda_n = \frac{2L}{n} for n = 1, 2, 3, \dots, and frequencies f_n = \frac{n v}{2L}. The fundamental mode has \lambda_1 = 2L and f_1 = v/(2L), featuring pressure nodes at both ends and a pressure antinode in the middle. The pressure variation for the nth mode in an open pipe can be expressed as p(x,t) = D_n \sin\left(\frac{n\pi x}{L}\right) \sin\left(\frac{n\pi v t}{L} + \theta_n\right), where D_n is the amplitude and \theta_n the phase; the general solution is a superposition over modes: p(x,t) = \sum_n D_n \sin\left(\frac{n\pi x}{L}\right) \sin\left(\frac{n\pi v t}{L} + \theta_n\right). This form ensures pressure nodes at x=0 and x=L. The v in the pipe's fluid, assuming an undergoing adiabatic compression, is v = \sqrt{\frac{\gamma P}{\rho}}, where \gamma is the adiabatic index, P the , and \rho the ; this derives from the for adiabatic processes, B = \gamma P, combined with v = \sqrt{B/\rho}. In real pipes, the ideal conditions are approximate due to the finite size of the open end; an end correction \Delta L \approx 0.6 r (where r is the pipe radius) must be added to the effective length L for more accurate frequencies. For a closed pipe, this correction applies only at the open end, while for an open pipe, it applies at both ends, yielding effective length L + 1.2 r. This empirical adjustment accounts for the antinode extending slightly beyond the physical open end.

Two-Dimensional Standing Wave with Rectangular Boundary

In a rectangular domain with dimensions L_x along the x-direction and L_y along the y-direction, the two-dimensional \frac{\partial^2 \psi}{\partial t^2} = c^2 \left( \frac{\partial^2 \psi}{\partial x^2} + \frac{\partial^2 \psi}{\partial y^2} \right) describes the propagation of waves, such as those on a vibrating or in an acoustic . Fixed boundaries, corresponding to Dirichlet conditions where \psi = 0 on all edges (x=0, L_x; y=0, L_y), imply nodes along the entire perimeter. The method of assumes a product \psi(x,y,t) = X(x) Y(y) T(t), which decouples the equation into three ordinary differential equations: one for the spatial dependence in x, one in y, and one for time. The boundary conditions yield sinusoidal solutions for the spatial parts: X(x) = \sin(k_x x) and Y(y) = \sin(k_y y), with wave numbers k_x = \frac{m \pi}{L_x} and k_y = \frac{n \pi}{L_y}, where m, n = 1, 2, [3, \dots](/page/3_Dots) are positive integers denoting the mode indices. The time dependence is , T(t) = A \cos(\omega_{mn} t) + B \sin(\omega_{mn} t), with given by \omega_{mn} = \pi c \sqrt{\left( \frac{m}{L_x} \right)^2 + \left( \frac{n}{L_y} \right)^2}, where c is the wave speed in the medium. Thus, the full mode shape is the product of two independent one-dimensional standing waves, \psi_{mn}(x,y,t) = \sin\left( \frac{m \pi x}{L_x} \right) \sin\left( \frac{n \pi y}{L_y} \right) [A \cos(\omega_{mn} t) + B \sin(\omega_{mn} t)]. Degenerate modes arise when distinct pairs (m,n) and (m',n') yield the same , i.e., \left( \frac{m}{L_x} \right)^2 + \left( \frac{n}{L_y} \right)^2 = \left( \frac{m'}{L_x} \right)^2 + \left( \frac{n'}{L_y} \right)^2. For a square domain where L_x = L_y, this occurs for pairs like (1,2) and (2,1), allowing linear combinations that form more complex nodal patterns. These solutions provide the foundation for analyzing phenomena such as of a rectangular and the axial, tangential, and oblique modes in rectangular room acoustics. In electromagnetic contexts, similar modes appear in rectangular cavities, influencing design.

Physical Properties

Nodes, Antinodes, and Mode Shapes

In standing waves, nodes are stationary points where the amplitude of oscillation is zero, resulting from complete destructive between the counter-propagating waves. Antinodes, conversely, are points of maximum , where constructive causes the medium to oscillate with the largest . These features arise in various media, such as or air columns, and their positions depend on the and boundary conditions; for instance, in a fixed at both ends, nodes occur at multiples of half-wavelengths from one end. Mode shapes describe the distinct spatial patterns of these nodes and antinodes in harmonic standing waves, with each mode corresponding to a specific multiple of the . The nth features n half-wavelengths fitting within the medium's length, resulting in n−1 nodes between the boundaries (plus nodes at the ends if fixed). For example, the fundamental (n=1) has no interior nodes and one antinode at the center, while the second (n=2) introduces one interior node at the , dividing the pattern into two oscillating segments. Higher modes thus exhibit increasingly complex geometries with more nodes and antinodes, enabling richer vibrational behaviors in systems like musical instruments. The spatial structure of standing waves can be visualized through the time-independent envelope, often expressed as \sin(kx), where k is the wave number, delineating the fixed positions of nodes (where \sin(kx) = 0) and antinodes (where \sin(kx) = \pm 1). This sinusoidal profile highlights the wave's stationary nature, with the medium oscillating vertically around the equilibrium while the pattern remains fixed in space. As detailed in the mathematical description, specific forms like \sin\left(\frac{n\pi x}{L}\right) for a finite of L define these shapes for each . Standing wave modes possess orthogonality, meaning their spatial functions are mutually independent and can be integrated to zero when different modes are multiplied, a property that underpins the decomposition of arbitrary initial displacements into a sum of modes via Fourier series. This allows any complex vibration to be expressed as a linear combination of these orthogonal modes, facilitating analysis in both classical and quantum contexts. In slightly perturbed systems, such as those with minor inhomogeneities or detuned frequencies, mode coupling can occur, leading to energy transfer between modes and phenomena like beating, where the superposition produces temporal amplitude modulation at the difference frequency. This instability contrasts with the stable, uncoupled modes in ideal conditions and is observed in plasma waves.

Standing Wave Ratio, Phase, and Energy Localization

In transmission lines and waveguides, the (SWR), also known as the voltage standing wave ratio (VSWR), quantifies the degree of impedance mismatch between the line and its load, arising from partial reflections of the wave. The SWR is defined as the ratio of the maximum to minimum voltage along the line and is calculated using the magnitude of the Γ, which represents the ratio of the reflected wave to the incident wave , via the formula: \text{VSWR} = \frac{1 + |\Gamma|}{1 - |\Gamma|} where |\Gamma| ranges from 0 (perfect match, no reflection) to 1 (total reflection). In modern RF engineering, a VSWR below 2:1 is typically considered acceptable for antenna systems, as it corresponds to less than 11% of the incident power being reflected, ensuring efficient energy transfer to the load. Partial reflections introduce a progressive shift along the standing wave , determined by the phase of the complex reflection coefficient Γ, which varies the positions of nodes and antinodes relative to a pure standing wave. This phase variation arises from the superposition of the incident and reflected waves, where the relative phase difference causes the envelope of the wave to modulate spatially, leading to a non-uniform that shifts with frequency or load changes. In a pure standing wave, where the incident and reflected waves have equal amplitudes (|\Gamma| = 1), there is no net energy transfer along the medium; instead, the total energy oscillates locally between kinetic and potential forms, with maximum kinetic energy at antinodes (where displacement is greatest) and maximum potential energy at nodes (where displacement is zero). This results in zero net energy flux, as the forward and backward power flows cancel exactly. In imperfect cases with partial standing waves (|\Gamma| < 1), the unequal amplitudes of the incident and reflected components introduce a traveling wave superposition, allowing net energy propagation toward the load despite the oscillatory standing pattern. The reflected power reduces overall efficiency, but the forward-propagating component delivers net energy, with the fraction of delivered power given by (1 - |\Gamma|^2).

Applications and Phenomena

Acoustic and Mechanical Resonances

Acoustic resonances occur when standing waves form in enclosed air volumes, amplifying sound at specific frequencies. A Helmholtz resonator, consisting of a cavity connected to the exterior by a narrow neck, supports a standing wave where the air in the neck oscillates like a mass on a spring, with the cavity acting as the spring, leading to a fundamental resonance frequency determined by the geometry and speed of sound. This configuration, first described by in 1863, is widely used in noise control and musical instruments due to its sharp resonance peak. In organ pipes, standing waves establish along the pipe's length, with resonance frequencies depending on whether the pipe is open or closed at one end, as outlined in the mathematical description of waves in pipes. For an open pipe, the fundamental mode has antinodes at both ends, producing a wavelength twice the pipe length and a frequency of f = v / (2L), where v is the speed of sound and L is the length; higher harmonics follow as integer multiples (n=1,2,3,...). These resonances drive the pipe's oscillation when excited by airflow, enabling the production of distinct musical tones in pipe organs. Mechanical resonances in solids involve standing waves on or within vibrating structures. In stringed instruments like the guitar and violin, standing waves form on taut strings fixed at both ends, with the fundamental mode having a single antinode in the middle and frequency f = (1/(2L)) \sqrt{T/\mu}, where T is tension, \mu is linear density, and L is length, as detailed in the finite string model. Plucking or bowing excites these modes, and the instrument's body resonates sympathetically to amplify the sound, with harmonics contributing to timbre. Chladni plates demonstrate two-dimensional standing waves in plates vibrated transversely, revealing nodal lines where sand accumulates due to zero displacement. These patterns emerge at resonant frequencies governed by the plate's shape and material properties, such as for a square plate where modes are products of sine functions along each dimension. First visualized by in the 18th century, the patterns illustrate flexural modes and are used to study plate vibrations in engineering. Seismic waves during earthquakes can produce standing modes in Earth's layered structure, acting as a waveguide where low-velocity zones trap waves, leading to resonant oscillations. Normal modes, or free oscillations of the entire planet, are standing waves excited by large earthquakes, with periods from tens to thousands of seconds corresponding to spheroidal and toroidal modes. These modes, observed globally via seismometers, provide insights into Earth's interior density and elasticity. Fault zones also support trapped standing waves, amplifying ground motion at specific frequencies.

Electromagnetic Standing Waves

Electromagnetic standing waves form when incident and reflected electromagnetic waves interfere within confined structures, resulting in stationary patterns of electric and magnetic field nodes and antinodes. These waves are fundamental to , , and microwave engineering, enabling phenomena from color production in thin films to resonance in cavities. Unlike mechanical waves, electromagnetic standing waves propagate at the speed of light in vacuum and exhibit no dispersion in free space, though material interactions can introduce frequency-dependent effects. In the visible light regime, standing waves manifest in thin films through multiple reflections at the film's boundaries, producing constructive or destructive interference that determines observed colors, such as the iridescent hues in soap bubbles or oil slicks. For more precise control, Fabry-Pérot etalons and cavities confine standing waves between two parallel, partially reflecting mirrors separated by distance L, where resonant frequencies occur when the cavity length accommodates an integer number of half-wavelengths. The spacing between these longitudinal modes, known as the free spectral range (FSR), is given by \Delta \nu = \frac{c}{2L}, with c the speed of light; this relation arises from the condition for phase-matching in the round-trip propagation. Such cavities achieve high finesse, enhancing transmission at resonances, as demonstrated in early optical experiments. At shorter wavelengths, X-ray standing waves emerge in crystalline lattices during , where the incident beam and its diffracted counterpart interfere to form periodic field patterns aligned with atomic planes. This interference satisfies the , which specify diffraction conditions via \mathbf{a}_i \cdot (\mathbf{k} - \mathbf{k}_0) = 2\pi h_i, \quad i = 1,2,3, relating the scattering vector \mathbf{k} - \mathbf{k}_0 to the reciprocal lattice vectors \mathbf{a}_i and integers h_i; these generalize n\lambda = 2d \sin\theta for multi-dimensional crystals. Pioneered in the 1910s, this effect allows precise probing of atomic positions, as X-rays penetrate crystals to depths of microns, creating standing waves that modulate fluorescence or photoelectron yields from surface adsorbates. In the microwave domain, standing waves in waveguides arise from boundary reflections, supporting transverse electric (TE) and transverse magnetic (TM) modes characterized by integer indices m and n that dictate field variations across the guide's cross-section. Each mode has a cutoff frequency f_c below which waves evanesce rather than propagate, given for a rectangular waveguide of dimensions a \times b by f_c = \frac{c}{2} \sqrt{\left(\frac{m}{a}\right)^2 + \left(\frac{n}{b}\right)^2}, ensuring single-mode operation in narrow frequency bands for applications like radar. Microwave cavities, formed by enclosing waveguides, sustain resonant standing waves at discrete frequencies, with mode patterns resembling solutions to two-dimensional rectangular boundary problems. Laser resonators rely on standing waves established between high-reflectivity mirrors, where counter-propagating fields form longitudinal modes spaced by the , selecting discrete emission frequencies from the gain medium. These modes, often Gaussian in profile, enable coherent output but can lead to spatial hole burning if multiple modes compete, reducing efficiency in single-frequency operation. Advancing into the 2020s, photonic crystals—periodic dielectric structures—have enabled engineered standing electromagnetic modes through bandgap engineering and topological protection, allowing defect-localized resonances for compact waveguides and sensors. Reviews highlight meta-crystal designs that manipulate EM parameters to create robust standing wave patterns, enhancing light-matter interactions in integrated photonics.

Quantum Mechanical Standing Waves

In quantum mechanics, the concept of standing waves extends to matter waves, as proposed by Louis de Broglie in his 1924 hypothesis that particles possess wave-like properties with wavelength λ = h/p, where h is Planck's constant and p is momentum. This idea suggested that electrons in atoms form standing de Broglie waves, leading to quantized orbits and discrete energy levels that explain atomic spectra. The particle in a box model illustrates this quantum standing wave behavior, where a particle of mass m is confined to a one-dimensional region of length L with infinite potential barriers at the ends. Solving the time-independent Schrödinger equation under these boundary conditions yields stationary standing wave solutions for the wave function: \psi_n(x) = \sqrt{\frac{2}{L}} \sin\left(\frac{n\pi x}{L}\right), where n = 1, 2, 3, ... is the quantum number, ensuring the wave function vanishes at x = 0 and x = L. The corresponding energies are quantized as E_n = \frac{n^2 h^2}{8 m L^2}, arising directly from the boundary-imposed standing wave condition, which prevents continuous energy values and produces discrete spectral lines observed in atomic emissions. Similar standing wave solutions emerge in the quantum harmonic oscillator, where a particle moves in a quadratic potential V(x) = (1/2) m ω² x², with ω the angular frequency. The Schrödinger equation here admits stationary states as standing waves, with wave functions involving Hermite polynomials modulated by a Gaussian envelope, resulting in equally spaced energy levels E_n = ħ ω (n + 1/2), n = 0, 1, 2, .... These solutions highlight how potential wells enforce standing wave patterns, quantizing vibrations in molecules and solids. In solid-state physics, electron waves in periodic crystal lattices form Bloch waves, which are standing wave-like superpositions of plane waves modulated by the lattice periodicity, enabling band structures and conductivity. Quantum dots, nanoscale semiconductor confinements acting as artificial atoms, exhibit discrete standing wave states analogous to the particle in a box, with tunable energy levels observed via resonant tunneling, facilitating applications in quantum computing and optoelectronics.

Fluid and Surface Wave Examples

Seiches represent a classic example of standing waves in enclosed or semi-enclosed bodies of water, such as lakes and harbors, where the water surface oscillates at natural resonant frequencies. These oscillations arise from the interference of incident and reflected waves, forming nodes and antinodes along the basin length. In lakes like , seiches are primarily forced by sustained winds that push water toward one end, causing it to slosh back and forth with periods ranging from minutes to hours, depending on basin geometry. Atmospheric pressure gradients can also contribute, amplifying the setup by altering the water level without direct wind shear. In harbors, seiches often result from long-period infragravity waves generated offshore by nonlinear interactions of wind waves, leading to resonant amplification if the harbor dimensions match a fraction of the wave wavelength. For instance, wind-driven seiches in Great Lakes harbors can reach amplitudes of several meters, posing risks to navigation and docking operations. The dynamics are governed by shallow-water wave equations, with damping from bottom friction and viscosity reducing amplitude over time. Faraday waves, also known as Faraday instability patterns, occur on the surface of a fluid layer subjected to vertical oscillatory forcing, producing subharmonic standing wave patterns at half the driving frequency. These patterns emerge when the vertical acceleration amplitude exceeds a critical , typically on the order of a few times for low-viscosity fluids, leading to hexagonal or striped arrays of surface undulations. The instability is parametric, with energy transfer from the oscillating container to the fluid surface via inertial forces. For deep fluid layers, the threshold acceleration satisfies a > 4g \left( \frac{\lambda}{2\pi} \right)^2, where \lambda is the of the standing pattern, highlighting the role of in stabilizing the flat surface below onset. Experimental studies confirm that near , patterns form via a , with subharmonic response dominating due to lower energy requirements compared to harmonic modes. These waves are observed in various fluids, including and silicone oils, and their nonlinear evolution can lead to complex spatiotemporal chaos at higher amplitudes. In stratified fluids, such as the where increases with depth due to and gradients, internal standing waves form at interfaces or within continuous layers, distinct from surface waves. These modes arise when progressive internal gravity waves reflect off boundaries like the seafloor or pycnocline, creating resonant oscillations with vertical displacements confined to regions of stable stratification. In oceanic basins, standing internal waves are excited by tidal currents interacting with topography, such as seamounts, or by wind forcing at the surface that propagates downward. For example, in semi-enclosed seas, the fundamental mode has a at mid-depth and antinodes near the surface and bottom, with periods matching basin-scale resonances on the order of hours. Laboratory experiments demonstrate parametric excitation of standing internal waves in linearly stratified tanks, where vertical oscillations of the container generate subharmonic modes, analogous to Faraday waves but driven by rather than . In the , these waves contribute to vertical mixing by breaking at critical Richardson numbers, influencing transport and maintenance. Recent studies in microgravity conditions for space applications have revealed altered dynamics of standing waves in fluids without buoyancy dominance. Experiments on the have demonstrated Faraday waves in multi-layer fluid systems, showing driven by and vibrations, with potential applications in fluid management such as propellant sloshing control. In the absence of , capillary forces govern wave propagation, allowing standing patterns to form in vibrated containers. Such research addresses challenges in long-duration missions and informs the development of vibration-isolated fluid systems for lunar or Martian habitats. Sloshing in containers, particularly propellant tanks in spacecraft, manifests as standing waves on the liquid free surface during acceleration or attitude changes, critically affecting vehicle stability and control. In partially filled tanks, lateral or rotational motions excite resonant modes, such as the fundamental sloshing mode with an antinode at the center and nodes near walls, leading to pressure oscillations and torque that can couple with the spacecraft's dynamics. NASA design handbooks emphasize modeling these as equivalent mechanical oscillators, with damping baffles introduced to suppress higher modes and prevent nutation growth. For cryogenic propellants like liquid oxygen, sloshing periods scale with tank dimensions and fill levels, often in the range of seconds, impacting upper-stage performance during coast phases. Recent analyses incorporate nonlinear effects, where wave breaking enhances energy dissipation, guiding baffle geometries to minimize slosh-induced loads in missions like Artemis.

References

  1. [1]
    Standing Wave - an overview | ScienceDirect Topics
    A standing wave is defined as a wave pattern formed when a significant proportion of a traveling wave is reflected, resulting in the superposition of ...
  2. [2]
    Standing waves review (article) - Khan Academy
    Term (symbol) Meaning. Standing wave. Waves which appear to be vibrating vertically without traveling horizontally. Created from waves with identical frequency ...
  3. [3]
    Standing Wave | Definition, Examples & Formula - Lesson - Study.com
    A standing wave is a wave that results in the interference of two waves with the same frequency. They collide in a contained medium.
  4. [4]
    Formation of Standing Waves - The Physics Classroom
    A standing wave pattern is a vibrational pattern created within a medium when the vibrational frequency of the source causes reflected waves from one end of ...
  5. [5]
    [PDF] Chapter 16 Waves in One Dimension
    ... superposition/superposition.html. Page 24. Slide 16-24. © 2015 Pearson ... (x, t) = 2A sin kx cos ωt = [2A sin kx] cos ωt. © 2015 Pearson Education, Inc ...
  6. [6]
    [PDF] Physics 1C
    They interfere according to the superposition principle. The resultant wave will be: y = (2A sin kx) cos ωt ... standing wave (1 anti-node, first harmonic) we get ...
  7. [7]
    None
    ### Summary of Historical Context on Standing Waves or Harmonics in Musical Strings (Pythagoras and Ancient Greeks)
  8. [8]
    Reflection of Sinusoidal Waves from Boundaries
    Oct 8, 2011 · As the incident wave train is reflected from the free end, a standing wave is quickly formed as the right-going incident wave is superposed with ...
  9. [9]
    Formation of Standing Waves - music.ucsd.edu
    Reflection along one dimension causes right and left travelling waves within the medium. Travelling waves interfere (sum) constructively and destructively ...
  10. [10]
    16.6: Standing Waves and Resonance - Maricopa Open Digital Press
    The waves are visible in the photo due to the reflection from a lamp. These waves are formed by the superposition of two or more traveling waves, such as ...
  11. [11]
    Standing Waves in Finite Continuous Medium
    The only simple solution of the wave equation, (6.1), that has stationary nodes and anti-nodes is a standing wave of the general form.
  12. [12]
    Lecture 13: Dispersive Medium, Phase Velocity, Group Velocity
    Prof. Lee discusses the phenomenon of dispersion in a realistic medium and the strategy to describe this kind of physics situation.
  13. [13]
    [PDF] LECTURE NOTES 4 - High Energy Physics
    Standing waves do not propagate in space, although they can / do evolve in time. (due to dispersion, dissipation and other processes). Let us consider a 1- ...
  14. [14]
    [PDF] Superposition and Standing Waves - Physics Courses
    6. Damping, and non–linear effects in the vibration turn the energy of vibration into internal energy. ... Thus, standing waves will be excited in the bow string.
  15. [15]
    Analyzing Waves on a String - Galileo
    Solving the Wave Equation​​ v2=Tμ. All traveling waves move at the same speed—and the speed is determined by the tension and the mass per unit length. We could ...
  16. [16]
    16.3 Wave Speed on a Stretched String – University Physics Volume 1
    Since the speed of a wave on a taunt string is proportional to the square root of the tension divided by the linear density, the wave speed would increase by 2 ...
  17. [17]
    [PDF] Chapter 8 - Traveling Waves - MIT OpenCourseWare
    In this chapter, we show how the same physics that leads to standing wave oscillations also gives rise to waves that move in space as well as time. We then go ...Missing: influence | Show results with:influence
  18. [18]
    Traveling Waves in Infinite Continuous Medium
    Traveling Waves in Infinite Continuous Medium. Consider solutions of the wave equation, (6.1), in an infinite medium. Such a medium does not possess any ...
  19. [19]
    [PDF] Waves - Oberlin College and Conservatory
    infinite string but still unrealistic. But if I clamp another end, that ... that standing wave leak out of instrument so that listeners can hear it.
  20. [20]
    Normal Modes of Uniform String
    The normal modes of a uniform string exhibit smooth sinusoidal spatial variation in the $x$ -direction, and that the angular frequency of the modes is directly ...
  21. [21]
    The Feynman Lectures on Physics Vol. I Ch. 49: Modes
    In short, if we assume that the string is infinite ... Then the final result is that the waves bouncing about in the box produce a standing-wave pattern, that is, ...
  22. [22]
    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 the higher frequencies are known as harmonics - these are integer ...
  23. [23]
    Strings, standing waves and harmonics - UNSW
    If you pluck the low E string anywhere except one third of the way along, the B string should start to vibrate, driven by the vibrations in the bridge from the ...
  24. [24]
    Applying Boundary Conditions to Standing Waves - Brilliant
    Boundary conditions can lead to standing waves, where certain points called nodes have zero displacement at all times and the maximum amplitude of the wave at ...
  25. [25]
    Boundary Conditions for Standing Waves (DP IB Physics)
    Dec 17, 2024 · Fixed at both ends. Free at both ends. One end fixed, the other free. At specific frequencies, known as natural frequencies, an integer number ...
  26. [26]
    Standing Waves - The Physics Hypertextbook
    When the medium has one fixed end and one free end the situation changes in an interesting way. A node will always form at the fixed end while an antinode will ...Missing: flagpole | Show results with:flagpole
  27. [27]
    Resonances of closed air columns - HyperPhysics
    In actual practice, the position of the antinode is slightly outside the open end, and an end correction of about 0.6 times the radius of the pipe should be ...
  28. [28]
    Sound Waves - Galileo
    For example, a standing wave in a pipe has the form s(x,t)=Asinkxsinωt, this would be for a pipe closed at x=0, so that the air doesn't move at x=0. The ...
  29. [29]
    [PDF] Chapter 12 Lecture Notes Formulas: v ≈ (331 + 0.60T ) m/s I ≡ P/A I ...
    In general λn = 2L/n, f = nv/2L n = 1,2,3,... for a tube open at both ends. So if we have an open tube, the frequency is determined by the velocity of the wave ...
  30. [30]
    Contemporary Challenges: Fall-2023 Homework 4 (SOLUTION)
    The pressure at the open end of a pipe is fixed at 1 atmosphere (this ... This would give a standing wave P=P0sinkxsinkvt+(1 atm) ...
  31. [31]
    Sound Waves in Ideal Gas
    $$\displaystyle v=\sqrt{\frac{\gamma\,p, (5.40). is a constant with the dimensions of velocity, which turns out to be the sound speed in the gas. (See Section ...
  32. [32]
    [PDF] Physics of Music PHY103 Worksheet #6 Set up for Flute Lab
    End correction = 0.61 R​​ 0.61 R Length L Open tube acts like a tube of length L + 0.61R+ 0.61R Page 2 You make an open flute from 2cm diameter bamboo pipe that ...
  33. [33]
    [PDF] The two dimensional wave equation - Trinity University
    utt = c2∆u = c2(uxx + uyy ). We assume the membrane lies over the rectangular region R = [0,a] × [0,b] and has fixed edges. u(x,y,0) = f (x,y), (x,y) ∈ R, ut(x ...
  34. [34]
    [PDF] Mathematical Musical Physics of the Wave Equation – Part 2
    We will discuss finding the eigen-solutions, etc. to the 3-D wave equation associated with longitudinal acoustic standing waves confined within the interior ...
  35. [35]
    Some Formulae Related to Room Acoustics - Duke Physics
    Normal Modes of a Rectalinear Room. The standing wave modes in a pipe closed at both ends have frequencies given by. eqn0.gif. where n = 1,2,3,. . . is an ...
  36. [36]
    [PDF] Chapter 9 – Cavities and Waveguides (9.2) Rectangular Cavity
    We can use the separation of variables technique to solve the 3D wave equation and obtain the ... This sort of construction means that we will have standing waves ...
  37. [37]
    Standing waves
    Nodes are the result of the meeting of a crest with a trough. This leads to a point of no displacement. Standing waves are also characterized by antinodes.
  38. [38]
    Superposition and Interference – Introductory Physics for the Health ...
    Nodes are points of no motion in standing waves. An antinode is the location of maximum amplitude of a standing wave. Waves on a string are resonant standing ...
  39. [39]
    Acoustics Chapter One: Standing Waves
    When two traveling sound waves are propagating in opposite directions, as a result of reflections inside a bounded system, be it a tube, string or room, the ...Missing: formation | Show results with:formation
  40. [40]
    Standing Waves - Physics
    The standing wave equation has the spatial part (kx) separated from the time part. It predicts that the string is totally flat at certain points in time, and it ...
  41. [41]
    [PDF] Nonlinear Trivelpiece-Gould waves: Frequency, functional form, and ...
    through nonlinear mode coupling. This implies ... standing wave, noting that if m is even then so ... Beating between these two eigenmodes is then ...
  42. [42]
    Voltage Standing Wave Ratio Definition and Formula - Analog Devices
    Nov 15, 2012 · VSWR is related to the reflection coefficient. A higher ratio depicts a larger mismatch, while 1:1 ratio is perfectly matched. This match or ...
  43. [43]
    Voltage standing wave ratio (VSWR) - Microwaves101
    The reflection coefficient, commonly denoted by the Greek letter gamma (Γ), can be calculated from the values of the complex load impedance and the transmission ...
  44. [44]
    5 Things to Consider when Choosing an Antenna - MobileMark
    Rating 5.0 (12) Nov 3, 2020 · A VSWR value under 2 is considered suitable for most antenna applications. A VSWR specification commonly adopted is a 2:1 VSWR, which means that ...
  45. [45]
    [PDF] Transmission Lines - Sandiego
    Thus, all the power must be reflected back, giving net power flow equal to zero ... (Actually the RMS value). This is called a partial standing wave. An ...
  46. [46]
    Energy Conservation
    Recall, from the previous section, that a standing wave is a superposition of two traveling waves, of equal amplitude and frequency, propagating in opposite ...Missing: continuous | Show results with:continuous
  47. [47]
    On the directionality of membrane coupled Helmholtz resonators ...
    Nov 13, 2024 · A Helmholtz resonator (HR), first proposed by Hermann von Helmholtz in 1862, consists of a rigid-walled cavity connected to a neck through which ...
  48. [48]
    The behavior of standing waves near the end of an open pipe with ...
    May 16, 2023 · In the simplest approximation, the open end of a pipe is considered to be the location of a node of an acoustic standing wave inside the pipe.
  49. [49]
    [PDF] Acoustics of Organ Pipes and Future Trends in the Research
    (A so-called standing wave occurs in a pipe when the sound waves reflected back and forth in the pipe are com- bined such that each location along the pipe axis ...
  50. [50]
    [PDF] Music in Terms of Science - arXiv
    Sep 17, 2012 · MUSICAL INSTRUMENTS. 6.1 Chordophones—string instruments. Chordophones are musical instruments in which standing waves that produce sound are ...
  51. [51]
    Exploration of Resonant Modes for Circular and Polygonal Chladni ...
    Each Chladni plate, depending on the material it is made of, as well as its size and shape, will have resonant modes, i.e., frequencies at which standing waves ...
  52. [52]
    [PDF] Mathematical Model of a Chladni Plate - BYU Engineering
    When driven by vibrating source, a Chladni plate develops 2D standing wave patterns that can be visualized when salt is poured onto the plate. In this paper, a ...
  53. [53]
    Earth's free oscillations excited by the 2011 Tohoku earthquake ...
    May 24, 2021 · A seismic normal mode does constitute such a signal—a harmonic standing wave phase-synchronous in time, having a spatial pattern of a surface ...
  54. [54]
    Analysis of Fault Zone Resonance Modes Recorded by a Dense ...
    Sep 24, 2020 · We present observations and modeling of spatial eigen-functions of resonating waves within fault zone waveguide, using data recorded on a dense seismic array.
  55. [55]
    Patterns of Faraday waves | Journal of Fluid Mechanics
    Dec 1, 2003 · Faraday waves are standing waves which arise through a parametric instability on the surface of a vertically oscillated fluid layer.
  56. [56]
    Parametric autoresonance in Faraday waves | Phys. Rev. E
    Jul 22, 2005 · We develop a theory of parametric excitation of weakly nonlinear standing gravity waves in a tank, which is under vertical vibrations with a slowly time- ...Article Text · INTRODUCTION · WEAKLY NONLINEAR... · DISCUSSION
  57. [57]
    Natural Frequencies of seiches in Lake Chapala - PMC
    Aug 14, 2019 · Theoretical calculation of seiche periods. The estimates of lake oscillation modes obtained from Merian's formula (Eq. 11) and approximation ...Missing: standing | Show results with:standing
  58. [58]
    Seiche modes in multi‐armed lakes - Brenner - 2018 - ASLO - Wiley
    Sep 4, 2018 · The effect of multiple arms on standing wave patterns, or seiches, present within a lake is not easy to predict. This study examines free seiche ...
  59. [59]
    27.7 Thin Film Interference – College Physics - UCF Pressbooks
    This interference is between light reflected from different surfaces of a thin film; thus, the effect is known as thin film interference.
  60. [60]
    [PDF] ATOMIC AND OPTICAL PHYSICS - Caltech PMA
    electromagnetic wave must be an integer number times the cavity's free spectral range ⌬␯fsr⬅c/2L, where L is the length of the cavity and c is the speed of ...
  61. [61]
    [PDF] Realization of a monolithic high-reflectivity cavity mirror from a single ...
    (FSR) νFSR and cavity linewidth ∆ν (full width at half maxi- mum). The FSR was νFSR = c/(2L) = (6.246 ± 0.13) GHz with c the speed of light and L = (24 ...
  62. [62]
    Theory and Applications of X-ray Standing Waves in Real Crystals
    Dec 4, 2000 · Theoretical aspects of x-ray standing wave method for investigation of the real structure of crystals are considered in this review paper.Missing: diffraction equations seminal
  63. [63]
    [PDF] The diffraction of X-rays by crystals - Nobel Lecture, September 6 ...
    von Laue had made some of his earliest experiments with a crystal of zinc sulphide, and had obtained results which proved that the diffracted pen- cils showed ...
  64. [64]
    [PDF] Waveguides
    However, TE and TM modes with non-zero cutoff frequencies do exist and place an upper limit on the usable bandwidth of the TEM mode. Similarly, in optical ...
  65. [65]
    [PDF] Part 5: Electromagnetic cavities and waveguides
    1897 Lord Rayleigh showed that two classes of waves are possible, "transverse electric" (TE) and "transverse magnetic" (TM). For each class, there is a minimum.
  66. [66]
    Standing-wave Resonators - RP Photonics
    The resonator modes of a linear resonator are always exactly perpendicular to the end mirrors; if such a resonator is somewhat misaligned, the modes adjust ...
  67. [67]
  68. [68]
    Three-Dimensional Topological Photonic Crystals (Invited Review)
    Dec 27, 2024 · In this review, concentrating on the novel boundary states unique in 3D systems, we provide a brief survey of the 3D topological photonic ...<|control11|><|separator|>
  69. [69]
    [PDF] Photonic meta-crystals
    Jul 8, 2025 · This review introduces photonic meta-crystals in sequence, based on the classification of electromagnetic parameters in metamaterials. Recent ...
  70. [70]
    [PDF] XXXV. A Tentative Theory of Light Quanta. By LOUIS DE BROGLIE
    I shall in the present paper assume the real existence of light quanta, and try to see how it would be possible to reconcile with it the strong experimental ...
  71. [71]
    [PDF] 3. Quantisation as an eigenvalue problem; by E. Schrödinger
    With this variation problem we substitute the quantum conditions. First of all, we will take for H the Hamiltonian function of the Keplerian motion and show ...
  72. [72]
    [PDF] An undulatory theory of the mechanics of atoms and molecules - ISY
    The paper gives an account of the author's work on a new form of quantum theory. §1. The Hamiltonian analogy between mechanics and optics. §2. The.
  73. [73]
    Observation of discrete electronic states in a zero-dimensional ...
    Feb 8, 1988 · The study observed discrete electronic states in a zero-dimensional system, a quantum dot, through resonant tunneling.