Fact-checked by Grok 2 weeks ago

Cnoidal wave

A cnoidal wave is a type of periodic, nonlinear wave that propagates in shallow water without changing form, characterized by long, flat troughs and narrow, sharp crests, and mathematically described using the Jacobi elliptic function cn (hence the name "cnoidal"). These waves represent an exact solution to the Korteweg–de Vries (KdV) equation, which models weakly nonlinear, dispersive waves in shallow water depths where the Ursell number (a dimensionless parameter measuring nonlinearity relative to dispersion) exceeds approximately 25–40, making them suitable for intermediate to shallow water conditions with finite amplitudes up to about half the water depth. Cnoidal wave theory originated in 1895 with the work of Diederik Korteweg and Gustav de Vries, who derived the KdV equation while studying solitary waves and periodic undular bores, building on earlier approximations by Joseph Boussinesq (1871) and Lord Rayleigh (1876) for long waves of finite amplitude. Subsequent refinements included systematic derivations by Reinhard Friedrichs and Joseph Keller in 1948, second-order extensions by Edward Laitone in 1960, and third-order formulations by James Chappelear in 1962, with higher-order solutions up to 24th order developed by researchers like Nishimura et al. in 1977 and practical fifth-order theories by John Fenton in 1979 and 1990. Key properties of cnoidal waves include their steady propagation speed, which depends on wave height, period, and water depth, and their transition to solitary waves as the elliptic modulus approaches 1 (yielding a single crest with infinite period) or to linear sinusoidal waves as the modulus approaches 0 (infinitesimal amplitude). The theory is particularly applicable to natural phenomena like tidal bores, river undulations, and nearshore wave fields, providing more accurate predictions than linear or Stokes wave theories in regimes where nonlinearity dominates dispersion, though it assumes irrotational, inviscid flow and neglects higher-order effects like wave breaking. Modern applications extend to coastal engineering for radiation stress calculations and numerical modeling of wave evolution.

Introduction

Definition and Context

Cnoidal waves represent exact periodic solutions to nonlinear wave equations arising in shallow water theory, where the surface elevation profile is described using the Jacobi elliptic cosine function, denoted as "cn," in contrast to the sinusoidal forms typical of linear wave theory. These waves emerge as stable, non-breaking periodic disturbances when nonlinear effects, such as wave steepening, are balanced by dispersive effects in depths much smaller than the . This balance, quantified by the Ursell number Ur = a λ² / h³ (where a is wave , λ , and h depth) exceeding approximately 25–40, allows for the propagation of wave trains with characteristic sharp crests and extended flat troughs, observed in real shallow- environments like coastal zones or channels. The mathematical form of a cnoidal wave's surface elevation is commonly expressed as \eta(x,t) = \eta_0 + A \, \mathrm{cn}^2 \left( \frac{K (x - c t)}{\lambda / 2}, m \right), where \eta_0 is the mean , A is the wave amplitude, c is the phase speed, \lambda is the , K is the complete of the first kind, and m (with $0 < m < 1) is the elliptic modulus that governs the wave's shape—yielding sharper crests as m approaches 1. This representation highlights the periodic nature of the solution, with the period tied to the elliptic function's properties. In distinction from other wave types, cnoidal waves apply specifically to shallow-water regimes and differ from Stokes waves, which model nonlinear periodic waves in deeper water using Fourier series expansions, or from solitary waves, which are localized pulses of infinite extent without periodicity. The crest sharpness in cnoidal waves, controlled by m, provides a tunable parameter for their morphology, enabling descriptions of a range of intermediate wave behaviors between linear and fully solitary limits. These solutions primarily arise from the , a foundational model for unidirectional shallow-water wave propagation.

Historical Development

The concept of cnoidal waves originated in the late 19th century through the work of Dutch mathematicians and his student , who sought to model the propagation of long waves in shallow water channels. Motivated by observations of tidal bores and the solitary waves reported by engineer in the 1830s and 1840s, Korteweg and de Vries derived periodic wave solutions as part of their investigation into nonlinear dispersive effects in rectangular canals. In their seminal 1895 paper, Korteweg and de Vries presented these periodic solutions to what is now known as the , expressing the wave profiles in terms of . The term "cnoidal wave" was coined by them, deriving from the (denoted cn), which captures the nonlinear periodic structure of the waves in shallow water. Subsequent developments in the mid-20th century built on this foundation, with British mathematician Tony Benjamin, American mathematician Jerry Bona, and Australian mathematician John Mahony addressing limitations in the original model's dispersion for long waves. In their 1972 paper, they proposed the Benjamin–Bona–Mahony equation as an alternative framework, better suited for simulating phenomena like undular bores in canals. Throughout the 20th century, cnoidal waves gained recognition as a cornerstone of nonlinear wave theory, bridging periodic structures and solitary waves, and profoundly influencing the development of soliton theory following numerical rediscoveries of the in the 1960s.

Governing Equations

Korteweg–de Vries Equation

The Korteweg–de Vries (KdV) equation was originally derived in 1895 by and to provide a theoretical explanation for the solitary waves observed by in 1834, though the equation also supports periodic wave solutions. Their work built on earlier approximations by , focusing on waves in a rectangular canal with rectangular cross-section. The KdV equation arises as an asymptotic model for unidirectional nonlinear shallow-water waves under specific assumptions: inviscid, irrotational flow; unidirectional propagation in the positive x-direction; small amplitude-to-depth ratio ε = a/h ≪ 1, where a is the wave amplitude and h is the undisturbed water depth; and long wavelength relative to depth, kh ≪ 1, where k is the wavenumber. These conditions ensure a balance between weak nonlinearity, which causes wave steepening, and weak dispersion, which promotes wave spreading. The derivation begins from the Euler equations for inviscid, incompressible, irrotational flow: the continuity equation ∇ · \vec{u} = 0, the momentum equations \partial_t \vec{u} + (\vec{u} · ∇) \vec{u} = -∇(p/ρ + gz), and kinematic/dynamic boundary conditions at the free surface z = h + η(x,t), where η is the surface elevation, alongside a rigid bottom at z = -h. Using the shallow-water approximation (horizontal scales much larger than depth) and a perturbation expansion in small parameters ε and δ = h/λ (with ε ~ δ² for balance), the velocity potential φ is expanded as φ = φ_0 + ε φ_1 + ..., leading to integrated forms for horizontal velocity u and surface elevation η. At leading order, the equations reduce to the linear shallow-water wave equation; higher-order terms introduce nonlinearity from the advective acceleration and dispersion from vertical structure in the velocity profile. The resulting dimensional KdV equation for the horizontal velocity u is u_t + u u_x + \alpha h^2 u_{xxx} = 0, where α = 1/3 for surface water waves under gravity. Equivalently, for the surface elevation η (with u ≈ √(gh) · (η/h) in the shallow-water limit), the form is η_t + c η_x + (3c/(2h)) η η_x + (c h^2 / 6) η_{xxx} = 0, where c = √(gh) is the linear long-wave speed. A common nondimensional version normalizes η by h, x by h, and t by h/√(gh), yielding \eta_t + \left(\eta + \frac{1}{2} \eta^2\right)_x + \frac{1}{6} \eta_{xxx} = 0, which expands to η_t + η_x + η η_x + (1/6) η_{xxx} = 0 and highlights the balance between linear advection, nonlinearity, and dispersion. The KdV equation possesses remarkable mathematical properties, including complete integrability via the inverse scattering transform, first demonstrated in 1967, which linearizes the nonlinear evolution through a scattering problem analogous to quantum mechanics. It admits an infinite number of conserved quantities, such as mass, momentum, energy, and higher-order integrals, ensuring stability and reversibility in its dynamics. These features arise from its bi-Hamiltonian structure and association with the Lax pair formalism, established in 1968, confirming its solubility for initial-value problems. Cnoidal waves emerge as exact periodic solutions to this equation.

Benjamin–Bona–Mahony Equation

The Benjamin–Bona–Mahony equation provides a refined mathematical model for the propagation of nonlinear dispersive waves, particularly addressing shortcomings in earlier approximations like the by incorporating a regularization term for improved stability. Its general formulation is u_t + u_x + u u_x - \delta u_{xxt} = 0, where u(x,t) denotes the wave elevation or velocity perturbation, and the term -\delta u_{xxt} regularizes the equation against ill-posedness for short waves by introducing a mixed space-time derivative. This equation was introduced in 1972 by T. B. Benjamin, J. L. Bona, and J. J. Mahony as an alternative model derived from the full governing inviscid, incompressible fluid flow. The derivation employs a systematic approximation for weakly nonlinear, long waves, replacing the purely spatial dispersion in prior models with the mixed derivative to mitigate high-frequency instabilities that render simulations of the prone to numerical blow-up. Compared to the Korteweg–de Vries equation, the Benjamin–Bona–Mahony equation offers superior performance for weakly nonlinear regimes, as its bounded dispersion relation enhances physical accuracy across a broader range of wavenumbers and prevents unphysical wave breaking in computational studies. Like the , it admits exact periodic solutions that generalize to cnoidal-like waves in shallow water, though they are more complex to express analytically. In the context of nondimensionalized surface water waves under shallow-water assumptions, the parameter is set to \delta = 1/3. The equation supports solitary wave solutions traveling at speed c = \sqrt{1 + A}, where A is the maximum amplitude (approximately c \approx 1 + A/2 for small A), maintaining shape without dispersion in this limit. Beyond surface waves, the Benjamin–Bona–Mahony equation finds applications in modeling internal waves in stratified fluids and as a continuum limit for integrable lattice equations, with integrability preserved in specific asymptotic regimes.

Wave Solutions

Surface Elevation Profile

The surface elevation of a cnoidal wave is described by the explicit mathematical expression \eta(x,t) = A \, \mathrm{cn}^2 \bigl( \beta (x - c t); m \bigr) + \mathrm{constant}, where \mathrm{cn}(u; m) is the Jacobi elliptic cosine function with modulus m (0 < m < 1), A is the wave amplitude, \beta = 2 K(m) / \lambda is the scaling parameter with K(m) the complete elliptic integral of the first kind and \lambda the wavelength, and c is the phase speed. This profile is symmetric about the crests and periodic in space, exhibiting sharper crests and progressively flatter troughs as the modulus m approaches 1, reflecting increasing nonlinearity in shallow water conditions. The periodicity arises from the properties of the elliptic function, with the spatial period \lambda tied to the argument such that the waveform repeats every \lambda in x for fixed t. The temporal period is given by T = 2 K(m) / (\beta c), corresponding to the interval for one full cycle in the traveling frame. Visually, the profile transitions smoothly from a near-sinusoidal oscillation when m \approx 0, resembling small-amplitude linear waves, to a tabular shape with extended flat regions in the troughs and steep rises to the crests when m \approx 1, capturing the essence of intermediate to highly nonlinear periodic waves in shallow water. The underlying Jacobi elliptic function \mathrm{cn}(u; m) is periodic with fundamental period $4 K(m) and satisfies the nonlinear differential equation \left( \frac{d \, \mathrm{cn} \, u}{du} \right)^2 = (1 - \mathrm{cn}^2 u) (1 - m \, \mathrm{cn}^2 u), which defines the oscillatory behavior and ensures the waveform's consistency with the governing dynamics of the . Consequently, \mathrm{cn}^2(u; m) inherits a period of $2 K(m), directly governing the repetition of the surface elevation profile.

Phase Speed and Periodicity

The phase speed c of a cnoidal wave, derived from the Korteweg–de Vries (KdV) equation, balances nonlinear steepening and dispersive spreading, and is given by c = \sqrt{gh} \left( 1 + \frac{A}{2h} - \frac{(m K'/K)^2}{3} \right), where g is gravitational acceleration, h is water depth, A is wave amplitude, m is the elliptic modulus ($0 < m < 1), K = K(m) is the complete elliptic integral of the first kind, and K' = K(1-m). This expression captures the enhancement from nonlinearity via the A/(2h) term and the reduction from dispersion via the elliptic ratio term, with the linear shallow-water speed \sqrt{gh} as the base. As the modulus m increases toward 1, the periods lengthen, transitioning toward the solitary wave limit where the wave train becomes infinitely extended. Unlike linear dispersive waves, where phase speed depends solely on wavenumber via c = \sqrt{gh} \tanh(kh) \approx \sqrt{gh} (1 - (kh)^2/3) for shallow water, the cnoidal phase speed incorporates nonlinearity, varying with both amplitude A and modulus m, which parameterizes the wave's sharpness and periodicity. The group velocity c_g for cnoidal waves follows from the dispersion relation as c_g = c - (\lambda / 2\pi) \, dc/dk, where k = 2\pi / \lambda is the fundamental wavenumber; alternatively, it can be obtained from the conserved quantities of the KdV equation, such as momentum and energy. Due to the integrability of the KdV equation, periodic cnoidal wave solutions are orbitally stable under small perturbations, with spectral stability confirmed for all moduli m in the periodic domain.

Limiting Cases

Solitary Wave Limit

As the elliptic modulus m approaches unity from below, the periodic cnoidal wave solution of the for water waves transitions to the solitary wave, or , solution. In this limit, the Jacobi elliptic cosine function \mathrm{cn}(u; m) approaches the hyperbolic secant function \mathrm{sech}(u), after appropriate scaling of the argument u. This yields the surface elevation profile \eta(x, t) = A \, \mathrm{sech}^2 \left[ \sqrt{\frac{3A}{4 h^3}} (x - c t) \right], where A is the wave amplitude, h is the undisturbed water depth, and c is the phase speed, representing the canonical for shallow-water waves. The phase speed in this solitary wave limit is c = \sqrt{g h} \left(1 + \frac{A}{2 h}\right), where g is gravitational acceleration, which exceeds the linear long-wave speed \sqrt{g h} and increases with amplitude, enabling overtaking of smaller waves by taller ones without distortion. Physically, the solitary wave limit corresponds to the wavelength \lambda or period tending to infinity as m \to 1, causing the wave to localize into a single, non-radiating pulse that maintains its form indefinitely, thus accounting for the emergence of stable solitary waves from evolving periodic trains under nonlinear dispersion. Compared to finite-period cnoidal waves, the solitary wave preserves essential nonlinear and dispersive balances—such as the \mathrm{sech}^2 profile shape—but eliminates spatial repetition, concentrating all wave energy into one isolated structure. Historically, derived cnoidal waves to theoretically explain John Scott Russell's observed "wave of translation," with the solitary wave emerging as the key limiting case of their periodic solutions in the infinite-wavelength regime.

Infinitesimal Amplitude Limit

In the infinitesimal amplitude limit, where the wave amplitude A approaches zero, the nonlinear parameter m of the cnoidal wave solution also tends to zero, causing the squared Jacobi elliptic cosine function \mathrm{cn}^2 to approximate \cos^2. This reduction yields a surface elevation profile of the form \eta \approx A \cos(k x - \omega t), representing a simple harmonic sinusoidal wave devoid of nonlinear distortions. At this limit, dispersion dominates the wave behavior without wave steepening, as nonlinear effects become negligible, aligning the cnoidal solution with the foundational linear theory for small-amplitude surface gravity waves. The phase speed in this regime converges to the linear dispersion relation c = \sqrt{\frac{g \tanh(k h)}{k}}, where g is gravitational acceleration, k is the wavenumber, and h is water depth. For shallow water conditions where k h \ll 1, this approximates c \approx \sqrt{g h} \left(1 - \frac{(k h)^2}{6}\right), capturing the leading dispersive correction to the non-dispersive shallow-water speed \sqrt{g h} while matching the infinitesimal limit of linear shallow-water theory. The wave period remains fixed by the linear relation \lambda = 2\pi / k, independent of amplitude, emphasizing the absence of amplitude-dependent nonlinearity. This transition marks the boundary where cnoidal waves effectively approximate the Airy linear wave solutions in finite depth or the initial terms of Stokes expansions in shallower regimes, serving as the baseline for weakly nonlinear extensions.

Derivations

From Approximate Equations

To derive cnoidal wave solutions from the , assume a traveling wave form u(x, t) = f(\xi), where \xi = x - c t and c > 0 is the wave speed. Substituting into the KdV equation u_t + 6 u u_x + u_{xxx} = 0 yields the ordinary differential equation (ODE) f''' + 6 f f' - c f' = 0, where primes denote derivatives with respect to \xi. Integrating once with respect to \xi gives f'' + 3 f^2 - c f = A, where A is the integration constant. To proceed, multiply this equation by f' and integrate again: \frac{1}{2} (f')^2 + V(f) = E, where V(f) = f^3 - \frac{c}{2} f^2 - A f is a cubic potential function, and E is another integration constant. For periodic solutions, choose A < 0 such that V(f) has three real roots f_1 < f_2 < f_3, with the solution oscillating between f_2 and f_3. Shifting variables to center at f_2 reduces the equation to the standard elliptic form (f - f_2)' = \sqrt{2 (E - V(f))}, which integrates to the cnoidal solution expressed in Jacobi elliptic functions: f(\xi) = f_2 + (f_3 - f_2) \operatorname{cn}^2 \left( \sqrt{(f_3 - f_1)(f_3 - f_2)} \, \xi / \sqrt{6} ; m \right), with elliptic modulus m = (f_3 - f_2)/(f_3 - f_1) \in (0, 1), and roots related to parameters via f_3 + f_2 + f_1 = c/3, f_3 f_2 + f_3 f_1 + f_2 f_1 = -A/3. The wave speed c and amplitude f_3 - f_2 are thus linked through m. For the Benjamin–Bona–Mahony (BBM) equation u_t + u_x + u u_x = u_{xxt}, the traveling wave assumption u(x, t) = f(\xi) with \xi = x - c t leads to a similar but nonlocal reduction: (1 - \partial_\xi^2) f' + f' + f f' - c f' = 0. Integrating yields an integro-differential equation, but periodic solutions can be constructed analogously to the KdV case by summing solitary waves via the Poisson summation formula, resulting in cnoidal forms f(\xi) = \beta_2 + (\beta_3 - \beta_2) \operatorname{cn}^2 \left( \sqrt{\frac{\beta_3 - \beta_1}{12 c}} \, \xi ; k \right), where the roots \beta_i depend on speed c > 1 and L, with k related to an adjusted coefficient that regularizes short waves compared to KdV. The solutions share the periodic structure but exhibit modified due to the nonlocal term. In the shallow-water context, the physical wave number k (spatial frequency) relates the parameters via k = \frac{\pi m}{K(m)} \sqrt{\frac{3 g}{A h}}, where g is gravitational acceleration, A is wave amplitude, h is water depth, and K(m) is the complete elliptic integral of the first kind; this ensures the solution's periodicity matches the physical wavelength. The cnoidal solutions form a one-parameter family indexed by the modulus m \in (0, 1), interpolating continuously between the infinitesimal-amplitude linear wave limit (m \to 0) and the solitary wave limit (m \to 1), providing a complete characterization of periodic traveling waves for these equations.

From Full Inviscid Equations

The derivation of cnoidal waves from the full inviscid Euler equations begins with the assumption of irrotational and beneath a , governed by a \phi(x, z, t) that satisfies \nabla^2 \phi = 0 in the fluid domain -h < z < \eta(x, t), where h is the undisturbed water depth and \eta(x, t) is the surface elevation. At the rigid bottom boundary z = -h, the vertical velocity vanishes, so \partial \phi / \partial z = 0. The nonlinear kinematic boundary condition at the free surface z = \eta enforces that fluid particles on the surface remain there: \partial \eta / \partial t + (\partial \phi / \partial x)(\partial \eta / \partial x) = \partial \phi / \partial z. The dynamic boundary condition, derived from the unsteady Bernoulli equation assuming constant atmospheric pressure, states that the pressure at the surface is zero: \partial \phi / \partial t + \frac{1}{2} |\nabla \phi|^2 + g \eta = 0 at z = \eta. For shallow-water conditions, where the wavenumber k satisfies kh \ll 1, these equations are expanded in a perturbation series in powers of kh, incorporating higher-order dispersive and nonlinear terms while satisfying the boundary conditions approximately at z = -h or via Taylor expansions at the mean surface level. Periodic solutions approximating cnoidal waves can be obtained through series expansions involving elliptic integrals or Fourier series, providing high-order approximations (e.g., up to fifth order) in the shallow-water regime. Early work by Joseph Boussinesq in 1872 used elliptic functions to approximate finite-amplitude long waves, influencing later developments. The resulting surface profile emerges in an implicit form, expressed via Jacobi elliptic functions such as \eta(x) \propto \mathrm{cn}^2(\theta | m), where m is the elliptic modulus related to wave steepness; this form recovers the cnoidal profile and converges to long-wave approximations in the limit of small amplitude or long period.

Physical Properties

Potential Energy

The potential energy of a cnoidal wave arises primarily from the gravitational displacement of the water surface and is calculated as the average over one wavelength of \frac{1}{2} \rho g \eta^2, where \rho is the water density, g is gravitational acceleration, and \eta(x) is the surface elevation profile. For a cnoidal wave of amplitude A and wavelength \lambda, the average density exceeds the linear value due to the nonlinear surface profile, which features sharper crests and flatter troughs compared to linear waves. In the linear limit as m \to 0, the potential energy reduces to the familiar \frac{1}{4} \rho g A^2 per unit horizontal area (or \frac{1}{4} \rho g A^2 \lambda per wavelength per unit width), where the total energy is equally partitioned between potential and kinetic components. For finite m > 0, the potential energy is higher due to the elevated crests concentrating more mass above the mean water level, while the energy density peaks at these crests, contributing to the wave's and influencing interactions with other waves or boundaries. In the KdV approximation, the total conserved energy (combining potential and kinetic contributions) per unit width is E = \int_0^\lambda \left( \frac{\eta^2}{2} - \frac{\eta^3}{6} + \frac{1}{6} \eta_x^2 \right) dx, which remains under and highlights the balance between nonlinear steepening and . (Note: this is in nondimensional form with appropriate for depth h=1 and g incorporated.) As m \to 1, the cnoidal wave approaches the solitary wave limit, where the total potential energy per unit width is E = \frac{4}{3} \rho g A^2 \sqrt{\frac{h^3}{3 g A}}, with h the undisturbed depth, reflecting the localized energy of the isolated . The total energy is approximately twice the in this approximation.

Effects of Surface Tension

is incorporated into the theory of cnoidal waves through the dynamic boundary condition at the , where an additional term σ κ accounts for the restoring force due to , with σ denoting the surface tension coefficient and κ ≈ -∂²η/∂x² the linearized of the surface elevation η. This modification extends the standard gravitational framework, leading to a capillary-gravity for linear waves: ω² = g k tanh(k h) + (σ/ρ) k³ tanh(k h), where g is , k is , h is depth, and ρ is . In the context of cnoidal waves, this alters the balance between nonlinearity and dispersion, particularly for finite-amplitude periodic solutions derived from the full inviscid equations or approximate models. For cnoidal waves, primarily impacts small-scale waves by stabilizing short wavelengths that would otherwise be unstable under gravity alone, thereby modifying the elliptic m required to achieve a given wave . Generalizations of the Korteweg-de Vries (KdV) equation incorporate a parameter, such as in extended Boussinesq or higher-order KdV models, which yield cnoidal-like solutions with adjusted phase speeds and profiles. These effects reduce the sharpness of wave crests compared to pure gravity cnoidal waves, as the capillary forces smooth out high-curvature regions, analogous to the widening observed in solitary wave limits. Surface tension becomes dominant in shallow-water capillary-gravity regimes, particularly for wavelengths λ on the order of millimeters (around the capillary length l_c ≈ 2.7 mm for water, where l_c = √(σ/(ρ g))), shifting the phase speed to c ≈ √{[(g + (σ/ρ) k²) tanh(k h)] / k}. In this scale, the dispersion relation emphasizes the k³ term, enabling stable periodic waves that transition from gravity-dominated to tension-dominated behavior. Solutions take extended cnoidal forms, often obtained via perturbation expansions around elliptic functions, which account for the interplay of gravity and tension in the nonlinear regime. However, surface tension effects are negligible for large-scale oceanic cnoidal waves, where the (σ/ρ) k³ term is small relative to g k due to long wavelengths (λ ≫ mm), rendering the standard gravity-only cnoidal theory sufficient. In contrast, tension plays a crucial role in laboratory experiments with controlled small-scale flows or microscale applications, such as in microfluidic devices, where it influences wave stability and energy dissipation.

Applications

Example in Shallow Water

A concrete numerical example of a cnoidal wave in shallow water can be considered for a channel with mean depth h = 1 m and wave amplitude A = 0.5 m, employing the modulus parameter m = 0.9. In this scenario, the phase speed c is about $1.25 \sqrt{g h}, where g denotes gravitational acceleration. The wavelength \lambda and period T can be determined from the theory's equations, typically yielding values consistent with shallow water conditions (e.g., T \approx \lambda / c). The wave profile features a maximum surface elevation \eta_{\max} = 0.5 m and a shallow trough below the mean water level, expressed in terms of the squared Jacobi elliptic cosine function \mathrm{cn}^2 with the complete elliptic integral of the first kind K(0.9) \approx 2.25. The horizontal velocity component u is approximated as u \approx \sqrt{g h} (A/h) \mathrm{cn}^2, while the vertical velocity remains small due to the shallow water regime. This parameter set simulates phenomena such as tidal bores or ship wakes in canals, where the wave maintains stability over long propagation distances. To evaluate the required elliptic functions numerically, methods such as expansions or modern computational libraries (e.g., those implementing arithmetic-geometric mean iterations) are employed.

Broader Implications

Cnoidal waves play a crucial role in modeling undular bores, which are oscillatory hydraulic jumps observed in shallow flows, providing insights into dissipation during bore propagation. In tsunami dynamics, these waves describe the upstream propagation of undular bores, aiding predictions of inundation patterns through extensions of the Korteweg– (KdV) equation that incorporate dissipation. For , cnoidal theory supports simulations of nonlinear wave propagation in nearshore regions using fully nonlinear Boussinesq equations, enabling accurate forecasting of wave transformation and run-up on beaches. Beyond water waves, cnoidal solutions appear as analogs in diverse nonlinear systems. In nonlinear optics, they model periodic structures arising from modulational instability in the nonlinear Schrödinger equation, describing pulse trains in optical fibers. In Bose-Einstein condensates, cnoidal waves represent density modulations in one-dimensional ultracold atomic gases, linking to phases with both superfluid and crystalline order. Recent studies have extended cnoidal wave theory to oblique propagation, deriving solutions for waves at angles to the flow direction in both and contexts, which enhances modeling of directional wave spreading. Interactions with background currents alter wave steepness and energy transfer, with opposing currents shortening wavelengths and amplifying heights in shallow simulations. In multi-layer fluids, internal cnoidal waves emerge in two-layer configurations, capturing rotational effects and defect-induced bursts relevant to oceanic stratification (as of 2024). Viscous effects introduce over boundaries, where energy dissipation in boundary layers reduces wave amplitude more significantly than in inviscid cases. In engineering applications, guide the design of wavemakers in flumes, where or flap mechanisms generate periodic trains for testing structural loads and wave breaking. These models also aid in predicting wave train evolution within harbors, informing layouts to mitigate and overtopping risks from incoming nonlinear . Theoretically, cnoidal waves serve as a bridge to integrable systems like the KdV equation, where they connect to N-soliton solutions through Bäcklund transformations and stability analyses, revealing interactions such as breathers formed by solitary wave perturbations on periodic backgrounds.

References

  1. [1]
    [PDF] The Cnoidal Theory of Water Waves - John Fenton Homepage
    Initially a history of cnoidal theory is given, and various contributions are described and reviewed.
  2. [2]
    Evolution of basic equations for nearshore wave field - PMC - NIH
    For very shallow water, cnoidal wave theory was developed by Korteweg and de Vries (1895), who derived an equation for shallow water waves called the KdV ...Missing: history | Show results with:history
  3. [3]
    [PDF] D.J. Korteweg and G. de Vries, On the change of form of long waves ...
    de Vries,. On the change of form of long waves advancing in a rectangular canal, and on a new type of long stationary waves,. Phil. Mag. (5) 39 (1895), 422–443.
  4. [4]
    XLI. On the change of form of long waves advancing in a rectangular ...
    (1895). XLI. On the change of form of long waves advancing in a rectangular canal, and on a new type of long stationary waves . The London, Edinburgh, and ...
  5. [5]
    [PDF] The Korteweg-de Vries equation: a historical essay
    The 1895 paper of Korteweg. & de Vries was excerpted and translated from de Vries's 1894 thesis. Korteweg & de Vries, unlike Boussinesq, emphasize at the outset ...
  6. [6]
    [PDF] some details of the history of the korteweg-de vries equation
    The paper is entitled: On the change of form of long waves advancing in a rectangular canal, and on a new type of long station- ary waves. It was published in ...
  7. [7]
    Model equations for long waves in nonlinear dispersive systems
    Several topics are studied concerning mathematical models for the unidirectional propagation of long waves in systems that manifest nonlinear and dispersive ...
  8. [8]
    [PDF] On the Origin of the Korteweg-de Vries Equation - arXiv
    Dec 8, 2011 · Abstract. The Korteweg-de Vries equation has a central place in a model for waves on shallow water and it is an example of the propagation ...
  9. [9]
    History and Origins of the Korteweg-de Vries Equation - SIAM.org
    Historically, however, the origins of the KdV equation are to be found much earlier, starting with the experiments of Scott Russell in 1834 [21] and the ...
  10. [10]
    [PDF] Derivation of the KdV equation for surface and internal waves
    We sketch here a different derivation of the Korteweg–de Vries equation, applicable to a wider range of problems, including internal waves and waves in ...
  11. [11]
    Benjamin-Bona-Mahony Equation - an overview - ScienceDirect.com
    To this end, the regularized long-wave equation (RLW) or Benjamin–Bona–Mahony equation (BBM) [1] was derived as a model to describe gravity water waves. It is ...
  12. [12]
    [PDF] Lecture 18: Wave-Mean Flow Interaction, Part I - WHOI GFD
    This cnoidal wave (88) contains two free parameters; we take these to be the amplitude a and the wavenumber k. We now use the cnoidal wave expression (88) to ...
  13. [13]
    [PDF] The orbital stability of the cnoidal waves of the Korteweg-de Vries ...
    Apr 8, 2010 · The cnoidal waves are the simplest of the periodic solutions of the KdV equation, represent the most general stationary (in a frame translating ...
  14. [14]
    [PDF] Korteweg-de Vries Equation - Aldebaran.cz
    Jan 11, 2008 · Korteweg and de Vries, published in 1895, in which they showed ... described by elliptic functions and commonly called cnoidal waves,. A ...
  15. [15]
    [PDF] Water Waves - MIT
    Dispersion relationship (6) uniquely relates ω and k, given h: ω = ω(k;h) or ... ( long waves or shallow water). 1 for kh >∼ 3, i.e. kh > π → h > λ. 2.
  16. [16]
    [PDF] Lecture 5: Waves in shallow water, part I: the theory - WHOI GFD
    form by introducing the Jacobi elliptical functions or cnoidal functions, which give the following set of solutions : ν(ξ,τ)=2p2cn2 p(ξ − 4p2τ + ξ0);κ + ν0. (37).<|control11|><|separator|>
  17. [17]
    None
    ### Summary of Cnoidal Wave Solutions for the Benjamin-Bona-Mahony (BBM) Equation
  18. [18]
    The Form of Standing Waves on the Surface of Running Water
    The Form of Standing Waves on the Surface of Running Water. Lord Rayleigh D.C.L., F.R.S., ... First published: November 1883. https://doi.org/10.1112/plms/s1 ...
  19. [19]
    [PDF] A high-order cnoidal wave theory
    Cnoidal wave solutions to fifth order are given so that a steady wave problem with known water depth, wave height and wave period or length may be solved to ...Missing: history | Show results with:history<|control11|><|separator|>
  20. [20]
    [PDF] Water Wave Mechanics for Engineers and Scientists
    ... cnoidal wave theories. The procedure for developing numerical wave theories to high order is described, as are the analytical and physical validities of ...
  21. [21]
    [PDF] Mathematical theory of irrotational translation waves
    Jacobian elliptic function, cosine amplitude. total energy of a wave per ... X wave length of a cnoidal wave (sections V-9, IS). p density of liquid ...
  22. [22]
    [PDF] A high-order cnoidal wave theory - John Fenton Homepage
    This equation, although accurate only to first order, can be used to examine three limits of interest, giving some insight into the nature of the cnoidal ...
  23. [23]
    [PDF] A Note on Solitary and Cnoidal Waves with Surface Tension. - DTIC
    If L =i0 relations (57), (52) (60) and (48) yield K - 1, y - h2 = 0 and h1 - 1. The cnoidal wave reduces then to the solitary wave considered in. Section 3. PFr ...Missing: phase | Show results with:phase
  24. [24]
    Asymptotic expansions for solitary gravity-capillary waves in two and ...
    Jun 24, 2009 · In equations (2.1) and (2.2), x=(x1,x2), k=(k1,k2), kx≡k⋅x and the constants g, ρ and σ denote gravity, density and surface tension, ...
  25. [25]
    (PDF) On Asymmetric generalized solitary gravity-capillary waves in ...
    The focus here is on the fifth order Korteweg–de Vries equation where three admissible one-parameter families of Whitham shocks are identified as solution ...
  26. [26]
    [PDF] Irrotational solitary waves with surface-tension
    solitary and cnoidal waves found by Korteweg and de Vries. For the exis ... It is found that the presence of the surface-tension widens the peak of the solitary ...
  27. [27]
    Contact Angle Measurement of Small Capillary Length Liquid in ...
    Apr 7, 2017 · The capillary length is defined as l cap = (γ /ρg)1/2, where γ is the liquid surface tension, ρ is the liquid density, and g is the ...
  28. [28]
    Cnoidal-type surface waves in deep water | Journal of Fluid Mechanics
    Jul 30, 2003 · Depending on the surface tension, the solitary wave involves one or two interfaces: a wave of depression; a wave of depression with a pocket of ...
  29. [29]
    On cnoidal waves and bores | Proceedings of the Royal ... - Journals
    The values of Q, R and S probably determine the wave-train uniquely; this is proved explicitly for long waves in a new presentation of the 'cnoidal wave' theory ...
  30. [30]
    On the undular bore generation for tsunami upstream propagation in ...
    In this study, a new theoretical model based on the Korteweg-de Vries (KdV)-Burgers type equation is presented for the generation of undular bore during the ...
  31. [31]
    [PDF] A fully nonlinear Boussinesq model for surface waves. Part 1. Highly ...
    Jan 28, 2016 · Fully nonlinear extensions of Boussinesq equations are derived to simulate surface wave propagation in coastal regions.
  32. [32]
    Nonlinear Stage of Modulational Instability in Repulsive Two ...
    Modulational instability (MI) is a fundamental phenomenon in the study of nonlinear dynamics, spanning diverse areas such as shallow water waves, optics, ...
  33. [33]
    Supersolidity of cnoidal waves in an ultracold Bose gas
    Feb 12, 2021 · A one-dimensional Bose-Einstein condensate may experience nonlinear periodic modulations known as cnoidal waves.<|separator|>
  34. [34]
    The Tantawy technique for modeling low-frequency ion-acoustic ...
    Oct 16, 2025 · This study investigates the propagation of both fractional and non-fractional ion-acoustic cnoidal waves (IACWs) in a magnetized degenerate ...
  35. [35]
    [PDF] CHAPTER 8 - Coastal Engineering Proceedings
    The computed wave angle and the Fourier coefficients for the input cnoidal wave were used to compute the surface displacement for oblique cnoidal waves.Missing: harbor | Show results with:harbor
  36. [36]
    Nonlinear Wave Evolution in Interaction with Currents and ... - MDPI
    Most of the energy in the cnoidal wave spectrum is confined in frequencies lower than 0.50 Hz where D m is consistently higher for opposing currents for ...
  37. [37]
    Evolution of internal cnoidal waves with local defects in a two-layer ...
    Evolution of internal cnoidal waves with local defects in a two-layer fluid with rotation.
  38. [38]
    The viscous damping of cnoidal waves | Journal of Fluid Mechanics
    Mar 29, 2006 · The viscous damping of cnoidal waves progressing over a smooth horizontal bed is investigated. First approximations are derived for the ...
  39. [39]
    Study of non-linear wave motions and wave forces on ship sections ...
    Citation Excerpt : In the laboratory experiments, cnoidal and sine wave trains were generated by the wavemaker, and four cases (two for cnoidal waves and two ...
  40. [40]
    [PDF] The development of a numerical model for the solution of the ...
    Accurate predictions of wave conditions within a harbour are an important factor in optimising its layout. Until recently the only reliable method available to ...
  41. [41]
    N-th Bäcklund transformation and soliton-cnoidal wave interaction ...
    The purpose of the paper is to investigate the residual symmetry, th BT, CRE solvability and soliton-cnoidal wave interaction solution of the combined KdV–nKdV ...