Fact-checked by Grok 2 weeks ago

Jacobi elliptic functions

Jacobi elliptic functions are a class of doubly periodic meromorphic functions in the complex plane, serving as inverses of elliptic integrals and generalizing the elementary trigonometric functions sine, cosine, and their hyperbolic counterparts. The three principal functions, denoted \operatorname{sn}(z,k), \operatorname{cn}(z,k), and \operatorname{dn}(z,k), depend on a complex argument z and a modulus parameter k with $0 \leq k \leq 1 for real-valued cases, where \operatorname{sn}(z,k) is defined such that z = \int_0^{\operatorname{sn}(z,k)} \frac{d t}{\sqrt{(1-t^2)(1-k^2 t^2)}}, \operatorname{cn}(z,k) = \sqrt{1 - \operatorname{sn}^2(z,k)}, and \operatorname{dn}(z,k) = \sqrt{1 - k^2 \operatorname{sn}^2(z,k)}. These functions satisfy fundamental identities analogous to Pythagorean relations, such as \operatorname{sn}^2(z,k) + \operatorname{cn}^2(z,k) = 1 and \operatorname{dn}^2(z,k) + k^2 \operatorname{sn}^2(z,k) = 1, and exhibit double periodicity with real period $4K(k) and imaginary period $4iK(k'), where K(k) is the complete elliptic integral of the first kind and k' = \sqrt{1 - k^2}. In the limiting cases, as k \to 0, they reduce to \operatorname{sn}(z,k) \to \sin z, \operatorname{cn}(z,k) \to \cos z, and \operatorname{dn}(z,k) \to 1; as k \to 1, they approach \operatorname{sn}(z,k) \to \tanh z, \operatorname{cn}(z,k) \to \operatorname{sech} z, and \operatorname{dn}(z,k) \to \operatorname{sech} z. Named after , who developed them in the late 1820s as part of his systematic study of elliptic integrals—initially publishing key results in —these functions provided a powerful framework for inverting integrals that had puzzled mathematicians since the . Jacobi's notation, refined by later mathematicians like , established \operatorname{sn}, \operatorname{cn}, and \operatorname{dn} as the standard forms, with twelve subsidiary functions (e.g., \operatorname{cd}(z,k) = \operatorname{cn}(z,k)/\operatorname{dn}(z,k)) derived from their ratios. Their theory, deeply connected to Jacobi theta functions via expressions like \operatorname{sn}(z,k) = \frac{\theta_3(0,q)}{\theta_2(0,q)} \frac{\theta_1(\pi z / 2K(k), q)}{\theta_4(\pi z / 2K(k), q)} where q = e^{-\pi K'(k)/K(k)} is the nome, underpins much of 19th-century . Beyond , Jacobi elliptic functions are indispensable in solving nonlinear differential equations arising in physics and , particularly for oscillatory systems where fail. For instance, they describe the exact motion of a simple for finite amplitudes, where the period depends on the swing angle via T = 4K(k)/\sqrt{g/l} with k = \sin(\theta_0/2), generalizing the small-angle approximation. Other applications include the of a on a rotating hoop, rotation (e.g., the via Euler's equations), and nonlinear wave solutions in fields like and . Their addition formulas, such as \operatorname{sn}(u+v,k) = \frac{\operatorname{sn} u \operatorname{cn} v \operatorname{dn} v + \operatorname{sn} v \operatorname{cn} u \operatorname{dn} u}{1 - k^2 \operatorname{sn}^2 u \operatorname{sn}^2 v}, enable explicit solutions to these problems and extend to modern contexts like integrable systems and in .

Overview and Notation

Overview

Jacobi elliptic functions were introduced by in his 1829 treatise Fundamenta Nova Theoriae Functionum Ellipticarum, where he developed them as doubly periodic elliptic functions extending the properties of to the context of elliptic integrals. These functions represent a bridge between elementary and the broader class of elliptic functions, enabling solutions to nonlinear differential equations and integrals that arise in problems where standard trigonometric identities are insufficient. The three primary Jacobi elliptic functions are denoted as \operatorname{sn}(u,k), \operatorname{cn}(u,k), and \operatorname{dn}(u,k), where u is the argument and k (with $0 < k < 1) is the elliptic modulus parameterizing the functions' behavior. As k \to 0, these functions reduce to the sine, cosine, and the constant function 1, respectively, while for k \to 1, they approach hyperbolic functions, illustrating their role as generalizations. Elliptic integrals, from which these functions are inversely defined, are nonelementary integrals of the form \int R(t, \sqrt{P(t)}) \, dt, where R is rational and P(t) is a cubic or quartic polynomial, such as those appearing in the arc length of an ellipse; this entry assumes basic familiarity with them but revisits key aspects later. In mathematics and its applications, Jacobi elliptic functions solve a range of nonlinear problems, particularly in physics, such as the exact motion of a simple pendulum beyond small-angle approximations. They also appear in engineering contexts like nonlinear oscillations and wave propagation. Modern numerical evaluation is facilitated by libraries including SciPy's ellipj function for standard precision computations and the Arb library for arbitrary-precision arithmetic supporting elliptic functions via theta-based implementations.

Notation

The Jacobi elliptic functions are denoted primarily by three functions: the sine amplitude \operatorname{sn}(u,k), cosine amplitude \operatorname{cn}(u,k), and delta amplitude \operatorname{dn}(u,k), where u is the complex argument and k (with $0 \leq k \leq 1) is the elliptic modulus. An alternative notation uses the parameter m = k^2 (with $0 \leq m \leq 1), expressed as \operatorname{sn}(u \mid m), \operatorname{cn}(u \mid m), and \operatorname{dn}(u \mid m); this form appears in historical texts and computational implementations. The complementary modulus is defined as k' = \sqrt{1 - k^2}, which determines the imaginary period of the functions. The quarter-period K(k), or complete elliptic integral of the first kind, is given by K(k) = \int_0^{\pi/2} \frac{\mathrm{d}\theta}{\sqrt{1 - k^2 \sin^2 \theta}}. The complementary quarter-period is K'(k) = K(k'). The associated functions are expressed in terms of the elliptic amplitude \operatorname{am}(u,k), defined such that \operatorname{sn}(u,k) = \sin(\operatorname{am}(u,k)); thus, \operatorname{cn}(u,k) = \cos(\operatorname{am}(u,k)) and \operatorname{dn}(u,k) = \sqrt{1 - k^2 \sin^2(\operatorname{am}(u,k))}. Derivatives of these functions are denoted by differentials, such as \operatorname{sn}'(u,k) = \operatorname{d} \operatorname{sn}(u,k)/\mathrm{d}u, rather than primes. Historically, Carl Gustav Jacob Jacobi introduced the twelve elliptic functions in 1827 without the modern letter-based notation, which was later standardized by Christoph Gudermann in 1838 for \operatorname{sn}, \operatorname{cn}, and \operatorname{dn}, and extended by James Glaisher in 1882 to include the reciprocals like \operatorname{ns}(u,k) = 1/\operatorname{sn}(u,k). Modern references differ slightly: the 1964 edition of Abramowitz and Stegun employs the parameter-based form \operatorname{sn}(u \mid m), while the NIST Digital Library of Mathematical Functions (2010) favors the modulus-based \operatorname{sn}(u,k). In software conventions, MATLAB's ellipj function computes \operatorname{sn}(u,m), \operatorname{cn}(u,m), and \operatorname{dn}(u,m) via [SN, CN, DN] = ellipj(U, M), where M = m = k^2 and U is the argument array, aligning with the parameter notation for numerical efficiency using the arithmetic-geometric mean algorithm.

Definitions

Via elliptic integrals

Elliptic integrals arise as non-elementary antiderivatives in problems such as computing the arc length of an ellipse, where the integral form involves the square root of a cubic or quartic polynomial. These integrals, first studied systematically in the 17th and 18th centuries by mathematicians like John Wallis and Leonhard Euler, cannot be expressed in terms of elementary functions and require special functions for inversion. The incomplete elliptic integral of the first kind is defined as F(\phi \mid k) = \int_0^\phi \frac{\mathrm{d}\theta}{\sqrt{1 - k^2 \sin^2 \theta}}, where \phi is the amplitude and k (with $0 < k < 1) is the elliptic modulus. This integral provides the foundational analytic definition for Jacobi elliptic functions through inversion: given u = F(\phi \mid k), the amplitude function is \phi = \operatorname{am}(u \mid k). The three principal Jacobi elliptic functions are then \operatorname{sn}(u \mid k) = \sin(\operatorname{am}(u \mid k)), \quad \operatorname{cn}(u \mid k) = \cos(\operatorname{am}(u \mid k)), \operatorname{dn}(u \mid k) = \sqrt{1 - k^2 \sin^2(\operatorname{am}(u \mid k))}. The complete elliptic integral of the first kind is the special case K(k) = F(\pi/2 \mid k), which determines the periods of the functions. Specifically, \operatorname{sn}(u + 4K(k) \mid k) = \operatorname{sn}(u \mid k), establishing $4K(k) as the real period, while the imaginary period involves the complementary complete integral K(k') = K(\sqrt{1 - k^2}). This integral-based approach originates from Carl Gustav Jacob Jacobi's seminal 1829 treatise , where he developed the theory of elliptic functions as inverses of elliptic integrals to resolve transformation problems in the field.

Geometric interpretation: the Jacobi ellipse

The Jacobi elliptic functions admit a geometric interpretation analogous to the parameterization of the unit circle by , but applied to an known as the Jacobi . Consider an defined by the equation \frac{x^2}{a^2} + \frac{y^2}{b^2} = 1 with semi-major axis a > b > 0 along the x-direction and k = \sqrt{1 - (b/a)^2}. The parametric equations for points on this are x = a \cn(u, k) and y = b \sn(u, k), where u is the parameter and the functions \sn and \cn are the Jacobi , respectively. This parameterization arises because \sn^2(u, k) + \cn^2(u, k) = 1, ensuring the points satisfy the equation upon scaling by a and b. The parameter \phi = \am(u, k), known as the amplitude, plays the role of the in this construction, with \sn(u, k) = \sin \phi and \cn(u, k) = \cos \phi. Thus, the full parametric form in terms of the eccentric anomaly is x = a \cos \phi and y = b \sin \phi, but reparameterized by u via the of the first kind: u = \int_0^\phi \frac{d\theta}{\sqrt{1 - k^2 \sin^2 \theta}}. As the parameter u increases uniformly, the point (x(u), y(u)) traces the ellipse, with the "speed" varying due to the nonlinear relation between u and \phi. In the k \to 0, the ellipse becomes a circle of radius a = b, \am(u, 0) = u, \sn(u, 0) = \sin u, and \cn(u, 0) = \cos u, recovering standard trigonometric parameterization. Conversely, as k \to 1, b \to 0, the ellipse flattens to the [-a, a] on the x-axis, and the functions approach forms: \sn(u, 1) = \tanh u, \cn(u, 1) = \sech u. This geometric viewpoint highlights the boundedness of the Jacobi elliptic functions, as |\sn(u, k)| \leq 1 and |\cn(u, k)| \leq 1 for real u and $0 \leq k \leq 1, mirroring the range of on the circle. The double periodicity of the functions corresponds to traversing the ellipse multiple times, providing an intuitive grasp of their oscillatory behavior in applications like motion or . Unlike arc-length parameterization, which would yield nonuniform u, this construction emphasizes the analogy to angular parameterization, facilitating conceptual understanding without delving into the underlying integrals.

Using Jacobi theta functions

The Jacobi theta functions provide a foundational framework for expressing the Jacobi elliptic functions through infinite products or sums, leveraging their quasi-periodic properties and modular transformations. The four Jacobi theta functions are defined as infinite series over the , with nome q = e^{i \pi \tau}, where \tau is a complex parameter in the upper half-plane (Im \tau > 0): \begin{align*} \theta_1(z \mid \tau) &= i \sum_{n=-\infty}^{\infty} (-1)^n q^{(n + 1/2)^2} e^{i (2n + 1) \pi z}, \\ \theta_2(z \mid \tau) &= \sum_{n=-\infty}^{\infty} q^{(n + 1/2)^2} e^{i (2n + 1) \pi z}, \\ \theta_3(z \mid \tau) &= \sum_{n=-\infty}^{\infty} q^{n^2} e^{2 i n \pi z}, \\ \theta_4(z \mid \tau) &= \sum_{n=-\infty}^{\infty} (-1)^n q^{n^2} e^{2 i n \pi z}. \end{align*} These functions exhibit modular properties under transformations of \tau, such as \tau \to -1/\tau, which underpin the periodicity of the resulting elliptic functions. For the Jacobi elliptic functions, the parameter k (modulus, $0 < k < 1) relates to \tau = i K'/K, where K = K(k) and K' = K(k') are the complete elliptic integrals of the first kind with k' = \sqrt{1 - k^2}, yielding q = e^{-\pi K'/K}. The functions are then expressed as ratios of theta functions with scaled complex argument \zeta = \frac{\pi u}{2 K}: \begin{align*} \operatorname{sn}(u, k) &= \frac{\theta_3(0, q)}{\theta_2(0, q)} \frac{\theta_1(\zeta, q)}{\theta_4(\zeta, q)}, \\ \operatorname{cn}(u, k) &= \frac{\theta_4(0, q)}{\theta_2(0, q)} \frac{\theta_2(\zeta, q)}{\theta_4(\zeta, q)}, \\ \operatorname{dn}(u, k) &= \frac{\theta_4(0, q)}{\theta_3(0, q)} \frac{\theta_3(\zeta, q)}{\theta_4(\zeta, q)}. \end{align*} This representation ensures the double periodicity of the elliptic functions over the lattice generated by $4K and $2 i K', with the theta ratios capturing the elliptic behavior through the zeros and poles inherent in \theta_1 and the other thetas. Two primary methods derive these expressions. The direct method proceeds via elliptic integrals by inverting the integral definition of u and expressing the amplitude \phi = \operatorname{am}(u, k) in terms of theta series expansions, equating coefficients to obtain the ratios. Alternatively, modular inversion, akin to Landen's transformation, exploits the theta functions' transformation laws under the modular group (e.g., \tau \to -1/\tau), which maps the period parallelogram and yields the elliptic functions through functional equations relating theta values at complementary moduli. These approaches highlight the deep connection between theta functions' modular invariance and the elliptic functions' geometry. In contrast to the Weierstrass elliptic functions, which are typically expressed via the Weierstrass \wp-function with a double pole and defined over a general lattice, the Jacobi forms via theta functions emphasize simple poles and a rectangular lattice scaled by K and K', facilitating applications in real-variable problems like pendulum motion.

Using Neville theta functions

Neville theta functions offer an alternative formulation for defining Jacobi elliptic functions through four real-valued functions, denoted \theta_{jk}(z, q) for appropriate indices j, k, expressed as infinite q-series expansions that depend only on real arguments when z and the nome q are real. These functions are constructed to mirror the properties of Jacobi theta functions but eliminate the need for complex variables in their series representations, making them particularly suitable for computational purposes. The nome q = \exp(-\pi K'/K) is derived from the complete elliptic integrals K and K', providing a bridge to the modulus k. A key expression using these functions is for the Jacobi sine: \sn(u, k) = \frac{\sqrt{m}}{\theta_{30}(0, q)} \cdot \frac{\theta_{41}(\nu, q)}{\theta_{11}(\nu, q)}, where m = k^2, \nu = u \sqrt{(1 + m)/(4 K^2)}, and K is the complete elliptic integral of the first kind. Similar expressions exist for \cn(u, k) and \dn(u, k), enabling the full set of Jacobi elliptic functions to be computed via ratios of these theta functions evaluated at scaled arguments. This representation provides computational advantages over traditional complex-based Jacobi theta formulations, as the real q-series converge rapidly for |q| < 1 and avoid numerical issues associated with complex arithmetic, enhancing stability in software implementations. These properties make Neville theta functions ideal for high-precision evaluations, particularly in scenarios where the modulus k approaches 0 or 1, where other methods may suffer from cancellation errors. The Neville theta functions were originally developed by E. H. Neville in the 1940s as part of his work on elliptic functions, with extensions for practical numerical use appearing in later decades. Modern arbitrary-precision libraries, such as mpmath in Python and Arb, leverage these real theta representations for robust computation of Jacobi elliptic functions across wide ranges of parameters.

Transformations

Imaginary transformations

The imaginary transformations of Jacobi elliptic functions relate the values of the functions at purely imaginary arguments to auxiliary elliptic functions evaluated at real arguments with the complementary modulus k' = \sqrt{1 - k^2}. These transformations, originally derived by , exhibit a hyperbolic character in the limit as k \to 0, where the functions approach trigonometric and hyperbolic counterparts. Specifically, for the principal functions, \operatorname{sn}(i u, k) = i \operatorname{sc}(u, k'), \quad \operatorname{cn}(i u, k) = \operatorname{nc}(u, k'), \quad \operatorname{dn}(i u, k) = \operatorname{dc}(u, k'), where the auxiliary functions are defined as ratios: \operatorname{sc}(u, k') = \operatorname{sn}(u, k') / \operatorname{cn}(u, k'), \operatorname{nc}(u, k') = 1 / \operatorname{cn}(u, k'), and \operatorname{dc}(u, k') = \operatorname{dn}(u, k') / \operatorname{cn}(u, k'). A complete set of these transformations extends to the nine other Jacobi elliptic functions, such as \operatorname{cd}(i u, k) = \operatorname{nd}(u, k') and \operatorname{sc}(i u, k) = i \operatorname{sn}(u, k'), preserving the double-periodic structure while shifting along the imaginary direction. These relations highlight the analytic continuation of the functions into the complex plane, where imaginary arguments map to real ones via the complementary modulus, facilitating numerical evaluation and revealing symmetries. The transformations can be derived from the integral definition of the Jacobi sine, \operatorname{sn}(u, k) = \sin \phi where u = \int_0^\phi (1 - k^2 \sin^2 \theta)^{-1/2} \, d\theta. Substituting an imaginary argument i v yields i v = \int_0^\psi (1 + k^2 \sin^2 \theta)^{-1/2} \, d\theta, which, upon reparameterization with the complementary modulus k', reduces to expressions involving \operatorname{sn}(v, k'), \operatorname{cn}(v, k'), and their ratios, confirming the formulas through differentiation or direct integration. Alternatively, using the representation in terms of Jacobi theta functions, the imaginary transformations follow from the modular transformations of the thetas under \tau \to -1/\tau, such as \vartheta_1(z | -1/\tau) = -i \sqrt{-i \tau} \exp(i z^2 / (\pi \tau)) \vartheta_1(z | \tau), which, when substituted into the quotient forms like \operatorname{sn}(u, k) \propto \vartheta_1(u/ (2K) | q) / \vartheta_3(u/ (2K) | q) (with nome q = e^{i \pi \tau}), yield the elliptic function identities after simplification. These theta-based derivations, established by Jacobi in 1828, underscore the deep connection between elliptic functions and modular forms. Regarding periodicity, the imaginary transformations imply that the function \operatorname{sn}(u, k) acquires an imaginary period of $2 i K', where K' = K(k') is the complete elliptic integral of the first kind for the complementary modulus, transforming the elliptic periods into a framework akin to hyperbolic functions along the imaginary axis. This shift alters the period parallelogram, with the real period $4K mapping to behaviors dominated by i times hyperbolic periods in the limit k' \to 0. In applications to the simple pendulum, these transformations arise when considering imaginary time \tau = i t, converting the oscillatory real-time solutions expressed via \operatorname{sn} and \operatorname{dn} into inverted pendulum trajectories or tunneling paths in quantum mechanics, where the modulus duality k \to k' links bounded motion to unbounded hyperbolic regimes, as seen in instanton calculations for barrier penetration.

Real transformations

Real transformations of Jacobi elliptic functions involve changes to the real argument u and modulus k that preserve the doubly periodic, elliptic character of the functions while relating them to instances with altered parameters. These transformations, such as the Landen and Gauss varieties, enable efficient computation, modulus adjustment, and connections to elliptic integrals by scaling the argument and modifying k. They differ from imaginary transformations by avoiding hyperbolic shifts and maintaining real ellipticity. The foundations of these real transformations trace back to Carl Gustav Jacob Jacobi's work in the 19th century, where he developed addition formulas that underpin double-angle and scaling relations for real arguments, as detailed in his seminal treatise on . Jacobi's real addition formulas, such as those for \operatorname{sn}(u + v, k), facilitated derivations of transformation laws by specializing to equal arguments, yielding double-angle identities like \operatorname{sn}(2u, k) = \frac{2 \operatorname{sn}(u, k) \operatorname{cn}(u, k) \operatorname{dn}(u, k)}{1 - k^2 \operatorname{sn}^4(u, k)}. These serve as basic real scaling mechanisms for \lambda = 2, relating the function at doubled argument to products of the original functions. The Landen transformation encompasses both descending and ascending forms, connecting Jacobi functions with different moduli via real argument rescaling. In the descending Landen transformation, the new modulus is k_1 = \frac{1 - k'}{1 + k'}, where k' = \sqrt{1 - k^2}, and the argument scales by \frac{1}{1 + k_1}: \begin{align*} \operatorname{sn}(z, k) &= \frac{(1 + k_1) \operatorname{sn}\left( \frac{z}{1 + k_1}, k_1 \right)}{1 + k_1 \operatorname{sn}^2\left( \frac{z}{1 + k_1}, k_1 \right)}, \\ \operatorname{cn}(z, k) &= \frac{\operatorname{cn}\left( \frac{z}{1 + k_1}, k_1 \right) \operatorname{dn}\left( \frac{z}{1 + k_1}, k_1 \right)}{1 + k_1 \operatorname{sn}^2\left( \frac{z}{1 + k_1}, k_1 \right)}, \\ \operatorname{dn}(z, k) &= \frac{\operatorname{dn}^2\left( \frac{z}{1 + k_1}, k_1 \right) - (1 - k_1)}{1 + k_1 - \operatorname{dn}^2\left( \frac{z}{1 + k_1}, k_1 \right)}. \end{align*} This relates functions with modulus k > k_1 through argument compression. The ascending reverses this process, with new modulus k_2 = \frac{2 \sqrt{k}}{1 + k} (so k_2^2 = \frac{4k}{(1 + k)^2}) and complementary k_2' = \frac{1 - k}{1 + k}, scaling the argument by \frac{1}{1 + k_2'}: \begin{align*} \operatorname{sn}(z, k) &= \frac{(1 + k_2') \operatorname{sn}\left( \frac{z}{1 + k_2'}, k_2 \right) \operatorname{cn}\left( \frac{z}{1 + k_2'}, k_2 \right)}{\operatorname{dn}\left( \frac{z}{1 + k_2'}, k_2 \right)}, \\ \operatorname{cn}(z, k) &= \frac{(1 + k_2') \left[ \operatorname{dn}^2\left( \frac{z}{1 + k_2'}, k_2 \right) - k_2' \right]}{k_2^2 \operatorname{dn}\left( \frac{z}{1 + k_2'}, k_2 \right)}, \\ \operatorname{dn}(z, k) &= \frac{(1 - k_2') \left[ \operatorname{dn}^2\left( \frac{z}{1 + k_2'}, k_2 \right) + k_2' \right]}{k_2^2 \operatorname{dn}\left( \frac{z}{1 + k_2'}, k_2 \right)}. \end{align*} This expands the argument and increases the modulus toward 1. The ascending form is also known as the Gauss transformation when expressed in terms of the original modulus, such as \operatorname{sn}((1 + k)u, k_2) = \frac{(1 + k) \operatorname{sn}(u, k)}{1 + k \operatorname{sn}^2(u, k)}, highlighting the scaling factor $1 + k. These transformations impact the quarter-period K(k) = \int_0^1 \frac{dt}{\sqrt{(1 - t^2)(1 - k^2 t^2)}}. For the descending Landen, K(k) = (1 + k_1) K(k_1), compressing the period as the modulus decreases. Conversely, the ascending (Gauss) transformation yields K(k) = (1 + k) K(k_2), where k_2 = \frac{2 \sqrt{k}}{1 + k}, expanding the period as the modulus increases; iteratively applying this relates K(k) to the arithmetic-geometric mean for numerical evaluation. More general real scalings, such as for arbitrary \lambda, follow from repeated application of double-angle formulas or representations, adjusting k' to maintain consistency.

Amplitude and other transformations

The Jacobi amplitude function, denoted \am(u, k), provides a geometric interpretation for the Jacobi elliptic functions by mapping the elliptic argument u to an angle \phi = \am(u, k) such that u = F(\phi, k), where F is the incomplete of the first kind. This function satisfies \sin(\am(u, k)) = \sn(u, k), \cos(\am(u, k)) = \cn(u, k), and \dn(u, k) = \sqrt{1 - k^2 \sin^2(\am(u, k))}. Transformations involving the amplitude often arise from addition theorems for the underlying elliptic functions, which can be used to derive expressions for \am(u + v, k). Specifically, \am(u + v, k) = \arcsin(\sn(u + v, k)), where the addition formula for \sn is \sn(u + v, k) = \frac{\sn u \ \cn v \ \dn v + \sn v \ \cn u \ \dn u}{1 - k^2 \sn^2 u \ \sn^2 v}. Similar addition formulas exist for \cn(u + v, k) and \dn(u + v, k). These relations allow computation of phase shifts in the , analogous to trigonometric identities, though the elliptic case lacks simple additivity due to the nonlinear dependence on the modulus k. For instance, expressing the addition directly in terms of amplitudes \phi = \am(u, k) and \psi = \am(v, k) involves tangent relations derived from the elliptic integral addition theorem, but remains algebraically complex. Modular inversion transformations relate Jacobi elliptic functions of modulus k > 1 to those of the reciprocal modulus $1/k < 1, facilitating numerical computation and analysis by reducing to the standard range $0 < k < 1. The key relations are \sn(u, k) = \frac{1}{k} \sn(ku, 1/k), \quad \operatorname{cn}(u, k) = \dn(ku, 1/k), \quad \dn(u, k) = \sqrt{\frac{k^2 - \operatorname{sn}^2(ku, 1/k)}{k^2}}. These follow from the reciprocal-modulus transformation for elliptic integrals, F(\phi, k) = \frac{1}{k} F(\chi, 1/k) with \sin \chi = k \sin \phi, and extend to the amplitude via \am(ku, 1/k) = \arcsin(k \sin(\am(u, k))). Jacobi elliptic functions admit Fourier series decompositions over their real period $4K(k), useful for approximations and analytic continuations. For instance, \operatorname{cn}(u, k) = \frac{2\pi}{K k} \sum_{n=0}^\infty \frac{ q^{n + 1/2} }{1 + q^{2n+1} } \cos \left( (2n+1) \frac{\pi u}{2 K} \right), where q = e^{-\pi K(k')/K(k)} is the nome, with analogous sine and cosine series for \sn(u, k) and \dn(u, k). These expansions converge rapidly for small q (large k) and highlight the periodic structure. The imaginary transformations (scaling arguments by i K'/K) and real transformations (reciprocal modulus scalings), combined with Landen transformations, generate the full group of modular transformations acting on the elliptic modulus via the parameter \tau = i K'/K. This group is isomorphic to the special linear group \mathrm{SL}(2, \mathbb{Z}), reflecting the underlying modular invariance of elliptic functions akin to that in theta function theory.

Jacobi hyperbola

The Jacobi hyperbola arises as the hyperbolic analog of the geometric interpretations associated with Jacobi elliptic functions in the limiting case where the modulus k \to 1^-. In this limit, the complete elliptic integral of the first kind K(k) diverges to infinity, and the functions degenerate to hyperbolic functions: \operatorname{sn}(u,1) = \tanh u, \operatorname{cn}(u,1) = \operatorname{sech} u, and \operatorname{dn}(u,1) = \operatorname{sech} u. Unlike the double periodicity of the elliptic case, these hyperbolic functions exhibit a single real period that becomes infinite, resulting in aperiodic behavior over the real line. This degeneration provides a parametric representation of the rectangular hyperbola x^2 - y^2 = 1. Specifically, setting x = 1/\operatorname{cn}(u,1) = \cosh u and y = \operatorname{sn}(u,1)/\operatorname{cn}(u,1) = \sinh u traces the hyperbola, where the parameter u corresponds to the hyperbolic arc length from the vertex. This construction parallels the elliptic case but replaces the bounded elliptic orbits with unbounded hyperbolic trajectories, reflecting the loss of the second period. These functions find applications in soliton theory, where the infinite period facilitates descriptions of localized, non-periodic wave solutions. For instance, the \operatorname{sech}^2 profile of the KdV soliton emerges as the k \to 1 limit of periodic Jacobi elliptic wave trains, enabling exact solutions for nonlinear wave propagation in dispersive media.

Minor functions

In addition to the three primary Jacobi elliptic functions \sn(u,k), \cn(u,k), and \dn(u,k), there are nine auxiliary functions known as minor functions, formed by taking reciprocals and ratios of the primaries. These minor functions facilitate compact notation in expressions and computations involving elliptic functions. For example, the reciprocal functions are defined as \ns(u,k) = 1/\sn(u,k), \nc(u,k) = 1/\cn(u,k), and \nd(u,k) = 1/\dn(u,k), while ratio functions include \dc(u,k) = \dn(u,k)/\cn(u,k), \sc(u,k) = \sn(u,k)/\cn(u,k), and \sd(u,k) = \sn(u,k)/\dn(u,k). The derivatives of the primary functions with respect to the argument u are expressed as products of the other primaries, serving an auxiliary role analogous to the minor functions in simplifying differential relations. Specifically, \frac{d}{du} \sn(u,k) = \cn(u,k) \dn(u,k), \frac{d}{du} \cn(u,k) = -\sn(u,k) \dn(u,k), \frac{d}{du} \dn(u,k) = -k^2 \sn(u,k) \cn(u,k). These derivative expressions, along with the minor functions, are termed "minor" in the literature to denote their supportive yet essential utility in derivations, a convention tracing to Jacobi's foundational work on elliptic functions. Minor functions are instrumental in simplifying algebraic identities and integral evaluations within elliptic theory. They appear prominently in integral representations of auxiliary quantities, such as the Jacobi epsilon function \E(u,k) = \int_0^u \dn^2(t,k) \, dt, where forms involving \dc, \nd, and \ns provide concise expressions for quasi-periodic behaviors and symmetry properties. The full system comprises the three primary functions and the nine minor functions, enumerated in the table below for clarity:
FunctionDefinition
\sn\sn(u,k)
\cn\cn(u,k)
\dn\dn(u,k)
\ns$1/\sn(u,k)
\nc$1/\cn(u,k)
\nd$1/\dn(u,k)
\sc\sn(u,k)/\cn(u,k)
\cs\cn(u,k)/\sn(u,k)
\sd\sn(u,k)/\dn(u,k)
\ds\dn(u,k)/\sn(u,k)
\cd\cn(u,k)/\dn(u,k)
\dc\dn(u,k)/\cn(u,k)
This nomenclature ensures mutual reciprocity, as the reciprocal of a minor function yields another in the set (e.g., $1/\sc(u,k) = \cs(u,k)).

Analytic Properties

Periodicity, poles, and residues

The Jacobi elliptic function \mathrm{sn}(u, k) exhibits double periodicity in the complex plane. It satisfies \mathrm{sn}(u + 4K, k) = \mathrm{sn}(u, k) and \mathrm{sn}(u + 2iK', k) = \mathrm{sn}(u, k), where K = K(k) is the complete elliptic integral of the first kind and K' = K(\sqrt{1 - k^2}). Additionally, it has half-period anti-symmetry \mathrm{sn}(u + 2K, k) = -\mathrm{sn}(u, k). The function \mathrm{sn}(u, k) has simple poles located at u = 2mK + i(2n + 1)K' for all integers m, n. These singularities are the only poles, and they form a lattice congruent to the principal pole at u = iK' modulo translations by $2K and $2iK'. The residue of \mathrm{sn}(u, k) at the principal pole u_0 = iK' is $1/k. Due to the anti-periodicity under translations by $2K and periodicity under translations by $2iK', the residue at a general pole u_0 = 2mK + i(2n + 1)K' is (-1)^{m}/k. The zeros of \mathrm{sn}(u, k) occur at the points of the period lattice u = 2mK + 2niK' for integers m, n. These are simple zeros, with the principal zero at u = 0. As meromorphic functions on the complex plane, the Jacobi elliptic functions extend naturally to the compact Riemann surface formed by the quotient \mathbb{C}/\Lambda, where \Lambda is the period lattice generated by $4K and $2iK'. This surface is a torus of genus 1.

Special values

The Jacobi elliptic functions sn(u,k), cn(u,k), and dn(u,k) admit simple explicit evaluations at the origin u=0, yielding sn(0,k)=0, cn(0,k)=1, and dn(0,k)=1 for any modulus k ∈ (0,1). At the quarter-period u=K(k), where K(k) is the complete elliptic integral of the first kind, the functions evaluate to sn(K(k),k)=1, cn(K(k),k)=0, and dn(K(k),k)=√(1-k²). In the singular limit as k → 0, the Jacobi functions reduce to elementary trigonometric forms: sn(u,k) → sin(u), cn(u,k) → cos(u), and dn(u,k) → 1. As k → 1, they approach hyperbolic functions: sn(u,k) → tanh(u), cn(u,k) → sech(u), and dn(u,k) → sech(u). A notable special case arises for the lemniscate modulus k=1/√2, where the complementary modulus k'=√(1-k²)=1/√2, and thus dn(K(1/√2),1/√2)=1/√2, while sn(K(1/√2),1/√2)=1 and cn(K(1/√2),1/√2)=0 as in the general case.
Modulus ksn(0,k)cn(0,k)dn(0,k)sn(K,k)cn(K,k)dn(K,k)
0011101
0.501110√0.75 ≈ 0.866
1/√2 ≈ 0.707011101/√2 ≈ 0.707
1011100
These evaluations at u=0 and u=K(k) hold symbolically for the selected moduli, with numerical approximations shown for illustration where k>0; the k=0 and k=1 rows reflect the singular limits.

Identities and Formulas

Basic relations and half-angle formulas

The Jacobi elliptic functions satisfy fundamental algebraic identities analogous to the in . Specifically, for the principal functions \operatorname{sn} u, \operatorname{cn} u, and \operatorname{dn} u with modulus k, \operatorname{sn}^2 u + \operatorname{cn}^2 u = 1, \quad k^2 \operatorname{sn}^2 u + \operatorname{dn}^2 u = 1. These relations follow directly from the integral definition of \operatorname{sn} u as the inverse of the elliptic integral of the first kind and the geometric interpretation of the functions on the elliptic curve. Half-argument formulas express the functions at u/2 in terms of those at u, facilitating recursive computations and series expansions. One standard form is \operatorname{sn}^2 \frac{u}{2} = \frac{1 - \operatorname{cn} u}{1 + \operatorname{dn} u}, with similar expressions for the cosine and delta analogs: \operatorname{cn}^2 \frac{u}{2} = \frac{-k'^2 + \operatorname{dn} u + k^2 \operatorname{cn} u}{k^2 (1 + \operatorname{cn} u)}, \quad \operatorname{dn}^2 \frac{u}{2} = \frac{\operatorname{dn} u + k^2 \operatorname{cn} u + k'^2}{1 + \operatorname{dn} u}, where k' = \sqrt{1 - k^2}. These can be derived by solving the double-angle formulas for half-arguments or via differentiation of the defining integral. The derivatives of the Jacobi functions are products of the other functions, reflecting their role as solutions to nonlinear equations. In particular, \frac{d}{du} \operatorname{sn} u = \operatorname{cn} u \, \operatorname{dn} u, \quad \frac{d}{du} \operatorname{cn} u = -\operatorname{sn} u \, \operatorname{dn} u, \quad \frac{d}{du} \operatorname{dn} u = -k^2 \operatorname{sn} u \, \operatorname{cn} u. These follow from implicit differentiation of the Pythagorean identities combined with the chain rule applied to the \operatorname{am} u. Indefinite integrals of the functions often reduce to logarithmic or trigonometric forms, though some involve s. For instance, \int \operatorname{sn} u \, du = \frac{1}{k} \ln \left( \operatorname{dn} u - k \operatorname{cn} u \right) + C, up to a constant and assuming the principal branch. An alternative expression in terms of the incomplete elliptic integral of the second kind E(\phi \mid k^2) is \int_0^u \operatorname{sn} t \, dt = \frac{1}{k^2} \left( u - E(\operatorname{am} u \mid k^2) \right), which highlights the connection to elliptic integrals and is useful for definite limits from zero. Proofs of these basic relations and formulas can be established rigorously using representations in terms of Jacobi theta functions, as detailed in classical analyses.

Addition theorems and K formulas

The addition theorems for Jacobi elliptic functions express the functions evaluated at the sum of arguments in terms of the functions at each individual argument. These theorems are analogous to the angle addition formulas for and are fundamental for deriving more complex identities. Specifically, for the sine amplitude function, \operatorname{sn}(u + v) = \frac{\operatorname{sn} u \, \operatorname{cn} v \, \operatorname{dn} v + \operatorname{sn} v \, \operatorname{cn} u \, \operatorname{dn} u}{1 - k^2 \operatorname{sn}^2 u \, \operatorname{sn}^2 v}, with similar expressions for the cosine amplitude and delta amplitude functions: \operatorname{cn}(u + v) = \frac{\operatorname{cn} u \, \operatorname{cn} v - \operatorname{sn} u \, \operatorname{dn} u \, \operatorname{sn} v \, \operatorname{dn} v}{1 - k^2 \operatorname{sn}^2 u \, \operatorname{sn}^2 v}, \operatorname{dn}(u + v) = \frac{\operatorname{dn} u \, \operatorname{dn} v - k^2 \operatorname{sn} u \, \operatorname{cn} u \, \operatorname{sn} v \, \operatorname{cn} v}{1 - k^2 \operatorname{sn}^2 u \, \operatorname{sn}^2 v}. These formulas hold for the modulus k with $0 < k < 1 and are derived from the geometric properties of elliptic integrals. Double-angle formulas arise as special cases by setting u = v in the addition theorems. For instance, \operatorname{sn}(2u) = \frac{2 \operatorname{sn} u \, \operatorname{cn} u \, \operatorname{dn} u}{1 - k^2 \operatorname{sn}^4 u}, with corresponding expressions for \operatorname{cn}(2u) and \operatorname{dn}(2u). These simplify computations for even multiples of the argument and connect to half-angle relations through inversion. The complete elliptic integral of the first kind, K(k), which defines the quarter-period of the Jacobi functions, admits a representation in terms of the Gaussian hypergeometric function: K(k) = \frac{\pi}{2} \, {}_2F_1\left(\frac{1}{2}, \frac{1}{2}; 1; k^2\right). This integral form, K(k) = \int_0^{\pi/2} (1 - k^2 \sin^2 \theta)^{-1/2} \, d\theta, underpins the periodicity of the functions, and the hypergeometric expression facilitates series expansions and asymptotic analysis. The addition theorems enable recurrences for multiple-angle evaluations, such as expressing \operatorname{sn}(nu) in terms of \operatorname{sn}((n-2)u) via repeated application, which is useful in solving nonlinear differential equations and in numerical algorithms for elliptic integrals. For example, such recurrences appear in the computation of elliptic integrals over multiple periods.

Relations between squares

The fundamental algebraic relations among the squares of the Jacobi elliptic functions are the Pythagorean-type identities \operatorname{sn}^2 z + \operatorname{cn}^2 z = 1, \operatorname{dn}^2 z + k^2 \operatorname{sn}^2 z = 1, which hold for any complex argument z and modulus k with $0 < k < 1. These relations stem from the geometric interpretation of the functions as coordinates on an and enable straightforward expressions for each square in terms of the others: \operatorname{cn}^2 z = 1 - \operatorname{sn}^2 z, \quad \operatorname{dn}^2 z = 1 - k^2 \operatorname{sn}^2 z, \quad \operatorname{sn}^2 z = \frac{1 - \operatorname{dn}^2 z}{k^2}. Since the identities are independent of the specific argument, they extend directly to sums and differences, yielding \operatorname{sn}^2(u + v) + \operatorname{cn}^2(u + v) = 1, \quad \operatorname{dn}^2(u + v) + k^2 \operatorname{sn}^2(u + v) = 1, and similarly for u - v. More involved relations arise from products involving squares across different arguments. For instance, the sum of the dn functions at summed and differenced arguments satisfies \operatorname{dn}(u + v) + \operatorname{dn}(u - v) = \frac{2 \operatorname{dn} u \operatorname{dn} v}{1 - k^2 \operatorname{sn}^2 u \operatorname{sn}^2 v}, derived from the addition theorem for dn. Such relations are particularly useful in reducing powers of the elliptic functions for evaluating definite integrals, such as those arising in mechanics and electrostatics. For example, higher even powers like \operatorname{sn}^4 z can be expressed as \operatorname{sn}^4 z = (1 - \operatorname{dn}^2 z)^2 / k^4 - 2(1 - \operatorname{dn}^2 z)/k^2 + 1, allowing integration by parts or substitution to lower-degree forms expressible in terms of complete elliptic integrals. This reduction technique, detailed in standard handbooks, avoids direct computation of higher-order terms and leverages the modular parameter k for numerical stability.

Differential Equations

As solutions to nonlinear ODEs

Jacobi elliptic functions provide exact solutions to a class of nonlinear ordinary differential equations (ODEs) that arise in physical and mathematical contexts where trigonometric or hyperbolic functions are insufficient due to the inherent nonlinearity. Unlike the sine and cosine functions, which solve linear harmonic oscillator equations such as \frac{d^2 x}{dt^2} + \omega^2 x = 0, Jacobi functions address equations with periodic but anharmonic potentials, such as those involving \sin x or quadratic terms in elliptic functions themselves. This capability stems from their double periodicity and bounded oscillation properties, making them ideal for modeling systems with conserved energy in nonlinear settings. A prominent example is the simple pendulum equation, \frac{d^2 \theta}{dt^2} + \frac{g}{l} \sin \theta = 0, where \theta(t) is the angular displacement, g is gravitational acceleration, and l is the pendulum length. For initial conditions with maximum angle \theta_0 < \pi, the exact solution is \theta(t) = 2 \operatorname{am}\left( \sqrt{\frac{g}{l}} t, k \right), with modulus k = \sin(\theta_0 / 2). Here, \operatorname{am}(u, k) is the Jacobi amplitude function, and the period of oscillation is T = 4 K(k) / \sqrt{g/l}, where K(k) is the complete elliptic integral of the first kind. This solution captures the full nonlinear dynamics, reducing to the small-angle approximation \theta(t) \approx \theta_0 \cos(\sqrt{g/l} t) as k \to 0. Another key application is the Lamé equation, \frac{d^2 y}{du^2} + \left[ h - n(n+1) k^2 \operatorname{sn}^2 u \right] y = 0, where n is an integer, h is the eigenvalue, and \operatorname{sn} u is the Jacobi sine function with modulus k. Solutions, known as Lamé functions, take the form y(u) = \operatorname{Ec}_m^n(u, k) or \operatorname{Es}_m^n(u, k), which are finite products or sums involving Jacobi elliptic functions like \operatorname{sn}^n u \cdot \operatorname{cn}^m u \cdot \operatorname{dn}^p u (with m + p = n - 2l for integers l). These solutions are periodic and exhibit band structure in their eigenvalues, relevant to quantum mechanics and separation of variables in ellipsoidal coordinates. In contrast to Weierstrass elliptic functions, which satisfy the autonomous cubic ODE \left( \frac{d \wp}{dz} \right)^2 = 4 \wp^3 - g_2 \wp - g_3 and are suited for algebraic-geometric interpretations via elliptic curves, Jacobi functions excel in explicit solutions to non-autonomous ODEs with trigonometric-like periodicity, such as the pendulum or Lamé forms. The nonlinearity in these equations prevents reduction to elementary functions, necessitating elliptic functions to invert the resulting elliptic integrals while preserving integrability. Historically, Carl Gustav Jacob Jacobi developed these functions in the 1820s–1830s, motivated by the need to invert elliptic integrals arising in pendulum motion and astronomical problems, laying foundations for modern integrable systems theory where such ODEs represent exactly solvable Hamiltonian dynamics.

Derivatives with respect to amplitude

The first derivatives of the Jacobi elliptic functions with respect to the amplitude u (with modulus k fixed) are given by \frac{d}{du} \operatorname{sn}(u,k) = \operatorname{cn}(u,k) \operatorname{dn}(u,k), \frac{d}{du} \operatorname{cn}(u,k) = -\operatorname{sn}(u,k) \operatorname{dn}(u,k), \frac{d}{du} \operatorname{dn}(u,k) = -k^2 \operatorname{sn}(u,k) \operatorname{cn}(u,k). These relations express each derivative as the product of the other two functions (up to sign and the factor k^2), highlighting the interdependence among \operatorname{sn}, \operatorname{cn}, and \operatorname{dn}. Differentiating these first-order formulas yields second derivatives, such as \frac{d^2}{du^2} \operatorname{sn}(u,k) = -(1 + k^2) \operatorname{sn}(u,k) + 2 k^2 \operatorname{sn}^3(u,k), with analogous expressions for \operatorname{cn} and \operatorname{dn}. Squaring the first derivative of \operatorname{sn}(u,k) produces the nonlinear first-order ordinary differential equation \left( \frac{d}{du} \operatorname{sn}(u,k) \right)^2 = (1 - \operatorname{sn}^2(u,k))(1 - k^2 \operatorname{sn}^2(u,k)), which encapsulates the defining dynamics of the function; similar equations hold for \operatorname{cn} and \operatorname{dn}. The derivative of the Jacobi amplitude function is \frac{d}{du} \operatorname{am}(u,k) = \operatorname{dn}(u,k). This relation links the amplitude directly to \operatorname{dn}, facilitating chain rule applications in compositions involving elliptic integrals. These derivative formulas are essential for numerical integration of elliptic integrals via their inverses and for developing series expansions of the functions, as they enable recursive computation and stability analysis in algorithms.

Series Expansions and Approximations

Expansion in terms of the nome

The nome q of the Jacobi elliptic functions is defined as q = \exp\left( -\pi \frac{K'(k)}{K(k)} \right), where K(k) and K'(k) = K(\sqrt{1-k^2}) are the complete elliptic integrals of the first kind. This parameter, with $0 < q < 1, facilitates expansions that generalize trigonometric Fourier series to the elliptic case. The Jacobi elliptic sine function admits an exact expression as a ratio of Jacobi theta functions: \operatorname{sn}(z, k) = \frac{\theta_3(0, q)}{\theta_2(0, q)} \frac{\theta_1(\zeta, q)}{\theta_4(\zeta, q)}, where \zeta = \frac{\pi z}{2 K(k)} and the theta functions are defined via their standard q-series or product representations. Similar theta-based formulas hold for \operatorname{cn}(z, k) and \operatorname{dn}(z, k): \operatorname{cn}(z, k) = \frac{\theta_4(0, q)}{\theta_2(0, q)} \frac{\theta_2(\zeta, q)}{\theta_4(\zeta, q)}, \quad \operatorname{dn}(z, k) = \frac{\theta_4(0, q)}{\theta_3(0, q)} \frac{\theta_3(\zeta, q)}{\theta_4(\zeta, q)}. These representations derive directly from the periodicity and quasi-periodicity properties of the theta functions, linking the elliptic functions to modular forms. A Fourier-like q-series expansion for \operatorname{sn}(z, k) is \operatorname{sn}(z, k) = \frac{2\pi}{K(k) k} \sum_{n=0}^{\infty} \frac{q^{n + 1/2}}{1 - q^{2n+1}} \sin\left( (2n+1) \zeta \right), with the same \zeta as above. This series converges when q \exp(2 |\Im \zeta|) < 1, which holds rapidly for small q (corresponding to small modulus k \approx 0), where q decays exponentially. For small k, the series provides high-precision evaluations of \operatorname{sn}(z, k) with few terms, making it valuable in applications requiring accurate elliptic function computations, such as in or . For large k near 1 (where q approaches 1 and direct convergence slows), post-2000 developments employ modular transformations of the underlying to map the problem to an equivalent form with a small effective nome, enabling rapid series convergence via the arithmetic-geometric mean or . These techniques, implemented in modern numerical libraries, extend the q-series utility across the full range of k.

Approximation via hyperbolic functions

Jacobi elliptic functions admit asymptotic approximations in terms of elementary trigonometric or hyperbolic functions when the modulus k approaches the extreme values 0 or 1, facilitating analytical insights and numerical estimates in limiting regimes. These approximations arise from the defining integral representations and are particularly useful for understanding the transition from periodic to aperiodic behavior. For small k, the functions approach their trigonometric counterparts, with perturbative corrections derived by Taylor expanding the integrand in the elliptic integral definition of the amplitude function \mathrm{am}(u,k). Specifically, \mathrm{am}(u,k) \approx u - \frac{k^2}{4} (u - \sin u \cos u) + O(k^4). Then, \mathrm{sn}(u,k) = \sin(\mathrm{am}(u,k)) \approx \sin u - \frac{k^2}{4} (u - \sin u \cos u) \cos u + O(k^4); \mathrm{cn}(u,k) = \cos(\mathrm{am}(u,k)) \approx \cos u + \frac{k^2}{4} (u - \sin u \cos u) \sin u + O(k^4); and \mathrm{dn}(u,k) \approx 1 - \frac{k^2}{2} \sin^2 u + O(k^4). These series, obtained by inverting the expanded elliptic integral u = \int_0^{\mathrm{am}(u,k)} \frac{dt}{\sqrt{1 - k^2 \sin^2 t}}, provide first-order perturbations beyond the leading trigonometric limits \mathrm{sn}(u,0) = \sin u, \mathrm{cn}(u,0) = \cos u, and \mathrm{dn}(u,0) = 1. As k \to 1^-, letting k' = \sqrt{1 - k^2} small, the functions degenerate to hyperbolic forms via a complementary modulus transformation, with corrections from expanding the integral in powers of k'^2. The amplitude satisfies \mathrm{am}(u,k) \approx \mathrm{gd}(u) - \frac{k'^2}{4} (u - \sinh u \cosh u) \sech u + O(k'^4), where \mathrm{gd}(u) is the Gudermannian function satisfying \sin(\mathrm{gd}(u)) = \tanh u and \cos(\mathrm{gd}(u)) = \sech u. Then, \mathrm{sn}(u,k) \approx \tanh u - \frac{k'^2}{4} (u - \sinh u \cosh u) \sech^2 u + O(k'^4); \mathrm{cn}(u,k) \approx \sech u + \frac{k'^2}{4} (u - \sinh u \cosh u) \sech u \tanh u + O(k'^4); and \mathrm{dn}(u,k) \approx \sech u + \frac{k'^2}{4} (u \sech u \tanh u - \sech^3 u) + O(k'^4). These derive from a first-order Taylor expansion of the sine and cosine around the Gudermannian, followed by the hyperbolic limits \mathrm{sn}(u,1) = \tanh u, \mathrm{cn}(u,1) = \mathrm{dn}(u,1) = \sech u, reflecting the infinite period as k \to 1. Such approximations, often obtained equivalently via theta function expansions in the nome q = e^{-\pi K'/K}, enable rapid evaluations without full elliptic computations, proving valuable in engineering contexts like approximating solutions to nonlinear oscillators or wave equations where exact elliptic forms are cumbersome.

Continued fraction representations

Continued fraction representations provide powerful tools for computing and analyzing Jacobi elliptic functions and associated elliptic integrals, offering rapid convergence properties especially when the complementary modulus k' = \sqrt{1 - k^2} is small. These representations were developed in the 19th century by and , who explored their theoretical foundations and applications in solving nonlinear differential equations and physical problems. Gauss's pioneering work on continued fractions for hypergeometric functions laid the groundwork, while Jacobi extended these ideas to elliptic functions, enabling more efficient numerical evaluations compared to power series expansions. For the complete elliptic integral of the first kind, a notable continued fraction is K(k) = \frac{\pi}{2} \left[ 1 + \frac{\left(\frac{1}{2} k'\right)^2}{1 + \frac{\left(\frac{3}{2} k'\right)^2}{1 + \frac{\left(\frac{5}{2} k'\right)^2}{1 + \ddots}}} \right], derived from the hypergeometric representation K(k) = \frac{\pi}{2} {}_2F_1\left(\frac{1}{2},\frac{1}{2};1;k^2\right) using Gauss's general continued fraction for the ratio of consecutive hypergeometric functions. This form exhibits rapid convergence for small k', making it advantageous for computational purposes where k is close to 1, as each term contributes significantly less than in the corresponding series expansion. Gauss's form for ratios of elliptic integrals, such as that between the complete elliptic integral of the second kind E(k) and K(k), is given by \frac{E(k)}{K(k)} = 1 + \frac{k^2}{2 + \frac{3k^2}{4 + \frac{5k^2}{6 + \ddots}}} \cdot \frac{1}{1 - k^2}, or equivalently, E(k) = K(k) \left[ 1 - \frac{k^2}{2 + \frac{3k^2}{4 + \frac{5k^2}{6 + \ddots}}} \right] \frac{1}{1 - k^2}. This representation allows for the simultaneous computation of E(k) and K(k) with high accuracy using modified Lentz's method or similar algorithms, highlighting the efficiency of continued fractions in avoiding overflow issues in series methods. For the Jacobi elliptic sine function sn(u,k), an infinite continued fraction can be derived using transformations involving the hyperbolic tangent function, particularly through Laplace transforms or addition theorems. One such form arises from the integral representation and yields rapid convergence for moderate u and k, aiding in numerical evaluations of periodic phenomena in physics. These 19th-century developments by Gauss and Jacobi underscore the enduring utility of continued fractions for both theoretical insights and practical computations in elliptic theory.

Computation and Inverses

Fast numerical computation

Efficient numerical evaluation of Jacobi elliptic functions, such as sn(u, k), cn(u, k), and dn(u, k), relies on algorithms that leverage their connections to elliptic integrals and theta functions, achieving high precision with controlled computational complexity. The arithmetic-geometric mean (AGM) iteration, originally developed by Gauss and Legendre, provides a foundational method for computing the complete elliptic integral of the first kind, K(k), which is essential for scaling the argument u and normalizing computations. This iterative process converges quadratically, requiring only O(log p) steps for p-bit precision, where each step involves simple arithmetic operations on pairs of values starting from a_0 = 1 and b_0 = sqrt(1 - k^2). Once K(k) is obtained, the AGM can be extended via ascending transformations or direct iteration to evaluate the amplitude function am(u, k), from which sn(u, k) = sin(am(u, k)), cn(u, k) = cos(am(u, k)), and dn(u, k) = sqrt(1 - k^2 sin^2(am(u, k))) follow immediately; this yields an overall complexity of O(1) evaluations after the initial O(log p) setup for fixed k. For scenarios with fixed modulus k, table-based methods precompute coefficients or interpolation tables derived from Chebyshev polynomial approximations of K(k) and E(k), enabling rapid evaluation of the functions for varying u without repeated integral computations. These approaches, which exploit the smoothness of the functions over u, can accelerate computations by 25 to 70 percent compared to general-purpose elliptic integral evaluations, particularly when evaluating series of arguments. When the modulus k is small, corresponding to a small nome q = exp(-π K'(k)/K(k)), the Jacobi elliptic functions can be computed via truncated q-series expansions of the underlying , as the series converge rapidly with few terms (e.g., O(1/q) terms for precision). This method is especially efficient for k near 0, where trigonometric limits apply, and achieves O(p^{1.5}) bit complexity through optimized rectangular summation techniques. Post-1990s advancements include Carlson's degenerate symmetric forms for elliptic integrals, which simplify numerical evaluation by reducing to hypergeometric series with arithmetic-geometric mean acceleration, extending efficiently to complex arguments and achieving O(p^{1.667}) complexity with dynamic precision control. Chebyshev approximations, fitted to minimax error over the modulus domain, provide another modern route: polynomials of degree 10-15 compute K(k) and E(k) with relative errors below 10^{-15}, followed by recursive or series-based steps for the incomplete functions, offering speeds comparable to nome expansions while maintaining high accuracy across k in [0,1). These techniques are implemented in libraries like the GNU Scientific Library (GSL), which employs AGM and series for real arguments with double precision; mpmath, which uses theta function q-series for arbitrary-precision arithmetic, supporting both real and complex evaluations; and SciPy, which utilizes Carlson and AGM methods for efficient computation.

Inverse Jacobi elliptic functions

The inverse Jacobi elliptic function corresponding to the sine amplitude function is defined via the incomplete elliptic integral of the first kind, denoted F(φ, k): F(\phi, k) = \int_0^\phi \frac{d\theta}{\sqrt{1 - k^2 \sin^2 \theta}}, where φ is the amplitude and 0 ≤ k < 1 is the elliptic modulus. This integral represents the value u such that sn(u, k) = sin φ, linking it directly to the forward Jacobi sine function. The relation to the amplitude function follows immediately: the inverse of the Jacobi amplitude am(u, k) = φ satisfies am^{-1}(φ, k) = F(φ, k). In the special case k = 0, F(φ, 0) = φ, while the inverse of sn(u, 0) = sin u reduces to the arcsine function, arcsin(x), for |x| ≤ 1. Numerical computation of these inverses often employs iterative methods for solving the underlying transcendental equations. Halley's method, which offers cubic convergence, is applied after an initial bisection refinement to invert the elliptic integral with respect to the amplitude φ, typically requiring only 2 iterations for high precision. Alternatively, series reversion techniques derive power series expansions for the inverses, such as for arcsp(z, k) (a form related to F), using symmetric elliptic integrals and homogeneous Legendre polynomials to achieve rapid convergence within specified radii, e.g., max{|Δ(p, q)|, |Δ(p, r)|} |z|^2 < 1. These inverse functions are essential for solving integrals that arise in amplitude determination, such as expressing the phase or argument in terms of accumulated elliptic integrals in periodic systems.

Applications

In map projections

Jacobi elliptic functions play a significant role in developing conformal map projections for ellipsoidal surfaces, enabling the preservation of local angles and shapes in cartographic representations of the Earth and other celestial bodies. Introduced in the 19th century through Carl Jacobi's lectures at Königsberg (1842–1843, published 1866), these functions provided a foundation for mapping irregular surfaces like triaxial ellipsoids, with renewed applications in the late 20th and early 21st centuries for asteroids such as 433 Eros and 25143 Itokawa. Early works by H.A. Schwarz (1864) and K. Weierstrass (1866) extended elliptic integrals to conformal mappings, while later adaptations by C.S. Peirce (1879) and A. Guyou (1887) incorporated Jacobi-inspired methods for polyhedral and rhombic projections of spheres and ellipsoids. In ellipsoidal cartography, Jacobi elliptic functions facilitate the transformation to auxiliary coordinates that maintain conformality. The authalic latitude \phi_2, an auxiliary latitude preserving surface area equivalence between the ellipsoid and a sphere, is given algebraically by \sin \phi_2 = \sqrt{1 - e^2} \sin \phi_1 / \sqrt{1 - e^2 \sin^2 \phi_1}, where e is the eccentricity and \phi_1 is the geodetic latitude; this does not directly involve Jacobi functions. However, the isometric latitude u, crucial for conformal mappings, is defined as the elliptic integral u = \int_0^{\phi_1} \frac{d \theta}{\sqrt{1 - e^2 \sin^2 \theta}} = F(\phi_1, e), and the geodetic latitude relates via \phi_1 = \am(u, e), with Jacobi sine \operatorname{sn}(u, e) = \sin \phi_1. The conformal latitude \chi, which ensures angle preservation, is derived from the isometric latitude as \chi = 2 \arctan \left( e^u \tan(\phi_1 / 2) \right) - \pi / 2 (approximate form; exact involves series or numerical inversion), parameterizing meridian arcs for projections. These latitudes allow ellipsoidal coordinates to be mapped conformally onto a plane or sphere. Projection coordinates in such systems are computed as x = \rho \cos \chi and y = \rho \sin \chi, where \rho is the distance from the projection center (derived from elliptic integrals of the first and third kinds) and \chi is the conformal latitude. For a triaxial ellipsoid defined by semi-axes a > b > c, the coordinates stem from elliptic integrals transformed via Jacobi functions: x = \int \frac{du}{\sqrt{1 - k_1^2 \sin^2 u}}, \quad y = \int \frac{dv}{(1 - k_2 \sin^2 v) \sqrt{1 - k_1^2 \sin^2 v}}, with moduli k_1 and k_2 determined by the ellipsoid's parameters. The Jacobi sine \operatorname{sn}(u, k) inverts these integrals to yield the angular components. These methods offer key advantages in modeling or triaxial Earth-like surfaces, as they preserve angles essential for and thematic while accommodating the ellipsoid's without excessive near poles or equators. For instance, the Jacobi conformal minimizes shape on small celestial bodies, producing maps suitable for geological analysis.

In physics and nonlinear dynamics

Jacobi elliptic functions provide exact solutions to the nonlinear governing the motion of a simple for large amplitudes, where the fails. The T of is given by T = 4 \sqrt{\frac{l}{g}} \, K(k), where l is the pendulum length, g is , k = \sin(\theta_0 / 2) with \theta_0 the maximum , and K(k) is the complete of the first kind. The angular displacement \theta(t) itself can be expressed as \theta(t) = 2 \operatorname{am}(u, k), where \operatorname{am} is the Jacobi amplitude function and u = \sqrt{g/l} \, t. This formulation captures the , with the increasing as \theta_0 approaches \pi. In nonlinear wave theory, Jacobi elliptic functions describe cnoidal waves, which are periodic analogs of solitons in shallow-water waves modeled by the Korteweg-de Vries (KdV) equation \partial_t \eta + c \partial_x \eta + \frac{3}{2} \eta \partial_x \eta + \frac{1}{6} h^2 \partial_x^3 \eta = 0, where \eta(x,t) is the water surface elevation, c is the linear wave speed, and h is the water depth. The exact solution is \eta(x,t) = \eta_0 + a \, \operatorname{cn}^2 \bigl( \kappa (x - c t), k \bigr), with a, k, \kappa, mean level \eta_0, and speed c related by dispersion relations derived from the equation parameters. As k \to 1, the cnoidal wave approaches a solitary wave, bridging periodic and solitary behaviors in integrable systems. This solution was first derived in the foundational work on long waves in rectangular canals. In nonlinear optics, Jacobi elliptic functions yield exact periodic solutions to the nonlinear Schrödinger equation (NLSE) i \partial_z \psi + \frac{1}{2} \partial_t^2 \psi + |\psi|^2 \psi = 0, which models pulse propagation in Kerr media with dispersion and self-phase modulation. These solutions, such as bright or dark periodic waves, take forms like \psi(z,t) = A \, \operatorname{cn}(B t, k) \, e^{i \phi(z)}, where parameters A, B, and phase \phi depend on the modulus k and initial conditions, enabling the study of modulation instability and optical rogue waves. Such elliptic representations extend beyond sech/tanh soliton limits, providing a framework for stable periodic structures in fiber optics and metamaterials. Jacobi elliptic functions also model dynamics in superconducting Josephson junctions, where the phase difference \phi across satisfies a pendulum-like \frac{\hbar C}{2e} \ddot{\phi} + \frac{\hbar}{2e R} \dot{\phi} + I_c \sin \phi = I, with C, resistance R, critical current I_c, and I. For underdamped cases near the plasma , the supercurrent is I(t) = I_c \, \operatorname{sn}(\omega t, k), where \omega is the and k reflects and levels. This snoidal form describes oscillations and fluxon motion, crucial for applications in SQUIDs and voltage standards. Jacobi elliptic functions appear in solutions to the one-dimensional Gross-Pitaevskii equation for repulsive Bose-Einstein condensates in traps, where the stationary density takes elliptic forms such as |\psi(x)|^2 \propto \operatorname{dn}^2(\kappa x, k).

References

  1. [1]
    DLMF: §22.2 Definitions ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    ### Mathematical Definitions of Jacobi Elliptic Functions
  2. [2]
    DLMF: §22.5 Special Values ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    ### Summary of Key Identities for Jacobi Elliptic Functions from https://dlmf.nist.gov/22.5
  3. [3]
    [PDF] A Brief History of Elliptic Functions
    Jan 24, 2019 · This thesis explores the history of elliptic functions, beginning with their ori- gins in elliptic integrals studied by Jacob Bernoulli and ...
  4. [4]
  5. [5]
    A primer on elliptic functions with applications in classical mechanics
    Nov 26, 2007 · They appear as solutions of many important problems in classical mechanics: the motion of a planar pendulum (Jacobi), the motion of a force-free ...
  6. [6]
    [PDF] Jacobi elliptic functions and the complete solution to the bead on the ...
    In this paper, we introduced Jacobi elliptic functions using integral inversion and discussed their properties and the meaning of their arguments ðu,kÞ. The ...
  7. [7]
    Euler's rigid rotators, Jacobi elliptic functions, and the Dzhanibekov ...
    Apr 1, 2021 · They are defined as geometrical ratios on a reference ellipse and developed geometrically without reference to power series or complex variables ...
  8. [8]
    22.15 Inverse Functions ‣ Properties ‣ Chapter 22 Jacobian Elliptic ...
    The inverse Jacobian elliptic functions can be defined in an analogous manner to the inverse trigonometric functions (§4.23).
  9. [9]
    Jacobi Elliptic Functions -- from Wolfram MathWorld
    The Jacobi elliptic functions are standard forms of elliptic functions. The three basic functions are denoted cn(u,k), dn(u,k), and sn(u,k), where k is known as
  10. [10]
    Elliptic Integral -- from Wolfram MathWorld
    Elliptic integrals can be viewed as generalizations of the inverse trigonometric functions and provide solutions to a wider class of problems.
  11. [11]
    22.19 Physical Applications
    Numerous other physical or engineering applications involving Jacobian elliptic functions, and their inverses, to problems of classical dynamics, electrostatics ...Missing: physics | Show results with:physics<|control11|><|separator|>
  12. [12]
    scipy.special.ellipj — SciPy v1.16.2 Manual
    Calculates the Jacobian elliptic functions of parameter m between 0 and 1, and real argument u. u array_like m array_like out tuple of ndarray, optional
  13. [13]
    acb_elliptic.h – elliptic integrals and functions of complex variables
    This module supports computation of elliptic (doubly periodic) functions, and their inverses, elliptic integrals.<|control11|><|separator|>
  14. [14]
    Abramowitz and Stegun. Page 567
    - **Notation Extracted from Abramowitz and Stegun, Page 567:**
  15. [15]
    Introduction to the Jacobi elliptic functions
    The modern notations for Jacobi functions were introduced later in the works of C. Gudermann (1838) (for functions , , and ) and J. Glaisher (1882) (for ...
  16. [16]
    ellipj - Jacobi elliptic functions - MATLAB - MathWorks
    This MATLAB function returns the Jacobi elliptic functions SN, CN, and DN evaluated for corresponding elements of argument U and parameter M.Description · Examples · Output Arguments · More About
  17. [17]
    Elliptic functions - MacTutor History of Mathematics
    The study of elliptical integrals can be said to start in 1655 when Wallis began to study the arc length of an ellipse.
  18. [18]
    22.2 Definitions ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    As a function of z , with fixed k , each of the 12 Jacobian elliptic functions is doubly periodic, having two periods whose ratio is not real. Each is ...
  19. [19]
    [PDF] arXiv:1201.4201v1 [physics.class-ph] 20 Jan 2012
    Jan 20, 2012 · In this section, we introduce Jacobi elliptic functions through elliptic geometry and integral inversion,5,12–16 and discuss the geometrical ...
  20. [20]
    DLMF: §20.15 Tables ‣ Computation ‣ Chapter 20 Theta Functions
    Tables of Neville's theta functions θ s ... n ⁡ ( x , q ) (see §20.1) and their logarithmic x -derivatives are given in Abramowitz and Stegun (1964, pp.
  21. [21]
    BOOK REVIEW Jacobian elliptic functions. By E. H. Neville. Oxford ...
    Jacobian elliptic functions. By E. H. Neville. Oxford University Press,. 1944. 16+331 pp. $7.50. In the development of the theory of elliptic functions ...
  22. [22]
    Elliptic functions — mpmath 1.3.0 documentation
    From a more modern point of view, an elliptic function is defined as a doubly periodic function, i.e. a function which satisfies. f ( z + 2 ω 1 ) = f ( z + 2 ω ...Missing: Whittaker Watson
  23. [23]
    DLMF: §22.6 Elementary Identities ‣ Properties ‣ Chapter 22 ...
    Annotations for §22.6 and Ch.22. Table 22.6.1: Jacobi's imaginary transformation of Jacobian elliptic functions. sn ⁡ ( i ⁢ z , k ) = i ⁢ sc ⁡ ( z , k ′ ) ...
  24. [24]
    [PDF] Evaluating Jacobi elliptic functions in the complex domain
    Nov 27, 2020 · Evaluating Jacobi elliptic functions in the complex domain ... We can then apply Jacobi's imaginary transformation [1, 8.152, 2nd line]:.
  25. [25]
    [PDF] Elliptic functions and Elliptic Integrals - UNCW
    Furthermore, we have the identities sn 2u + cn 2u = 1, k2 sn 2u + dn 2u = 1. Derivatives Derivatives of the Jacobi elliptic functions are easily found. First, ...
  26. [26]
    [PDF] A universal identity for theta functions of degree eight and applications
    Jacobi's triple product identity in the next theorem is one of the most fundamental results in the theory of elliptic theta functions and q-series, which can be ...
  27. [27]
    [PDF] Duality symmetries behind solutions of the classical simple pendulum
    (4), ii) the second option consists in solving the equation directly in terms of the imaginary time y. ... the Jacobi elliptic functions [2]. For instance ...
  28. [28]
    DLMF: §22.7 Landen Transformations ‣ Properties ‣ Chapter 22 ...
    dn ⁡ ( z , k ) : Jacobian elliptic function, z : complex, k : modulus, k 2 : change of variable and k 2 ′ : change of variable; A&S Ref: 16.14.4; Permalink ...Missing: imaginary | Show results with:imaginary
  29. [29]
  30. [30]
    [PDF] Generalized Landen Transformation Formulas for Jacobi Elliptic ...
    Landen transformation formulas, which connect Jacobi elliptic functions with different modulus pa- rameters, were first obtained over two hundred years ago ...
  31. [31]
    Jacobi Amplitude -- from Wolfram MathWorld
    The variable phi (also denoted am(u,k)) used in elliptic functions and elliptic integrals is called the amplitude (or Jacobi amplitude).
  32. [32]
    Jacobi Elliptic Functions
    The Jacobi elliptic functions are standard forms of Elliptic Functions. The three basic functions are denoted , , and , where is known as the Modulus. In terms ...
  33. [33]
    [PDF] Elliptic Modular Forms and Their Applications
    The special linear group SL(2, R) acts on H in the standard way by Möbius transformations (or fractional linear transformations): γ = a b. c d. : H → H , z → γz ...
  34. [34]
    Jacobi's Imaginary Transformation -- from Wolfram MathWorld
    Jacobi's imaginary transformations relate elliptic functions to other elliptic functions of the same type but having different arguments.Missing: derivation | Show results with:derivation
  35. [35]
    22.13 Derivatives and Differential Equations
    Table 22.13.1: Derivatives of Jacobian elliptic functions with respect to variable. ... Note that each derivative in Table 22.13.1 is a constant multiple of the ...Missing: ellipse | Show results with:ellipse
  36. [36]
    DLMF: §22.16 Related Functions ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    ### Summary of Relations to Theta Functions for Jacobi Elliptic Functions
  37. [37]
    DLMF: §22.4 Periods, Poles, and Zeros ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    ### Summary of Periods of Jacobi Elliptic Functions Related to Complete Elliptic Integral K(k)
  38. [38]
    DLMF: Chapter 22 Jacobian Elliptic Functions
    The references used for the mathematical properties in this chapter are Armitage and Eberlein (2006) , Bowman (1953) , Copson (1935) , Lawden (1989) , McKean ...
  39. [39]
  40. [40]
    DLMF: §22.8 Addition Theorems ‣ Properties ‣ Chapter 22 ...
    §22.8 Addition Theorems. ⓘ. Keywords: Jacobian elliptic functions, addition theorems; Permalink: http://dlmf.nist.gov/22.8; See also: Annotations for Ch.22 ...
  41. [41]
    DLMF: §15.9 Relations to Other Functions ‣ Properties ‣ Chapter ...
    Any hypergeometric function for which a quadratic transformation exists can be expressed in terms of associated Legendre functions or Ferrers functions.
  42. [42]
    22.14 Integrals ‣ Properties ‣ Chapter 22 Jacobian Elliptic Functions
    The indefinite integral of a 4th power can be expressed as a complete elliptic integral, a polynomial in Jacobian functions, and the integration variable.
  43. [43]
    DLMF: §29.2 Differential Equations ‣ Lamé Functions ‣ Chapter 29 ...
    Lamé's Equation where k and ν are real parameters such that 0 < k < 1 and ν ≥ − 1 2. For sn ( z , k ) see §22.2. This equation has regular singularities.
  44. [44]
    [PDF] Math 213a (Fall 2024) Yum-Tong Siu 1 ELLIPTIC FUNCTIONS ...
    Before we introduce the addition formula for the Jacobian elliptic sine function, we first discuss how we guess what it should be like by studying a special way.
  45. [45]
    22.11 Fourier and Hyperbolic Series
    Symbols: dn ⁡ ( z , k ) : Jacobian elliptic function, π : the ratio of the circumference of a circle to its diameter, K ⁡ ( k ) : Legendre's complete elliptic ...
  46. [46]
    DLMF: §20.7 Identities ‣ Properties ‣ Chapter 20 Theta Functions
    This reference also gives the eleven additional identities for the permutations of the four theta functions.
  47. [47]
    Elliptic functions — mpmath 1.2.0 documentation
    Feb 1, 2021 · The canonical elliptic functions are the Jacobi elliptic functions. More broadly, this section includes quasi-doubly periodic functions ...
  48. [48]
    None
    ### Summary of Series Expansions for Jacobian Elliptic Functions
  49. [49]
    [PDF] Evaluation of the Elliptic Integrals - Al-Azhar Bulletin of Science
    Elliptic integrals, recursive computations algorithms, continued fraction, trigonometric series expansions. ... F(k, π/2) ≡ K(k) ≡ K and E(k, π/2) ≡ E(k) ≡ E. (3).
  50. [50]
    [PDF] SOME CONTINUED FRACTION EXPANSIONS OF LAPLACE ...
    In a 1907 paper, L. Rogers used two methods to obtain continued fractions for cer- tain Laplace transforms of Jacobi elliptic functions.
  51. [51]
    22.20 Methods of Computation
    Jacobi's epsilon function can be computed from its representation (22.16.30) in terms of theta functions and complete elliptic integrals; compare §20.14.
  52. [52]
    [PDF] The Arithmetic-Geometric Mean and Fast Computation of ...
    Apr 23, 2003 · tal function is the arithmetic-geometric mean iteration of Gauss and Legendre for computing complete elliptic integrals. This is where we ...
  53. [53]
    [PDF] Numerical Evaluation of Elliptic Functions, Elliptic Integrals ... - arXiv
    Jun 18, 2018 · The inverses of Jacobi's elliptic functions can be com- puted similarly, but at this time they are not implemented in Arb. Carlson gives ...
  54. [54]
    Fast computation of complete elliptic integrals and Jacobian elliptic ...
    Oct 25, 2009 · We created a fast method to calculate the complete elliptic integral of the first and second kinds, K(m) and E(m), for the standard domain of the elliptic ...<|control11|><|separator|>
  55. [55]
    Fast computation of Jacobian elliptic functions and incomplete ...
    The new method is around 25 times faster than the method using the incomplete elliptic integral of general kind and around 70 times faster than the method using ...
  56. [56]
    Numerical computation of real or complex elliptic integrals
    Algorithms for numerical computation of symmetric elliptic integrals of all three kinds are improved in several ways and extended to complex values of the.
  57. [57]
    Special Functions — GSL 2.8 documentation - GNU.org
    The Jacobian Elliptic functions are defined in Abramowitz & Stegun, Chapter 16. The functions are declared in the header file gsl_sf_elljac.h . int ...
  58. [58]
    Numerical Inversion of General Incomplete Elliptic Integral
    Aug 6, 2025 · We present a numerical method to invert a general incomplete elliptic integral with respect to its argument and/or amplitude.Missing: reversion | Show results with:reversion
  59. [59]
    power series for inverse Jacobian elliptic functions
    Dec 11, 2007 · POWER SERIES FOR INVERSE JACOBIAN ELLIPTIC FUNCTIONS. 1617. For a proof of the stronger statement that each inequality involves the radius of.<|control11|><|separator|>
  60. [60]
    (PDF) Jacobi Conformal Projection of the Triaxial Ellipsoid
    Sep 30, 2016 · In this paper a new technique for recalculating geographic coordinates of a triaxial ellipsoid to elliptical and then to rectangular ...<|separator|>
  61. [61]
    [PDF] Elliptic functions applied to conformal world maps. Department of ...
    Jan 1, 2006 · SOME NUMERICAL VALUES OB TEE ABELIAN FUNCTIONS. We have already seen that with g(0) - 1 g(@ -etkK, h(K) = u (R) ==e-, g(-. '0,. 9(2m =o, h(2K) ...
  62. [62]
    [PDF] solving the nonlinear pendulum equation
    In this paper we have derived a closed-form solution for the angular displace- ment of an ideal simple pendulum in terms of the Jacobi elliptic function sn ...
  63. [63]
    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 ...
  64. [64]
    Stationary solutions for the 1 + 1 nonlinear Schrödinger equation ...
    Jul 29, 2013 · We shall frequently revisit the K 2 = 1 case, as it will greatly simplify some calculations. Such solutions, based around the tanh solution, ...
  65. [65]
    Exact analytical solution of the problem of current-carrying states of ...
    Nov 20, 2007 · The obtained solutions describe the current-carrying states of the Josephson junction of arbitrary length W ≡ 2 L ∊ ( 0 , ∞ ) in the presence of ...