Fact-checked by Grok 2 weeks ago

Kirchhoff equations

The Kirchhoff equations are a system of six first-order ordinary differential equations that govern the translational and rotational motion of a rigid body immersed in an unbounded ideal fluid, characterized as incompressible, inviscid, and at rest at infinity. These equations couple the rigid body's dynamics with the fluid's hydrodynamic response, where the fluid motion is irrotational and determined solely by the body's velocity, leading to added mass and inertial effects without dissipative forces. Formulated by the German physicist Gustav Robert Kirchhoff, they represent a foundational model in theoretical fluid dynamics and rigid body mechanics. In their , the Kirchhoff equations are expressed in terms of the body's linear \mathbf{p} and \mathbf{M} relative to a fixed point, with the H(\mathbf{M}, \mathbf{p}) representing the total of the body-fluid system as a positive . The equations read: \dot{\mathbf{p}} = \mathbf{p} \times \boldsymbol{\omega}, \quad \dot{\mathbf{M}} = \mathbf{M} \times \boldsymbol{\omega} + \mathbf{p} \times \mathbf{u}, where \boldsymbol{\omega} = \nabla_{\mathbf{M}} H and \mathbf{u} = \nabla_{\mathbf{p}} H are the angular and linear velocities, respectively, and external forces or torques can be added for more general cases. This structure endows the system with a Hamiltonian formulation on the e(3)^* of the , preserving key integrals such as the of linear |\mathbf{p}|^2 and the scalar product \mathbf{M} \cdot \mathbf{p}. Kirchhoff's original derivation appeared in his lectures on mathematical physics, building on potential flow theory to compute hydrodynamic loads via the added-mass tensor, which depends on the body's geometry. The model assumes no circulation or vorticity in the fluid unless introduced externally, making it ideal for analyzing inertial motions like falling or rotating bodies in water or air without friction. Extensions of the equations have since incorporated gravity, circulation, deformable bodies, and viscous effects, finding applications in marine engineering, aerospace, and numerical simulations of fluid-structure interactions. Despite their idealizations, the Kirchhoff equations remain a benchmark for understanding coupled body-fluid dynamics and reveal rich behaviors, including integrable cases and chaotic regimes under perturbations.

Background Concepts

Rigid Body Motion

A in is defined as a system of material points such that the distances between any pair of points remain invariant under the action of applied forces, ensuring no relative deformation occurs. This rigid constraint limits the configuration space to : three translational degrees corresponding to the position of the center of mass and three rotational degrees describing the body's in space. The motion of a is analyzed using two coordinate systems: an inertial reference frame fixed in space, in which Newton's laws hold without fictitious forces, and a body-fixed frame attached to the body that rotates with it. The relative orientation between these frames is captured by an orthogonal R, which maps vectors from the body frame to the inertial frame and satisfies R^T R = I with \det(R) = 1, preserving lengths and orientations. The translational dynamics of the rigid body are encapsulated in the linear momentum equation for the center of mass: m \dot{\mathbf{v}} = \mathbf{F}, where m is the total mass, \mathbf{v} is the velocity of the center of mass in the inertial frame, and \mathbf{F} is the resultant external force acting on the body. This follows directly from integrating Newton's second law over the body's mass distribution. Rotational dynamics in the body-fixed frame, assuming principal axes alignment, are governed by Euler's equations: \mathbf{I} \dot{\boldsymbol{\omega}} + \boldsymbol{\omega} \times (\mathbf{I} \boldsymbol{\omega}) = \boldsymbol{\Gamma}, where \mathbf{I} is the inertia tensor (diagonal in principal coordinates), \boldsymbol{\omega} is the angular velocity vector, and \boldsymbol{\Gamma} is the external torque vector, all expressed in the body frame. These equations arise from the time derivative of angular momentum in the rotating frame and were first systematically derived by Leonhard Euler in his foundational work on rigid body motion.

Ideal Fluid Dynamics

In the context of Kirchhoff equations, an fluid is characterized by three key assumptions: it is inviscid, meaning it experiences no shear stresses or ; incompressible, satisfying the \nabla \cdot \mathbf{u} = 0; and irrotational, with \nabla \times \mathbf{u} = 0, which permits the introduction of a \phi such that \mathbf{u} = \nabla \phi and the potential satisfies \nabla^2 \phi = 0. These properties simplify the governing equations for the fluid surrounding a moving , allowing the focus to remain on inertial and forces without dissipative effects. The boundary conditions for the fluid potential are essential to couple the fluid motion with the . On the surface of the body, the no-penetration condition requires that the normal component of the fluid velocity matches the normal component of the body's velocity: \mathbf{u} \cdot \mathbf{n} = \mathbf{v} \cdot \mathbf{n}, or equivalently \frac{\partial \phi}{\partial n} = \mathbf{v} \cdot \mathbf{n}. Far from the body, at , the fluid is assumed to be at rest, so \mathbf{u} \to 0 and \phi \to 0. These conditions ensure that the fluid does not cross the body surface while decaying appropriately in the unbounded domain. For unsteady irrotational flow, the pressure in the ideal fluid is governed by the integrated form of the Euler equations, known as Bernoulli's equation: \frac{\partial \phi}{\partial t} + \frac{1}{2} |\nabla \phi|^2 + \frac{p}{\rho} + g z = F(t), where p is the pressure, \rho is the constant fluid density, g is the gravitational acceleration, z is the vertical coordinate, and F(t) is an arbitrary function of time. This equation relates the unsteady potential, kinetic energy per unit mass, pressure head, and gravitational potential, providing the pressure distribution necessary for computing hydrodynamic forces on the body. The kinetic energy of the ideal fluid, T_f, can be expressed in terms of the velocity potential using the divergence theorem, reducing the volume integral over the infinite fluid domain to a surface integral over the body: T_f = -\frac{1}{2} \rho \int_S \phi \frac{\partial \phi}{\partial n} \, dS, where the integration is performed over the body surface S. This form highlights the dependence of the fluid's energy on the body's motion through the boundary values of \phi and its normal derivative, establishing a foundation for variational approaches in the coupled body-fluid system.

Historical Development

Kirchhoff's Original Formulation

In 1876, presented the foundational equations for the motion of a immersed in an ideal fluid during his lectures on at the University of Berlin, which were subsequently published as the first volume of Vorlesungen über Mathematische Physik: Mechanik. These lectures, delivered over the 1876–1877 academic year, marked a significant advancement in theoretical by integrating with rigid body motion. Kirchhoff's approach built upon earlier developments in hydrodynamics, including variational principles introduced by Alfred Clebsch in the 1850s, extending the framework to coupled systems where the fluid's response influences the body's trajectory. Central to Kirchhoff's formulation was the extension of Hamilton's to encompass both the and the surrounding fluid, treating the total of the system as the key quantity in deriving the dynamics. This variational motivation allowed for a systematic of the equations, emphasizing conservation laws inherent in the ideal fluid setting. The resulting equations express the body's through time derivatives of partial derivatives of the system's with respect to linear and angular velocities, capturing the interplay between inertial forces and fluid-induced effects. Kirchhoff's model rested on several key assumptions to simplify the complex : the fluid is infinite in extent, irrotational (with zero ), incompressible, and inviscid (non-viscous), remaining at rest far from the body. Additionally, the rigid body's motion occurs without —where vapor bubbles form due to low —or the of surfaces, ensuring the fluid fully wets the body and maintains continuity. These idealizations facilitated analytical tractability while highlighting fundamental hydrodynamic influences on .

Extensions by Clebsch and Others

Alfred Clebsch developed a for the steady motion of an incompressible fluid in his 1859 paper, building on his earlier 1857 work, which introduced potentials to describe fluid motion without relying solely on Euler's equations. This approach provided a framework for deriving the through a formulation, emphasizing the role of Clebsch potentials in representing the velocity field. Clebsch's contributions laid important groundwork for later formulations, including Kirchhoff's. In the late 19th century, this variational method was adapted to extend Kirchhoff's original equations for motion in ideal fluids to cases involving rotational flows. The adaptation incorporated Clebsch potentials \psi and \chi to account for , allowing the velocity field to be expressed as \mathbf{u} = \nabla \phi + \psi \nabla \chi, where \phi is the for the irrotational part and the curl term introduces \boldsymbol{\omega} = \nabla \psi \times \nabla \chi. This Kirchhoff-Clebsch form generalized the framework to more complex fluid states while preserving the variational structure. Helmholtz's 1858 work on vortex theorems and the conservation of circulation influenced these extensions by providing a of the velocity into irrotational and vortical components, which aligned with the use of Clebsch potentials in the Kirchhoff framework. In the 20th century, Horace Lamb's 1932 treatise on hydrodynamics further developed these ideas, particularly in analyzing the stability of rigid body motions in fluids with , incorporating the extended Kirchhoff equations to study perturbations and equilibrium configurations.

Mathematical Formulation

Kinetic Energy and Lagrangian

The total kinetic energy T of a rigid body immersed in an ideal incompressible fluid is the sum of the body's kinetic energy T_b and the fluid's kinetic energy T_f. The body's contribution is given by T_b = \frac{1}{2} m \| \mathbf{v} \|^2 + \frac{1}{2} \boldsymbol{\omega}^T \tilde{I} \boldsymbol{\omega}, where m is the body's , \mathbf{v} is its linear of of mass, \boldsymbol{\omega} is its , and \tilde{I} is the inertia tensor relative to the center of mass. The fluid's kinetic energy, assuming irrotational flow, is expressed using the \phi that satisfies in the exterior fluid domain V: T_f = \frac{1}{2} \rho \int_V \| \nabla \phi \|^2 \, dV, where \rho is the fluid density. By applying Green's second identity and the boundary conditions on the body surface S, this reduces to the equivalent T_f = -\frac{1}{2} \rho \int_S \phi \frac{\partial \phi}{\partial n} \, dS. The potential \phi is determined by the body's motion and vanishes at infinity, ensuring the fluid's contribution captures the hydrodynamic inertia. In the Kirchhoff framework, the total T = T_b + T_f is formulated in terms of associated with the body's motion, typically the linear velocity \mathbf{v} and \boldsymbol{\omega} in a body-fixed . This yields a for the L, which for the ideal case without approximates L \approx T when V is negligible (e.g., for motion where effects are absent or balanced). In general, L = T - V, with V accounting for if relevant. The explicit expression in body-fixed coordinates is L = \frac{1}{2} (\mathbf{A} \boldsymbol{\omega}, \boldsymbol{\omega}) + (\mathbf{B} \boldsymbol{\omega}, \mathbf{v}) + \frac{1}{2} (\mathbf{C} \mathbf{v}, \mathbf{v}) + (\mathbf{k}, \boldsymbol{\omega}) + (\mathbf{l}, \mathbf{v}), where \mathbf{A}, \mathbf{B}, and \mathbf{C} are symmetric matrices representing the added rotational , cross-coupling, and added translational , respectively; these depend solely on the body's and the properties. The linear terms (\mathbf{k}, \boldsymbol{\omega}) and (\mathbf{l}, \mathbf{v}) arise in extensions involving circulation around the body or non-zero far-field flow. The in the Kirchhoff framework are derived variationally using , which states that the is stationary: \delta \int_{t_1}^{t_2} L \, dt = 0, for admissible variations in the body's and that vanish at the endpoints t_1 and t_2. This incorporates the coupled body-fluid dynamics through the expressions, leading to the conservation of total linear and in the absence of external forces.

Equations of Motion

The core Kirchhoff equations in quasi-Lagrangian form describe the coupled translational and rotational of a immersed in an ideal incompressible fluid, derived from the total T of the body-fluid system. These equations are expressed as \frac{d}{dt} \left( \frac{\partial T}{\partial \omega} \right) = \frac{\partial T}{\partial \omega} \times \omega + \frac{\partial T}{\partial v} \times v + Q_h + Q, \frac{d}{dt} \left( \frac{\partial T}{\partial v} \right) = \frac{\partial T}{\partial v} \times \omega + F_h + F, where \mathbf{v} and \boldsymbol{\omega} denote the linear and angular velocity vectors of the body in the body-fixed frame, \mathbf{Q} and \mathbf{F} represent external torque and force vectors, and \mathbf{Q}_h and \mathbf{F}_h are the corresponding hydrodynamic contributions from fluid pressure integrated over the body surface. The left-hand sides correspond to the time derivatives of the angular and linear momenta conjugate to \boldsymbol{\omega} and \mathbf{v}, while the cross-product terms on the right-hand sides capture the geometric coupling between rotational and translational motions, manifesting as Coriolis-like effects in the non-inertial body frame. The hydrodynamic terms \mathbf{Q}_h and \mathbf{F}_h depend quadratically on the velocities through the added mass and inertia contributions encoded in T. For fully submerged bodies where external forces and torques are absent or exactly balanced (e.g., by Archimedean in a uniform ), the Kirchhoff-Clebsch variant yields a self-contained system by absorbing the hydrodynamic effects directly into the functional. In this case, the equations simplify to \frac{d}{dt} \left( \frac{\partial L}{\partial \boldsymbol{\omega}} \right) = \frac{\partial L}{\partial \boldsymbol{\omega}} \times \boldsymbol{\omega} + \frac{\partial L}{\partial \mathbf{v}} \times \mathbf{v}, \frac{d}{dt} \left( \frac{\partial L}{\partial \mathbf{v}} \right) = \frac{\partial L}{\partial \mathbf{v}} \times \boldsymbol{\omega}, with L representing the augmented kinetic energy that includes both the body's intrinsic inertia and the fluid's added mass tensor. These represent the Euler-Lagrange equations for a Lagrangian L = T (kinetic energy only, absent potential terms in the ideal fluid approximation). Solving these nonlinear ordinary differential equations for the six degrees-of-freedom (6-DOF) motion typically requires techniques, such as Runge-Kutta methods, with the and cross-coupling inertia matrices precomputed analytically or via boundary-element methods for the given body geometry. This approach enables simulation of the body's trajectory and orientation, assuming irrotational fluid flow and neglecting .

Hydrodynamic Interactions

Added Mass and Inertia Tensor

In the context of a moving through an ideal incompressible , the , also known as virtual mass, represents the additional inertia imparted to the body due to the acceleration of the surrounding . This effect arises because the body's motion induces a in the that must be accounted for in the dynamics, effectively increasing the body's inertial response as if its mass were augmented. In Kirchhoff's formulation, this added inertia is captured through tensorial quantities derived from theory, ensuring the reflect the coupled body-fluid interaction without or . For translational motion, the hydrodynamic force component due to takes the form \mathbf{F}_h = -m_a \dot{\mathbf{v}} - \boldsymbol{\omega} \times m_a \mathbf{v}, where m_a is the added mass scalar, \dot{\mathbf{v}} is the , and \boldsymbol{\omega} is the ; the second accounts for the convective in a rotating . In general, m_a is replaced by the added mass tensor \mathbf{M}_a, a symmetric 3×3 whose depend on the body's geometry and the fluid density \rho, yielding F_{h,j} = -\dot{U}_i (M_a)_{ij} - \varepsilon_{jkl} U_i \Omega_k (M_a)_{li}, where \mathbf{U} is the velocity vector and \varepsilon_{jkl} is the . For a sphere of volume V, the tensor is isotropic with m_a = \frac{1}{2} \rho V, meaning the added mass equals half the displaced fluid mass in any direction. The rotational analog involves an added inertia tensor \mathbf{I}_a, such that the total inertia experienced by the body is \tilde{\mathbf{I}} = \mathbf{I}_b + \mathbf{I}_a, where \mathbf{I}_b is the body's intrinsic tensor about its . The fluid contribution \mathbf{I}_a appears in the as a \frac{1}{2} (\mathbf{A} \boldsymbol{\omega}, \boldsymbol{\omega}), with \mathbf{A} being the symmetric added rotational tensor, whose elements scale with \rho times a geometric factor involving the body's and shape. For symmetric bodies aligned with principal axes, \mathbf{I}_a is diagonal, simplifying the rotational dynamics. Coupling between translation and rotation introduces cross terms via a tensor \mathbf{B}, contributing to the kinetic energy as (\mathbf{B} \boldsymbol{\omega}, \mathbf{v}); these terms are zero for bodies with sufficient symmetry, such as spheres or ellipsoids aligned with their principal axes, but become non-zero for asymmetric geometries, leading to off-diagonal elements in the overall 6×6 inertia matrix that mix linear and angular motions. These tensors are computed using by solving the Laplace equation \nabla^2 \phi = 0 in the domain exterior to the body, subject to the \frac{\partial \phi}{\partial n} = \mathbf{u} \cdot \mathbf{n} on the body surface, where \mathbf{u} is a unit in the of interest and \mathbf{n} is the outward normal. The tensor components are then obtained from surface integrals, such as (M_a)_{ij} = -\rho \oint_S \phi^{(j)} n_i \, dS, where \phi^{(j)} is the potential for unit in the j-. For ellipsoids with semi-axes a, b, c, closed-form expressions exist for the diagonal elements; for instance, the translational coefficient along the a-axis is A_{11} = \frac{\alpha_0}{2 - \alpha_0}, where \alpha_0 is an depending on the aspect ratios, yielding explicit values like A_{11} = 0.5 for a . These appear in the as additional terms from the .

Forces and Torques from Fluid Pressure

In ideal fluid dynamics, the hydrodynamic \mathbf{F}_h acting on a submerged arises from the integral of the p over the body surface S, given by \mathbf{F}_h = -\int_S p \mathbf{n} \, dS, where \mathbf{n} is the outward to the surface. Similarly, the hydrodynamic \mathbf{Q}_h about the body's is \mathbf{Q}_h = -\int_S p \mathbf{r} \times \mathbf{n} \, dS, with \mathbf{r} denoting the position vector from the center of mass to the surface element. These expressions follow from the fundamental property that, in an inviscid , the only stresses are pressures, neglecting viscous . The pressure p on the surface is determined from the unsteady Bernoulli equation for irrotational flow, where the \phi satisfies \nabla^2 \phi = 0 in the fluid domain exterior to the body. Neglecting gravity, the pressure takes the form p = -\rho \left( \frac{\partial \phi}{\partial t} + \frac{1}{2} |\nabla \phi|^2 \right), with \rho as the constant fluid density; this equation integrates the Euler equations along streamlines and holds instantaneously throughout the fluid. Substituting this into the force and torque integrals yields \mathbf{F}_h = \rho \int_S \left( \frac{\partial \phi}{\partial t} + \frac{1}{2} |\nabla \phi|^2 \right) \mathbf{n} \, dS, \quad \mathbf{Q}_h = \rho \int_S \left( \frac{\partial \phi}{\partial t} + \frac{1}{2} |\nabla \phi|^2 \right) \mathbf{r} \times \mathbf{n} \, dS. This formulation allows a natural of the hydrodynamic loads into unsteady and steady contributions. The unsteady term, -\rho \int_S \frac{\partial \phi}{\partial t} \mathbf{n} \, dS, corresponds to forces and torques proportional to the body's linear and angular accelerations, manifesting as effects that augment the body's inertia. The steady term, \rho \int_S \frac{1}{2} |\nabla \phi|^2 \mathbf{n} \, dS, arises from the convective acceleration in the Euler equations and reflects quadratic dependencies on the body's velocity and through the potential \phi. For irrotational, in an unbounded domain, the potential \phi is uniquely determined by the instantaneous motion (translation and ), satisfying the no-penetration boundary condition \mathbf{n} \cdot \nabla \phi = \mathbf{v} \cdot \mathbf{n} on S, where \mathbf{v} is the at each point. Consequently, both \mathbf{F}_h and \mathbf{Q}_h are fully specified by the body's alone, without external influences, their incorporation as external terms in the Kirchhoff equations of motion. A notable special case occurs for steady, uniform translation of the body in an , where the symmetrizes such that \mathbf{F}_h = 0, a result known as D'Alembert's paradox that highlights the absence of drag in potential .

Applications and Limitations

Modeling Submerged Rigid Bodies

The Kirchhoff equations for modeling fully submerged rigid bodies assume an ideal fluid that is incompressible, irrotational, and inviscid, with no effects due to deep submergence. Under these conditions, the and added inertia tensors remain constant, as the body's geometry is fixed and independent of its pose or motion. This simplifies the hydrodynamic interactions to linear terms in the velocity and acceleration, allowing the equations to capture the coupled translational and rotational dynamics in (6-DOF) without time-varying fluid contributions from waves or boundaries. A representative example is a translating in a uniform flow, where leads to decoupled translational and rotational equations. For translation, the is half the displaced fluid mass, given by m_a = \frac{1}{2} \rho \frac{4}{3} \pi r^3, where \rho is the fluid density and r is the ; the rotational added tensor is isotropic and similarly scales with the displaced volume. In this case, the Kirchhoff equations reduce to independent ordinary differential equations (ODEs) for linear velocity \mathbf{v} and \boldsymbol{\omega}, with the effective incorporating the constant added mass to predict steady-state drift and spin without cross-coupling. Numerical solutions of the Kirchhoff equations for 6-DOF trajectories typically employ time-stepping integrators such as explicit Runge-Kutta methods to handle the nonlinear coupling between , , and orientation. These schemes advance the (including \mathbf{\eta}, \boldsymbol{\nu}, and quaternion representation for rotation) over discrete time steps, with added mass tensors precomputed via boundary element methods for arbitrary shapes. Adaptations of open-source CFD software like have been used to estimate these tensors in solvers before integrating the ODEs, enabling efficient simulation of complex maneuvers for underwater vehicles. Experimental validation of the Kirchhoff model for submerged bodies often involves towing tank tests at high Reynolds numbers (or as high as feasible in model scale) to approximate inviscid conditions by minimizing the relative influence of viscous effects. In such setups, a scaled is towed at controlled speeds, and measured trajectories are compared to numerical predictions, showing good agreement in added mass-dominated regimes for bodies like towed underwater vehicles. These comparisons confirm the model's accuracy for predicting , , and yaw motions, with discrepancies primarily arising from unmodeled at higher speeds.

Extensions to Viscous and Rotational Flows

The inviscid assumption underlying the original Kirchhoff equations leads to significant limitations in real flows, particularly at high Reynolds numbers where viscous effects dominate despite the flow appearing nearly inviscid. This results in , predicting zero drag on a moving steadily through the , which contradicts experimental observations of substantial form and . For surface-piercing , the model also fails to capture wave generation and radiation, as it neglects dynamics and the resulting . To incorporate viscous effects, extensions couple the Kirchhoff framework with the Navier-Stokes equations, often through boundary layer approximations or full viscous flow solvers that resolve the wake and separation regions. This approach allows simulation of body-fluid interactions in time-dependent viscous flows, capturing added-mass forces alongside viscous drag and lift, as demonstrated in numerical studies of ellipsoidal bubbles at Reynolds numbers up to 3000, where wake instabilities drive path deviations. For low-speed slender bodies, such as underwater vehicles or offshore structures, the Morison equation provides a practical viscous extension by combining the inertial term from potential flow added mass (derived from Kirchhoff-like formulations) with a quadratic drag term proportional to relative velocity squared, enabling efficient prediction of total hydrodynamic loads without full viscous resolution. The drag coefficient in this equation is empirically tuned based on body geometry and flow conditions, addressing the inviscid model's inability to model dissipation. For rotational flows with , Clebsch potentials \psi and \chi are introduced to represent the velocity field as \mathbf{u} = \nabla \phi + \psi \nabla \chi, where \phi is the potential and the curl term \nabla \times \mathbf{u} = \nabla \psi \times \nabla \chi encodes . This modifies the in the Kirchhoff formulation to include terms accounting for circulation and while preserving the variational structure. Such extensions are particularly useful for analyzing impulsive starts or unsteady motions where external interacts with the body, extending the irrotational assumption. Extensions to free surface effects address wave-body interactions by incorporating linearized boundary conditions into the , such as the combined condition \frac{\partial^2 \phi}{\partial t^2} + [g](/page/Gravity) \frac{\partial \phi}{\partial z} = 0 at z=0, where g is . The Froude-Krylov forces arise from the integral of undisturbed incident pressure over the body's wetted surface, providing the excitation component in ship motion equations, while and terms handle scattered . For ships with forward speed, the Korsmeyer-Wu method employs a three-dimensional approach with a low-order Green function to solve the efficiently, capturing unsteady forces including wave-making in restricted waters or multi-body interactions. Modern developments since 2000 integrate Kirchhoff-based potential flow with (CFD) solvers for hybrid simulations of viscous, multiphase flows around moving bodies. These approaches use immersed boundary methods or overset grids to couple inviscid exterior solutions with viscous near-body regions, improving accuracy for high-Reynolds-number applications like or platforms while reducing computational cost compared to full Navier-Stokes resolutions.

References

  1. [1]
    [PDF] periodic solutions of kirchhoff's equations for
    In the Kirchhoff problem on the motion of a rigid body in an ideal incompressible fluid it would be useful to analyze the stability of spatial motions in τ3 ...
  2. [2]
    [PDF] A new solving procedure for the Kelvin–Kirchhoff equations in case ...
    1. Introduction, equations of motion. Kelvin–Kirchhoff equations describe the dynamics of rigid body motion in an ideal fluid [Kirchhoff, 1877] ...<|control11|><|separator|>
  3. [3]
    Generalized Kirchhoff equations for a deformable body moving in a ...
    The classical Kirchhoff's method provides an efficient way of calculating the hydrodynamical loads (forces and moments) acting on a rigid body moving with ...
  4. [4]
    Rigid Body Motion - an overview | ScienceDirect Topics
    A rigid body is an object in which the distance between any two points remains the same. By definition, a rigid body does not deform, or change shape. In this ...Missing: authoritative | Show results with:authoritative
  5. [5]
    [PDF] Euler's Equations - 3D Rigid Body Dynamics - MIT OpenCourseWare
    We now turn to the task of deriving the general equations of motion for a three-dimensional rigid body. These equations are referred to as Euler's equations ...
  6. [6]
    [PDF] Newton-Euler Equations
    The Euler Equation ~M = P ~MG = IG~↵ represents the rotation motion of the body about its center of mass. In other words, the rotational motion of the rigid ...
  7. [7]
    (PDF) Leonhard Euler and the mechanics of rigid bodies
    Oct 28, 2025 · In this work we present the original ideas and the construction of the rigid bodies theory realised by Leonhard Euler between 1738 and 1775.
  8. [8]
    [PDF] Hydrodynamics
    ... Kirchhoff's elliptic vortex 250. 159. Vortices in a curved stratum of fluid ... Sitzungsberichte, t. lii., p. 214 (1865). t Berl. Monatsler., July 19 ...
  9. [9]
    Some remarks on the integrability of the equations of motion of a ...
    G. Kirchhoff, Vorlesungen über Mathematische Physik, Vol. 1, Mechanik, Teubner, Leipzig (1876). · A. Clebsch, Math. Ann.,3, 238–262 (1871). · V. Steklov, Math.
  10. [10]
    A few remarks about integrability of the equations of motion of a rigid ...
    [1]. G. Kirchhoff. Vorlesungen über mathematische Physik. Mechanik, 1, Teubner, Leipzig (1876) · [2]. L. Dirichlet. Monatsber. Berliner Acad., 30 (1852), p. · [3].
  11. [11]
    [PDF] control of underwater vehicles in inviscid fluids. i: irrotational ... - HAL
    In most of the papers devoted to that issue, the fluid is assumed to be inviscid, incompressible and irrotational, and the rigid body. (the vehicle) is ...
  12. [12]
    (PDF) Generalization of the Kelvin-Kirchhoff equations for the motion ...
    Aug 6, 2025 · The classical Kelvin–Kirchhoff equations of motion of a rigid body in an inviscid, unbounded fluid are generalized for the case where other ...Missing: standard | Show results with:standard
  13. [13]
    Comment on Clebsch's 1857 and 1859 papers on using ...
    Aug 2, 2021 · The purpose of Clebsch in his 1857 paper is to find a variational principle characterising the steady motion of an incompressible fluid. In his ...
  14. [14]
    Application of Clebsch variables to fluid-body interaction in ...
    Jul 31, 2014 · The method is based on the decomposition of the velocity field into vortical and potential components making use of Clebsch variables.
  15. [15]
    [PDF] Stability and instability in nineteenth-century fluid mechanics
    The stability or instability of a few basic flows was conjectured, debated, and sometimes proved in the nineteenth century.<|control11|><|separator|>
  16. [16]
    [PDF] Ueber die Bewegung eines Rotationskörpers in einer Flüssigkeit.
    Titel: Ueber die Bewegung eines Rotationskörpers in einer Flüssigkeit. Autor: Kirchhoff, G. Jahr: 1869.Missing: Gleichgewicht schwimmenden
  17. [17]
    [PDF] 23 APPENDIX 2: ADDED MASS VIA LAGRANGIAN DYNAMICS
    The Kirchoff relations follow directly from combining the three action variations. The kinetic co-energy we developed is that of the fluid, and so the ...Missing: quadratic | Show results with:quadratic
  18. [18]
    [PDF] Newton–Euler, Lagrange and Kirchhoff formulations of rigid body ...
    Jul 1, 2016 · Kirchhoff equations are a useful tool in rigid body dy- namics: they are well known and widely used in many fields of applied mathematics, ...
  19. [19]
    [PDF] Chaos and Integrability in Ideal Body-Fluid Interactions
    Sep 8, 2011 · The form of these equations are derived by combining the classical equations for free body motion in ideal fluid of G. Kirchhoff and Lord Kelvin ...
  20. [20]
    [PDF] 2.016 Hydrodynamics Added Mass
    Added mass forces can arise in one direction due to motion in a different direction, and thus we can end up with a 6 x 6 matrix of added mass coefficients.Missing: Kirchhoff | Show results with:Kirchhoff
  21. [21]
    [PDF] Maneuvering Hydrodynamics of Ellipsoidal Forms
    It shou1d be also mentioned that for e11ipsoida1 bodies there exist c1osed-form express ions for the three trans1a- tory added-mass ... (Lamb 1932, p. 153 ...
  22. [22]
    Attitude & Position Tracking of a Underwater Rigid Body
    Underwater dynamics of rigid bodies are modelled using the 'Kirchhoff Tensor', assuming the fluid to be irrotational, inviscid and incompressible.Missing: submerged assumptions
  23. [23]
    (PDF) Underwater Rigid Body Dynamics - ResearchGate
    Aug 9, 2025 · We show that the motion of rigid bodies under water can be realistically simulated by replacing the usual inertia tensor and scalar mass by ...
  24. [24]
    [PDF] 2.016 Hydrodynamics 0.1 Derivation of Added Mass around a Sphere
    Since force equals mass times acceleration, we can think of the additional force in terms of an imaginary added mass of the object in the fluid.Missing: Kirchhoff | Show results with:Kirchhoff
  25. [25]
    Modelling and Manoeuvrability Design of Autonomous Underwater ...
    Software solves the nonlinear secondary differential equation by implementing numerical method solution using explicit Runge-Kutta formulations. The results ...<|separator|>
  26. [26]
    Experimental investigation on a two-part underwater towed system
    Aug 6, 2025 · Measurements of towed body motion in a laboratory towing tank have validated the mathematical model used as a basis for predicting body and ...
  27. [27]
    Numerical Simulation of Two-Part Underwater Towing System
    A lumped mass-spring system was used for modeling the cable. The formulation for the towed body was based on Kirchhoff's equations for motion of a rigid body.
  28. [28]
    D'Alembert's Paradox | SIAM Review
    The problem of two-dimensional incompressible laminar flow past a bluff body at large Reynolds number ( R ) is discussed. The governing equations are the Navier ...
  29. [29]
    [PDF] Development of Simplified Formula for Froude-Krylov Force - ClassNK
    The linear Froude-Krylov force is defined by the following equation as the integral of the velocity potential of the incident wave on the surface SH of the ...
  30. [30]
    The generalized Kirchhoff equations and their application to the ...
    Aug 7, 2025 · The generalized Kirchhoff equations and their application to the interaction between a rigid body and an arbitrary time-dependent viscous flow.
  31. [31]
    [PDF] Chapter 6 RIGID BODY DYNAMICS - TU Delft OpenCourseWare
    ... force : Fz(t) = FA(t) +FF (t). = Fa sin(!t + "Fz) pitch moment : Mµ(t) = r fFA(t) ¡ FF(t)g = Ma sin(!t + "Mz). (7.176). Heave Oscillations. The linear equations ...Missing: F_h = "-m_a" dot
  32. [32]
    [PDF] 13.42 04/01/04: Morrison's Equation
    Apr 1, 2004 · The mass and drag coefficients are found using the diameter of the cylinder and the normal velocity as in the general (non-inclined) case. The ...Missing: Kirchhoff | Show results with:Kirchhoff
  33. [33]
    [PDF] Ship Interactions in Arbitrary Channels - DSpace@MIT
    A 3-D panel method approach will be used in this study. This study is an extension of the work done by Korsmeyer, Lee and Newman (1993) [9] and Wu (1995) [20]; ...