Fact-checked by Grok 2 weeks ago

Rayleigh dissipation function

The Rayleigh dissipation function, also known as Rayleigh's dissipation function, is a scalar quantity in that quantifies the rate of dissipation due to frictional forces linearly proportional to the velocities of the components. Introduced by John William Strutt, 3rd Baron Rayleigh, in his 1873 paper "Some general theorems relating to vibrations," it provides an elegant method to incorporate such dissipative effects into the variational principles of mechanics without explicitly treating the non-conservative forces as external terms. Formally defined for a with q_i and velocities \dot{q}_i as R = \frac{1}{2} \sum_{i,j} b_{ij} \dot{q}_i \dot{q}_j, where b_{ij} are coefficients related to the friction matrix, the function yields the dissipative generalized forces via Q_j^{\text{diss}} = -\frac{\partial R}{\partial \dot{q}_j}. This function was originally developed in the context of acoustic vibrations and fluid dissipation, as elaborated in Rayleigh's seminal 1877–1878 treatise The Theory of Sound, where it modeled viscous losses in wave propagation. In , it modifies the Euler-Lagrange equations for non-conservative systems to \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_j} \right) - \frac{\partial L}{\partial q_j} + \frac{\partial R}{\partial \dot{q}_j} = Q_j^{\text{ext}}, where L = T - V is the , T the , V the , and Q_j^{\text{ext}} any additional non-dissipative generalized forces. For viscous , the diagonal form R = \frac{1}{2} \sum_i b_i \dot{q}_i^2 is common, with b_i representing damping coefficients, ensuring the dissipation rate equals $2R. The Rayleigh dissipation function has broad applications beyond acoustics, including , control systems, and , where it facilitates the analysis of damped oscillators and multi-body systems with . Extensions to nonlinear friction, such as or velocity-dependent dry , have been developed, allowing its use in more complex scenarios like on conveyor belts or rotating machinery.

Definition and Mathematical Formulation

Definition

The Rayleigh dissipation function is a scalar quantity introduced by Lord Rayleigh in his 1873 paper "Some general theorems relating to ," as elaborated in his 1877 treatise The Theory of Sound, to describe energy losses in mechanical systems due to frictional forces that are proportional to . It serves as a potential-like function for dissipative effects, particularly those arising from or linear , enabling a unified treatment within the framework of . Conceptually, the function represents half the instantaneous rate of energy dissipation in the system, capturing how frictional forces convert into or other irreversible forms. For instance, in the case of a single particle moving through a viscous medium where the drag force opposes motion and scales linearly with speed, the dissipation function quantifies the power lost as the particle works against this resistance. This approach highlights the quadratic dependence on velocities inherent to linear friction, distinguishing it from conservative forces derived from potentials. In , the Rayleigh dissipation function facilitates the handling of non-conservative forces by generating generalized forces that can be directly incorporated into the , avoiding the need for separate force balance terms. This makes it particularly useful for systems with multiple , where dissipation may couple different coordinates. The function's is independent of the choice of coordinates, allowing it to be expressed naturally in suitable for the system's description.

Mathematical Expression

The Rayleigh dissipation function R is generally expressed in quadratic form as R = \frac{1}{2} \sum_{i,j} c_{ij} \dot{q}_i \dot{q}_j, where q_i are the , \dot{q}_i are the corresponding generalized velocities, and c_{ij} are the dissipation coefficients forming a symmetric positive semi-definite to ensure R \geq 0 for physical systems. In Cartesian coordinates for a of N particles subject to velocity-proportional , the dissipation function takes the form R = \frac{1}{2} \sum_{i=1}^N \left( k_{x} v_{i,x}^2 + k_{y} v_{i,y}^2 + k_{z} v_{i,z}^2 \right), where v_{i,\alpha} (\alpha = x,y,z) are the components of the i-th particle, and k_{\alpha} are positive coefficients; this corresponds to frictional forces \vec{F}_{f,i} = -k \vec{v}_i for isotropic with k_x = k_y = k_z = k. (Goldstein et al., 2002, p. 22) The R relates directly to the power in the , equaling half the instantaneous of work done against the frictional forces, such that $2R = \sum_i -\vec{F}_{f,i} \cdot \vec{v}_i, where the negative sign accounts for the dissipative nature of \vec{F}_{f,i}. A key property of R is that it is homogeneous of degree 2 in the velocities \dot{q}_i, meaning R(\lambda \dot{q}) = \lambda^2 R(\dot{q}) for any scalar \lambda > 0; this homogeneity ensures the function scales appropriately for dissipative forces linear in velocity, consistent with the and the resulting generalized forces Q_k = -\partial R / \partial \dot{q}_k.

Incorporation into Lagrangian Mechanics

Generalized Forces from Dissipation

In , the dissipative generalized forces arising from the Rayleigh dissipation function R are defined as Q_k^{\text{diss}} = -\frac{\partial R}{\partial \dot{q}_k}, where q_k are the and \dot{q}_k their time derivatives. This formulation allows the incorporation of velocity-dependent dissipative effects, such as or , into the without explicitly treating them as external forces. For a simple viscous acting on a coordinate x with c, the takes the R = \frac{1}{2} c \dot{x}^2. The yields \frac{\partial R}{\partial \dot{x}} = c \dot{x}, so the dissipative force is Q^{\text{diss}} = -c \dot{x}, representing the opposing force proportional to . These forces are inherently non-conservative and depend explicitly on the generalized velocities, in contrast to conservative forces derived from a potential V(q), which depend only on positions. In multi-degree-of-freedom systems, the Rayleigh is typically in the velocities, leading to a vector form of the dissipative forces \vec{Q}^{\text{diss}} = -C \dot{\vec{q}}, where C is the symmetric matrix with elements corresponding to the between coordinates.

Modified Equations of Motion

The Rayleigh dissipation function R extends the standard Euler-Lagrange equations to account for linear velocity-dependent dissipative forces. The modified take the form \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_k} \right) - \frac{\partial L}{\partial q_k} + \frac{\partial R}{\partial \dot{q}_k} = 0, where L = T - V is the , with T denoting the and V the , and the index k labels the q_k. This modification is equivalent to incorporating explicit dissipative generalized forces Q_k^{\text{diss}} = -\frac{\partial R}{\partial \dot{q}_k} into the unmodified Euler-Lagrange equations, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{q}_k} \right) - \frac{\partial L}{\partial q_k} = Q_k^{\text{diss}}. The term Q_k^{\text{diss}} arises directly from the of R with respect to the generalized velocities \dot{q}_k, ensuring that the dissipative effects are systematically included without altering the conservative structure of the . In the Hamiltonian formalism, the Rayleigh dissipation function introduces a corresponding dissipative contribution to the . Specifically, it modifies the evolution equation by adding the term Q_k^{\text{diss}} = -\frac{\partial R}{\partial \dot{q}_k} to \dot{p}_k = -\frac{\partial [H](/page/Hamiltonian)}{\partial q_k}, where [H](/page/Hamiltonian) is the and p_k are the conjugate , while the coordinate evolution \dot{q}_k = \frac{\partial [H](/page/Hamiltonian)}{\partial p_k} remains unchanged. This approach offers significant advantages for systems involving linear dissipation, as it streamlines the inclusion of velocity-proportional forces compared to deriving and adding them explicitly, especially in formulations with intricate or high .

Derivation

Principle of Virtual Work

The principle of virtual work provides a foundational approach to incorporating dissipative forces into the dynamics of mechanical systems, extending D'Alembert's principle to non-conservative cases. For a system subject to dissipative forces \vec{F}_f, the virtual work done by these forces during an infinitesimal virtual displacement \delta \vec{r} is given by \delta W^{\text{diss}} = \sum \vec{F}_f \cdot \delta \vec{r}. In generalized coordinates q_k, this takes the form \delta W^{\text{diss}} = \sum_k Q_k^{\text{diss}} \delta q_k, where Q_k^{\text{diss}} are the generalized dissipative forces. Assuming the dissipative forces are linear in the generalized velocities \dot{q}_k, as in viscous \vec{F}_f = -\mathbf{C} \vec{v}, the generalized forces derive as Q_k^{\text{diss}} = -\sum_j b_{kj} \dot{q}_j = -\partial R / \partial \dot{q}_k, where R = \frac{1}{2} \sum_{i,j} b_{ij} \dot{q}_i \dot{q}_j and b_{ij} are the elements of the . This partial derivative integrates R into the . The quadratic form of R arises from the linearity of the dissipative forces in , leading to a bilinear expression for the work in terms of velocities. The Rayleigh dissipation function R relates to energy dissipation through the power lost to , given by \sum_k Q_k^{\text{diss}} \dot{q}_k = -2R. This relation justifies the conventional factor of $1/2 in the definition of R, ensuring that the power dissipated by the frictional forces equals $2R, consistent with the of R in velocities. For instance, in a with linear velocity-dependent dissipation, the instantaneous power loss aligns with twice the value of R, balancing the energy extraction from the . The bilinearity of the frictional contributions ensures that R is a homogeneous of the \dot{q}_k, capturing the symmetric coupling between in multi-particle or continuous systems with linear .

Assumptions for Linear Dissipation

The Rayleigh dissipation function applies specifically to systems where dissipative forces are linear in the velocities, typically modeled as \mathbf{F}_f = -\mathbf{C} \mathbf{v}, with \mathbf{C} a positive semi-definite damping matrix representing viscous friction. This linearity assumption, originating from Rayleigh's analysis of resistant forces proportional to velocity, excludes nonlinear dissipative effects such as Coulomb dry friction, which depends on the direction of motion rather than velocity magnitude. The formulation further assumes and homogeneity in the process, implying that the coefficients in \mathbf{C} are uniform across the system and exhibit no directional bias or spatial variation. These conditions ensure the can be captured by a scalar in velocities, without dependence on position coordinates, which simplifies incorporation into the framework via the principle of . Validity is restricted to regimes of small perturbations or low velocities, where higher-order terms like quadratic drag become negligible compared to linear contributions. In such cases, the Rayleigh function effectively models weak without requiring explicit force terms in the . Key limitations include its inability to handle position-dependent or conservative-like dissipation, where mimics potential forces, and time-varying coefficients in \mathbf{C}, which would violate the velocity-only dependence. Consequently, while the function guarantees monotonic decrease in the system's total —manifesting as a non-positive time derivative of —the approach to is not assured and depends on additional conditions.

Applications

Damped Harmonic Oscillators

The Rayleigh dissipation function finds a straightforward application in the analysis of a single degree-of-freedom damped , a fundamental model in representing systems like a mass-spring setup immersed in a viscous medium. The for the undamped system is given by L = \frac{1}{2} m \dot{x}^2 - \frac{1}{2} k x^2, where m is the mass, k is the spring constant, x is the displacement from equilibrium, and \dot{x} is the velocity. To incorporate viscous damping, which exerts a force proportional to velocity F = -c \dot{x} with damping coefficient c, the Rayleigh dissipation function is defined as R = \frac{1}{2} c \dot{x}^2. This quadratic form in velocity captures the linear dissipative forces symmetrically. Substituting into the modified Lagrange equation of motion, \frac{d}{dt} \left( \frac{\partial L}{\partial \dot{x}} \right) - \frac{\partial L}{\partial x} + \frac{\partial R}{\partial \dot{x}} = 0, yields the standard second-order differential equation m \ddot{x} + c \dot{x} + k x = 0, or equivalently, \ddot{x} + 2\gamma \dot{x} + \omega_0^2 x = 0, where \gamma = c/(2m) is the damping rate and \omega_0 = \sqrt{k/m} is the natural frequency. The nature of the solutions depends on the damping regime, determined by the ratio \gamma / \omega_0. For underdamping (\gamma < \omega_0), the motion is oscillatory with decaying amplitude, described by x(t) = A e^{-\gamma t} \cos(\omega_d t + \phi), where \omega_d = \sqrt{\omega_0^2 - \gamma^2}. Critical damping occurs at \gamma = \omega_0, yielding the fastest non-oscillatory return to equilibrium via x(t) = (A + B t) e^{-\gamma t}. Overdamping (\gamma > \omega_0) results in purely exponential decay without oscillations, x(t) = A e^{-(\gamma - \sqrt{\gamma^2 - \omega_0^2}) t} + B e^{-(\gamma + \sqrt{\gamma^2 - \omega_0^2}) t}. These behaviors illustrate how the dissipation function systematically accounts for energy loss in the oscillator's dynamics. The physical interpretation of the dissipation is evident in the energy balance: the rate of change of the total E = \frac{1}{2} m \dot{x}^2 + \frac{1}{2} k x^2 is \frac{dE}{dt} = -2R = -c \dot{x}^2, confirming that the power dissipated equals the negative of twice the dissipation function, leading to of the amplitude in the underdamped case as e^{-\gamma t}. This relation highlights the function's role in quantifying irreversible energy loss to the surroundings, such as through viscous , without altering the conservative structure of the .

Viscous Fluid Dynamics

In viscous fluid dynamics, the Rayleigh dissipation function plays a crucial role in modeling the energy losses due to shear viscosity within the framework of the Navier-Stokes equations. The dissipation term in the equation for incompressible Newtonian fluids appears as the negative of the integral of twice the dynamic viscosity μ times the square of the strain rate tensor components, which quantifies the irreversible conversion of to . This term is analogous to the continuous-media form of the Rayleigh dissipation function, expressed as R = \frac{1}{4} \int \mu \left( \nabla \mathbf{v} + (\nabla \mathbf{v})^T \right)^2 \, dV, where \mathbf{v} is the velocity field, such that the total dissipation rate equals $2R. A representative example is Poiseuille flow, the steady of a viscous driven by a through a cylindrical or between parallel plates, where the Rayleigh dissipation function captures the shear viscosity losses across the cross-section. In this case, the velocity profile is parabolic, with the strain rate peaking at the walls, leading to maximum dissipation there; the function R integrates these local losses to yield the total power dissipated, equaling the pressure drop times the . For non-isothermal extensions, such as in polymeric fluids, the dissipation function enters the equation to account for temperature rises due to viscous heating. In variational formulations of incompressible flows, the Rayleigh dissipation function modifies the action principle by incorporating viscous terms into the , enabling the of the Navier-Stokes equations through stationarity conditions that minimize energy dissipation. This approach aligns with the Helmholtz minimum dissipation theorem for steady s, where the function ensures the velocity field minimizes R subject to boundary conditions. Physically, the Rayleigh dissipation function quantifies the arising from the irreversibility of viscous processes, particularly in low-Reynolds-number regimes where inertial effects are negligible and dissipation dominates the flow behavior. In such creeping flows, like around obstacles, R provides a measure of the thermodynamic inefficiency, with generation rate proportional to $2R / T, where T is .

History and Extensions

Rayleigh's Original Contribution

John William Strutt, the third Baron , first introduced the concept of the dissipation function in his paper titled "Some General Theorems Relating to Vibrations," presented to the London Mathematical Society on June 12, 1873, and published in the Proceedings of the London Mathematical Society. In this work, Rayleigh addressed the challenge of incorporating dissipative effects into the study of vibrating systems, building on earlier variational approaches to mechanics. He proposed the dissipation function as a mathematical tool to quantify energy loss in systems subject to frictional forces, marking a significant advance in handling non-conservative forces within frameworks. Rayleigh's original formulation arose in the broader context of acoustics, where he sought to model the of waves due to internal friction from and heat conduction. This idea was elaborated in his comprehensive two-volume treatise The Theory of Sound, published between 1877 and 1878, which became a foundational text in the field. In this book, Rayleigh applied the dissipation function to analyze how viscous drag and thermal gradients in fluids like air lead to progressive damping of acoustic disturbances, providing a rigorous of wave propagation under dissipative conditions. The key insight of Rayleigh's contribution was his observation that dissipative forces linear in velocity could be encapsulated in a quadratic dissipation function, Φ, such that the power dissipated equals twice this function, mirroring the role of T in conservative systems. This analogy allowed dissipative terms to be incorporated symmetrically into the , facilitating the use of variational principles for vibratory problems. For early applications, Rayleigh demonstrated its utility in deriving attenuation rates for plane waves in air, showing that the absorption coefficient depends on the medium's and thermal conductivity, with explicit expressions for the of wave over distance. These calculations highlighted the function's practical value in predicting the limited range of audible in dissipative .

Modern Generalizations

Since the , the Rayleigh dissipation function has been generalized through the introduction of dissipation potentials, which extend the to nonlinear dissipative forces. In this approach, a dissipation potential D(\dot{\mathbf{q}}) is defined such that the dissipative generalized forces are given by \mathbf{Q}^{\text{diss}} = -\frac{\partial D}{\partial \dot{\mathbf{q}}}. When D is in the velocities, \frac{1}{2} \sum_{i,j} c_{ij} \dot{q}_i \dot{q}_j, it recovers the original Rayleigh function R = 2D, yielding linear velocity-proportional forces; however, more general forms of D allow for nonlinearities, such as linear D in |\dot{q}| for dry friction or cubic D in \dot{q} for forces. This generalization, foundational to the of generalized standard materials, enables modeling of complex dissipative behaviors in solids and fluids while preserving variational structure in the . State-dependent extensions further broaden applicability by allowing D to depend not only on velocities but also on positions \mathbf{q} or other state variables, accommodating spatially varying or velocity-nonlinear friction. Such forms trace back to early generalizations in non-holonomic dynamics via the Gibbs-Appell equations, where dissipation is incorporated into higher-order work expressions, and have been refined in modern control theory for systems with position-dependent damping. In robotics, these generalized dissipation potentials facilitate Lyapunov-based stability analysis for manipulators subject to nonlinear friction, where a Lyapunov function V = T + U (kinetic plus potential energy) yields \dot{V} = -\frac{\partial D}{\partial \dot{\mathbf{q}}} \cdot \dot{\mathbf{q}} \leq 0, ensuring asymptotic stability under feedback control laws that dominate unmodeled nonlinearities. Representative examples include hybrid position/force control of robotic arms with Coulomb friction and sliding mode control of multi-fingered grippers, where the potential structure simplifies proof of robustness. Recent developments integrate these generalizations with through Rayleighian functionals, which combine rates with variational principles to describe far-from-equilibrium dynamics in and active systems. The Rayleighian \mathcal{R} = \dot{F} + \Phi, where \dot{F} is the rate and \Phi the function, is minimized to yield evolution equations, linking mechanical potentials to Onsager reciprocity and bracket formalisms for irreversible processes. 2024 analyses quantify 's role in flows and porous media, extending the framework to continuum nonequilibrium settings. These advances highlight the function's enduring role in unifying and .

References

  1. [1]
    [PDF] Rayleigh's dissipation function at work - arXiv
    Feb 28, 2015 · Abstract. It is shown that the Rayleigh's dissipation function can be successfully ap- plied in the solution of mechanical problems ...Missing: primary | Show results with:primary
  2. [2]
  3. [3]
    10.4: Rayleigh's Dissipation Function - Physics LibreTexts
    Feb 27, 2021 · The Rayleigh dissipation function is an elegant way to include linear velocity-dependent dissipative forces in both Lagrangian and Hamiltonian mechanics.Missing: primary | Show results with:primary
  4. [4]
    7.3.1 Coupling of damped systems and the first-order method
    known as the Rayleigh dissipation function, is half the rate of dissipation of energy. This is another quadratic expression, in which we can choose the ...<|control11|><|separator|>
  5. [5]
    [PDF] Lagrange's Equations
    where the are the damping coefficients, which are dissipative in nature → result in a loss of energy ij c. • Now define the Rayleigh dissipation function. 1.
  6. [6]
    [PDF] Analytical Dynamics: Lagrange's Equation and its Application
    Apr 9, 2017 · 2 (rel. to x1). Rayleigh's Dissipation Function: The form of the dissipation function follows that of the potential energy: D = 1.<|control11|><|separator|>
  7. [7]
    [PDF] The Rayleigh dissipation function
    The Rayleigh dissipation function, as we will see, will allow us to describe the non- conservative force of friction, hence we'll focus on non-conservative ...Missing: source | Show results with:source
  8. [8]
    None
    ### Summary of Rayleigh Dissipation Function Derivation Using Virtual Work
  9. [9]
    [PDF] Damped Harmonic Oscillator 1. General Form of the Lagrangian
    Oct 23, 2013 · Rayleigh's dissipation function allows a gen- eral linear damping term to be incorporated into the Lagrangian formalism. then d dt. µ∂L. ∂ ˙q. ¶.
  10. [10]
    [PDF] Using Onsager principle as an approximation tool for complicated ...
    Sep 28, 2019 · The Rayleigh dissipation function is given by Φ(˙u), which is defined as the half of the total energy dissipation rate. If the system is.
  11. [11]
  12. [12]
    Application of recursive Gibbs–Appell formulation in deriving the ...
    Rayleigh's dissipation function of the system and its derivatives. An important class of nonconservative forces is the class of forces that dissipate energy ...
  13. [13]
    Nonlinear control of a fully actuated robotic hand using high-order ...
    represents the Rayleigh dissipation function, capturing energy dissipation (e.g., damping effects), defined as 3: (3). Here, represents the damping ...<|separator|>
  14. [14]
  15. [15]
    Dissipation in nonequilibrium thermodynamics and its connection to ...
    Jan 4, 2024 · We examine quantitatively the role of dissipation in nonequilibrium thermodynamics and its connection to variational principles and the Rayleighian functional.