Rayleigh dissipation function

In physics, the Rayleigh dissipation function, named after Lord Rayleigh, is a function used to handle the effects of velocity-proportional frictional forces in Lagrangian mechanics. It was first introduced by him in 1873.[1] If the frictional force on a particle with velocity v {\displaystyle {\vec {v}}} can be written as F f = k v {\displaystyle {\vec {F}}_{f}=-{\vec {k}}\cdot {\vec {v}}} , the Rayleigh dissipation function can be defined for a system of N {\displaystyle N} particles as

R ( v ) = 1 2 i = 1 N ( k x v i , x 2 + k y v i , y 2 + k z v i , z 2 ) . {\displaystyle R(v)={\frac {1}{2}}\sum _{i=1}^{N}(k_{x}v_{i,x}^{2}+k_{y}v_{i,y}^{2}+k_{z}v_{i,z}^{2}).}

This function represents half of the rate of energy dissipation of the system through friction. The force of friction is negative the velocity gradient of the dissipation function, F f = v R ( v ) {\displaystyle {\vec {F}}_{f}=-\nabla _{v}R(v)} , analogous to a force being equal to the negative position gradient of a potential. This relationship is represented in terms of the set of generalized coordinates q i = { q 1 , q 2 , q n } {\displaystyle q_{i}=\left\{q_{1},q_{2},\ldots q_{n}\right\}} as

F f = R q ˙ i {\displaystyle {\vec {F}}_{f}=-{\frac {\partial R}{\partial {\dot {q}}_{i}}}} .

As friction is not conservative, it is included in the Q i {\displaystyle Q_{i}} term of Lagrange's equations,

d d t L q i ˙ L q i = Q i {\displaystyle {\frac {d}{dt}}{\frac {\partial L}{\partial {\dot {q_{i}}}}}-{\frac {\partial L}{\partial q_{i}}}=Q_{i}} .

Applying of the value of the frictional force described by generalized coordinates into the Euler-Lagrange equations gives (see [2])

d d t ( L q i ˙ ) L q i = R q ˙ i {\displaystyle {\frac {d}{dt}}{\big (}{\frac {\partial L}{\partial {\dot {q_{i}}}}}{\big )}-{\frac {\partial L}{\partial q_{i}}}=-{\frac {\partial R}{\partial {\dot {q}}_{i}}}} .

Rayleigh writes the Lagrangian L {\displaystyle L} as kinetic energy T {\displaystyle T} minus potential energy V {\displaystyle V} , which yields Rayleigh's Eqn. (26) from 1873.

d d t ( T q i ˙ ) + R q ˙ i + V q i = 0 {\displaystyle {\frac {d}{dt}}{\big (}{\frac {\partial T}{\partial {\dot {q_{i}}}}}{\big )}+{\frac {\partial R}{\partial {\dot {q}}_{i}}}+{\frac {\partial V}{\partial q_{i}}}=0} .

Since the 1970s the name Rayleigh dissipation potential for R {\displaystyle R} is more common. Moreover, the original theory is generalized from quadratic functions q R ( q ˙ ˙ ) = 1 2 q ˙ V q ˙ {\displaystyle q\mapsto R({\dot {\dot {q}}})={\frac {1}{2}}{\dot {q}}\cdot \mathbb {V} {\dot {q}}} to dissipation potentials that are depending on q {\displaystyle q} (then called state dependence) and are non-quadratic, which leads to nonlinear friction laws like in Coulomb friction or in plasticity. The main assumption is then, that the mapping q ˙ R ( q , q ˙ ) {\displaystyle {\dot {q}}\mapsto R(q,{\dot {q}})} is convex and satisfies 0 = R ( q , 0 ) R ( q , q ˙ ) {\displaystyle 0=R(q,0)\leq R(q,{\dot {q}})} , see e.g. [3] [4] [5]


References

  1. ^ Rayleigh, Lord (1873). "Some general theorems relating to vibrations". Proc. London Math. Soc. s1-4: 357–368. doi:10.1112/plms/s1-4.1.357.
  2. ^ Goldstein, Herbert (1980). Classical Mechanics (2nd ed.). Reading, MA: Addison-Wesley. p. 24. ISBN 0-201-02918-9.
  3. ^ Moreau, Jean Jacques (1971). "Fonctions de résistance et fonctions de dissipation". Travaux du Séminaire d'Analyse Convexe, Montpellier (Exposé no. 6): (See page 6.3 for "fonction de resistance").
  4. ^ Lebon, Georgy; Jou, David; Casas-Vàzquez, Jos\'e (2008). Understanding Non-equilibrium Thermodynamics. Springer-Verlag. p. (See Chapter 10.2 for dissipation potentials).
  5. ^ Mielke, Alexander (2023). "An introduction to the analysis of gradient systems". p. (See Definition 3.1 on page 25 for dissipation potentials). arXiv:2306.05026 [math-ph].