<<

Nick Ouellette 113

The Electromagnetic Lagrangian and Hamiltonian

Up to this point, we have only applied the Lagrangian and Hamiltonian formalisms to velocity independent forces and . When we introduce a velocity dependence, we can no longer say that

L = T V, − where L is the Lagrangian, T is the kinetic energy, and V is the energy. We must redefine the Lagrangian to be the function that gives the right answer when fed into the Euler-Lagrange equation. As an example of this, let us consider a charged particle moving in an electromagnetic field. The particle feels a force given by the law: v F = q E + B , (1)  c ×  where we have used cgs units. Note that q is the of the particle and not a generalized coordinate. Maxwell’s equations and vector calculus let us express the electric and magnetic fields in terms of the scalar potential φ and the A: 1 ∂A E = φ −∇ − c ∂t B = A ∇ × Inserting these expressions into the Lorentz force law, we get

1 ∂A 1 F = q  φ + v ( A) (2) −∇ − c ∂t c × ∇ × We now use the vector identity

B ( C) = (B C) (B )C (C )B C ( B) (3) × ∇ × ∇ · − · ∇ − · ∇ − × ∇ × and the fact that v is not an explicit function of position to write

v ( A) = (v A) (v )A. (4) × ∇ × ∇ · − · ∇ Now, writing the total derivative of the vector potential as dA ∂A ∂A dx ∂A dy ∂A dz ∂A = + + + = + (v )A, (5) dt ∂t ∂x dt ∂y dt ∂z dt ∂t · ∇ we can put the Lorentz force law into the form

1 1 dA F = q  φ + (v A)  . (6) −∇ c ∇ · − c dt Now, even though we cannot say that L = T V because of the velocity dependence, there is a more general form of the Euler-Lagrange equation we can− use. The kinetic energy is defined by 1 T = m x˙ 2. X 2 i i i

1 Putting this into Newton’s second law, we obtain

d d ∂T Fi = (mix˙ i) =   . (7) dt dt ∂x˙ i

If we now replace the xi with generalized coordinates qi, some algebra shows that

∂T ∂xj = mjx˙ j   . (8) ∂q˙ X ∂q i j i

Taking a time derivative, more algebra shows that

d ∂T ∂T   = Qi + , (9) dt ∂q˙i ∂qi

where the Qi are the components of the generalized force and are given by ∂x Q = F j . (10) i X j ∂q j i

Rearranging equation (9), we get d ∂T ∂T Qi =   , (11) dt ∂q˙i − ∂qi which is the most general form of the Euler-Lagrange equation. In Cartesian coordinates, it reduces to

d ∂T ∂T Fi =   . dt ∂x˙ i − ∂xi If the system is conservative, the we can write

F = V, −∇ and we come up with d ∂L ∂L   = 0, (12) dt ∂x˙ j − ∂xj which is the familiar Euler-Lagrange equation for L = T V . If, however, the potential (call it U) is velocity-dependent, we can still write an equation of the exact− same form as equation (12) with L = T U if the velocity-dependent force is of the form −

∂U d ∂U Fj = +   . (13) −∂xj dt ∂x˙ j We can put the Lorentz force law into this form by being clever. First, we write

dAj d ∂ =  (v A) , dt dt ∂vj ·

since the will pick out only the jth component of the dot product. Now, since the scalar potential is independent of the velocity, we can add on a term containing it inside the partial derivative:

dAj d ∂ =  (v A qφ) . dt dt ∂vj · −

2 This lets us write the Lorentz force law as ∂ q d ∂ q Fx = qφ (v A) +  qφ (v A)  (14) −∂x  − c ·  dt ∂vx  − c ·  which in turn tells us that our generalized potential must be q U = qφ (v A). (15) − c · This finally gives us the electromagnetic Lagrangian: 1 q L = T U = mv v qφ + (v A). (16) − 2 · − c · The Lagrangian in turn gives us the canonical momentum: ∂L q p = = mv + A. (17) ∂v c Using the canonical momentum and the Lagrangian, we can construct the electromagnetic Hamiltonian:

1 q 2 H = p v L = p A + qφ. (18) · − 2m  − c 

3