Solutions

Classical Mechanics - The Theoretical Minimum

Exercise list

Lecture 8

Exercise 8.1

Exercise 8.2

Lecture 10

Exercise 10.1

Exercise 10.2

Exercise 10.3

Lecture 11

Exercise 11.1

Exercise 11.2

Exercise 11.3

Exercise 11.4

Exercise 11.5

Appendix 1

Exercise A1.1

Lecture 8

Exercise 8.1

Start with the Lagrangian \dfrac{m\dot{x}^2}{2}-\dfrac{k}{2}x^2 and show that if you make the change in variables q=(k m )^{1/4}x, the Lagrangian has the form of Eq. (14). What is the connection among k, m and \omega ?

Considering the harmonic oscillator:

\mathcal L = \frac{m}{2}\dot x^2 - \frac{k}{2} x^2

Performing the substitution q = (k \, m)^{{}^1{\mskip -5mu/\mskip -3mu}_4}\,x:

\begin{array}{lll} q = (k \, m)^{\frac{1}{4}}\,x & \Rightarrow & x^2 = \frac{1}{\sqrt{m \, k}} \, q^2\\ \dot q = (k \, m)^{\frac{1}{4}}\,\dot x & \Rightarrow & \dot x = \frac{1}{\sqrt{m \, k}} \, \dot q^2 \end{array}

The Lagrangian becomes:

\mathcal L = \frac{m}{2\sqrt{m \, k}}\dot q^2 - \frac{k}{2\sqrt{m \, k}} q^2 = \frac{1}{2}\sqrt{\frac{k}{m}}\dot q^2 - \frac{1}{2}\sqrt{\frac{m}{k}} q^2

Therefore, written as function of \omega = \sqrt{\frac{k}{m}} the Lagrangian is:

\mathcal L = \frac{1}{2\omega}\dot q^2 -\frac{\omega^2}{2} q^2

Exercise 8.2

Starting with Eq. (14), calculate the Hamiltonian in terms of p and q.

Written as function of \omega = \sqrt{\frac{k}{m}} the Lagrangian is:

\mathcal L = \frac{1}{2\omega}\dot q^2 -\frac{\omega^2}{2} q^2

\dot q can be derived as function of q and p:

\begin{aligned} & p = \dfrac{\dot q}{\omega} \quad \Rightarrow \quad \dot q = \omega\,p \end{aligned}

The Hamiltonian can be then written:

\mathcal H = \omega p^2 -\frac{1}{2\omega}\omega^2 p^2 + \frac{\omega^2}{2} q^2 = \frac{\omega}{2}\left( p^2 + q^2 \right)

This quantity, the energy E, is conserved, and a point which starts with a given value of energy H(p,q) = c it remains on that energy.

Lecture 10

Exercise 10.1

Prove Eq. (14).

From the definition of the Poisson bracket:

\{F,G\} \equiv \sum_i \left( \frac{\partial F}{\partial q_i} \frac{\partial G}{\partial p_i} - \frac{\partial F}{\partial p_i} \frac{\partial G}{\partial q_i} \right)

For a function F(q,p):

\{F(q,p),p_i\} = \sum_j \left( \frac{\partial F(q,p)}{\partial q_j} \frac{\partial p_i}{\partial p_j} - \frac{\partial F(q,p)}{\partial p_j}\frac{\partial p_i}{\partial q_j} \right) = \frac{\partial F(q,p)}{\partial q_i}

since in the sum, all terms \frac{\partial p_i}{\partial q_j} vanish (i.e., each right product), and also all terms \frac{\partial p_i}{\partial p_j} with i \neq j in the left product. Thus, only the terms with j=i remains from the sum, where \frac{\partial p_i}{\partial p_i} = 1.

Exercise 10.2

Hamilton’s equations can be written in the form \dot{q} = \{q, \mathcal H\} and \dot{p} = \{p, \mathcal H\}. Assume that the Hamiltonian has the form \mathcal H = \dfrac1{2m}p^2+V(q). Using only the PB axioms, prove Newton’s equations of motion.

Using just the definition of the Poisson brackets, computing \dot q:

\begin{aligned} \dot q & = \left\{ q, \mathcal H \right\} = \left\{ q, \frac{1}{2m} p^2 + V(q) \right\} = \left\{ q, \frac{1}{2m} p^2 \right\} + \left\{ q, V(q) \right\} \\ & = \frac{1}{2m} \left\{ q, p^2 \right\} + 0 = \frac{1}{2m} \left[ p\left\{ q, p \right\} + p\left\{ q, p \right\} \right] = \frac{p}{m} \end{aligned}

Computing \dot p:

\begin{aligned} \dot p & = \left\{ p, \mathcal H \right\} = \left\{ p, \frac{1}{2m} p^2 + V(q) \right\} = \frac{1}{2m} \left\{ p, p^2 \right\} + \left\{ p, V(q) \right\}\\ & = 0 - \left\{ V(q), p \right\} = -\frac{\mathrm dV}{\mathrm dq} \end{aligned}

Taking the derivative with respect of time of the first equation and using the second it is possible to derive Newton’s equation of motion:

\begin{aligned} & \ddot q = \dfrac{\dot p}{m} = -\frac{1}{m} \frac{\mathrm dV}{\mathrm dq} \\ & m \, \ddot q = -\frac{\mathrm dV}{\mathrm dq} = F_q \\ & F_q = m \, a_q \end{aligned}

Exercise 10.3

Using the definition of PB’s and the axioms, work out the PB’s in Equations (19). In each expression, look for things in the parentheses that have non-zero Poisson Brackets with the coordinate x, y or z. For example, in the first PB, x has a nonzero PB with p_x.

Angular momentum is the quantity which is conserved when there are rotation symmetry.

Rotation of a particle by \theta

\begin{aligned} & \delta x = - y\,\varepsilon \\ & \delta y = x\,\varepsilon \\ & \delta z = 0 \end{aligned}

The last equation is simply stating that there is no change in z for a rotation in the plane. Looking at the momentum \mathbf p it possible to write a similar set for each component:

\begin{aligned} & \delta p_x = - p_y\,\varepsilon \\ & \delta p_y = p_x\,\varepsilon \\ & \delta p_z = 0 \end{aligned}

If there is a symmetry there is a conserved quantity:

\mathcal Q = \sum_i p_i \,f_i(\mathbf q) = x\,p_y - y\,p_y = L_z

This is the angular momentum on the z direction, perpendicular to the xy plane. The other components in the three dimensional space of \mathbf L can be derived in a similar way, and will have indexes in a cyclic fashion x \rightarrow y \rightarrow z:

\begin{aligned} & L_z = x\,p_y - y\, p_x \\ & L_x = y\,p_z - z\, p_y \\ & L_y = z\,p_x - x\, p_z \end{aligned}

Using the cross product, it can be written in a short form as:

\mathbf L = \mathbf r \times \mathbf p

Using the Poisson brackets and take them between x, y, z and L_z:

\begin{aligned} \left\{ x, L_z \right\} = & \left\{ x, x\,p_y - y\, p_x \right\} = \left\{ x, - y\, p_x \right\} = -y \left\{ x, p_x \right\} = -y \\ \left\{ y, L_z \right\} = & \left\{ y, x\,p_y - y\, p_x \right\} = \left\{ y, x \, p_y \right\} = x \left\{ y, p_y \right\} = x \\ \left\{ z, L_z \right\} = & \left\{ z, x\,p_y - y\, p_x \right\} = 0 \end{aligned}

These quantity are the infinitesimal change \delta divided by the fix quantity \varepsilon.

Calculating the Poisson bracket with respect to momentum:

\begin{aligned} \left\{ p_x, L_z \right\} = & \left\{ p_x, x\,p_y - y\, p_x \right\} = \left\{ p_x, x\, p_y \right\} = p_y \left\{ p_x, y \right\} = -p_y \\ \left\{ p_y, L_z \right\} = & \left\{ p_y, x\,p_y - y\, p_x \right\} = \left\{ p_y, -y\, p_x \right\} = -p_x \left\{ p_y, x \right\} = p_x \\ \left\{ p_z, L_z \right\} = & \left\{ p_z, x\,p_y - y\, p_x \right\} = 0 \end{aligned}

This has been calculated for rotation invariance; it is possible to do a similar analysis for translation invariance, and the quantity which is conserved in that case is the momentum.

\delta q = \varepsilon \quad f = 1 \quad \Rightarrow \quad \mathcal Q = p

Calculating the Poisson bracket between different component of the angular momentum:

\begin{aligned} & \left\{ L_z, L_x \right\} = \left\{ x\,p_y - y\,p_x, y\,p_z - z\,p_y \right\} = z\,p_x - x\,p_z = L_y \\ & \left\{ L_x, L_y \right\} = \left\{ y\,p_z - z\,p_y, z\,p_x - x\,p_z \right\} = x\,p_y - z\,p_x = L_z \\ & \left\{ L_y, L_z \right\} = \left\{ z\,p_x - x\,p_z, x\,p_y - y\,p_x \right\} = y\,p_x - x\,p_y = L_x \end{aligned}

Lecture 11

Exercise 11.1

Confirm Eq. (3). Also prove that V_iA_j - V_jA_i = \sum_k \epsilon_{ijk}\left( \mathbf{V} \times \mathbf{A} \right)_i.

The cross product of two vectors \mathbf{V} and \mathbf{A} (\mathbf{V} \times \mathbf{A}), is a vector which the i^{th} component can be expressed using the Levi-Civita symbol \epsilon_{ijk} and the components of the vectors \mathbf{V} and \mathbf{A}. The Levi-Civita symbol is defined as 1 for even permutations of (i, j, k), -1 for odd permutations, and 0 if any two indices are equal.

The i^{th} component of the cross product is defined as:

(\mathbf{V} \times \mathbf{A})_i = \sum_{j=1}^{3} \sum_{k=1}^{3} \epsilon_{ijk} V_j A_k

For the first component:

(\mathbf{V} \times \mathbf{A})_1 = \sum_{j=1}^{3} \sum_{k=1}^{3} \epsilon_{1jk} V_j A_k

Expanding the sums, we specifically get terms involving \epsilon_{123}, \epsilon_{132}, …, leading to:

(\mathbf{V} \times \mathbf{A})_1 = \epsilon_{123} V_2 A_3 + \epsilon_{132} V_3 A_2

Given \epsilon_{123} = 1 (cyclic permutation) and \epsilon_{132} = -1 (anti-cyclic permutation), it simplifies to:

(\mathbf{V} \times \mathbf{A})_1 = V_2 A_3 - V_3 A_2

This process can be analogously applied for i=2 and i=3, leading to the full cross product vector.

The expression for the second component (\mathbf{V} \times \mathbf{A})_2 can be derived similarly:

(\mathbf{V} \times \mathbf{A})_2 = \sum_{j=1}^{3} \sum_{k=1}^{3} \epsilon_{2jk} V_j A_k

Expanding the sums and considering the non-zero terms involving \epsilon_{213}, \epsilon_{231}, …:

(\mathbf{V} \times \mathbf{A})_2 = \epsilon_{213} V_1 A_3 + \epsilon_{231} V_3 A_1

Given \epsilon_{213} = -1 (anti-cyclic permutation) and \epsilon_{231} = 1 (cyclic permutation), it simplifies to:

(\mathbf{V} \times \mathbf{A})_2 = -V_1 A_3 + V_3 A_1

Applying the same approach for the third component (\mathbf{V} \times \mathbf{A})_3:

(\mathbf{V} \times \mathbf{A})_3 = \sum_{j=1}^{3} \sum_{k=1}^{3} \epsilon_{3jk} V_j A_k

Considering the relevant terms with non-zero \epsilon values:

(\mathbf{V} \times \mathbf{A})_3 = \epsilon_{312} V_1 A_2 + \epsilon_{321} V_2 A_1

Given \epsilon_{312} = 1 (cyclic permutation) and \epsilon_{321} = -1 (anti-cyclic permutation), this simplifies to:

(\mathbf{V} \times \mathbf{A})_3 = V_1 A_2 - V_2 A_1

Hence, the cross product \mathbf{V} \times \mathbf{A} in component form is:

\mathbf{V} \times \mathbf{A} = \left( V_2 A_3 - V_3 A_2, V_3 A_1 - V_1 A_3, V_1 A_2 - V_2 A_1 \right)

The second part of the exercise is to prove:

V_i A_j - V_j A_i = \sum_k \epsilon_{ijk} (\mathbf{V} \times \mathbf{A})_k

Substituting (\mathbf{V} \times \mathbf{A})_k with its definition in terms of \epsilon_{kmn} and vector components:

V_i A_j - V_j A_i = \sum_k \epsilon_{ijk} \left( \sum_{m=1}^{3} \sum_{n=1}^{3} \epsilon_{kmn} V_m A_n \right)

Simplifying using the property of the Levi-Civita symbol:

V_i A_j - V_j A_i = \sum_{m=1}^{3} \sum_{n=1}^{3} \left( \sum_k \epsilon_{ijk} \epsilon_{kmn} \right) V_m A_n

Applying the Levi-Civita contraction identity:

V_i A_j - V_j A_i = \sum_{m=1}^{3} \sum_{n=1}^{3} (\delta_{im}\delta_{jn} - \delta_{in}\delta_{jm}) V_m A_n

Recognizing that only terms where m=i and n=j (and vice versa) contribute:

V_i A_j - V_j A_i = V_i A_j - V_j A_i

Exercise 11.2

Prove Eq. (4).

Let V(x, y, z) be a scalar field. The gradient of V, denoted as \nabla V, in Cartesian coordinates is given by:

\nabla V = \frac{\partial V}{\partial x} \mathbf{i} + \frac{\partial V}{\partial y} \mathbf{j} + \frac{\partial V}{\partial z} \mathbf{k}

The curl of a vector field \mathbf{F} = M\mathbf{i} + N\mathbf{j} + P\mathbf{k} is defined as:

\nabla \times \mathbf{F} = \left( \frac{\partial P}{\partial y} - \frac{\partial N}{\partial z} \right) \mathbf{i} + \left( \frac{\partial M}{\partial z} - \frac{\partial P}{\partial x} \right) \mathbf{j} + \left( \frac{\partial N}{\partial x} - \frac{\partial M}{\partial y} \right) \mathbf{k}

Applying the curl operator to \nabla V:

\nabla \times (\nabla V) = \left( \frac{\partial}{\partial y}\left(\frac{\partial V}{\partial z}\right) - \frac{\partial}{\partial z}\left(\frac{\partial V}{\partial y}\right) \right) \mathbf{i} + \left( \frac{\partial}{\partial z}\left(\frac{\partial V}{\partial x}\right) - \frac{\partial}{\partial x}\left(\frac{\partial V}{\partial z}\right) \right) \mathbf{j} + \left( \frac{\partial}{\partial x}\left(\frac{\partial V}{\partial y}\right) - \frac{\partial}{\partial y}\left(\frac{\partial V}{\partial x}\right) \right) \mathbf{k}

Given that the order of partial differentiation is interchangeable for continuous and twice differentiable functions (Schwarz’s theorem), we have:

\frac{\partial^2 V}{\partial y \partial z} = \frac{\partial^2 V}{\partial z \partial y}, \quad \frac{\partial^2 V}{\partial z \partial x} = \frac{\partial^2 V}{\partial x \partial z}, \quad \frac{\partial^2 V}{\partial x \partial y} = \frac{\partial^2 V}{\partial y \partial x}

Therefore, each term in the expression for \nabla \times (\nabla V) cancels out, leading to:

\nabla \times (\nabla V) = 0 \mathbf{i} + 0 \mathbf{j} + 0 \mathbf{k} = \mathbf{0}

This proves that the curl of the gradient of any scalar field V(x, y, z) is always zero.

Exercise 11.3

Show that the vector potentials in Equations (8) and Equations (9) both give the same uniform magnetic field. This means that the two differ by a gradient. Find the scalar whose gradient, when added to Equations (8), gives Equations (9).

Considering \mathbf A:

\begin{aligned} & A_x = 0 \\ & A_y = b\,x \\ & A_z = 0 \end{aligned}

Taking the curl, the only derivative is \frac{\partial A_y}{\partial x} = b, and taking the curl is:

B_z = \frac{\partial A_y}{\partial x} - \frac{\partial A_x}{\partial y} = b - (0) = b

Considering \mathbf A{'} instead:

\begin{aligned} & A_x{'} = -b\,y \\ & A_y{'} = 0 \\ & A_z{'} = 0 \end{aligned}

Taking again the curl:

B_z = \frac{\partial A_y{'}}{\partial x} - \frac{\partial A_x{'}}{\partial y} = (0) - -b = b So these gives the same uniform magnetic field oriented in the z direction and therefore they differ by a gradient:

\mathbf A = \mathbf A{'} + \nabla V

Therefore:

\nabla V = \mathbf A - \mathbf A{'} since \frac{\partial V}{\partial x} = - by, \frac{\partial V}{\partial y} = -bx and \frac{\partial V}{\partial z} = 0, the potential function which satisfies these equation and has this gradient is:

V(x,y,z) = -bxy

Exercise 11.4

Using the Hamiltonian, Eq. (24), work out Hamilton’s equations of motion and show that you just get back to the Newton-Lorentz equation of motion.

The Newton-Lorentz equation of motion are:

m\mathbf{a} = \frac{e}{c}(\mathbf{v} \times\mathbf{B})

Considering a Lagrangian for a particle moving in a constant magnetic field:

\mathcal L = \frac{m}{2}\left( \dot x^2 + \dot y^2 + \dot z^2 \right) +\frac{e}{c}\left( \dot x A_x + \dot y A_y + \dot z A_z \right)

From this Lagrangian, it is possible to compute the Hamiltonian:

\mathcal H = \sum_i \left(p_i\dot q_i\right) - \mathcal L

In general, the Hamiltonian is a function or p and q and therefore it is necessary to solve the equation for the canonical momentum to retrieve \dot x and use it in the Hamiltonian:

\dot x_i = \frac{1}{m}\left(p_i - \frac{e}{c}A_i \right)

and using the canonical momentum Lagrangian in case of magnetic field:

\begin{aligned} \mathcal H & = \sum_i \left(p_i\dot x_i\right) - \sum_i \left[\frac{m}{2}\left( \dot x_i \right)^2 + \frac{e}{c} \dot x_i A_i \right] \\ & = \sum_i \left[\left(p_i\dot x_i\right) - \frac{m}{2}\left( \dot x_i \right)^2 - \frac{e}{c} \dot x_i A_i \right] \\ & = \sum_i \left[p_i \left( \frac{p_i}{m} - \frac{e}{m\,c}A_i \right) - \frac{m}{2}\left( \frac{p_i}{m} - \frac{e}{m\,c}A_i \right)^2 - \frac{e}{c} \left(\frac{p_i}{m} - \frac{e}{m\,c}A_i \right) A_i \right] \\ & = \sum_i \left[p_i \left( \frac{p_i}{m} - \frac{e}{m\,c}A_i \right) - \frac{m}{2}\left( \frac{p_i}{m} - \frac{e}{m\,c}A_i \right)^2 - \frac{e}{c} \left(\frac{p_i}{m} - \frac{e}{m\,c}A_i \right) A_i \right] \\ & = \sum_i \left[\frac{p_i^2}{m} - \frac{e\,p_i}{m\,c}A_i - \frac{p_i^2}{2\,m} + \frac{e\,p_i}{c\,m}A_i - \frac{e^2}{2\,m\,c^2}A_i^2 - \frac{e}{m\,c}p_iA_i + \frac{e^2}{m\,c^2}A_i^2 \right] \\ & = \sum_i \left[\frac{p_i^2}{2\,m} - \frac{e}{m\,c}p_iA_i + \frac{e^2}{2\,m\,c^2}A_i^2 \right] \\ & = \sum_i \left[\frac{1}{2\,m} \left(p_i^2 - \frac{2\,e}{c}p_iA_i + \frac{e^2}{c^2}A_i^2 \right) \right] \\ & = \sum_i \left[\frac{1}{2\,m} \left(p_i- \frac{e}{c}A_i \right)^2 \right] \end{aligned}

Incidentally, the Hamiltonian is just the kinetic energy:

\mathcal H = \sum_i \left[\frac{1}{2\,m} \left(p_i- \frac{e}{c}A_i \right)^2 \right] = \sum_i \frac{1}{2}m\ \dot x_i^2 = \frac{1}{2}m\,v^2

From the Hamiltonian it is possible to compute the equations of motion:

\begin{array}{c} \dfrac{\partial H}{\partial p_i} = \dot q_i \\[10pt] \dfrac{\partial H}{\partial q_i} = - \dot p_i \end{array}

Substituting:

\begin{aligned} \dot x_i & = \dfrac{\partial H}{\partial p_i} = \frac{1}{m}\left( p_i - \frac{e}{c}A_i \right) \\ \dot p_i & = - \dfrac{\partial H}{\partial x_i} = - \frac{1}{m} \sum_j \left( p_j - \frac{e}{c}A_j \right) - \frac{e}{c}\frac{ \partial A_j}{\partial x_i} \\ & = \frac{e}{c} \sum_j \left( p_j - \frac{e}{c}A_j \right) \frac{ \partial A_j}{\partial x_i} = \frac{e}{c} \sum_j \frac{ \partial A_j}{\partial x_i} \dot x_j \end{aligned}

The first equation results in this way because \mathbf A(\mathbf x) is not function of \mathbf p and for the i component it just depends from p_i; for the second the definition of \dot x_i.

It is now possible to write the equation of motion:

\begin{aligned} m\,a_x & = m\,\ddot x_i = \dot p_i -\frac{e}{c}\dot A_i = \dot p_i - \sum_j \frac{e}{c}\frac{\partial A_i}{\partial x_j}\dot x_j = \frac{e}{c} \sum_j \frac{ \partial A_j}{\partial x_i} \dot x_j - \sum_j \frac{e}{c}\frac{\partial A_i}{\partial x_j}\dot x_j \\ & = \frac{e}{c} \sum_{j\neq i} \left( \frac{ \partial A_j}{\partial x_i} - \frac{\partial A_i}{\partial x_j} \right) \dot x_j = \frac{e}{c} \left( \dot {\mathbf x} \times \mathbf B \right) \end{aligned}

This results coincides with the Newton-Lorentz equations of motion.

Exercise 11.5

Show that in the x, y plane, the solution to Eq. (25) and the solution to Eq. (26) are a circular orbit with the center of the orbit being anywhere on the plane. Find the radius of the orbit in terms of the velocity.

Considering a uniform magnetic field which lies in the z direction with a magnitude b; is possible to chose either A_x = -b\,y or A_y = b\,x, both are leading to the same equations of motions.

Starting from the case:

A_x = 0; \quad A_y = b\,x \quad A_z = 0

The Hamiltonian is:

\mathcal H = \sum_i \left[\frac{1}{2\,m} \left(p_i- \frac{e}{c}A_i \right)^2 \right] =\frac{1}{2\,m} \left[ p_x^2 + \left(p_y - \frac{e}{c}b\,x \right)^2 + p_z^2 \right]

As pointed out previously, the energy is conserved; since y and z are cyclic coordinates, p_y and p_z will be conserved. Since A_z = 0, p_z = m\,v_z and therefore the velocity in the z direction will be constant; however, p_y = c is not equivalent to the conservation of the velocity in the y direction, as the canonical momentum conjugated to y is p_y = m\,v_y + \frac{e}{c}b\,x.

\frac{\mathrm d}{\mathrm dt}\left( m\,v_y + \frac{e}{c}b\,x \right) = m\,a_y + \frac{e}{c}b\,v_x = 0

Therefore:

a_y = - \frac{e}{c} \frac{b}{m} v_x

Since the Hamiltonian has an explicit dependence on x, it is not possible to use a symmetry to compute p_x and therefore it is necessary to use Hamilton equations; it is however possible to change the gauge and use a different Hamiltonian in which x is now a cyclic coordinate. Considering the case:

A_x = -b\,y; \quad A_y = 0 \quad A_z = 0

The Hamiltonian is now:

\mathcal H = \sum_i \left[\frac{1}{2\,m} \left(p_i- \frac{e}{c}A_i \right)^2 \right] =\frac{1}{2\,m} \left[ \left(p_x + \frac{e}{c}b\,y \right)^2 + p_y^2 + p_z^2 \right]

Now the momentum conjugated to x is conserved:

\frac{\mathrm d}{\mathrm dt}\left( m\,v_x - \frac{e}{c}b\,y \right) = m\,a_x - \frac{e}{c}b\,v_y = 0

Therefore:

a_x = \frac{e}{c} \frac{b}{m} v_y

These are the Newton-Lorentz equations in the special case of uniform magnetic field.

Since there is no acceleration in the z direction, it is possible to assume that the velocity is zero and just consider the motion in the xy plane; the equations of motion are the one just derived:

\begin{aligned} & m\,a_x = \frac{e}{c} b\,v_y = \frac{e}{c} (\mathbf v \times \mathbf B)_x \\ & m\,a_y = -\frac{e}{c}b\, v_y = \frac{e}{c} (\mathbf v \times \mathbf B)_y \\ &m\,\mathbf a = \frac{e}{c} (\mathbf v \times \mathbf B) \end{aligned}

The acceleration is perpendicular to the velocity, but the magnitude of the velocity is constant, and also the magnitude of the field \mathbf B is constant, therefore the magnitude of the cross product is also constant; therefore it moves on a circle:

m\frac{v^2}{r} = \frac{e}{c}\,v\,b

This give the radius of the orbit:

r = \frac{m\,c}{e\,b}|v|

The position of the center is not explicit, so there is an infinite number of circular orbits, each for every point, with the radius that increase for higher velocities.

Appendix 1

Exercise A1.1

Show that Eq. (17) above, is a consequence of Equations (3) from Lecture 2.

From Lecture 2 the components of the velocity and the acceleration are:

\begin{aligned} & x(t) = r\,\cos(\omega\,t) \\ & y(t) = r\,\sin(\omega\,t) \\ & a_x(t) = -r\,\omega^2 \cos(\omega\,t) \\ & a_y(t) = -r\, \omega^2 \sin(\omega\,t) \end{aligned}

Therefore magnitude of the acceleration is:

a = \sqrt{a_x^2 + a_y^2} = \sqrt{r^2 \omega^4 \cos(\omega\,t)^2 + r^2 \omega^4 \sin(\omega\,t)^2} = \omega^2r

Go to the top of the page