Geometry Topology and Physics 1-1 Analytical Mechanics

博客分类: 格物 阅读次数: comments

Geometry Topology and Physics 1-1 Analytical Mechanics

Analytical mechanics

Newtonian mechanics

A conserved force is the gradient of a scalar function: $\mathbf{F}=-\nabla V(\mathbf{x})$.

Energy of a system only with conserved forces is conserved:

\[\begin{equation} E = \frac{m}{2}\left(\frac{\mathrm{d}\mathbf{x}}{\mathrm{d}t}\right)^2 + V(\mathbf{x}),\ \frac{\mathrm{d}E}{\mathrm{d}t} = 0. \tag{eq:NE} \end{equation}\]

Any one-dimensional force F(x) only explicitly dependent on x is conserved:

\[\begin{equation} V(x) = -\int^x F(\xi) \mathrm{d}\xi. \tag{eq:NP} \end{equation}\]

Lagrangian formalism

State of the system is described by generalized coordinates ${q_i} (i = 1,\cdots,N)$ in the configuration space M (a manifold). Generalized velocity is $\dot{q_i}$.

Principle of the least action

Integral form

The action is a functional from the trajectory $q(t)$ to a real number, defined by the integral of the Lagrangian from the initial to the final state of the system:

\[\begin{equation} S[q(t), \dot{q}(t)] = \int_{t_i}^{t_f} L(q,\dot{q}) \mathrm{d}t. \tag{eq:HA} \end{equation}\]

Hamilton’s principle/principle of the least action: The physically realized trajectory corresponds to an extremum of the action.

Variation form

Suppose q(t) is a path realizing an extremum of S. Consider a variation δq(t) of the trajectory such that $\delta q(t_i) = \delta q(t_f) = 0$. Under this variation,

\[\begin{equation} \begin{split} \delta S &= \int_{t_i}^{t_f} L(q+\delta q, \dot{q}+\delta \dot{q}) - L(q, \dot{q}) \mathrm{d}t\\ &= \int_{t_i}^{t_f}\frac{\partial L}{\partial q}\delta q\mathrm{d}t+\int_{t_i}^{t_f}\frac{\partial L}{\partial \dot{q}}\delta \dot{q}\mathrm{d}t + \mathcal{O}(\delta q^2) + \mathcal{O}(\delta \dot{q}^2)\\ &= \int_{t_i}^{t_f}\frac{\partial L}{\partial q}\delta q\mathrm{d}t+\int_{t_i}^{t_f}\frac{\partial L}{\partial \dot{q}}\mathrm{d}\delta q + \mathcal{O}(\delta q^2)\\ &= \int_{t_i}^{t_f}\frac{\partial L}{\partial q}\delta q\mathrm{d}t-\int_{t_i}^{t_f}\delta q\mathrm{d}\frac{\partial L}{\partial \dot{q}} + \mathcal{O}(\delta q^2)\\ &= \int_{t_i}^{t_f}\delta q\ \mathrm{d}t\ \left(\frac{\partial L}{\partial q}-\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{q}}\right) + \mathcal{O}(\delta q^2)\\ &\approx \int_{t_i}^{t_f}\delta q\ \mathrm{d}t\ \left(\frac{\partial L}{\partial q}-\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{q}}\right). \end{split} \tag{eq:PLA} \end{equation}\]

Euler-Lagrange equation

q(t) is an extremum, so $\delta S=0\ \forall\ \delta q$, which implies

\[\begin{equation} \frac{\partial L}{\partial q_k}-\frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial \dot{q}_k}=0\quad \text{for }k=1,\cdots,N, \tag{eq:ELE} \end{equation}\]

which is the Euler-Lagrange equation.

Using generalized momentum $p=\dfrac{\partial L}{\partial \dot{q}}$,

\[\begin{equation} \frac{\partial L}{\partial q_k}=\frac{\mathrm{d}p_k}{\mathrm{d}t}. \tag{eq:ELEk} \end{equation}\]

Let $L=\frac{1}{2}m\dot{q}^2-V(q)$, then (eq:ELEk) reduces to $p = m\dot{q},\ m\ddot{q}+\frac{\partial V}{\partial q}=0$.

Functional derivative

Using a single degree of freedom s or t, the parameter of the configuration trajectory, a functional derivative, considering the variation from the stationary path q(s) to another path q(t), can be defined by

\[\begin{equation} \begin{split} \frac{\delta S[q,\dot{q}]}{\delta q(s)}: &=\lim_{\varepsilon\to 0}\frac{1}{\varepsilon} \left\{S[q(t),\dot{q}(t)]-S[q(s),\dot{q}(s)]\right\}\\ &=\lim_{\varepsilon\to 0}\frac{1}{\varepsilon} \left\{ S[q(s)+\varepsilon\delta(t-s), \dot{q}(s)+\varepsilon\frac{\mathrm{d}} {\mathrm{d}t}\delta(t-s)] -S[q(s),\dot{q}(s)] \right\}, \end{split} \tag{eq:FD} \end{equation}\]

where $\varepsilon\delta(t-s)$ is a parametric expression of $\delta q$. $\varepsilon$ can be thought of a distance in configuration space between the two trajectories q(s), q(t), where they have parameter values of s and t, and $\delta(t-s)$ the unit vector pointing from point q(s) to point q(t). Note that the s and t here represent not a single value of the parameter (time), but the whole process of them ranging from the initial to the final state.

Therefore, substituting $\delta q=\varepsilon\delta(t-s)$ into (eq:PLA) yields

\[\begin{equation} \begin{split} \delta S &= \int_{t_i}^{t_f}\mathrm{d}t\ \varepsilon\ \left( \frac{\partial L}{\partial q}(s) -\frac{\mathrm{d}}{\mathrm{d}t} \frac{\partial L}{\partial \dot{q}}(s) \right) \delta(t-s) + \mathcal{O}(\varepsilon^2). \end{split} \tag{eq:PPLA} \end{equation}\]

In this flavor the Euler-Lagrangian Equation may be written as

\[\begin{equation} \frac{\partial L}{\partial q}(s) = \frac{\mathrm{d}}{\mathrm{d}t}\frac{\partial L}{\partial\dot{q}}(s) \tag{eq:PELE} \end{equation}\]

Symmetry of Lagrangian

Cyclic coordinate and momentum conservation

If Lagrangian L is independent of $q_k$ (but not $\dot{q}_k$), then $q_k$ is called cyclic. The conjugating momentum is conserved, according to (eq:PELE).

Symmetry

Infinitesimal symmetry operation: $q_k(t)\to q_k(t)+\delta q_k(t)$ on the path $q_k(t)$ from $t_i$ till $t_f$. Since L is independent of $q_k$, this Lagrangian should remain constant under the operation, at any point of the path.

For a particle in a spherically symmetric potential, the longitude angle $\phi$ is a cyclic coordinate. The corresponding conservative momentum is the angular momentum around the z axis.

Remarks:

Hamiltonian formalism

Hamiltonian is the Legendre transformation of the Lagrangian with respect to $\dot{q}$, into the phase space $(q_k,p_k)$.

\[\tag{eq:Hd} H(q,p):=\sum_k p_k\dot{q}_k - L(q,\dot{q}),\]

This transformation requires the Jacobian determinant to be non-zero.

\[\color{red}\det{\mathcal{J}} = \det{\frac{\partial p_i}{\partial \dot{q}_j}} = \det{\frac{\partial^2 L}{\partial \dot{q}_j\partial\dot{q}_i}} \neq 0. \tag{eq:detJ}\]

For Lagrangians without cross terms between the velocities like $L = \frac{1}{2}\sum_i{m_i\dot{q}_i^2}-V$, (eq:detJ) is simply $m_i\neq 0$.

An intuitive example is:

$L=\sum_k\frac{1}{2}m\dot{q}_k^2-V(q) = T-V$

$L+H = \sum_k p_k\dot{q_k}=\sum_k m\dot{q}_k^2 = 2T$

$H = \sum_k\frac{1}{2}m\dot{q}_k^2+V(q) = T+V$

Under an infinitesimal variation of $q_k$ and $p_k$,

\[\begin{equation} \begin{split} \delta H &= \sum_k\left[ \delta p_k\dot{q}_k + p_k\delta\dot{q}_k - \frac{\partial L}{\partial q_k}\delta q_k - \frac{\partial L}{\partial \dot{q}_k}\delta \dot{q}_k \right]\\ &= \sum_k\left[ \delta p_k\dot{q}_k - \frac{\partial L}{\partial q_k}\delta q_k \right], \end{split} \tag{eq:delH} \end{equation}\]

corresponding terms

\[\frac{\partial H}{\partial p_k} = \dot{q}_k,\quad \frac{\partial H}{\partial q_k} = -\frac{\partial L}{\partial q_k}. \tag{eq:HE1}\]

Substituting the Euler-Lagrange equation, we get

\[\frac{\partial H}{\partial p_k} = \dot{q}_k,\quad \frac{\partial H}{\partial q_k} = -\dot{p}_k, \tag{eq:HE2}\]

Substituting $L = \frac{1}{2}m\dot{\mathbf{q}}^2-V{\dot{\mathbf{q}}}$, then (eq:HE2) correspond to definition of momentum and Newton’s equation respectively.

Poisson bracket

\[[A,B] := \sum_k\left( \frac{\partial A}{\partial q_k}\frac{\partial B}{\partial p_k} -\frac{\partial A}{\partial p_k}\frac{\partial B}{\partial q_k}. \right) \tag{eq:PB}\]

The Poisson bracket is a Lie bracket, namely it is linear, antisymmetric, and Jacobi identical.

Jacobi identity: (see for Landau’s book)

\[[[A,B],C] + [[C,A],B] + [[B,C],A] = 0. \tag{eq:JI}\]

Fundamental Poisson brackets

\[[p_i,p_j]=[q_i,q_j]=0,\quad [q_i,p_j]=\delta_{ij}. \tag{eq:FPB}\]

Using Poisson brackets, the time development of $A(q,p)$ can be written as

\[\frac{\mathrm{d}A}{\mathrm{d}t}=[A,H]. \tag{eq:AP}\]

From (eq:AP) :

  1. $[A,H]=0$ implies that A is conserved.
  2. Hamilton’s equations can be written as
\[\frac{\partial H}{\partial p_k} = [q_k,H],\quad \frac{\partial H}{\partial q_k} = -[p_k,H]. \tag{eq:HEP}\]

Noether’s theorem

Suppose the Hamiltonian of a system is invariant under an infinitesimal coordinate transformation $q_k\to q’_k=q_k+\varepsilon f_k(q)$, then

\[Q=\sum_k p_kf_k(q) \tag{eq:NT}\]

is conserved.

Proof

The coordinate transform yields

\[\begin{equation} \begin{split} \frac{\partial q_i'}{\partial q_j} &= \frac{\partial}{\partial q_j}\left[q_i+\varepsilon f_i(q)\right] = \delta_{ij} + \varepsilon\frac{\partial f_i(q)}{\partial q_j} + \mathcal{O}(\varepsilon ^2),\\ \frac{\partial q_i}{\partial q_j'} &= \frac{\partial}{\partial q_j'}\left[q_i'-\varepsilon f_i(q)\right] = \delta_{ij} - \varepsilon\frac{\partial f_i(q)}{\partial q_j'} + \mathcal{O}(\varepsilon ^2), \end{split} \tag{eq:Lij} \end{equation}\]

where

\[\begin{equation} \begin{split} \frac{\partial f_i(q)}{\partial q_j'} &=\sum_k \frac{\partial q_k}{\partial q_j'} \frac{\partial f_i(q)}{\partial q_k}\\ &=\sum_k\left[ \delta_{kj} - \varepsilon\frac{\partial f_k(q)}{\partial q_j'} + \mathcal{O}(\varepsilon ^2) \right] \frac{\partial f_i(q)}{\partial q_k}\\ &=\frac{\partial f_i(q)}{\partial q_j} + \mathcal{O}(\varepsilon), \end{split} \tag{eq:Lji0} \end{equation}\]

so

\[\frac{\partial q_i}{\partial q_j'} = \frac{\partial}{\partial q_j'}\left[q_i'-\varepsilon f_i(q)\right] = \delta_{ij} - \varepsilon\frac{\partial f_i(q)}{\partial q_j} + \mathcal{O}(\varepsilon ^2). \tag{eq:Lji1}\]

Using (eq:Lji1), since $H$ is invariant, the momentum transforms as

\[\begin{equation} \begin{split} p_k\to p_k' &= -\frac{\partial H}{\partial q_k'} = -\sum_l \frac{\partial q_l}{\partial q'_k} \frac{\partial H}{\partial q_l}\\ &= -\sum_l \left[ \delta_{kl} - \varepsilon\frac{\partial f_l(q)}{\partial q_k} + \mathcal{O}(\varepsilon ^2) \right] \frac{\partial H}{\partial q_l}\\ &= -\frac{\partial H}{\partial q_k} + \varepsilon\frac{\partial f_l(q)}{\partial q_k} \frac{\partial H}{\partial q_l} + \mathcal{O}(\varepsilon ^2)\\ &= p_k - \varepsilon\frac{\partial f_l(q)}{\partial q_k} q_l + \mathcal{O}(\varepsilon ^2). \end{split} \tag{eq:MT} \end{equation}\]

The Hamiltonian is invariant implies that:

\[\begin{equation} \begin{split} 0 &=H(q_k',p_k')-H(q_k,p_k)\\ &=\sum_k\left[ \frac{\partial H}{\partial q_k} \varepsilon f_k(q) - \frac{\partial H}{\partial p_k} \varepsilon \sum_j p_j\frac{\partial f_j(q)}{\partial q_k} \right]\\ &=\varepsilon \sum_k\left[ \frac{\partial H}{\partial q_k} \frac{\partial Q}{\partial p_k} - \frac{\partial H}{\partial p_k} \frac{\partial Q}{\partial q_k} \right]\\ &= [H, Q] = \frac{\mathrm{d}Q}{\mathrm{d}t}. \end{split} \tag{eq:HQ} \end{equation}\]

which shows that Q is conserved.$\qquad\square$

Noether’s theorem shows that to find a conserved quantity is equivalent to finding a transformation which leaves the Hamiltonian invariant.

Generator of the transform

Since $q_i$ and $p_k$ between each other, $\frac{\partial q_i}{\partial p_k}=0$, the Poisson bracket

\[[q_i, Q] = \sum_k\left[ \frac{\partial q_i}{\partial q_k} \frac{\partial Q}{\partial p_k} - \frac{\partial q_i}{\partial p_k} \frac{\partial Q}{\partial q_k} \right] = \sum_k\delta_{ik} f_k = f_i(q), \tag{eq:qiQ}\]

which shows that $\delta q_i = \varepsilon f_i(q)=\varepsilon[q_i, Q]$. In this sense $Q$ is called the generator of the transformation $\delta q_i$.

System 1D particle 2D particle
Lagrangian $L = \frac{m\dot{r}^2}{2}$ $L=\frac{1}{2}m(\dot{r}^2+r^2\dot{\theta}^2)-V(r)$
Hamiltonian $H = \frac{p^2}{2m}$ $H=\frac{p_r^2}{2m}+\frac{p_{\theta}^2}{2mr^2}+V(r)$
Symmetry Translation Rotation around $z$
Transform of q $r\to r + \varepsilon$ $\theta\to \theta+\varepsilon$
Transform of p $p\to p$ $p_{\theta}\to p_{\theta}$
Generator Q $p$ $p_{\theta}=mr^2\dot{\theta}$
Conservation law Momentum Angular momentum around $z$

References

This article is based on:

  1. Nakahara, Mikio. Geometry, Topology, and Physics. 2nd ed. Graduate Student Series in Physics. Bristol ; Philadelphia: Institute of Physics Publishing, 2003.