28

The usual Schrödinger Lagrangian is $$ \tag 1 i(\psi^{*}\partial_{t}\psi ) + \frac{1}{2m} \psi^{*}(\nabla^2)\psi, $$ which gives the correct equations of motion, with conjugate momentum for $\psi^{*}$ vanishing. This Lagrangian density is not real but differs from a real Lagrangian density $$ \tag 2 \frac{i}{2}(\psi^{*}\partial_{t}\psi -\psi \partial_{t}\psi^{*} ) + \frac{1}{2m} \psi^{*}(\nabla^2)\psi $$ by a total derivative.

My trouble is that these two Lagrangian densities lead to different conjugate momenta and hence when setting equal time commutation relations, I am getting different results (a factor of 2 is causing the problem). I can rescale the fields but then my Hamiltonian also changes. Then applying Heisenberg equation of motion, I don't get the operator Schrödinger equation.

Is it possible to work with the real Lagrangian density and somehow get the correct commutation relations? I would have expected two Lagrangians differing by total derivative terms to give identical commutation relations (since canonical transformations preserve them). But perhaps I am making some very simple error. Unless all conjugate momenta are equivalent for two Lagrangians differing by total derivatives, how does one choose the correct one?

I guess the same thing happens for other first order systems like Dirac Lagrangian also.

Roger V.
  • 58,522
user5468
  • 381
  • 4
    I don't have time for a detailed reply to your question, but it may help to have a look at the end of Sec. 7.2 in Weinberg's textbook (vol. 1). He discusses the effect of adding a total time derivative to the Lagrangian and shows that while modifying the canonical momentum, it does not affect the commutation relations of the theory. – Tomáš Brauner Sep 30 '11 at 10:17
  • This may help: https://arxiv.org/abs/hep-th/0301052 – Quillo Apr 07 '21 at 19:43

1 Answers1

25

Here we will for simplicity only consider the Schrödinger system. We will assume that

$$\phi~=~(\phi^1+i\phi^2)/\sqrt{2} \tag{A}$$

is a bosonic complex field, and that

$$\phi^*~=~(\phi^1-i\phi^2)/\sqrt{2} \tag{B} $$

is the complex conjugate, where $\phi^a$ are the two real component fields, $a=1,2$. [Note the change in notation $\psi\longrightarrow\phi$ as compared with the OP's question (v1).]

  1. The Lagrangian density

$${\cal L}~:=~ i\phi^{*}\dot{\phi} + \frac{1}{2m} \phi^* \nabla^2\phi \tag{C} $$

for the Schrödinger field $\phi$ is already on the Hamiltonian form

$${\cal L}~=~ \pi\dot{\phi} - {\cal H}. \tag{D} $$

Simply define complex momentum

$$\pi~:=~i \phi^{\ast}, \tag{E} $$

and Hamiltonian density

$${\cal H}~:=~-\frac{1}{2m} \phi^{\ast} \nabla^2\phi. \tag{F} $$

More generally, this identification is a simple example of the Faddeev-Jackiw method.

  1. Recall that the Euler-Lagrange equations do not change by adding a $4$-divergence $d_{\mu}\Lambda^{\mu}$ to the Lagrangian density

$${\cal L} ~~\longrightarrow~~ {\cal L}^{\prime}~:=~{\cal L} + d_{\mu}\Lambda^{\mu},\tag{G}$$

cf. e.g. this Phys.SE post. [We use the symbol $d_{\mu}$ (rather than $\partial_{\mu}$) to stress the fact that the derivative $d_{\mu}$ is a total derivative, which involves both implicit differentiation through the field variables $\phi^a(x)$, and explicit differentiation wrt. $x^{\mu}$.] Therefore, we can (via spatial integration by parts) choose an equivalent Hamiltonian density

$$\begin{align}{\cal H} ~~\longrightarrow~~ {\cal H}^{\prime}~:=~&\frac{1}{2m}|\nabla\phi|^2\cr ~=~&\frac{1}{4m}(\nabla\phi^1)^2 +\frac{1}{4m}(\nabla\phi^2)^2,\end{align}\tag{H} $$

and we can (via temporal integrations by part) choose an equivalent kinetic term

$$\begin{align} i\phi^*\dot{\phi}~=~ \pi\dot{\phi} ~~\longrightarrow~&~ \frac{1}{2}(\pi\dot{\phi}-\phi\dot{\pi})\cr ~=~& \frac{i}{2}(\phi^*\dot{\phi}-\phi\dot{\phi}^*)\cr ~=~&\frac{1}{2}(\phi^2\dot{\phi}^1-\phi^1\dot{\phi}^2)\cr ~~\longrightarrow~&~\phi^2\dot{\phi}^1. \end{align}\tag{I} $$

The last expression shows (in accordance with the Faddeev-Jackiw method) that

$$ \text{The second component }\phi^2 \\ \text{ is the momenta for the first component }\phi^1. \tag{J}$$

  1. Alternatively, we can perform a Dirac-Bergmann analysis$^1$ directly. Consider for instance the Lagrangian density

$${\cal L}^{\prime}~=~ (\alpha+\frac{1}{2})\phi^2\dot{\phi}^1+(\alpha-\frac{1}{2})\phi^1\dot{\phi}^2 - {\cal H}^{\prime},\tag{K} $$

where $\alpha$ is an arbitrary real number. [The term $d(\phi^1\phi^2)/ dt$, which is multiplied by $\alpha$ in ${\cal L}^{\prime}$, is a total time derivative.] Let us check that the quantization procedure does not depend on this parameter $\alpha$. We introduce canonical Poisson brackets

$$\begin{align} \{\phi^a({\bf x},t),\phi^b({\bf y},t)\}_{PB} ~=~&0, \cr \{\phi^a({\bf x},t),\pi_b({\bf y},t)\}_{PB} ~=~&\delta^a_b ~ \delta^3 ({\bf x}-{\bf y}), \cr \{\pi_a({\bf x},t),\pi_b({\bf y},t)\}_{PB} ~=~&0,\end{align} \tag{L}$$

in the standard way. The canonical momenta $\pi_a$ are defined as

$$\begin{align} \pi_1~:=~&\frac{\partial {\cal L}^{\prime}}{\partial \dot{\phi}^1} ~=~(\alpha+\frac{1}{2})\phi^2,\cr \pi_2~:=~&\frac{\partial {\cal L}^{\prime}}{\partial \dot{\phi}^2} ~=~(\alpha-\frac{1}{2})\phi^1.\end{align}\tag{M}$$

These two definitions produce two primary constraints

$$\begin{align}\chi_1~:=~&\pi_1-(\alpha+\frac{1}{2})\phi^2~\approx~0,\cr \chi_2~:=~&\pi_2-(\alpha-\frac{1}{2})\phi^1~\approx~0,\end{align}\tag{N}$$

where the $\approx$ sign means equal modulo constraints. The two constraints are of second-class, because

$$ \{\chi_2({\bf x},t),\chi_1({\bf y},t)\}_{PB}~=~\delta^3 ({\bf x}-{\bf y})~\neq~0. \tag{O} $$

Thus the Poisson bracket should be replaced by the Dirac bracket. [There are no secondary constraints, because

$$\begin{align} \dot{\chi}_a({\bf x},t) ~=~&\{\chi_a({\bf x},t), H^{\prime}(t)\}_{DB} ~=~ 0, \cr H^{\prime}(t)~:=~& \int d^3y \ {\cal H}^{\prime}({\bf y},t),\end{align} \tag{P} $$

are automatically satisfied.] The Dirac bracket between the two $\phi^a$'s is

$$\{\phi^1({\bf x},t),\phi^2({\bf y},t)\}_{DB}~=~\delta^3 ({\bf x}-{\bf y}), \tag{Q}$$

leading to the same conclusion (J) as the Faddeev-Jackiw method. Note that the eqs. (O) and (Q) are independent of the parameter $\alpha$.

  1. In all cases, the canonical equal-time commutator relations for the corresponding operators become

$$\begin{align} [\hat{\phi}^1({\bf x},t), \hat{\phi}^2({\bf y},t)] ~=~& i\hbar {\bf 1}~\delta^3 ({\bf x}-{\bf y}), \cr [\hat{\phi}({\bf x},t), \hat{\phi}^{\dagger}({\bf y},t)] ~=~& \hbar {\bf 1}~\delta^3 ({\bf x}-{\bf y}), \cr [\hat{\phi}({\bf x},t), \hat{\pi}({\bf y},t)] ~=~& i\hbar {\bf 1}~\delta^3 ({\bf x}-{\bf y}).\end{align} \tag{R}$$

--

$^1$ See, e.g., M. Henneaux and C. Teitelboim, Quantization of Gauge Systems, 1992.

Qmechanic
  • 201,751
  • Thanks a lot to Qmechanic for very detailed answer. I really appreciate the help. – user5468 Oct 08 '11 at 17:39
  • 1
    Notes for later: The Schrödinger action for multiple particles is $$I[\psi]~=~\int! dt~\left[ \prod_{j=1}^N d^3{\bf x}_j \right] {\cal L}.$$ – Qmechanic Jun 26 '18 at 14:02
  • 1
    Notes for later: The Schrödinger Lagrangian density is $$ {\cal L} ~=~\frac{i\hbar}{2}(\psi^{\ast}\dot{\psi}-\dot{\psi}^{\ast}\psi) - \sum_{j=1}^N\frac{\hbar^2}{2m_j}|\nabla_j\psi|^2 -V |\psi|^2$$ $$~=~ -\rho\dot{S}-\sum_{j=1}^N\frac{\hbar^2}{2m_j}(\nabla_j\sqrt{\rho})^2-\sum_{j=1}^N\frac{\rho}{2m_j}(\nabla_j S)^2 -\rho V,$$ where we rewrote the wavefunction $\psi=\sqrt{\rho}\exp\left(\frac{i}{\hbar}S\right)$ in "polar" coordinates $\rho$ and $S$. – Qmechanic Jun 26 '18 at 14:12
  • 1
    Notes for later: Case $N=1$: Note that $\rho$ and $S$ are canonical variables ${\rho({\bf x}),S({\bf y})}=\delta^3({\bf x}-{\bf y})$, and that ${\cal L}$ is already on Hamiltonian first-order form. The Schrödinger picture suggests second quantization, cf. e.g. this Phys.SE post. – Qmechanic Jun 26 '18 at 14:59
  • 1
    Notes for later: The EL eqs. wrt. $\rho$ and $S$ are $$\dot{S} - \sum_{j=1}^N\frac{\hbar^2}{2m_j\sqrt{\rho}}\nabla_j^2\sqrt{\rho} +\sum_{j=1}^N\frac{1}{m_j}(\nabla_jS)^2 + V ~\approx~ 0\qquad\text{and}\qquad \dot{\rho} + \sum_{j=1}^N\frac{1}{m_j}\nabla_j \cdot (\rho \nabla_jS) ~\approx~ 0,$$ respectively. The Madelung equations with ${\bf u}_j :=\frac{1}{m_j}\nabla_jS$ are straightforward consequences (and analogues of Navier–Stokes eqs). See also de Broglie–Bohm pilot-wave theory. – Qmechanic Jun 29 '18 at 16:55
  • @Qmechanic The Lagrangian form that you give differs form Eq. (C) in the answer. I agree with it. – my2cts Nov 19 '21 at 14:28
  • Notes for later: Free 1+1D Schrödinger Lagrangian density $\quad{\cal L}=\frac{1}{2}\left(\pi\dot{\phi}-\phi\dot{\pi}\right)-{\cal H}$, $\quad{\cal H}=\frac{1}{4m}\left(\phi^{\prime 2}+\pi^{\prime 2}\right)$, $\quad\psi=\frac{\phi+i\pi}{\sqrt{2}}$. $\phi$-translation symmetry has Noether charge density $\rho=\frac{\pi}{2}$. $\pi$-translation symmetry has Noether charge density $\rho=-\frac{\phi}{2}$. Global $U(1)$ phase symmetry $\delta_0\psi=-i\epsilon\psi$, $\delta_0\phi=\epsilon\pi$, $\delta_0\pi=-\epsilon\phi$ has Noether charge density $\rho={\cal I}=|\psi|^2=\frac{\phi^2+\pi^2}{2}$. – Qmechanic Jun 03 '23 at 11:48
  • $t$-translation symmetry has Noether charge density $\rho={\cal H}$. $\quad\delta_0\phi=-\frac{\epsilon}{2m}\pi^{\prime\prime}$, $\quad\delta_0\pi=\frac{\epsilon}{2m}\phi^{\prime\prime}$. $\quad\delta_0\psi\approx \epsilon\dot{\psi}$. $x$-translation symmetry has Noether charge density $\rho={\cal P}=\frac{\pi\phi^{\prime}-\phi\pi^{\prime}}{2}$. $\quad\delta_0\psi=\epsilon\psi^{\prime}$. There is a worldsheet representation of algebra ${\cal H}\to\partial_t$, ${\cal P}\to\partial_x$, ${\cal I}\to\hat{\bf 1}$. Roughly $\quad H\sim \frac{1}{2m}P^2$. $\quad D\sim PG\sim I$. $\quad C\sim G^2$. – Qmechanic Jun 03 '23 at 11:55
  • $\quad{P[f],H[g]}=\frac{1}{4m}\int! dx \left((f\phi^{\prime}+(f\phi)^{\prime})(\phi^{\prime}g)^{\prime} +(\phi\leftrightarrow \pi) \right)\longrightarrow 0$. $\quad{H[f],I[g] }=\frac{1}{2m}\int! dx \left(f\phi^{\prime}(g\pi)^{\prime}-f\pi^{\prime}(g\phi)^{\prime}\right)\longrightarrow 0$. $\quad{I[f],P[g] }=\int! dx \left(f\phi\phi^{\prime}g+f\phi(\phi g)^{\prime}+(\phi\leftrightarrow \pi) \right)\longrightarrow 0$. Field theoretic problems with boundary terms. Bulk target space algebra: $\quad{P,D}=P$, $\quad{H,D}=2H$, $\quad{P,G}=mI$, etc. – Qmechanic Jun 03 '23 at 12:34
  • Dilation symmetry $\delta t=-2\epsilon t$, $\delta x=-\epsilon x$, $\delta\psi=\frac{\epsilon}{2}\psi$, $\delta_0\psi=\frac{\epsilon}{2}\psi+2\epsilon t\dot{\psi}+\epsilon x\psi^{\prime}$ has Noether charge density $\rho={\cal D}=x{\cal P}+2t{\cal H}$. – Qmechanic Jun 03 '23 at 13:41
  • Galilean boost symmetry $\delta x =-\epsilon t$, $\delta\psi=-i\epsilon mx \psi$, $\delta\phi=\epsilon mx \pi$, $\delta\pi=-\epsilon mx \phi$, $\delta_0\phi=\epsilon mx \pi+\epsilon t \phi^{\prime}$, $\delta_0\pi=-\epsilon mx \phi+\epsilon t \pi^{\prime}$ has Noether charge density $\rho={\cal G}=mx{\cal I} +t{\cal P}$. – Qmechanic Jun 03 '23 at 15:13
  • Schrödinger group. Conformal/expansion symmetry $\delta t=-\epsilon t^2$, $\delta x=-\epsilon xt$, $\delta\psi=-\frac{i\epsilon}{2}mx^2\psi$, $\delta\phi=\frac{\epsilon}{2}mx^2\pi$, $\delta\pi=-\frac{\epsilon}{2}mx^2\phi$ has Noether charge density $\rho={\cal C}=mx^2{\cal I} +tx{\cal P}+t^2{\cal H}$. https://doi.org/10.1103/PhysRevD.5.377 eq. (2.13). https://arxiv.org/abs/1304.7293 top of p. 6. https://lbt.usal.es/wp-content/uploads/2016/06/PAPER.pdf eq. (13). – Qmechanic Jun 04 '23 at 08:38
  • Symmetries of SHO: $\quad L_H=i\bar{z}\dot{z}-\omega|z|^2 =i\bar{z}e^{-i\omega t} \frac{d}{dt}(e^{i\omega t}z)$. $\quad \mathbb{L}_H=i\bar{z}\mathrm{d}z-\omega|z|^2\mathrm{d}t =i\bar{z}e^{-i\omega t}\mathrm{d}(e^{i\omega t}z)$. Liouville-integrable. Solution: $e^{i\omega t}z\approx {\rm const}$. M. Lutzky https://doi.org/10.1088/0305-4470/11/2/005 Lagrangian formulation. The symmetry vector field is assumed to not depend on velocity. Eq. (24) has only 5 symmetries. – Qmechanic Jun 22 '23 at 14:30
  • Infinitesimal quasi-symmetries: 1. $t$-translation: $\delta t=\epsilon$. 2. $U(1)$-phase $\delta z = i\epsilon z$. 3. $\delta z = e^{-i\omega t}(\epsilon_1+i\epsilon_2)$. 4. Local reparametrization trivial gauge symmetry: $\delta t =\epsilon(t)$, $\delta z =-i\omega z \epsilon(t)$, $\delta_0 z = -(i\omega z+\dot{z})\epsilon(t)$. Noether charges: 1. Hamiltonian: $\omega|z|^2$. 2.$|z|^2$. 3. Initial conditions: $e^{-i\omega t}\bar{z}$. 4. None. So the most general constant of motion seems to be $f(|z|^2,e^{i\omega t}z, e^{-i\omega t}\bar{z})$ $=f(e^{i\omega t}z, e^{-i\omega t}\bar{z})$. – Qmechanic Jun 24 '23 at 07:14