Sunday, January 31, 2016

cosmology - If the universe is flat, does that imply that the Big Bang produced an infinite amount of energy?


Too much density and the universe is closed, analogous to a sphere in four dimensions: you travel in a straight line and you end up where you started. Too little and you have a saddle: not sure about the destination if you travel in a straight line. Just the right amount and the topology is flat. The flat topology is infinite: you travel in a straight line forever.


If the topology is flat (and at this point all evidence indicates that it is to within 0.4%), then multiplying the critical density by an infinite amount of cubic meters gives you an infinite energy/stress.$$\rho_{CRIT}\space kg\space m^{-3}\times \infty\space m^3=\infty\space kg$$


Is that a reasonable interpretation?




Why can't the Klein-Gordon equation explain the hydrogen atom but the Dirac equation does?


Why can't the Klein-Gordon equation with a Couloumb potential describe the hydrogen atom?



Why can the first order Dirac equation explain it? What are the failures?




quantum mechanics - How does the energy of a single photon manifest itself as the frequency of the classical EM wave?


We know that the energy of the photon is given by $\hbar\omega$ and it so happens that this exact $\omega$ is the frequency we would obtain from the many photon EM wave. How does one relate the two?


Is the energy spacing between states of a harmonic oscillator related to the coherent state formed by the system? If so, how does one extend this to the case of EM waves?



Answer



There is a standard way to quantize (non-relativisticly) the EM field. Based on the classical energy density $$H = \frac{1}{8\pi}\int\! d^3r \left[|\vec{E}(\vec{r})|^2+|\vec{B}(\vec{r})|^2\right]$$ We write everything in terms of the vector potential $$\vec{A}(\vec{r},t)$$ which we expand in plain waves $\vec{q}_{\vec{k}}(t) e^{i\vec{k} \vec{r}}$. Note that we do not make any assumptions about the time-dependent, or the frequency. That will come out naturally. Then from the Hamiltonian is $$ H = \sum_{\vec{k}} |\dot{\vec{q}}_{\vec{k}}|^2 + \omega_{\vec{k}} |\vec{q}_{\vec{k}}|^2$$ with $\omega_{\vec{k}} = c|\vec{k}|$. Now we quantize these modes, with $p=\dot{q}$ the conjugate momenta of $q$, and you see that the EM field is described as sum of Harmonic oscillators. The creation and annihilation operators of these "Harmonic oscillators" add or remove a quanta of $\hbar\omega_{\vec{k}}$ from the field, and these are the photons.


You can work out how $\vec{E}$ and $\vec{B}$ look in terms of these fields, and you get that $\vec{E}$ has expansion in modes that propagate like the calssical fields, that is with an exponent of $i(\vec{k}\vec{r}-\omega_\vec{k} t)$.


A nice point here is that for a state of the field with a well-defined number of photons, the expectation value of both the electric and magnetic field is zero (just like that the momenta and position of the Harmonic oscillator is zero for a state with well defined $n$). You need a coherent state in order to describe the classical limit. In fact - this is the origin of the term "coherent state", as it was born in quantum optics!



More detailed equations can be found in the wikipedia page on the subject.


quantum mechanics - What is the usefulness of the Wigner-Eckart theorem?


I am doing some self-study in between undergrad and grad school and I came across the beastly Wigner-Eckart theorem in Sakurai's Modern Quantum Mechanics. I was wondering if someone could tell me why it is useful and perhaps just help me understand a bit more about it. I have had two years of undergrad mechanics and I think I have a reasonably firm grasp of the earlier material out of Sakurai, so don't be afraid to get a little technical.



Answer




I will not get into theoretical details -- Luboš ad Marek did that better than I'm able to.
Let me give an example instead: suppose that we need to calculate this integral:


$\int d\Omega (Y_{3m_1})^*Y_{2m_2}Y_{1m_3}$


Here $Y_{lm}$ -- are spherical harmonics and we integrate over the sphere $d\Omega=\sin\theta d\theta d\phi$.


This kind of integrals appear over and over in, say, spectroscopy problems. Let us calculate it for $m_1=m_2=m_3=0$:


$\int d\Omega (Y_{30})^*Y_{20}Y_{10} = \frac{\sqrt{105}}{32\sqrt{\pi^3}}\int d\Omega \cos\theta\,(1-3\cos^2\theta)(3\cos\theta-5\cos^3\theta)=$


$ = \frac{\sqrt{105}}{32\sqrt{\pi^3}}\cdot 2\pi \int d\theta\,\left(3\cos^2\theta\sin\theta-14\cos^4\theta\sin\theta+15\cos^6\theta\sin\theta\right)=\frac{3}{2}\sqrt{\frac{3}{35\pi}}$


Hard work, huh? The problem is that we usually need to evaluate this for all values of $m_i$. That is 7*5*3 = 105 integrals. So instead of doing all of them we got to exploit their symmetry. And that's exactly where the Wigner-Eckart theorem is useful:


$\int d\Omega (Y_{3m_1})^*Y_{2m_2}Y_{1m_3} = \langle l=3,m_1| Y_{2m_2} | l=1,m_3\rangle = C_{m_1m_2m_3}^{3\,2\,1}(3||Y_2||1)$


$C_{m_1m_2m_3}^{j_1j_2j_3}$ -- are the Clebsch-Gordan coefficients



$(3||Y_2||1)$ -- is the reduced matrix element which we can derive from our expression for $m_1=m_2=m_3=0$:


$\frac{3}{2}\sqrt{\frac{3}{35\pi}} = C_{0\,0\,0}^{3\,2\,1}(3||Y_2||1)\quad \Rightarrow \quad (3||Y_2||1)=\frac{1}{2}\sqrt{\frac{3}{\pi}}$


So the final answer for our integral is:


$\int d\Omega(Y_{3m_1})^*Y_{2m_2}Y_{1m_3}=\sqrt{\frac{3}{4\pi}}C_{m_1m_2m_3}^{3\,2\,1}$


It is reduced to calculation of the Clebsch-Gordan coefficient and there are a lot of, tables, programs, reduction and summation formulae to work with them.


Saturday, January 30, 2016

quantum mechanics - Where does ket vector live in rigged Hilbert space?


Let's say rigged Hilbert space $(\mathcal{S},\mathcal{H},\mathcal{S}^{*})$ in Gelfand triple. Where would ket vector live in? Would it be $\mathcal{S}$? That is what I thought, but https://arxiv.org/abs/quant-ph/0502053 suggests that it is actually in anti-dual space of $\mathcal{S}$. (Page 3) So is this article right, and I am wrong?



Answer




The paper is correct. Note that in a rigged Hilbert space $(\mathcal S, \mathcal H, \mathcal S^*)$, we have that $\mathcal S \subset \mathcal H \subset \mathcal S^*$. That is, $\mathcal S$ (the set of so-called test functions) is a subset of $\mathcal H$. The only reason the rigged Hilbert space construction is required in the first place is that ket vectors corresponding to definite states of continuous observables like $X$ and $P$ aren't actually elements of $\mathcal H$, which means they're definitely not elements of a subset of $\mathcal H$.




I'll be more explicit to address your additional question. Let $\mathcal H$ be $L^2(\mathbb R)$, which is (roughly) the space of square integrable functions $f:\mathbb R \rightarrow \mathbb C$. Additionally, let $\mathcal S$ be the space of rapidly-decaying smooth functions, defined as follows:


$$\mathcal S := \left\{f \in C^\infty(\mathbb R) : \forall m,n\in\mathbb N, \sup_x\left| x^n \cdot \frac{d^mf}{dx^m }\right|<\infty \right\}$$


Essentially, $\mathcal S$ is the space of all functions to which you can apply the position and momentum operators as many times as you want, with the result still being bounded. It's not difficult to show that $\mathcal S\subset L^2(\mathbb R)$. It's less clear that $\mathcal S$ is dense in $L^2(\mathbb R)$, but that also happens to be true.


A linear functional on $\mathcal S$ is a map $\varphi:\mathcal S \rightarrow \mathbb C$ such that for all $f,g\in\mathcal S$ and $\lambda\in\mathbb C$,



  • $\varphi(f + g) = \varphi(f)+\varphi(g)$

  • $\varphi(\lambda f) = \lambda \varphi(f)$



An antilinear functional is the same, except $\phi(\lambda f)=\bar{\lambda} \phi(f)$ where the bar denotes complex conjugation.


The space of all linear functionals (which we will identify as bras) on $\mathcal S$ is called the dual space $S'$, while the set of all antilinear functionals (which we will identify as kets) on $\mathcal S$ is called the antidual space $S^*$.


Observe that any element $f\in\mathcal H$ can be identified with a linear functional $\varphi_f \equiv \langle f, \bullet \rangle$, which acts on some $g\in\mathcal S$ as follows: $$\varphi_f(g) = \langle f,g\rangle$$ Furthermore, it can be identified with an antilinear functional $\phi_f \equiv \langle \bullet , f \rangle$ as well: $$\phi_f(g) = \langle g,f\rangle$$ This implies at least that $\mathcal H \subseteq S'$ and $\mathcal H \subseteq S^*$. However, the spaces $S'$ and $S^*$ are much bigger than $\mathcal H$.


Momentum Eigenfunctions


Observe that the function $e^{ikx}$, which is not square integrable and thus not an element of $\mathcal H$, can be identified with the dual space element $\varphi_k$ and the antidual space element $\phi_k$ where for all $g\in \mathcal S$,


$$\varphi_k(g) = \int_{-\infty}^\infty \overline{e^{ikx}} g(x) dx = \int_{-\infty}^\infty e^{-ikx}g(x) dx$$ and $$\phi_k(g) = \int_{-\infty}^\infty \overline{g(x)} e^{ikx} dx $$


You are more used to the bra-ket notation, in which $\varphi_k \equiv \langle k|$ and $\phi_k \equiv |k\rangle$.


Position Eigenfunctions


$S'$ and $S^*$ also contain elements that don't correspond to functions at all. The Dirac delta distribution $\delta_a$ is deceptively simple - it just evaluates a function at $a$. Define the linear functional $\delta_a$ and the antilinear functional $\delta^*_a$ simply as


$$\delta_a (g) = g(a)$$ $$\delta^*_a (g) = \overline{g(a)}$$



You are again more familiar with the notation $\delta_x \equiv \langle x|$ and $\delta^*_x \equiv |x\rangle$.


particle physics - Random directions of the trajectories in a cloud chamber



It is easy to construct a home-made cloud chamber. In observation, one finds that the directions of the trajectories are quite random. Does this mean that all the particles detected are secondary cosmic rays?



Answer



It turns out there is a lot of interesting physics involved in explaining this.


Cosmic rays are about 90% protons and 10% alpha particles (with small amounts of other objects). However these mostly get stopped by collisions in the upper atmosphere. The collisions mostly create muons, so the cosmic rays we see at ground level are mostly muons. Strictly speaking these are already secondary cosmic rays, so I guess that would make the tracks you observe in the cloud chamber tertiary rays, but lets overlook this nicety for now.


The tracks you see in the cloud chamber start when a high energy cosmic ray muon collides with and ionises an air molecule. The collision ejects an electron from the molecule and it's that electron that leaves the track. So the track starts at the site of the muon-molecule collision and fades out after a few cm as the electron loses energy.


So the question is why the high energy muon doesn't leave a track while the low energy electron does. And the answer is (as I've implied by mentioning the energy) that in general the energy loss per unit length is higher for low energy particles like the electrons that it is for high energy particles like the muons i.e. the electrons interact more strongly with the air in the chamber than the muons do.


The energy loss per unit length is described by the Bethe formula. Rather than just quote the complicated and largely incomprehensible formula let me just show the graph of energy loss as a function of energy from that Wikipedia article:


Bethe formula


The energy of the muons created by the cosmic rays is around 4GeV at the Earth's surface so that's off the right hand end of the graph meaning that those muons interact only weakly. Offhand I don't know the energy of the electrons that leave the tracks, but I would guess they are around $1-10$ MeV and therefore they interact far more strongly.


thermodynamics - Maxwell-Boltzmann distribution, average speed in one direction



Consider an ideal gas obeying the Maxwell-Boltzmann distribution i.e


$$f(v) = \bigg(\frac{m}{2 \pi k_{B} T}\bigg)^{3/2} \exp \left(-\frac{m v^{2}}{2 k_{B} T} \right) \, .$$


The probability distribution in 3D velocity space ($v^{2} = v_{x}^2 + v_{y}^2 + v_{z}^2$). How might you determine the average speed the particles are moving at, $\langle |v_{z}| \rangle$, in one direction?


Additionally if my ideal gas is now confined to a hemisphere in velocity space i.e we have the conditions $- \infty \leq v_{x}, v_{y} \leq \infty$ and $ 0 \leq v_{z} \leq \infty$ but it still has a Maxwell Boltzmann velocity distribution (except I think the normalization factor on $f(v)$ should change) then what is the average speed, or velocity, in the z direction $\langle v_{z} \rangle$, will this be the same as $|\langle v_{z}| \rangle$ from the previous answer?




Friday, January 29, 2016

quantum mechanics - Definite Parity of Solutions to a Schrödinger Equation with even Potential?


I am reading up on the Schrödinger equation and I quote:



Because the potential is symmetric under $x\to-x$, we expect that there will be solutions of definite parity.



Could someone kindly explain why this is true? And perhaps also what it means physically?



Answer



Good question! First you need to know that parity refers to the behavior of a physical system, or one of the mathematical functions that describe such a system, under reflection. There are two "kinds" of parity:



  • If $f(x) = f(-x)$, we say the function $f$ has even parity


  • If $f(x) = -f(-x)$, we say the function $f$ has odd parity


Of course, for most functions, neither of those conditions are true, and in that case we would say the function $f$ has indefinite parity.


Now, have a look at the time-independent Schrödinger equation in 1D:


$$-\frac{\hbar^2}{2m}\frac{\mathrm{d}^2}{\mathrm{d}x^2}\psi(x) + V(x)\psi(x) = E\psi(x)$$


and notice what happens when you reflect $x\to -x$:


$$-\frac{\hbar^2}{2m}\frac{\mathrm{d}^2}{\mathrm{d}x^2}\psi(-x) + V(-x)\psi(-x) = E\psi(-x)$$


If you have a symmetric (even) potential, $V(x) = V(-x)$, this is exactly the same as the original equation except that we've transformed $\psi(x) \to \psi(-x)$. Since the two functions $\psi(x)$ and $\psi(-x)$ satisfy the same equation, you should get the same solutions for them, except for an overall multiplicative constant; in other words,


$$\psi(x) = a\psi(-x)$$


Normalizing $\psi$ requires that $|a| = 1$, which leaves two possibilities: $a = +1$ (even parity) and $a = -1$ (odd parity).



As for what this means physically, it tells you that whenever you have a symmetric potential, you should be able to find a basis of eigenstates which have definite even or odd parity (though I haven't proved that here,* only made it seem reasonable). In practice, you get linear combinations of eigenstates with different parities, so the actual state may not actually be symmetric (or antisymmetric) around the origin, but it does at least tell you that if your potential is symmetric, you could construct a symmetric (or antisymmetric) state. That's not guaranteed otherwise. You'd probably have to get input from someone else as to what exactly definite-parity states are used for, though, since that's out of my area of expertise (unless you care about parity of elementary particles, which is rather weirder).




*There is a parity operator $P$ that reverses the orientation of the space: $Pf(x) = f(-x)$. Functions of definite parity are eigenfunctions of this operator. I believe you can demonstrate the existence of a definite-parity eigenbasis by showing that $[H,P] = 0$.


New direction vector after collision of spheres



I have a volume in 3-space in which random spheres are spawned in motion. They have the following attributes to them:



  • position known (in three axes)

  • a direction vector (in three axes)

  • a scalar speed

  • a radius (all radii are equal)



I can determine if a collision is made from objects position and radius using the below formula:


$$d = \sqrt{(x_2-x_1)^2 + (y_2-y_1)^2 + (z_2-z_1)^2} - 2r$$


If $d = 0$, then a collision has happened. What I want to know is how to determine the new direction vector after impact. I'm not working with angles.


I considered that if spheres $A$ and $B$ collided, and $A_x > B_x$, then $A$ is to the right of $B$, thus the new direction of movement for $A$ would be $D_x = -D_{old x}$, same thing for $B$, but I'm not sure if this is entirely correct.


If it makes a difference, there's no air resistance or friction.



Answer



Consider the motion of the two spheres in their center of mass frame: that is, the frame of reference moving at a velocity


$$\vec{v_f}=\frac{\vec{v_1}+\vec{v_2}}{2}$$


Subtract that velocity from the "stationary world" coordinates; then perform the analysis below. Finally, add that velocity back to the result.



At the moment the two spheres touch, draw a line between their centers. Along that direction their velocities are reversed; the other component of their velocity (perpendicular to that axis) is unchanged.


So if particle 1 has initial vector $\vec{v_1}$, and the direction of the line connecting the centers at the moment of collision is $\vec{r}$ (normalized), then the velocity along the connecting line is


$$\vec{v_{1a}}=\left(\vec{v_1}\cdot\vec{r}\right)\;\vec{r} $$


and the velocity perpendicular to the connecting line is


$$\vec{v_{1p}} = \vec{v_1}-\vec{v_{1a}}$$


After collision, the "along" velocity is reversed, and the new velocity is


$$\vec{v_{after}}=\vec{v_{1p}} - \vec{v_{1a}}= \vec{v_1}-2\vec{v_{1a}}$$


Thursday, January 28, 2016

quantum mechanics - What is the physical intuition behind the fact that 'energy is not continuous'?


First of all I am a novice regarding my knowledge of quantum mechanics. But curiously I do want to know what is the problem if energy is continuous like spontaneously flowing tap water.


In fact I actually don't know what is continuous referring here. What does this statement mean? Energy is the ability to do work. So, what is the problem with continuous energy?



Plank solved the problem of Classical physics by chopping up the energy into discrete particles which possess energy proportional to the frequency of the radiation.



So, what is this? What led Planck to chop the energy? What is the physical intuition behind this? Please help giving me a math-free explanation.



Answer



Let me first give you an example on what 'continuous' and 'discrete' are, and then show you how it relates to energy.



Let's say water is flowing in a stream. We say that the water flow is 'continuous' since we don't see induvidual 'blocks' or 'lumps' of water flowing one after the other in the stream. All we see is one continuous indivisible 'body' of water which negotiates rocks and ridges with ease by simply 'flowing' over them.


It may seem apparent at first that energy also behaves in the same manner. The way we think of energy untuitively is that it is some sort of invisible continuous fluid which floats in space from body to body. The evidence to support this intuition physically is not scanty.


Look at the example of visible spectra. This is what you get when you pass a beam of white light through a prism. Prior to the 20th century, the theory that explained the properties of light satisfactorily was the Wave theory of light. According to it, each color of light corresponds to a particular wavelength, and thus also corresponds to a particular energy with which it moves. Since this spectra is continuous, it means that the white light has all energy levels or colors, there are no gaps or sudden color changes. Can you make out exactly where, for example, yellow turns orange ? enter image description here


But then take a look at this spectra too: enter image description here


Do you see bold, discrete and distinct lines of red and blue ? This is a hydrogen spectra. The colors are the different wavelengths of light emitted by hydrogen gas filled in a tube when current is passed through it. These distinct lines baffled 19th century physicists. They couldn't understand how only particular 'discrete' wavelengths of light could be emitted by hydrogen atoms and not 'continuously' like the visible spectra.


At the start of the 20th century, Max Planck was conducting an experiment concerning black body radiation. By chance, he observed that in the energy values he tabulated, the values were always integral multiples of $h\nu$, where $\nu$ is the frequency of radiation emitted by the black body. It conlusively meant that energy could only be emitted in specified amounts and had an elementary, basic unit much like the 'lumps' of water I described in the second paragraph. If energy is transmitted in space from body to body, it means that that energy is made up of such elementary 'lumps' called 'quanta' (later they termed it as 'photon', a particle-like entity), much like all mass is made up of 'atoms'.


The 'discrete' wavelengths in the hydrogen spectra was due to the atoms collectively emitting 'discrete' photons of specified energy and not of any energy (as in the continuous spectrum) as electrons jumped to lower energy levels from higher ones.


Another question may arise: 'Why' does energy have to only be transmitted in discrete, specified quanta ? The answer ultimately is that it 'just is'. Later in the 20th century, Einstein showed that energy and mass are two sides of the same coin. So one may also argue that: Just as mass has a quanta called 'atom', energy also may have such a quanta in the form of 'photons'. Eitherwise, none of these can be given as 'the one and only reason' since it is purely a question of philosophy, not physics.


P.S: As julian fernandez said, it will take a really long time to fully type out the origins of quantum mechanics. What I have given is just a brief intro to the field.


Hope it helps !



electricity - Why do bulbs glow brighter when connected in parallel?


Consider a circuit powered by a battery. If light bulbs are attached in parallel, the current will be divided across all of them. But if the light bulbs are connected in series, the current will be the same in all of them. Then it looks like the bulbs should be brighter when connected in series, but actually, they are brighter when connected in parallel. Why is that?




Answer



The bulbs will only appear brighter if the available current to the system is not limited. In that case the series bulbs will have a lower voltage across each individual bulb and they will appear dimmer. If the power input to the circuit is a constant than the total wattage output from all bulbs is also constant and the bulbs will all appear the same (assuming the filaments for the bulbs are all identical resistance).


In a typical simple circuit the power source will be a battery which attempts to hold a constant voltage across the circuit. In this case the voltage across the bulbs in parallel will be equal to the voltage of the battery and the current through the bulb will be defined by $V = IR$ where $R$ is the resistance of the filament. This means more current (and thus more power) will be drawn from a battery into the parallel circuit than a series one and the parallel circuit will appear brighter (but will drain your battery faster).


classical mechanics - How did Feynman derive the physics of medallion vs. plate wobble rate?


I am referring to this:



Within a week I was in the cafeteria and some guy, fooling around, throws a plate in the air. As the plate went up in the air I saw it wobble, and I noticed the red medallion of Cornell on the plate going around. It was pretty obvious to me that the medallion went around faster than the wobbling.



I had nothing to do, so I start to figure out the motion of the rotating plate. I discover that when the angle is very slight, the medallion rotates twice as fast as the wobble rate - two to one. It came out of a complicated equation! Then I thought, ``Is there some way I can see in a more fundamental way, by looking at the forces or the dynamics, why it's two to one?''


I don't remember how I did it, but I ultimately worked out what the motion of the mass particles is, and how all the accelerations balance to make it come out two to one.



Anyone knows how to derive the two-to-one relationship?



Answer



While I cannot possibly claim to know what Feynman did, here is a fairly-simple way of deriving the 2:1 ratio (which incidentally Feynman remembered backwards in that quote; the wobble is faster than the spin).


The viewpoint I'll take is that it's a given that the spin rate and wobble rate are constant, then ask why they are related. We'll work with a thin plate and a small wobble angle, $\theta$.


If the wobble is around a vertical axis, the spin axis (symmetry axis) of the plate points mostly vertically, but off to the side by an angle $\theta$. If the spin is much faster than the wobble, the direction of the angular momentum is almost exactly aligned with the spin axis. As time goes on, the wobble carries the spin axis around and so the angular momentum precesses. This is impossible because angular momentum must be conserved.


Evidently, the wobble must be just fast enough that the angular momentum of the wobbling adds to the spin angular momentum so the total angular momentum of the entire disk is vertical; this way we get no precession as the disk wobbles.


By symmetry, the plate has a principal axis along its spin axis (moment of inertia $I_s$), and the other two principal axes are degenerate, in the plane of the disk (moment of inertia $I_p$). Let's write the angular velocity of the plate as $\vec{\omega} = \vec{\omega}_s + \vec{\omega}_p$, denoting angular velocity components along the spin axis and in the plate. The wobble angular velocity will be denoted $\omega_w$.



We would like to ensure that the horizontal components of total angular momentum are zero. The spin part of the angular velocity gives horizontal angular momentum $L_{s,h} = \omega_s I_s \sin\theta \approx \omega_s I_s \theta$


There is also a horizontal component from the angular momentum in the plane of the disk, which is $L_{p,h} = \omega_p I_p \cos\theta \approx \omega_p I_p$


These must have equal magnitude to sum to zero, so $\frac{\omega_s \theta}{\omega_p} = \frac{I_p}{I_s}$


Finally, we need to relate the angular velocity component in the plane to the angular velocity of the wobble. One way to do this is to note that for a radius r, the left hand edge of the disk (ignoring the spinning) wobbles up and down a distance $r \theta$ over the course of a wobble, and so has a maximum speed of $r\theta \omega_w$. The maximum speed induced by the angular velocity in the plane is $r \omega_p$, so $\omega_p = \theta \omega_w$.


We find


$\frac{\omega_s}{\omega_w} = \frac{I_p}{I_s}$


A quick integral gives that this ratio is the factor of two that Feynman found.


As this is a famous anecdote, there are many solutions based on finding the equations of motion or otherwise analyzing the dynamics already online. I thought it would be nice to add a simple view of the factor of two based solely on a conservation law.


I originally wrote this answer on Quora.


general relativity - Does gravity slow the speed that light travels?


Does gravity slow the speed that light travels? Can we actual measure the time it takes light from the sun to reach us? Is that light delayed as it climbs out of the sun's gravity well?



Answer



This is one of those questions that is more subtle than it seems. In GR the velocity of light is only locally equal to $c$, and we (approximately) Schwarzschild observers do see the speed of light change as light moves to or away from a black hole (or any gravity well). Famously, the speed that radially moving light travels falls to zero at the event horizon. So the answer to your first question is that yes gravity does slow the light reaching us from the Sun.


To be more precise about this, we can measure the Schwarzschild radius $r$ by measuring the circumference of a circular orbit round the Sun and dividing by 2$\pi$. We can also measure the circumference of the Sun and calculate its radius, and from these values calculate the distance from our position to the Sun's surface. If we do this we'll find the average speed of light over this distance is less than $c$.


However suppose we measured the distance to the Sun's surface with a (long) tape measure. We'd get a value bigger than the one calculated in the paragraph above, and if we use this distance to calculate the speed of the light from the Sun we'd get an average speed of $c$.


So I suppose the only accurate answer to your question is: it depends.


Re your other question, assuming the spacetime around the Sun is described by the Schwarzschild metric, the time dilation at the surface of the Sun is given by:



$$ \text{time dilation factor} = \frac{1}{\sqrt{1 - r_s/r}} $$


where $r_s$ is the radius of a black hole with the mass of the Sun and $r$ is the radius of the Sun. The former is about 3,000m and the latter about 700,000,000m so I calculate the time dilation factor to be around 1.000002 and this is too small to measure directly.


However you can interpret gravitational lensing to be due to changes in the speed of light, and since we can measure the gravitational lensing due to the Sun you can argue we have measured its effect on the speed of light. This isn't really true as what gravitational lensing measures is the spacetime curvature. However the change in the speed of light (measured by a Schwarzschild observer) is an aspect of this.


Wednesday, January 27, 2016

experimental physics - On the right-angled fork track of alpha particles


In high school we were told that the idea "alpha particles are actually helium nuclei" came from observing the right-angled fork track, produced by placing an alpha source in a diffusion cloud chamber filled with helium.


From the point of view of the Geiger-Marsden Scattering Experiment, an alpha particle hitting a helium nucleus is an extremely rare event. Besides, there is no guarantee at all that the alpha particles would not hit the nuclei of other gas molecules in the chamber.


So my question is: How did people discover this extremely rare event (sit there and stare at the chamber for hours?), and why were they so sure that it was a helium nucleus that the alpha particle hit?




Tuesday, January 26, 2016

particle physics - Why is the laboratory frame energy always greater than the center of mass frame energy?


I have been looking for an answer to 'Why is the laboratory frame energy always greater than the center of mass frame energy during collisions?'.


A lot of resources provided mathematical explanations. I wanted to get some perspectives in terms of physics.


Could you explain this?




voltage - Are square wave harmonics real-life phenomena or just mathematical abstractions?



Based on my limited knowledge, it is my understanding that square waves can be mathematically broken down into an infinite sum of sinusoidal waves (of different amplitudes, frequency, etc) . This is very interesting, and mathematically it makes sense, however, this concept starts breaking down for me when I start hearing about how all square waves (no matter how they are generated) abide by that rule.


For example, sometimes, I stumble upon articles about electronics where the author mentions how a square voltage wave being impressed into a circuit will be felt by the circuit as a bunch of sinusoidal waves of multiple amplitudes, frequencies, etc. Really? How can this be? I understand that if a square wave is built using sinusoidal waves then it makes sense that the circuit will feel all the sinusoidal waves that the square wave is made of, however, some square waves are made from just “on” / “off” transitions (nowhere do we inject sinusoidal waves to that square wave) so I don’t see how those particular square waves can be felt by the circuit as any type of sinusoidal waves.


So what is going on here? Could someone please help me understand this a little better?




fluid dynamics - Wavy stream of liquid


While pouring a liquid into a glass from a bottle, some streams have a wavy shape, like the one in the following photo:


example


What causes the stream to be of such a shape?



Answer




On the one hand, as the liquid is not being poured very slowly, there are different velocities in the incoming liquid: it's faster at the top, and, it seems, at the back. On the other hand, the cohesive forces (resulting in particular in surface tension at the surface of the flow) are sufficient here to prevent separation of some liquid from the rest — splattering. Then it is as if you had two masses going at different speeds (the liquid at the back and at the front) linked by some piece of rubber (cohesive forces) : they turn around one another.


The smooth appearance due to surface tension gives the final touch for a beautiful flow. You'll probably enjoy taking a tour in John Bush's gallery if you like this, http://math.mit.edu/~bush/?page_id=252. A very relevant example is the case of colliding viscous jets, pdf and below, which exemplifies beautifully the present discussion at low flow rate (left) but gives much more impressive results at higher flow rates! (right)


http://i.imgur.com/HCHZWo2.jpg


thermal radiation - Rayleigh-Jeans Law


My question is simple, why do we believe Rayleigh-Jeans law to be absurd? Is the Ultra-violet catastrophe incorrect or is it only because we can not create or know of a perfect emitter? I am a bit confused as to why Quanta was necessary.




Monday, January 25, 2016

What exactly does No cloning mean, in the context of Quantum Computing?


I am trying to get an intuitive idea of how the No-Cloning theorem affects Quantum computation. My understanding is that given a qubit $Q$ in superposition $Q_0 \left| 0 \right> + Q_1 \left| 1 \right>$, NCT states another Qubit $S$ cannot be designed such that $S$ is equivalent to the state of $Q$.


Now the catch is, what does Equivalent mean? It could mean either that:





  1. $S = S_0 \left| 0 \right> + S_1 \left| 1 \right>$ such that $S_0 = Q_0, S_1 = Q_1$.




  2. Or it could mean that $S = Q$, meaning that if $S$ is observed to be some value ( for example 0) then $Q$ MUST be that same value, and vice versa.




So it seems that point 2, occurs anyways in entangled systems (particularly cat-states), so I can eliminate that option and conclude that that No Cloning states, given a qubit $Q$, it's impossible to make another qubit $S$ such that:


$S = S_0 \left| 0 \right> + S_1 \left| 1 \right>$ such that $S_0 = Q_0, S_1 = Q_1$.


Is this correct?



Answer




You need to use a more precise notion of the cloning process, in order to understand the general statement and its repercussions. I will give you some outline here (mainly following the explanations of B. Schumacher and M. Westmoreland given in the reference), with an emphasis on the most important aspects of it, but to fully appreciate the importance of the No-cloning Thm I highly recommend looking through the various ways you can prove it (I can show you some ways of proving it in this post, if you will see it necessary).


Main statement:



  • No-cloning theorem states that no unitary cloning machine exists that would clone arbitrary initial states.

  • A softer version would be: Quantum information cannot be copied exactly.


Repercussions if arbitrary cloning was possible: (not following any specific order)



  • If a hypothetical device exists that could duplicate the state of a quantum system, then an eavesdropper would be able to break the security of the $BB84$ key distribution protocol.



  • A cloning machine would allow to create multi-copy states $|x\rangle^{\otimes n}$ from a single state $|x\rangle.$ But take another single state $|y\rangle,$ create its corresponding multi-state $|y\rangle^{\otimes n},$ and you can overcome the basic distinguishability limitations of states in Quantum Mechanics, as multi-copy states can be better distinguished (correct term would be more reliably) than single states.


    Recall that the distinguishability of two states $x,y$ is given by their amount of overlap, i.e. $|\langle x | y\rangle|,$ the closer this is to vanishing, the better we can distinguish between the states. (If you're not familiar with the concept of multi-states being more reliably distinguishable, let me know).




  • The no-cloning theorem guarantees the no-communication theorem and thus prevents faster than-light communication using entangled states. (the no-communication theorem basically says that: if two parties have systems $A$ and $B$ respectively, and suppose their joint state is entangled $|\psi^{AB}\rangle,$ then: the two parties cannot transfer information to each other either by: choosing different measurements for their respective systems, or evolving their systems using different unitary time evolution operators.)




More precise definition of the cloning problem:


There are three elements involved, the initial state (input) to be copied $(1)$, a blank state onto which we want to create the copy $(2)$ and a machine that plays the role of the cloning device $(3)$. The composite system is then $(123).$


Suppose the state of $(2)$ is $|0\rangle,$ state of $(1)$ being $|\Phi\rangle$ and the starting state of $(3)$ is $|D_i\rangle.$ Let us denote the action of the cloning device by the unitary operator $U.$ It is important to point out that the starting state of the composite system $(23)$ and the action $U$ is independent of the state to be copied, i.e. system $(1).$ Our starting composite state is then:



$$ |123\rangle_i = |\Phi\rangle \otimes |0\rangle \otimes |D_i\rangle $$ By applying $U$, thus after cloning, the state of $(1)$ is unchanged, but upon success of the cloning, the state of $(2)$ must be exactly that of $(1).$ So


$$ U|123\rangle_i = |\Phi\rangle \otimes |\Phi\rangle \otimes |D_f\rangle $$


Given this description, the no-cloning theorem says that such $U$ does not exist for arbitrary states of $(1).$


Hints on the proofs:



  • One way would be to use the principle of superposition to show that such cloning is not possible, by showing that if the device is to work for two orthogonal states, it would create entangled outputs for their superposition. (thus the subsystems are no longer even in pure states)

  • Another way would be going back to the concept of distinguishability between non-orthogonal states, and using the fact that unitary time evolution preserve inner products, thus showing that the cloning device is impossible as it would allow considerable improvement on the distinguishability.




Reference: A highly recommended reference, also for further reading on all this matter, would be the book of Quantum Processes, systems & Information by Benjamin Schumacher and Michael Westmoreland. (Of relevance here, are chapters 4 and 7.)



particle physics - What would happen if an incredibly high energy photon passed through a human body?


What would happen if an incredibly high energy photon passed through a human body (or any other material)? When I say incredibly high energy photon, I mean higher than any known energy scales... higher than electronic energy scales, higher than nuclear transition energies, etc...


In my mind, I pictured the photon tearing apart whatever it passed through. However, browsing Wikipedia's article on the Photoelectric Effect, I noticed that the cross section went as $1/E^p$, where I think $E$ is energy and $p$ is some number. So, the probability of an atom being ionized by a high energy photon actually drops with the photon energy (although I don't know what approximations went into this calculation... first order perturbation theory?). If the other light-matter interaction processes also have cross sections of the form $1/E^p$, then an incredibly high energy photon would pass through us without harming us!


Or maybe this question is unanswerable because we don't yet understand physics at such high energy scales. What are your thoughts on this? What do you think would happen if an incredibly high energy photon passed through a human body?




cosmology - What's wrong with the Big Spin Model?


A Dr. Serkan Zorba has a paper on arxiv in which he considers, what if the universe is actually slowly rotating? This gives rise to centrifugal and Coriolis forces on a galactic scale that seem to perfectly explain the effects of dark energy and dark matter. Not only that, but it explains a few other observations in cosmology.


But I haven't been able to find anyone else commenting on his model, and it's been several years now. Is there something wrong with it that allows us to dismiss it out of hand? Or do other cosmologists just not care?




Sunday, January 24, 2016

general relativity - Lagrangian for Relativistic Dust derivation questions


In most GR textbooks, one derives the stress energy tensor for relativistic dust: $$ T_{\mu\nu} = \rho v_\mu v_\nu $$ And then one puts this on the right hand side of the Einstein's equations. I would like to derive this from some action. I have read through all standard GR textbooks, and the only one which talks about this is General Theory of Relativity by P. Dirac. If you don't have the book, let me quickly reproduce the derivation here. The action for the relativistic dust is: $$ S_M = -\int \rho c \sqrt{v_\mu v^\mu} \sqrt{ |\det g| } d^4 x = -\int c \sqrt{p_\mu p^\mu} d^4 x $$ where $p^\mu=\rho v^\mu \sqrt{ |\det g| }$ is the 4-momentum density. Now we vary with respect to $g^{\mu\nu}$ as follows: $$ \delta S_M = - \delta \int c \sqrt{{p}_\mu {p}^\mu} d^4 x = $$ $$ = - \int c {\delta(g^{\mu\nu} {p}_\mu {p}_\nu) \over 2\sqrt{{p}_\alpha {p}^\alpha}} d^4 x = $$ $$ = - \int c { {p}_\mu {p}_\nu \over 2\sqrt{{p}_\alpha {p}^\alpha}} \delta(g^{\mu\nu}) d^4 x = $$ $$ = - \int c { \rho v_\mu \rho v_\nu \sqrt{ |\det g| }^2 \over 2 \rho c \sqrt{ |\det g| } } \delta(g^{\mu\nu}) d^4 x = $$ $$ = - \int {1\over2} \rho v_\mu v_\nu \delta(g^{\mu\nu}) \sqrt{ |\det g| } d^4 x $$ From which we calculate the stress energy tensor using the standard GR formula for it: $$ T_{\mu\nu} = - {2\over\sqrt{ |\det g| }}{\delta S_M\over\delta g^{\mu\nu}} = $$ $$ = - {2\over\sqrt{ |\det g| }} \left( -{1\over2} \rho v_\mu v_\nu \sqrt{ |\det g| } \right)= $$ $$ = \rho v_\mu v_\nu $$


If we vary with respect to $x^\mu$, we obtain the geodesic equation (the calculation is lengthy, see for example here, or the Dirac book). I'll be happy to clarify any of the derivations above if needed. Now my questions:


1) Why isn't this in every GR textbook? Is there some problem with the derivation?


2) The geodesic equation follows from this action $S_M$. Standard way to derive the geodesic equation is to maximize the proper time $$\tau = \int d \tau = \int \sqrt{-{1\over c^2} d s^2} = \int \sqrt{-{1\over c^2} g_{\mu\nu} d x^\mu d x^\nu} .$$ Is there some relation between this $\tau$ and $S_M$ since both give us the same geodesic equation?


3) Is it correct to simply say, that all GR for relativistic dust (without electromagnetism) follows from this action: $$ S = {c^4\over 16\pi G} \int R \sqrt{ |\det g_{\mu\nu}| } d^4 x -\int c \sqrt{p_\mu p^\mu} d^4 x $$ when varied with respect to $g^{\mu\nu}$ it gives the Einstein's equations with the $T_{\mu\nu} = \rho v_\mu v_\nu$ tensor on the right hand side, when varied with respect to $x^\mu$ it gives us the geodesic equation (for each particle of the dust).



4) Why do we need to hide the $\sqrt{ |\det g_{\mu\nu}| }$ in the 4-momentum density and do not vary it? Dirac says that it is because $\rho$ and $v^\mu$ are not independent quantities when varying, but I don't understand the argument.


5) The standard way is to use the Hilbert action, the action for the elmag field, the stress energy tensor for dust and derive the geodesic equation as a conservation of the stress energy tensor, that follows from the conservation of the Einstein's tensor. How is this approach related to the above? Isn't it physically better to simply postulate the total action and derive everything from it?


Note: Dirac shows how to incorporate electromagnetism by simply using the standard action for it and the procedure above then gives the correct elmag. tensor on the right hand side of Einstein's equations and the Lorentz force at the right hand side of the geodesic equation, as well as Maxwell's equations for the elmag field.



Answer



The dust stress-energy is a sum of independent particle stress energies. Each particle is going along with a 4-velocity $v^\mu$, and this means it is carrying a momentum $mv^\mu$ in the direction $v^\mu$, or an implusive stress tensor (momentum flow) of $T^{\mu\nu}=mv^\mu v^\nu$ at the location of the particle (this needs to be multiplied by a delta function at the location of the particle). Adding this stress tensor contribution up over a continuous mass distribution of particles reproduces the stress tensor of a dust, and you can figure it out by just thinking about how much momentum the particles of the dust carry across an infinitesimal surface in each unit of time. So this is a very simple thing physically.


Question 1: why is this not in books other than Dirac's?


Generally people do a derivation similar to this in all GR books, but they do it particle by particle, instead of summing over all the particles in the dust. The particle action is the length of the worldline, setting the variation to zero gives the equation of motion (the calculation is completely parallel to Dirac's, since each particle in the dust has an action which is independently equal to its worldline length), and looking at the metric derivative gives the stress tensor source term for Einstein equations. So it's really the same as a single particle.


The likely reason Dirac is not comfortable with using a single particle action is because GR is not compatible with point sources. A point source of gravity is inconsistent, it's inside its own Schwartschild horizon. One modern point of view is to view the point source as an infinitesimal black hole, but then you need to establish that the action of a black hole can be taken to be the length of its approximate worldline, which is a complicated vacuum equation calculation. The issue is sidestepped if you have a smeared out stress-energy fluid like in a dust, so Dirac probably felt more comfortable deriving it this way, although it introduces some minor extra complications.


Question 2: Relation between the dust action and the arclength action


The arclength action for a point particle is



$$S(X) = \int d\tau |\dot{X}(\tau)| = \int d^4x \int d\tau \;\;\delta^4(x-X(\tau)) \sqrt{\dot{X}^\mu(\tau)\dot{X}^\nu(\tau)g_{\mu\nu}(X(\tau))} $$


Where the second expression is just a formal integration over space which collapses onto the particle trajectory, to describe where this action is located.


Label each particle of the dust by its initial position $\sigma$, which varies along some initial data surface. For each $\sigma$, consider the integral lines of the vector field $V^\mu$, and call this $x^\mu(\sigma,\tau)$, the dust trajectories. You can write the dust action as the sum of the individual particle actions


$$ S = \int d^4x \int d^3\sigma \rho(\sigma) S(x^\mu(a,\tau) $$


Where S(a) is the arclength action for each particle.


note that the physical space-time density $\rho(x)$ is related to the initial hypersurface density $\rho(\sigma)$ by


$$ \rho(x) = \int d\sigma d\tau \rho(\sigma) \delta^4(x-X(\sigma,\tau))$$


If you get rid of the $(\tau,\sigma)$ integrals in the action using the delta functions, you recover Dirac's action. When you vary the individual dust trajectories, you get the geodesic equation. When you vary $g_{\mu\nu}$ you get the individual particle stresses contributing to the total stress energy. This type of thing appears in all modern GR books.


The number of particles $\rho(a)$ must not be varied in the Lagrangian formulation--- the total number of particles in the dust must be conserved. So you can vary the trajectories, but not the number of particles.


Question 3: Is Dirac's action complete for dust/gravity?



It gives the dust equation of motion and the Einstein equations sourced with the right dust stress-energy, so yes.


Question 4: What's up with varying $g_{\mu\nu}$ holding $P^\mu$ fixed?


This is a little subtler--- when you vary $g_{\mu\nu}$ you vary the volume of space time as well as the local metric, but you have to do the variation holding the number of particles fixed. If you vary $g_{\mu\nu}$ at constant $\rho$ you get the wrong variation--- you will reduce the number of particles as you decrease the total volume of spacetime. This means that $\rho$ needs to transform as a density when you vary $g$.


To avoid dealing with density, Dirac uses the fact that the $P^{\mu}$ is a conserved quantity locally, in the absence of gravity, so it transforms as a simple vector. The density variation of $\rho$ is absorbed by the factor of $\sqrt{g}$ included in the definition $P$. So there is no extra volume variation for $P^{\mu}$. You can do it without Dirac's simplification, and you get the same answer, but its a little more algebra and less conceptually illuminating.


Question 5: What's the best way to derive everything?


I don't think that the action principle way is necessarily better, because the macroscopic equation of the sort that describe a dust do not have to follow from a Lagrangian. They can have dissipation. So if you couple a dissipating Navier Stokes fluid to gravity, you still expect to have a good equation of motion with viscosity, but you don't expect a Lagrangian description when the viscosity is nonzero.


The stresses due to viscosity aren't Lagrangian, but they still gravitate, and there is no reason to leave out certain stresses because they are not fundamental. In Dirac's time, there was a push to try to make the classical theories as complete as possible, because they were guides to a more complete theory. So Dirac might have wanted to consider a model where the electron is a continuum dust model, for example, and then it is important to have a Lagrangian. The modern theories of quantum gravity have advanced to the point where I think it is not necessary to be as pedantic, and mildly inconsistent things like the point particle are fine in intermediate steps. So there is no reason to prefer a Lagrangian approach for this sort of thing, although it is nice when it exists.


electromagnetism - Is it safe to use any wireless device during a lightning storm?


I need "educated" reasons whether it is safe to use any wireless device during a lightning storm.


Most people said don't use it but they cannot explain why.




quantum mechanics - Finding classical action in tunneling problem


In QM: I am trying to show that the minimum action for a classical path going between two potential wells (centered at $\pm L$) in a dbl-well potential is $$S_{classical} = \int_{-L}^{L} dx' \sqrt{2mV(x')}\tag{1}$$ where $V(x')$ is the inverted potential (in imaginary time $t' \rightarrow -it$).



I know that $$S = \int_{0}^{t''} dt'\, (1/2 m \dot{x'}^2 + V(x'))\tag{2}$$ which is the action in imaginary time. Furthermore $$x'(0) = - L\qquad\text{and}\qquad x'(t'') = L,\tag{3}$$ where $x'$ is the corresponding 'transformed position'.


I am told to use energy conservation and the fact that the minimum energy is zero, but I am still unsure how to tackle this problem. I am unexperiened with the least-action principle, so this might be why I am struggling.



Answer



Good question. What you see here is the procedure known as the WKB approximation. Let's start from scratch and proceed slowly. Consider the 1D path integral given by $$Z = \int_{-\infty}^{\infty}\exp(i\mathcal{S}/\hbar)\ \text{dx} = \int_{-\infty}^\infty\exp\left(\frac{i}{\hbar}\int_0^{t_0}\mathcal{L}(x,\dot{x})\text{dt}\right)\text{dx}.$$ Let's look at the Lagrangian closely. Suppose that the Lagrangian has the form $$\mathcal{L}(x,\dot{x}) = \frac{1}{2}m\dot{x}^2-V(x)$$ where the potential looks like two wells stitched together. For visual sake, suppose the potential looks like $$V(x)=\frac{1}{2}(x-L)^2(x+L)^2$$ so that when $x=L$ or $x=-L$ the potential is $V(x)=0$. Graphically, it could look like the following image posted by another physics stack exchange post.


enter image description here


Anyway, because of energy conservation, we know that $E=0$ but don't forget that $E=\mathcal{H}$, the Hamiltonian, and since $\mathcal{H}$ and $\mathcal{L}$ are related by $$\mathcal{H}=\dot{x}p-L,$$ we can solve the above equation using $\mathcal{H}=0$, again we can do this because $E=0$. The above equation actually comes from the Legendre transform of $\mathcal{L}$ and you can find out more about it here. Anyway, returning to the problem we could solve the above equation to find \begin{align} L &= \dot{x}p\iff\\ \frac{1}{2}m\dot{x}^2-V(x)&=\dot{x}p \iff \\ -V(x) &= \dot{x}p - \frac{1}{2}m\dot{x}^2 \iff\\ -V(x) &= \dot{x}p - \frac{1}{2}\dot{x}(m\dot{x})\iff\\ -V(x) &= \dot{x}p - \frac{1}{2}\dot{x}p = \frac{1}{2}\dot{x}p\iff\\ -V(x) &= \frac{p^2}{2m} \implies \\ &\boxed{p = i\sqrt{2mV(x)}} \end{align} which is imaginary. Now let's go back to the path integral and plug things in to see what happens. \begin{align}Z &= \int_{-\infty}^{\infty}\exp(i\mathcal{S}/\hbar)\ \text{dx} = \int_{-\infty}^\infty\exp\left(\frac{i}{\hbar}\int_0^{t_0}\mathcal{L}(x,\dot{x})\text{dt}\right)\text{dx}\\ &= \int_{-\infty}^\infty\exp\left(\frac{i}{\hbar}\int_0^{t_0}\left[\dot{x}p-\mathcal{H}\right]\text{dt}\right)\text{dx}=\int_{-\infty}^\infty\exp\left(\frac{i}{\hbar}\int_0^{t_0}\left[\dot{x}p-0\right]\text{dt}\right)\text{dx}\\ &=\int_{-\infty}^\infty\exp\left(\frac{i}{\hbar}\int_0^{t_0}\left[\dot{x}i\sqrt{2mV(x)}\right]\text{dt}\right)\text{dx}\\ &=\int_{-\infty}^\infty\exp\left(\frac{-1}{\hbar}\int_0^{t_0}\sqrt{2mV(x)}\frac{\text{d}x(\text{t})}{\text{dt}}\text{dt}\right)\text{dx} \end{align} Now from calculus, we know that the differential of $x$ is given by $$\text{d}x(\text{t})=\frac{\partial{x(\text{t})}}{{\partial\text{t}}}\text{dt}=\frac{\text{d}x(\text{t})}{\text{dt}}\text{dt}$$ and then letting $x(t=0)=-L$ and $x(t=t_0)=+L$ we have $$Z = \int_{-\infty}^\infty\exp\left(\frac{-1}{\hbar}\int_{-L}^{+L}\sqrt{2mV(x)}\text{dx}\right)\text{d}x = \int_{-\infty}^{\infty}\exp\left(\frac{-1}{\hbar}\mathcal{S}_{classical}\right)\text{dx}$$ Boom. The classical action pops right out.


homework and exercises - Nonlinear spring $F=-kx^3$


A nonlinear spring whose restoring force is given by $F=-kx^3$ where $x$ is the displacement from equilibrium , is stretched a distance $A$. Attached to its end is a mass $m$. Calculate....(I can do that) ..suppose the amplitude of oscillation is increased, what happens to the period?


Here's what I think: If the amplitude is increased the spring posses more total energy, at equilibrium the spring is traveling faster than before because it posses more kinetic energy. I think in the spring travels faster when it's at a similar displacement from equilibrium, but it has to travel more distance, so I can't conclude anything.


I was think about solving,


$$mx''=-kx^3$$


But realized this is a very hard job.


Any ideas? Thanks.



Answer



The potential energy is $U\left(x\right) = kx^4/4$ since $-d/dx\left(kx^4/4\right) = -kx^3 = F$, and the energy $$ E = \frac{1}{2}m\left(\frac{dx}{dt}\right)^2 + \frac{1}{4}kx^4 $$ is conserved.



From the above you can show that $$ \begin{eqnarray} dt &=& \pm \ dx \sqrt{\frac{m}{2E}}\left(1-\frac{k}{4E}x^4\right)^{-1/2} \\ &=& \pm \ dx \sqrt{\frac{2m}{k}} \ A^{-2} \left[1-\left(\frac{x}{A}\right)^4\right]^{-1/2} \end{eqnarray} $$ where the amplitude $A = \left(4E / k\right)^{1/4}$ can be found from setting $dx/dt = 0$ in the expression for the energy and solving for $x$.


The period is then $$ \begin{eqnarray} T &=& 4 \sqrt{\frac{2m}{k}} \ A^{-2} \int_0^A dx \left[1-\left(\frac{x}{A}\right)^4\right]^{-1/2} \\ &=& 4 \sqrt{\frac{2m}{k}} \ A^{-1} \int_0^1 du \left(1-u^4\right)^{-1/2} \\ &=& \left(4 \sqrt{\frac{2m}{k}} I\right) A^{-1} \\ &\propto& A^{-1} \end{eqnarray} $$ where $u = x/A$ and $I = \int_0^1 du \left(1-u^4\right)^{-1/2} \approx 1.31$ (see this).


You can repeat the above for a more general potential energy $U\left(x\right) = \alpha \left|x\right|^n$, where you should find that


$$ dt = \pm \ dx \sqrt{\frac{m}{2\alpha}} \ A^{-n/2} \left[1-\left(\frac{\left|x\right|}{A}\right)^n\right]^{-1/2} $$


and


$$ \begin{eqnarray} T_n &=& \left(4 \sqrt{\frac{m}{2\alpha}} I_n\right) A^{1-n/2} \\ &\propto& A^{1-n/2} \end{eqnarray} $$


where


$$ I_n = \int_0^1 du \left(1-u^n\right)^{-1/2} $$


can be evaluated in terms of gamma functions (see this).


This is in agreement with the above for $\alpha = k/4$ and $n=4$, and with Landau and Lifshitz's Mechanics problem 2a of section 12 (page 27), where they find that $T_n \propto E^{1/n-1/2} \propto A^{1-n/2}$.



Saturday, January 23, 2016

What is the largest sphere of liquid water that could exist in space held together under it's own gravity?



What is the largest theoretical sphere of liquid water that could exist in space held together under it's own gravity? I've always wondered if a planet the size of earth could exist as a single volume of water? If in orbit of a star (which I guess it would need to be to remain a liquid), would the star's UV radiation evaporate such a watery world? Do the laws of physics permit enormous volumes of liquid water to float around in space?



Answer



Planets made mostly of water almost certainly exist, and at least one may have been detected.


However, such bodies will not be made entirely of liquid. In fact they will be mostly solid, even if the temperature is very high. This is because water can form more than one form of ice. The kind we're familiar with forms at low temperatures and is less dense than water. But there are several other kinds that form at very high pressures instead. These "exotic" ices are heavier than liquid water, so will sink to the bottom. You can get a feel for the pressures involved by looking at a phase diagram of water, e.g. here or here.


Earth's oceans aren't deep enough for such pressures to exist, but if the whole planet were made of water, the planet would be composed mostly of "exotic" ices, covered by a relatively thin layer of liquid (though this would still be a lot deeper than Earth's oceans). In fact, Europa, Titan and many of the moons of the outer Solar system are probably more or less like this, except that they have a layer of normal water ice on top of the liquid ocean, because they're so cold.


Another answer mentions the possibility that there would be a smooth transition from water vapour to liquid water, rather than a clear ocean surface as there is on Earth. Whether this is true or not depends on the temperature. The critical point for water is at about $647\,\mathrm{K}$, or $374^\circ\mathrm{C}$. If the temperature is higher than this then there will not be a phase transition between the atmosphere and the liquid, but if it's lower then there will, just as there is on Earth.



With those details out of the way we can address the question of stability. Your original question was what's the largest body of liquid water that can hold itself together under its own gravity. However, I don't think such a body can exist at all. If it did, it would have to be rather small in order to prevent the formation of exotic ices in its interior. But that would mean its gravity would be quite low, so it wouldn't have an atmosphere. The low pressure at the surface would cause the water to evaporate or boil, and the molecules of the resulting gas would easily be moving fast enough to escape the body's gravitational pull. (Of course, things are different if we allow the surface to be frozen - see the outer Solar system moons - but I'm assuming you're specifically interested in a body with a liquid surface.)


However, if we allow the interior to be composed of exotic ices rather than liquid, it gets a lot easier. Then all we need is for it to be big enough that it's escape velocity is higher than the velocity of water vapour molecules. Anything from around Earth-sized upwards should do the trick. I don't think there's an upper limit to the size, except that if it's really, really big then nuclei in its core will start to fuse and it will turn into a star.


You mention UV radiation causing a loss of the water, and this is an issue for a body the size of Earth. UV radiation can split water into hydrogen and oxygen, and the hydrogen atoms have a much higher velocity, making it easier for them to escape into space. On Earth we're fortunate to have an oxidising atmosphere, which tends to turn the hydrogen atoms back into molecules before they can escape. If it wasn't for that, the Earth's oceans would already have disappeared.


However, if we start with a pure $\mathrm{H_2O}$ planet that's big enough, eventually it too will have an oxygen atmosphere. Not because of photosynthesis, but simply because as the hydrogen escapes it leaves oxygen behind. Given sufficient time I would expect this to lead to a protective oxygen atmosphere that prevents further escape of hydrogen.


The only detail left is how to get a surface temperature that's in the right range for the surface to be liquid. This depends on the distance from the star, but also on the composition of the atmosphere. On Earth, the atmosphere's composition is mostly due to the action of the biosphere, which keeps the temperature regulated in just the right range for water to be liquid. Perhaps it's possible to imagine life on such a water world, in the form of photosynthesising algae-like organisms, which might play a similar role.


How strong is electron degeneracy pressure?


I'm trying to get some specific numbers for electron degeneracy that I can understand, using a concrete example.


Take for example this portion of carbon crystal:


carbon crystal lattice




  1. Exactly how much energy would be required to overcome the electron degeneracy pressure of these 18 carbon atoms, in a theoretical laboratory capable of such things.

  2. I believe that neutrinos would be emitted as a byproduct of the collapse, is that correct? What other reactions if any would be needed to fully account for conservation of energy/momentum etc?




quantum mechanics - What does a Field Theory mean?


What exactly is a field theory?


How do we classify theories as field theories and non field theories?


EDIT:



  1. After reading the answers I am under the impression that almost every theory is a field theory. Is this correct?

  2. What is difference between QM and QFT?



Answer



Update to address new questions.





  1. The answer to this question is no. At least if you take the question purely formally. Only theories such as classical field theory, quantum field theory and continuum mechanics are field theories (you generally recognize them by having continuous degrees of freedom; also they usually have the word field in the title :-)). But physically, lots of different theories may be equivalent, or may be approximations of some other theory, so there are many connections among them (this is the point I was trying to illustrate, but maybe I overemphasized it).




  2. Difference between QM and QFT is essentially the same as between classical mechanics and classical field theory. In the mechanics you have just a few particles (or more generally, small number of degrees of freedom), while fields have an infinite number of degrees of freedom. Naturally, field theories are a lot harder than the corresponding mechanics. But there is a connection I already mentioned: you can see what happens when you let the number of particles grow arbitrarily large. This system will then essentially behave as a field theory. So in a sense, we can say that field theory is a large $N$ (number of degrees of freedom) limit of the corresponding mechanical theory. Of course, this view is very simplified, but I don't want to get too technical here.






Field theory is a theory that studies fields. Now what is a field? I suppose everyone should be familiar with at least some of them, e.g. gravitational or electromagnetic (EM) field.



Now, how do you recognize that object is a field? Well, essentially, you look at how complicated the object is. To make this more precise: main objects of study of classical mechanics are point particles. All you need to keep track of them is just few parameters (position, velocity). On the other hand, consider the EM field: you need to keep track of the data (electric and magnetic field vector) in every point of the universe, so there is infinitely many parameters of this system! This is what I meant by system being large: you need a lot of data to describe it.


Now, it might seem that something is amiss. You do need a lot of data to describe real objects (just think of how many atoms there are in the grain of sand). So are ordinary objects fields? Yes and no, both answers are correct depending on your point of view. If you consider a massive object as essentially being described by few parameters (like center of mass velocity and moment of inertia) and completely ignore all information about atoms then it's clearly not a field. Nevertheless, at the microscopic level, atoms wiggle around and even the grain of sand really is as complicated object as any EM field (not to mention that atoms themselves produce EM field), so it's certainly correct to call them that.


Now let us see where our definition of field takes us. Let's talk about quantum mechanics for a while. What about two quantum particles? Is it a field? Well, clearly not. What about three? Still not. And what if we keep adding particles so that there will be a huge number of them? Well, it turns out that we'll get a quantum field! This is precisely the correspondence between e.g. photons and quantum EM field. You can either look at EM field as being described by vector of electric and magnetic field at every point as in the classical case, or you can instead reorganize your data so that you keep track of what kind of photons you have. It's useful to carry both pictures in head and use the more appropriate one.


There is also a subject of continuum mechanics. There you can also start with particles (describing atoms in some real object, e.g. water) and because there are so many of them, you can again reorganize your data, consider the object as being essentially continuous (which real objects surely are at least unless you look at them with a microscope), and instead describe them by parameters such as pressure and temperature at every point.


To summarize: the field theory is essentially about dealing with large objects. However, when we are looking at the problem with particle hat on, we usually don't say it's a field. For instance, when describing real objects as consisting of atoms, we are usually talking about statistical mechanics, or condensed matter physics. Only when we move to the realm of continuum mechanics, we say that there are fields.


There is much more to be said on the topic but this post got already too long so I'll stop here. If you have any questions, ask away!


condensed matter - Why are metals malleable and ductile?


Why are metals malleable and ductile? These two properties seem to be related. Is there a microscopic understanding of these properties possible?



Answer



Let's draw a comparison with ceramics, which—just as metals are generally ductile—are generally brittle.



First, note that crystals (and metals and ceramics are both generally polycrystalline) can deform through dislocation motion. A dislocation is a line defect that carries plasticity through a crystal. The classic analogy is moving a rug by kicking a wrinkle down its length. You don't need to deform the entire crystal at once; you just need to sweep one (or many) dislocations through the material, breaking a relatively small number of bonds at a time.


Here's a simple illustration of a curved dislocation carrying shear through a crystal; the passage of the dislocation leaves a new permanent step:


enter image description here


So this is a very convenient way to achieve permanent deformation. However, it's much easier to break these bonds in metals than in ceramics because the metallic bonds in the former are weaker than the ionic/covalent bonds in the latter (as evidenced by the fact that ceramics are generally refractory, i.e., they have high melting temperatures). In particular, the delocalized nature of the electrons in metals allows dislocation to slip by easily. This equates to ductility/malleability. (The two terms are identical for this discussion; they differ only in the type of loading conditions that result in easy deformation.)


Additionally, in metals with a face-centered-cubic crystalline structure (think gold or copper, for example), the structural symmetry provides many possible slip planes along which dislocations can easily propagate. This equates to even greater ductility/malleability.


Here's an illustration of a face-centered-cubic structure; the close packing of atoms on multiple planes allows dislocations to hop only short distances, greatly easing their passage:


enter image description here


In contrast, dislocation motion is so strongly hindered in ceramics (because the bonds are directional and the charges are rigidly fixed) that it may take less energy to simply break all the bonds at once, corresponding to bulk fracture and brittleness.


One consequence of these microscopic differences between metals and ceramics is the way that they respond to cracks or flaws. A sharp crack produces a stress concentration, essentially because the stress field has to twist sharply around it. In a metal, this stress concentration isn't much of a problem—some dislocations will move, resulting in plastic deformation and blunting of the crack tip. This option is much less likely in a ceramic because of the impediments to dislocation motion. It may just be easier to break the bonds permanently and form a new open surface at the formerly high-stress area. This is the mechanism of crack propagation, and if the crack continues to propagate, you get bulk fracture.


Collapse in Quantum Field Theory?



I do not want answers telling me that wave-function collapse is not real and decoherence is the answer (I know the situation with that). I am asking a question purely on the basis if wave-function collapse is the correct method. My question is: in normal quantum mechanics superposition of the state (position, momentum) exists until the wave-function collapses (how or why it collapses is not important in this question), now in quantum field theory we can also have superposition as in the superposition of Fock space states with different particle number. Can the superposition also collapse here under collapse interpretations of quantum mechanics/quantum field theory?


Laymans answers would be mainly appreciated...




Friday, January 22, 2016

kinematics - When this CD is spun so fast that it shatters, why does the warped shape move slower than the surface?


In this video where a CD is spun abnormally fast to the point of shattering, one notices from the writing on the CD that the surface is spinning faster than the warp "shape" is. In other words, the shape the CD has warp into is moving in a circle, but not quite as fast as the surface of the CD. What's the explanation behind this de-synchronization?



I would expect the warp shape to either not spin much at all, or spin at the same speed as the surface of the CD.




dirac matrices - Are Lifshitz and Berestetskii right in this case?



In the Quantum electrodynamics book (look at the problem) its authors Lifshitz and Berestetskei claim that operator of charge conjugation $\hat {C} = -\alpha_{2}$ in Majorana basis transforms as $\hat {C}^{M} = \alpha_{2}$. Here (they started from the standart (Dirac) representation of the gamma-matrices) $$ \alpha_{2} = \begin{pmatrix} 0 & \sigma_{y}\\ \sigma_{y} & 0\end{pmatrix}, \quad \beta = \begin{pmatrix} 1 & 0\\ 0 & -1\end{pmatrix}, $$ $$ \hat {C}^{M} = \hat {U}^{+} \hat {C} \hat {U}, \quad \hat {U} = \frac{1}{\sqrt{2}}(\alpha_{2} + \beta ) = \frac{1}{\sqrt{2}}\begin{pmatrix} 1 & \sigma_{y}\\ \sigma_{y} & -1\end{pmatrix} = \hat {U}^{+}. $$ I tried to get their result, but I only got $$ \hat {C}^{M} = \frac{1}{2}\begin{pmatrix} 1 & \sigma_{y}\\ \sigma_{y} & -1\end{pmatrix}\begin{pmatrix} 0 & \sigma_{y}\\ \sigma_{y} & 0\end{pmatrix}\begin{pmatrix} 1 & \sigma_{y}\\ \sigma_{y} & -1\end{pmatrix} = \begin{pmatrix} 1 & 0\\ 0 & -1\end{pmatrix}, $$ because $\sigma_{y}^{2} = 1$.


Where is the mistake?


An edit.


I found the mistake. I used the wrong definition of transformation of charge conjugation operator under unitary spinor transformation. The correct one is $\hat {C}^{'} = \hat {U}\hat {C}\hat {U}^{T}$.




Thursday, January 21, 2016

electricity - What happens to capacitor’s charge when the plates are moved further apart?



In my physics textbook there is an example of using capacitor switches in computer keyboard:



Pressing the key pushes two capacitor plates closer together, increasing their capacitance. A larger capacitor can hold more charge, so a momentary current carries charge from the battery (or power supply) to the capacitor. This current is sensed, and the keystroke is then recorded.



That makes perfect sense, and is kind of neat. What I am curious about, is what happens to that extra charge afterwards. Is there some sort of discharge mechanism? I suppose, that would be also necessary to differentiate between single keystrokes and continuous depression (register stroke current, then register the discharge). What would happen to the capacitor if there was no such discharge mechanism, but its capacitance was suddenly reduced?


If capacitance is reduced, and the charge stays the same, then, according to $Q = C \Delta V_C$, the difference of potentials on plates of capacitor should increase and exceed that of a power supply thus reversing the current. Is that what is happening, and the keystrokes are recorded by sensing not only the existence of the current, but also its direction?




atmospheric science - Are the air particles in today's wind on earth (more or less) the same as the air 2/3 billion years ago?


So the air pressure on earth remains relatively constant, right?


Is there air gained or lost through transitions of any sort? e.g. plumes going out into space, earth gaining new air particles with gravity from air in the solar system.



I was curious to know whether in general the particles that brushed up against the dinosaurs, and against Homo erectus's face, are the same as the wind today, ebbing and flowing as pressure variation directs it. Is this true? :)


I suppose air goes into trees, and into mammals, and recycles around. It seems like the recyclability rate must be very very high.


Is this logic correct? It's all very fascinating!



Answer



The atmosphere of the Earth is mainly composed of nitrogen (N2, 78%) and oxygen (O2, 21%) molecules, which together make up about 99% of its total volume. The remaining 1% contains all sorts of other stuff like argon, water and carbon dioxide, but let's ignore those for now.


As you probably know, the oxygen we breathe is produced by plants from water and carbon dioxide as a byproduct of photosynthesis. Conversely, animals (including humans) use the oxygen to burn organic compounds (like sugars, fats and proteins) back into water and carbon dioxide, obtaining energy in the process. So do many bacteria and fungi, too, and some oxygen also gets burned in abiotic processes like wildfires and the oxidization of minerals.


The result is that oxygen cycles pretty rapidly in and out of the atmosphere. According to the Wikipedia article I just linked to, the average time an oxygen molecule spends in the atmosphere before being burned or otherwise removed from the air is around 4,500 years. The most recent known Homo erectus fossil dates from about 143,000 years ago, so the probability that a particular oxygen molecule hitting your face today has been around since that time is roughly $\exp(- 143000 / 4500) = \exp(-31.78) \approx 1.58 \times 10^{-14}$, i.e. basically zero.


Of course, the oxygen atoms used for respiration don't disappear anywhere: they just become part of the water and carbon dioxide molecules. Those that end up in carbon dioxide usually get photosynthesized back into free oxygen pretty soon, unless they happen to get trapped in a carbonate sediment or something like that. The oxygen atoms that end up in water, on the other hand, may spend quite a long time in the oceans before being recycled back into the air; if I'm not mistaken, the total amount of oxygen in the hydrosphere is about 1000 times the amount in the atmosphere, so the mean cycle time should also be about 1000 times longer. Still, eventually, even the oxygen in the oceans gets cycled back into the atmosphere. Thus, while the oxygen molecules you breathe might not have been around for more than a few thousand years, the atoms they consist of have been around since long before the dinosaurs.




How about nitrogen, then? Perhaps a bit surprisingly, given how inert nitrogen generally is, it's also actively cycled by the biosphere. Unfortunately, the actual rate at which this cycling occurs seems to be still poorly understood, which makes estimating the mean cycle times difficult.



If I'm reading these tables correctly, the annual (natural) nitrogen flux in and out of the atmosphere is estimated to be somewhere between 40 and 400 teragrams per year, while the total atmospheric nitrogen content is about 4 zettagrams.


This would put the mean lifetime of a nitrogen molecule in the atmosphere somewhere between 10 million and 100 million years, well above the time since Homo erectus first appeared (about 1.8 million years ago). Thus, it seems that most of the air molecules around you have probably been around since the days of Homo erectus, and some of them might even have been present during the age of the dinosaurs, which ended about 66 million years ago.


everyday life - Why 8 am is darker than 12 pm?


Why is 8 am (just after sunrise) darker than 12 pm (noon)? Does it have something to do with the geometry of spheres? Or is it due to the atmosphere?



Answer




The main factor is geometry. The Sun doesn't change, and it emits a constant "radiant intensity", defined as "luminous flux per solid angle"


$I=\frac{dF}{d\Omega}$


And the illumination depends on it. Imagine a disc on the ground. Of course it is much more illuminated if the source is opposite to it. Explicitly, the illumination is


$ \frac{dF}{dS_2}=\frac{Id\Omega}{dS_2}=\frac{I\cdot \cos\alpha_2}{r^2}$


And so the illumination directly depends on the cosine of the angle of incidence. That's why you want to sunbath perpendicularly.


Flux, irradiance and solid angle


Besides this, the atmospehre plays also a role. When the rays are not perpendicular, they must travel along more thickness, so they are very slightly mittigated (thats why you shouldn't sunbath too much at noon). But this effect is much more noticeable as for the scattering of wavelenghts. That's why sunset and sunrises look red. Check this image:


https://thesciencegeek01.files.wordpress.com/2015/09/redsunsetdiagram1.png


quantum mechanics - Spin in magnetic field and eigenvalues


We have some arbitrary quantum state, lets say $$\vert\Psi\rangle=\alpha_{1}\vert\uparrow\rangle+\alpha_{2}\vert\downarrow\rangle= \begin{pmatrix} \alpha_{1} \\ \alpha_{2} \\ \end{pmatrix}$$. And we act on it with some operator, whatever is appropriate math for it, lets say in this case particular case a linear combination of Pauli matrices $$ \sigma_{n} = \begin{pmatrix} n_z & n_x-in_y \\ n_x+in_y & -n_z \\ \end{pmatrix}$$ My understanding of physical process, that is happening here, is as follows. We have some arbitrary electron, with operator we put on magnetic field in arbitrary direction and eigenvalues make sure we calculate expectation value correctly. So eigenvalues are measurement result. (?)


On the other hand operator acts on a vector and we get a new state. (Rotates the state on a Bloch sphere?) So eigenvalues change basis vectors and we rotate a state.


How does the operators relate to rotations on Bloch sphere? In my interpretation eigenvalue existence mess up probabilities.


Addition: It seems I can't connect the idea of an observable (mathematical construction for expectation value - to get measurement results) and idea of an unitary operator which changes state of a system (due to magnetic field). How are they related? Are they one and the same?



Answer




I would re-state the problem slightly to conform to conventional notation. In any discussion of magnetic resonance -- of spin 1/2 particles, be they electrons, or say, protons-- the direction of the magnetic (polarizing) field is always chosen as +z in a laboratory coordinate frame. Then the two stationary states, which (for convenience) I write as |+> and |->, correspond to the spin aligned with its z-component of angular momentum parallel or anti-parallel to the polarizing field.


The stationary states are in fact eigenstates of the z component Pauli matrix, with respect to the current laboratory coordinate frame.


Then the operator you illustrate is a linear combination of all three Pauli matrices in this frame. Any Pauli matrix (or linear combination thereof) acts mathematically as the generator of an infinitesimal rotation of the spin. Your conventional x, y, and z Pauli matrices each individually generate infinitesimal rotations of the spin about these axes. The best discussion I know of this is Messiah's book (as noted in an earlier response, above); you may also want to consult M. E. Rose 'Theory of Angular Momentum,' which is available as a Dover reprint.


Rotating an eigenket gives you a new state, but it doesn't change your basis, which is defined essentially by the direction of the polarizing magnetic field. You can rotate the spins any way you like, but your result will still be expressed as a linear combination of those same eigenkets, as long as you don't physically change the direction of the field.


So far we have dealt exclusively in quantum mechanical rotation.


Once we bring in the 'Bloch sphere', we need to introduce the Bloch equations, which are classical. In fact, the Bloch equations without relaxation are exactly generators of infinitesimal rotations, but for 3 dimensional classical vectors, specifically (in the case of magnetic resonance) the magnetization. I usually (for convenience) try to keep the classical and quantum views separate in my mind-- so I think of the Pauli matrices as rotating an individual spin, and the Bloch matrices as rotating a bulk magnetization vector comprising the resultant of many millions of individual spins.


However, I emphasize that the there is no logical requirement to view things this way. The simple view is that the Pauli matrices give a quantum rotation and the Bloch matrices give a classical rotation. Nonetheless, you will see purely quantum discussions which refer also to the Bloch sphere-- I don't think this should cause you any confusion.


I hope this helps a bit; do not hesitate to request clarification.


Wednesday, January 20, 2016

lagrangian formalism - Constrained Hamiltonian systems: spin 1/2 particle


I am trying to apply the Constrained Hamiltonian Systems theory on relativistic particles. For what concerns the scalar particle there is no issue. Indeed, I have the action \begin{equation} S=-m\int d\tau \sqrt{-\dot{x}^\mu \dot{x}_\mu}\tag{1} \end{equation} and computing the momentum \begin{equation} p_\mu=\frac{m\dot{x}_\mu}{\sqrt{-\dot{x}^2}}\tag{2} \end{equation} I see that it satisfies the constraint $p_\mu p^\mu+m^2=0$. I then proceed to quantize the system with the Dirac method.


I am finding issues with the relativistic massless spin 1/2 particle. Indeed, it is described by the space-time coordinates $x^\mu$ and by the real grassmann variables $\psi^\mu$, according to my notes. The action should take the form \begin{equation} S=\int d\tau \space \dot{x}^\mu\dot{x}_\mu+\frac{i}{2}\psi_\mu\dot{\psi}^\mu\tag{3} \end{equation} which exhibits a supersymmetry on the worldline, the supersymmetric conserved charge being $Q=\psi^\mu p_\mu$. According to the lecturer I should find the constraints \begin{equation} H=\frac{1}{2}p^2, \quad Q=\psi^\mu p_\mu\tag{4} \end{equation} i.e. the dynamics on the phase space should take place on hypersurfaces $$H=0, Q=0.\tag{5}$$


My question: How can I derive these constraints? They should arise simply with the definition of momenta, but, having no constants to work with, I'm left with \begin{equation} p_\mu=\dot{x}_\mu\\ \Pi_\mu=\frac{i}{2}\dot{\psi}_\mu\tag{6} \end{equation} and I don't know what to do with them. I see that, in principle, the first constraint is obtained by setting $m=0$ in the constraint of the scalar particle for example, but what if I want to derive it without the previous knowledge? And what about $Q$?


Edit: By intuition, knowing that the model exhibits a ${\cal N}=1$ supersymmetry, I may understand that the dynamics must take place on a surface such that $H=const$ and $Q=const$ (then I could set the constant to zero without lack of generality?), being $Q$ and $H$ conserved charges. Is it the only way to find these constraints? Should I need this previous knowledge about supersymmetry to study the model? I think I should be able to find these constraints just by looking at the Lagrangian itself.



Answer






  1. We consider here the massless case $m=0$. Let us start from the Lagrangian$^1$ $$L_0~=~\frac{\dot{x}^2}{2e} +\frac{i}{2}\psi_{\mu}\dot{\psi}^{\mu} \tag{A}$$ with an einbein field $e$, cf. e.g. this Phys.SE post. If we introduce the momentum $$ p_{\mu}~=~\frac{\partial L_0}{\partial \dot{x}^{\mu}} ~=~\frac{\dot{x}_{\mu}}{e},\tag{B}$$ the corresponding Legendre transformation $\dot{x}^{\mu}\leftrightarrow p_{\mu}$ yields a first-order Lagrangian $$ L_1~=~p_{\mu}\dot{x}^{\mu} +\frac{i}{2}\psi_{\mu}\dot{\psi}^{\mu}-eH, \qquad H~:=~\frac{p^2}{2}. \tag{C}$$ This explains OP's first constraint $H\approx 0$, which is indirectly due to world-line (WL) reparametrization invariance, cf. this Phys.SE post.




  2. It is unnecessary to introduce momentum for the fermions $\psi^{\mu}$ as the Lagrangian $L_1$ is already on first-order form, cf. the Faddeev-Jackiw method.




  3. The Lagrangian $L_1$ has a global super quasisymmetry. The infinitesimal transformation $$ \delta x^{\mu}~=~i\varepsilon\psi^{\mu}, \qquad \delta \psi^{\mu}~=~-\varepsilon p^{\mu}, \qquad \delta p^{\mu}~=~ 0 , \qquad \delta e~=~ 0, \tag{D}$$ changes the Lagrangian with a total derivative $$ \delta L_1~=~\ldots~=~i\dot{\varepsilon}Q+\frac{i}{2}\frac{d(\varepsilon Q)}{d\tau}, \qquad Q~:=~p_{\mu}\psi^{\mu}, \tag{E}$$ for $\tau$-independent Grassmann-odd infinitesimal parameter $\varepsilon$.





  4. OP's other constraint $Q\approx0$ arises by gauging the SUSY, i.e. $\delta L_1$ should be a total derivative for an arbitrary function $\varepsilon(\tau)$. On reason to do this is given in Ref. 2 below eq. (3.3):



    Because of the time component of the field $\psi^{\mu}$ there is a possibility that negative norm states may appear in the physical spectrum. In order to decouple them we require an additional invariance and, inspired by the Neveu-Schwarz-Ramond model, it seems natural to demand invariance under local supergauge transformations.





  5. Concretely, we impose $Q\approx0$ with the help of a Lagrange multiplier $\chi$. This leads to the Lagrangian $$ L_2~=~L_1-i \chi Q~=~p_{\mu}\dot{x}^{\mu} +\frac{i}{2}\psi_{\mu}\dot{\psi}^{\mu}-eH - i \chi Q .\tag{F}$$




  6. Let us mention for completeness that in order to have gauged super quasisymmetry of the new Lagrangian $L_2$, the previous transformation $\delta e= 0$ needs to be modified into $$ \delta e~=~2i\chi\varepsilon, \qquad \delta \chi~=~\dot{\varepsilon}.\tag{G}$$





  7. An alternative perspective is the replacement $$L_2~=~ L_1|_{\dot{x}\to Dx}\tag{H}$$ of the ordinary derivative $$\dot{x}^{\mu}\quad\longrightarrow\quad Dx^{\mu}~:=~\dot{x}^{\mu} -i\chi \psi^{\mu}\tag{I}$$ with a gauge-covariant derivative $Dx^{\mu}$. Here $\chi$ is a compensating gauge field. The gauge-covariant derivative transforms as $$ \delta Dx^{\mu}~=~i\varepsilon(\dot{\psi}^{\mu}-\chi p^{\mu}).\tag{J}$$




References:




  1. F. Bastianelli, Constrained hamiltonian systems and relativistic particles, 2017 lecture notes; Section 2.2.





  2. L. Brink, P. Di Vecchia & P. Howe, Nucl. Phys. B118 (1977) 76; Below eq. (3.3).




  3. C.M. Hull & J.-L. Vazquez-Bello, arXiv:hep-th/9308022; Chapter 2, p. 7-8.




--


$^1$ Conventions: We use the Minkowski sign convention $(-,+,+,+)$ and we work in units where $c=1$.


classical mechanics - Moment of a force about a given axis (Torque) - Scalar or vectorial?

I am studying Statics and saw that: The moment of a force about a given axis (or Torque) is defined by the equation: $M_X = (\vec r \times \...