Q: What is the probability that two randomly chosen people will have been born on the same day?

Physicist: 0.0035% or about 1 in 28,500.

Two randomly chosen people.

The most recent, complete information I could find is the 2000 American census.  There they have the number of people of a particular age sorted into five year blocks, so I assumed a “step function” for the age distribution (which isn’t terrible).  Also, for the 85+ block, it turns out that 80 year olds live to 90 on average, so I figure the average number of days people live in the 85+ block is more or less the same as the others.

This qualifies as a “useless statistic”.   It’s fairly wrong world wide, quite wrong when you consider the people you work with (probably), even more wrong at a high school graduation, and really wrong in a maternity ward.  Even worse, it changes from year to year (well… decade to decade).

But it does line up well with the approximately 28,600 day average human life span (which is actually a little surprising).

Posted in -- By the Physicist, Probability | 4 Comments

Q: Which is a better approach to quantum mechanics: Copenhagen or Many Worlds?

Physicist: The Copenhagen interpretation requires that new laws be created that, in addition to being impossible, are completely unnecessary, physically unfeasible, and utterly unjustifiable.  The basic laws of quantum mechanics, when applied at all scales, give you the Many Worlds interpretation.  No fancy rules, no awkward questions.  Even better, what seems to be wave function collapse in Copenhagen is actually described by the Many Worlds approach.  So why choose the Copenhagen interpretation over Many Worlds?  No damn reason at all.

Many Worlds vs. Copenhagen. On the one hand there are different versions of you, and on the other you've got vaguely defined mental powers.

If you’re already familiar with the basic ideas behind the measurement problem, and the divide between Copenhagen and MW, feel free to skip down to the word “catawampus“.

Bell’s theorem demonstrated that either information travels faster than light (Bohm) or particles exist in many states simultaneously (everybody else).  There is a lot wrong with Bohm’s theories, not the least of which is that it doesn’t have a relativistic formulation (which may not sound too bad, but that’s really damning in physics circles).

The best example is the double slit experiment: shine coherent light on two slits and look at the pattern that’s projected onto a screen.  You get a nice interference pattern of dark and light bands on the screen, which is fine since light is a wave (don’t stress about wave/particleness here).  What’s messed up is that even when the light intensity is turned way down to just one photon at a time, the photon still tends to land where the bright bands were, and almost never lands where the dark bands were.

You have to repeat the single-photon double-slit experiment a lot to see the pattern, but there it is. In order to get these interference fringes from two slits the photon must interfere with itself. It's going through both.

Alright, so everybody has come to accept that very small things have no problem being in several states simultaneously (if you don’t buy it, then just take it as read).  But here’s the weirdness.  It turns out that particles only exhibit the super-position properties when “you” can’t tell the difference between the various states.  As soon as “you” can single one out, then suddenly the super-positionness goes away.  I’m putting quotes (“”) around “you”, because “you” could be a person, or a computer, or even another particle.  Just a weakness of the language.

This is the measurement problem.  Things can be in several states at the same time, but only if nothing observes* them.  The moment they’re observed* they are found to be in only one state, and they continue on their merry way in only that one state.  The most straight forward (brute force) example is covering one slit and noting that the interference bands disappear.  By covering one slit you’re “observing*” that the photon is going through the other.  Most examples are more complex and subtle.

(* this is another weakness of the language.  “Observe” would seem to imply a hierarchy; “this observes that”.  A better term might be “interact with”.)

Here’s where the split comes up.  Copenhagenists will say that something, usually size, or complexity, or random collapse, or, worst of all, consciousness, causes “wave function collapse” (all but one state disappears).

Many Worlds adherents will say that wave function collapse never happens, and that the trick is realizing that both the particles involved and the observer* are in multiple states.  After the observation* the multiple states of the particle are now entangled with the multiple states of the observer*.  In the double-slit test you’d have the two states: “photon goes right and observer sees it go right” and “photon goes left and observer sees it go left”.  Both states persist, and there’s no collapse.

 

Catawampus: You’ll often hear that there’s no experiment that can be done to prove which approach is the correct one.  I’m of the opinion that the experiments have already been done, but that most people (myself included) don’t like the results.  However, among people who have stopped to consider the options (and there aren’t many good reasons to do so), most of us have decided to accept the results and move on.

The big advantage behind the Copenhagen interpretation is that it makes people (like you!) important, and different from particles.  <sarcasm>Sure, they may be in multiple states, but I’m definitely in exactly one state.  Unlike particles, people can tell the difference.</sarcasm>

It’s creepy to think that there are different versions of yourself “out there” doing “stuff”, but it’s awesome to assume that you’re special and that your mind (not brain) has some kind of power over reality.

There is a version of the Copenhagen interpretation of quantum mechanics where consciousness (usually human, sometimes God, sometimes Gaea, …) plays a key role.  I feel pretty comfortable dismissing it out of hand.  There’s a post here that talks about it a little.

Time and again, we’ve managed to show that larger and larger objects can be in multiple states, using the double slit experiment or variations of it.  At last check, the double slit experiment was successfully preformed on C60F48, which has fully 108 atoms, or 2,424 protons, neutrons, and electrons.  The entire molecule (actually, thousands of them) actually interfered with itself, demonstrating the ability to be in multiple states.

Which raises the question: what’s the damn problem?  Everything that can be tested has demonstrated quantum superposition, so why not just extend that to “everything obeys the same quantum mechanical laws, including superposition.”?  Why not indeed?

One may be tempted to say “the physics at small scales is just different!”.  Fair enough.  However, there are no physical laws that work differently on different scales.  For example, at very small scales water acts like honey, and to swim you need to use things like flagella.  At the other end of the scale (our scale) water behaves… like water, and things like fins and flippers suddenly work really well, but flagella don’t.  However, the same physical laws (specifically, the Navier-Stokes equation) govern everything.

More generally, all laws apply at all scales, it’s just a question of degree.  Relativity works at all velocities, but you don’t notice the weird effects until you’re moving really fast.  What we call “Newton’s laws” are just an approximation that work at low speeds.

If the Copenhagen “size argument” (that larger objects somehow have different laws) holds up, it’ll be the first of its kind.

So how can the many worlds hypothesis hold up?  Either we’re in multiple states or we’re not, and (most) people don’t feel like they’re in more than one state.

In very much the same sense that relativity includes Newtonian dynamics as a special case, the Many World hypothesis actually contains Copenhagenism.

Normally in quantum mechanics, when you’re trying to predict the behavior of a system, you just let all the components evolve in time according to the Schrödinger equation.  If you follow a particular state or object, then you’ll find that it experiences wave function collapse all over the place.

That is to say: If you require (artificially) that some particular thing must remain in one state (or small set of states) by disregarding its other states, then that object will (seem) to see wave function collapse.

I should repeat that, because it bears repeating:

According to the Many Worlds approach, the individual states of an object witness wave function collapse all the time, but taken as a whole, there is no collapse.

Here’s an simplified example with very little physical basis, but that hopefully gets the point across.

In what follows the \square are collision points.  Either the particles pass each other, or bounce off each other and change direction.  They can’t both go down the same path.  The \boxminus are splitters.  When a particle hits this it has a 50% chance of passing, and 50% chance of reflecting.

To calculate the probabilities of how the two quantum particles will come out of the machine you have to sum over ever possible path. You have to take into account, not only every "choice", but every interaction.

This map shows how two particle move through some machine.  Either they bounce off each other at the first \square and you have blue-up/red-down, or they pass each other and you have blue-down/red-up.  These two states then go on to the splitters and so on.  There is no collapse, and every possibility (every path, every interaction) is included (like the double slit).

But what if you were the to track the blue particle?  Put your foot down and insisted that it remain in just one state and take just one path through the machine?  Even more profound, what is the effect on the red particle (from the new one-path-point-of-view of the blue particle)?

Easy enough, pick one of the blue paths (worlds) above.

1) The particles start. 2) In this world they pass each other. They didn't have to, they just happened too. 3) Since we're insisting the blue particle stays on one path, it must pick a path: reflect. The red particle, however, is free to take multiple paths, and does. 4) In this world the blue particle bounces off the red particle at the second box. But this interaction means that the blue particle now knows where the red particle is. Collapse!

This story is just one of the stories encompassed in the Many Worlds picture.

Here’s a slightly different one.  What would happen if the blue particle went through the splitter instead of reflecting?

1&2) Same as last time. 3) Since we insisted that the blue particle must be in one state, it must either reflect or pass. In this case it passes. The red takes both available paths. 4) The blue particle can't interact with the red, so this time the red particle is free to take even more paths simultaneously (as far as this version of the blue particle is concerned).

So the big point is that the Copenhagen wave function collapse is strictly an illusion created by restricting your attention to a particular state of one object.  So why is our attention stuck on just one state?  We’re not special, just victims of conservation laws.

Looking down on the double slit experiment from outside you can ask questions like “what is the probability that the photon will go through each slot?”.  You have no “givens” to affect your probabilities so you say “50/50”, and you’re right.  The photon goes through both, but since there’s only one photon (conserved number of photons), it does it in a particular (some what obvious) way: it combines the states “left/not right” and “not left/right”.

Now say you’re presented with two doors.  You also can’t be in the state “right and left”.  Now when you go through one door, because you’re interacting with yourself, you have givens that affect the probabilities.  Ask yourself, “What is the probability that I went through the left door, given that I just went through the right door?”  Zero, baby!

The version of you that when through the left door will be able to make a very similar calculation.

Finally (without going into too much detail), the Copenhagen interpretation also violates a number of very straight forward physical laws.  Conservation of information (supported by everything else, including logic), time reversibility (again, everything else), and information flowing backward through time (spacelike information exchange or “spooky action at a distance”) are only the most direct and grievous examples.

Posted in -- By the Physicist, Paranoia, Philosophical, Physics, Quantum Theory, Skepticism | 58 Comments

Q: Why is our vision blurred underwater?

Physicist: The speed of light depends on the medium it moves through.  So as light moves from air to glass, for example, it slows down.  Because light is a little spread out, speeding up or slowing down can make it change direction.  When the boundary between mediums is slanted, one side of the wave changes speed first, causing the wave to “swing around”.  This change in direction is called “refraction”.

Sleds are faster on snow than pavement. So when the right runner hits the asphalt it slows down, and the sled turns right. As it happens, the rider tends to turn left.

It’s worth noting (because this is a common confusion), that when the speed of light slows down it doesn’t have any profound repercussions for relativity or time travel or anything like that.  In fact the reason it slows down is that it’s being absorbed and re-emitted by atoms as it moves through the material.  In between atoms it’s still as fast as ever.  Picture it as the fastest, stop-and-go traffic evers.

Back to the point.  All lenses use refraction to change the direction of light.  The surface is curved in such a way that all the light from one direction is directed to a single point (on the focal plane).  So without refraction, you’ve got no lensing.

Physicists measure the speed of light through materials using the “index of refraction”, which is defined as n=\frac{c}{v} where c is the usual speed of light and v is the speed in the material.  So, if n=2, then light is traveling through that material at half speed.  In general, the shallower the incoming angle and the greater the difference in n’s, the more light will bend.  The exact relationship is summed up in Snell’s law.

For those of you keeping score; “Willy Snell” is the second nerdiest name in physics, second only to “Norbert Wiener“.

There's a big difference between the n's of air and your cornea, so the lens bends light very effectively. But, there's very little difference between the n of your cornea and the n of water, so the lens barely has any effect at all.

Light travels through water, the human cornea, and the “aqueous humour” at about the same speed, so they all have nearly equal n’s.  Because the index of refraction is so similar, light barely changes direction at all.  It’s like sledding from one kind of snow to another, nearly identical, kind of snow.  Normally, the light spreading out from a single point in your field of view gets pulled back together, into a single point again, on the back of the eye (top part of the picture).

However, without a functioning lens, the light from a particular point (in your field of view) keeps spreading out, and is projected onto a large area in your eye.  That means that you can’t tell exactly where the point is in reality.  Objects near each other end up smearing over the same areas on the back of the eye, etc.  In other words; it’s blurry.

By the way, in a functioning eye, there’s very little abstraction going on in there.  If you were small enough to crawl inside some one’s eye (only try this with close friends) you would actually see a picture of whatever they’re looking at projected onto the back wall of the eye.

Being under water is about as close to not having a cornea as you can come without surgery.  In fact, the only reason that you’re able to see anything other than color is that the iris (being fairly small compared to the rest of your eye) forms a pinhole camera.  (The image above, though beautiful, is not to scale)

What about goggles?  There’s still a couple of medium changes (water, plastic, air) so light still bends.  But goggles are held a little away from the eye, so that most of the light you’re seeing has hit the surface head on (meaning a small angle change) instead of at a shallow angle (which causes a big angle change).  You’ll notice that if you look at things near the edge of your goggles they’ll be blurrier.  So, that’s why.

The shallower the incoming angle, the greater the bending and also the greater the lensing effect, which in this case serves to make things more blurry. Luckily, when we look at things we tend to turn our head toward them, and the angle of the light from straight ahead isn't changed much.

What’s most important is that the lens in your eye is allowed to do what it does, and that requires the proper curvature (check) and the proper difference in indexes of refraction (n=1 and n=1.4), which requires air.  Those creatures lucky enough to be able to see both in water and air do it with a combination of mucus goggles, and corneas capable of changing shape dramatically.

Posted in -- By the Physicist, Biology, Physics | 9 Comments

Q: In the NEC “faster than light” experiment, did they really make something go faster than light?

The original question was: I was reading an article by Lijun Wang the experiment I think is called the NEC experiment.  Basically what the guy does is speed up light so that it goes faster than the than the speed of light.  If anything goes faster than light does that mean it goes backwards in time?  Is this experiment genuine?

Here’s a description of the experiment.


Physicist: This experiment showed up in the popular media about ten years ago, and it’s still bothering people today.  The very short answer is: nope!  No physical object ever travels faster than light, and yet the experiment is real. This is pretty tricky, but it comes down to a distinction between “phase velocity” (light speed, “c”) and “group velocity” (depends on the medium).  The experiment essentially requires a “standing wave” in the test chamber, with waves at many frequencies. These light waves interact in such a way that they create peaks, like when you pluck two slightly-out-of-tune guitar strings and you hear a pulsing sound.  These peaks are usually called “wave groups” or “wave packets”.  The peaks move around, and you can more or less keep track of where they are (more exactly; you can infer where they were after the fact). The pivotal point is that the experiment isn’t about turning on some kind of laser, that sends out a pulse, that suddenly moves faster than light.  The experiment is about peaks made out of standing waves, and the standing waves are already all over the inside of the test chamber when the experiment begins. In the middle of the camber is a specially constructed cloud with extremely weird and strong optical properties, that changes how fast light of various frequencies move through it (slowing only, but to various degrees).  Specifically, the lower the frequency, the slower it’s allowed to propagate through the cloud.  This is a weird set up, and it’s difficult to produce.

Waves of different frequencies travel at different speeds and that gives rise to "groups" or "wave packets". The red boxes are examples of how waves of different frequencies can add up. The blue box shows what happens to various waves as they pass through the cloud.

The above image was jacked from the original author’s follow up explanatory paper. A good estimate for the group velocity in general is v_g=\frac{\partial\omega}{\partial k}, where \omega is the angular frequency and k is the wave number.  The take home message from this is; by creating a material that has a very strong relationship between \omega and k you can dial up the group velocity to ridiculous levels (in this case -310c).  For comparison, the group velocity of light in water differs from the phase velocity by less than 1% (which is different from 31,000%). So, by changing how the different frequencies move through the cloud you change how the peaks move around (making them faster than light in this case), however nothing is actually moving faster than light, just a non-information-carrying effect. A much simpler example of this sort of thing is the “scissor paradox”.  Although the execution is different, the underlying idea is exactly the same. When you close a pair of scissors the point where the blades intersect is moving much faster than either blade, and if you were to construct a pair of scissors large and fast enough, then the point of intersection could easily move faster than light, but it wouldn’t be able to carry any kind of information (in part) because you’d already see the slower-than-light blades moving long before the intersection point got to you.  Also, it’s not like you can mail a letter by attaching it to an intersection.  An intersection isn’t an actual “thing”.

Nothing can physically move faster than light. However, abstract points, like the point of intersection or even the peak of a wave packet, can move as fast as they want.

This NEC thing is a similar “paradox”.  And again, the author was kind enough to explain this whole thing here. To answer the follow up, “does going faster than light mean you go back in time?”: maybe.  Probably.  But, it’s never come up.

Posted in -- By the Physicist, Physics, Relativity, Skepticism | 15 Comments

Q: How does a Tesla coil work?

Physicist: Stripped down to it’s most essential parts, a Tesla coil is a wire sticking out of the ground. To get sparks to fly out of the top the rest of the machine “sloshes” electrons up and down the wire.
The picture you should have in your head is a long bathtub, open to the ocean on one end.  The machinery of the Tesla coil is like some dude in the bathtub sliding back and forth, splashing water (electrons) out of the closed end, while the tub is refilled from the ocean (ground).

The electricity in the primary coil is what’s doing the pushing, and the electricity in the secondary coil is what’s being pushed.  To understand how the driving mechanism works requires a new metaphor and some answer gravy.

Aside from inspiring fear, Tesla coils are useless. Truly, Tesla was a genius. The strange shape is an attempt to avoid arching from the torus to the primary coil, which is bad.

Answer gravy: To get sparks to really fly you need very high voltage (up to several million volts) at a fairly exact frequency.  The current that flows up and down the secondary coil, and sloshes out the top, has a high resonant frequency (~MHz, unless the coil is ridiculously huge) that you really can’t do much about.  But the current coming out of the wall has a frequency of only 60 Hz (50 Hz for our Old World readers).

One possible circuit configuration for a Tesla coil.

So how do you change frequencies?  The answer is you “pluck” the primary coil.  For example: If you pick a guitar string once a second you have a frequency of 1 Hz, but the string vibrates on its own at whatever frequency it’s made for (~10 kHz).

The AC mains have a low frequency (60 Hz) while the secondary coil needs to be driven at a high frequency (~1,000,000 Hz).  That means that the secondary will slosh back and forth thousands of times every time the current from the wall turns over just once.  Since the fast part of the circuit is so much faster than the slow part, you can just pretend that the current from the transformer is DC (direct current = 0 Hz).

The secret to plucking is to change the circuit’s “shape” using a spark gap.  Spark gaps have some pretty slick properties.  They have an essentially infinite resistance until a high enough voltage is applied across them, at which point they spark (hence the name).  The spark you see is the air being pulled apart and ionized.  Now ionized gas is a really good conductor, so a spark is like instantly closing a switch.

Also, spark gaps are the cheapest circuit element evar.  Can you cut a wire?  Now you got a gap!

Also, adding spark gaps to a device is one of the quickest ways to bridge the divide between regular and mad science.

The transformer on the left forces charge to build up in the capacitor on the top. But the voltage across a capacitor is proportional to the amount of charge it's holding, so eventually the voltage is high enough to trip the spark gap.

The only job that the slow part of the circuit has is to charge the capacitor (pull back the string).  When the spark gap sparks (pluck!) the fast part of the circuit takes over, and the slow part is essentially ignored until all the energy is exhausted by exciting the secondary coil (string vibrates and slows).

With the spark gap active the charge can flow out of the capacitor and swing back and forth many times, very fast (thousands to millions of times per second). The current through the primary coil then drives current up and down the secondary, causing electrons to "overflow" from the top of the Tesla coil. The "overflow" is a delight to children of all ages.

As current flows through the primary it creates a voltage across the secondary that’s so high that electricity actually flies out of the top of the coil, despite having nowhere in particular to go.  It generally takes at least several hundred thousand volts to make that happen.

The loop in the picture above forms an RLC circuit with a high resonant frequency (that matches the frequency dictated by the secondary).  As the energy in this system runs out the voltage needed to maintain the spark gap (which is much less than the voltage needed to start it) is lost, and the whole thing returns to the slow, charging phase.

Since the power supply oscillates at 60 Hz, the whole system briefly turns off 120 times every second (the voltage is +, 0, -, 0, +, 0, …).  For this reason Tesla coils have a very loud 120 Hz hum that sounds “staticy” and ominous, as opposed to Jacob’s ladders which are continuous, and tend to sound more like “tearing”.  Connoisseurs, I’m sure, will agree.

Posted in -- By the Physicist, Engineering, Physics | 53 Comments

Q: What are Feynman diagrams, how are they used (theoretically & practically), and are there alternative/competing diagrams to Feynman’s?

Physicist: Feynman diagrams are primarily a way to keep track of what you’re doing.  Physicists aren’t geniuses or anything, and they get distracted pretty easily.

When you’re trying to calculate the probability of a particular particle interaction you’ll find yourself integrating over (adding up) every possible position and momentum of every involved particle.  Moreover, when there are several different ways that you can get a particular result you need to keep track of all the different ways that the result can happen.
Feynman diagrams are not technically necessary, but they do help you keep track of variables and permutations.  It turns out that particle physics is a lot more complicated than you’d think.  That holds true even if you think it’s really, really complicated.

Feynman diagrams were created alongside the field of quantum electrodynamics and (as far as I know) no one has suggested another diagram system (if it ain’t broke…).  They allow you to write down the important integrals in the form of a picture, which makes them far more understandable to just about anybody.
Ultimately, once a diagram has been made, and an integral is created and evaluated from it, you end up with the probability amplitude of that particular particle interaction taking place.

The Feynman diagram for the Coulomb interaction (electric force), along with the parts of the Feynman integral they correspond too. Every part of this is really nasty. For example, that "g" is actually 16 numbers.

The electric force (what physicists call the “Coulomb force” to look smart) is mediated by photons.  That is to say, particles with charge push or pull on each other using photons.  The diagram above is the “first order Feynman diagram” for two electrons repelling each other.  The probability amplitude of two electrons with momentum p and k pushing off of each other and flying off again with momentum q and l is given by:

A=(-ie)^2\bar{u}(q)\gamma^\mu u(p)\frac{-ig_{\mu\nu}}{(q-p)^2}\bar{u}(l)\gamma^\nu u(k)\delta(p+k-q-l)

The squiggly line is the exchanged (virtual) photon and the solid lines are the electrons.  The diagram generally represents time as the vertical direction and space as the horizontal, but it turns out that on extremely short time and distance scales it’s not super important.

If you’re wondering which particles are virtual and which are real: virtual particles are the ones stuck inside the diagram and real particles are the ones going in and coming out (they might go on to be detected somewhere).

The reason this is a “first order” diagram is because this isn’t the only way this interaction can happen, merely the most likely.  The less likely interactions are also all happening at the same time, they just have less weight.  Super-positions and whatnot.  Quantum mech is good fun.

There are a lot of ways this interaction could have gone down, and each one has it's own equation (many are very similar). The diagrams help you to keep track of which one is which.

The more complicated an interaction is, the less likely it is, and the less it adds to the total probability amplitude.  In the above picture:

b,c) What if two photons are exchanged instead of just one?

d) What if the exchanged photon creates a virtual electron/positron pair en route?

e,f) What if one of the electrons has a virtual photon “up in the air” during the photon exchange?

g,h,i,j) What if one of the electrons emits and reabsorbs a virtual photon right before or after the exchange?

Feynman diagrams apply to a lot more than just electromagnetic interactions.  For example; neutron decay!  For some reason this is the third mention of neutron decay on this website (1 and 2).

A neutron spontaneously emits a virtual W-boson which goes on to turn into an electron and anti-electron neutrino.

Unfortunately, if you really want to learn how to use Feynman diagrams you should get a BS in physics and then take several classes in quantum field theory in grad school.  This stuff is crazy hard.

Posted in -- By the Physicist, Combinatorics, Equations, Particle Physics, Physics, Quantum Theory | 11 Comments