Jan 16 2013

Mark Kelly is right about gun policy. His 85% statistic is bullshit

ScreenClip

Not even close.

I strongly support gun control. This country has gun laws dictated by a relatively small group of highly vocal lunatics who believe all manner of strange things, from doomsday scenarios to byzantine and ghastly conspiracy theories. Any sane political theory dictates that if we take sensible steps to reduce the number of semi-automatic firearms in the US it would probably greatly decrease our homicide rate.

Nevertheless, Mark Kelly and others who declare that 85% of all children killed in the world with guns are killed here in the US are promoting a pants-on-fire lie that is off by probably more than one order of magnitude. We on the left rightly criticize those on the right for making horrifically huge and stupid misstatements. Let's not become that.

Where does the 85% figure even come from? My hope is that anyone who hears it would be immediately incredulous, given that Americans make up only about 4.2% of the world's population and that deplorable violence is pervasive throughout the developing world. Though no citation was immediately apparent, a little Googling turned up a study by Erin Richardson and David Hemenway from 2010, published in what was then the Journal of Trauma, Injury, Infection, and Critical Care (I found the full text here). The authors took data from the 23 richest nations (Australia, Austria, Canada, the Czech Republic, Finland, France, Germany, Hungary, Iceland, Italy, Japan, Luxembourg, Netherlands, New Zealand, Norway, Portugal, Slovakia, Spain, Sweden, England/Wales, North Ireland, Scotland, and the US) furnished by the World Health Organization and filtered the data according to the ICD-10 classification system of causes of death, such as firearm-related homicides, suicides, and several other variants. The data was from 2003, when the US still was under the 1994 Assault Weapons ban, which restricted magazine size and banned many types of firearms.

Richardson and Hemenway found that 87% of all children aged 0 to 14 killed by firearms in these 23 countries were US children. 80% of all firearm deaths in this group occurred in the US. The US overall had a rate 19.5 times higher than the other 22 countries, driven mainly by guns.

Ok, good, stop there. Just say "compared to other developed countries, our laws are ridiculous and it's literally killing our citizens". Great, factual argument.

What about the original claim? The study cited could not possibly have been carried out for all 196 countries in the world. However, the UN (specifically the UNODC) collects data on all homicides in a country, and the data was particularly well-filled-out in 2008 (only 18 countries had no data). I calculated overall homicide rates for the world vs the US using data from 2008.

Whereas the US had a rate 6.9 times higher than the 22 other rich countries, it had a below average rate for the world. The US share of homicides in 2008 was 3.57%, whereas its share of population (surveyed) in 2008 was 4.66%.

98 countries on the list had higher homicide rates than we do. The US's rate was 5.4 homicides per 100000 people. Contrast that with Burundi (21.7), Ethiopia (25.5), or Kenya (20.1). Ivory Coast (cote d'Ivoire) had 10,801 homicides in a population of only 19 million people, a rate of 56.9 per 100000. Jamaica: 59.5, El Salvador 51.9, Honduras, for Christ's sake, 61.3 (4473 homicides in a population of only 7.3 million!).

The statement Mark Kelly made is probably off by a factor of 20. Liberals, myself included, are proud of the fact that we are, for the most part, the party of intellectuals, operating in reality and fact that Fox News and their ilk ignore. Great. Let's act that way.

Dec 28 2012

Thermo for Normals (part 28): How things radiate

Every object, including you, radiates all the time. The radiation comes in the form of light, also known as electromagnetic radiation. Most of the time, this light isn't visible. Right now I'm radiating, and almost all of what is coming off is infra-red (IR), though there is also a tiny amount of radio, and microwaves. Your eyes can't see light with such a long wavelength, so in a dark room, you would see me using an infra-red camera, but not a regular visible-light camera. Hotter things, such as filaments in light bulbs, do radiate light that can be seen. The surface of the Sun, at around 6000 Kelvin, also creates ultraviolet (UV) rays. And the blazing hot corona of the sun emits x-rays, and gamma rays, in addition to all of these previous kinds.

We normally talk about the different types of light as distinct objects. IR is "heat" radiation, because normally when we encounter it that's what it's doing: heating us. Then there's what most people call light, but which I call "visible light"; that kind of light is between 400 nm and 700 nm in wavelength, or between about 4.28 and cycles per second. 1 cycle per second is called a Hertz (Hz). And if you have of them, that's a terahertz (THz). So visible light is 428 to 750 THz. Your eye was adapted to collect light of this type, so that's what you see.

There are two things to consider. 1. What kinds of light does an object at temperature radiate? and 2. How much light does it radiate? Radiation is constantly providing cooling and warming of objects, so we need to know the answers.

A substance is always made up of atoms, and the atoms all have electrons. These electrons orbit the nucleus, but only at certain energies. These energies form discrete steps. When the atoms knock into each other, which they are constantly doing, there's a chance that the electron gets promoted to a higher step. If it picks up a lot of thermal energy from the atoms, it can suddenly jump up a few steps. After a short time, it will fall back down. If it falls down, a particle of light is generated, called a photon. Since the electron loses energy during that fall, the photon must be carrying away energy.

That's right, photons, which are what light is made of, have energy. The energy of a photon depends on its frequency or wavelength. If its frequency is then its energy is . is just a constant, named for Max Planck, and called Planck's constant. If you'd prefer wavelength, it's , where is the wavelength and is the speed of light. Light is characterized only by frequency (or wavelength). The different types of light (radio, microwave, IR, visible, UV, x-ray, and gamma ray) are determined solely by , and so by energy. Thus, the type of light emitted will have something to do with energy the electrons pick up, and hence on the temperature. On the other hand, the brightness only depends on how many photons there are. Brighter light, more photons. So what we have to figure out is how many photons of each type are made.

Here I'm going to write the amount of radiation emitted per second (the power) at each frequency for a body at temperature . The only way I know to derive it is very advanced (a 4th year undergraduate physics major problem), so I'm not going to. After I write it, we can see its general features, and then we can try to figure out the total power radiated. Here it is:

This answers both questions at once. If you want to know the amount of light at some frequency, you just plug in and to this*. If you want to know the total energy being radiated, you just sum up the energies for all the possible frequencies, which we'll do next time. Here is called the emissivity, and it depends upon the object itself. If an object is perfectly absorbing, and reflects nothing, then will be 1 for all frequencies. Otherwise, is somewhere between 0 and 1, and it might depend on frequency.

(*The above is actually per unit area. The bigger the object, the bigger the surface area, and the more photons. Also, since this is a distribution, we have to take an area to get an intensity, but this is a mathematical kink that the reader should not be worried about)

Here's what the spectrum looks like for an object at 300K:
Blackbody300KFor simplicity we'll say this is a black object like coal. The area under the curve represents the total brightness, or the number of photons coming off. The range on the -axis is all in the "far infrared" segment of the electromagnetic spectrum. All of the frequencies lower than this are squished into the far left of the graph, so they aren't visible. Here's a close-up of the left part of the graph:
Blackbody300KLowerThere is very little radio being emitted, a tiny amount of microwaves, and then we get into the Far IR. It's very tempting to just say that objects at room temperature only radiate IR. The fraction of light emitted in the radio/microwave range is only 0.0005%, which I determined by comparing the area under the curve for the radio and microwave parts to the area under the curve for the rest. It's not nothing, but it's quite small.

Now let's look at what happens for an object at twice the temperature, 600 K, which is hotter than your oven gets:
Blackbody300600KWow, the 600 K object radiates a lot more light! I would say, just to eyeball it, that the area under the green curve is more than 20 times larger than the blue curve, and the area, as I've said, corresponds to more photons ("brighter", in a sense, except you can't actually see it). It also goes to larger frequencies than the 300 K object, but it's still very much in the infrared region. It has not even gotten to the "near infrared" region, which is above 300 THz (and which some insects can see). So we think that probably things have to be quite hot before they start to radiate light we can see. Notice that the peak frequency (the frequency that the most photons come off with) has also moved to the right, from about 18 to 40 THz. We'll think about that again later.

Now, let's increase the temperature until we see the object start to glow red. Here's the spectrum at 1600 K:
Blackbody1600KThe tail on the right is finally barely overlapping the visible. A huge, huge majority of the light is still in the Far IR part. This thing is still mostly radiating heat, rather than useful visible light. But your eyes are fairly sensitive, so you can actually see an object start to glow red at 1600 K (I have seen it myself). Now let's heat up to 2100 K:
Blackbody2100KAt 2100 K, you can see that there is significant overlap of the curve with red, and some overlapping with green. Red and green makes yellow, so as you increase the temperature, it turns red hot, and then yellow hot. It never turns "green hot", because the tail always has to start at infrared and go down. Any time green is emitted, red is also emitted. And, of course, once we turn the temperature up even higher, there will be blue photons, and we start to get white light, which is a combination of red, green, and blue.

Now let's check out the surface of the Sun, which is about 6000 K:
Blackbody6000KSurfSunHere we still have a huge amount of IR radiation. A little less than half of the light coming off is still "heat" radiation, which can warm us up but can't be seen. The visible region is well represented, which means we have a significant amount of white light that is just a little bit less blue than it is red and green. Finally, we start to have significant amounts of radiation in the UV range. This is both the near-UV (which include UVA and UVB photons), and some far UV. Above UV would be X-rays and gamma rays. The surface of the sun does not create these; instead, they originate in the upper layer of hot gas around the sun called the Corona, for which the physics may be a bit different.

The phenomenon of objects radiating visible light when they get very hot is called incandescence. Regular old light bulbs use incandescence to create light. Here's what the electromagnetic spectrum for a tungsten filament in a light bulb ( equal to about 3000 K) looks like:
Blackbody3000KLightBulbThat sure seems like a dumb way to make light. Only about 20% of the photons are useful for seeing. The rest are just being radiated out as heat, warming up the glass bulb, the fixture, and your house. It is for this reason that it makes sense to ban incandescent light bulbs. The government has not yet seen fit to do so, but it should. We have much smarter ways of doing this.

Next time we'll look more carefully at the total amount of radiation coming off something at temperature , and we'll also start to investigate what happens when the object is not totally absorbing. We'll look at how things receive radiation, and so why temperature has to be given "in the shade" to be comparable from day to day. We'll see that we feel warmer when this radiation is incident upon us. And we'll see how reflection moderates how hot things get when sitting out in the Sun. Eventually, we'll be able to start talking about the phenomenon of climate change using this information.

Dec 07 2012

Thermo for Normals (part 27): Diffusion and sound

The other day, my wife cooked up some microwavable popcorn while I was down the hall in the bedroom. The sound of the popcorn being made traveled seemingly instantly, but it took a full 5 minutes before the buttery smell came down the hall. Why is one so slow and the other so fast? Both of these rates have something to do with the air. But what?

By now we have a pretty good picture of what is happening in a gas. There are about or so gas molecules in every cubic centimeter. These are flying around at relatively high speeds, and the average speed is something like , or about 500 m/s (about 1100 miles per hour) at room temperature. That is fast...damn fast! Yet somehow the aromatic molecules coming off the popcorn take 5 whole minutes to traverse the approximately 10 meters from the microwave to the bedroom.

Of course, the answer is that the molecules are constantly running into each other. If there were no gas in between the microwave and the bedroom, the molecule would get there in s. But as they try to make their way, they collide with all of the molecules around them, and this makes them bounce off in all sorts of directions. They do eventually make it, but it takes quite awhile.

Let's come up with some quantities that would be useful in thinking about this. A molecule in a gas at temperature experiences a lot of collisions every second. Let's call (that's tau) the mean time between collisions. As an unrealistic example, if a molecule experiences 50 collisions in 10 seconds, the mean collision time is 0.2 s. That's the average time for a molecule to go without running into anything. In air at room temperature this time is about seconds, or 0.5 nanoseconds. That doesn't mean that every molecule goes that long before colliding. Sometimes a molecule will collide much sooner, and sometimes much later. But on average, that's the time.

There is also the average distance that the molecule travels before running into another molecule, called the mean free path, . The gas molecules are separated by only about 3.3 nanometers, so you might thinking that the mean free path is around that value. But amazingly, the mean free path is about 93 nm. The gas molecules travel much farther than one would think before running into something. Imagine standing in a room where there are 3 ft separating you from everybody else on average, everyone walking around and bumping into each other. Then imagine I told you that if you started walking in one direction, you would travel 93 feet without running into somebody. Crazy! But gas molecules are so small, and the forces so short ranged, compared to the spacing between them that it's true. This is why I say that molecules in a gas don't run into each other very often: compared with a liquid they travel a really long distance before running into something.

Suppose we focus in on two imaginary boxes in the air. Box 1 has 9 "special" molecules (solid circles) among its around 25 "normal" molecules (empty circles). Box 2 has only 3 special molecules for the same amount of air. This is not an equilibrium condition. Randomly, the special molecules will travel to the right to even things out. This tendency is called diffusion. The special molecules in box 1 can travel to the right past the imaginary wall. However, note that the special molecules in box 2 can also jump back into box 1. It's just more likely that the ones from box 1 will jump to box 2, since there are more of them.
The net rate of flow of molecules from box 1 to box 2 we call . It will depend, obviously, on how many special molecules there are. Let's call the number of special molecules in box 1 (per unit volume) and the number in box 2 . You notice that the boxes in the picture are of a size , the mean free path. That means that during a time , on average will cross the imaginary barrier from the left to the right, since the distance they can travel is about . The ones in box 2 travel from the right to the left are . is the average speed. So the net flow rate from box 1 to box 2 is the difference between these divided by the time:

Now, the difference turns out to be approximately the derivative, or the gradient of special molecules, times . That is, it's how steeply the number of special molecules drops off per meter. So

I've actually fudged the derivation by not considering that the molecules can travel at all angles. That would have given me an extra factor of 1/3 out front. But it adds significant complexity for very little improvement of understanding.

Let us call the factor the diffusion coefficient, with symbol . Then

which is quite tidy. You can look up the diffusion coefficient for air, which is about 0.2 cm/s. (You could also calculate it by evaluating with the mean free path and mean velocity of a gas, but it's simpler and more accurate to just look it up).

To figure out how long it will take the special buttery molecules to travel the 10 m length of the hallway, we can make 100 million little boxes of width nm. Initially, the far left boxes will have the microwave that has all of that wonderful aromatic special molecules, and the far right boxes have my nose and no special molecules. The density that a person can smell and recognize a molecule is about 0.1-10 parts per million, and a strong smell is about 10-50 parts per million. So the question is, if there's a strong smell initially of 50 ppm in boxes 1 through 5,000,000 (where the microwave is), when does the density get up to 0.1 ppm in box 100,000,000? Because that's when I'll smell it.

We can have a computer do this for us. What you have to do is go to each box, calculate the difference in concentration between the adjacent boxes, plug that into , and then transfer the molecules. But this calculation will take too long, even on a modern computer. So, we'd do better to cut down the number of boxes to something sensible, like 100. The distance is 10 m, so each box will represent 10 cm. Box 1 has buttery particles per cubic centimeter at the beginning. We'll assume that the popcorn keeps releasing buttery molecules and maintains box 1 at molecules per cubic centimeter. After 1 full second, you see the distribution of buttery particles looks like this:
You can see that the popcorn can be smelled only at about half a meter (charitably) after 1 second. Meanwhile, the sound took only about 20 milliseconds to reach all the way to the bedroom! After 10 s, it the distribution looks like I would say this was probably be distinguishable at about 1.5 meters. Finally, after 100 s (almost 2 minutes), it looks like this:
This can be smelled at about 5 meters, halfway to the bedroom. So roughly 5 minutes is totally reasonable! It takes a very long time for particles to diffuse in a gas.

So now let's turn to the sound. Once the microwave starts making noise, it's pushing the air molecules next to it. Those molecules are pushed into a smaller volume and create a local high-pressure system. They then collide with the ones near them along the same direction as they were pushed initially. The collision doesn't take very long at all, since they collide many many times a second. But what traveled down the hallway was not air. Nearly all of the gas stayed in the kitchen. What was moving was a propagating wave of air pressure. If you could look at the air while the sound was traveling, you'd see a pattern of high pressure and low pressure air in between the microwave and my ear. It was this pattern of compressions that traveled to my eardrum and vibrated it, which sent an electrical signal to my brain that there was noise. The equilibrium pressure is normal atmospheric pressure, 101 kPa (or 1 atmosphere, or 760 torr, or 14.7 PSI ...). It looks something like this:
This is what the wave looks like in real space (what you would see if you could see gas molecules). The wave shown could be propagating to the left or to the right. Where the wave has peaks, the pressure is high and the molecules are crowded together. Where the wave has troughs (low points), the pressure is low and the molecules are farther away from each other than normal. If the wave repeats itself in space, the distance between peaks is the wavelength. For this wave, the wavelength looks to be about 6 meters, which would be a 56 Hertz sound (this is close to the lowest A on a piano).

In the figure, you can see that there are imaginary boxes of high and low pressure. What if the boxes were so small that their length was only 1 mean free path? Then the molecules in the high pressure boxes would jump out spontaneously to the low pressure ones in a time , and the sound would not propagate, because those molecules didn't shove the ones in the low pressure box enough to keep a wave going. Thus, the possible wavelengths of sound are ones where the wavelength is much bigger than a mean free path. That's very interesting. What if I reduce the pressure in the room so that the mean free path gets bigger? What if I decrease the pressure until the mean free path is 1 cm. That happens at about 10 milli-torr (1.333 kPa). Then there is no sound! In a vacuum, such as outer space or the Moon, no sound can propagate. There is some gas there, but it's a much lower pressure than 1.333 kPa, so no audible sound will propagate.

Incidentally, I've way exaggerated the vertical scale on the graph. A difference of 1% from atmospheric pressure would be an ungodly loud noise. A difference in pressure of kPa is 60 decibels. That's not deafening, but it's not quiet either. That's incredibly small! A 60 dB noise corresponds to only 20 billionths of a percent change in pressure. The loudness, or intensity, of a sound is given in decibels in terms of the pressure change from atmosphere by


where is kPa. To give you a feel for that, a whisper is about 0 decibels, and 100 decibels is the threshold of pain. Dogs can bark louder than 100 dB! A normal rock concert is 80-90 decibels. Since it's a logarithmic scale, an increase of 10 decibels is 10 times the loudness.

We can try to guess what determines the speed of sound in a gas. This gets fairly technical.

Important!

The pressure at any point can be written in terms of atmospheric pressure and the extra pressure :


The pressure has got to depend somehow on the density, . We can call the function that relates them for now. So, , , and . In an ideal gas, of course, this is simple. Since the number density is related to the mass density


where is the molar mass and is the number of moles. Thus,


But that's ok, we'll just leave it for now.

We can define the wave function, , to be the displacement of a molecule at from where it was before the wave came about. If we look at a parcel of air that extends from to , then at time the left side of it has moved and the right side has moved to . But, crucially, the mass of the parcel stays the same. Since it originally had density and length , we can set the masses before and after equal to each other:


(The area is the same on both sides of the equation, and is therefore irrelevant.) Simplifying,


Let's put in


and neglect the final term as very small,

On the other hand, the second derivative of the wave function, , has got to be the acceleration of the parcel. For surely the only net motion of the particles is due to that. The pressure on either side of the parcel is on the left and on the right, so using Newton's second law,


or


(The last step has instead of because is constant.)

Finally, let us define such that


where the prime denotes derivative with respect to density. That being the case, we can also write in terms of pressure:

Now that we have all the pieces assembled, we can derive a differential equation for . First, eliminate pressure from II using III:


or


Then use I for :


or, finally,



The final equation above is called the wave equation. All classical waves---waves on strings, water waves, light, sound---obey this equation. The important part about it, though, is that the factor in front of the derivative is the speed squared. That is, the speed of sound is . Luckily, for an ideal gas we can find out what is.

Important!

For if you see (III) above, it's just the derivative of with respect to . Now, we cannot use the ideal gas law, because we don't know what will do under differentiation. But assuming that the sound propagates without any heat, we can use the relationship that is a constant. Since , we have


Thus, . Taking one derivative with respect to gives


Simple! Now, for an ideal gas, we can substitute for the value I gave way back at the beginning of this mess:


Thus,


Recall that is just the mass per mole.

We're done! According to this, the speed of sound in air is


For air, is about 1.4, and the molar mass is 28.97 grams per mole. Using K and J/K mol, we can get that the speed of sound is 347 m/s. This is pretty good. The actual value is closer to 340 m/s.

Note a couple things about this. First, it's off by 2%. You might be fine with that (I usually would be too), but we should always think about where we went wrong. Where we went wrong a little bit is that air does not really act like an ideal gas. The gas molecules are a bit stickier than an ideal gas is. The second thing to note is that sound moves slightly faster in warmer air, which is kind of interesting.

Finally, the speed of sound actually goes up for lighter air ( goes up if goes down). A cavity's resonant frequency goes up with the speed of sound, so if you breathe in helium, which is lighter than air per mole, you increase the speed of sound, and therefore you increase the resonant frequency of your windpipe. This is why helium makes you sound high pitched.

In the end, what determines the speed of sound in any medium, not just in air, is how pressure changes with density. That is, if you squish the mass into a smaller space by changing , how much does the pressure change? For a gas, the answer is not that much. You can squish it quite a bit without a giant change in pressure. So the speed of sound is about 340 m/s, which is really pretty slow. But if you squish a liquid down, its pressure goes up by quite a bit. So we expect the speed of sound in water to be faster, and indeed it is (about 5 times faster). And solids are the least squishable, and have an astounding speed of sound 20 times that of the speed in air. The vibration of an oncoming train on its tracks can be felt long before the sound of the train can be heard.

Dec 05 2012

'Skyfall' is Bond at its worst

One of the people who accompanied me to see the latest Bond film, Skyfall, had the misfortune that it was the only Bond movie she had seen. This is a shame, because Skyfall is Bond at its worst: offensive, misogynist, uninspired, bloated, and even corny, with cringe-inducing writing, a couple poor performances, and lackluster direction.

The film couldn't be more different than 2006's sublime Casino Royale, less of a sequel than a total deconstruction of the franchise. That film, directed by the otherwise undistinguished Martin Campbell not only introduced the best actor to portray the role (Daniel Craig, who is also good in the most recent film), but also treated the material as what it was: filled with cruel sadism, real tragedy, but also real tenderness. Vesper (Eva Green) wasn't a plot device, a walking vagina that made sure 007 was motivated onto his next scenic destination. She was a character, not something discarded so that the other characters could get back to shooting each other.

Contrast this with Skyfall's Sévérine (Bérénice Marlohe). Within the span of 20 minutes, she is introduced, given a marginal back story, screws Bond, and is unceremoniously killed. Her body is barely shown, nobody expresses the merest of reactions to her death, and she isn't mentioned again throughout the remaining (ponderous) length of the film. Director Sam Mendes, with his shots, shows us that she is of no import: we get no close shot of the body, only a distance shot and a medium reaction shot of Bond marginally disappointed that she's dead. At least Goldfinger had the decency to gild a whore before it killed and discarded her like garbage.

What's particularly disturbing about Skyfall is how much of a return to form the most recent film is to the terrible history of the franchise. In Casino Royale a particularly memorable and obvious scene takes place between Bond and a bartender shortly after Bond is defeated at poker:

Bond: Vodka-martini.
Bartender: Shaken or stirred?
James Bond: Do I look like I give a damn?

Between this deliberate bit of iconoclasm and the revealing shot of Craig's Bond emerging in a speedo from the ocean, we immediately know that the franchise understands its obsolescence and wants to atone. Gone are the days of posh Roger Moore gallivanting around the globe womanizing and cracking lame jokes. This is a serious story set in the real world, where gender politics exists and murder matters.

Instead, with Skyfall we get Bond fighting a giant mute asian stereotype in a komodo dragon pit, and using the dragon as a step-stool to escape the pit. He playfully innuendos with Moneypenny before going to see (the male) M to get his next assignment. He fucks two women that apparently neither he nor we care about. He grabs onto the bottom of an elevator for no discernible reason. He drives a classic Astin Martin with ejector seat and machine guns built in it. The villain has his own private giant island base. It almost reads as a parody if one didn't know better.

From a writing perspective the film is a bit of a mess. It opens with a failed operation to retrieve a stolen disk containing compromising information about many secret operatives embedded around the world. Bond is shot and presumed dead, and we are meant to believe that during his recovery he picks up a dependency on drugs and alcohol. MI6's headquarters is bombed (good thing that run down cabana gets CNN!), so he returns to help find the disk and stop whoever perpetrated the attack, and it's revealed that it's someone who is targeting M (Judi Dench, seemingly the only holdover from the Pierce Brosnan days). That someone turns out to be Silva (a terrible Javier Bardem), a comically stereotypical gay ex-agent who hates M for leaving him in enemy hands after he was captured several years earlier.

Bond tracks down the man who took the disk and kills him. A casino chip on his body leads him to a casino where he meets Sévérine. She tells him to meet her at her boat, and they ride to Silva's island base, where Bond calls in the military and apprehends him (oh, and he kills Sévérine, but who cares, right?). As it happens, this was Silva's plan all along, as his machine hacks MI6's system allowing him to escape and shoot up parliament. Bond takes M to his childhood home and tells Q to lead Silva to them, while Albert Finney mulls around as an anachronistic caretaker, seemingly there for comic relief. There is a standoff at the old house, and Silva is killed, but so is M.

It's troubling that Bond's drug abuse and alcoholism is a mere plot point; the film has no patience for the actual implications of such a terrible disease, so why bring them up? There is such a thing is functional alcoholism, but it's tragic and debilitating, and if the film was going to mention it so many times, it oughtn't be so easily tossed aside (at some point, Bond is just able to shoot straight again---that isn't how chemical dependency works).

Visually, it's just not a well-shot movie. Sam Mendes has chops: Road to Perdition was as hard-hitting as it comes, and American Beauty, though not on the same level, was also quite good. But the director shoots this movie like TV, and at some points it seemed like I was just watching a BBC drama on a big screen. He has indulgently long establishing shots of London rooftops, but isn't able to frame his characters in any meaningful way. The action scenes are competent at best, never reproducing anything like the visceral camerawork of John Woo or Paul Greengrass. By the time we reach the eponymous estate, one thinks of seeing Miss Marple in the background. The two principals stand in frame staring off into space, a composition representing nothing that I can fathom. As such, Mendes fails to establish any real connection between Bond and M, though he seems to think he is supposed to.

Skyfall has a few moments that could have held real promise. A far-off look by the woman Bond has shacked up with as she recognizes his damaging substance abuse. The trembling Sévérine describing her terrible plight, as desperate a tale as one can imagine. The withering career of an old cold war veteran, M, struggling with bureaucrats who think she's disposable. And maybe even Bond himself, dealing with his exploitation as an orphan by M. But none of these is explored in any depth whatsoever. It's a real missed opportunity, and a real shame.

Nov 20 2012

Thermo for Normals (part 26): A gas in detail

A molecule in a gas is being knocked into by lots of surrounding gas molecules. This is a very complex, almost random-seeming process. And yet we know that at higher temperatures the gas molecules simply must be moving faster on average. There must be a relationship between the temperature and the average energy. Moreover, we should be able to get some idea of how fast these molecules are moving.

The molecule is in a gas at temperature , and so its probability of having energy at some time is proportional to . But let's stop to think about what this means for a minute. The thing is, there are lots of ways to have a given energy . Suppose we choose eV. An oxygen molecule would have 1 eV of energy if it was traveling to the right at 2454 meters per second. But it would also have 1 eV if it were traveling to the left at 2454 m/s. If it could only move to the left and right, then, the probability of having 1 eV would have to be counted twice. That's in contrast to having zero energy. There's only one way to have zero energy. We would only count that once.

Once we let atoms start to move in 3 dimensions, now suddenly there are a lot of ways to have a certain energy. There are a bunch of different angles, all with the same speed, that would have energy . Even though the probability of having a certain energy is , we actually have to add up a bunch of the probabilities for every energy, since there are lots of ways to have a certain energy. That is, except for 0. At 0, there is only one way, no matter how many dimensions there are. So even though the probability initially looked large at , there's actually only a very small chance of having 0 energy.

Now, we can easily change energy into velocity, providing there's no potential energy. Energy is just one-half times mass times the velocity squared. A given particle can travel in all 3 directions, which we call , , and . Then it travels along the direction with velocity , along with , and along with . If , for instance, is positive, that means it moves to the right, and if is negative it's moving to the left. The energy is and the speed is . So now what we want to know is, what is the probability of having speed in a gas? (Note: you can read Wikipedia's entry on this if you want. Their derivation is bullshit.)

To make it simpler, we'll assume that velocity only comes in certain units (this is actually true). And we'll just form a grid and look at it. Here it is drawn in only two dimensions.
Each point represents a velocity the molecule can have. The one in the very center is and . You can see the tiny circle drawn around it. That circle represents zero speed. And it only holds one point, because, after all, there is only one way to have speed 0: the velocity in each dimension has to be zero. Now, the next circle going out from that is a bigger speed. All of the points enclosed by the next circle, but outside of the first one, are at a slightly higher speed. There are about 8 of them. The probability of having that energy is proportional to , and . But that probability applies to every point in there. So the probability of having speed one step above zero goes down because of , but goes up because we multiply by 8. Then we increase again; that's the third circle. The probability of having that speed is times the number of points in that circle (I count 36). The circles get a bigger radius, and so they form strips of more and more speed/energy every step.

But you might see what's happening. Even though the number of points is going up, the probability is going down every step.

So now all we have to do to form the probability of a certain speed is to count the number of points between one circle and the circle inside that, and multiply that by . If I do that for all the speeds (in three dimensions, even though the above graph only shows two), and then divide by the total, I get the probability as a function of :
So the probability of having no energy is not a maximum at zero. In fact, it's tiny! Then it rises to a maximum, and then it tails off to the right. At 300 K, the peak is at pretty low energy, and it tails off kind of quickly. At 900 K, though, the peak happens at a higher energy, and there are lots of molecules at higher energies. Duh! Of course, that's what we expect for a gas. If the temperature is higher more atoms are running around faster.

The graph shown above is called the Maxwell-Boltzmann distribution. It tells you, for a gas, how many of the gas molecules have a certain speed, based only on what kind of gas it is (the mass of the molecules) and the temperature. That is a hell of a lot of information from just those two things! Imagine trying to guess how many people made $25,000 per year just by knowing that the country is the United States and the GDP is 12 trillion dollars. How accurate do you think your guess would be?

If I'm going to want to get averages and stuff, we'll want an actual mathematical expression for the distribution, so I'll derive it here. As always, skip past it if you like.

Important!

The density of points we can call . The volume between concentric spheres is . Thus, the probability is


Let's absorb all constants into , since they don't really matter.


As always, we have to normalize this so that the probability of having some speed at all is 1. That also tells us what is. This is not an elementary integral, but it can be done with Feynman's trick of differentiating under the integral sign, and then using another handy trick for integrating a Gaussian. If we let , then


Now we integrate from 0 to and set that equal to 1:


We can't find the anti-derivative of the function , since it's not an elementary function. But we can do the definite integral


except we'll actually do .


Convert this to polar coordinates, so that and Then, goes from 0 to and goes from to , so


This integral is elementary; just make a substitution , so that . Then,


Thus, . Going back to our integral, we are only integrating from 0 to , so we take half of this:


Differentiating gives


so


and finally


Therefore the actual distribution, properly normalized, is

With this, now we can tell all the important statistics of a gas. How fast are the molecules moving on average? We can calculate that! I'll do it later, but the answer is

That's it. So, for argon gas at room temperature, the average speed is 398 meters per second, or about 890 miles per hour. Oxygen atoms in the air are traveling about 995 miles per hour. That is damn fast.

What about the average energy? For that we need to find the average of (which we usually write . A line over it means "average"). We can calculate that too. The answer is that the average of is . So the average energy of a gas molecule, so long as it only has kinetic energy, is


The average energy of a gas molecule in a monatomic gas is just three-halves . This will not work for diatomic gases like oxygen or certainly for something like steam. This is because diatomic or polyatomic gases have more ways of having an energy : they have ways to rotate, and the chemical bonds can vibrate. So being at a temperature would imply that there was more energy in the gas. We'll come back to that.

Now, a monatomic gas is made up of a lot of atoms, and not all of them have the same energy. But since they have on average, the internal energy of the gas is just the number of gas molecules times this, or


If you look at that, since we know that heat is added as energy, we see that if we want to increase the temperature by 1 K then we have to add of heat. That means that, by definition, the heat capacity of a monatomic gas is . Let's look at the heat capacities of 1 mole ( atoms) of some monatomic gases:

Heat capacities at constant volume for monatomic gases.

Gas (J/mol K)
He 12.5
Ar 12.5
Ne 12.7
Kr 12.3

And what do we think it should be?


Pretty damn good!

All of this came from one simple result, that (at least for one dimension), the probability that a molecule in a gas at temperature has energy is . Everything else was just figuring out how many ways there were to be at that energy, and this told us how the speeds are distributed. It's quite amazing.

Important!

I promised to calculate the average of and of , so here it is. The average value is the weighted sum, just like before. We multiply each speed by the probability of that speed and sum over all possible speeds.


That can be done pretty easily. Let's go back to , and we'll make a substitution , so that . Then


The latter integral can be done by parts:


Then,


For the average of , we take


We have to use Feynman's derivative trick again.


The integral is just . So,

Older posts «

» Newer posts