Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 4

Feynman diagram 1

This episode is supported by curiosity stream. Quantum Field Theory is stunningly successful at
describing the smaller scales of reality. But its equations are also stunningly complex. A lot of the genius
in qf T's development was in finding brilliant hacks to make these equations workable. The most famous
of these are the incredible Fineman diagrams

the equations of quantum field theory allow us to calculate the behavior of subatomic particles by
expressing them as vibrations in quantum fields. But even the most elegant and complete formulations
of quantum field theory, like the Dirac equation, or fireman's path integral, become impossibly
complicated when we try to use them on anything but the most simple systems that physicists tend to
interpret that's impossible as idea to try and try they did. First they express these impossible equations
in approximate but solvable forms. Then, they tackled the pesky infinities that kept appearing in these
new approximate equations. Finally, the entire mess was ordered into a system that mere humans could
deal with using the famous Fineman diagrams. To give you an idea of how messy quantum field theory
can be. Let's look at what should be a simple phenomenon. Electron scattering when two electrons repel
each other. In old fashioned classical electrodynamics, we think of each electron as producing an
electromagnetic field, that field then exerts a repulsive force on the other electron. At least in the
simplest cases, the coulomb equation governing this subatomic billiard shot is really easy to solve. But in
quantum field theory, specifically, quantum electrodynamics or QED, the story is very different. We
think of the electromagnetic field as existing everywhere in space, whether or not there's an electron
present vibrations in the EM field called photons, what we experience as light, the electron itself is just
an excitation of vibration in a different field, the electron field, and the electron and em fields are
connected. vibrations in one can cause vibrations in the other. This is how QED describes electron
scattering when electron excites a photon, and that photon delivers a bit of the first electrons
momentum to the second electron. It's arguable exactly how real that exchange photon is. In fact, we
call it a virtual photon. And it only exists long enough to communicate this force. There are other types
of virtual particle whose existence is similarly ambiguous. We'll get back to those in another episode.
This is a good time to introduce our first Fineman diagram. The brilliant Richard Fineman developed
these pictorial tools to organize the painful mathematics of quantum field theory. But they also serve to
give a general idea of what these interactions look like. In a Fineman diagram. One Direction is for time,
in this case, up the other axis represents space, although the actual distances aren't relevant. Here, we
see two electrons entering in the beginning and moving towards each other. They exchange a virtual
photon this squiggly line here, and the two electrons move apart at the end. But Fineman diagrams
aren't really just drawings of the interaction. They're actually equations in disguise. Each part of the
Fineman diagram represents a chunk of the math. Incoming lines are associated with the initial electron
states and outgoing lines represent the final electron states. The squiggle represents the quantized
fuelled excitation of the photon, and the connecting points, the vertices represent the absorption and
emission of the photon. The equation you string together from this one diagram represents all of the
ways that two electrons can deflect involving only a single virtual photon. And from that equation, it's
possible to perfectly calculate the effect of that simple exchange. And fortunately, real electron
scattering at a quantum level is a good deal more complicated than this. For that reason, this simple
calculation gives the wrong repulsive effect between two electrons. If we observe two electrons
bouncing off each other, all we really see is two electrons going in and two electrons going out. The
Quantum event around the scattering is a mystery. There are literally infinite ways that scattering could
have occurred. In fact, according to some interpretations, all infinite intermediate events that lead to
the same final result do happen, sort of. We talked about this weirdness when we discussed the
Fineman path integral recently. Just as with the path integral, to perfectly calculate the scattering of two
electrons, we need to add up all of the ways the electrons can be scattered. And this is where Fineman
diagrams start to come in handy, because they keep track of the different families of possibilities. For
example, the electrons might exchange just a single virtual photon. But they might also exchange two or
three or more free electrons might also emit and reabsorb virtual photon, or any of those photons might
do something crazy, like momentarily split into a virtual anti particle particle pair. Those last two events
are actually hugely complicating, as we'll see. With infinite possible interactions behind this one simple
process, a perfectly complete quantum field theoretic solution is impossible. But if you can't do
something perfectly, maybe near enough is good enough. This is the philosophy behind perturbation
theory, an absolutely essential tool to solving quantum field theory problems. The idea is that if the
correct equation is unsolvable, just find a similar equation that you can solve, then make small
modifications to it per term. So it's a bit closer to the equation that you want. It'll never be exact, but it
might get you pretty close. In the case of electron scattering, the most likely interaction is the exchange
of a single photon. Every other way to scatter the electrons contributes less to the probability of the
event. In fact, the more complicated the interaction, the less it contributes. Here, Fineman diagrams are
indispensable. It turns out the probability amplitude of a particular interaction depends on the number
of connections, or vertices in the diagram, every additional vertex in an interaction reduces its
contribution to the probability by a factor of around 100. So the most probable interaction for electron
scattering is the simple case of one photon exchange with its two vertices, a three vertex interaction
would contribute about 1% of the probability of the main to Vertex interaction. However, it turns out
that for electron scattering, there are no three vertex interactions. However, there are several
interactions that include four vertices, and each contributes about 1% of 1% of the to Vertex interaction.
And this is true, even though those complex interactions are very different to each other. They include
exchanging two virtual photons, or one electron emitting and reabsorbing a virtual photon, or the
exchange photon momentarily exciting a virtual electron positron pair, and more complicated
interactions add even less to the probability. So with Fineman diagrams, you very quickly get an idea of
which are the important additions to your equation, and which you can ignore perturbation theory with
the help of Fineman diagrams make the calculation possible. But that doesn't mean we're done.
Including all of these weird intermediate states really opens up a can of worms. This is especially true for
so called loop interactions, like when a photon momentarily becomes a virtual particle, anti-particle pair
and then reverts to a photon again, or when a single electron emits and reabsorbs the same photon.
This latter case can be thought of as the electron causing a constant disturbance in the EM field.
Electrons are constantly interacting with virtual photons. This impedes the motion of the electrons and
actually increases its effective mass. The effect is called self-energy. Trying to calculate the self-energy
correction to an electron mass, using quantum electrodynamics, the result is infinitely extra mass.

This sounds like a problem. To calculate the mass correction due to one of these self energy loops, you
need to add up all possible photon energies. But those energies can be arbitrarily large, sending the self
energy and hence the mass to infinity. In reality, something must limit the maximum energy of these
photons. We don't know what that limit is. The answer probably lies within a theory of quantum gravity,
which we don't yet have. But just as with perturbation theory, physicists found a cunning trick to get
around this mathematical inconvenience. It's called renormalization. Obviously, electrons do not have
infinite mass. And we know that because we've measured that mess. Although any measurement we
make actually includes some of this self energy. So our measurements are never of the fundamental or
The mass of the electron. And that is where the trick lies. Instead of trying to start with the
unmeasurable fundamental mass of the electron and solve the equations from there, you fold in a term
for the self energy corrected mass based on your measurement. In a sense, you capture the theoretical
infinite terms within an experimental finite number. This renormalization trick can be used to eliminate
many of the infinities that arise in quantum field theory. For example, the infinite shielding of electric
charge due to virtual particle anti particle peers popping into and out of existence. However, you pay a
price for renormalization. For every infinity you want to get rid of, you have to measure some property
in the lab. That means the theory can't predict that particular property from scratch, it can only make
predictions of other properties relative to your lab measurements. Nonetheless, renormalization saved
quantum field theory from this plague of infinities. Fineman diagrams successfully describe everything
from particle scattering, self energy interactions, matter antiemetic creation and annihilation to all sorts
of decay processes. We'll go further into the nuts and bolts of Fineman diagrams. In an upcoming
challenge episode, a set of relatively straightforward rules governs what diagrams are possible. And
these rules make Fineman doodles, an incredibly powerful tool for using quantum field theory. To
predict the behavior of the subatomic world. The results lead to the Standard Model of particle physics.
In future episodes,

we'll talk more about what is now the most complete description we have for the smaller scales of space
time. This episode is brought to you by curiosity stream, a subscription streaming service that offers
documentaries and nonfiction titles from some of the world's best filmmakers, including exclusive
originals. The two episode documentary, The Ultimate formula gives a really nice history of the
development of quantum field theory. get unlimited access today. And for our audience, the first two
months are free if you sign up at curiosity stream comm slash spacetime and use the promo code space
time during the signup process. As always, a huge thanks to our supporters on Patreon. This week, we'd
like to give an extra big thanks to Eugene Lawson, who is contributing at the Big Bang level. Eugene, it's
an incredible help and you make our jobs much easier. And another reminder to current or would be
Patreon patrons we made some spacetime eclipse glasses. Super handy for not going blind watching the
great American eclipse in August, we're sending out a set to every Patreon contributor at the $5 level or
above, and then include anyone who signs up at or increases to the $5 level in the month of July. At
least until we run out of glasses. It'll be first come first served. Last week we talked about Richard
Fineman his brilliant contribution to the development of quantum field theory with his path integral
formulation. You guys had a lot to say. Christian has asked how firemans path integral method which is
compatible with special relativity can derive Schrodinger equation when Schrodinger equation is not
compatible. So the deal is that trading is equation is a special case of a more general formulation of
quantum mechanics. In Schrodinger equation, all the particles are tracked according to one universal
master clock. In fireman's approach. Each particle is tracked according to its own proper time clock,
which can vary in its tick speed depending on how fast the particle is traveling. Derivation of Schrodinger
from Fineman requires approximating all of the separate proper time coordinates to give a single time
coordinate, that approximation is okay at low speeds that breaks when things get close to the speed of
light. psychische points out that the final probability for a particle journey is the square of the length of
the complex probability amplitude vector. And yeah, that's right. Probability is the square of the
probability amplitude. That's the Born rule right there. However, the sense I wanted to relay is that the
total probability depends on the length of the summed probability amplitudes, so the square of the real
plus the square of the complex components. So I also correctly points out that the individual paths don't
have different probability amplitude lengths taken separately, but rather they're pointing in the complex
vector space rotates so that each path ends differently to the total probability. Lowfat points out that
the wildly divergent paths would require superluminal speeds to reach their destination at the same
time. As the straight line paths Yeah, that's right. Fineman didn't limit particle velocity to the speed of
light by applying the principle of least action in the determination of probability amplitudes. It turns out
that the crazy paths, including the superluminal ones cancel out and add little to the probability

You might also like