Sie sind auf Seite 1von 116

There are few things more iconic of particle physics than Feynman diagrams.

These little
figures of squiggly show up prominently on particle physicists’ chalkboards alongside scribbled
equations. Here’s a ‘typical’ example from a previous post.

The simplicity of these diagrams has a certain aesthetic appeal, though as one might imagine
there are many layers of meaning behind them. The good news is that’s it’s really easy to
understand the first few layers and today you will learn how to draw your own Feynman
diagrams and interpret their physical meaning.

You do not need to know any fancy-schmancy math or physics to do this!

That’s right. I know a lot of people are intimidated by physics: don’t be! Today there will be no
equations, just non-threatening squiggly lines. Even school children can learn how to draw
Feynman diagrams (and, I hope, some cool science). Particle physics: fun for the whole

family.

For now, think of this as a game. You’ll need a piece of paper and a pen/pencil. The rules are as
follows (read these carefully):

1. You can draw two kinds of lines, a straight line with an arrow or a wiggly line:

You can draw these pointing in any direction.

2. You may only connect these lines if you have two lines with arrows meeting a single
wiggly line.

Note that the orientation of the arrows is important! You must have exactly one arrow
going into the vertex and exactly one arrow coming out.

3. Your diagram should only contain connected pieces. That is every line must connect to at
least one vertex. There shouldn’t be any disconnected part of the diagram.
In the image above the diagram on the left is allowed while the one on the right is not
since the top and bottom parts don’t connect.

4. What’s really important are the endpoints of each line, so we can get rid of excess curves.
You should treat each line as a shoelace and pull each line taut to make them nice and
neat. They should be as straight as possible. (But the wiggly line stays wiggly!)

That’s it! Those are the rules of the game. Any diagram you can draw that passes these rules is a
valid Feynman diagram. We will call this game QED. Take some time now to draw a few
diagrams. Beware of a few common pitfalls of diagrams that do not work (can you see why?):

After a while, you might notice a few patterns emerging. For example, you could count the
number of external lines (one free end) versus the number of internal lines (both ends attached to
a vertex).

 How are the number of external lines related to the number of internal lines and vertices?

 If I tell you the number of external lines with arrows point inward, can you tell me the
number of external lines with arrows pointing outward? Does a similar relation hole for
the number of external wiggly lines?

 If you keep following the arrowed lines, is it possible to end on some internal vertex?
 Did you consider diagrams that contain closed loops? If not, do your answers to the
above two questions change?

I won’t answer these questions for you, at least not in this post. Take some time to really play
with these diagrams. There’s a lot of intuition you can develop with this “QED” game. After a
while, you’ll have a pleasantly silly-looking piece of paper and you’ll be ready to move on to the
next discussion:

What does it all mean?

Now we get to some physics. Each line in rule (1) is called a particle. (Aha!) The vertex in rule
(2) is called an interaction. The rules above are an outline for a theory of particles and their
interactions. We called it QED, which is short for quantum electrodynamics. The lines with
arrows are matter particles (“fermions”). The wiggly line is a force particle (“boson”) which, in
this case, mediates electromagnetic interactions: it is the photon.

The diagrams tell a story about how a set of particles interact. We read the diagrams from left to
right, so if you have up-and-down lines you should shift them a little so they slant in either
direction. This left-to-right reading is important since it determines our interpretation of the
diagrams. Matter particles with arrows pointing from left to right are electrons. Matter particles
with arrows pointing in the other direction are positrons (antimatter!). In fact, you can think
about the arrow as pointing in the direction of the flow of electric charge. As a summary, we our
particle content is:

(e+ is a positron, e- is an electron, and the gamma is a photon… think of a gamma ray.)

From this we can make a few important remarks:

 The interaction with a photon shown above secretly includes information about the
conservation of electric charge: for every arrow coming in, there must be an arrow
coming out.

 But wait: we can also rotate the interaction so that it tells a different story. Here are a few
examples of the different ways one can interpret the single interaction (reading from left
to right):
These are to be interpreted as: (1) an electron emits a photon and keeps going, (2) a
positron absorbs a photon and keeps going, (3) an electron and positron annihilate into a
photon, (4) a photon spontaneously “pair produces” an electron and positron.

On the left side of a diagram we have “incoming particles,” these are the particles that are about
to crash into each other to do something interesting. For example, at the LHC these ‘incoming
particles’ are the quarks and gluons that live inside the accelerated protons. On the right side of a
diagram we have “outgoing particles,” these are the things which are detected after an interesting
interaction.

For the theory above, we can imagine an electron/positron collider like the the
old LEP and SLACfacilities. In these experiments an electron and positron collide and the
resulting outgoing particles are detected. In our simple QED theory, what kinds of “experimental
signatures” (outgoing particle configurations) could they measure? (e.g. is it possible to have a
signature of a single electron with two positrons? Are there constraints on how many photons
come out?)

So we see that the external lines correspond to incoming or outgoing particles. What about the
internal lines? These represent virtual particles that are never directly observed. They are created
quantum mechanically and disappear quantum mechanically, serving only the purpose of
allowing a given set of interactions to occur to allow the incoming particles to turn into the
outgoing particles. We’ll have a lot to say about these guys in future posts. Here’s an example
where we have a virtual photon mediating the interaction between an electron and a positron.

In the first diagram the electron and positron annihilate into a photon which then produces
another electron-positron pair. In the second diagram an electron tosses a photon to a nearby
positron (without ever touching the positron). This all meshes with the idea that force particles
are just weird quantum objects which mediate forces. However, our theory treats force and
matter particles on equal footing. We could draw diagrams where there are photons in the
external state and electrons are virtual:
This is a process where light (the photon) and an electron bounce off each other and is
called Compton scattering. Note, by the way, that I didn’t bother to slant the vertical virtual
particle in the second diagram. This is because it doesn’t matter whether we interpret it as a
virtual electron or a virtual positron: we can either say (1) that the electron emits a photon and
then scatters off of the incoming photon, or (2) we can say that the incoming photon pair
produced with the resulting positron annihilating with the electron to form an outgoing photon:

Anyway, this is the basic idea of Feynman diagrams. They allow us to write down what
interactions are possible. We will see later that in fact there is a much more mathematical
interpretation of these diagrams that produces the mathematical expressions that predict the
probability of these interactions to occur, and so there is actually some rather complicated
mathematics “under the hood.” However, just like a work of art, it’s perfectly acceptable to
appreciate these diagrams at face value as diagrams of particle interactions. In subsequent posts
we’ll develop more techniques and use this to talk about some really interesting physics, but until
then let me close with a quick “frequently asked questions”:

1. What is the significance of the x and y axes?


These are really spacetime diagrams that outline the “trajectory” of particles. By reading
these diagrams from left to right, we interpret the x axis as time. You can think of each
vertical slice as a moment in time. The y axis is roughly the space direction.

2. So are you telling me that the particles travel in straight lines?


No, but it’s easy to mistakenly believe this if you take the diagrams too seriously.
The path that particles take through actual space is determined not only by the
interactions (which are captured by Feynman diagrams), but the kinematics (which is
not). For example, one would still have to impose things like momentum and energy
conservation. The point of the Feynman diagram is to understand the interactions along a
particle’s path, not the actual trajectory of the particle in space.

3. Does this mean that positrons are just electrons moving backwards in time?
In the early days of quantum electrodynamics this seemed to be an idea that people liked
to say once in a while because it sounds neat. Diagrammatically (and in some sense
mathematically) one can take this interpretation, but it doesn’t really buy you anything.
Among other more technical reasons, this viewpoint is rather counterproductive because
the mathematical framework of quantum field theory is built upon the idea of causality.

4. What does it mean that a set of incoming particles and outgoing particles can have
multiple diagrams?
In the examples above of two-to-two scattering I showed two different diagrams that take
the in-state and produce the required out-state. In fact, there are an infinite set of such
diagrams. (Can you draw a few more?) Quantum mechanically, one has to sum over all
the different ways to get from the in state to the out state. This should sound familiar: it’s
just the usual sum over paths in the double slit experiment that we discussed before.
We’ll have plenty more to say about this, but the idea is that one has to add the
mathematical expressions associated with each diagram just like we had to sum numbers
associated with each path in the double slit experiment.

5. What is the significance of rules 3 and 4?


Rule 3 says that we’re only going to care about one particular chain of interactions. We
don’t care about additional particles which don’t interact or additional independent chains
of interactions. Rule 4 just makes the diagrams easier to read. Occasionally we’ll have to
draw curvy lines or even lines that “slide under” other lines.

6. Where do the rules come from?


The rules that we gave above (called Feynman rules) are essentially the definition of a
theory of particle physics. More completely, the rules should also include a few numbers
associated with the parameters of the theory (e.g. the masses of the particles, how
strongly they couple), but we won’t worry about these. Graduate students in particle
physics spent much of their first year learning how to carefully extract the diagrammatic
rules from mathematical expressions (and then how to use the diagrams to do more
math), but the physical content of the theory is most intuitively understood by looking at
the diagrams directly and ignoring the math. If you’re really curious, the expression from
which one obtains the rules looks something like this (from TD Gutierrez), though that’s
a deliberately “scary-looking” formulation.

We’ll develop more intuition about these diagrams and eventually get to some LHC physics, but
hopefully this will get the ball rolling!
More Feynman Diagrams

In a previous post we learned how to draw Feynman diagrams by drawing lines and connecting
them. We started with a set of rules for how one could draw diagrams:

We could draw lines with arrows or wiggly lines and we were only permitted to join them using
intersections (vertices) of the above form. These are the rules of the game. We then said that the
arrowed lines are electrons (if the arrow goes from left to right) and positrons (if the arrow
points in the opposite direction) while the wiggly lines are photons. The choice of rules is what
we call a “model of particle interactions,” and in particular we developed what is
called quantum electrodynamics, which is physics-talk for “the theory of electrons and
photons.”

Where did it all come from?

One question you could ask now is: “Where did these rules come from? Why do they prohibit
me from drawing diagrams with three wiggly lines intersecting?”

The short answer is that those are just the rules that we chose. Technically they came from a
more mathematical formulation of the theory. It is not obvious at all, but the reason why we only
allow that one particular vertex is that it is the only interaction that both respects the
(1) spacetime (“Lorentz”) symmetry and (2) internal ‘gauge’ symmetry of the theory. This is an
unsatisfying answer, but we’ll gradually build up more complicated theories that should help
shed some light on this. Just for fun, here’s the mathematical expression that encodes the same
information as the Feynman rules above: [caution: I know this is an equation, but do not be
scared!]

Without going into details, the represents the electron (the bar turns it into a positron) while
the A is the photon. The number e is the ‘electric coupling’ and determines the charge of the
electron. Because equations can be intimidating, we won’t worry about them here. In fact our
goal will be to go in the opposite direction: we will see that we can learn quite a lot
by only looking at Feynman diagrams and never doing any complicated math. The important
point is that our cute rules for how to connect lines really captures most of the physics encoded
in these ugly equations.
Now a quick parenthetical note because I’m sure some of you are curious: In the equation above,
the is a kind of derivative. Derivatives tell us about how things change, and in fact this
term tells us about how the electron propagates through space. The e?A term tells us how the
photon couples to the electron. The m term is the electron’s mass. We’ll have more to say about
this down the road when we discuss the Higgs boson. Finally, the Fs are the “field strength” of
the photon: it is the analog of the derivative term for the electron and tells us how the photon
propagates through space. In fact, these F’s encode the electric and magnetic fields.

[Extra credit for advanced readers: notice that the electron mass term looks like the Feynman
rule for a two-electron interaction with coupling strength m. You can see this by looking at the
electron-electron-photon term and removing the photon.]

What we can learn from just looking at the rules

We learned that we could use our lines and intersections to draw diagrams that represent particle
interactions. If you haven’t already, I encourage you to grab a piece of scratch paper and play
with these Feynman rules. A good game to play is asking yourself whether a certain initial state
can ever give you a certain final state. Here are a few exercises:

1. You start with one electron. Can you ever end up with a final state positron? [Answer:
yes! Draw one such diagram.]

2. If you start with one electron, can you ever end up with more final state positrons than
final state electrons? [Answer: no! Draw diagrams until you’re convinced it’s
impossible.]

3. Draw a diagram where an electron and a photon interact to produce 3 electrons, 2


positrons, and 2 photons. Draw a few more to get a feel for how many different ways one
can do this.

4. If you start with a photon, can you end up with a final state of only multiple photons?
[This is actually a trick question; the answer is no but this is a rather subtle quantum
mechanical effect that’s beyond our scope. You should be able to draw a diagram think
that the answer is ‘yes.’]

So here’s what you should get out of this: Feynman rules are a nice way to learn what kinds of
particle interactions can and cannot occur. (e.g. questions 1 and 2) In fact, the lesson you should
have gleaned is that there is a conservation of electric charge in each diagram coming to the
conservation of electric charge in each intersection. You can also see how complicated
interactions can be reduced to simple interactions with “virtual particles” (intermediate particles
that don’t appear in the initial state). We are able to do this simply by stating the Feynman rules
of our theory and playing with drawings. No math or fancy technical background required.

Summing diagrams: an analogy to summing paths


There’s a lot more one could do with Feynman diagrams, such as calculating probabilities for
interactions to occur. Actually doing this requires more formal math and physics background, but
there’s still a lot that we can learn conceptually.

For example, there were two simple diagram that we could draw that represented the scattering
of an electron and a positron off of one another:

We recall that we can describe these interactions in words by “reading” them from left to right:

 The first diagram shows an electron and a positron annihilating into a photon, which then
“pair produces” into another electron and positron.

 The second diagram shows an electron and a positron interacting by sending a photon
between them. This is definitely a different process since the electron and positron never
actually touch, unlike the first diagram.

Remember that these diagrams are actually shorthand for complex numbers. The numbers
represent the probability for each these processes to occur. In order to calculate
the full probability that an electron and a positron will bounce off of one another, we have to add
together these contributions as complex numbers.

What does this mean? This is just quantum mechanics at work! Recall another old post about the
double slit experiment. We learned that quantum mechanics tells us that objects take all paths
between an initial observed state to a final observed state. Thus if you see a particle at point A,
the probability for it to show up at point B is given by the sum of the probability amplitudes for
each intermediate path.

The sum of diagrams above is a generalization of the exact same idea. Our initial observed state
is an electron and a positron. Each of these have some fixed [and observed] momentum. If you
want to calculate the probability that these would interact and produce an electron and positron
of some other momentum (e.g. they bounce off each other and head off in opposite directions),
then one not only has to sum over the different intermediate paths, but also the different
intermediate interactions.
Again, a pause for the big picture: we’re not actually going to calculate anything since for most
people, this isn’t as fun as drawing diagrams. But even just describing what one would calculate,
we can see how things reduce to our simple picture of quantum mechanics: the double slit
experiment.

Momentum Conservation

Each initial and final state particle has a well-defined momentum. (By ‘momentum’ I also
include the particle’s total energy.) As one could guess, any physical diagram must satisfy the
conservation of momentum. In fact, this is built into each intersection: we assume that the sum of
the momentum going into each intersection (i.e. from the left) is equal to the momentum going
out of it (to the right). Thus you cannot have two very low energy initial state electrons scattering
into something with a very high energy final state.

Perhaps more obviously, this means that you cannot have diagrams where “nothing” turns into
stuff, or something turns into nothing:

One will note that both of these diagrams are technically allowed by our diagrammatic Feynman
rules. We thus have to impose momentum conservation as an additional Feynman rule.

Here’s an exercise for slightly more advanced readers who know special relativity: convince
yourself that momentum conservation prohibits any diagrams that only contain a single
interaction, e.g.

A hint: consider going into the rest frame of the particles.

[A much simpler exercise for everyone: “read” each of these diagrams from left to right and
describe what’s going on in words. Even though these are all variations of the rule for
intersecting lines, how do these three diagrams differ in physical interpretation?]

It is straightforward to see that in the electron-positron scattering diagrams above, the


momentum of the intermediate photon is fully determined by the momenta of the external
particles. For example, in the first diagram the photon momentum must be the sum of the initial
particle momenta. (An an exercise for advanced readers again: convince yourself that the
intermediate photon is not ‘on shell’, i.e. the square of its 4-momentum doesn’t equal zero. This
is okay because the photon is a virtual particle.)

Loop diagrams: a prelude for things to come

Now I’d like to pause to mention an ‘advanced topic’ that we’ll get to in a future post. If you’ve
been diligent and have played with drawing different kinds of Feynman diagrams, you’ll have
noticed that you can also draw diagrams that have closed loops, such as:

We call such graphs loop diagrams for obvious reasons. Diagrams without loops are called tree
diagrams. It turns out that loop diagrams are rather special and introduce a few ‘deep’ issues
that I’ll only mention in passing for now: (some of these are a bit ‘advanced’, don’t worry if
they’re a little vague — we’ll come back to them later)

 The above diagram is a contribution to the electron-positron scattering process that we


considered above. You should be able to convince yourself that there are in fact
an infinite number of contributions for each interaction between a given final and initial
state particles given by drawing more loops in creative ways. This sounds weird, but
remember that there were also an infinite number of paths between any two points when
we studied the “infinite-slit” experiment.

 For those of with some calculus background: what we’re actually doing is a Taylor
expansion. What is our expansion parameter? The electromagnetic coupling e (in the
equation we wrote above). In other words, we are expanding in the number of vertices.
Each vertex gives a factor of e (which is a small number), so that the full result is actually
very well approximated by only taking into account tree diagrams.

 In light of our comment about momentum conservation, you should convince yourself
that the “loop” particles (which are completely virtual) can have any arbitrarily large
momentum. This is in contrast to intermediate particles in tree diagrams whose
momentum is constrained by the external momenta. This is actually rather interesting:
this means that the loops are sensitive to physics at higher energy scales.

 In light of our understanding of quantum mechanics, we see that even for a single loop
diagram we have to sum over an infinite number of possible loop momenta. This can be a
problem: a sum over an infinite set of numbers can itself be infinite. Thus we worry that
the calculation of our quantum mechanical probabilities may end up giving nonsensical
results (what does it mean for a probability to be infinite?). In fact, it turns out that this is
very deeply related to the idea that the loops are sensitive to physics at higher energies.
We will discuss all of this more thoroughly when we get to the Higgs boson.

Loop diagrams can be very tedious to calculate. In fact, they’re the bane of most graduate
students’ lives when they first learn quantum field theory. Fortunately we’re not going to
calculate any of them and, for now, will just marvel at their ability to complicate things.

Action-at-a-distance: Attractive and Repulsive Forces

Our discussion has become a little technical, so let’s take a step back and see how some of these
pieces come together. We recall that photons are not only particle of light, but they are the
intermediate ‘force particles’ which mediate the electric force between electrons (and positrons).
The cartoon picture of these force particles is that charged particles “toss” photons back and
forth when they interact. One can imagine, as pictured in the Particle Adventure, two electrons as
basketball players tossing a ball between them while standing on ice. The momentum of the ball
being tossed back and forth translates into a motion of the particles away from each other.

This always begs the question: how the heck are we supposed to understand forces that
causes particles to attract? (For example, an electron and a positron.) The cartoon picture
doesn’t make sense anymore!

There are a few ways to answer this question. First of all, “forces” are classical descriptions
of quantum phenomena. In order to properly derive a force, one should find a way to construct
the potential energy of a system and see what kind of particle motion causes it to decrease. There
is a way to do this from the quantum perspective, and it turns out to give exactly the correct
behavior. For those with some background in quantum mechanics, I would suggest the first 30 or
so pages of Zee’s textbook.

However, I promised you no calculations. So let me try to motivate this more heuristically. We
should recall that classically, in the presence of a force field, momentum is not conserved (the
force causes acceleration). Our Feynman rules, however, explicitly require momentum
conservation. So we could cast our question in a different way: how can our quantum theory
gives us any kind of force?

To go from a quantum (virtual particle) picture to a classical (force) picture, we have to


somehow include the effect of many quantum particles to generate a macroscopic phenomenon.
What actually happens when an electron and a positron are attracted to each other over long
distances is something like this:
This almost looks like one of the diagrams we considered above, except now there are several
final state photons. The electron and the positron move towards each other by shedding photons
off into space. This is precisely what a stranded astronaut would do to get back to her space
shuttle: throw a wrench in the opposite direction and let conservation of momentum do its job.

Now you say: “That’s crazy! My physics textbook says that oppositely charged particles attract
and that’s it — there’s no mention of a bunch of extra photons.” Well, your physics textbook
also says that there is something else: the electromagnetic field. The extra photons in the
quantum mechanical picture precisely set up the electromagnetic field in the classical picture!
This sounds weird, but this is what we mean when we say that the photon is the quantum carrier
of the electromagnetic force: it is the quantum of the electric field.

Now I have dodged the question about how a force ‘knows’ whether it should be attractive or
repulsive. Thus far I’ve only explained why this could happen. The question now is how do the
photons know to be emitted in such a way that the particles attract or repel? I don’t have simple
explanation for this; the most straightforward way to determine this from first principles is to
actually do the calculation — which we’ve promised not to do. What we will do is motivate why
electron-electron scattering (repulsive force) should be different from electron-positron
scattering (attractive force). The simplest way to note that these two process should behave
differently is that they are described by different Feynman diagrams!

Electron-positron scattering is mediated by the diagrams we discussed above:

Whereas electron-electron scattering is described by a different pair of graphs:


To be certain they look similar, especially the second diagrams, the actual calculation (which we
won’t do!) gives different results. To get the classical behavior one has to include the emission of
photons that become the electromagnetic field. The key result is that the quantum probability
amplitude for electron-electron scattering prefers to emit photons in such a way that the particles
repel, while the quantum probability amplitude for electron-positron scattering prefers them to
attract.

Next time…

QED + μ: introducing the muon

It’s time to return to our ongoing exploration of the structure of the Standard Model. Our primary
tools are Feynman diagrams, which we introduced in previous posts (part 1, part 2). By now
we’ve already familiarized ourselves with quantum electrodynamics (QED): the theory of
electrons, positrons, and photons. Now we’re going to start adding on pieces to build up the
Standard Model. We’ll start with the muon, portrayed below by Los Angeles artist Julie Peasley.
(These handmade plushes can be found at her website, The Particle Zoo.)
We’re all familiar with the electron. Allow me to introduce its heavier cousin, the muon (μ).
Where did this muon come from? Or, as Nobel Prize winner I. I. Rabi once asked, “Who ordered
that?” (This is still an unanswered question!) Besides its mass, the muon has the same
fundamental properties as the electron: it has the same charge, feels the same forces, and—like
the electron—has an anti-particle partner.

Feynman rules for QED+μ

This makes it really easy to extend our Feynman rules. We’ll call our theory “QED+μ,” quantum
electrodynamics with an additional particle. We just have to write the rules for two copies of
QED:

Let’s recall how to interpret this. The three lines tell us that we have three kinds or particles in
the theory: electrons (e), muons (μ), and photons (γ). Recall that the matter particles, the ones
whose lines have an arrow, also have antiparticles. We indicate antiparticles by arrows pointing
in the wrong direction when we read the diagrams from left-to-right. The vertex rules tell us that
we have two kinds of interactions: a photon can either interact with two electrons or two muons.
It’s important to note that we cannot have photon couplings that mix electrons and muons. In
terms of conservation laws, we say that electron and muon number are each conserved. For
example, in the theory we’ve developed so far, you cannot have a muon decay into an electron
and a photon. (We’ll introduce these sorts of interactions next time when we discuss electroweak
theory.)

Exercise: Is the following diagram allowed in QED + μ?

Answer: Yes! But doesn’t this violate conservation of electron and muon number? You start out
with two e‘s on the left and end up with two μ’s. Hint: what are the arrows telling you?

Once you’ve convinced yourself that the above diagram doesn’t violate electron or muon
conservation, let me remark that this is an easy way to produce muons at low energy electron
colliders. You just smash an electron against a positron and sometimes you’ll end up with a
muon-antimuon pair which you can detect experimentally.

Exercise: when we previously did electron-positron to electron-positron scattering, we had to


include two diagrams. Why is there only one diagram for eμ to eμ? Hint: draw the two diagrams
for ee to ee and check if the Feynman rules still allow both diagrams if we convert the final
states to muons.

Detecting muons, some collider physics

If you think about this a little, you might wonder: if electrons and muons are so similar, how can
experimentalists distinguish between them at a collider? Seth and Mike might scold me for
skipping over some information about the interaction of charged particles through matter, but
one simple way to distinguish muons from electrons is to measure their energy and momenta.
We know that (away from a potential) a particle’s energy is the sum of its kinetic energy plus it’s
mass energy added in quadrature E2=m2c4+p2c2 (this is the “real” version of E=mc2). Since
muons are heavier than electrons, we can just check the mass of the particle by plugging in the
measured energy and momentum.

Actually, this is an oversimplified picture. In order not to annoy the other US/LHC bloggers, I’d
better provide a slightly less oversimplified “cartoon.” Electrons are light, so let’s imagine that
they’re ping pong balls. On the other hand, muons are heavy, so let’s imagine them as bowling
balls. As you probably know, the LHC detectors are big and full of stuff… by that I mean atoms,
which in turn are made up of a nucleus and a cloud of electrons. We can thus imagine a sea of
ping-pong balls (think of a Chuck-E-Cheese ball pit). When electrons hit this ball pit, they end up
distributing all of their energy into the other balls. This happens in the electromagnetic
calorimeter, or ECAL. “Calor” is Latin for heat, so you can guess that the ECAL is really just a
big fancy thermometer that measures the energy that the electron dissipates. Muons on the other
hand, are bowling balls that are so massive that they just barrel straight through the ball pit to get
to the other side. Here’s a very scientific illustration:

I hope we don’t get any comments saying, “oh man, muons are jerks.” In fact, they’re quite the
opposite: muons are the only Standard Model particles that make it all the way to the outside of
the detector, making it easy for us to identify them. In fact, the big distinctive toroidal magnets
on the ATLAS detector below are there to bend the path of muons to help the outermost detector
determine the muon momentum by measuring the curvature of their trail.

Exercise: [for those who want to do some actual calculations, requires a high school physics
background] Convince yourself that this heuristic picture is correct by calculating the final
momenta of a ball colliding elastically with (a) a ball of the same mass and (b) a ball of much
lighter mass.
ATLAS toroidal magnets. Image from the Interactions.org Image Bank

Neat things that muons can do

Let me make a few more semi-historical remarks: our QED+μ model is just a theoretical toy.
Historically, scientists knew immediately that something was weird about the muon: unlike
electrons, it decayed into other particles and seemed to interact with mesons in unusual ways. In
fact, for a while people thought that muons were a kind of meson. These differences ended up
being a harbinger of something more interesting: the weak force.

Exercise: convince yourself that our Feynman rules for QED+μ do not allow muon decay, i.e. μ
turning into non-μ stuff.

Muons are generated in the sky when cosmic rays hit atoms of the upper atmosphere. These rain
down onto the Earth and force us to put our dark matter experiments deep underground to avoid
their ‘noise.’ What’s really neat, however, is that the fact that muons make it to the surface of the
Earth is a rousing experimental check of relativity. We know that muons at rest decay in
microseconds. In this time, it seems like there’s no way for them to traverse the kilometers (about
4 km) between the Earth and its upper atmosphere; even if they were traveling at the speed of
light! (c ~ 3. 108 m/s). What’s happening is the phenomenon of time dilation!

Introducing the tau (via the Socratic method)

Exercise: the Standard Model actually has another cousin of the electron, the tau (τ), leading to
three charged leptons in total. Write down the Feynman rules for the theory QED+μ+τ, i.e. the
theory of electrons, muons, and taus interacting via photons. Make sure that electron, muon, and
tau number are all conserved. Draw the diagram for tau production in an electron-positron
collider.
Exercise: Above we argued that muons are special because they barrel right through our
detectors like bowling balls through an array of ping pong balls. Taus are even heavier, shouldn’t
they also make it to the outside of the detector?

Answer: This was a bit of a trick question. The logic is correct that sufficiently energetic taus
should make it all the way to the outside of the detector in our QED+μ+τ theory. However, this
is not the full story for electrons, muons, and taus (collectively known as leptons) in the
Standard Model. Like muons, taus are unstable and will decay. In fact, they decay much more
quickly than muons because they have more mass and can decay into stuff (they have more
“phase space”). While muons are like bowling balls barreling through the detector, taus are more
like grenades that burst into hadronic “shrapnel” inside the calorimeters. They are usually very
difficult to reconstruct from the data.

A preview of things to come:

Now we’re very familiar with putting together multiple copies of QED. For now, there are only
three copies we have to worry about. It is an open question why this is the case. The existence of
at least three copies, however, turns out to be significant for the imbalance of matter and anti-
matter in the universe. In the next post we’ll introduce the weak force and really see what we
can do with these leptons.

I’m currently in the middle of my “Advancement to Candidacy” exam, so my posts might be a


little more delayed than usual this month. By the end of it, however, I hope to be blogging as an

official PhD candidate.

Erratum: virtual particles

I wanted to correct a misleading statement I made in my previous QED post: I discussed the
visualization of virtual particles as balls that two kids toss back and forth while standing on
frictionless ice. Conservation of momentum causes the two kids to slide apart as they throw and
catch the ball, generating what we observe macroscopically as a repulsive force. We mentioned
that it’s more difficult to see how this could give rise to an attractive force. I suggested that this
is a phenomenon coming from the accumulated effect of many quantum exchanges. While this is
true, there is a simpler way to understand this: pretend the ball has negative momentum! Since
the particle is virtual, it is inherently quantum mechanical and needn’t have ‘on-shell’ (physical)
momentum. Thus one could imagine tossing the ball with negative momentum, causing one to be
deflected in the same direction as the ball was tossed. Similarly, catching the ball with negative
momentum would push one in the direction that the ball came from.

The Z boson and resonances


Hello everyone! Let’s continue our ongoing investigation of the particles and interactions of the
Standard Model. For those that are just joining us or have forgotten, the previous installments of
our adventure can be found at the following links: Part 1, Part 2, Part 3.

Up to this point we’ve familiarized ourselves the Feynman rules—which are shorthand for
particle content and interactions—for the theory of electrons and photons (quantum
electrodynamics, or QED). We then saw how the rules changed if we added another electron-like
particle, the muon ?. The theory looked very similar: it was just two copies of QED, except
sometimes a a high-energy electron and positron collision could produce a muon and anti-muon
pair. At the end of the last post we also thought about what would happen if we added a third
copy of electrons.

Let’s make another seemingly innocuous generalization: instead of adding more matter particles,
let’s add another force particle. In fact, let’s add the simplest new force particle we could think
of: a heavy version of the photon. This particular particle is called the Z boson. Here’s a plush
rendition made by The Particle Zoo:

Feynman rules for QED+?+Z

Our particle content now includes electrons, muons, photons, and Z bosons. We draw their lines
as follows:
Recall that anti-electrons (positrons) and anti-muons are represented by arrows pointing in the
opposite direction.

Question: What about anti-photons and anti-Z bosons?


Answer: Photons and Z bosons don’t have any charge and turn out to be their own anti-particles.
(This is usually true of force particles, but we will see later that the W bosons, cousins of the Z,
have electric charge.)

The theory isn’t interesting until we explain how these particles interact with each other. We thus
make the straightforward generalization from QED and allow the Z to have the same interactions
as the photon:

What I mean by this is that the squiggly line can either be a photon or a Z. Thus we see that we
have the following four possible vertices:

1. two electrons and a photon

2. two electrons and a Z

3. two muons and a photon

4. two muons and a Z

Question: What are the conservation laws of this theory?


Answer: The conservation laws are exactly the same as in QED+?: conservation of electron
number (# electrons – # positrons) and conservation of muon number (#muons – #anti-muons).
Thus the total electron number and muon number coming out of a diagram must be the same as
was going into it. This is because the new interactions we introduced also preserve these
numbers, so we haven’t broken any of the symmetries of our previous theory. (We will see that
the W boson breaks these conservation laws!) We also have the usual conservation laws: energy,
momentum, angular momentum.

Resonances

So far this seems like a familiar story. However, our theory now has enough structure to teach us
something important about the kind of physics done at colliders like the LHC. We started out by
saying that the Z boson is heavy, roughly 91 GeV. This is almost a hundred times heavier than a
muon (and 20,000 times heavier than an electron). From our Feynman rules above we can see
that the Z is unstable: it will decay into two electrons or two muons via its fundamental
interactions.

Question: The photon has the same interactions as the Z, why isn’t it unstable?
[Hint: kinematics! Namely, energy conservation.]

In fact, because electrons and muons are so much lighter, the Z is very happy to decay quickly
into them. It turns out that the Z decays so quickly that we don’t have any chance of detecting
them directly! We can only hope to look for traces of the Z in its decay products. In particular,
let’s consider the following process: an electron positron pair annihilate into a Z, which then
decays into a muon anti-muon pair.

The Z boson here is virtual—it only exists quantum mechanically and is never directly
measured. In fact, because it is virtual this process occurs even when the electrons are not
energetic enough to produce a physical Z boson, via E=mc2. However, it turns out that something
very special when the electrons have just enough energy to produce a physical Z: the process
goes “on shell” and is greatly enhanced! The reason for this is that the expression for the
quantum mechanical rate includes terms that look like (this should be taken as a fact which we
will not prove):

where M is the mass of the Z boson, p is essentially the net energy of the two electrons, and ? is a
small number (the ‘decay width of the Z‘). When the electrons have just enough energy, p2–M2 =
0 and so the fraction looks like i/?. For a small ?, this is a big factor and the rate for this diagram
dominates over all other diagrams with the same initial and final states. Recall that quantum
mechanics tells us that we have to sum all such diagrams; now we see that only the diagram with
an intermediate Z is relevant in this regime.

Question: What other diagrams contribute? Related question: why did we choose this particular
process to demonstrate this phenomenon?
Answer: The other diagram that contributes is the same as above but with the Z replaced by a
photon. There are two reasons why we needed to consider ee ? Z ? ??. First, an intermediate
photon would have M = 0, so p2–M2 will never vanish and we’d never hit the resonance (recall
that the electrons have energy tied up in their mass, so p ? 2m where m is the electron mass).
Second, we consider a muon final state because this way we don’t have to consider background
from, for example:

These are called t-channel diagrams and do not


have a big enhancement; these diagrams never have a time slice (we read time from left-to-right)
where only a Z exists. (For the record, the diagrams which do get enhanced at p2–M2 = 0 are
called s-channel for no particularly good reason.)

Intuitively, what’s happening is that the electrons are resonating with the Z boson field: they’re
“tickling” the Z boson potential in just the right way to make it want to spit out a particle.
Resonance is a very common idea in physics: my favorite example is a microwave—the
electromagnetic waves resonate with the electric dipole moment of water molecules.

Detecting the Z boson

This idea of resonance gives us a simple handle to detect the Z boson even if it decays before it
can reach our detectors. Let’s consider an electron-positron collider. We can control the initial
energy of the electron-positron collision (p in the expression above). If we scan over a range of
initial energies and keep track of the total rate of ?? final states, then we should notice a big
increase when we hit the resonance. In fact, things are even better since the position of the
resonance tells us the mass of the Z.

Below is a plot of the resonance from the LEP collaboration (Fig 1.2 from hep-ex/0509008):
Different patches of points correspond to different experiments. The x-axis is the collision energy
(what we called p), while the y-axis is the rate at which the particular final states were observed.
(Instead of ee? ?? this particular plot shows ee ? hadrons, but the idea is exactly the same.) A
nice, brief historical discussion of the Z discovery can be found in the August ’08 issue
of Symmetry Magazine, which includes the following reproduction of James Rohlf’s hand-
drawn plot of the first four Z boson candidate events:

[When is the last time any of the US LHC bloggers plotted data by hand?]

In fact, one way to search for new physics at the LHC is to do this simple bump hunting: as we
scan over energies, we keep an eye out for resonances that we didn’t expect. The location of the
bump tells us the mass of the intermediate particle. This, unfortunately, though we’ve accurately
described the ‘big idea,’ it is somewhat of a simplified story. In the case of the electron-positron
collider, there are some effects from initial- and final-state radiation that smear out the actual
energy fed into the Z boson. In the case of the LHC the things that actually collide aren’t actually
the protons, but rather the quarks and gluons that make up the protons—and the fraction of the
total proton energy that goes into each colliding object is actually unknown. This is what is
usually meant when people say that “hadron colliders are messy.” It turns out that one can turn
this on its head and use it to our advantage; we’ll get to this story eventually.

Neutrinos

Time for another dose of particles for the people (eh, working title). In previous installments
(Part 1, Part 2, Part 3, Part 4) we started a basic theory (QED; electrons and photons) and added
on muons, taus, and the Z boson. Now we’re going to add on a set of particles that have recently
made some news, the neutrino.

Here’s the Particle Zoo‘s depiction of an electron-neutrino:

There are, in fact, three types of neutrino: one to pair with each of our electron-like particles.
Thus in addition to the electron-neutrino, we also have the muon-neutrino and the tau-neutrino.
As their name suggests, neutrinos are neutral and have no electric charge. Further, they’re
extremely light.

The fact that neutrinos don’t have any charge means that they don’t couple to photons, i.e. there
are no Feynman rules for neutrinos to interact with photons. In fact, the only particle we’ve met
so far that does interact with the neutrino is the Z boson, with the following Feynman rules:
Exercise: Consider a theory with only neutrinos and Z bosons so that we only have the Feynman
rules above. Check that this looks just like three copies of QED (a theory of electrons and
photons).

Question: How is the theory of only neutrinos and Z bosons different from “three copies of
QED?”
Answer: Unlike the photon, the Z boson has mass! This means that the Z boson doesn’t produce
a long-range force like electromagnetism. We’ll discuss this soon when we introduce
the W boson and explain that the W and Z together mediate the so-called “weak nuclear force.”

Exercise: Draw the Feynman diagrams for an electron and positron annihilating into a neutrino
and anti-neutrino. What are the possible final states? (e.g. can you have a muon-neutrino and
anti-muon neutrino? Can you have an electron neutrino and an anti-tau neutrino?) Given that
neutrinos don’t interact electrically and that the Z boson interacts very weakly, what do you think
this would look like in a particle detector? (Consider the significance of the phrase “missing
energy.”)

This should start to sound very boring!

If you’re starting to get bored because we keep writing down the same QED-like theory, then
you’re keeping up. So far we’ve introduced all of the basic players in the game, but we haven’t
told them how to interact with each other in exciting ways: don’t worry! We’ll get to this in the
next post on the W boson.

Let’s recap how boring we have been:

 We started with a theory of electrons and photons called QED.

 We then “doubled” the theory by adding muons which were heavier electrons that
coupled in the same way to photons. Then we “tripled” the theory by adding taus, which
are yet another heavy copy of electrons.

 Next we added a new force particle, the Z. This is a heavy version of the photon (with a
weaker interaction strength), but otherwise our Feynman rules again seemed like a
doubling of the rules in the previous step. (We now have 6 “copies” of QED.)

 Now we’ve added three neutrinos, which only interact with the Z in a way that looks just
like QED. We now have 9 “copies” of QED.
I promise things will get a lot more exciting very soon. First, here’s a pop quiz to make sure
you’ve been paying attention:

Question: Can you draw a diagram where an electron decays into any number of neutrinos?
Why not?

Some properties of neutrinos

We don’t quite have the full story of neutrinos yet, but here’s a glimpse of what’s to come:

 Those familiar with chemistry will know that neutrinos are produced in beta-decay
processes.

 There is a neutrino for each electron-like particle. This is not a coincidence.

 One of the great experimental discoveries in the past 15 years was that neutrinos have [a
very tiny] mass. It turns out that this is related to another remarkable property: neutrinos
change identity! An electron neutrino can spontaneously turn into a muon or a tau
neutrino. What’s even more remarkable is that this turns out to have a very deep
connection to the difference between matter and antimatter. This is something we’ll have
a lot to say about very soon.

 Because neutrinos are so light they played a key role in the early universe. As the
universe cooled down from the big bang, heavy particles could no longer be produced by
the ambient thermal energy. This left only neutrinos and photons buzzing around to
redistribute energy. This turned out to play an important role in the formation of galaxies
from quantum fluctuations.

Remarks about neutrino history

In the interests of getting to the electroweak model of leptons, I will not do justice to the rich and
fascinating history of neutrino physics. Here are a few highlights that I’ve found interesting.

 The Super Kamiokande detector in Japan was originally built to look for signals of
proton decay that is predicted by many models of grand unification. These proton decay
signals were never found (and are still being searched for), but in 1998 Super-K made a
breakthrough observation of neutrino oscillation.

 Neutrino oscillation solved the solar neutrino problem.

 More recently, last month the OPERA experiment at the Gran Sasso Laboratory in Italy
found further evidence for neutrino oscillation by directly observing a tau-
neutrino coming from a beam of muon neutrinos which had traveled 730 km from
CERN.
 One of the great theorists of the 1900s, Wolfgang Pauli, postulated the existence of a
neutral, light particle to explain apparent violations to energy conservation coming from
nuclear decays. He called the proposed particle a “neutron,” but also noted that it would
be extremely difficult to detect directly. Later Chadwick discovered the neutron
(what we call the neutron) but it was clearly too heavy to be Pauli’s “neutron,” so Fermi
renamed the latter to be the neutrino (“little neutral one”). Here’s a nice Logbook article
in Symmetry Magazine about Pauli’s original postulate that such a particle should exist.

 Neutrino physics has become one of the focus points of Fermilab’s research program into
the ‘intensity frontier.’ The general idea is to generate a beam of high-energy neutrinos
(using the Tevatron’s proton beam) and shoot it towards targets at different distances (up
to 450 miles away in Minnesota!). Because neutrinos are so weakly interacting, they pass
harmlessly through the earth at a slight downward angle until a small number of them
interact with large underground detectors at the target site.

 There are lots of neat proposals about interesting things one can do with neutrinos. To the
best of my knowledge, most of these are still in the “interesting idea” phase, but it’s a
nice example of potential spin-off technologies from fundamental research. Some
examples include

1. Probing geological activity deep underground, or even forecasting earthquakes.

2. One-way communication with deep ocean submarines.

3. Non-intrusive nuclear reactor inspection to check if nuclear reactors were being


used to produce weapons-grade plutonium.

4. Even more dramatically, neutralization of nuclear weapons.

he W boson: mixing things up

For those of you who have been following our foray into the particle content of the Standard
Model, this is where thing become exciting. We now introduce the W boson and present a nearly-
complete picture of what we know about leptons.

We’re picking up right where we left off, so if you need a refresher, please refer to previous
installments where we introduce Feynman rules and several particles: Part 1, Part 2, Part 3, Part
4, Part 5

The W is actually two particles: one with positive charge and one with negative charge. This is
similar to every electron having a positron anti-partner. Here’s the Particle Zoo’s depiction of
the W boson:
Together with the Z boson, the Ws mediate the weak [nuclear] force. You might remember this
force from chemistry: it is responsible for the radioactive decay of heavy nuclei into lighter
nuclei. We’ll draw the Feynman diagram for β-decay below. First we need Feynman rules.

Feynman Rules for the W: Interactions with leptons

Here are the Feynman rules for how the W interacts with the leptons. Recall that there are three
charged leptons (electron, muon, tau) and three neutrinos (one for each charged lepton).
In addition, there are also the same rules with the arrows pointing in opposite directions, for a
total of 18 vertices. Note that we’ve written plus-or-minus for the W, but we always use
the W with the correct charge to satisfy charge conservation.

Quick exercise: remind yourself why the rules above are different from those with arrows
pointing in the opposite direction. Hint: think of these as simple Feynman diagrams that we read
from left to right. Think about particles and anti-particles.

In words: the W connects any charged lepton to any neutrino. As shorthand, we can write these
rules as:

Here we’ve written a curly-L to mean “[charged]


lepton” and a νi to mean a neutrino of the ith type, where i can be electron/muon/tau.

Exercise: What are the symmetries of the theory? In other words, what are the conserved
quantities? Compare this to our previous theory of leptons without the W.

Answer: Electric charge is conserved, as we should expect. However, we no longer individually


conserve the number of electrons. Similarly, we no longer conserve the number of muons, taus,
electron-neutrinos, etc. However, the total lepton number is still conserved: the number of
leptons (electrons, muons, neutrinos, etc.) minus the number of anti-leptons stays the same
before and after any interaction.

Really neat fact #1: The W can mix up electron-like things (electrons and electron-neutrinos)
with not-electron-like things (e.g. muons, tau-neutrinos). The W is special in the Standard Model
because it can mix different kinds of particles. The “electron-ness” or “muon-neutrino-ness” (and
so forth) of a particle is often called its flavor. We say that the W mediates flavor-
changing processes. Flavor physics (of quarks) is the focus of the LHCb experiment at CERN.

Exercise: Draw a few diagrams that violate electron number. [If it’s not clear, convince yourself
that you cannot have such effects without a W in your theory.]

Answer: here’s one example: a muon decaying into an electron and a neutrino-antineutrino pair.
(Bonus question: what is the charge of the W?)
Remark (update 7 July): In the comments below Mori and Stephen point out that in the ‘vanilla’
Standard Model, leptons don’t have flavor-changing couplings to the W as I’ve drawn above.
This is technically true, at least before one includes the phenomena of neutrino-oscillations (only
definitively confirmed in 1998). In the presentation here I am assuming that such interactions
take place, which is a small modification from the “most minimal” Standard Model. Such
effects must take place due to the neutrino oscillation phenomena. We will discuss this in a
future post on neutrino-less double beta decay.

Feynman Rules for the W: Interactions with other force particles

There are additional Feynman rules. In fact, you should have already guessed one them: because
the Wis electrically charged, it interacts with the photon! Thus we have the additional Feynman
rule:

[Update, Aug 9: note that for these vertices I’ve used the convention that all of the bosons are in-
coming. Thus these are not Feynman Diagrams representing physical processes, they’re just
vertices which we can convert into diagrams or pieces or diagrams. For example, the
above vertex has an incoming photon, incoming W+, and an incoming W-. If we wanted
the diagram for a W+ emitting a photon (W+ -> W+ photon), then we would swap the incoming
W- for an outgoing W+ (they’re sort of antiparticles).]

This turns out to only be the tip of the iceberg. We can replace the photon with a Z (as one would
expect since the Z is a heavy cousin of the photon) to get another three-force-particle vertex:
Finally, we can even construct four force-particle vertices. Note that each of these satisfies
charge conservation!

These four-force-particle vertices are usually smaller than any of the previous vertices, so we
won’t spend too much time thinking about them.

Really neat fact #2: We see that the W introduces a whole new kind of Feynman rule: force
particles interacting with other force particles without any matter particles! (In fancy words:
gauge bosons interacting with other gauge bosons without any fermions.)

Remarks

1. The most interesting feature of the W is that it can change fermion flavors, i.e. it can not
only connect a lepton and a neutrino, but it can connect a lepton of one type with a
neutrino of a different type. One very strong experimental constraint on flavor physics
comes from the decay μ→eϒ (muon decaying to electron and photon). As an exercise,
draw a Feynman diagram contributing to this process. (Hint: you’ll need to have
a W boson and you’ll end up with a closed loop.)

2. It is worth noting, however, that these flavor-changing effects tend to be smaller than
flavor-conserving effects. In other words, a W is more likely to decay into an electron and
an electron-neutrino rather than an electron and a tau-neutrino. We’ll discuss how much
smaller these effects are later.

3. W bosons are rather heavy—around 80 GeV, slightly lighter than the Z but
still much heavier than any of the leptons. Thus, as we learned from the Z, it decays
before it can be directly observed in a detector.

4. The W was discovered at the UA1 and UA2 experiments at CERN in the 80s. Their
discovery was a real experimental triumph: as you now know from the Feynman rules
above, the W decays into a lepton and a neutrino—the latter of which cannot be
directly detected! This prevents experimentalists from observing a nice resonance as
they did for the Z boson a few months later. They used a slightly modified technique
based on a quantity called “transverse mass” to search for a smeared-out resonance using
only the information about the observed lepton. Generalizations of this technique are still
being developed today to search for supersymmetry! (For experts: see this recent review
article on LHC kinematics.)
5. The W boson only talks to left-handed particles. This is a remarkable fact that turns out
to be related to the difference between matter and antimatter. For a proper introduction,
check out this slightly-more-detailed post.

Exercise: Now that we’ve developed quite a bit of formalism with Feynman rules, try drawing
diagrams corresponding to W boson production at a lepton collider. Assume the initial particles
are an electron and positron. Draw a few diagrams that produce W bosons. “Finish” each
diagram by allowing any heavy bosons (Z, W) to decay into leptons.

What is the simplest diagram that includes a W boson? Is the final state observable in a detector?
(Remember: neutrinos aren’t directly observable.) What general properties do you notice in
diagrams that both (1) include a W boson and (2) have a detectable final state (at least one
charged lepton)?

Can you draw diagrams where the W boson is produced in pairs? Can you draw diagrams where
the W boson is produced by itself?

Hints: You should have at least one diagram where the W is the only intermediate particle. You
should also play with diagrams with both the fermion-fermion-boson vertices and the three-
boson vertices. You may also use the four-boson vertices, but note that these are smaller effects.

Remark: Try this exercise, you’ll really start to get a handle for drawing diagrams for more
complicated processes. Plus, this is precisely the thought process when physicists think about
how to detect new particles. As an additional remark, this is not quite how the W was
discovered—CERN used proton-antiproton collisons, which we’ll get to when we discuss
quantum chromodynamics.

Relating this to chemistry

Before closing our introduction to the W boson, let’s remark on its role in chemistry and
simultaneously give a preview for the weak interactions of quarks. You’ll recall that in chemistry
one could have β decay:

neutron → proton + electron + anti-neutrino

This converts one atom into an isotope of another atom. Let’s see how this works at the level of
subatomic particles.

Protons and neutrons are made out of up and down type quarks. Up quarks (u) have electric
charge +2/3 and down quarks (d) have electric charge -1/3. As we will see when we properly
introduce the quarks, up and down quarks have the same relationship as electron-neutrinos and
electrons. Thus we can expect a coupling between the up, down, and W boson.
A neutron is composed of two down quarks and an up quark (ddu) while a proton is composed of
two up quarks and a down quark (uud). [Check that the electric charges add up to what you
expect!] The diagram that converts a neutron to a proton is then:

Update: As reader Cris pointed out to me in an e-mail, the W should have negative charge and
should decay into an electron and anti-neutrino!

Because the W boson is much heavier than the up and down quarks—in fact, it’s much heavier
than the entire proton—it is necessarily a virtual particle that can only exist for a short time. One
can imagine that the system has to ‘borrow’ energy to create the W so that the Heisenberg
uncertainty principle tells us that it has to give back the energy very quickly. Thus the W can’t
travel very far before decaying and we say that it is a “short range force.” Thus sometimes the
weak force is called the weak nuclear force. Compare this to photons, which have no mass and
hence are a “long range force.”

[We now know, however, that it is not intrinsically a nuclear force (in our theory above we never
mentioned quarks or nuclei), and further its ‘weakness’ is related to the mass of the W making it
a short-range force.]

Meet the quarks

One of the most important experiments in the history of physics was the Rutherford
experiment where “alpha particles” were shot at a sheet of gold foil. The way that the particles
scattered off the foil was a tell-tale signature that atoms contained a dense nucleus of positive
charge. This is one of the guiding principles of high-energy experiments:

If you smash things together at high enough energies, you probe the substructure of those
particles.

When people say that the LHC is a machine colliding protons at 14 TeV, what they really mean
is that it’s a quark/gluon collider since these are the subnuclear particles which make up protons.
In this post we’ll begin a discussion about what these subatomic particles are and why they’re so
different from any of the other particles we’ve met.
(Regina mentioned QCD in her last post—I think “subtracting the effects of QCD,” loosely
phrased, is one of the ‘problems’ that both theorists and experimentalists often struggle with.)

This post is part of a series introducing the Standard Model through Feynman diagrams. An
index of these entries can be found on the original post. In this post we’ll just go over the matter
particles in QCD. (I’m experimenting with more frequent—but shorter—posts.)

A (partial) periodic table for QCD

The theory that describes quarks and gluons is called Quantum Chromodynamics, or QCD for
short. QCD is a part of the Standard Model, but for this post we’ll focus on just QCD by itself.
Quarks are the fermions—the matter particles—of the theory. There are six quarks, which come
in three “families” (columns in the table below):

The quarks have cute names: the up, down,


charm, strange, top, and bottom. Historically the t and b quarks have also been called “truth” and
“beauty,” but—for reasons I don’t quite understand—those names have fallen out of use, thus
sparing what would have been an endless parade of puns in academic paper titles.

The top row (u,c,t) is composed of particles with +2/3 electric charge while the bottom row is
composed of particles of -1/3 charge. These are funny values since we’re used to protons and
electrons with charges +1 and -1 respectively. On the one hand this is a historical effect: if we
measured the quark charges first we would have said that

 the down quark has charge -1

 the up quark has charge +2

 the electron has charge -3

 and the proton has charge +3

It’s just a definition of how much is one “unit” of charge. However, the fact that the quark and
lepton charges have these particular ratios is a numerical curiosity, since it is suggestive (for
reasons we won’t go into here) of something called grand unification. (It’s not really as “grand”
as it sounds.)

One quark, two quark, red quark, blue quark?

I drew the above diagram very suggestively: there are actually three quarks for each letter above.
We name these quarks according to colors: thus we can have a red up quark, a blue up quark,
and a greenup quark, and similarly with each of the five quarks. Let me stress that actual quarks
are not actually“colored” in the conventional sense! These are just names that physicists use.

The ‘colors’ are really a kind of “chromodynamic” charge. What does this mean? Recall in QED
(usual electromagnetism) that the electron’s electric charge means that it can couple to photons.
In other words, you can draw Feynman diagrams where photons and electrons interact. This is
precisely what we did in my first post on the subject. In QED we just had two kinds of charge:
positive and negative. When you bring a positive and negative charge together, they become
neutral. In QCD we generalize this notion by having three kinds of charge, and bringing
all three charges together gives you something neutral. (Weird!)

The naming of different kinds of quarks according to colors is actually very clever and is based
on the way that colored light mixes. In particular, we know that equal parts of red + green + blue
= white. We interpret “white” as “color neutral,” meaning having no “color charge.”

There’s a second way to get something color neutral: you can add something of one color with
it’s “anti-color.” (You can formalize these in color theory, but this would take us a bit off
course.) For example, the “anti-color” of red is cyan. So we could have red + “anti-red” (cyan) =
color neutral.

If we don’t see them, are quarks real?

The point of all of these “color mixing” analogies is that [at low energies], QCD is a strongly
coupledforce. In fact, we often just call it the strong force. It’s responsible for holding together
protons and neutrons. In fact, QCD is so strong that it forces all “color-charged” states to find
each other and become color neutral. We’ll get into some details about this in follow up posts
when we introduce the QCD force particles, the gluons. For now, you should believe (with a hint
of scientific skepticism) that there is no such thing as a “free quark.” Nobody has ever picked up
a quark and examined it to determine its properties. As far as you, me, the LHC, and everyone
else is concerned, quarks are always tied up in bound states.

There are two kinds of bound states:

 Bound states of 3 quarks: these are called baryons. You already know two: the proton
and the neutron. The proton is a combination (uud) while the neutron is a combination
(ddu). For homework, check that the electric charges add up to be +1 and 0. Because
these have to be color neutral, we know that the quark colors have to sum according to
red + green + blue.

 Bound states of a quark and an anti-quark: these are called mesons. These are color-
neutral because you have a color + it’s anti-color. The lightest mesons are called pions
and are composed of up and down quarks. For example, the π+ meson looks something
like (u anti-d). (Check to make sure you agree that it has +1 electric charge.)

Collectively these bound states are called hadrons. In the real world (i.e. in our particle
detectors) we only see hadrons because any free quarks automatically get paired up with either
anti-quarks or two other quarks. (Where do these quarks come from? We’ll discuss that soon!)

This seems to lead to an almost philosophical question: if quarks are always tied up in hadrons,
how do we know they really exist?

A neat historical fact: Murray Gell-Mann and Yuval Ne’eman, progenitors of the quark model,
originally proposed quarks as a mathematical tool to understand the properties of hadrons;
largely because we’d found lots of hadrons, but no isolated quarks. For a period in the 60s people
would do calculations with quarks as abstract objects with no physical relevance.

Why we believe that quarks are real

This seems to lead to an almost philosophical question: if quarks are always tied up in hadrons,
how do we know they really exist? Fortunately, we are physicists, not philosophers. Just as
Rutherford first probed the structure of the atomic nucleus by smashing high energy alpha
particles (which were themselves nuclei), the deep inelastic scattering experiments at the
Stanford Linear Accelerator Center (joint with MIT and Caltech) in the 60s collided electrons
into liquid hydrogen/deuterium targets and revealed the quark substructure of the proton.

A discussion of deep inelastic scattering could easily span several blog posts by itself. (Indeed, it
could span several weeks in a graduate quantum field theory course!) I hope to get back to this in
the future, since it was truly one of the important discoveries of the second half of the twentieth
century. To whet your appetites, I’ll only draw the Feynman diagram for the process:
This is unlabeled, but by now you should see what’s going on. The particle on top is the electron
that interacts with the proton, which is drawn as the three quark lines on the bottom left. The
circle (technically called a “blob” in the literature) represents some QCD interactions between
the three quarks (holding them together). The electron interacts with a quark through some kind
of force particle, the wiggly line. For simplicity you can assume that it is a photon (for
homework, think about what is different if it’s a W). We’ve drawn the quark that interacts as the
isolated line coming out of the blob.

This quark is somewhat special because it’s the particle that the electron recoils against. This
means that it gets a big kick in energy, which can knock it out of the proton. As I mentioned
above, this quark is now “free” — but not for long! It has to hadronize into more complicated
QCD objects, mesons or baryons. The spectrum of outgoing particles gives clues about what
actually happened inside the diagram.

We’ve just glossed over the surface of this diagram: there is a lot of very deep (no pun intended)
physics involved here. (These sorts of processes are also a notorious pain in the you-know-where
to calculate the first time one meets them in graduate courses.)

(By the way: the typical interactions of interest at the LHC are similar to the diagram above, only
with two protons interacting!)

A hint of group theory and unification

I would be negligent not to mention some of the symmetry of the matter content of the Standard
Model. Let’s take a look at all of the fermions that we’ve met so far:
There are all sorts of fantastic patterns that one can glean from things that we’ve learned in these
blog posts alone!

The top two rows are quarks (each with three different colors), while the bottom two rows are
leptons. Each row has a different electric charge. Each column carries the same
properties, except that each successive column is heavier than the previous one. We learned that
the W boson mediates decays between the columns, and since heavy things decay into lighter
things, most of our universe is made up of exclusively the first column.

There are other patterns we can see. For example:

 When we first met QED, we only needed one type of particle, say the electron. We knew
that electrons and anti-electrons (positrons) could interact with a photon.

 When we met the weak force (the W boson), we needed to introduce another type or
particle: the neutrino. An electron and an anti-neutrino could interact with a W boson.

 Now we’ve met the strong force, QCD. In our next post we’ll meet the force particle, the
gluon. What I’ve already told you, though, is that there are three kinds of particles that
interact with QCD: red, green, and blue. In order to form something neutral, you need all
three color charges to cancel.

There’s a very deep mathematical reason why we get this one-two-three kind of counting: it
comes from the underlying “gauge symmetry” of the Standard Model. The mathematical field
of group theory is (a rough definition) the study of how symmetries can manifest themselves.
Each type of force in the Standard Model is associated with a particular “symmetry group.”
Without knowing what these names mean, it should not surprise you if I told you that the
symmetry group of the Standard Model is: U(1) SU(2) SU(3). There’s that one-two-three
counting!

It turns out that this is also very suggestive of grand unification. The main thrust of the idea is
that all three forces actually fit together in a nice way into a single force which is represented by
a single “symmetry group,” say, SU(5). In such a scheme, each column in the “periodic table”
above can actually be “derived” from the mathematical properties of the GUT (grand unified
theory) group. So in the same way that QCD told us we needed three colors, the GUT group
would tell us that matter must come in sets composed of quarks with three colors, a charged
lepton, and a neutrino; all together!

By the way, while they sound similar, don’t confuse “grand unified theories” with a “theory of
everything.” The former are theories of particle physics, while the latter try to unify particle
physics with gravity (e.g. string theory). Grand unified theories are actually fairly mundane and I
think most physicists suspect that whatever completes the Standard Model should
somehow eventually unify (though there has been no direct experimental evidence yet).
“Theories of everything” are far more speculative by comparison.

Where we’ll go from here?

I seem to have failed in my attempt to write shorter blog posts, but this has been a quick intro to
QCD. Hopefully I can write up a few more posts describing gluons, confinement, and hadrons.

For all of you LHC fans out there: QCD is really important. (For all of you LHC scientists out
there, you already know that the correct phrase is, “QCD is really annoying.”) When we say that
SLAC/Brookhaven discovered the charm quark or that Fermilab discovered the top quark,
nobody actually bottled up a quark and presented it to the Nobel Prize committee. Our detectors
see hadrons, and the properties of particular processes like deep inelastic scattering allow us to
learn somewhat indirectly about the substructure of these hadrons to learn about the existence of
quarks. This, in general, is really, really, really hard—both experimentally and theoretically.

Thanks everyone,
Flip, US LHC Blogs

(By the way, if there are particle physics topics that people want to hear about, feel free to leave
suggestions in the comments of the blog. I can’t promise that I’ll be able to discuss all of them,
but I do appreciate feedback and suggestions. Don’t worry, I’ll get to the Higgs boson
eventually… first I want to discuss the particles that we have discovered!)

World of Glue
I’m a bit overdue for my next post introducing the Standard Model through Feynman diagrams,
so let’s continue our discussion of the theory the subnuclear “strong” force, quantum
chromodynamics, or QCD. This is the force that binds quarks into protons and neutrons, and the
source of many theoretical and experimental headaches at the LHC.

For a listing of all the posts in this series, see the original post here.

Last time we met the quarks and “color,” a kind of charge that can take three values: red, green,
and blue. Just as the electromagnetic force (which has only one kind of charge) pulls negatively
charged electrons to positively charged protons to form electrically neutral atoms, QCD forces
the quarks to be confined into color-neutral bound states:

 Three quarks, one of each color. These are called baryons. Common examples are the
proton and neutron.

 A quark and an anti-quark of the same color (e.g. a red and anti-red quark). These are
called mesons.

Collectively baryons and mesons are called hadrons. Because the strong force is so strong, these
color-neutral bound states are formed very quickly and they are all we see in our detectors. Also
like the electromagnetic force, the strong force is mediated by a particle: a boson called
the gluon, represented in plush form below (courtesy of the Particle Zoo)

Gluons are so named because they glue together mesons and baryons into bound states. In
Feynman diagrams we draw gluons as curly lines. Here’s our prototypical gluon-quark vertex:
We see that the gluon takes an incoming quark of one color and turns it into an outgoing quark of
another color. According to the rules associated with Feynman diagrams, we can move lines
around (always keeping the orientation of arrows relative to the vertex the same!) and interpret
this vertex as

 A red quark and a blue anti-quark (with color anti-blue) annihilate into a gluon

 A red quark emits a gluon and turns into a blue quark

 A red quark absorbs a gluon and turns into a blue quark

 A gluon decays into a blue quark and a red anti-quark

As a simple homework, you can come up with the two interpretations that I’ve left out.

Let us make the following caveats:

1. The quarks needn’t have different colors, but one arrow has to be pointing in while the
other is pointing out. (The quarks carry electric charge, and remember that one way to
understand the arrows is as the flow of electric charge.)

2. The quarks involved in the vertex can have any flavor (up, down, strange, charm, bottom,
top), but both must have the same flavor. This is because QCD is “flavor blind,” it treats
all of the flavors equally and doesn’t mix them up. (Compare this to the W boson!)

3. The interpretations for the vertex above are all correct, except that none of them are
allowed kinematically. This is because the gluon is massless. In other words,
conservation of energy and momentum are violated if you consider those processes. This
is for precisely the same reason that we couldn’t have single-vertex photon diagrams in
QED.

Homework: up an down quark scattering by gluons. Draw all the diagrams for the following
processes that are mediated by a single gluon (i.e. only contain a single internal gluon line)

1. uu → uu (one diagram)

2. u anti-u → u anti-u (two diagrams)


3. u anti-u → d anti-d (one diagram)

4. u anti-d → u anti-d (one diagram)

You may assign colors as necessary (explain why it matters or does not matter). Why is it
impossible to draw a u anti-d → d anti-u diagram (note that this process is allowed if you replace
the gluon by a W)? [Hint: you might want to review some of our discussions
about QED and muons; the diagrams are all very similar with photons replaced by gluons!]

We can continue to make analogies to QED. We explained that the QED vertex had to be charge
neutral: an arrow pointing inwards carries some electric charge, while the arrow pointing
outward must carry the same electric charge. The gluon vertex above is electrically neutral in this
sense, but does not seem to be color neutral! It brings in some red-charge while splitting out
some blue charge.

The resolution is that gluons themselves carry color charge! This is now very different from
QED. It’s a little bit like the W boson, which carried electric charge and so could interact with
photons. We can see from the vertex above that the gluon must, in fact, carry two charges: in that
example the gluon carries an incoming blue charge and and outgoing red charge; or, in other
words charge blue and charge anti-red. These are the charges that it must carry in order for the
vertex to be color neutral.

Thus there are actually many types of gluons which we can classify according to the color and
anti-color which they can carry. Since there are three colors (and correspondingly three anti-
colors), we expect there to be nine types of gluons. However, for somewhat mathematical
reasons, it turns out that there are only eight.

The mathematical reason is that the gluons are in the adjoint representation of SU(3) and so
number only 32-1 = 8. They are associated with the space of traceless Hermitian 3×3 matrices.
The “missing” gluon is the quantum superposition of “red/anti-red + blue/anti-blue + green/anti-
green.” If that’s all gibberish to you, then that’s okay—these are little details that we won’t need.

The fact that gluons themselves carry color charge means something very important: gluons feel
the strong force, which means that gluons interact with other gluons! In other words, we can
draw three- and four- gluon vertices:
There are no five or higher vertices, but as homework you can convince yourself that from these
vertices you can draw diagrams with any number of external gluons. In fancy schmancy
mathematical language, we say that QCD is non-Abelian because the force mediators interact
with themselves. (In fact, the weak force is also non-Abelian, but its story is one of broken
dreams which we will get to when we meet the Higgs.)

Now, you might wonder: if the strong force is so strong that gluons bind quarks together, and if
gluons also interact with themselves, is it possible for gluons to bind each other into some kind
of bound state? The answer is yes, though we have yet to confirm this experimentally. The bound
states are called glueballs and can be pretty complicated objects. Theoretically we have good
reasons to believe that they should exist (and eventually decay into mesons and baryons), and
very sophisticated simulations of QCD have also suggested that they should exist… but
experimentally they are very hard to see in a detector and we have yet to confirm any glueball
signature. Very recently some theorists have suggested that there might have been hints at the
BES collider in Beijing.

Gluon hunting, however, is something of a lower energy frontier since our predictions for the
lightest glueball masses are around 1.7 GeV; so don’t expect anything from the LHC on this. It’s
worth remarking, on the other hand, about a mathematical issue regarding a world of glue. (This
is a bad pun for a computer game that I like.) The question is: if the universe had no matter
particles and only gluons—which would then form glueball bound states—are there any
massless particles observable in nature? Sure, the gluons themselves are massless, but they’re
not observable states; only glueballs are observable. Everything we know about QCD—which
isn’t the whole story—suggests that glueballs always have some non-zero mass, but we don’t
know how to prove this. This question, in fact, is one of the Clay Mathematics Millennium Prize
Problems, making it literally a million dollar question.

Next time: the wonderful world of hadrons and what we can actually detect at the LHC.

QCD and Confinement

Now that we’ve met quarks and gluons, what I should do is describe how they interact with
the othersectors of the Standard Model: how do they talk to the leptons and gauge bosons
(photon, W, Z) that we met in the rest of this series on Feynman diagrams. I’ll have to put this off
a little bit longer, since there’s still quite a lot to be said about the “fundamental problem” of
QCD:

The high energy degrees of freedom (quarks and gluons) are not what we see at low energies
(hadrons).

Colliders like the LHC smash protons together at high energies so that the point-like interactions
are between quarks and gluons. By the time these quarks and gluons scatter into the LHC
detectors, however, they have now “dressed” themselves into hadronic bound states. This is the
phenomenon of confinement.

As a very rough starting point, we can think about how protons and electrons are bound into the
hydrogen atom. Here the electric potential attracts the proton and electron to one another. We
can draw the electric field lines something like this:

These are just like the patterns of iron filings near a bar magnet. The field lines are, of course,
just a macroscopic effect set up by lots and lots of photons, but we’re in a regime where we’re
justified in taking a “semi-classical” approximation. In fact, we could have drawn the same field
lines for gravity. They are all a manifestation of the radially symmetric 1/r potential. We can try
to extend this analogy to QCD. Instead of a proton and electron attracted by the electric force,
let’s draw an up quark and a down quark attracted by the color (chromodynamic) force.

This looks exactly the same as the electric picture above, but instead of photons setting up a
classical field, we imagine a macroscopic configuration of gluons. But wait a second! There’s no
such thing as a macroscopic configuration of gluons! We never talk about long range classical
chromodynamic forces.

Something is wrong with this picture. We could guess that maybe the chromodynamic force law
takes a different form than the usual V(r) ~ 1/r potential for electricity and gravity. This is indeed
a step in the right direction. In fact, the chomodynamic potential is linear: V(r)~ r. But what does
this all mean?

By the way, the form of the potential is often referred to as the phase of the theory. The “usual”
1/r potential that we’re used to in classical physics is known as the Coulomb phase. Here we’ll
explain what it means that QCD is in the confining phase. Just for fun, let me mention another
type of phase called the Higgs phase, which describes the weak force and is related to the
generation of fermion masses.

Okay, so I’ve just alluded to a bunch of physics jargon. We can do better. The main question we
want to answer is: how is QCD different from the electric force? Well, thing about electricity is
that I can pull an electron off of its proton. Similarly, a satellite orbiting Earth can turn on its
thrusters and escape out of the solar system. This is the key difference between electricity (and
gravity) and QCD. As we pull the electron far away from the proton, then the field lines near the
proton “forget” about the electron altogether. (Eventually, the field lines all reach the electron,
but they’re weak.)

QCD is different. The as we pull apart the quarks, the force is that pulls them back together
becomes stronger energy stored in the gluon field gets larger. The potential difference
gets larger and it takes more energy to keep those quarks separated, something like a spring. So
we can imagine pulling the quarks apart further and further. You should imagine the look of
anguish on my face as I’m putting all of my strength into trying to pull these two quarks apart—
every centimeter I pull they want to spring back towards one another with even more force…

… stores more and more energy in the gluon field. (This is the opposite of QED, where the
energy decreases as I pull the electron from the proton! Errata: 10/23, this statement is
incorrect! See the comments below. Thanks to readers Josh, Leon, Tim, and Heisenberg for
pointing this out!) Think of those springy “expander” chest exercise machines. Sometimes we
call this long, narrow set of field lines a flux tube. If we continued this way and kept pulling,
then classical physics would tell us that we can get generate arbitrarily large energy!
Something has to give. Classically cannot pull two quarks apart.

Errata (10/22): Many thanks to Andreas Kronfeld for pointing out an embarrassing error: as I
pull the quarks apart the force doesn’t increase—since the potential is linear V(r) ~ r, the force
is constant, F(r) ~ -V'(r) ~ constant. Physicists often make this mistake when speaking to the
public because in the back of their minds they’re thinking of a quantum mechanical property of
QCD called asymptotic freedom in which the coupling “constant” of QCD actually increases as
one goes to large distances (so it’s not much of a constant). As Andreas notes, this phenomenon
isn’t the relevant physics in the confining phase so we’ll leave it for another time, since a proper
explanation would require another post entirely. I’ve corrected my incorrect sentences above.
Thanks, Andreas!
What actually happens is that quantum mechanics steps in. At some point, as I’m pulling these
quarks apart, the energy in the gluon field becomes larger than the mass energy of a quark anti-
quark pair. Thus it is energetically favorable for the gluons to produce a quark–anti-quark pair:

From the sketch above, this pair production reduces the energy in the gluon field. In other words,
we turned one long flux tube into two shorter flux tubes. Yet another way to say this is to think
of the virtual(quantum mechanical) quark/anti-quark pairs popping in and out of the vacuum,
spontaneously appearing and then annihilating. When the energy in the gluon field gets very
large, though, the gluons are able to pull apart the quark/anti-quark pair before they can
annihilate, thus making the virtual quarks physical.

This is remarkably different behavior from QED, where we could just pull off an electron and
send it far away. In QCD, you can start with a meson (quark–anti-quark pair) and try to pull apart
its constituents. Instead of being able to do this, however, you inadvertently break the meson not
into two quarks, but into two mesons. Because of this, at low energies one cannot observe
individual quarks, they immediately confine (or hadronize) into hadronic bound states.

Some context

This idea of confinement is what made the quark model so hard to swallow when it was first
proposed: what is the use of such a model if one of the predictions is that we can’t observe the
constituents? Indeed, for a long time people thought of the quark model as just a mathematical
trick to determine relations between hadrons—but that “quarks” themselves were not physical.

On the other hand, imagine how bizarre this confinement phenomenon must have
seemed without the quark model. As you try to pull apart a meson, instead of observing
“smaller” objects, you end up pulling out two versions of the same type of object! How could it
have been that inside one meson is two mesons? This would be like a Russian matryoshka doll
where the smaller dolls are the same size as the larger ones—how can they fit? (Part of the
failure here is classical intuition.) This sort of confusion led to the S–matrix or “bootstrap”
program in the 60s where people thought to replace quantum field theory with something where
the distinction “composite” versus “elementary” particles was dropped. The rise of QCD showed
that this was the wrong direction for the problem and that the “conservative” approach of
keeping quantum theory was able to give a very accurate description of the underlying physics.
In some sense the S-matrix program is a famous “red herring” in the history of particle physics.
However, it is a curious historical note—and more and more so a curious scientific note—that
this ‘red herring’ ended up planting some of the seeds for the development of string theory,
which was originally developed to try to explain hadrons! The “flux tubes” above were
associated with the “strings” in this proto-string theory. With the advent of QCD, people realized
that string theory doesn’t describe the strong force, but seemed to have some of the ingredients
for one of the “holy grails” of theoretical physics, a theory of quantum gravity.

These days string theory as a “theory of everything” is still up in the air, as it turns out that there
are some deep and difficult-to-answer questions about string theory’s predictions. On the other
hand, the theory has made some very remarkable progress in directions other than the
“fundamental theory of everything.” In particular, one idea called the AdS/CFT correspondence
has had profound impacts on the structure of quantum field theories independent of whether or
not string theory is the “final theory.” (We won’t describe what the AdS/CFT correspondence is
in this post, but part of it has to do with the distinction between elementary and composite
states.) One of the things we hope to extract from the AdS/CFT idea is a way to describe theories
which are strongly coupled, which is a fancy phrase for confining. In this way, some branches
of stringy research is finding its way back to its hadronic origins.

Even more remarkable, there has been a return to ideas similar to the S-matrix program in recent
research directions involving the calculation of scattering amplitudes. While the original aim of
this research was to solve problems within quantum field theory—namely calculations in QCD—
some people have started to think about it again as a framework beyond quantum field theory.

High scale, low scale, and something in-between

This is an issue of energy scales. At high energies, we are probing short distance physics so that
the actual “hard collisions” at the LHC aren’t between protons, but quarks and gluons. On the
other hand, at low energies these “fundamental” particles always confine into “composite”
particles like mesons and these are the stable states. Indeed, we can smash quarks and gluons
together at high energies, but the QCD stuff that reaches the outer parts of the experimental
detectors are things like mesons.

In fact, there’s an intermediate energy scale that is even more important. What is happening
between the picture of the “high energy” quark and the “low energy meson?” The quark barrels
through the inner parts of the detector, it can radiate energy by emitting gluons.

… These gluons can produce quark/anti-quark pairs


… which themselves can produce gluons
… etc., etc.

At each step, the energy of the quarks and gluons decrease, but the number of particles increases.
Eventually the energy is such that the “free quarks” cannot prevent the inevitable and they must
hadronize. Because there are so many, however, there are a lot of mesons barreling through the
detector. The detector is essentially a block of dense material which can measure the energy
deposited into it, and what it ‘sees’ is a “shower” of energy in a particular direction. This is what
we call a jet, and it is the signature of a high energy quark or gluon that shot off in a particular
direction and eventually hadronizes. Here’s a picture that I borrowed from a CDF talk:

Read the picture from the bottom up:

1. First two protons collide… by which we really mean the quarks and gluons inside the
proton interact.

2. High energy quarks and gluons spit off other quark/gluons and increase in number

3. Doing this reduces their energy so that eventually the quarks and gluons must confine
(hadronize) into mesons

4. … which eventually deposit most of their energy into the detector (calorimeter)

Jets are important signatures at high energy colliders and are a primary handle for understanding
the high energy interactions that we seek to better understand at the LHC. In order to measure the
energy and momentum of the initial high energy quark, for example, one must be able to
measure all of the energy and momentum from the spray of particles in the jet, while taking into
account the small cracks between detecting materials as well as any sneaky mesons which may
have escaped the detector. (This is the hadronic analog of the electromagnetic calorimeter that
Christine recently described.)

Now you can at least heuristically see why this information can be so hard to extract. First the
actual particles that are interacting at high energies are different from the particles that exist at
low energies. Secondly, even individual high-energy quarks and gluons lead to a big messy
experimental signature that require careful analysis to extract even “basic” information about the
original particle

“Known knowns” of the Standard Model

This is the tenth (or so) post about Feynman diagrams, there’s an index to the entire series in
the first post.

There is a famous quote by former Secretary of Defense Donald Rumsfeld that really applies to
particle physicists:

There are known knowns.


These are things we know that we know.
There are known unknowns.
That is to say, there are things that we know we don’t know.
But there are also unknown unknowns.
There are things we don’t know we don’t know.

Ignoring originally intended context, this statement describes not only the current status of the
Standard Model, but accurately captures all of our hopes and dreams about the LHC.

 We have “known knowns” for which our theories have remarkable agreement with
experiment. In this post I’d like to summarize some of these in the language of Feynman
diagrams.

 There are also “known unknowns” where our theories break down and we need
something new. This is what most of my research focuses on and what I’d like to write
about in the near future.

 Finally, what’s most exciting for us is the chance to trek into unexplored territory and
find something completely unexpected—“unknown unknowns.”

Today let’s focus on the “known knowns,” the things that we’re pretty sure we understand.
There’s a very important caveat that we need to make regarding what we mean by “pretty sure,”
but we’ll get to that at the bottom. The “known knowns” are what we call the Standard
Model of particle physics*, a name that says much about its repeated experimental
confirmations.

* — a small caveat: there’s actually one “known unknown” that is assumed to be part of the
Standard Model, that’s the Higgs boson. The Higgs is currently one of the most famous yet-to-
be-discovered particle and will be the focus of a future post. In the meanwhile, Burton managed
to take a few charming photos of the elusive boson in his recent post.
First, let’s start by reviewing the matter particles of the Standard Model. These are
called fermions and they are the “nouns” of our story.

MATTER PARTICLES: THE FERMIONS

We can arrange these in a handy little chart, something like a periodic table for particle physics:

Let’s focus on only the highlighted first column. This contains all of the ‘normal’ matter particles
that make up nearly all matter in the universe and whose interactions explain everything we need
to know about chemistry (and arguably everything built on it).

The top two particles are the up and down quarks. These are the guys which make up the proton
(uud) and neutron (udd). As indicated in the chart, both the up and down quarks come in three
“colors.” These aren’t literally colors of the electromagnetic spectrum, but a handy mnemonic
for different copies of the quarks.

Below the up and down we have the electron and the electron-neutrino (?e), these are
collectively known as leptons. The electron is the usual particle whose “cloud” surrounds an
atom and whose interactions is largely responsible for most of chemistry. The electron-
neutrino is the electron’s ghostly cousin; it only interacts very weakly and is nearly massless.

As we said, this first column (u, d, e, and ?e) is enough to explain just about all atomic
phenomena. It’s something of a surprise, then, that we have two more columns of particles that
have nearly identical properties as their horizontal neighbors. The only difference is that as you
move to the right on the chart above, the particles become heavier. Thus the charm quark (c) is a
copy of the up quark that turns out to be 500 times heavier. The top quark (t) is heavier still;
weighing in at over 172 GeV, it is the heaviest known elementary particle. The siblings of the
down quark are the strange (s) and bottom (b) quarks; these have historically played a key role
in flavor physics, a field which will soon benefit from the LHCbexperiment. Each of these
quarks all come in three colors, for a total of 2 types x 3 colors x 3 columns = 18 fundamental
quarks. Finally, the electrons and neutrinos come with copies named muon (?) and tau(?). It’s
worth remarking that we don’t yet know if the muon and tau neutrinos are heavier than the
electron-neutrino. (Neutrino physics has become one of Fermilab’s major research areas.)

So those are all of the particles. As we mentioned in our first post, we can draw these as solid
lines with an arrow going through them. You can see that there are two types of leptons (e.g.
electron-like and neutrino) and two types of quarks (up-like and down-like), as well as several
copies of these particles. In addition, each particle comes with an antiparticle of opposite electric
charge. I won’t go into details about antimatter, but see this previous post for a very thorough
(but hopefully still accessible) description.

You can think of them as nouns. We now want to give them verbs to describe how they can
interact with one another. To do this, we introduce force particles (bosons) and provide
the Feynman rules to describe how the particles interact with one another. By stringing together
various particle lines to the vertices describing interactions, we end up with a Feynman
diagram that tells the story of a particle interaction. (This is the “sentence” formed from the
fermion nouns and the boson verbs.)

We will refer to these forces by the ‘theories’ that describe them, but they are all part of the
larger Standard Model framework.

QUANTUM ELECTRODYNAMICS

The simplest force to describe is QED, the theory of electricity and magnetism as mediated by
the photon. (Yes, this is just the “particle” of light!) Like all force particles, we draw the photon
as a wiggly line. We drew the fundamental vertex describing the coupling of an electron to the
photon in one of our earliest Feynman diagram posts,

For historical reasons, physicists often write the photons as a gamma, ?. Photons are massless,
which means they can travel long distances and large numbers of them can set up macroscopic
electromagnetic fields. As we described in our first post, you are free to move the endpoints of
the vertex freely. At the end of the day, however, you must have one arrowed line coming into
the vertex and one arrowed line coming out. This is just electric charge conservation.

In addition to the electron, however, all charged particles interact with the photon the same
vertex. This means that all of the particles above, except the neutrinos, have this vertex. For
example, we can have an “uu?” vertex where we just replace the e‘s above by u‘s.

QED is responsible for electricity and magnetism and all of the good stuff that comes along with
it (like… electronics, computers, and the US LHC blog).

QUANTUM FLAVORDYNAMICS

This is a somewhat antiquated name for the weak force which is responsible for radioactivity
(among other things). There are two types of force particle associated with the weak force:
the Z boson and the W boson. Z bosons are heavier copies of photons, so we can just take the
Feynman rule above and change the ? to a Z. Unlike photons, however, the Z boson can also
interact with neutrinos. The presence of the Z plays an important role in the mathematical
consistency of the Standard Model, but for our present purposes they’re a little bit boring since
they seem like chubby photon wanna-be’s.

On the other hand, the W boson is something different. The W carries electric charge and will
connect particles of different types (in such a way that conserves overall charge at each vertex).
We can draw the lepton vertices as:

We have written a curly-L to mean a charged lepton (e, ?, ?) and ?i to mean any neutrino (?e, ??,
??). An explicit set of rules can be found here. In addition to these, the quarks also couple to the
W in precisely the same way: just replace the charged lepton and neutrino by an up-type quark
and a down-type quark respectively. The different copies of the up, down, electron, and electron-
neutrino are called flavors. The W boson is special because it mediates interactions between
different particle flavors. Note that it does not mix quarks with leptons.

Because the W is charged, it also couples to photons:


It also couples to the Z, since the Z just wants to be a copy-cat photon:

Finally, the W also participates in some four-boson interactions (which will not be so important
to us):

QUANTUM CHROMODYNAMICS

Finally, we arrive at QCD: the theory of “strong force.” QCD is responsible for binding quarks
together into baryons (e.g. protons and neutrons) and mesons (quark–anti-quark pairs). The
strong force is mediated by gluons, which we draw as curly lines. Gluons couple to particles
with color, so they only interact with the quarks. The fundamental quark-gluon interaction takes
the form

The quarks must be of the same flavor (e.g. the vertex may look like up-up-gluon but not up-
down gluon) but may be of different colors. Just as the photon vertex had to be charge-neutral,
the gluon vertex must also be color-neutral. Thus we say that the gluon carries a color and an
anti-color; e.g. red/anti-blue. For reasons related to group theory, there are a total of eight gluons
rather than the nine that one might expect. Further, because gluons carry color, they interact with
themselves:

QCD—besides holding matter together and being a rich topic in itself—is responsible for all
sorts of head aches from both theoretical and experimental particle physicists. On the
experimental side it means that individual quarks and gluons appear as complicated hadronic jets
in particle colliders (see, e.g. Jim’s latest post). On the theoretical side the issue of strong
coupling (and the related idea of confinement) means that the usual ‘perturbative’ techniques to
actually calculate the rate for a process quickly becomes messy and intractable. Fortunately,
there are clever techniques on both fronts that we can use to make progress.

THE MISSING PIECE: THE HIGGS BOSON

Everything we’ve reviewed so far are known knowns, these are parts of our theory that have been
tested and retested and give good agreement with all known experiments. There are a few
unknown parameters such as the precise masses of the neutrinos, but these are essentially just
numbers that have to be measured and plugged into the existing theory.

There’s one missing piece that we know must either show up, or something like it must show up:
the Higgs boson. I’d like to dedicate an entire post to the Higgs later, so suffice it to say for now
that the Higgs is an integral part of the Standard Model. In fact, it is intimately related to the
weak sector. The importance of the Higgs boson is something called electroweak symmetry
breaking. This is a process that explains why particles have the masses that they do and why
the W, Z, and photon should be so interwoven. More importantly, the entire structure of the
Standard Model breaks down unless something like the Higgs boson exists to induce electroweak
symmetry breaking: the mathematical machinery behind these diagrams end up giving
nonsensical results like probabilities that are larger than 100%. Incidentally, this catastrophic
nonsenical behavior begins at roughly the TeV scale—precisely the reason why this is the energy
scale that the LHC is probing, and precisely the reason why we expect it to find something.

A fancy way of describing the Standard Model is that there are actually four Higgs bosons, but
three of them are “eaten” by the W and Z bosons when they become massive. (This is called the
Goldstone mechanism, but you can think of it as the Grimm’s Fairy Tale of particle physics.)
This has led snarky physicists to say things like, “Higgs boson? We’ve already found three of
them!”
THEORIES AND EFFECTIVE THEORIES

By specifying the above particles and stating how the Higgs induces electroweak symmetry
breaking, one specifies everything about the theory up to particular numbers that just have to be
measured. This is not actually that much information; the structure of quantum mechanics and
special relativity fixes everything else: how to write down predictions for different kinds of
processes between these particles.

But now something seems weird: we’ve been able to check and cross-check the Standard Model
in several different ways. Now, however, I’m telling you that there’s this one last missing
piece—the Higgs boson—which is really really important… but we haven’t found it yet. If
that’s true, how the heck can we be so sure about our tests of the Standard Model? How can
these be “known knowns” when we’re missing the most important part of the theory?

More generally, it should seem funny to say that we “know” anything with any certainty in
science! After all, part of the excitement of the LHC is the hope that the data will contradict the
Standard Model and force us to search for a more fundamental description of Nature. The basis
of the scientific method is that a theory is only as good as the last experiment which checked it,
and there are good reasons to believe that the Standard Model breaks down at some scale. If this
is the case, then how can we actually “know” anything within the soon-to-be-overthrown
Standard Model paradigm?

The key point here is that the Standard Model is something called an effective theory. It
captures almost everything we need to know about physics below, say, 200 GeV, but doesn’t
necessarily make any promises about what is above that scale. In fact, the sicknesses that the
Standard Model suffers from when we remove the Higgs boson (or something like it) are just the
theory’s way of telling us, “hey, I’m no longer valid here!”

This is not as weird as one might think. Consider the classical electromagnetic field of a point
particle: it is a well known curiosity to any high school student that the potential at the exact
location of the point source is infinity. Does that mean that an electron has infinite energy? No!
In fact, this seemingly nonsensical prediction is classical electromagnetism telling us that
something new has to fix it. That something new is quantum mechanics and the existence of
antiparticles, as we previously discussed.

This doesn’t mean that the effective theory is no good, it only means that it breaks down above
some region of validity. Despite the existence of quantum mechanics, the lessons we learn from
high school physics were still enough for us to navigate space probes to explore the solar
system. We just shouldn’t expect to trust Newtonian mechanics when describing subatomic
particles. There’s actually a rather precise sense in which a quantum field theory is “effective,”
but that’s a technical matter that shouldn’t obfuscate the physical intuition presented here.
For physicists: the theory of the Standard Model without a Higgs is a type of non-linear sigma
model (NL?M). This accurately describes a theory of massive vector bosons but suffers from a
breakdown of unitarity. The Higgs is the linear completion of the NL?M that increases the
theory’s cutoff. In fact, this makes the theory manifestly unitary, but does not address the
hierarchy problem. For an excellent pedagogical discussion, see Nima Arkani-Hamed’s PiTP
2010 lectures.

WHERE WE GO FROM HERE

The particles and interactions we’ve described here (except the Higgs) are objects and processes
that we have actually produced and observed in the lab. We have a theory that describes all of it
in a nice and compact way, and that theory requires something like the Higgs boson to make
sense at high energies.

That doesn’t mean that there aren’t lots of open questions. We said that the Higgs is related to
something called “electroweak symmetry breaking.” It is still unknown why this happens.
Further, we have good reason to expect the Higgs to appear in the 115 – 200 GeV range, but
theoretically it takes a “natural” value at the Planck mass (1019 GeV!). Why should the Higgs be
so much lighter than its “natural” value? What particle explains dark matter? Why is there more
matter than anti-matter in the universe?

While the Higgs might be the last piece of the Standard Model, discovering the Higgs (or
something like it!) is just the beginning of an even longer and more exciting story. This is at the
heart of my own research interests, and involves really neat-sounding ideas
like supersymmetry and extra dimensions

hen Feynman Diagrams Fail

We’ve gone pretty far with our series of posts about learning particle physics through Feynman
diagrams. In our last post we summarized the Feynman rules for all of the known particles of the
Standard Model. Now it’s time to fess up a little about the shortcomings of the Feynman diagram
approach to calculations; in doing so, we’ll learn a little more about what Feynman diagrams
actually represent as well as the kinds of physics that we must work with at a machine like the
LHC.

WHEN ONE DIAGRAM ISN’T ENOUGH

Recall that mesons are bound states of quarks and anti-quarks which are confined by the strong
force. This binding force is very non-perturbative; in other words, the math behind our Feynman
diagrams is not the right tool to analyze it. Let’s go into more detail about what this means.
Consider the simplest Feynman diagram one might draw to describe the gluon-mediated
interaction between a quark and an anti-quark:
Easy, right? Well, one thing that we have glossed over in our discussions of Feynman diagrams
so far is that we can also draw much more complicated diagrams. For example, using the QCD
Feynman rules we can draw something much uglier:

This is another physical contribution to the interaction between a quark and an anti-quark. It
should be clear that one can draw arbitrarily many diagrams of this form, each more and more
complicated than the last. What does this all mean?

Each Feynman diagram represents a term in a mathematical expression. The sum of these terms
gives the complete probability amplitude for the process to occur. The really complicated
diagrams usually give a much smaller contribution than the simple diagrams. For example,
each photon vertex additional internal photon line (edit Dec 11, thanks ChriSp and Lubos) gives
a factor of roughly α=1/137 to the diagram’s contribution to the overall probability. (There are
some subtleties here that are mentioned in the comments.) Thus it is usually fine to just take the
simplest diagrams and calculate those. The contribution from more complicated diagrams are
then very small corrections that are only important to calculate when experiments reach that level
of precision. For those with some calculus background, this should sound familiar: it is simply a
Taylor expansion. (In fact, most of physics is about making the right Tayor expansion.)

However, QCD defies this approximation. It turns out that the simplest diagrams do not give the
dominant contribution! It turns out that both the simple diagram and the complicated diagram
above give roughly the same contribution. One has to include many complicated diagrams to
obtain a good approximate calculation. And by “many,” I mean almost all of them… and
“almost all” of an infinite number of diagrams is quite a lot. For various reasons, these
complicated diagrams are very difficult to calculate and at the moment our normal approach is
useless.

There’s a lot of current research pushing this direction (e.g. so-called holographic techniques and
recent progress on scattering amplitudes), but let’s move on to what we can do.

QCD AND THE LATTICE


`Surely,’ said I, `surely that is something at my window lattice;
Let me see then, what thereat is, and this mystery explore –
— Edgar Allen Poe, “The Raven”

A different tool that we can use is called Lattice QCD. I can’t go into much detail about this
since it’s rather far from my area of expertise, but the idea is that instead using Feynman
diagrams to calculate processes perturbatively—i.e. only taking the simplest diagrams—we can
use computers to numerically solve for a related quantity. This related quantity is called
the partition function and is a mathematical object from which one can straightforwardly
calculate probability amplitudes. (I only mention the fancy name because it is completely
analogous to an object of the same name that one meets in thermodynamics.)

The point is that the lattice techniques are non-perturbative in the sense that we don’t calculate
individual diagrams, we simultaneously calculate all diagrams. The trade-off is that one has to
put spacetime on a lattice so that the calculations are actually done on a four-dimensional hyper-
cube. The accuracy of this approximation depend on the lattice size and spacing relative to the
physics that you want to study. (Engineers will be familiar with this idea from the use of Fourier
transforms.) As usual, a picture is worth a thousand words; suppose we wanted to study the
Mona Lisa:

The first image is the original. The second image comes from putting the image on a lattice, you
see that we lose details about small things. Because things with small wavelengths have high
energies, we call this an ultraviolet (UV) cutoff. The third image comes from having a smaller
canvas size so that we cannot see the big picture of the entire image. Because things with big
wavelengths have low energies, we call this an IR cutoff. The final image is meant to convey the
limitations imposed by the combination of the UV and IR cutoffs; in other words, the restrictions
from using a lattice of finite size and finite lattice spacing.

If you’re interested in only the broad features the Mona Lisa’s face, then the lattice depiction
above isn’t so bad. Of course, if you are a fine art critic, then the loss of small and large scale
information is unforgiveable. Currently, lattice techniques have a UV cutoff of around 3 GeV
and an IR cutoff of about 30 MeV; this makes them very useful for calculating information about
transitions between charm (mass = 1.2 GeV) and strange quarks (mass = 100 MeV).
TRANSLATING FROM THEORY TO EXPERIMENT (AND BACK)

When I was an undergraduate, I always used to be flummoxed that theorists would always draw
these deceptively simple looking Feynman diagrams on their chalkboards, while experimentalists
had very complicated plots and graphs to represent the same physics. Indeed, you can tell
whether a scientific paper or talk has been written by a theorist or an experimentalist based on
whether it includes more Feynman diagrams or histograms. (This seems to be changing a bit as
the theory community has made a concerted effort over the past decade to learn to the lingo of
the LHC. As Seth pointed out, this is an ongoing process.)

There’s an reason for this: experimental data is very different from writing down new models of
particle interactions. I encourage you to go check out the sample event displays
from CMS and ATLAS on the Symmetry Breaking blog for a fantastic and accessible discussion
of what it all means. I can imagine fellow bloggers Jim and Burton spending a lot of time
looking at similar event displays! (Or maybe not; I suspect that an actual analysis focuses more
on accumulated data over many events rather than individual events.) As a theorist, on the other

hand, I seem to be left with my chalkboard connecting squiggly lines to one another.

Once again, part of the reason why we speak such different languages is non-perturbativity. One
cannot take the straightforward Feynman diagram approach and use it when there is all sorts of
strongly-coupled gunk flying around. For example, here’s a diagram for electron–positron
scattering from Dieter Zeppenfeld’s PiTP 2005 lectures:

The part in black, which is labeled “hard scattering,” is what a theorist would draw. As a test of
your Feynman diagram see if you can “read” the following: This diagram represents an electron
and positron annihilating into a Z boson, which then decays into a top–anti-top pair. The brown
lines also show the subsequent decay of each top into a W and (anti-)bottom quark.
Great, that much we’ve learned from our previous posts. The big question is: what’s all that
other junk?!That, my friend, is the result of QCD. You can see that the pink lines are gluons
which are emitted from the final state quarks. These gluons can sprout off other gluons or quark–
anti-quark pairs. All of these quarks and gluons must then hadronize into color-neutral hadron
states, mostly mesons. These are shown as the grey blobs. These hadrons can in turn decay into
other hadrons, depicted by yellow blobs. Most of all of this happens before any of the particles
reach the detector. Needless to say, there are many, many similar diagrams which should all be
calculated to give an accurate prediction.

In fact, for the LHC it’s even more complicated since even the initial states are colored and so
they also spit off gluons (“hadronic junk”). Here’s a picture just to show how ridiculous these
process look at a particle-by-particle level:

Let me just remark that the two dark gray blobs are the incoming protons. The big red blob
represents all of the gluons that these protons emit. Note that the actual “hard interaction,” i.e.
the “core process” is gluon-gluon scattering. This is a bit of a subtle point, but at very high
energies, the actual point-like objects which are interacting are gluons, not the quarks that make
up the proton!

All of this hadronic junk ends up being sprayed through the experiments’ detectors. If the origin
of some of the hadronic junk comes from a high-energy colored particle (e.g. a quark that came
from the decay of a new heavy TeV-scale particle), then they are collimated into cones that are
pointing in roughly the same direction called a jet, (image from Gavin Salam’s 2010 lectures at
Cargese)
Some terminology: parton refers to either a quark or gluon, LO means “leading-order”, NLO
means “next-to-leading order.” The parton shower is the stage in which partons can radiate
more low-energy partons, which then confine into hadrons. Now one can start to see how to
connect our simple Feynman diagrams to the neat looking event reconstructions at the LHC:
(image from Gavin Salam’s lectures again)

Everything except for the black lines are examples of what one would actually read off of an
event display. This is meant to be a cross-section of the interaction point of the beamline. The
blue lines come from a tracking chamber, basically layers of silicon chips that detect the
passage of charged particles. The yellow and pink bars are readings from the calorimeters,
which tell how much energy is deposited into chunks of dense material. Note how ‘messy’ the
event looks experimentally: all of those hadrons obscure the so-called hard
scattering underlying event (edit Dec 11, thanks to ChriSp), which is what we draw with
Feynman diagrams.

So here’s the situation: theorists can calculate the “hard scattering” or “underlying event” (black
lines in the two diagrams above), but all of the QCD-induced stuff that happens after the hard
scattering is beyond our Feynman diagram techniques and cannot be calculated from first
principles. Fortunately, most of the non-peturbative effects can again be accounted for using
computers. The real question is given an underlying event (a Feynman diagram), how many
times will the final state particles turn into a range of different hadrons configurations. This time
one uses Monte-Carlo techniques where instead of calculating the probabilities of each hadronic
final state, the computer randomly generates these final states according to some pre-defined
probability distribution. If we run such a simulation over and over again, then we end up with a
simulated distribution of events which should match experiments relatively well.

One might wonder why this technique should work. It seems like we’re cheating—where did
these “pre-defined” probability distributions come from? Aren’t these what we want to calculate
in the first place? The answer is that these probability distributions come from experiments
themselves. This isn’t cheating since the experiments reflect data about low-energy physics. This
is well known territory that we really understand. In fact, everything in this business of hadronic
junk is low-energy physics. The whole point is that the only missing information is the high-
energy underlying event hard scattering (ed. Dec 11)—but fortunately that’s the part that
we can calculate! The fact this works is a straightforward result of “decoupling,” or the idea
that physics at different scales shouldn’t affect one another. (In this case physicists often say that
the hadronic part of the calculation “factorizes.”)

To summarize: theorists can calculate the underlying event hard scattering (ed. Dec 11) for their
favorite pet models of new physics. This is not the whole story, since it doesn’t reflect what’s
actually observed at a hadron collider. It’s not possible to calculate what happens next from first
principles, but fortunately this isn’t necessary, we can just use well-known probability
distributions to simulate many events and predict what the model of new physics would predict
in a large data set from an actual experiment. Now that we’re working our way into the LHC era,
clever theorists and experimentalists are working on new ways to go the other way around and
take the experimental signatures to try to recreate the underlying model.

As a kid I remember learning over and over again how a bill becomes a law. What we’ve shown
here is how a model of physics (a bunch of Feynman rules) becomes a prediction at a hadron
collider! (And along the way we’ve hopefully learned a lot about what Feynman diagrams are
and how we deal with physics that can’t be described by them.)

An Idiosyncratic Introduction to the Higgs

A different presentation of the Higgs

There have been several very clever attempts to explain the Higgs to a general audience using
analogies; one of my favorites is a CERN comic based on an explanation by David Miller.
Science-by-analogy, however, is a notoriously delicate tightrope to traverse. Instead, we’ll take a
different approach and jump straight into the physics. We can do this because we’ve already laid
down the ground work to use Feynman diagrams to describe particle interactions.
In the next few posts we’ll proceed as we did with the other particles of the Standard Model and
learn how to draw diagrams involving the Higgs. We’ll see what makes the Higgs special from
the diagrammatic point of view, and then gradually unpack the deeper ideas associated with it.
The approach will be idiosyncratic, but I think it is closer to the way particle physicists really
think about some of the big ideas in our field.

This first post we’ll start very innocently. We’ll present simplified Feynman rules for the Higgs
and then use them to discuss how we expect to produce the Higgs at the LHC. In follow-up posts
we’ll refine our Feynman rules to learn more about the nature of mass and the phenomenon
called electroweak symmetry breaking.

Feynman Rules (simplified)

First off, a dashed line represents the propagation of a Higgs boson:

You can already guess that there’s something different going


on since we haven’t seen this kind of line before. Previously, we drew matter particles
(fermions) as solid lines with arrows and force particles (gauge bosons) as wiggly lines. The
Higgs is indeed a boson, but it’s different from the gauge bosons that we’ve already met: the
photon, W, Z, and gluon. To understand this difference, let’s go into a little more depth on this:

 Gauge bosons, things which mediate “fundamental” forces, carry angular momentum,
or spin. Gauge bosons carry one unit of spin; roughly this means if you rotate a photon
by 360 degrees, it returns to the same quantum mechanical state.

 Fermions, matter particles, also carry angular momentum. However, unlike gauge
bosons, they carry only half a unit of spin: you have to rotate an electron by 720 degrees
to get the same quantum state. (Weird!)

 The Higgs boson is a scalar boson, which means it has no spin. You can rotate it by any
amount and it will be the same state. All scalar particles are bosons, but they don’t
mediate “fundamental” forces in the way that gauge bosons do.

This notion of spin is completely quantum mechanical, and it is a theorem that any particle with
whole number spin is a boson (“force particle”) and any particle with fractional spin is
a fermion (“matter particle”). It’s not worth dwelling too much about what kind of ‘force’ the
Higgs mediates—it turns out that there are much more interesting things afoot.

Now let’s ask how the Higgs interacts with other particles. There are two Feynman rules that we
can write down right away:
Here we see that the Higgs can interact with either a pair of fermions or a pair of gauge bosons.
This means, for example, that a Higgs can decay into an electron/positron pair (or, more likely, a
quark/anti-quark pair). For reasons that will become clear later, let’s say that the Higgs can
interact with any Standard Model particle with mass. Thus it does not interact with the photon or
gluon, and for argument’s sake we can ignore the interaction with the neutrino.

The interaction with fermions is something that we’re used to: it looks just like every other
fermion vertex we’ve written down: one fermion coming in, one fermion coming out, and some
kind of boson. This reflects the conservation of fermion number. We’ll see later that because the
Higgs is a scalar, there’s actually something sneaky happening here.

Finally, the Higgs also interacts with itself via a four-point interaction: (This is similar to the
four-gluon vertex of QCD.)

There are actually lots of subtleties that we’ve not mentioned and a few more Feynman rules to
throw in, but we’ll get to these in the next post when we will see what happens with the Higgs
gets a “vacuum expectation value”. Please, no comments yet about how I’m totally missing the
point… we’ll get to it all gradually, I promise.

Higgs Production

Thus far all we’ve been doing is laying the groundwork in preparation for a discussion of the
neat things that make the Higgs special. Even before we get into that stuff, though, we can use
what we’ve already learned to talk about how we hope to produce the Higgs at the LHC. This is
an exercise in drawing Feynman diagrams. (Review the old Feynman diagram posts if
necessary!)

The general problem is this: at the LHC, we’re smashing protons into one another. The protons
are each made up of a goop of quarks, antiquarks, and gluons. This is important: the protons
are more than just three quarks! As we mentioned before, protons are terribly non-perturbative
objects. Virtual (anti-)quarks and gluons are being produced and reabsorbed all over the place. It
turns out that the main processes that produce Higgs bosons from proton collisions comes from
the interaction of these virtual particles!

One of the main “production channels” at the LHC is the following gluon fusion diagram:

This is kind of a funny diagram because there’s a closed loop in the middle. (This makes it a very
quantum effect… and somewhat more tricky to actually calculate.) What’s happening is that a
gluon from one proton and a gluon from the other proton interact to form a Higgs. However,
because the gluons don’t directly interact with a Higgs, they have to do so through quarks. It
turns out that the top quark—which is heaviest—has the strongest interaction with the Higgs, so
the virtual quarks here are tops.

Another way to get a Higgs is associated production with a top pair. The diagram looks like
this:

Here gluons again produce a Higgs through top quarks. This time, however, a top quark and an
anti-top quark are also produced along with the Higgs. We can draw a similar diagram without
the gluons:
This is called vector fusion, because virtual W or Z bosons produce a Higgs. Note that we have
two quarks being produced as well.

Finally, there is associated production with a W or Z. As homework you can fill in the particle
labels assuming the final gauge boson is either W or Z:

There are other ways of producing a Higgs out of a proton-proton collision, but these are the
dominant processes. While we know a lot about the properties of a Standard Model Higgs, we
still don’t know its mass. It turns out that the relative rates of these processes depends on the
Higgs mass, as can be seen in the plot below (from the “Tevatron-for-LHC” report):
The
horizontal axis is the hypothetical HIggs mass, while the vertical axis measure the cross
section for Higgs production by the various labeled processes. For our purposes, the cross section
is basically the rate at which these processes occur. (Experimentally, we know that a Standard
Model Higgs should have a mass between about 115 GeV and 200 Gev.) We can see that the gg
→ h is the dominant production mechanism throughout the range of possible Higgs masses—but
this is only half of the story.
We don’t actually directly measure the Higgs in our detectors because it decays into lighter
Standard Model particles. The particular rate at which it decays to different final states
(“branching ratios”) are plotted above, image from CDF. This means we have to tell our
detectors to look for the decay products of the Higgs in addition to the extra stuff that comes out
of producing the Higgs in the first place. For example, in associated production with a top pair,
we have gg → tth. Each of the tops decay into a bquark, a lepton, and a neutrino (can you draw
the diagram showing this?), while the Higgs also decays—say, into a pair of b quarks. (For now
I’m not distinguishing quarks and anti-quarks.) This means that one channel we have to look for
is the rather cumbersome decay,

gg → tth →blν blν bb

Not only is this a lot of junk to look for in the final state (each of the b quarks hadronizes into a
jet), but there are all sorts of other Standard Model processes which give the same final state!
Thus if we simply counted the number of “four jets, two leptons, and missing energy
(neutrinos)” events, we wouldn’t only be counting Higgs production events, but also a bunch of
other background events which have nothing to do with the Higgs. One has to predict the rate of
these background events and subtract them from the experimental count. (Not to mention the
task of dealing with experimental uncertainties and possible mis-measurements!)

The punchline is that it can be very tricky to search for the Higgs and that this search is very
dependent on the Higgs mass. This is why we may have to wait a few years before the LHC has
enough data to say something definitive about the Higgs boson. (I’ve been somewhat terse here,
but my main point is to give a flavor of the Higgs search at the LHC rather than explain it in any
detail.)

As a single concrete example, consider the gluon fusion production channel, gg → h. This seems
nice since there’s no extra particles in the production process. However, from the plot above, we
can see that for relatively light masses (less than 140 GeV) the Higgs will want to decay
into b quarks. This is no good experimentally since the signal for this has hopelessly large
background from non-Higgs events.

In fact, rather counter intuitively, that one of the best ways to use gluon-fusion to search for a
light-mass Higgs is to look for instances where it decays into a pair of photons! This
is really weird since the Higgs doesn’t interact directly with photons, so this process must occur
through virtual quarks, just like the Higgs-gluon coupling above. As the branching ratio chart
above shows, this is a very rare process: the Higgs doesn’t want to decay into photons very often.
However, the upshot is that there aren’t many things in the Standard Model which can mimic this
“two photon” signal so that there is very little background. You can see that this stops working if
the Higgs is too heavy since the decay rate into photons shrinks very quickly.

Next time

In our next post we’ll introduce a completely new type of Feynman rule representing the Higgs
“vacuum expectation value.” In doing so we’ll sort out what we really mean when we say that a
particle has mass and continue our march towards the fascinating topic of electroweak symmetry
breaking (“the Higgs mechanism”).

Higgs and the vacuum: Viva la “vev”

Hello everyone! Recently we’ve been looking at the Feynman rules for the Higgs boson. Last
time I posted, we started to make a very suggestive connection between the Higgs and the origin
of mass. We noted that the Higgs has a special trick up it’s sleeve: it has a Feynman rule that
allows a Higgs line to terminate:

This allowed us to draw diagrams with two fermions or two


gauge bosons attached to a terminated Higgs:
We made the bold claim that these diagrams should be interpreted as masses for the particles
attached to the Higgs. We’ll explore this interpretation in a later post, but for now let’s better
understand why we should have this odd Feynman rule and what it means.

Quantum Fields Forever

Before we get into that, though, we need to get back to one of the fundamental ideas of quantum
physics: wave-particle duality. Douglas Hofstadter’s famous ambrigram summarizes the well-
known statement in pop-science:

Amigram by Douglas Hofstadter, image used according to the Creative Commons Attribution-
Share Alike 3.0 Unported license.

Wave-particle duality is one of those well-known buzz words of quantum mechanics. Is light a
particle or a wave? Is an electron a particle or a wave? If these things are waves, then what are
they waves of?

In high energy physics we’re usually interested in small things that move very quickly, so we use
the framework of quantum field theory (QFT) which is the marriage of quantum mechanics
(which describes “small” things) and special relativity (which describes “fast” things). The
quantum ‘waves’ are waves in the quantum field associated with a particle. The loose
interpretation of the field is the probability that you might find a particle there.

A slightly more technical explanation: the whole framework of QFT is based on the idea of
causality. You can’t have an effect happen before the cause, but special relativity messes with
our notion of before-and-after. Thus we impose that particle interactions must be local in
spacetime; the vertices in our Feynman rules really represent a specific place at a specific time.
The objects which we wish to describe are honest-to-goodness particles, but a local description
of quantum particles is naturally packaged in terms of fields. For a nice discussion (at the level of
advanced undergrads), see the first half hour of this lecture by N. Arkani-Hamed at Perimeter.

So the quantum field is a mathematical object which we construct which tells us how likely it is
that there’s a particle at each point in space and time. Most of the time the quantum field is pretty
much zero: the vacuum of space is more-or-less empty. We can imagine a particle as a ripple in
the quantum field, which former US LHC blogger Sue Ann Koay very nicely depicted thusly:

Sue Ann Koay's depiction of a quantum field. Ripples in the field should be interpreted as
particles. Here we have two particles interacting. (For experts: Sue pointed out the ISR in the
image.)

Sometimes ripples can excite others ripples (perhaps in other quantum fields), this is precisely
what’s happening when we draw a Feynman diagram that describes the interaction of different
particles.

The vacuum and the ‘Higgs phase’

Now we get to the idea of the vacuum—space when there isn’t any stuff in it. Usually when you
think of the vacuum of empty space you’re supposed to think of nothingness. It turns out that the
vacuum is a rather busy place on small scales because of quantum fluctuations: there are virtual
particle–anti-particle pairs that keep popping into existence and then annihilating. Further still,
vacuum is also filled with cosmic microwave background radiation at 2.725 Kelvin. But for now
we’re going to ignore both of these effects. It turns out that there’s something much more
surprising about the vacuum:

It’s full of Higgs bosons.

The quantum field for normal particle species like electrons or quarks is zero everywhere except
where there are particles moving around. Particles are wiggles on top of this zero value. The
Higgs is different because the value of its quantum field in the vacuum is not zero. We say that it
has a vacuum expectation value, or “vev” for short. It is precisely this Higgs vev which is
represented by the crossed out Higgs line in our Feynman rules.
A loose interpretation for the Higgs vev is a background probability for there to be a Higgs
boson at any given point in spacetime. These “background” Higgs bosons carry no momentum,
but they can interact with other particles as we saw above:

The cross means that instead of a ‘physical’ Higgs particle, the dashed line corresponds to an
interaction with one of these background Higgses. In this sense, we are swimming in a sea of
Higgs. Our interactions with the Higgs are what give us mass, though this statement will perhaps
only make sense after we spend some time in a later post understanding what mass really is.

A good question to ask is why the Higgs has a vacuum expectation value. This is the result of
something called electroweak symmetry breaking and is related to the unification of the
electromagnetic force and the weak force, i.e. somehow the Higgs is part of a broader story about
unification of the fundamental forces.

Often people will say that the universe is in a ‘Higgs phase,’ a phrase which draws on very
elegant connections between the quantum field theory of particles and the statistical field theory
of condensed matter systems. Just as we can discuss phase transitions between liquid and gas
states (or more complicated phases), we can also discuss how the universe underwent an
electroweak phase transition which led to the Higgs vev that lends masses to our favorite
particles.

Next time…

When we continue our story of the Higgs, we’ll start to better understand the relation of the
Higgs vev with the mass of the other Standard Model particles and will learn more about
electroweak symmetry breaking.

elicity, Chirality, Mass, and the Higgs

We’ve been discussing the Higgs (its interactions, its role in particle mass, and its vacuum
expectation value) as part of our ongoing series on understanding the Standard Model with
Feynman diagrams. Now I’d like to take a post to discuss a very subtle feature of the Standard
Model: its chiral structure and the meaning of “mass.” This post is a little bit different in
character from the others, but it goes over some very subtle features of particle physics and I
would really like to explain them carefully because they’re important for understanding the
entire scaffolding of the Standard Model.

My goal is to explain the sense in which the Standard Model is “chiral” and what that means. In
order to do this, we’ll first learn about a related idea, helicity, which is related to a particle’s
spin. We’ll then use this as an intuitive step to understanding the more abstract notion
of chirality, and then see how masses affect chiral theories and what this all has to do with the
Higgs.

Helicity

Fact: every matter particle (electrons, quarks, etc.) is spinning, i.e. each matter particle carries
some intrinsic angular momentum.

Let me make the caveat that this spin is an inherently quantum mechanical property of
fundamental particles! There’s really no classical sense in which there’s a little sphere spinning
like a top. Nevertheless, this turns out to be a useful cartoon picture of what’s going on:

This is our spinning particle. The red arrow indicates the direction of the particle’s spin. The gray
arrow indicates the direction that the particle is moving. I’ve drawn a face on the particle just to
show it spinning.

The red arrow (indicating spin) and the gray arrow (indicating direction of motion) defines an
orientation, or a handedness. The particular particle above is “right-handed” because it’s the
same orientation as your right hand: if your thumb points in the direction of the gray arrow, then
your fingers wrap in the direction of the red arrow. Physicists call this “handedness”
the helicity of a particle.

To be clear, we can also draw the right-handed particle moving in the opposite direction (to the
left):
Note that the direction of the spin (the red arrow) also had to change. You can confirm that if you
point your thumb in the opposite direction, your fingers will also wrap in the opposite direction.

Sounds good? Okay, now we can also imagine a particle that is left-handed (or “left helicity”).
For reference here’s a depiction of a left-handed particle moving in each direction; to help
distinguish between left- and right-handed spins, I’ve given left-handed particles a blue arrow:

[Confirm that these two particles are different from the red-arrowed particles!]

An observation: note that if you only flip the direction of the gray arrow, you end up with a
particle with the opposite handedness. This is precisely the reason why the person staring back at
you in the mirror is left-handed (if you are right-handed)!

Thus far we’re restricting ourselves to matter particles (fermions). There’s a similar story for
force particles (gauge bosons), but there’s an additional twist that will deserve special attention.
The Higgs boson is another special case since it doesn’t have spin, but this actually ties into the
gauge boson story.

Once we specify that we have a particular type of fermion, say an electron, we automatically
have a left-helicity and a right-helicity version.

Helicity, Relativity, and Mass

Now let’s start to think about the meaning of mass. There are a lot of ways to think about mass.
For example, it is perhaps most intuitive to associate mass with how ‘heavy’ a particle is. We’ll
take a different point of view that is inspired by special relativity.

A massless particle (like the photon) travels at the speed of light and you can never catch up to
it. There is no “rest frame” in which a massless particle is at rest. The analogy for this is driving
on the freeway: if you are driving at the same speed as the car in the lane next to you, then it
appears as if the car next to you is not moving (relative to you). Just replace the car with a
particle.

On the other hand, a massive particle travels at less than the speed of light so that you can (in
principle) match its velocity so that the particle is at rest relative to you. In fact, you can
move faster than a massive particle so that it looks like the particle is traveling in
the opposite direction (this flips the direction of the gray arrow). Note that the direction of its
spin (the red arrow) does not change! However, we already noted that flipping only the particle’s
direction—and not its spin—changes the particle’s helicity:

Here we’ve drawn the particle with a blue arrow because it has gone from being right-handed to
left-handed. Clearly this is the same particle: all that we’ve done is gone to a different reference
frame and principles of special relativity say that any reference frame is valid.

Okay, so here’s the point so far: mass is a something that tells us whether or not helicity is an
“intrinsic” property of the particle. If a particle is massless, then its helicity has a fixed value in
all reference frames. On the other hand, if a particle has any mass, then helicity is not an intrinsic
property since different observers (in valid reference frames) can measure different values for the
helicity (left- or right-helicity). So even though helicity is something which is easy to visualize, it
is not a “fundamental” property of most particles.
Now a good question to ask is: Is there some property of a particle related to the helicity
which isintrinsic to the particle? In other words, is there some property which

1. is equivalent to helicity in the massless limit

2. is something which all observers in valid reference frames would measure to be the same
for a given particle.

The good news is that such a property exists, it is called chirality. The bad news is that it’s a bit
more abstract. However, this is where a lot of the subtlety of the Standard Model lives, and I
think it’s best to just go through it carefully.

Chirality

Chirality and helicity are very closely related ideas. Just as we say that a particle can have left- or
right-handed helicity, we also say that a particle can have left- or right-handed chirality. As we
said above, for massless particles the chirality and helicity are the same. A massless left-chiral
particle also has left-helicity.

However, a massive particle has a specific chirality. A massive left-chiral particle may have
either left- or right-helicity depending on your reference frame relative to the particle. In all
reference frames the particle will still be left-chiral, no matter what helicity it is.

Unfortunately, chirality is a bit trickier to define. It is an inherently quantum mechanical sense in


which a particle is left- or right-handed. For now let us focus on fermions, which are “spin one-
half.” Recall that this means that if you rotate an electron by 360 degrees, you don’t get the same
quantum mechanical state: you get the same state up to a minus sign! This minus sign is related
to quantum interference. A fermion’s chirality tells you how it gets to this minus sign in terms of
a complex number:
What happens when you rotate a left- vs right-chiral fermion 360 degree about its direction of
motion. Both particles pick up a -1, but the left-chiral fermion goes one way around the complex
plane, while the right-chiral fermion goes the other way. The circle on the right represents the
complex phase of the particle’s quantum state; as we rotate a particle, the value of the phase
moves along the circle. Rotating the particle 360 degrees only brings you halfway around the
circle in a direction that depends on the chirality of the fermion.

The physical meaning of this is the phase of the particle’s wavefunction. When you rotate a
fermion, its quantum wavefunction is shifted in a way that depends on the fermion’s chirality:

Rotating a fermion shifts its quantum wavefunction. Left- and right-chiral fermions are shifted in
opposite directions. This is a purely quantum phenomenon.

We don’t have to worry too much about the meaning of this quantum mechanical phase shift.
The point is that chirality is related in a “deep” way to the particle’s inherent quantum
properties. We’ll see below that this notion of chirality has more dramatic effects when we
introduce mass.

Some technical remarks: The broad procedure being outlined in the last two sections can be
understood in terms of group theory. What we claim is that massive and massless particles
transform under different [unitary] representations of the Poincaré group. The notion of fermion
chirality refers to the two types of spin-1/2 representations of the Poincaré group. In the brief
discussion above, I tried to explain the difference by looking at the effect of a rotation about
the z-axis, which is generated by ±σ3/2.

The take home message here is that particles with different chiralities are really different
particles. If we have a particle with left-handed helicity, then we know that there should also be a
version of the particle with right-handed helicity. On the other hand, a particle with left-
handed chiralityneedn’t have a right-chiral partner. (But it will certainly furnish both helicities
either way.) Bear with me on this, because this is really where the magic of the Higgs shows up
in the Standard Model.

Chiral theories
[6/20/11: the following 2 paragraphs were edited and augmented slightly for better clarity.
Thanks to Bjorn and Jack C. for comments. 4/8/17: corrected “right-chiral positron” to “left-
chiral positron” and analogously for anti-positrons; further clarification to text and images;
thanks to Martha Lindeman, Ph.D.]

One of the funny features of the Standard Model is that it is a chiral theory, which means that
left-chiral and right-chiral particles behave differently. In particular, the W bosons will only talk
to electrons (left-chiral electrons and right-chiral anti-electrons) and refuses to talk to positrons
(left-chiral positrons and right-chiral anti-positrons). You should stop and think about this for a
moment: nature discriminates between left- and right-chiral particles! (Of course, biologists are
already very familiar with this from the ‘chirality’ of amino acids.)

Note that Nature is still, in some sense, symmetric with respect to left- and right-helicity. In the
case where everything is massless, the chirality and helicity of a particle are the same.
The W will couple to both a left- and right-helicity particles: the electron and anti-electron.
However, it still ignores the positrons. In other words, the W will couple to a charge -1 left-
handed particle (the electron), but does not couple to a charge -1 right-handed particle (the anti-
positron). This is a very subtle point!

Technical remark: the difference between chirality and helicity is one of the very subtle points
when one is first learning field theory. The mathematical difference can be seen just by looking
at the form of the helicity and chirality operators. Intuitively, helicity is something which can be
directly measured (by looking at angular momentum) whereas chirality is associated with the
transformation under the Lorentz group (e.g. the quantum mechanical phase under a rotation).

In order to really drive this point home, let me reintroduce two particles to you: the electron and
the “anti-positron.” We often say that the positron is the anti-partner of the electron, so shouldn’t
these two particles be the same? No! The real story is actually more subtle—though some of this
depends on what people mean by ‘positron,’ here we are making a useful, if unconventional,
definition. The electron is a left-chiral particle while the positron is a right-chiral particle. Both
have electric charge -1, but they are two completely different particles.
Electrons (left-chiral) and anti-positrons (right-chiral) have the same electric charge but are two
completely different particles, as evidenced by the positron’s mustache.

How different are these particles? The electron can couple to a neutrino through the W-boson,
while the anti-positron cannot. Why does the W only talk to the (left-chiral) electron? That’s just
the way the Standard Model is constructed; the left-chiral electron is charged under the weak
force whereas the right-chiral anti-positron is not. So let us be clear: the electron and the anti-
positron are not the same particle! Even though they both have the same charge, they have
different chirality and the electron can talk to a W, whereas the anti-positron cannot.

For now let us assume that all of these particles are massless so that these chirality states can be
identified with their helicity states. Further, at this stage, the electron has its own anti-particle (an
“anti-electron”) which has right-chirality which couples to the W boson. The anti-positron also
has a different antiparticle which we call the positron (the same as an “anti-anti-positron”)
and has left-chirality but does not couple to the W boson. We thus have a total of four particles
(plus the four with opposite helicities):

The electron, anti-electron, anti-positron, and positron.

Technical remark: the left- & right-helicity electrons and left- & right-helicity anti-
positrons are the four components of the Dirac spinor for the object which we normally call the
electron (in the mass basis). Similarly, the left- & right-helicity anti-electrons and left- & right-
helicity positrons for the conjugate Dirac spinor which represents what we normally call the
positron (in the mass basis).

Important summary: [6/20/11: added this section to address some lingering confusion; thanks
to David and James from CV, and Steve. 6/29: Thanks to Rainer for pointing out a mistake in 3
and 4 below (‘left’ and ‘right’ were swapped).] We’re bending the usual nomenclature for
pedagogical reasons—the things which we are calling the “electron” and “positron” (and their
anti-partners) are not the “physical” electron in, say, the Hydrogen atom. We will see below how
these two ideas are connected. Thus far, the key point is that there are fourdistinct particles:

1. Electron: left-chiral, charge -1, can interact with the W

2. Anti-electron: right-chiral, charge +1, can interact with the W

3. Positron: left-chiral, charge +1, cannot interact with the W


4. Anti-positron: right-chiral, charge -1, cannot interact with the W.

We’re using names “electron” and “positron” to distinguish between the particles which couple
to the W and those that don’t. The conventional language in particle physics is to call these the
left-handed (chirality) electron and the right-handed (chirality) electron. But I wanted to use a
different notation to emphasize that these are not related to one another by parity (space
inversion, or reversing angular momentum).

Masses mix different particles!

Now here’s the magical step: masses cause different particles to “mix” with one another.

Recall that we explained that mass could be understood as a particle “bumping up against the
Higgs boson’s vacuum expectation value (vev).” We drew crosses in the fermion lines of
Feynman diagrams to represent a particle interacting with the Higgs vev, where each cross is
really a truncated Higgs line. Let us now show explicitly what particles are appearing in these
diagrams:

An “electron” propagating in space and interacting with the Higgs field. Note that the Higgs-
induced mass term connects an electron with an anti-positron. This means that the two types of
particles are exhibiting quantum mixing.

[6/25: this paragraph added for clarity] Note that in this picture the blue arrow
represents helicity(it is conserved), whereas the mustache (or non-mustache) represents chirality.
The mass insertions flip chirality, but maintain helicity.

This is very important; two completely different particles (the electron and the anti-positron) are
swapping back and forth. What does this mean? The physical thing which is propagating through
space is a mixture of the two particles. When you observe the particle at one point, it may be an
electron, but if you observe it a moment later, the very same particle might manifest itself as an
anti-positron! This should sound very familiar, it’s the exact same story as neutrino mixing (or,
similarly, meson mixing).

Let us call this propagating particle is a “physical electron.” The mass-basis-electron can either
be an electron or an anti-positron when you observe it; it is a quantum mixture of both.
The W boson only interacts with the “physical electron” through its electron component and does
not interact with the anti-positron component. Similarly, we can define a “physical positron”
which is the mixture of the positron and anti-electron. Now I need to clarify the language a bit.
When people usually refer to an electron, what they really mean is the mass-basis-electron, not
the “electron which interacts with W.” It’s easiest to see this as a picture:

The “physical electron” (what most people mean when they say “electron”) is a combination of
an electron and an anti-positron. Note that the electron and the anti-positron have different
interactions (e.g. the electron can interact with the W boson); the physical electron inherits the
interactions of both particles.

Note that we can now say that the “physical electron” and the “physical positron” are
antiparticles of one another. This is clear since the two particles which combine to make up a
physical electron are the antiparticles of the two particles which combine to make up the physical
positron. Further, let me pause to remark that in all of the above discussion, one could have
replaced the electron and positron with any other Standard Model matter particle (except the
neutrino, see below). [The electron and positron are handy examples because the positron has a
name other than anti-electron, which would have introduced language ambiguities.]

Technical remarks: To match to the parlance used in particle physics:

1. The “electron” (interacts with the W) is called eL, or the left-chiral electron

2. The “anti-positron” (does not interact with the W) is called eR, or the right-chiral
electron. [6/25: corrected and updated, thanks to those who left comments about
this] Note that I very carefully said that this is a right-chiral electron, not a right-
helicity electron. In order to conserve angular momentum, the helicities of
the eL and eR have to match. This means that one of these particles has opposite helicity
and chirality—and this is the whole point of distinguishing helicity from chirality!

3. The “physical electron” is usually just called the electron, e, or mass-basis electron

The analogy to flavor mixing should be taken literally. These are different particles that can
propagate into one another in exactly the same way that different flavors are different particles
that propagate into one another. Note that the mixing angle is controlled by the ratio of the
energy to the mass and is 45 degrees in the non-relativistic limit. [6/22: thanks to Rainer P. for
correcting me on this.] Also, the “physical electron” now contains twice the physical degrees of
freedom as the electron and anti-positron. This is just the observation that a Dirac mass combines
two 2-component Weyl spinors into a 4-component Dirac spinor.

When one first learns quantum field theory, one usually glosses over all of these details because
one can work directly in the mass basis where all fermions are Dirac spinors and all mass
insertions are re-summed in the propagators. However, the chiral structure of the Standard Model
is telling us that the underlying theory is written in terms of two-component [chiral] Weyl
spinors and the Higgs induces the mixing into Dirac spinors. For those that want to learn the two-
component formalism in gory detail, I strongly recommend the recent review by Dreiner, Haber,
and Martin.

What this all has to do with the Higgs

We have now learned that masses are responsible for mixing between different types of particles.
The mass terms combine two a priori particles (electron and anti-positron) into a single particle
(physical electron). [See a very old post where I tried—I think unsuccessfully—to convey similar
ideas.] The reason why we’ve gone through this entire rigmarole is to say that ordinarily, two
unrelated particles don’t want to be mixed up into one another.

The reason for this is that particles can only mix if they carry the same quantum properties.
You’ll note, for example, that the electron and the anti-positron both had the same electric charge
(-1). It would have been impossible for the electron and anti-electron to mix because they have
different electric charges. However, the electron carries a weak charge because it couples to
the W boson, whereas the anti-positron carries no weak charge. Thus these two particles
should not be able to mix. In highfalutin language, one might say that this mass term is
prohibited by “gauge invariance,” where the word “gauge” refers to the W as a gauge boson. This
is a consequence of the Standard Model being a chiral theory.

The reason why this unlikely mixing is allowed is because of the Higgs vev. The Higgs carries
weak charge. When it obtains a vacuum expectation value, it “breaks” the conservation of weak
charge and allows the electron to mix with the anti-positron, even though they have different
weak charges. Or, in other words, the vacuum expectation value of the Higgs “soaks up” the
difference in weak charge between the electron and anti-positron.

So now the mystery of the Higgs boson continues. First we said that the Higgs somehow gives
particle masses. We then said that these masses are generated by the Higgs vacuum expectation
value. In this post we took a detour to explain what this mass really does and got a glimpse of
why the Higgs vev was necessary to allow this mass. The next step is to finally address how this
Higgs managed to obtain a vacuum expectation value, and what it means that it “breaks” weak
charge. This phenomenon is called electroweak symmetry breaking, and is one of the primary
motivations for theories of new physics beyond the Standard Model.

Addendum: Majorana masses


Okay, this is somewhat outside of our main discussion, but I feel obligated to mention it. The
kind of fermion mass that we discussed above is called a Dirac mass. This is a type of mass that
connects two different particles (electron and anti-positron). It is also possible to have a mass
that connects two of the same kind of particle, this is called a Majorana mass. This type of mass
is forbidden for particles that have any type of charge. For example, an electron and an anti-
electron cannot mix because they have opposite electric charge, as we discussed above. There is,
however, one type of matter particle in the Standard Model which does not carry any charge: the
neutrino! (Neutrinos do carry weak charge, but this is “soaked up” by the Higgs vev.)

Within the Standard Model, Majorana masses are special for neutrinos. They mix neutrinos with
anti-neutrinos so that the “physical neutrino” is its own antiparticle. (In fancy language, we’d say
the neutrino is a Majorana fermion, or is described by a Weyl spinor rather than a Dirac spinor.)
It is also possible for the neutrino to have both a Majorana and a Dirac mass. (The latter would
require additional “mustached” neutrinos to play the role of the positron.) This would have some
interesting consequences. As we suggested above, the Dirac mass is associated with the non-
conservation of weak charge due to the Higgs, thus Dirac masses are typically “small.” (Nature
doesn’t like it when things which ought to be conserved are not.) Majorana masses, on the other
hand, do not cause any charge non-conservation and can be arbitrarily large. The “see-saw”
between these two masses can lead to a natural explanation for why neutrinos are so much lighter
than the other Standard Model fermions, though for the moment this is a conjecture which is
outside of the range of present experiments.

The Birds and the Bs

`Yesterday marked the beginning of the HEP summer conference season with EPS-HEP 2011,
which is particularly exciting since the LHC now has enough luminosity (accumulated data) to
start seeing hints of new physics. As Ken pointed out, the Tevatron’s new lower bound on the
Bs → μμ decay rate seemed to be a harbinger of things to come (Experts can check out
the official paper, the CDF public page, and the excellent summaries by Tommaso
Dorigo and Jester.).

Somewhat unfortunately, the first LHCb results on this process do not confirm the CDF excess,
though they are not yet mutually exclusive. Instead of delving too much into this particular
result, I’d like to give some background to motivate why it’s interesting to those of us looking
for new physics. This requires a lesson in “the birds and the Bs”—of course, by this I mean B
mesons and the so-called ‘penguin’ diagrams.

THE BS MESON: WHY IT’S SPECIAL


It's a terrible pun, I know.

A Bs meson is a bound state of a bottom anti-quark and strange quark; it’s sort of like a
“molecule” of quarks. There are all sorts of mesons that one could imagine by sticking together
different quarks and anti-quarks, but the Bs meson and it’s lighter cousin, the Bd meson, are
particularly interesting characters in the spectrum of all possible mesons.

The reason is that both the Bs and the Bd are neutral particles, and it turns out that they mix
quantum mechanically with their antiparticles, which we call the Bs and Bd. This mixing is the
exact same kind of flavor phenomenon that we described when we mentioned “Neapolitan”
neutrinos and is analogous to the mixing of chiralities in a massive fermion. Recall that
properties like “bottom-ness” or “strangeness” are referred to as flavor. Going from a Bs to
a Bs changes the “number of bottom quarks” from -1 to +1 and the “number of strange quarks”
from +1 to -1, so such effects are called flavor-changing.

To help clarify things, here’s an example diagram that encodes this quantum mixing:

The ui refers to any up-type quark.

Any neutral meson can mix—or “oscillate”—into its antiparticle, but the B mesons are special
because of their lifetime. Recall that mesons are unstable and decay, so unlike neutrinos, we
can’t just wait for a while to see if they oscillate into something interesting. Some mesons live
for too long and their oscillation phenomena get ‘washed out’ before we get to observe them.
Other mesons don’t live long enough and decay before they have a chance to oscillate at all. But
B mesons—oh, wonderful Goldilocks B mesons—they have a lifetime and oscillation time that
are roughly of the same magnitude. This means that by measuring their decays and relative
decay rates we can learn about how these mesons mix, i.e. we can learn about the underlying
flavor structure of the Standard Model.

Historical remark: The Bd meson is special for another reason: by a coincidence, we can
produce them rather copiously. The reason is that the Bd meson mass just happens to be just
under half of the mass of the Upsilon 4S particle, ϒ(4S), which just happens to decay into a Bd–
Bd pair. Thus, by the power of resonances, we can collide electrons and positrons to produce lots
of upsilons, which then decay in to lots of B mesons. For the past decade flavor physics focused
around these ‘B factories,’ mainly the BaBar detector at SLAC and Belle in Japan. BaBar has
since been retired, while Belle is being upgraded to “Super Belle.” For the meanwhile, the
current torch-bearer for B-physics is LHCb.

THE CDF AND LHCB RESULTS: BS → MU MU

It turns out that there are interesting flavor-changing effects even without considering meson
mixing, but rather in the decay of the B meson itself. For example, we can modify the previous
diagram to consider the decay of a Bs meson into a muon/anti-muon pair:

This is still a flavor-changing decay since the net strangeness (+1) and bottom-ness (-1) is not
preserved; but note that the lepton flavor is conserved since the muon/anti-muon pair have no net
muon number. (As an exercise: try drawing the other diagrams that contribute; the trick is that
you need W bosons to change flavor.) You could also replace muons by electrons or taus, but
those decays are much harder to detect experimentally. As a rule of thumb muons are really nice
final state particles since they make it all the way through the detector and one has a decent shot
at getting good momentum measurements.

It turns out that this decay is extremely rare. For the Bs meson, the Standard Model predicts a
dimuon branching ratio of around 3 × 10-9, which means that a Bs will only decay into two
muons 0.0000003% of the time… clearly in order to accurately measure the actual rate one needs
to produce a lot of B mesons.

In fact, until recently, we simply did not have enough observed B meson decays to even estimate
the true dimuon decay rate. The ‘B factories’ of the past decade were only able to put upper
limits on this rate. In fact, this decay is one of the main motivations for LHCb, which was
designed to be the first experiment that would be sensitive enough to probe the Standard Model
decay rate. (This means that if the decay rate is at least at the Standard Model rate, then LHCb
will see it.)

The exciting news from CDF last week was that—for the first time—they appeared to have been
able to set a lower bound on the dimuon decay rate of the Bs meson. (The Bd meson has a smaller
decay rate and CDF was unable to set a lower bound.) The lower bound is still statistically
consistent with the Standard Model rate, but the suggested (‘central value’) rate was 1.8 × 10-8. If
this is true, then it would be a fairly strong signal for new physics beyond the Standard Model.
The 90% confidence level range from CDF is:

4.6 × 10-9 < BR(Bs → μ+μ–) < 3.9 × 10-8.

Unfortunately, today’s new result from LHCb didn’t detect an excess with which it could set a
lower bound and could only set a 90% confidence upper bound,

BR(Bs → μ+μ–) < 1.3 × 10-8.

This goes down to 1.2 × 10-8 when including 2010 data. The bounds are not yet at odds with one
another, but many people were hoping that LHCb would have been able to confirm the CDF
excess in dimuon events. The analyses of the two experiments seem to be fairly similar, so there
isn’t too much wiggle room to think that the different results just come from having different
experiments.

More data will clarify the situation; LHCb should accumulate enough data to prove branching
ratios down to the Standard Model prediction of 3 × 10-9. Unfortunately CDF will not be able to
reach that sensitivity.

NEW PHYSICS IN LOOPS

Now that we’re up to date with the experimental status of Bs → μμ, let’s figure out why it’s so
interesting from a theoretical point of view. One thing you might have noticed from the
“box” Feynman diagrams above is that they involve a closed loop. An interesting thing about
closed loops in Feynman diagrams is that they can probe physics at much higher energies than
one would naively expect.

The reason for this is that the particles running in the loop do not have their momenta fixed in
terms of the momenta of the external particles. You can see this for yourself by assigning
momenta (call them p1, p2, … , etc.) to each particle line and (following the usual Feynman
rules) impose momentum conservation at each vertex. You’ll find that there is an unconstrained
momentum that goes around the loop. Because this momentum is unspecified, the laws of
quantum physics say that one must add together the contributions from all possible momenta.
Thus it turns out that even though the Bs meson mass is around 5 GeV, the dimuon decay is
sensitive to particles that are a hundred times heavier.
Note that unlike other processes where we study new physics by directly producing it and
watching it decay, in low-energy loop diagrams one only intuits the presence of new particles
through their virtual effects (quantum interference). I’ll leave the details for another time, but
here are a few facts that you can assume for now:

1. Loop diagrams can be sensitive to new heavy particles through quantum interference.

2. Processes which only occur through loop diagrams are often suppressed. (This is partly
why the Standard Model branching ratio for Bs → μμ is so small.)

3. In the Standard Model, all flavor-changing neutral currents (FCNC)—i.e. all flavor-
changing processes whose intermediate states carry no net electric charge—only occur at
loop level. (Recall that the electrically-charged W bosons can change flavor, but the
electrically neutral Z bosons cannot. Similarly, note that there is no way to draw a Bs →
μμ diagram in the Standard Model without including a loop.)

4. Thus, processes with a flavor-changing neutral current (such as Bs → μμ) are fruitful
places to look for new physics effects that only show up at loop level. If there were a
non-loop level (“tree level”) contribution from the Standard Model, then the loop-induced
new physics effects would tend to be drowned out because they are only small
corrections to the tree-level result. However, since there are no FCNCs in the Standard
Model, the new physics contributions have a ‘fighting change’ at having a big effect
relative to the Standard Model result.

5. Semi-technical remark, for experts: indeed, for Bs → μμ the Standard Model diagrams
are additionally suppressed by a GIM suppression (as is the case for FCNCs) as well as
helicity suppression (the B meson is a pseudoscalar, so the final states require a muon
mass insertion).

So the punchline is that Bs → μμ is a really fertile place to hope to see some deviation from the
Standard Model branching ratio due to new physics.

INTRODUCING THE PENGUIN

I would be remiss if I didn’t mention the “penguin diagram” and its role in physics. You can
learn about the penguin’s silly etymology in its Wikipedia article; suffice it for me to ‘wow’ you
with a picture of an autographed paper from one of the penguin’s progenitors:
A copy of the original "penguin" paper, autographed by John Ellis.

The main idea is that penguin diagrams are flavor-changing loops that involve two fermions and
a neutral gauge boson. For example, the b→s penguin takes the form (no, it doesn’t look much
like a penguin)

You should have guessed that in the Standard Model, the wiggly line on top has to be a W boson
in order for the fermion line to change flavors. The photon could also be a Z boson, a gluon, or
even a Higgs boson. If we allow the boson to decay into a pair of muons, we obtain a diagram
that contributes to Bs → μμ.

Some intuition for why the penguin takes this particular form: as mentioned above, any flavor-
changing neutral transition in the Standard Model requires a loop. So we start by drawing a
diagram with a W loop. This is fine, but because the b quark is so much heavier than the s quark,
the diagram does not conserve energy. We need to have a third particle which carries away the
difference in energy between the b and the s, so we allow the loop to emit a gauge boson. And
thus we have the diagram above.

Thus, in addition to the box diagrams above, there are penguin diagrams which contribute to
Bs → μμ. As a nice ‘homework’ exercise, you can try drawing all of the penguins that contribute
to this process in the Standard Model. (Most of the work is relabeling diagrams for different
internal states.)

[Remark, 6/23: my colleague Monika points out that it’s ironic that I drew the b, s, photon
penguin since this penguin doesn’t actually contribute to the dimuon decay! (For experts: the
reason is the Ward identity.) ]

SUPERSYMMETRY AND THE BS → MU MU PENGUIN

Finally, I’d like to give an example of a new physics scenario where we would expect that
penguins containing new particles give a large contribution to the Bs → μμ branching ratio. It
turns out that this happens quite often in models of supersymmetry or, more generally, ‘two
Higgs doublet models.’

If neither of those words mean anything to you, then all you have to know is that these models
have not just one, but two independent Higgs particles which obtain separate vacuum expectation
values (vevs). The punchline is that there is a free parameter in such theories called tan β which
measures the ratio of the two vevs, and that for large values of tan β, the Bs → μμ branching ratio
goes like (tan β)6 … which can be quite large and can dwarf the Standard Model contribution.

Added 6/23, because I couldn't help it: a supersymmetric penguin. Corny image from one of my
talks.

[What follows is mostly for ‘experts,’ my apologies.]


On a slightly more technical note, it’s not often well explained why this branching ratio goes like
the sixthpower of tan β, so I did want to point this out for anyone who was curious. There are
three sources of tan β in the amplitude; these all appear in the neutral Higgs diagram:

Each blue dot is a factor of tan β. The Yukawa couplings at each Higgs vertex goes like the
fermion mass divided by the Higgs vev. For the down-type quarks and leptons, this gives a factor
of m/v ~ 1/cos β ~ tan β for large tan β. An additional factor of comes from the mixing between
the s and b quarks, which also goes like the Yukawa coupling. (This is the blue dot on
the s quark leg.) Hence one has three powers of tan β in the amplitude, and thus six powers of tan
β in the branching ratio.

OUTLOOK

While the LHCb result was somewhat sobering, we can still cross our fingers and hope that there
is still an excess to be discovered in the near future. The LHC shuts down for repairs at the end
of next year; this should provide ample data for LHCb to probe all the way down to the Standard
Model expectation value for this process. Meanwhile, it seems that while I’ve been writing this
post there have been intriguing hints of a Higgs (also via our editor)… [edit, 6/23: Aidan put
up an excellent intro to these results]

[Many thanks to the experimentalists with whom I’ve had useful discussions about this.]

he spin of gauge bosons: vector particles

Particles have an inherent spin. We explored the case of fermions (“spin-1/2”) in a recent post
on helicity and chirality. Now we’ll extend this to the case of vector (“spin-1”) particles which
describe gauge bosons—force particles.

By now regular US LHC readers are probably familiar with the idea that there are two kinds of
particles in nature: fermions (matter particles) and bosons (force particles). The matter particles
are the ‘nouns’ of the Standard Model. The ‘verbs’ are the bosons which mediate forces between
these particles. The Standard Model bosons are the photon, gluon, W, Z, and the Higgs. The first
four (the gauge bosons of the fundamental forces) are what we call vector particles because of
the way they spin.
An arrow that represents spin

You might remember the usual high school definition of a vector: an object that has a direction
and a magnitude. More colloquially, it’s something that you can draw as an arrow. Great. What
does this have to do with force particles?

In our recent investigation of spin-1/2 fermions, the punchline was that chiral (massless)
fermions either spin clockwise or counter-clockwise relative to their direction of motion. We can
convert this into an arrow by identifying the spin axis. Take your right hand and wrap your
fingers around the direction of rotation. The direction of your thumb is an arrow that identifies
the helicity of the fermion, it is a ‘spin vector.’ In the following cartoon, the gray arrows
represent the direction of motion (right) and the big colored arrows give the spin vector.

You can see that a particle has either spin up (red: spin points in the same direction as motion)
or spin down (blue: spin points in the opposite direction as motion). It should not surprise you
that we can write down a two-component mathematical object that describes a particle. Such an
object is called a spinor, but it’s really just a special kind of vector. In can be represented this
way:

ψ = ( spin up , spin down )

As you can see, there’s one slot that contains information about the particle when it is spin up
and another slot that contains information about the particle when it is spin down. It’s really just
a list with two entries.

Don’t panic! We’re not going to do any actual math in this post, but it will be instructive—and
relatively painless—to see what the mathematical objects look like. This is the difference
between taking a look at the cockpit of a jet versus actually flying it.
All you have to appreciate at this point is that we’ve described fermions (spin-1/2 particles) in
terms of an arrow that determine its spin. Further, we can describe this object as a two-
component ‘spinor.’

For experts: a spinor is a vector (“fundamental representation”) of the group SL(2,C), which is
the universal cover of the Lorentz group. The point here is that we’re looking
at projective representations of the Lorentz group (quantum mechanics says that we’re allowed
to transform up to a phase). The existence of a projective representation of a group is closely tied
to its topology (whether or not it is simply connected); the Lorentz group is not simply
connected, it is doubly connected. The objects with projective phase -1 (i.e. that pick up a minus
sign after a 360 degree rotation) are precisely the half-integer spinor representations, i.e. the
fermions.

Relativity and spin

Why did we bother writing the spinor as two components? Why not just work with one
component at a time: we pick up a fermion and if it’s spin up we use one object and if it’s spin
down we use another.

This, however, doesn’t work. To see why, we can imagine what happens if we take
the sameparticle but change the observer. You can imagine driving next to a spin-up particle on
the freeway, and then accelerating past it. Relative to you, the particle reverses its direction of
motion so that it becomes a spin-down particle.

What does this mean? In order to account for relativity (different observers see different things)
we must describe the particle simultaneously in terms of being spin-up and spin-down. To
describe this effect mathematically, we would perform a transformation on the spinor which
changes the spin up component into the spin down component.

Remark: I’m cheating a little because I’m implicitly referring to a massive fermion while
referring to the two-component spinor of a massless fermion. Experts can imagine that I’m
referring to a Majorana fermion, non-experts can ignore this because the punchline is the same
and there’s not much to be gained by being more rigorous at this stage.
In fact, to a mathematician, this is the whole point of constructing vectors: they’re things which
know how to transform properly when you rotate them. In this way they are intimately linked to
the symmetries of spacetime: we should know how particles behave when we grab them and
rotate them.

Spin-1 (vector) particles

Now that we’ve reviewed spin-1/2 (fermions), let’s move on to spin-1: these are the vector
particles and include the gauge bosons of the Standard Model. Unlike the spin-1/2 particles,
whose spin arrows must be parallel to the direction of motion, vector particles can have their
spin point in any direction. (This is subject to some constraints that we’ll get to below.) We
know how to write arrows in three dimensions: you just write down the coordinates of the arrow
tip:

3D arrow = (x-component, y-component, z-component)

When we take into account special relativity, however, we must work instead
in four dimensional spacetime, i.e. we need a vector with four components (sometimes called
a four-vector, see Brian’s recent post). The reason for this is in addition to rotating our vector,
we can also boost the observer—this is precisely what we did in the example above where we
drove past a particle on the freeway—so that we need to be able to include the length contraction
and time dilation effects that occur in special relativity. Heuristically, these are rotations into the
time direction.

So now we’ve defined vector particles to be those whose spin can be described by an arrow
pointing in four dimensions. A photon, for example, can thus be represented as:

Aμ = (A0, A1, A2, A3)

Here we’ve used the standard convention of labeling the x, y, and z directions by 1, 2, and 3.
The A0 corresponds to the component of the spin in the time direction. What does this all mean?
The (spin) vector associated with a spin-1 particle has a more common name: the polarization of
the particle.

You’ve probably heard of polarized light: the electric (and hence also the magnetic) field is fixed
to oscillate along only one axis; this is the basis for polarized sunglasses. Here’s a heuristic
drawing of electromagnetic radiation from a dipole (from Wikipedia, CC-BY-SA license):
The polarization of a photon refers to the same idea. As mentioned in Brian’s post, the electric
and magnetic fields are given by derivatives of the vector potential A. This vector potential is
exactly the same thing that we have specified above; in a sense, a photon is a quantum of the
vector potential.

Four vectors are too big

Now we get to a very important point: we’ve argued based on spacetime symmetry that we
should be using these four-component vectors to describe particles like photons. Unfortunately, it
turns out that four components are too many! In other words, there are some photon polarizations
that we could write down which are not physical!

Here we’ll describe one reason why this is true; we will again appeal to special relativity. One of
the tenets of special relativity is that you cannot travel faster than the speed of light. Further, we
know that photons are massless and thus travel at exactly the speed of light. Now consider a
photon with is spinning in the same direction as its motion (i.e. the spin vector is perpendicular
to the page):

In this case the bottom part of the photon (blue) is moving opposite the direction of motion and
so travels slightly slower than the speed of light. On the other hand, the top part of the photon is
moving with the photon and thus would be moving faster than the speed of light!

This is a big no-no, and so we cannot have any photons polarized in this way. Our four-
component vector contains more information than the physical photon. Or more
accurately: being able to write down our theory in a way that manifestly respects spacetime
symmetry comes at the cost of introducing extra, non-physical degrees of freedom in how we
describe some of our particles.

(If we removed this degree of freedom and worked with three-component vectors, then our
mathematical formalism doesn’t have enough room to describe how the particle behaves under
rotations and boosts.)
Fortunately, when we put four-component photons through the machinery of quantum field
theory, we automatically get rid of these unphysical polarizations. (Quantum field theory is
really just quantum mechanics that knows about special relativity.)

Gauge invariance: four vectors are still too big

Now I’d like to introduce one of the key ideas of particle physics. It turns out that even after
removing the unphysical ‘faster than light’ polarization of the photon, we still have too many
degrees of freedom. A massless particle only has two polarizations: spin-up or spin-down. Thus
our photon still has one extra degree of freedom!

The resolution to this problem is incredibly subtle: some of the polarizations that we could write
down using a four-vector are physically identical. I don’t just mean that they give the same
numbers when you do the math, I mean that they literally describe the same physical state. In
other words, there is a redundancy in this four-vector description of particles! Just as the case of
the unphysical polarization above, this redundancy is the cost of writing things in a way which
manifestly respects spacetime symmetry. This redundancy is called gauge invariance.

Gauge invariance is a big topic that deserves its own post—I’m still thinking of a good way to
present it—but the “gauge” refers to the same thing in term “gauge boson.” This gauge
invariance (redundancy in our description of physics) is intimately linked to the fundamental
forces of our theory.

Remark, massive particles: Unlike the massless photon, which has two polarizations,
the W and Z bosons have three polarizations. Heuristically the third polarization corresponds to
the particle spinning in the direction of motion which wasn’t allowed for massles particles that
travel at the speed of light. It is still true, however, that there is still a gauge redundancy in the
four-component description for the thee-polarization massive gauge bosons.

For experts: at this point, I should probably mention that the mathematical object
which really describes gauge bosons aren’t vectors, but rather co-vectors, or (one-)forms. One
way to see this is that these are objects that get integrated over in the action. The distinction is
mostly pedantic, but a lot of the power of differential geometry and topology is manifested when
one treats gauge theory in its ‘natural’ language of fiber bundles. For more prosaic goals, we can
write down Maxwell’s equations in an even more compact form: d*F = j. (Even more compact
than Brian’s notation! 🙂 )

Wigner’s classification
Let me take a step back to address the ‘big picture.’ In this post
I’ve tried to give a hint of a classification of “irreducible [unitarity] representations of the
Poincaré group” by Hungarian mathematical physicist Eugene Wigner in the late 1930s.

At the heart of this program is a definition of what we really mean by ‘particle.’ A particle is
something with transforms in a definite way under the symmeties of spacetime, which we call
the Poincaré group. Wigner developed a systematic way to write down all of the
‘representations’ of the Poincaré group that describe quantum particles; these representations are
what we mean by spin-1, spin-1/2, etc.

In addition to these two examples, there are fields which do nothing under spacetime
symmetries: these are the spin-0 scalar fields, such as the Higgs boson. If we treated gravity
quantum mechanically, then the graviton would be a spin-2 [antisymmetric] tensor field. If
nature is supersymmetric, then the graviton would also have a spin-3/2 gravitino partner. Each
of these different spin fields is represented by a mathematical object with different numbers of
components that mix into one another when you do a spacetime transformation (e.g. rotations,
boosts).

In principle one can construct higher spin fields, e.g. spin-3, but there are good reasons to believe
that such particles would not be manifested in nature. These reasons basically say that those
particles wouldn’t be able to interact with any of the lower-spin particles (there’s no “conserved
current” to which they may couple).

Next time: there are a few other physics (and some non-physics) topics that I’d like to blog
about in the near future, but I will eventually get back to this burning question about the meaning
of gauge symmetry. From there we can then talk about electroweak symmetry breaking, is the
main reason why we need the Higgs boson (or something like it) in nature. (For those who have
been wondering why I haven’t been writing about the Higgs—this is why! We need to go over
more background to do things properly.)

Who ate the Higgs?

While one of the priorities of the LHC is to find the Higgs boson (also see Aidan’s rebuttal), it
should also be pointed out that we have already discovered three quarters of the Standard Model
Higgs. Just don’t expect to hear about this in the New York Times; this isn’t breaking news—
we’ve known about this “three quarters” of a Higgs for nearly two decades now. In fact, these
three quarters of a Higgs live inside the belly of two beasts: the Z and W bosons!

What the heck do I mean by all this? What is “three quarters” of a particle? What does the Higgs
have to do with the Z and the W? And to what extent have we or haven’t we
discovered the Higgs boson? These are all part a subtle piece of the Standard Model story that
we are now in an excellent position to decipher.

What we will find is that there’s not one, but four Higgs bosons in the Standard Model. Three of
them are absorbed—or eaten—by the Z and W bosons when they become massive. (This is very
different from the way matter particles obtain mass!) In this sense the discovery of
massive Z and W bosons was also a discovery of these three Higgs bosons. The fourth Higgs is
what we call the Higgs boson and its discovery (or non-discovery) will reveal crucial details
about the limits of the Standard Model.

THE DIFFERENCE BETWEEN MASSLESS AND MASSIVE VECTORS

In the not-so-recent past we delved into some of the nitty-gritty of vector bosons such as the
force particles of the Standard Model. We saw that relativity forces us to describe these particles
with four-component mathematical objects. But alas, such objects are redundant because they
encode more polarization states than are physically present. For example, a photon can’t spin in
the direction of motion (longitudinal polarization) since this would mean part of the field is
traveling faster than the speed of light.

Now, what do we mean by polarization anyway? We’d previously seen that polarizations are
different ways a quantum particle can spin. In fact, each polarization state can be thought of as
an independent particle, or an independent “degree of freedom.” In this sense there are two
photons: one which has a left-handed polarization and one with a right-handed polarization.

Because massive particles (which travel slower than light) can have a longitudinal polarization,
they have an extra degree of freedom compared to massless particles. So repeat after me:

The difference between massless force particles (like the photon and gluon) and massive force
particles (like the W and Z) is the longitudinal degree of freedom.

Since a “degree of freedom” is something like an independent particle, what we’re really saying
is that the Wand Z seem to have an “extra particle’s worth of particle” in them compared to the
photon and gluon. We will see that this poetic language is also technically correct.

The mass of a force particle is important for large scale physics: the reason why Maxwell was
able to write down a classical theory of electromagnetism in the 19th century is because the
photon has no mass and hence can create macroscopic fields. The W and Z on the other hand, are
heavy and can only mediate short-range forces—it costs energy for particles to exchange heavy
force particles.

MASSIVE VECTORS ARE A PROBLEM

The fact that the W and Z are massless is also important for the following reason:

In the early days of quantum field theory, massive vector particles didn’t seem to make any
sense!

The details don’t matter, but the punchline is that the very mathematical consistency of a typical
theory with massive vector particles breaks down at high energies. You can ask a well-posed
physical question—what’s the probability of Ws to scatter off one another—and it is as if the
theory itself realizes that something isn’t right and gives up halfway through, leaving your
calculations in tatters. It seemed like massive vector particles just weren’t allowed.

If that’s the case, then how can the W and Z bosons be massive? Contrary to lyrics to a popular
Lady Gaga song, the W and Z bosons were not “born this way.” Force particles naturally appear
in theories as massless particles. From our arguments above, we now know that the difference
between a massless and a massive particle is a single, extra longitudinal degree of freedom.
Somehow we need to find extra longitudinal degrees of freedom to lend to the W and Z.

Technical remark & update (10 Oct): As a commenter has pointed out below, I should be more
careful in how I phrase this. Theories of massive vectors (essentially nonlinear sigma models)
only become non-unitary at tree-level so that we say they lose “perturbative unitarity.” This on
its own is not a problem and certainly doesn’t mean that the they is “mathematically
inconsistent” since they become strongly coupled and get large corrections from higher order
terms. What we do lose is calculability and one has to wonder if there’s a better description of
the physics at those scales. Many thanks to the ‘anonymous’ commenter for calling me out on

this.

LET THEM EAT GOLDSTONE BOSONS

Jeffrey Goldstone, image from his MIT webpage

Where can this extra degree of freedom come from? One very nice resolution to this puzzle is
called the Higgs mechanism. The main idea is that vector particles can simply annex another
particle to make up the “extra particle’s worth of particle” it needs to become massive. We’ll see
how this works below, but what’s really fantastic is that this is one of the very few known ways
to obtain a mathematically consistent theory of massive vector particles.

So what are these extra particles?

Since particles with spin carry at least two degrees of freedom, this “extra longitudinal degree of
freedom” can only come from a spin-less (or scalar) particle. Such a particle has to somehow be
connected to the force particles that want to absorb it, so it should be charged under the weak
force. (For example, neutrinos are uncharged under electromagnetism since they don’t talk to
photons, but they are charged under the weak force since they talk to the W and Z bosons.)

Further, this particle has to obtain a vacuum expectation value (“vev”). Those of you who have
been following along with our series on Feynman diagrams will already be familiar with this,
though we’re now approaching the topic from a different direction.

In general, particles that can be combined with massless force particles to form massive force
particles are called Goldstone bosons (or Nambu-Goldstone bosons including one of the 2008
Nobel prize winners) after Jeffrey Goldstone, pictured to the right. The Goldstone theorem states
that

A theory with spontaneous symmetry breaking has massless scalar particle in its spectrum.

For now don’t worry about any of these words other than the fact that this gives a condition for
which there must be scalar particles in a theory. We’ll get back to the details below and we’ll see
that these scalar particles, the Goldstone bosons, are precisely the scalars which massless force
particles can absorb in order to become massive.

So now we arrive at another aphorism in physics:

Force particles can eat Goldstone bosons to become massive.

In light of this terminology, perhaps a more appropriate cartoon of this is to draw the Goldstone
particle as a popular type of fish-shaped cracker…

A W eating a Goldstonefish cracker... get it? (I hope we don't get sued for this.)

Technical remarks for experts: (corrected Oct 10 thanks to anon.) The “mathematical
inconsistency” of a generic theory of massive vectors is the non-unitarity of tree-level WW
scattering. This isn’t really an inconsistency since the theory of massive vectors has a cutoff; as
one approaches the cutoff loop-level diagrams give large corrections to the amplitude and the
theory becomes strongly coupled. While this isn’t a technical necessity for new physics, it is at
least a very compelling reason to suspect that there is at least a better description.

In the Standard Model this is done perturbatively. The tree-level cross section for WW scattering
increases with energy but is unitarized by the Higgs boson.

Saying that force particles are “born massless” is a particular viewpoint that lends itself to this
UV completion by linearization of the nonlinear sigma model associated with a
phenomenological theory of massive vectors. This isn’t the only game in town. For example, one
can treat the ρ meson is a vector that can be understood as the massive gauge boson of a `hidden’
gauge symmetry in the chiral Lagrangian. The UV completion of such a theory is not a Higgs,
but the appearance of the bound quarks that compose the ρ. The analogs of this kind of UV
completion in the Standard Model are technicolor, composite Higgs, and Higgs-less models.

FOUR HIGGSES: A DIFFERENT KIND OF REDUNDANCY

Okay, so we have three massive gauge bosons: the W+, W–, and Z. Each one of these has two
transverse polarizations (right- and left-handed) in addition to a longitudinal polarization. This
means we need threeGoldstone bosons to feed them. Where do these particles come from? The
answer should be no surprise, the Higgs.

Indeed, you might think I’m selling you the Standard Model like an informercial:

If you buy now, the Standard Model comes with not one, not two, not even three, but four—
count them, four—Higgs bosons!

Four Higgs bosons?! That’s an awful lot of Higgs. But it turns out this is exactly what we have:
we call them the H+, H–, H0, and h. As you can see, two of them are charged (you can guess
these will be eaten by the Ws), two are uncharged. Here’s they are:

"The Four Higgses of the Standard Model," biblical pun intended

Where did all of these Higgses come from? And why did our theory just happen to have enough
of them? These four Higgses are all manifestations of a different kind of redundancy
called gauge symmetry. The name is related to gauge bosons, the name we give to force
particles.

When we described vector particles, we said that our mathematical structure was redundant: our
four-component objects have too many degrees of freedom than the physical objects they
represented. One redundancy came from the restriction that massless particles can have no
longitudinal polarization. This brings us down from 4 degrees of freedom to 3. However, we
know that massless particles only have two polarizations—we have to remove one more
polarization. (Similarly for massive particles, which have 3, not 4, degrees of freedom.) This left-
over redundancy is precisely what we mean by gauge symmetry.

For those with some calculus-based physics background: this is related to the fact that the
electromagnetic field can be written as derivatives of a potential. This means the potential is
defined up to an constant. This overall constant (more generally, a total derivative) is a gauge
symmetry. To connect to the quantum picture, we previously mentioned that the vector potential
is the classical analog of the 4-vector describing the photon polarization.

Technical remark: in some sense, this gauge symmetry is not a ‘symmetry’ at all but an
overspecification of a physical state such that distinct 4-vectors may describe identical state.
(Compare this to a symmetry where different states yield the same physics.)

Gauge symmetry doesn’t just explain the redundancy in the vector particles, but it also imposes a
redundancy in any matter particles that are charged under the associated force. In particular, the
gauge symmetry associated with the weak force requires that the Higgs is described by a two
component complex-valued object. Since a complex number contains two real numbers, this
means the Higgs is really composed of four distinct particles—the four particles we met above.

Now let’s get back to the statement of Goldstone’s theorem that we gave above:

A theory with spontaneous symmetry breaking has massless scalar particle in its spectrum.

We’re already happy with the implications of having a scalar. Let’s unpack the rest of this
sentence. The hefty phrase is “spontaneous symmetry breaking.” This is a big idea that
deserves its own blog post, but in our present case (the Standard Model) we’ll be “breaking” the
gauge symmetry associated with the W and Zbosons.

What happens is that one of the Higges (in fact, this is “the Higgs,” the one called h) gets
a vacuum expectation value. This means that everywhere in spacetime there Higgs field is “on.”
However, the Higgs carries weak charge—so if it is “on” everywhere, then something must be
‘broken’ with this gauge symmetry… the universe is no longer symmetric since there’s a
preferred weak charge (the charge of the Higgs, h).

For reasons that we’ll postpone for another time, Goldstone’s theorem then implies that the other
Higgses serve as Goldstone bosons. That is, the H+, H–, and H0 can be eaten by the W+, W–,
and Z respectively, thus providing the third polarization required for a massive vector particle
(and doing so in a way that is mathematically consistent at high energies).
Three of the four Higgses are Goldstones and are eaten by the W and Z.

EPILOGUE

There are still a few things that I haven’t told you. I haven’t explained why there was exactly one
Goldstone particle for each heavy force particle. Further, I haven’t explained why it turned out
that each Goldstone particle had the same electric charge as the force particle that ate it. And
while we’re at it, I haven’t said anything about why the photon should be massless while
the W and Z bosons gain mass—they’re close cousins and you may wonder why the photon
couldn’t have just gone off and eaten the h.

Alas, all of these things will have to wait for a future post on what we really mean
by electroweak symmetry breaking.

What we have done is shown how gauge symmetry and the Higgs are related to the mass of force
particles. We’ve seen that the Higgs gives masses to vector bosons in a way that is very different
from the way it gives masses to fermions. Fermions never “ate” any part of the Higgs
but bounced off its vacuum expectation value, while the weak gauge bosons feasted on three-
fourths of the Higgs! This difference is related to the way that relativity restricts the behavior of
spin-one particles versus spin–one-half particles.

Finally, while we’ve shown that we’ve indeed discovered “3/4th of the Standard Model Higgs,”
that there is a reason why the remaining Higgs is special and called the Higgs—it’s the specific
degree of freedom which obtains the vacuum expectation value which breaks the gauge
symmetry (allowing its siblings to be eaten). The discovery of the Higgs would shed light on the
physics that induces this so-called electroweak symmetry breaking, while a non-discovery
of the Higgs would lead us to consider alternate explanations for what resolves the mathematical
inconsistencies in WW scattering at high energies.

Why do we expect a Higgs boson? Part II: Unitarization of Vector Boson Scattering
Hi everyone—it’s time that I wrap up some old posts about the Higgs boson. Last December’s
tantalizing results may end up being the first signals of the real deal and the physics community
is eagerly awaiting the combined results to be announce at the Rencontres de
Moriond conference next month. So now would be a great time to remind ourselves of why
we’re making such a big deal out of the Higgs.

Review of the story so far

Since it’s been a while since I’ve posted (sorry about that!), let’s review the main points that
we’ve developed so far. See the linked posts for a reminder of the ideas behind the words and
pictures.

There’s not only one, but four particles associated with the Higgs. Three of these
particles “eaten” by the Wand Z bosons to become massive; they form the “longitudinal
polarization” of those massive particles. The fourth particle—the one we really mean when we
refer to The Higgs boson—is responsible for electroweak symmetry breaking. A cartoon picture
would look something like this:
The solid line is a one-dimensional version of the Higgs potential. The x-axis represents the
Higgs “vacuum expectation value,” or vev. For any value other than zero, this means that the
Higgs field is “on” at every point in spacetime, allowing fermions to bounce off of it and
hence become massive. The y-axis is the potential energy cost of the Higgs taking a particular
vacuum value—we see that to minimize this energy, the Higgs wants to roll down to a non-zero
vev.

Actually, because the Higgs vev can be any complex number, a more realistic picture is to plot
the Higgs potential over the complex plane:
Now the minimum of the potential is a circle and the Higgs can pick any value.
Higgs particles are quantum excitations—or ripples—of the Higgs field. Quantum
excitations which push along this circle are called Goldstone bosons, and these represent the
parts of the Higgs which are eaten by the gauge bosons. Here’s an example:

Of course, in the Standard Model we know there are three Goldstone bosons (one each for
the W+, W-, and Z), so there must be three “flat directions” in the Higgs potential. Unfortunately,

I cannot fit this many dimensions into a 2D picture. The remaining Higgs particle is the
excitation in the not-flat direction:

Usually all of this is said rather glibly:

The Higgs boson is the particle which is responsible for giving mass.

A better reason for why we need the Higgs

The above story is nice, but you would be perfectly justified if you thought it sounded like a bit
of overkill. Why do we need all of this fancy machinery with Goldstone bosons and these funny
“Mexican hat” potentials? Couldn’t we have just had a theory that started out with massive
gauge bosons without needing any of this fancy “electroweak symmetry breaking” footwork?

It turns out that this is the main reason why we need the Higgs-or-something-like it. It turns out
that if we tried to build the Standard Model without it, then something very nefarious happens.
To see what happens, we’ll appeal to some Feynman diagrams, which you may want to review if
you’re rusty.

Suppose you wanted to study the scattering of two W bosons off of one another. In the Standard
Model you would draw the following diagrams:

There are other diagrams, but these two will be sufficient for our purposes. You can draw the rest
of the diagrams for homework, there should be three more that have at most one virtual particle.
In the first diagram, the two W bosons annihilate into a virtual Z boson or a photon (γ) which
subsequently decay back into two W bosons. In the second diagram it’s the same story, only now
the W bosons annihilate into a virtual Higgs particle.

Recall that these diagrams are shorthand for mathematical expressions for the probability that
the W bosons to scatter off of one another. If you always include the sum of the virtual Z/photon
diagrams with the virtual Higgs diagram, then everything is well behaved. On the other hand, if
you ignored the Higgs and onlyincluded the Z/photon diagram, then the mathematical
expressions do not behave.

By this I mean that the probability keeps growing and growing with energy like the monsters that
fight the Power Rangers. If you smash the two W bosons together at higher and higher energies,
the number associated with this diagram gets bigger and bigger. If these numbers get too big,
then it would seem that probability isn’t conserved—we’d get probabilities larger than 100%, a
mathematical inconsistency. That’s a problem that not even the Power Rangers could handle.

Mathematics doesn’t actually break down in this scenario—what really happens in our “no
Higgs” theory is something more subtle but also disturbing: the theory becomes non-
perturbative (or “strongly coupled”). In other words, the theory enters a regime where Feynman
diagrams fail. The simple diagram above no longer accurately represents the W scattering
process because of large corrections from additional diagrams which are more “quantum,” i.e.
they have more unobserved internal virtual particles. For example:
In addition to this diagram we would also have even more involved diagrams with even more
virtual particles which also give big corrections:

And so forth until you have more diagrams than you can calculate in a lifetime (even with a
computer!). Usually these “very quantum” diagrams are negligible compared to the simpler
diagrams, but in the non-perturbative regime each successive diagram is almost as important as
the previous. Our usual tools fail us. Our “no Higgs theory” avoids mathematical inconsistency,
but at the steep cost of losing predictivity.

Now let me be totally clear: there’s nothing “wrong” with this scenario… nature may very well
have chosen this path. In fact, we know at least one example where it has: the theory of quarks
and gluons (QCD) at low energies is non-perturbative. But this is just telling us that the
“particles” that we see at those energies aren’t quarks and gluons since they’re too tightly bound
together: the relevant particles at those energies are mesons and baryons (e.g.pions and
protons). Even though QCD—a theory of quarks and gluons—breaks down as a calculational
tool, nature allowed us to describe physics in terms of perfectly well behaved (perturbative)
“bound state” objects like mesons in aneffective theory of QCD. The old adage is true: when
nature closes a door, it opens a window.

So if we took our “no Higgs” theory seriously, we’d be in an uncomfortable situation. The theory
at high energies would become “strongly coupled” and non-perturbative just like QCD at low
energies. It turns out that for W boson scattering, this happens at around the TeV scale, which
means that we should be seeing hints of the substructure of the Standard Model electroweak
gauge bosons—which we do not. (Incidentally, the signatures of such a scenario would likely
involve something that behaves somewhat like the Standard Model Higgs.)

On the other hand, if we had the Higgs and we proposed the “electroweak symmetry breaking”
story above, then this is never a problem. The probability for W boson scattering doesn’t grow
uncontrollably and the theory remains well behaved and perturbative.

GOLDSTONE LIBERATION AT HIGH ENERGIES


The way that the Higgs mechanism saves us is somewhat technical and falls under the name of
the Goldstone Boson Equivalence Theorem. The main point is that our massive gauge
bosons—the ones which misbehave if there were no Higgs—are actually a pair of particles: a
massless gauge boson and a massless Higgs/Goldstone particle which was “eaten” so that
the combined particle is massive. One cute way of showing this is to show the W boson eating
Gold[stone]fish:

Indeed, at low energies the combined “massless W plus Goldstone” particle behaves just like a
massive W. A good question right now is “low compared to what?” The answer is the Higgs
vacuum expectation value (vev), i.e. the energy scale at which electroweak symmetry is broken.

However, at very high energies compared to the Higgs vev, we should expect these two particles
to behave independently again. This is a very intuitive statement: it would be very disruptive if
your cell phone rang at a “low energy” classical music concert and people would be very
affected by this; they would shake their heads at you disapprovingly. However, at a “high
energy” heavy metal concert, nobody would even hear your cell phone ring.

Thus at high energies, the “massless W plus Goldstone” system really behaves like two different
particles. In a sense, the Goldstone is being liberated from the massive gauge boson:

Now it turns out that the massless W is perfectly well behaved so that at high energies. Further,
the set of all four Higgses together (the three Goldstones that were eaten and the Higgs) are also
perfectly well behaved. However, if you separate the four Higgses, then each individual piece
behaves poorly. This is fine, since the the four Higgses come as a package deal when we write
our theory.
What electroweak symmetry breaking really does is that it mixes up these Higgses with the
massless gauge bosons. Since this is just a reshuffling of the same particles into different
combinations, the entire combined theory is still well behaved. This good behavior, though,
hinges on the fact that even though we’ve separated the four Higgses, all four of them are still in
the theory.

This is why the Higgs (the one we’re looking for) is so important: the good behavior of the
Standard Model depends on it. In fact, it turns out that any well behaved theory with massive
gauge bosons must have come from some kind of Higgs-like mechanism. In jargon, we say that
the Higgs unitarizes longitudinal gauge boson scattering.

For advanced readers: What’s happening here is that the theory of a complex scalar Higgs
doublet is perfectly well behaved. However, when we write the theory nonlinearly (e.g. chiral
perturbation theory, nonlinear sigma model) to incorporate electroweak symmetry breaking, we
say something like: H(x) = (v+h(x)) exp (i π(x)/v). The π’s are the Goldstone bosons. If we
ignore the Higgs, h, we’re doing gross violence to the well behaved complex scalar doublet.
Further, we’re left with a non-renormalizable theory with dimensionful couplings that have
powers of 1/v all over the place. Just by dimensional analysis, you can see that scattering cross
sections for these Goldstones (i.e. the longitudinal modes of the gauge bosons) must scale like a
positive power of the energy. In this sense, the problem of “unitarizing W boson scattering” is
really the same as UV completing a non-renormalizable effective theory. [I thank Javi S. for
filling in this gap in my education.]

CAVEAT: HIGGS VERSUS HIGGS-LIKE

I want to make one important caveat: all that I’ve argued here is that we need something to play
the role of the Higgs in order to “restore” the “four well behaved Higgses.” While the Standard
Model gives a simple candidate for this, there are other theories beyond the Standard Model that
give alternate candidates. For example, the Higgs itself might be a “meson” formed out of some
strongly coupled new physics. There are even “Higgsless” theories in which this “unitarization”
occurs due to the exchange of new gauge bosons. But the point is that there needs to
be something that plays the role of the Higgs in the above story.

Particle Paparazzi: the private lives of the Standard Model particles (summary)
I wanted to take a break from our ongoing discussion of Feynman diagrams and the Standard
Model to highlight what we’ve learned the true identities of the particles we now know and love.
THE “FAKE” PARTICLE PERIODIC TABLE

These days most people who read Scientific American or Dennis Overbye’s articles in the NY
Times can list off the Standard Model particles off the top of their heads. Their list would look
something like this (which we posted earlier):
These are, of course, the matter particles. Savvy science fans will also list off the photon, W, Z,
and gluon as the Standard Model force particles. With these particles you can explain a whole
swath of nuclear physics and chemistry. But the chart above doesn’t shed light on the
electroweak (Higgs!) physics that leads to this structure.
For this reason, these aren’t the particles that physicists use when describing the theoretical
structure of the Standard Model and its extensions. Instead, we distinguish between things like
the “left-handed electron” and “right-handed anti-positron,” or the “transverse W polarizations”
and versus the longitudinal polarization which is really a Goldstone boson stolen from a group of
[four!] Higgses.
Instead of a lengthy recap of previous blog posts, let’s try to summarize with the right cartoon
pictures. In doing so, we’ll get to learn something about the true hidden identities of the Standard
Model particles and how they really interact with one another.
THE TRUE IDENTITIES OF THE STANDARD MODEL PARTICLES

Let’s start with the electron. We learned that the electron is really a mixture of
two chiral particles: a massless “left-handed electron” and a massless “right-handed anti-
positron.” These two particles bounce off of the Higgs vacuum expectation value (vev) an
convert into one another, leading to the massive particle that we just called the “electron” above.
(We denote the right-handed matter particle with a mustache to highlight that it’s really a totally
different object.) The same story is true for the electrons two siblings, the muon and tau, and also
for each of their antiparticles.
What about the neutrino? In the Standard Model, the neutrino is massless. We now know they
have very small mass, but we’ll stick to the vanilla version of the theory. For this reason, the
neutrino doesn’t have to bounce off of the Higgs vev and should be identified with only a left-
handed neutrino.

I drew speed lines on the neutrino because the Standard Model neutrino travels at the speed of
light due to being massless. I mention this to remind you that the reason why the massive
electron could swap between being a “left-handed electron” (yellow particle above) and a “right-
handed anti-positron” (green particle with a mustache) is that we could imagine being in a
reference frame where the electron spins in the opposite direction, say by speeding past the
electron along its trajectory. We can’t do that with the massless neutrino because it’s going at the
speed of light; thus all neutrinos that we observe are left-handed.

The Standard Model quarks are all massive and behave just like the electron: they are a mixture
of “left-handed quarks” (yellow) and right-handed anti-quarks (green with mustache). Also
shown below are the different “color charge” that each quark comes in: red, blue, and green. This
has nothing to do with actual colors in the visible spectrum, rather it’s a way of describing
objects with three types of charge rather than just two (positive or negative).
There are two types of quarks: up-type (up, charm, top) and down-type (down, strange, bottom);
they differ by their electric charge. Up-type and down-type quarks interact with one another
through the massive Wboson. Part of the theoretical structure of the Standard Model is
that the W boson only talks to the left-handed particles (the yellow guys with no mustache)—we
say that the Standard Model is chiral (compare this to the definition in chemistry and biology).
The W comes along with its also-massive cousin, the Z boson. We now know that massive gauge
bosons come about when previously-massless gauge bosons “eat” a Goldstone boson. These
‘eaten’ Goldstone bosons are three of the four parts of the Standard Model Higgs, leaving one
neutral particle which we call theHiggs. This entire process by which force particles have
become massive has a fancy name, electroweak symmetry breaking.

In fact, in the picture above we show a very important feature of electroweak symmetry
breaking: in addition to H0, the Z boson is also a mixture of a third W boson (called W3 above)
and something we called the Bboson. Like its charged siblings, the W3 only talks to left-handed
particles, though the B is more sociable and will happily talk to both left- and right-handed
particles. Note, however, that even here the Standard Model is biased: the “B charge” of a left-
handed electron (yellow, no mustache) is different from the “B charge” of a right-handed
electron!
What happened to the other leftover parts of the W3 and B? They form another force particle, the
photon! The interesting thing about the photon is that it’s the electroweak force particle that
didn’t eat any Goldstones—it’s the combination of gauge bosons that survived “electroweak
symmetry breaking” without putting on weight. We say that “electroweak symmetry has broken
to electromagnetism.” Unlike the full electroweak theory, electromagnetism is left/right
democratic in how it talks to matter particles.

Since the photon has not eaten any Goldstone boson, it remains massless and travels at the speed
of light. This sounds silly since it is light, so I should say that it travels at the “universal speed of
all things which have no mass.” Speaking of which, there’s one more force particle that has
nothing to do with electroweak symmetry and also is massless, the gluon:

There are actually eight gluons coming from different color and anti-color combinations, i.e.
corresponding to different combinations of quarks and anti-quark colors that may interact with
the gluon. (Astute readers will wonder why there aren’t nine combinations… this is due to a
subtlety in group theory!)
WHAT DOES THIS ALL BUY US?

So we see that the Standard Model is actually a bit more complicated than the “fake” version that
we showed at the top of the page. Even though it might not seem like it, the theory of this “more
complicated” Standard Model is actually rather elegant and minimal. I should also say that
calling the simple table a “fake” is too harsh: that is indeed an accurate description of the
Standard Model after “electroweak symmetry breaking,” but it doesn’t illuminate the rich
structure of interactions included in the theory.
For example, one would have never understood why nuclear beta decays (mediated by the W)
always produce left-handed neutrinos. It also wouldn’t have been clear how “longitudinal vector
boson scattering is rendered unitary” at high energies—in other words, the description of certain
types of gauge boson scattering breaks down without something like a Higgs to keep the
calculations under control.
The first example is a real matching of observed phenomenon to a theoretical framework. The
second example shows how our theory is prediction about missing pieces that it needs to make
sense. While it may sound dry, these are important points—this is part of the Scientific Method:
we use observations of natural phenomena and build hypotheses (theories) that make predictions.
We can then go and build and LHC (… as well as LEP and the Tevatron) to confirm or refute
these predictions—the data from these experiments then feed back in to revise (or overthrow) our
theories.
So now that we’re more or less up to speed with the moving parts of the Standard Model, we can
push forward to figure out why we believe it should still break down at TeV-scale energies and
give some hint of even more fundamental organizing principles. (This is the really exciting part!)

Das könnte Ihnen auch gefallen