1: \chapter{Introduction}
2: \label{ch:intro}
3:
4: Approximations abound in physics. Classical mechanics is merely an approximation that works well at large distances and small speeds. If distances become too short, we must revert to quantum mechanics. If the speeds become too fast, we enter the realm of relativity. If the distances are short and the speeds are fast, then the approximations supplied by quantum mechanics and relativity are no longer valid and we must combine the two to get quantum field theory. Even quantum field theory, which is the basis for the Standard Model and nearly all of particle physics theory, is likely only an approximation. It may break down at even higher energies and need to be replaced by something else like string theory.
5:
6: Just because an approximation is not correct {\em everywhere} does not mean it is useless and should be discarded. In fact, the opposite is true. Since it is correct {\em somewhere}, it should be embraced. Approximations provide a way of isolating and controlling our ignorance. Identify only the necessary details and throw the garbage away! The hard part, of course, is identifying what is necessary so we do not throw away the baby with the bath water.
7:
8: The heart of this thesis is approximations. We analyze the low-energy quantum mechanical three-body problem using delta function potentials. For sufficiently low energies, the particles cannot discern the details of the interaction. Therefore, any suitably-adjusted, short-range potential should provide an accurate approximation to the true underlying potential. All knowledge of the true potential is characterized by a few parameters which can be fit to experimental data. This allows us to study general behavior that is valid for any low energy system.
9:
10: I shall first provide several examples of approximations in areas ranging from classical electrostatics to effective field theories. This will hopefully illustrate the basic concepts used throughout the rest of our work. Readers familiar with these ideas may wish to skip ahead to Section \ref{sec:mywork} where we discuss in more detail the problem considered in this dissertation.
11:
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%
13: %% MULTIPOLE EXPANSION %%
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%
15: \section{Multipole Expansion}
16:
17: The electrostatic potential created at some point $\vec{r}$ by a charge distribution is given by the formula
18: %
19: \begin{equation}
20: V(\vec{r}) = \frac{1}{4 \pi \epsilon_0} \int d^3\vec{r_{~}}' \, \frac{\rho(\vec{r}')}{\left| \vec{r} - {\vec{r_{~}}}' \right|} ,
21: \end{equation}
22: %
23: \noindent where $\rho$ is the charge density of the distribution. Suppose that all of the charge is contained within some sphere of radius $l$ ($r' \le l$), and that the point at which we calculate the potential is very far away ($l \ll r$). In this case, we can expand the denominator in the integral
24: %
25: \begin{equation}
26: \frac{1}{\left| \vec{r} - \vec{r_{~}}' \right|} = \frac{1}{r} \left[ 1 + \frac{r'}{r} \cos(\theta) + \left(\frac{r'}{r}\right)^2 \left(\frac{3}{2}\cos^2(\theta) - \frac{1}{2}\right) + \cdots \right] ,
27: \end{equation}
28: %
29: \noindent where $\theta$ is the angle between the vectors $\vec{r}$ and $\vec{r_{~}}'$. This leads to approximations for the potential that result from truncating the multipole expansion \cite{Jackson:EM}:
30: %
31: \begin{equation}
32: V(\vec{r}) = \frac{1}{4 \pi \epsilon_0} \left[ \frac{1}{r} \int d^3\vec{r_{~}}' \, \rho(\vec{r_{~}}') + \frac{1}{r^2} \int d^3\vec{r_{~}}' \, r' \cos(\theta) \, \rho(\vec{r_{~}}') + \cdots \right] .
33: \end{equation}
34: %
35: \noindent The first term in the expansion is called the monopole term. It represents the potential that would be created if all of the charge was concentrated at one point. For large distances, it makes sense that the distribution would look like a point charge, and the first term reflects this. However, if the total amount of charge is zero, then the monopole term is also zero. Yet the charges must still create some potential. This potential is approximated by the second piece called the dipole term. It is the potential that would be created by a dipole at $\vec{r_{~}}' = 0$. The full potential is built from all of these terms. It behaves somewhat like a monopole, and somewhat like a dipole, and somewhat like a quadrupole, etc.
36:
37: Notice that as we move farther away from the charges, all of the terms decrease in strength, but some decrease faster than others. For a non-zero total charge, the dominant term is the monopole term at sufficiently large distance, so it is called the {\em leading order} term. The dipole term is then called the {\em next-to-leading order} term, or equivalently the {\em first order correction} to the leading order term. If the total charge is zero, then the dipole term is the leading order term with the quadrupole moment providing the first order correction.
38:
39: Even if we stay at a fixed radius, the higher terms in the expansion still contribute less and less. Consider the dipole term. The integration involves $r'$ which we know to be less than or equal to $l$. We expect that the integral would be roughly equal to $l$ times some charge. The whole term then looks like $l/r^2$. This is smaller than the leading term, which behaves like $1/r$, by a factor of $l/r \ll 1$. We say that this dipole term is of order $\mathcal{O}(l/r)$ compared to the leading term.
40:
41: The small quantity $l/r$ acts as an expansion parameter for the potential. Each additional term is smaller and smaller. If we desire some accuracy $\epsilon$ in our calculation, we only need to include a finite number of terms from the expansion. Let $N$ be the integer such that $(l/r)^N < \epsilon$. Then any term in the expansion of $\mathcal{O}((l/r)^n)$ for $n > N$ may be dropped. For example, if $N = 2$, we keep only the monopole and dipole terms.
42:
43: The expansion hinges upon the fact that $l$ and $r$ are widely separated length scales. If we considered distances where $r \sim l$, then each term of the expansion would be about as large as all the others. They all contribute equally, and the expansion breaks down, reflecting the fact that we are now considering distances short enough to discern the details of the charge distribution. Thus, there is a limit imposed on how small $r$ may be. Exceed this limit, and the approximation is worthless. Even without the $r > l$ restriction, we find that the individual terms in the multipole expansion diverge as $r \rightarrow 0$ even if the true potential never diverges. This provides additional proof that the multipole approximation is no good at very short distances.
44:
45:
46: %%%%%%%%%%%%%%%%%%%%%%%%%
47: %% QUANTUM MECHANICS %%
48: %%%%%%%%%%%%%%%%%%%%%%%%%
49: \section{Quantum Mechanics}
50: \label{sec:introQM}
51:
52: The Schr\"odinger equation in position space is
53: %
54: \begin{equation}
55: \left[ -\frac{\nabla^2}{2m} + V(\vec{r}) \right] \psi(\vec{r}) = E \, \psi(\vec{r}) .
56: \end{equation}
57: %
58: \noindent We will choose units so that $\hbar = 1$. Let us consider the case of a spherically symmetric potential and look at low-energy S-wave scattering from this potential. We will not cover scattering theory in great detail, but rather try to treat the subject quite simply.
59:
60: If the potential is zero, then solutions to the equation for $E > 0$ are easily found. There are two linearly independent solutions which we take to be incoming and outgoing spherical waves. Any S-wave solution is written as a linear combination of these two solutions:
61: %
62: \begin{equation}
63: \psi(r) = A \frac{\me^{ikr}}{r} + B \frac{\me^{-ikr}}{r} .
64: \end{equation}
65: %
66: \noindent The energy for this wavefunction is $E = k^2/2m$. For a non-zero short-range potential, this solution will still be valid at large distances where we can neglect the interaction. So we can view scattering as a spherical wave approaching the potential from $r = \infty$, interacting with the potential, then leaving the potential and returning to infinity. Since the probability associated with the incoming and outgoing waves must be conserved, the constants $A$ and $B$ can only differ by a phase, $A = -B \me^{2 \mi \delta_0(k)}$. The function $\delta_0(k)$ is called the phase shift, and the $k$ dependence is shown to illustrate the fact that the phase can depend upon the energy of the wave.
67:
68: Analytic properties of the phase shift $\delta_0(k)$ for very low momentum can be used to show that the quantity $k \, \cot(\delta_0(k))$ has a well-defined expansion about the point $k = 0$. This expansion, known as the effective range expansion, is written as
69: %
70: \begin{equation}
71: k \, \cot(\delta_0(k)) = - \frac{1}{a} + \frac{1}{2} r_e k^2 + \cdots .
72: \end{equation}
73: %
74: \noindent For a more detailed derivation of this expansion, see Ref.~\cite{Taylor:scatter} or \cite{Newton:scatter}.
75:
76: The parameters $a$, $r_e$, etc.~contain information about the details of the potential with which the spherical wave interacts. The parameter $a$ is known as the scattering length, while $r_e$ is known as the effective range. In general, an infinite number of terms appear in the exact expansion, but we can truncate the series for small $k$ to achieve a desired accuracy with a finite number of terms.
77:
78: In a sense, we are using low-energy scattering to probe some of the properties of the potential. While we can use this to determine some of the details, we cannot determine all of them (unless we measure the phase shift exactly for all $k$). There will always be infinitely many potentials that have the same scattering length. Even if we did distinguish some of them by different effective ranges, there would still be an infinite number of potentials that share both $a$ and $r_e$.
79:
80: Fortunately, this is an asset and not a hindrance. What this means is that the low-energy behavior is insensitive to the detailed form of the potential, and any two potentials that give the same parameters are equally good approximations. In fact, sometimes it is not even necessary to find a potential to use as an approximation. The effective range expansion can be used to make predictions, and other derived quantities can be written in terms of the scattering length and other parameters. By experimental measurements, we can determine $a$ and $r_e$ and then use them to predict other quantities.
81:
82: Let us look at an example to make some of these ideas clearer. We shall take the potential to be a square well
83: %
84: \begin{equation}
85: V(r) = -V_0 \, \theta(1/\Lambda - r).
86: \end{equation}
87: %
88: \noindent Inside the well, the wavefunction has the form
89: %
90: \begin{equation}
91: \psi(r) = A \sin(K r) ,
92: \end{equation}
93: %
94: \noindent where
95: %
96: \begin{equation}
97: K = \sqrt{2 m (E + V_0)} .
98: \end{equation}
99: %
100: \noindent Outside the well, the solution takes the form
101: %
102: \begin{equation}
103: \psi(r) = B \sin(k r + \delta_0) ,
104: \end{equation}
105: %
106: \noindent where
107: %
108: \begin{equation}
109: k = \sqrt{2 m E} .
110: \end{equation}
111: %
112: \noindent Since the quantity $\partial \ln(\psi(r))/\partial r$ must be continuous,
113: %
114: \begin{equation}
115: K \cot(K/\Lambda) = k \cot(k/\Lambda + \delta_0) \label{eqn:boundary}.
116: \end{equation}
117: %
118: \noindent For very low energies, we can expand both sides of Eq.~(\ref{eqn:boundary}) in powers of $k$ and then match terms. Notice that the phase shift occurs on one side, and $V_0$ on the other. This ensures that when we match expansions, the parameters for $\delta_0$ will be written in terms of the parameters for the potential.
119:
120: Without going into details, we present the results of this expansion:
121: %
122: \begin{eqnarray}
123: \frac{1}{a} & = & - \frac{g \Lambda}{\tan(g) - g} \label{eqn:a},
124: \\
125: r_e & = & \frac{(1 - a \Lambda)^2}{a^2 \Lambda^3} \label{eqn:r2},
126: \end{eqnarray}
127: %
128: \noindent where $g = \sqrt{2 m V_0}/\Lambda$. The one condition required to truncate the expansion is that $k \ll \Lambda$, which should come as no surprise. A particle of energy $k^2/2m$ has a wavelength of about $1/k$. If we demand insensitivity to the short distance behavior, then the wavelength must be much greater than the range of the interaction. High energies probe small distances, but low energies do not. The parameter $\Lambda$ acts as a momentum cutoff. Incidently, this square well potential can be used to approximate any other short-range potential.
129:
130: By choosing $g$, and hence $V_0$, to be a function of $\Lambda$, we can make the scattering length $a$ independent of the cutoff by always making sure that Eq.~(\ref{eqn:a}) is satisfied for the same constant $a$. This defines a whole set of potentials that give identical scattering lengths. The only difference between them is the value of $r_e$. As $\Lambda$ changes, so will the effective range. This is where the errors in a simple square well approximation come in. If our potential had another free coupling to adjust, then we could typically make $r_e$ independent of the cutoff, and the error moves to the next parameter. So we see that we can define many different potentials (one for each value of $\Lambda$) that can serve as equally accurate approximations. We can even go so far as to let $\Lambda \rightarrow \infty$ and end up with a delta function potential well.
131:
132: We should also take a moment to look at the relative size of our parameters. The scattering length appears to be around $1/\Lambda$, and this would imply that $r_e$ is also about $1/\Lambda$. [We say that $a$ and $r_e$ are both $\mathcal{O}(1/\Lambda)$.] This is called a ``natural'' theory. All of the parameters have their lengths set by the underlying length scale of the problem, in this case $1/\Lambda$. However, it is possible that $g$ could be fine-tuned to make $a$ much larger than $1/\Lambda$. Even though $a \Lambda \gg 1$, the size of $r_e$ would still be $\mathcal{O}(1/\Lambda)$. The effective range is still set by the range of the potential. However, this too could be made unnaturally large if there is a second parameter in our theory that we could fine-tune.
133:
134: In addition to looking at low-energy scattering states, we could also look at low-energy bound states. The condition for having a bound state of energy $-B$ is
135: %
136: \begin{equation}
137: \sqrt{g^2 \Lambda^2 - 2 m B} \cot\left(\sqrt{g^2 \Lambda^2 - 2 m B}/\Lambda\right) = - \sqrt{2 m B} \label{eqn:SWbound}.
138: \end{equation}
139: %
140: \noindent Treating $mB/\Lambda^2$ as a small quantity, we expand the left-hand side to leading order to obtain the relation
141: %
142: \begin{equation}
143: B = \frac{1}{2m} (g \Lambda \cot(g))^2 = \frac{1}{2m} \left(\frac{\Lambda}{1 - a\Lambda}\right)^2 \label{eqn:SWboundapprox}.
144: \end{equation}
145: %
146: \noindent For very large $a$, the binding energy behaves like $1/ma^2$. This shows the connection between large scattering lengths and shallow bound states. If we wanted to expand the equation for $B$ to higher orders, we would find that for large $a$ it is expanded in powers of $1/a\Lambda \sim r_e/a$. Once again, our expansion parameter is set by two widely separated length scales.
147:
148: For a natural theory, the scattering length would be set by the scale $1/\Lambda$. In this case, Eq.~(\ref{eqn:SWboundapprox}) implies that the binding energy is also set by the potential's length scale: $B \sim \Lambda^2/2m$. This argument should not be taken too seriously because such a binding energy would violate the assumption that $mB/\Lambda^2 \ll 1$, which was required to expand Eq.~(\ref{eqn:SWbound}) in the first place. The leading approximation for the binding energy also diverges when $a \Lambda = 1$ showing that it breaks down for a natural theory. Nonetheless, the fact that $B \sim \Lambda^2/2m$ remains valid as can be seen by comparing Eqs.~(\ref{eqn:a}) and (\ref{eqn:SWbound}) for the case $a \sim 1/\Lambda$.
149:
150: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
151: %% EFFECTIVE FIELD THEORY %%
152: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
153: \section{Effective Field Theory}
154:
155: We now examine the interaction of two identical particles from the field theory point-of-view. This requires some prior knowledge of field theory Lagrangians and how Feynman rules are derived from them.\footnote{See, for example, Ref.~\cite{Weinberg:QFT1}} Once again, we will sacrifice rigor for simplicity and simply state many results so that readers unfamiliar with perturbative field theory can follow the discussion. The important things at this point are not the mathematical steps, but the ideas behind them.
156:
157: Suppose we have two identical particles interacting via some unknown short-range potential and we wish to approximate the behavior. We will start the approximation with the Lagrangian
158: %
159: \begin{equation}
160: \mathcal{L} = \phi^*(\vec{x}) \left( i \frac{\partial}{\partial t} + \frac{\nabla^2}{2 m} \right) \phi(\vec{x}) + C_0 \left[\phi^*(\vec{x}) \phi(\vec{x})\right]^2 \label{eqn:L1} ,
161: \end{equation}
162: %
163: \noindent where $C_0$ is the coupling constant for the two-body interaction. We require that all interactions in the Lagrangian be local operators. That is, they are the product of fields at the same point. These operators can be thought of as contact interactions since they act only at a single point.
164:
165: If we expand the scattering amplitude perturbatively using this interaction, the first term will consist of a single vertex with a value of $- \mi C_0$. The next term will contain two vertices, two propagators, and a single loop integration. Unfortunately, the loop integral behaves like $\int dk$ which diverges linearly. If we try calculating the next term, we find that it diverges quadratically. Each successive term has a more severe divergence than the last. It appears that our hope of a perturbative expansion is lost.
166:
167: Or is it? The reason we encounter this divergence is because we have states with arbitrarily large differences in momenta coupled to one another. However, our approximation is not meant to be used for arbitrarily high momentum. It is only valid for low momenta that cannot resolve the internal details of the interaction. For the moment, let us impose a cutoff $\Lambda$ on the momentum values. This represents the point at which our approximation breaks down and knowledge of the true potential becomes necessary. In the center-of-mass frame, the scattering amplitude as a function of the momentum $p$ for the first two perturbative terms turns out to be
168: %
169: \begin{equation}
170: \mathcal{A}(p) = -C_0 - \frac{C_0^2 m}{4 \pi^2} \left( \Lambda - \mi \pi p /2 \right) \label{eqn:Ap}.
171: \end{equation}
172: %
173: \noindent To calculate $\mathcal{A}$, we would need to know the values of $C_0$ and $\Lambda$. Yet the reason we are using this approximation is because we do not know the underlying potential, and we want our results to be insensitive to such details as the exact value of $\Lambda$.
174:
175: The solution is to let $C_0$ be a function of the cutoff. At $p = 0$,
176: %
177: \begin{equation}
178: \mathcal{A}(0) = -C_0 - \frac{C_0^2 m}{4 \pi^2} \Lambda ,
179: \end{equation}
180: %
181: \noindent which is related to the two-body scattering length $a$ by
182: %
183: \begin{equation}
184: a = - \frac{m \mathcal{A}(0)}{8 \pi} = \frac{m C_0}{8 \pi} \left( 1 + \frac{C_0 m}{4 \pi^2} \Lambda \right) \label{eqn:aC0expand}.
185: \end{equation}
186: %
187: \noindent If we invert this equation to get a perturbative expression for $C_0$ as a function of $a$, we obtain
188: %
189: \begin{equation}
190: C_0 = \frac{8 \pi a}{m} \left( 1 - \frac{2 a \Lambda}{\pi} \right) .
191: \end{equation}
192: %
193: \noindent This allows us to rewrite Eq.~(\ref{eqn:Ap}) as
194: %
195: \begin{equation}
196: \mathcal{A}(p) = - \frac{8 \pi a}{m} \left( 1 - \mi a p\right) \label{eqn:Ap2}.
197: \end{equation}
198: %
199: \noindent Our originally divergent amplitude has now been cast in terms of the finite scattering length, which we can determine from experiment.
200:
201: This procedure that we have just sketched is an example of {\em renormalization}. We {\em regulate} any divergent integrals with some sort of cutoff, allow the coupling constants to be functions of this cutoff, and then choose the form of the couplings to remove this cutoff from physical quantities. Because the coupling constants change as the cutoff changes, they are said to ``run with the cutoff'' and may be referred to as ``running couplings.'' In our example, all reference to $\Lambda$ is eliminated, at least up to the perturbative order in the couplings we have expanded in. In general, this will not be the case. Instead, there will remain some cutoff dependence involving inverse powers of $\Lambda$. The cutoff can then be taken to infinity, leaving a finite result.
202:
203: Notice that the second term in Eq.~(\ref{eqn:Ap2}) contains the product $a p$. Our approximation is only valid for low momentum, such that $p^{-1}$ is much greater than the range of the underlying interaction. If this is a natural theory, we expect that $a$ will be of the same order as the interaction range and hence $ap \ll 1$. In other words, we can expand our amplitude as a power series in $ap$, and we have seen just that. The $-\mi a p$ term is much smaller than the first term of $\mathcal{O}(1)$.
204:
205: We have only included one interaction in the Lagrangian so far. With just this, we cannot produce a theory with a fixed, non-zero effective range. To control the effective range, we shall add another operator of the form $ C_2 \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right) \cdot \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right)$. Since it contains derivatives of the fields, it is sometimes referred to as a ``derivative interaction.'' If we were to add its contributions to $\mathcal{A}(p)$, we would find divergences that need to be regulated. This can be done once again by applying the cutoff $\Lambda$ and then properly choosing $C_2(\Lambda)$. This allows us to control the effective range $r_e$, and after renormalization, the divergent quantities can be rewritten in powers of $r_e p$. For a natural theory, this is the same order as $a p$ and can also be used as an expansion parameter.
206:
207: This example has helped to introduce the general concepts of effective field theory (EFT) \cite{Weinberg:PhysicaA,Lepage:GeneralEFT,Kaplan:GeneralEFT}. The main idea behind EFT is that the details of the underlying physics are embodied in the coupling constants of the theory. By relating the couplings to physical quantities (like scattering length and effective range), other calculations can be written in terms of these physical parameters without any knowledge of the detailed underlying theory. In addition, interactions containing more and more derivatives or fields contribute less and less to the overall result at low energies, just as higher order terms in the multipole expansion contribute less and less to the electrostatic potential far from the source.
208:
209: To see this, we must look at the dimension of each operator. We will assume that the range $R$ of the underlying interaction sets the scale of the terms in our Lagrangian density:
210: %
211: \begin{equation}
212: \mathcal{L} = \phi^*(\vec{x}) \left( i \frac{\partial}{\partial t} + \frac{\nabla^2}{2 m} \right) \phi(\vec{x}) + C_0 \left(\phi^*(\vec{x}) \phi(\vec{x})\right)^2 + C_2 \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right) \cdot \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right).
213: \end{equation}
214: %
215: \noindent If that is the case, then we would expect any length variables like $x$ to scale like $R$, which we shall write as $[x] = R$. This implies that $[\partial/\partial x] = R^{-1}$ and hence $[\nabla] = R^{-1}$. By these arguments, the kinetic energy operator would scale like $[\nabla^2/2m] = m^{-1} R^{-2}$. A look at our Lagrangian density shows that the time derivative must also scale in the same way: $[\partial/\partial t] = m^{-1} R^{-2}$. This relation then implies that the time variable scales as $[t] = m R^2$.
216:
217: To find the scaling behavior of the fields and couplings, we observe that the action $S = \int dt \int d^3x \, \mathcal{L}$ is dimensionless. Since $[dt \, d^3x] = [t x^3] = m R^5$, we must have $[\mathcal{L}] = m^{-1} R^{-5}$. Therefore, every term that makes up a part of $\mathcal{L}$ must also scale in the same way:
218: %
219: \begin{eqnarray}
220: \left[\phi^*(\vec{x}) \left( i \frac{\partial}{\partial t} + \frac{\nabla^2}{2 m} \right) \phi(\vec{x})\right] & = & m^{-1} R^{-5} \nonumber,
221: \\
222: \left[C_0 \left(\phi^*(\vec{x}) \phi(\vec{x})\right)^2\right] & = & m^{-1} R^{-5} \nonumber,
223: \\
224: \left[C_2 \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right) \cdot \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right)\right] & = & m^{-1} R^{-5} \nonumber.
225: \end{eqnarray}
226: %
227: \noindent From this, we can easily draw the following relations:
228: %
229: \begin{eqnarray}
230: \left[ \phi \right] & = & R^{-3/2} \nonumber,
231: \\
232: \left[ C_0 \right] & = & m^{-1} R \nonumber,
233: \\
234: \left[ C_2 \right] & = & m^{-1} R^3 \nonumber.
235: \end{eqnarray}
236:
237: Since the couplings are supposed to approximate the details of the short-range behavior, we expect that their scale should be set by the scale of the short-range interaction. This is in agreement with the leading order approximation to the scattering length in Eq.~(\ref{eqn:aC0expand}) where $a \propto m C_0 \sim R$. Notice the factor of $m^{-1}$ in both $C_0$ and $C_2$. This same factor occurs in every coupling. Every interaction operator will consist of fields ($\phi$), derivatives ($\nabla$), and some coupling. Every field scales as $R^{-3/2}$, while every derivative scales like $R^{-1}$. If the fields and derivatives combined scale like $R^{-N}$, then the coupling must scale like $m^{-1} R^{N-5}$.
238:
239: Of course, any dimensionless quantity containing the couplings must have another scale to balance out the powers of $R$. This scale comes from the momenta of the process being studied. If the system has a typical momentum of $p$, then a dimensionless quantity would be a function of $p R$. Hence, if $p R \ll 1$, we can expand in powers of $p R$. An operator of dimension $R^N$ has a coupling proportional to $R^{N-5}$ and contributes at order $(pR)^{N-5}$. This allows us to classify the operators and decide which Feynman diagrams are needed to achieve any given accuracy.
240:
241: Returning to our example, a single $C_0$ vertex adds a term of order $R \sim a$ to $\mathcal{A}$. Two $C_0$ vertices contribute $R^2 p \sim a^2 p$, while three add $R^3 p^2 \sim a^3 p^2$. At this point, we notice that a single $C_2$ vertex is of order $R^3 p^2$, implying that we must include $C_2$ if we want to calculate $C_0$ effects beyond the first two orders. Because the effects of the $C_2$ interaction do not contribute to leading order, we classify the operator $\vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right) \cdot \vec{\nabla}\left( \phi^*(\vec{x}) \phi(\vec{x})\right)$ as {\em irrelevant}. The ``irrelevant'' qualifier in the term irrelevant operator does not carry the usual meaning, implying that this operator is not needed at all. As we have seen, it is vital if we desire a cutoff-independent effective range. Rather, we use the term irrelevant operator to mean that the operator's effects are smaller than the leading order effects. The renormalization group provides a more precise definition based on the scaling behavior of the operator about a fixed point, but this is more technical than required for our discussion.
242:
243: We now have a well-defined expansion method. We can calculate increasingly accurate results by adding additional operators, which means fitting more couplings to experimental data. One might wonder why this is any different from curve fitting. Why not create some arbitrary potential that includes a bunch of adjustable parameters, and then do some sort of least-squares fit to experimental data to determine the optimal parameter values? The difference is that EFT provides a systematic way of improving the result, instead of merely guessing a potential that may or may not improve things. For any given system, we can write down local operators that could appear as interactions, provided that they satisfy all symmetries of the system. These operators can then be classified based on their dimension, and the point at which they each contribute to the expansion is known.
244:
245: There are cases where the above expansion method is not as useful as might be expected. The most obvious is when $a$ is very large. Since we require $a p \ll 1$, a large scattering length severely limits the range of momentum in which our expansion might converge. Another case would be the study of bound states. The typical momentum in a bound state is of order $1/a$, so once again the expansion would not converge. The solution in these cases is to treat $a p$ as $\mathcal{O}(1)$ and sum all terms. If this is possible, then we can still treat powers of $r_e p$ perturbatively for the case where $r_e \ll a$, and we have an increased momentum range over which the expansion is valid.
246:
247: Fortunately in our example, the contributions to $\mathcal{A}$ from terms containing only $C_0$ form a geometric series. The result of the summation is
248: %
249: \begin{equation}
250: \mathcal{A}(p) = - \left( \frac{8 \pi}{m} \right) \left[\left( \frac{8 \pi}{C_0 m} - \frac{2 \Lambda}{\pi} \right) + \mi p\right]^{-1} ,
251: \end{equation}
252: %
253: \noindent which leads to a scattering length of
254: %
255: \begin{equation}
256: a = \left( \frac{8 \pi}{C_0 m} - \frac{2 \Lambda}{\pi} \right)^{-1}.
257: \end{equation}
258: %
259: \noindent To keep the scattering length cutoff independent, we choose $C_0(\Lambda)$ to be
260: %
261: \begin{equation}
262: C_0(\Lambda) = \left(\frac{8 \pi a}{m}\right)\left(1 + \frac{2 a \Lambda}{\pi}\right)^{-1} .
263: \end{equation}
264: %
265: \noindent If we let $a \rightarrow \infty$ in the above equation, then we see that $C_0$ approaches $(4 \pi^2)/(m \Lambda)$. It is by fine-tuning $C_0$ to be close to this value that we find a system with a very large scattering length. This also illustrates that the running coupling $C_0$ possesses a ``fixed point.'' It is common to consider the behavior of the dimensionless versions of the couplings as $\Lambda \rightarrow \infty$. In this case, $\Lambda C_0 \rightarrow 4 \pi^2/m$ which is a fixed constant value. This is a typical behavior for many couplings. When we consider the three-body problem later, we will find that the three-body coupling exhibits another type of behavior: a limit cycle. Limit-cycle behavior simply means that the coupling approaches a periodic function as $\Lambda \rightarrow \infty$.
266:
267:
268: %%%%%%%%%%%%%%%
269: %% MY WORK %%
270: %%%%%%%%%%%%%%%
271: \section{Three-Body Problem}
272: \label{sec:mywork}
273:
274: The EFT approach has been applied with great success to many low-energy two-body systems \cite{Seki:EFT2body,Bedaque:EFT2body,VanKolck:EFT2body,Beane:EFT2body}. A logical next step would be the application of EFT techniques to the three-body problem. Recent investigations of such systems by Bedaque, Hammer, and van Kolck have turned up some surprising results \cite{hammer:orig}. For three identical bosons, dimensional analysis suggests that only the leading order two-body interaction is needed to obtain the leading order three-body results. While this is indeed sufficient to remove any integration divergences in the equations, cutoff sensitivity still remains in the form of a non-unique $\Lambda \rightarrow \infty$ limit. The remedy is to introduce a three-body contact interaction of higher dimension, indicating that short-distance effects are playing an important role. In addition, the coupling of the three-body interaction is found to exhibit a limit-cycle behavior, unlike the more common fixed-point behavior of running coupling constants.
275:
276: I choose to investigate low-energy three-body systems in a slightly different manner from that used by Bedaque, Hammer, and van Kolck. For sufficiently low energies, the potential $V(\vec{r})$ is replaced by a series of point-like potentials, namely products of $\delta^3(\vec{r})$ and its derivatives. As in the case of EFT, the coupling constants multiplying these terms approximate the short-range details of $V(\vec{r})$ and must be fit to experimental quantities such as scattering length and effective range.
277:
278: The focus of my calculation is on the bound-state equation, which provides a complimentary approach to the scattering methods employed in other works \cite{ksw:largea,hammer:orig,kolck:largea}. My work closely follows unpublished calculations by Kenneth Wilson who developed the basic method and tools that I employ \cite{Wilson:notes}. I use a combination of perturbative expansion and numerical methods that not only provides an independent verification of previous results but also offers several possible advantages. There is no need to introduce a composite field, such as the ``dimeron'' used in Ref.~\cite{hammer:orig}, nor is it necessary to resum Feynman diagrams for the dimeron's propagator. I introduce a technique for expanding the bound state equation in inverse powers of the cutoff, giving us the ability to examine cutoff-dependent behavior order by order. Additionally, the numerical methods I use allow computation to high accuracy, typically 11 to 12 digits in the couplings and binding energies.
279:
280: In this dissertation, I derive the expanded set of three-body integral equations and present the results of my numerical calculations. I numerically ``prove'' that a three-body coupling term is necessary and sufficient to renormalize the three-body bound-state spectrum. The three-body coupling exhibits limit-cycle behavior, and I present an argument for why this should be expected. I also calculate the three-body bound-state spectrum and how it changes with the two-body bound-state energy. This data is an essential part of computing Efimov's function for three-body bound states to high accuracy. This function can be used to determine the binding energies of three-body bound states when there is a large scattering length \cite{efimov:2}. Efimov's function is currently being used to study Feshbach resonances \cite{Braaten:2002er}.
281:
282:
283: %%%%%%%%%%%%%%%%%%%
284: %% NEW RESULTS %%
285: %%%%%%%%%%%%%%%%%%%
286: \section{New Results}
287:
288: To the best of the author's knowledge, this dissertation contains several new results, some of which are extensions to work done by Wilson \cite{Wilson:notes}:
289: \begin{itemize}
290: \item Correction of minor errors in the leading order equations and the derivation of the equations for first order corrections.
291: \item Modification of Wilson's original leading order equations to maintain high accuracy when $B_2 \simeq B_3$ in numerical calculations.
292: \item Detailed explanation of the derivation of the three-body integral equations and the use of Wilson's methods in that derivation.
293: \item Proof of and corrections to Wilson's assertion that certain integrals which are encountered when computing corrections to the leading order results are identically zero.
294: \item Derivation of new integration limits to simplify numerical calculations.
295: \item Implementation of numerical methods to solve leading order equations.
296: \item Use of leading order equations to compute Efimov's universal function to high accuracy.
297: \item Use of leading order equations to analytically and numerically verify similar work by others.
298: \end{itemize}
299:
300:
301: %%%%%%%%%%%%%%%%
302: %% OVERVIEW %%
303: %%%%%%%%%%%%%%%%
304: \section{Organization of Dissertation}
305:
306: After outlining the use of the $\delta^3(\vec{r})$ potential in the context of the two-body problem, I derive some formulas in Chapter 2 that will be of use later in the analysis of the three-body problem. Chapter 3 gives the derivation of the three-body bound-state equation and outlines some assumptions made in my calculations. I digress briefly in Chapter 4 to explain the method used for expanding the equations perturbatively and then apply the technique in Chapter 5. This results in the final form of the equations used for numerical computation. Chapter 6 outlines some details of the numerical technique and discusses some of the issues that arise from the need for high accuracy. Following this is a discussion of analytic and numerical results in Chapter 7. Chapter 8 is devoted to introducing Efimov's function and then numerically computing it using the leading order equations. I conclude in Chapter 9 with a brief discussion of areas for future work.