hep-ph0507111/eft.tex
1: \chapter{Effective Theories of Strong Interactions}
2: \label{chap:eft}
3: 
4: \newcommand{\glqq}{\raisebox{.2ex}{$\scriptscriptstyle\gg$}}
5: \newcommand{\grqq}{\raisebox{.2ex}{$\scriptscriptstyle\ll$}}
6: 
7: \begin{quote}
8: \small\singlespace\raggedright 
9: \emph{Theoria} bezeichnet die rein empfangende, von aller \glqq praktischen\grqq\ Bezweckung des t\"{a}tigen Lebens durchaus unabh\"{a}ngige Zuwendung zur Wirklichkeit. Man mag diese Zuwendung \glqq uninteressiert\grqq\ nennen---wenn hiermit nichts anderes ausgeschlossen sein soll als jegliches auf Dienlichkeiten und Belange gerichtete Absehen. Im \"{u}brigen ist hier auf h\"{o}chst entschiedene Weise Interesse, Beteiligung, Aufmerksamheit, Zielsetzung. \emph{Theoria} und \emph{contemplatio} zielen mit ihrer vollen Energie dahin---freilich: \emph{ausschlie\ss lich} dahin---, da\ss\ die ins Auge gefa\ss te Wirklichkeit offenbar und deutlich werde, da\ss\ sie sich zeige und enth\"{u}lle; sie zielen auf Wahrheit und nichts sonst. \\ 
10: \flushright\vspace{-12pt}\emph{Josef Pieper, \emph{Gl\"{u}ck und Kontemplation}} \\
11: \bigskip
12: \raggedright
13: \emph{Theoria} has to do with the purely receptive approach to reality, one altogether independent of all practical aims in active life. We may call this approach ``disinterested,'' in that it is altogether divorced from utilitarian ends. In all other respects, however, \emph{theoria} emphatically involves interest, participation, attention, purposiveness. \emph{Theoria} and \emph{contemplatio} devote their full energy to revealing, clarifying, and making manifest the reality which has been sighted; they aim at truth and nothing else. \\
14: \flushright\vspace{-12pt}\emph{Josef Pieper, \emph{Happiness \& Contemplation}}
15: \end{quote}
16: In this chapter we review the basic features of several effective theories for the strong interactions---heavy quark effective theory, soft-collinear effective theory, and non-relativistic QCD---following developments of these theories in recent years. This forms the background for the discussion of the applications of these effective theories pursued in the subsequent chapters.
17: 
18: \section{Quantum Chromodynamics}
19: \label{sec:QCD}
20: 
21: \begin{quote}
22: \small\singlespace\raggedright It is a lovely language, but it takes a very long time to say anything in it, because we do not say anything in it, unless it is worth taking a long time to say, and to listen to. \\
23: \flushright\vspace{-12pt}\emph{Treebeard, in \emph{The Two Towers}, by J.R.R. Tolkien}
24: \end{quote}
25: Quantum Chromodynamics (QCD) is the theory of the strong interactions between quarks and gluons, which bind together to make protons, neutrons, and other hadrons. The theory accounts successfully for many hadronic phenomena, especially those occurring at relatively large energies, for example, in the bombardment of protons by highly energetic electrons in deep inelastic scattering, revealing the pointlike substructure of the proton. QCD is governed by the Lagrangian density:
26: \begin{equation}
27: \label{QCDLag}
28: \mathcal{L}_{\text{QCD}} = \bar\psi_q(i\Dslash - m_q)\psi_q - \frac{1}{4}G^A_{\mu\nu}G^{A\mu\nu},
29: \end{equation}
30: where we sum over quark flavors $q$ and $SU(3)$ generators $A$. The covariant derivative is $D_\mu = \partial_\mu - igA_\mu^A T^A$.
31: 
32: QCD allows for reliable quantitative predictions at large energies due to the phenomenon of asymptotic freedom \cite{Politzer:1973,Gross:1973id}. The observed strength of the interaction between quarks and gluons is characterized by a coupling constant $g_s$, which is a function of energy, becoming small at large energies and large at small energies. Physical quantities can be calculated as a perturbation series in powers of $g_s$, or, rather, $\alpha_s = g_s^2/4\pi$, which is a reliable procedure as long as $\alpha_s\ll1$. 
33: 
34: At small energies, however, such as the scale at which quarks and gluons bind together into light hadrons (protons, pions, etc.), the coupling constant is large, $\alpha_s\gtrsim 1$, and perturbation theory breaks down completely. Even in processes involving strongly-interacting particles at large energies, because the particles used or observed directly in experiments are hadrons, not free quarks and gluons, these low-energy binding effects contaminate the analysis, preventing completely precise calculation of the observable quantities. However, it is often possible to separate the perturbatively-calculable large energy phenomena from the low-energy hadronization effects in a way that preserves much predictive power. Imagine calculating the total rate for $Z$ bosons to decay to hadrons, $\Gamma(Z\rightarrow\text{hadrons})$. In perturbation theory, we begin with the $Z$ coupling to quarks:
35: \begin{equation}
36: \mathcal{L}_{Z\bar q q} = \bar\psi_q\gamma_\mu(g_V + g_A\gamma_5)\psi_q Z^\mu.
37: \end{equation}
38: We can calculate the decay rate for $Z$ to $q\bar q$, $\Gamma(Z\rightarrow q\bar q)$. This process produces a $\bar q q$ pair moving back-to-back with energy $M_Z/2$. This is much larger than the scale of hadronization, $\Lambda_{\text{QCD}}$. The quark and antiquark move far apart before they hadronize. We do not know how to calculate the dynamics of this hadronization, but since we know it must happen with probability 1, we can approximate the total hadronic decay rate of the $Z$ by:
39: \begin{equation}
40: \Gamma(Z\rightarrow\text{hadrons}) = \Gamma(Z\rightarrow\bar q q),
41: \end{equation}
42: to leading order in $\alpha_s(M_Z)$. We can compute to next order in $\alpha_s(M_Z)$ by including the rate $\Gamma(Z\rightarrow\bar q q g)$. And so on. Because we have separated the parton-level physics from the longer-range, lower-energy physics of hadronization, we are able to make a reliable quantitative prediction that is a very good approximation to reality. We say we have ``factorized'' the total hadronic decay rate of the $Z$---into $\Gamma(Z\rightarrow\text{partons})$ and the probability for partons to hadronize, which is 1.
43: 
44: For more complicated observables, we seek to perform similar factorizations, although the low-energy, nonperturbative quantities which appear will not, in general, be so simple as ``1''. These may be objects such as parton distribution functions, meson light-cone wave functions, etc. Although such quantities are not calculable in perturbation theory, they may appear in more than one physical observable, thus providing some remnant of predictive power. They may also be calculable numerically in lattice QCD.
45: 
46: Proving factorization and identifying the relevant nonperturbative quantities is a major thrust in the direction of modern research in QCD. There are two main camps of research. There are those who attack the problem directly in QCD, the so-called perturbative QCD or QCD factorization approach \cite{StermanTasi}. Then there are those who quail at the enormity of full QCD and attempt to work instead with a simpler version of the theory. This is the approach of \emph{effective field theory} \cite{EFTnotes,Rothstein:2003mp}, which we adopt in this thesis.
47: 
48: \section{Example: Heavy Quark Effective Theory}
49: 
50: In an effective field theory one identifies a small parameter determined by the relevant physics in a given problem in which to expand the full theory to a given order of approximation. Equivalently one integrates out the degrees of freedom living at an energy scale much larger than those relevant in the physical problem at hand.\footnote{The presentation in this and the following section are heavily influenced by the lectures presented in Ref.~\cite{SCETlectures}.}
51: 
52: 
53: As an illustrative example and a prelude to the effective field theories we consider later in this thesis, we overview the \emph{heavy quark effective theory} (HQET) \cite{HQET}, an effective field theory describing hadrons containing one heavy quark, namely, charm ($c$) or bottom ($b$).\footnote{We limit our attention, not surprisingly, to heavy hadrons containing two ($Q\bar q$) or three ($Qqq$), not five \cite{Stewart:2004pd,Wessling:2004ag}, quarks and antiquarks.} In such a hadron, the heavy $c$ or $b$ quark moves with a relatively slow velocity compared to the other light quark(s) accompanying it. The typical momenta of the constituent partons is of order $p\sim\lqcd$. Thus, the velocity of the heavy quark $Q$ is of the order $v\sim\Lambda_{\text{QCD}}/m_Q$. We can expand the Lagrangian of full QCD in Eq.~(\ref{QCDLag}) for heavy quarks in powers of this small velocity $v$.
54: 
55: Physically, we imagine that the interaction of a heavy quark with low-energy gluons causes only small fluctuations of its momentum, $p = m_Q v + k$, with $k\sim\lqcd$. We want the effective theory to describe these small fluctuations. First, we remove from the full QCD heavy quark field $Q(x)$ the dependence on the large momentum $p = m_Q v$, defining a new field $Q_v$:
56: \begin{equation}
57: Q(x) = \sum_{v}e^{-im_Qv\cdot x}Q_v(x),
58: \end{equation}
59: summing over labels $v$. In terms of the new fields, the quark part of the QCD Lagrangian becomes:
60: \begin{equation}
61: \label{QvLag}
62: \mathcal{L}_Q = \bar Q_v(x)(i\Dslash + m_Q\vslash - m_Q)Q_v(x),
63: \end{equation}
64: summing implicitly over $v$.\footnote{Strictly speaking, we should have summed over separate labels $v,v'$ for the two heavy quark fields, but interactions with soft gluons cannot change these label velocities; hence we sum over only a single velocity. This simplification will not occur in SCET.} Now write $Q_v(x)$ as the sum  $Q_v(x) = h_v(x) + H_v(x)$, where:
65: \begin{equation}
66: h_v(x) = \frac{1+\vslash}{2}Q_v(x),\qquad H_v(x) = \frac{1-\vslash}{2}Q_v(x).
67: \end{equation}
68: Then,  the heavy quark Lagrangian (\ref{QvLag}) becomes:
69: \begin{equation}
70: \label{hHLag}
71: \begin{split}
72: \mathcal{L}_Q &= \bar h_v(x)i\Dslash h_v(x) + \bar H_v(x)(i\Dslash - 2m_Q)H_v(x) \\&\quad + \bar H_v(x)i\Dslash h_v(x) + \bar h_v(x)i\Dslash H_v(x).
73: \end{split}
74: \end{equation}
75: Derivatives acting on the fields $h_v,H_v$ produce momenta of order $\lqcd/m_Q$, so the kinetic term for $H_v(x)$ is suppressed relative the leading term quadratic in $H_v$, $-2m_Q\bar H_v(x) H_v(x)$. $H_v$ therefore is not a dynamical field and can be integrated out by using the equation of motion (obtained by varying with respect to $\bar H_v$):
76: \begin{equation}
77: (i\Dslash - 2m_Q)H_v = -i\Dslash h_v.
78: \end{equation}
79: Substituting into Eq.~(\ref{hHLag}), we obtain
80: \begin{equation}
81: \mathcal{L}_Q = \bar h_v(x)i\Dslash h_v(x) - \bar h_v(x) i\Dslash\frac{1}{i\Dslash-2m_Q}i\Dslash h_v(x).
82: \end{equation}
83: Expanding in powers of $1/m_Q$:
84: \begin{equation}
85: \label{hvLag}
86: \mathcal{L}_Q = \bar h_v(x)\left(iv\mcdot D - \frac{1}{2m_Q}\Dslash\Dslash + \cdots\right)h_v(x),
87: \end{equation}
88: having used $P_v \gamma^\mu P_v = v^\mu$, where $P_v = (1+\vslash)/2$. The dots denote higher-order terms in the $1/m_Q$ expansion.
89: 
90: Keeping terms only to a fixed order in $1/m_Q$ in Eq.~(\ref{hvLag}) defines the Lagrangian of \emph{heavy quark effective theory}. The leading order term is very simple:
91: \begin{equation}
92: \mathcal{L}_{\text{HQET}} = \bar h_v(x) (iv\mcdot D) h_v(x),
93: \end{equation}
94: giving the heavy quark propagator
95: \begin{equation}
96: i\frac{1+\vslash}{2}\frac{1}{v\mcdot k + i\epsilon}
97: \end{equation}
98: and the heavy quark-gluon interaction vertex
99: \begin{equation}
100: \label{HQETvertex}
101: -ig T^A v^\mu.
102: \end{equation}
103: This leading-order Lagrangian exhibits more symmetries than evident in the full QCD Lagrangian. First, it is independent of the heavy quark mass, giving rise to a flavor symmetry between $b$ and $c$ quarks. Second, containing no Dirac matrices, it also exhibits spin symmetry. This heavy quark spin-flavor symmetry is thus an approximate symmetry of full QCD made evident only by the effective theory expansion in $1/m_Q$. It allows for powerful predictions for processes involving different hadrons containing heavy quarks which are related by spin-flavor symmetry. Violations of this symmetry at higher order in $1/m_Q$ can be systematically calculated by keeping more terms in the HQET Lagrangian (\ref{hvLag})\footnote{Or estimated by being more clever \cite{Dorsten:2003ru}.}.
104: 
105: This relatively simple example of HQET illustrates two key features we will seek in formulating any effective field theory:
106: \begin{itemize}
107: \item Extra, approximate symmetries not obvious in the full theory but easily identified at leading order(s) in the the effective theory.
108: 
109: \item The simplification of interactions between heavy (hard) particles and soft ones, as in Eq.~(\ref{HQETvertex}).
110: \end{itemize}
111: The first property grants us more predictive power in relating apparently different physical processes to one another, while more sophisticated exploitation of the latter will simplify proofs of factorization of perturbative and nonperturbative contributions to physical observables. We now embark on this task in the effective theory truly of interest in this thesis.
112: 
113: \section{Soft-Collinear Effective Theory}
114: 
115: Consider a process in which there are hadrons moving with very large energy compared to their invariant mass, for instance, in the decay of $Z$ bosons to light hadrons as considered in Sec.~\ref{sec:QCD}, or in $B$ decays such as $B\rightarrow D\pi$ or $B\rightarrow \pi\pi$. We can formulate an effective field theory for quarks and gluons moving on such collinear trajectories and the soft partons with which they interact---the \emph{soft-collinear effective theory} \cite{SCET1a,SCET1b}.
116: 
117: \subsection{Effective Theory for Inclusive Decays---SCET$_{\rm I}$}
118: 
119: We first identify the relevant energy scales and small parameters to use in formulating the effective theory. Recall that in HQET, we took this to be the heavy quark mass $m_Q$ and velocity $v\sim\lqcd/m_Q$. Here, we begin by defining two light-cone vectors, $n$ and $\bar n$, which satisfy
120: \begin{equation}
121: n^2 = \bar n^2 = 0,\quad n\mcdot\bar n = 2,
122: \end{equation}
123: for example, $n = (1,0,0,1)$ and $\bar n = (1,0,0,-1)$. Vectors can be decomposed along these two light-cone directions and the orthogonal transverse directions:
124: \begin{equation}
125: V^\mu = \frac{\bar n\mcdot V}{2} n^\mu + \frac{n\mcdot V}{2}\bar n^\mu + V_\perp^\mu.
126: \end{equation}
127: We thus define the light-cone components $V^+ = n\mcdot V$, $V^- = \bar n\mcdot V$. For a collinear particle, one light-cone component of its momentum will be large, the other small. Its momentum will scale in powers of some small parameter $\lambda$ as:
128: \begin{equation}
129: (p^+,p^-,p_\perp)\sim Q(\lambda^2,1,\lambda),
130: \end{equation}
131: where $Q$ is the large scale governing the physical process being studied, on which the definition of $\lambda$ also depends. The collinear particles also interact with soft, or ultrasoft, particles, with momenta scaling as:
132: \begin{subequations}
133: \begin{align}
134: \text{soft:}&\quad p_s\sim Q\lambda \\
135: \text{ultrasoft:}&\quad p_{us}\sim Q\lambda^2.
136: \end{align}
137: \end{subequations}
138: The effective theory for collinear and (ultra)soft particles with these scalings is called SCET$_{\rm I}$. Another choice of scalings gives the theory SCET$_{\rm II}$, to be introduced in the next section.
139: 
140: We form the \SCETa\ Lagrangian essentially by expanding the QCD Lagrangian in powers of $\lambda$. We start with the Lagrangian for collinear quarks, which we take to be massless. In QCD,
141: \begin{equation}
142: \label{masslessQCD}
143: \mathcal{L}_q = \bar q i\Dslash q.
144: \end{equation}
145: We seek to describe fluctuations in the momenta of the collinear particles about its collinear trajectory. So, for a collinear quark of momentum $p$, we write:
146: \begin{equation}
147: \label{decomposition}
148: p = \tilde p + k,
149: \end{equation}
150: where $\tilde p$ contains the large (\emph{label}) components of the momentum,
151: \begin{equation}
152: \tilde p^\mu = \tilde p^-\frac{n^\mu}{2} + \tilde p_\perp,
153: \end{equation}
154: and the \emph{residual} momentum $k\sim\lqcd$ represents the fluctuations about the label momentum.
155: 
156: As in HQET, we extract the large momentum fluctuations from the full QCD field, defining a new field $q_{n,p}$:
157: \begin{equation}
158: q(x) = \sum_{\tilde p} e^{-i\tilde p\cdot x}q_{n,p}(x).
159: \end{equation}
160: The QCD Lagrangian (\ref{masslessQCD}) then becomes:
161: \begin{equation}
162: \label{qnpLag}
163: \mathcal{L}_q = \sum_{\tilde p,\tilde p'} e^{-i(\tilde p - \tilde p')\cdot x}\bar q_{n,p'}(x)(\diracslash{\tilde p} + i\Dslash)q_{n,p}(x).
164: \end{equation}
165: Again as in HQET, we project out the large and small components of Dirac field, writing $q_{n,p} = \xi_{n,p} + \Xi_{n,p}$, where:
166: \begin{equation}
167: \xi_{n,p} = \frac{\nslash\bnslash}{4}q_{n,p},\quad \Xi_{n,p} = \frac{\bnslash\nslash}{4}q_{n,p},
168: \end{equation}
169: in terms of which the Lagrangian (\ref{qnpLag}) becomes:
170: \begin{equation}
171: \label{xiXiLag}
172: \begin{split}
173: \mathcal{L}_q = \sum_{\tilde p,\tilde p'}e^{-i(\tilde p - \tilde p')\cdot x}\biggl[&\bar\xi_{n,p'}(x)\frac{\bnslash}{2}in\mcdot D\xi_{n,p}(x) + \overline{\Xi}_{n,p'}(x)\frac{\nslash}{2}(\tilde p^- + i\bar n\mcdot D)\Xi_{n,p}(x) \\
174: &+ \bar\xi_{n,p'}(x)(\diracslash{\tilde p}_\perp + i\Dslash_\perp)\Xi_{n,p}(x) + \overline{\Xi}_{n,p'}(x)(\diracslash{\tilde p}_\perp + i\Dslash_\perp)\xi_{n,p}(x)\biggr].
175: \end{split}
176: \end{equation}
177: Again, as in HQET, we have a derivative, $\bar n\cdot\partial$, acting on $\Xi_{n,p}$, which is suppressed relative to the term containing the label momentum $\tilde p^-$. So the field $\Xi_{n,p}$ is not dynamical, and we eliminate it using its equation of motion:
178: \begin{equation}
179: \Xi_{n,p}(x) = \frac{1}{\tilde p^- + i\bar n\mcdot D}(\diracslash{\tilde p}_\perp + i\Dslash_\perp)\frac{\bnslash}{2}\xi_{n,p}(x)
180: \end{equation}
181: Inserting this into (\ref{xiXiLag}) leaves us with the Lagrangian:
182: \begin{equation}
183: \label{xiLag}
184: \mathcal{L}_\xi = \sum_{\tilde p,\tilde p'}e^{-i(\tilde p - \tilde p')\cdot x}\bar\xi_{n,p'}\biggl[in\mcdot D + (\diracslash{\tilde p}_\perp + i\Dslash_\perp)\frac{1}{\tilde p^- + i\bar n\mcdot D}(\diracslash{\tilde p}_\perp + i\Dslash_\perp)\biggr]\frac{\bnslash}{2}\xi_{n,p}(x).
185: \end{equation}
186: To sort out the interactions between gluons and collinear quarks, we must first distinguish between the collinear and (ultra)soft gluon fields. The gluon fields appearing in the covariant derivatives in the collinear quark Lagrangian (\ref{xiLag}) can be split up:
187: \begin{equation}
188: A = A^c + A^s + A^{us},
189: \end{equation}
190: where the collinear, soft, and ultrasoft fields are assigned the power countingas:
191: \begin{equation}
192: \label{Apowers}
193: A^c\sim Q(\lambda^2,1,\lambda),\quad A^s\sim Q(\lambda,\lambda,\lambda),\quad A^{us}\sim Q(\lambda^2,\lambda^2,\lambda^2),
194: \end{equation}
195: to match the scalings of the corresponding momenta. Note that interactions of soft gluons with collinear quarks leave the quark with the momentum scaling $Q(\lambda,1,\lambda)$, which does not exist in the effective theory. So soft gluons should not appear in the Lagrangian (\ref{xiLag}). Collinear quarks do interact with ultrasoft gluons. The effective theory collinear gluon fields $A_{n,q}^c$ are defined by:
196: \begin{equation}
197: \label{Acdef}
198: A^c(x) = \sum_{\tilde q}e^{-i\tilde q\cdot x}A_{n,q}^c(x),
199: \end{equation}
200: factoring out the large label momentum $\tilde q$. 
201: \setlength{\unitlength}{1pt}
202: \begin{figure}[]
203: \begin{align*}
204: \parbox{42mm}{
205: \begin{fmffile}{xi}
206: \begin{fmfgraph*}(100,30)
207: \fmfpen{thin}
208: \fmfleft{i}
209: \fmfright{o}
210: \fmf{dashes_arrow}{i,o}
211: \end{fmfgraph*}
212: \end{fmffile}}
213: &= i\frac{\Diracslash{\bar n}}{2}\frac{\bar n\cdot \tilde p}{n\cdot k\,\bar n\cdot \tilde p + \tilde p_\perp^2 + i\epsilon} \\ & \\
214: \raisebox{-2cm}{
215: \begin{fmffile}{xiAu}
216: \begin{fmfgraph*}(100,80)
217: \fmfleft{xi1}
218: \fmfright{xi2}
219: \fmftop{Au}
220: \fmf{dashes_arrow}{xi1,v}
221: \fmf{dashes_arrow}{v,xi2}
222: \fmf{gluon,tension=0}{v,Au}
223: \fmflabel{$\noexpand\mu,A$}{Au}
224: \end{fmfgraph*}
225: \end{fmffile}}
226: &= igT^An_\mu\frac{\Diracslash{\bar n}}{2} \\
227: \raisebox{-2cm}{
228: \begin{fmffile}{xiAc}
229: \begin{fmfgraph*}(100,80)
230: \fmfleft{xi1}
231: \fmfright{xi2}
232: \fmftop{Ac}
233: \fmf{dashes_arrow}{xi1,v}
234: \fmf{dashes_arrow}{v,xi2}
235: \fmf{gluon,tension=0}{v,Ac}
236: \fmf{plain,tension=0}{v,Ac}
237: \fmflabel{$p$}{xi1}
238: \fmflabel{$p'$}{xi2}
239: \fmflabel{$\noexpand\mu,A$}{Ac}
240: \end{fmfgraph*}
241: \end{fmffile}}
242: &= igT^A\left(n_\mu + \frac{\gamma^\perp_\mu\diracslash{\tilde p}_\perp}{\bar n\cdot\tilde p} + \frac{\diracslash{\tilde p}'_\perp\gamma^\perp_\mu}{\bar n\cdot \tilde p'} - \frac{\diracslash{\tilde p}'_\perp\diracslash{\tilde p}_\perp}{\bar n\cdot \tilde p\ \bar n\cdot \tilde p'}\bar n_\mu\right)\frac{\Diracslash{\bar n}}{2}
243: \end{align*}
244: \setlength{\unitlength}{1mm}
245: \begin{picture}(0,0)(0,0)
246: \put(24,79){$\tilde p + k$}
247: \put(21,21){$\tilde p$}
248: \put(41,21){$\tilde p'$}
249: \end{picture}
250: \vspace{-1cm}
251: \caption[Feynman rules involving collinear quarks in \SCETa.]{Feynman rules involving collinear quarks in \SCETa. The collinear particles are shown with label momenta $\tilde p,\tilde p'$ and residual momentum $k$.}
252: \label{scetrules}
253: \end{figure}
254: 
255: In the Lagrangian (\ref{xiLag}), the components $D_\perp$ and $\bar n\cdot D$ of the covariant derivative contain both ultrasoft and collinear gluons. Due to the power counting in Eq.~(\ref{Apowers}), the ultrasoft gluon fields give a subdominant contribution compared to that of the collinear gluons. Thus, the correct collinear quark Lagrangian in \SCETa\ at leading order in $\lambda$ is:
256: \begin{equation}
257: \label{SCETLag1}
258: \mathcal{L}_{\text{\SCETa}} = \sum_{\tilde p,\tilde p'}e^{-i(\tilde p - \tilde p')\mcdot x}\bar\xi_{n,p'}\biggl[in\mcdot D + (\diracslash{\tilde p}_\perp + i\Dslash_\perp^c)\frac{1}{\tilde p^- + i\bar n\mcdot D^c}(\diracslash{\tilde p}_\perp + i\Dslash_\perp^c)\biggr]\frac{\bnslash}{2}\xi_{n,p}(x),
259: \end{equation}
260: where $D^c = \partial - igA^c$ contains only the collinear gluon field, while $n\mcdot D$ still contains both collinear and ultrasoft fields. Some of the Feynman rules arising from this Lagrangian are shown in Fig.~\ref{scetrules}.
261: 
262: 
263: The Lagrangian (\ref{SCETLag1}) may be written in a still more compact form by the introduction of the \emph{label operators} \cite{Bauer:2001ct}:
264: \begin{subequations}
265: \label{labelopdefs}
266: \begin{align}
267: \bar n\mcdot\mathcal{P} \xi_{n,p} &= \bar n\mcdot \tilde p\,\xi_{n,p} \\
268: \mathcal{P}_\perp^\mu\xi_{n,p} &= \tilde p_\perp^\mu\xi_{n,p},
269: \end{align}
270: \end{subequations}
271: and similarly for label operators acting on collinear gluon fields. In the second, messier, term of the Lagrangian (\ref{SCETLag1}), the ordinary derivatives which appear hit everything to their right. Thus, they bring down extra factors of the label momenta of the collinear gluons, because of the exponential factor in Eq.~(\ref{Acdef}). For example,
272: \begin{equation}
273: \bar n\mcdot\partial A^c_\perp(x) = \bar n\mcdot\partial\sum_{\tilde q}e^{-i\tilde q\cdot x}A_{n,q}^{c\perp}(x) = \sum_{\tilde q} e^{-i\tilde q\cdot x}\bar n\mcdot\tilde qA_{n,q}^{c\perp}(x).
274: \end{equation}
275: We can account for derivatives on the exponential factors and eliminate the explicit label momenta appearing in the Lagrangian (\ref{SCETLag1}) by making use of the label operators in (\ref{labelopdefs}):
276: \begin{equation}
277: \label{SCETLag2}
278: \mathcal{L}_{\text{\SCETa}} = \bar\xi(x)\biggl[in\mcdot D + i\Dslash^c_\perp \frac{1}{i\bar n\mcdot D^c}i\Dslash^c_\perp\biggr]\frac{\bnslash}{2}\xi_n(x),
279: \end{equation}
280: where now
281: \begin{equation}
282: D^c_\mu = \mathcal{P}_\mu - igA^c_\mu,
283: \end{equation}
284: and
285: \begin{equation}
286: \xi_n(x) = \sum_{\tilde p}e^{-i\tilde p\cdot x}\xi_{n,p}(x).
287: \end{equation}
288: Finally, note that this Lagrangian contains the collinear gluon field $\bar n\cdot A^c$ in the denominator of the second term, giving rise to couplings of collinear quarks to an arbitrary number of $\bar n\mcdot A^c$ gluons. These infinitely many couplings can be resummed into the form of \emph{Wilson lines}:
289: \begin{equation}
290: W_n(x) = P\exp\biggl[ig\int_{-\infty}^x ds\,\bar n\mcdot A^c(x)\biggr],
291: \end{equation}
292: where $P$ denotes path ordering. Using the property 
293: \begin{equation}
294: W_n(x)\bar n\mcdot\mathcal{P} W_n^\dag(x) = \bar n\mcdot\mathcal{P} -i g\bar n\mcdot A^c(x) = \bar n\mcdot D^c,
295: \end{equation}
296: we can express the \SCETa\ Lagrangian (\ref{SCETLag2}) as:
297: \begin{equation}
298: \label{SCETLag3}
299: \mathcal{L}_{\text{\SCETa}} = \bar\xi_n(x)\biggl[in\mcdot D + i\Dslash^c_\perp W_n(x)\frac{1}{i\bar n\mcdot\mathcal{P}}W_n^\dag(x)i\Dslash^c_\perp\biggr]\frac{\bnslash}{2}\xi_n(x).
300: \end{equation}
301: It is also possible to argue that invariance under gauge transformations of collinear field requires this precise form of the Lagrangian, which we do not go through here. 
302: 
303: Similar analysis leads also to the effective Lagrangian for collinear gluons \cite{SCET3}:
304: \begin{equation}
305: \label{SCETgluonLag}
306: \mathcal{L}_c^{(g)} = \sum_{\tilde q,\tilde q'}e^{-i(\tilde q - \tilde q')\cdot x}\frac{1}{2g^2}\Tr\left\{\left[i\mathcal{D}^\mu + gA_{n,q},i\mathcal{D}^\nu + gA_{n,q'}^\nu\right]\right\}^2,
307: \end{equation}
308: where
309: \begin{equation}
310: \mathcal{D}^\mu = \bar n\mcdot\mathcal{P}\frac{n^\mu}{2} + \mathcal{P}_\perp^\mu + in\mcdot D\frac{\bar n^\mu}{2}.
311: \end{equation}
312: Ghost fields and gauge-dependent terms have been ignored in Eq.~(\ref{SCETgluonLag}).
313: 
314: \subsection{Decoupling Ultrasoft Fields}
315: 
316: We could take Eq.~(\ref{SCETLag3}) as the final form of our Lagrangian for \SCETa. However, one more manipulation simplifies greatly the separation of hard and soft (nonperturbative) physics in the effective theory \cite{SCET3}. First, introduce the ultrasoft Wilson lines:
317: \begin{equation}
318: Y_n(x) = P\exp\biggl[ig\int_{-\infty}^x ds\,n\mcdot A_{us}(ns)\biggr].
319: \end{equation}
320:  Rewrite the fields in (\ref{SCETLag3}) as:
321: \begin{subequations}
322: \label{BPSredef}
323: \begin{align}
324: \xi_n(x) &= Y_n(x)\xi_n^{(0)}(x) \\
325: A_n(x) &= Y_n(x)A_n^{(0)}(x)Y_n^\dag(x) \\
326: W_n(x) &= Y_n(x)W_n^{(0)}(x)Y_n^\dag(x).
327: \end{align}
328: \end{subequations}
329: (The last rule actually follows from the second.) From the property
330: \begin{equation}
331: [(n\mcdot\partial - ig n\mcdot A_{us})Y_n(x)] = 0,
332: \end{equation}
333: we find that upon making the replacements (\ref{BPSredef}) in the \SCETa\ Lagrangian (\ref{SCETLag3}), the term containing the ultrasoft gluon disappears! That is,
334: \begin{equation}
335: \bar\xi_n(x)in\mcdot D\frac{\bnslash}{2}\xi_n(x) \rightarrow \bar\xi_n^{(0)}(x)in\cdot(\partial - igA_n^{(0)})\frac{\bnslash}{2}\xi_n^{(0)}(x).
336: \end{equation}
337: Thus, in terms of the fields $\xi_n^{(0)},A_n^{(0_)}$, there are no couplings of collinear particles to ultrasoft gluons at leading order in \SCETa. The same decoupling occurs in the gluon Lagrangian (\ref{SCETgluonLag}). Since nonperturbative physics is governed by interactions with ultrasoft gluons, this decoupling greatly simplifies the separation of perturbative and nonperturbative contributions to physical observables in the effective theory.
338: 
339: Ultrasoft gluons have not entirely disappeared from the theory, however. When we match currents or operators mediating various decays from QCD onto \SCETa, we must include ultrasoft Wilson lines $Y_n(x)$ whenever we have a collinear field in the operator, according to Eqs.~(\ref{BPSredef}). (Equivalently, we must ensure that all operators are invariant under ultrasoft gauge transformations.) These rules will become apparent in the applications presented in subsequent chapters.
340: 
341: 
342: \subsection{Effective Theory for Exclusive Decays---\SCETb}
343: 
344: We now have all the tools in the effective theory required to describe inclusive decays such as in $Z$ decays to hadrons in Chap.~\ref{chap:jet}. Before proceeding to this application, however, let us introduce the novel features of a second version of SCET---\SCETb\ \cite{Bauer:2002aj}--- required to analyze exclusive decays as in Chap.~\ref{chap:ups} on radiative exclusive $\Upsilon$ decays.
345: 
346: To see the inadequacy of \SCETa\ in analyzing an exclusive decay producing energetic light hadrons, such as $\Upsilon\rightarrow\gamma\pi\pi$, consider the typical invariant mass of these light hadrons. For a particular light hadron, such as the pion, the invariant mass squared is of the order $\sim\lqcd^2$. The typical momentum of a collinear particle in \SCETa, however, scales as:
347: \begin{equation}
348: (p^+, p^-, p_\perp) \sim Q(\lambda^2,1,\lambda),
349: \end{equation}
350: where $\lambda = \sqrt{\lqcd/Q}$. The invariant mass squared of such a particle is:
351: \begin{equation}
352: p^2 = p^+ p^- - p_\perp^2 \sim Q^2\lambda^2 \sim Q\lqcd,
353: \end{equation}
354: which is too large to represent a light particle with $m^2\sim \lqcd^2$.
355: 
356: The correct effective theory is formed by using instead the parameter $\eta = \lqcd/Q$ and collinear fields which scale as:
357: \begin{equation}
358: \label{collscalingII}
359: (p^+, p^-, p_\perp) \sim Q(\eta^2,1,\eta),
360: \end{equation}
361: giving an invariant mass squared of $p^2\sim Q^2\eta^2 = \lqcd^2$, which is now correct for an exclusive light hadron. These collinear fields interact with soft fields, whose momenta scale as:
362: \begin{equation}
363: \label{softscaling}
364: p_{s}\sim Q(\eta,\eta,\eta),
365: \end{equation}
366: with invariant mass squared $p_s^2\sim \lqcd^2$. The effective theory for collinear and soft degrees of freedom with these scalings is called \SCETb.
367: 
368: The soft fields of \SCETb\ are in fact the same as the ultrasoft fields in \SCETa. The collinear fields, however, live at a lower energy scale in \SCETb\ than they do in \SCETa. For this reason, the collinear fields in \SCETa\ are often called \emph{hard-collinear} fields to distinguish them from the collinear fields of \SCETb. The introduction of another degree of freedom, the \emph{soft-collinear messenger} mode, was introduced in Refs.~\cite{Becher:2003qh,Becher:2003kh} so that all infrared divergences in one-loop graphs in \SCETb\ and QCD would match, which can also be accomplished by introducing a proper infrared regulator in the effective theory, as in Ref.~\cite{Bauer:2003td}. We do not evaluate any loop graphs with these divergences in the subsequent chapters, and so blissfully ignore these modes or regulators.
369: 
370: To work in \SCETb, one can imagine matching directly from QCD onto \SCETb. However, it is more convenient to match QCD first onto \SCETa, and then match \SCETa\ onto \SCETb, since we have already done the hard work of completing the first step. In the second step, we take the collinear fields of \SCETa, and lower their off-shellness from the scale $\sqrt{Q\lqcd}$ to $\lqcd$, putting them in \SCETb. Interactions with ultrasoft gluons are already removed from the \SCETa\ collinear quark and gluon Lagrangians via the field redefinition (\ref{BPSredef}). We simply change ultrasoft Wilson lines $Y_n(x)$ in \SCETa\ into soft Wilson lines $S_n(x)$ in \SCETb\, which look the same as in \SCETa\, but contain the soft gluon fields of \SCETb\ with the scaling (\ref{softscaling}).
371: 
372: Note that in \SCETb, there could not be interactions of collinear quarks with soft gluons anyway, because the soft gluon knocks the collinear quark off shell, due to the scalings in (\ref{collscalingII}) and (\ref{softscaling}).
373: 
374: We will make use of \SCETb\ in Chap.~\ref{chap:ups}, where the procedures described above will be illustrated by explicit example.
375: 
376: \subsection{Reparametrization Invariance}
377: \label{sec:RPI}
378: 
379: The introduction of the light-cone vectors $n,\bar n$ along which to decompose felds and momenta in the effective theory breaks the Lorentz invariance of the original theory of QCD. The Lorentz symmetry manifests itself in the effective theory by means of invariance under redefinitions of the vectors $n,\bar n$ and of the label momenta assigned to collinear particles. These features are known as \emph{reparametrization invariance} (RPI) \cite{Manohar:2002fd}.
380: 
381: We focus here on the theory \SCETa\ for simplicity. The first invariance requirement arises from the fact that the decomposition of collinear momenta into label and residual components is not unique. We wrote in Eq.~(\ref{decomposition}) for a collinear momentum $p$,
382: \begin{equation}
383: p = \tilde p + k,
384: \end{equation}
385: where $\tilde p$ is contains the large $\mathcal{O}(Q)$ and $\mathcal{O}(Q\lambda)$ label momenta, and $k$ is the $\mathcal{O}(Q\lambda^2)$ residual momentum. Shifting an amount of momentum of order $Q\lambda^2$ between the label and residual momentum should yield an equivalent description of the collinear physics. The main consequence of this invariance, together with gauge invariance, is that operators in the effective theory containing derivatives must appear in the combination:
386: \begin{equation}
387: \label{DcDus}
388: \mathcal{D}^\mu = D_c^\mu + D_{us}^\mu,
389: \end{equation}
390: where $D_c^\mu = \mathcal{P}^\mu - igA_{n,q}^\mu$, and $D_{us}^\mu = \partial^\mu - ig A_{us}^\mu$. This condition relates operators at different orders in $\lambda$, since the terms contained in Eq.~(\ref{DcDus}) have different scalings in $\lambda$. In this thesis we will focus on the constructions of operators only at leading order in $\lambda$, so this constraint will not be crucial.
391: 
392: More relevant constraints arise from the second type of invariance, that of redefinitions of the light-cone vectors $n,\bar n$ themselves. These vectors enter the very definition of the SCET Lagrangian, and the theory must remain invariant for equivalent choices of $n,\bar n$, which are those that leave invariant the conditions:
393: \begin{equation}
394: n^2 = \bar n^2 = 0,\qquad n\mcdot\bar n = 2,
395: \end{equation}
396: must yield an equivalent theory. There are three classes of transformations we can make. In the first two, we fiddle just with the transverse components of one or the other of $n,\bar n$:
397: \begin{equation}
398: \text{Type I:}\quad
399: \begin{cases}
400: n_\mu \rightarrow n_\mu + \Delta^\perp_\mu \\
401: \bar n_\mu\rightarrow\bar n_\mu
402: \end{cases}
403: \qquad
404: \text{Type II:}
405: \quad
406: \begin{cases}
407: n_\mu \rightarrow n_\mu \\
408: \bar n_\mu\rightarrow\bar n_\mu + \varepsilon^\perp_\mu
409: \end{cases}
410: ,
411: \end{equation}
412: and in the last class we mutually rescale the vectors:
413: \begin{equation}
414: \text{Type III:}
415: \quad
416: \begin{cases}
417: n_\mu \rightarrow (1+\alpha)n_\mu \\
418: \bar n_\mu\rightarrow(1-\alpha)\bar n_\mu 
419: \end{cases},
420: \end{equation}
421: The parameters $\Delta^\perp,\varepsilon^\perp,\alpha$ are infinitesimal, but can be assigned a particular scaling in $\lambda$. Starting with type-II transformations, consider the change induced in the light-cone components of a vector $V^\mu$. We start with the decomposition:
422: \begin{equation}
423: V^\mu = \bar n\mcdot V\frac{n^\mu}{2} + n\mcdot V\frac{\bar n^\mu}{2} + V_\perp^\mu\ .
424: \end{equation}
425: Under the transformation, the components become:
426: \begin{subequations}
427: \begin{align}
428: n\mcdot V &\rightarrow n\mcdot V \\
429: \bar n\mcdot V &\rightarrow (\bar n + \varepsilon_\perp)\mcdot V = \bar n\mcdot V + \varepsilon_\perp\mcdot V_\perp \\
430: V_\perp^\mu &= V^\mu - \bar n\mcdot V\frac{n^\mu}{2} - n\mcdot V\frac{\bar n^\mu}{2} \nonumber \\
431: &\rightarrow V^\mu - (\bar n + \varepsilon_\perp)\mcdot V\frac{n^\mu}{2} - n\mcdot V\frac{\bar n^\mu + \varepsilon_\perp^\mu}{2} \\
432: &= V_\perp^\mu - \varepsilon_\perp\mcdot V_\perp\frac{n^\mu}{2} - n\mcdot V\frac{\varepsilon_\perp^\mu}{2}\ . \nonumber
433: \end{align}
434: \end{subequations}
435: If $V$ is, say, a collinear momentum or field with scaling $(V^+,V^-,V_\perp)\sim Q(\lambda^2,1,\lambda)$, then we see from the shifts in $\bar n\mcdot V$ and $V_\perp$ that we can assign a scaling of 1 to the parameter $\varepsilon_\perp$ without messing up the power counting of $p$. Similar examination of type-I and type-III transformation reveals that $\Delta_\perp$ cannot have a scaling larger then $\lambda$, while $\alpha$ can also be order 1. Therefore, type-II and type-III RPI can constrain operators at the same order in $\lambda$, while type-I will relate operators at different orders in $\lambda$. We will make use of type-II and type-III RPI to restrict the number of operators which can appear in the description of $\Upsilon$ decays in Chap.~\ref{chap:ups}.
436: 
437: 
438: 
439: 
440: \section{Non-Relativistic QCD}
441: 
442: We have described effective theories in QCD for hadrons containing a single heavy quark (HQET) and for light, energetic hadrons (SCET). We make use in this thesis of one more effective theory, that for mesons containing a heavy quark-antiquark pair, $Q\bar Q$. In such a meson, both heavy partons move with relatively small velocity. Thus, we are led to expand QCD about its non-relativistic limit---hence, the effective theory of \emph{non-relativistic QCD} (NRQCD) \cite{Bodwin:1995jh,Braaten:1996ix,Luke:2000kz}.
443: 
444: In NRQCD, we separate the fields in the QCD Lagrangian which create quarks and antiquarks, into the two-component spinor fields $\psi$ and $\chi$, respectively. The small expansion in NRQCD is the heavy quark velocity $v$. This parameter now determines two separate physical scales, not just one: the heavy quark three-momentum is of the order $\abs{\vect{p}}\sim m_Q v$, while the kinetic energy is of order $E\sim m_Q v^2$.
445: 
446: To form the NRQCD Lagrangian for heavy quarks, begin again in QCD:
447: \begin{equation}
448: \label{QCDLagQ}
449: \mathcal{L}_Q = \bar\Psi_Q(i\Dslash - m_Q)\Psi_Q.
450: \end{equation}
451: The Dirac field $\Psi_Q$ contains fields creating/annihilating both quarks and antiquarks:
452: \begin{equation}
453: \Psi_Q = \begin{pmatrix} \psi \\ \chi \end{pmatrix},
454: \end{equation} 
455: in terms of which the Lagrangian (\ref{QCDLagQ}) is:
456: \begin{equation}
457: \begin{split}
458: \mathcal{L}_Q = &\psi^\dag(iD_0 - m_Q)\psi + \chi^\dag(iD_0 + m_Q)\chi \\
459: &+ \psi^\dag i\boldsigma\cdot\vect{D}\chi + \chi^\dag i\boldsigma\cdot\vect{D}\psi,
460: \end{split}
461: \end{equation}
462: having used the conventions for Dirac matrices:
463: \begin{equation}
464: \gamma^0 = 
465: \begin{pmatrix} 
466: 1 & 0 \\ 
467: 0 & -1 
468: \end{pmatrix},\quad 
469: \gamma^i = 
470: \begin{pmatrix} 
471: 0 & \sigma^i \\
472: -\sigma^i & 0
473: \end{pmatrix}.
474: \end{equation}
475: We can decouple the $\psi$ and $\chi$ fields by the following transformation:
476: \begin{equation}
477: \Psi_Q\rightarrow\exp\left(\frac{i\boldgamma\cdot\vect{D}}{2m_Q}\right)\Psi_Q,
478: \end{equation}
479: which turns (\ref{QCDLagQ}) into:
480: \begin{equation}
481: \mathcal{L}_Q\rightarrow
482: \begin{pmatrix} \psi^\dag & \chi^\dag \end{pmatrix}
483: \begin{pmatrix} 
484: -m_Q + iD_0 + \frac{\vect{D}^2}{2m_Q} & 0 \\
485: 0 & m_Q + iD_0 - \frac{\vect{D}^2}{2m_Q}
486: \end{pmatrix}
487: \begin{pmatrix} \psi \\ \chi \end{pmatrix},
488: \end{equation}
489: up to terms suppressed by higher powers of $v$. Removing the dependence on large label momenta $\vect{p}$ from the full-theory fields:
490: \begin{equation}
491: \psi(x) = e^{-i(m_Q t-\vect{p}\cdot\vect{x})}\psi_{\vect{p}}(x),\quad \chi(x) = e^{i(m_Q t+\vect{p}\cdot\vect{x})}\psi_{\vect{p}}(x),
492: \end{equation}
493: and keeping the dominant terms in $v$, we obtain the leading-order Lagrangian for quarks and antiquarks in NRQCD:
494: \begin{equation}
495: \mathcal{L}_{\text{NRQCD}} = \psi_{\vect{p}}^\dag(x)\left(iD^0 - \frac{\vect{p}^2}{2m_Q}\right)\psi_{\vect{p}}(x) + \chi_{\vect{p}}^\dag(x)\left(iD^0 + \frac{\vect{p}^2}{2m_Q}\right)\chi_{\vect{p}}(x).
496: \end{equation}
497: The fields $\psi_{\vect{p}},\chi_{\vect{p}}$ describe fluctuations about the label momentum $\vect{p}$ of order $m_Qv^2$:
498: \begin{equation}
499: \vect{P} = \vect{p} + \vect{k},\quad E = k^0,
500: \end{equation}
501: where $\vect{P}$ is the total heavy quark or antiquark three-momentum, and the residual momentum $k = (k^0,\vect{k})\sim Mv^2$.
502: 
503: The construction of operators to describe quarkonium decays in this effective theory is performed in Chap.~\ref{chap:ups}.