hep-ph0208203/sm.tex
1: \chapter{The Basic Concepts}
2: \label{ch:SM}
3: 
4: In this chapter we briefly introduce the basic concepts needed for doing calculations in elementary particle physics. We present the Standard Model in a nutshell and introduce the concepts of renormalization, renormalization group, operator product expansion, and effective theories. We assume that the reader is familiar with quantum field and gauge theories and refer to the pertinent textbooks \cite{qft}.
5: 
6: 
7: % =========================================
8: % =      The Standard Model               =
9: % =========================================
10: 
11: \section{The Standard Model}
12: \label{sec:SM}
13: 
14: As mentioned already in the introduction, the Standard Model (SM) is a comprehensive theory of particle interactions. Its success in giving a complete and correct description of all non-gravitational physics tested so far is unprecedented. In the following we give a short introduction into this beautiful theory. We want to introduce the ``Old Standard Model,'' i.e. the one where neutrinos are massless. The recent evidence for neutrino masses, coming from the observation of neutrino oscillations \cite{SK}, has no direct consequences for our work.
15: 
16: The Standard Model is made up of the Glashow-Salam-Weinberg Model \cite{GSWSM} of electroweak interaction and Quantum Chromodynamics (QCD) \cite{QCD}. It is based on the principle of gauge symmetry. The Lagrangian of a gauge theory is invariant under local ``gauge'' transformations of a symmetry group. Such a symmetry can be used to generate dynamics - the {\em gauge interactions.} The prototype gauge theory is quantum electrodynamics (QED) with its Abelian $U(1)$ local symmetry. It is believed that all fundamental interactions are described by some form of gauge theory.
17: 
18: The gauge group of strong interactions is the non-Abelian group $SU(3)_C$ which has eight generators. These correspond to the gluons that communicate the strong force between objects carrying colour charge - therefore the ``C'' as subscript. Since the gluons themselves are coloured, they can directly interact with each other, which leads to the phenomena of {\em ``asymptotic freedom''} and {\em ``confinement.''} At short distances, the coupling constant $\alpha_s$ becomes small. This allows us to compute colour interactions using perturbative techniques and turns QCD into a quantitative calculational scheme. For long distances on the other hand, the coupling gets large, which causes the quarks to be ``confined'' into colourless hadrons. In the words of Yuri Dokshitzer: ``QCD, the marvellous theory of the strong interactions, has a split personality. It embodies `hard' and `soft' physics, both being hard subjects, the softer ones being the hardest.'' \cite{YD}
19: 
20: Electroweak interaction is based on the gauge group $SU(2)_L \otimes U(1)_Y$ - ``L'' stands for ``left'' and ``Y'' denotes the hypercharge - which is spontaneously broken to $U(1)_{QED}$. This is achieved through the non vanishing vacuum expectation value of a scalar isospin doublet {\em Higgs} field \cite{Higgs}
21: \begin{equation}
22: \label{higgs}
23:   \phi=\left(\begin{array}{c}\phi^+\\ \phi^0\end{array}\right)
24: \end{equation}
25: Three of the four scalar degrees of freedom of the Higgs field give masses to the $W$ and $Z$ bosons. The remaining manifests itself in a massive neutral spin zero boson, the physical Higgs boson. It is the only particle of the Standard Model which lacks direct experimental detection. The current lower limit on its mass is 114.1 GeV at the $95\%$ confidence level \cite{mH}. From electroweak precision data there is much evidence for a light Higgs. But as soon as such a light Higgs is found, this gives birth to the {\em hierarchy problem.} A scalar (Higgs) mass is not protected by gauge or chiral symmetries so we expect $m_H \approx \Lambda \approx 10^{16}\,\mbox{GeV}$ if we do not want to fine-tune the bare Higgs mass against the mass aquired from quantum effects. Why should $m_H$ be much smaller than $\Lambda$?
26: 
27: Fermions are the building blocks of matter. In the SM they appear in three generations which differ only in their masses. The two species of fundamental fermions are leptons and quarks. With regard to the gauge group $SU(2)_L$ the quarks and leptons can be classed in left-handed doublets and right-handed singlets.
28: \begin{displaymath}
29:   \begin{array}{p{2.3cm}ccccc}
30:     \mbox{Quarks:}&
31:     \left(\begin{array}{c} u\\ d\:\!' \end{array} \right)_{\!\!\!L} & \qquad \qquad &
32:     \left(\begin{array}{c} c\\ s\:\!' \end{array} \right)_{\!\!\!L} & \qquad \qquad &
33:     \left(\begin{array}{c} t\\ b\:\!' \end{array} \right)_{\!\!\!L} \\
34:      & u_R & \qquad \qquad & c_R & \qquad \qquad & t_R \\
35:      & d_R & \qquad \qquad & s_R & \qquad \qquad & b_R \\
36:     \\
37:     \mbox{Leptons:}&
38:     \left(\begin{array}{c} \nu_e\\ e^- \end{array} \right)_{\!\!\!L}
39:     & \qquad \qquad &
40:     \left(\begin{array}{c} \nu_\mu\\ \mu^- \end{array}
41:     \right)_{\!\!\!L} & \qquad \qquad &
42:     \left(\begin{array}{c} \nu_\tau\\ \tau^- \end{array}
43:     \right)_{\!\!\!L} \\
44:      & e_R & \qquad \qquad & \mu_R & \qquad \qquad & \tau_R
45:    \end{array}
46: \end{displaymath}
47: Within the Old Standard Model there are no right-handed neutrinos. The quarks carry colour charge and transform as $SU(3)_C$ triplets whereas the colourless leptons are $SU(3)_C$ singlets.
48: 
49: Fermion masses are generated via a Yukawa interaction $\overline \psi(x)\phi(x)\psi(x)$ with the Higgs field (\ref{higgs}). Using global unitary transformations in flavour space, the Yukawa interaction can be diagonalized to obtain the physical mass eigenstates
50: \begin{equation}
51: \label{defCKM}
52:   \left( \begin{array}{c} d\:\!' \\ s\:\!' \\ b\:\!' \end{array} \right) =
53:   \underbrace{\left( \begin{array}{ccc} V_{ud} & V_{us} & V_{ub} \\
54:       V_{cd} & V_{cs} & V_{cb} \\
55:       V_{td} & V_{ts} & V_{tb} \end{array} \right)}_{\displaystyle \equiv V_\mathrm{CKM}} \cdot
56:   \left( \begin{array}{c} d \\ s \\ b \end{array} \right)
57: \end{equation}
58: The non-diagonal elements of the Cabibbo-Kobayashi-Maskawa matrix \cite{CKM} allow for transitions between the quark generations in the charged quark current
59: \begin{equation}
60: \label{cc}
61:   J^+_\mu = \left( \overline{u}, \overline{c}, \overline{t} \right)_L
62:   \gamma_\mu V_\mathrm{CKM} \left( \begin{array}{c}
63:       d\\s\\b\end{array}\right)_{\!\!\!L}
64: \end{equation}
65: Unitarity of $V_\mathrm{CKM}$ guaranties the absence of flavour-changing-neutral-current (FCNC) processes at tree level. This Glashow-Iliopoulos-Maiani (GIM) mechanism \cite{GIM} would forbid FCNC transitions even beyond the tree level if we had exact horizontal flavour symmetry which assures the equality of quark masses of a given charge. Such a symmetry is in nature obviously broken by the different quark masses so that at the one-loop level effective $b\to s$, $b\to d$, or $s\to d$ processes like $B\to X_{s}\gamma$ or $K\to\pi\nu\bar\nu$ can appear.
66: 
67: 
68: A unitary complex $N\times N$ matrix can be described by $N^2$ real parameters. If this matrix mixes $N$ families each with two quarks one can remove $2N-1$ phases through a redefinition of the quark states, which leaves the Lagrangian invariant. Because an orthogonal $N\times N$ matrix has $N(N-1)/2$ real parameters (angles) we are left with $N^2-(2N-1)-N(N-1)/2=(N-1)(N-2)/2$ independent physical phases in the quark mixing matrix. Therefore, it is real if it mixes two generations only. The three-generation CKM matrix, however, has to be described by three angles and one complex phase. The latter one is the only source of CP violation within the Standard Model if we desist from the possibility that $\theta_\mathrm{QCD}\neq 0$. But these CP-violating effects can show up only if really all three generations of the Standard Model are involved in the process. Typically this is the case if one considers loop contributions of weak interaction, like box or penguin diagrams as in Fig.~\ref{fig:boxpen}.
69: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70: \begin{figure}
71:   \begin{center}
72:     \psfig{figure=boxpen.ps}
73:   \end{center}
74:   \caption{Example of a box and penguin diagram in the full theory. CP violating effects are possible because quarks of all three generations are involved in the processes.\label{fig:boxpen}}
75: \end{figure}
76: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77: 
78: In principle there are many different ways of parametrizing the CKM matrix. For practical purposes most useful is the so called Wolfenstein parametrization \cite{Wolfen}
79: \begin{equation}
80: \label{wolfen}
81: V_\mathrm{CKM}=\left(
82:   \begin{array}{ccc} 
83:     1-\frac{\lambda^2}{2} & \lambda & A\lambda^3 (\rho-i\eta) \\
84:     -\lambda & 1-\frac{\lambda^2}{2} & A\lambda^2 \\
85:     A \lambda^3 (1-\rho-i\eta) & -A\lambda^2 & 1
86:   \end{array}\right)
87: \end{equation}
88: which is an expansion to ${\cal O}(\lambda^3)$ in the small parameter $\lambda=|V_{us}|\approx 0.22$. It is possible to improve the Wolfenstein parametrization to include higher orders of $\lambda$ \cite{impwolf}. In essence $\rho$ and $\eta$ are replaced with $\bar\rho=\rho(1-\lambda^2/2)$ and $\bar\eta=\eta(1-\lambda^2/2)$, respectively.
89: 
90: For phenomenological studies of CP-violating effects, the so called standard unitarity triangle (UT) plays a special role. It is a graphical representation of one of the six unitarity relations, namely
91: \begin{equation}
92: \label{ut}
93: V_{ud}V_{ub}^*+V_{cd}V_{cb}^*+V_{td}V_{tb}^*=0
94: \end{equation}
95: in the complex $(\rho,\eta)$ plane. This unitarity relation involves simultaneously the elements $V_{ub}$, $V_{cb}$, and $V_{td}$ which are under extensive discussion at present. The area of this and all other unitarity triangles equals half the absolute value of $J_{CP}=\mathrm{Im}(V_{us}^{}V_{cb}^{}V_{ub}^*V_{cs}^*)$, the Jarlskog measure of CP violation \cite{JCP}. Usually, one chooses a phase convention where $V_{cd}^{}V_{cb}^*$ is real and rescales the above equation with $|V_{cd}^{}V_{cb}^*|=A\lambda^3$. This leads to the triangle in Figure \ref{fig:UT} with a base of unit length and the apex $(\bar\rho,\bar\eta)$.
96: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97: \begin{figure}
98:   \begin{center}
99:     \input{ut.pstex_t}
100:   \end{center}
101:   \caption{Unitarity Triangle\label{fig:UT}}
102: \end{figure}
103: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
104: A phase transformation in (\ref{ut}) only rotates the triangle, but leaves its form unchanged. Therefore, the angles and sides of the unitarity triangle are physical observables and can be measured. Much effort was and is put into the determination of the UT parameters. One tries to measure as many parameters as possible. The consistency of the various measurements tests the consequences of unitarity in the three generation Standard Model. Any discrepancy with the SM expectations would imply the presence of new channels or particles contributing to the decay under consideration. So far, all experimental results are consistent with the Standard Model picture \cite{CKMfitter,UTfit}. The state-of-the-art ``frequentists'' result for the unitarity triangle from the 2002 Winter conferences is displayed in Fig.~\ref{fig:ckmfit} \cite{CKMfitter}.
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106: \begin{figure}
107: %   \epsfxsize=10cm
108:    \centerline{\epsffile{ckmfitter.eps}}
109: \caption{The Range-fit result of the global CKM fit from CKMfitter (status May 2002) \cite{CKMfitter}.\label{fig:ckmfit}}
110: \end{figure}
111: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
112: Actually the good agreement of measurements with the Kobayashi-Maskawa mechanism gives rise to some theoretical puzzles: the KM mechanism for example does explain neither the cosmic baryon asymmetry nor the smallness of $\theta_\mathrm{QCD}$ and basically all extensions of the Standard Model introduce a large number of new CP-violating phases.
113: 
114: 
115: 
116: % =========================================
117: % =      Renormalization                  =
118: % =========================================
119: 
120: \section{Renormalization and Renormalization Group}
121: \label{sec:ren}
122: 
123: Given the Lagrangian of a theory one can deduce the Feynman rules by means of which amplitudes of the processes occuring in this theory can be calculated in perturbation theory. In Feynman diagrams with internal loops, however, one often encounters ultraviolet divergences. This is because the momentum variable of the virtual particle in the loop integration ranges from zero to infinity. The theory of renormalization is a prescription which allows us to consistently isolate and remove all these infinities from the physically measurable quantities. A two-step procedure is necessary.
124: 
125: First, one {\em regulates} the theory. That is, one modifies it in a way that observable quantities are finite and well defined to all orders in perturbation theory. We are then free to manipulate formally these quantities, which are divergent only when the regularization is removed. The most straightforward way to make the integrals finite is to introduce a momentum cutoff. But this violates for example Lorentz invariance or the Ward identities. A regularization method that preserves all symmetries of a gauge theory is {\em dimensional regularization} \cite{HV,dimreg}. The basic idea is to compute the Feynman diagram as an analytic function of the dimensionality of space time $D=4-2\epsilon$. For sufficiently small $D$, any loop-momentum integral will converge. The singularities are extracted as poles for $\epsilon\to 0$.
126: 
127: Potential problems of dimensional regularization concern the treatment of $\gamma_5$ in $D\neq 4$ dimensions. The definition
128: \begin{equation}
129: \label{gamma5}
130: \gamma_5=\frac{i}{4!}\varepsilon_{\kappa\lambda\mu\nu}\gamma^\kappa \gamma^\lambda \gamma^\mu \gamma^\nu
131: \end{equation}
132: with $\epsilon_{\kappa\lambda\mu\nu}$ the completely antisymmetric tensor in four dimensions, cannot straightforwardly be translated to $D\neq 4$ dimensions. In the so called ``naive dimensional regularization'' (NDR) scheme \cite{NDR} the metric tensor is generalized to $D$ dimensions: $g_\mu^\mu=D$, and the $\gamma$ matrices obey the same anticommuting rules as in four dimensions. Even if these rules are algebraically inconsistent \cite{BM}, the NDR scheme gives correct results provided one can avoid the calculation of traces like ${\rm tr}(\gamma_5 \gamma_\kappa \gamma_\lambda \gamma_\mu \gamma_\nu)$ \cite{BW}.
133: 
134: The scheme originally proposed by 't Hooft and Veltman (HV scheme) \cite{HV} allows a consistent formulation of dimensional regularization even when $\gamma_5$ couplings are present \cite{BM}. Besides the $D$- and 4-dimensional metric tensors $g$ and $\tilde g$ one introduces the $-2\varepsilon$-dimensional tensor $\hat g$. One can then split the $D$-dimensional Dirac matrix $\gamma_\mu$ into a 4- and a $-2\varepsilon$-dimensional part $\tilde\gamma_\mu$ and $\hat\gamma_\mu$ which separately obey anticommutation relations with the appropriate metric tensors. A $\gamma_5$ can be introduced which anticommutes with $\tilde\gamma$ but commutes with $\hat\gamma$. The price one has to pay for a consistent dimensional regularization scheme is a substantial increase in the complexity of calculations. 
135: 
136: The second step in our programme to eliminate the infinities from a theory is {\em renormalization}. This is the process of relating the unphysical {\em (bare)} and physical {\em (renormalized)} parameters like couplings $g$ or masses $m$ and rewrite observables as functions of the physical quantities. The renormalization procedure hides all divergences in a redefinition of the fields and parameters in the Lagrangian, i.e.
137: \begin{equation}
138:   \label{ren}
139:   \begin{array}{lcl}
140:    g_0=Z_g g \mu^\varepsilon & \qquad & m_0=Z_m m\\
141:    q_0=Z_q^{1/2}q            & \qquad & A_0^\mu=Z_3^{1/2}A^\mu
142:   \end{array}
143: \end{equation}
144: and thus guaranties that measurable quantities stay finite. The index ``0'' indicates bare quantities. Introducing the parameter $\mu$ with dimension of mass in (\ref{ren}) is necessary to keep the coupling dimensionless. The factors $Z$ are the renormalization constants. The renormalization process is performed recursively in powers of the coupling constant $g$. If at every order of perturbation theory all divergences are reabsorbed in $Z$'s, the theory is called ``renormalizable''. Theories with gauge symmetries, like the Standard Model, are renormalizable. This is true even if the gauge symmetry is spontaneously broken via the Higgs mechanism because gauge invariance of the Lagrangian is conserved \cite{GtH}.
145: 
146: Renormalization can be straightforwardly implemented via the counter-term method. According to (\ref{ren}) the unrenormalized quantities are reexpressed through the renormalized ones in the original Lagrangian. Thus
147: \begin{equation}
148: \label{lcounter}
149: {\cal L}_0={\cal L}+{\cal L}_\mathrm{counter}
150: \end{equation}
151: The counter terms ${\cal L}_\mathrm{counter}$ are proportional to $(Z-1)$ and can be treated as new interaction terms. For these new interactions Feynman rules can be derived and the renormalization constants $Z_i$ are determined such that the contributions from these new interactions cancel the divergences in the Green functions. This fixes the renormalization constants only up to an arbitrary subtraction of finite parts. Different finite parts define different {\em renormalization schemes.} In the Minimal Subtraction (MS) scheme only the divergences and no finite parts are subtracted \cite{MS}. The modified MS scheme ($\overline{\rm MS}$) \cite{BBDM} defines the finite parts such that terms $\ln 4\pi-\gamma_E$, the artifacts of dimensional regularization, vanish. This can be achieved if one calculates with
152: \begin{equation}
153: \label{mumsbar}
154:   \mu_{\overline{\mathrm{MS}}}=\frac{\mu \, e^{\gamma_E/2}}{\sqrt{4\pi}}
155: \end{equation}
156: instead of $\mu$ and performs minimal subtraction afterwards. We will exclusively work with the $\overline{\rm MS}$ scheme in the following.
157: 
158: Every renormalization procedure necessitates to introduce a dimensionful parameter $\mu$ into the theory. Even after renormalization the theoretical predictions depend on this {\em renormalization scale} $\mu$. At this momentum scale the renormalization prescriptions, which the parameters of a renormalized field theory depend on, are applied. One ``defines the theory at the scale $\mu$.'' The bare parameters are $\mu$-independent. To determine the renormalized parameters from experiment, a specific choice of $\mu$ is necessary: $g\equiv g(\mu)$, $m\equiv m(\mu)$, $q\equiv q(\mu)$. Different values of $\mu$ define different parameter sets $g(\mu)$, $m(\mu)$, $q(\mu)$. The set of all tranformations that relates parameter sets with different $\mu$ is called {\em renormalization group} (RG).
159: 
160: The scale dependence of the renormalized parameters can be obtained from the $\mu$-independence of the bare ones. In QCD we get from (\ref{ren}) the {\em renormalization group equations} (RGE) for the {\em running coupling} and the {\em running mass}
161: \begin{eqnarray}
162:   \label{RGEg}
163:   \frac{dg(\mu)}{d\ln \mu} &=& \beta\!\left(g(\mu),\varepsilon\right)\\
164:   \label{RGEm}
165:   \frac{dm(\mu)}{d\ln \mu} &=&-\gamma_m\left(g(\mu)\right) m(\mu)
166: \end{eqnarray}
167: with the $\beta$-function
168: \begin{equation}
169: \label{beta}
170:   \beta\!\left(g(\mu),\varepsilon\right)=-\varepsilon g \underbrace{-\frac{1}{Z_g}\frac{dZ_g}{d \ln\mu}}_{\displaystyle =: \beta(g)}
171: \end{equation}
172: and the anomalous dimension of the mass operator
173: \begin{equation}
174: \label{gamma}
175:   \gamma_m\left(g(\mu)\right)=\frac{1}{Z_m}\frac{dZ_m}{d \ln\mu}
176: \end{equation}
177: Calculating to two-loop accuracy we get
178: \begin{eqnarray}
179:   \beta(g)&=& -\frac{g^3}{16\pi^2}\beta_0 -\frac{g^5}{(16\pi^2)^2}\beta_1\\
180:   \gamma_m(\alpha_s) &=& \frac{\alpha_s}{4\pi}\gamma_m^{(0)} +\left(\frac{\alpha_s}{4\pi}\right)^2\gamma_m^{(1)}
181: \end{eqnarray}
182: where
183: \begin{eqnarray}\label{betadef}
184:   \begin{array}{p{4cm}p{6.5cm}}
185:     $\displaystyle \beta_0=\frac{11N-2f}{3}$ & $\displaystyle \beta_1=\frac{34}{3}N^2 -\frac{10}{3}N f-2C_F f$
186:   \end{array}\\ \label{gammadef}
187:   \begin{array}{p{4cm}p{6.5cm}}
188:     $\displaystyle \gamma_m^{(0)}=6C_F$ & $\displaystyle \gamma_m^{(1)}=C_F\left(3C_F+\frac{97}{3}N-\frac{10}{3}f\right)$
189:   \end{array}\\ \label{alphasCFdef}
190:   \begin{array}{p{4cm}p{6.5cm}}
191:     $\displaystyle \alpha_s(\mu)=\frac{g^2(\mu)}{4\pi}$ & $\displaystyle C_F=\frac{N^2-1}{2N}$
192:   \end{array}
193: \end{eqnarray}
194: with $N$ the number of colours and $f$ the number of active flavours. The solutions for $\alpha_s(\mu)$ and $m(\mu)$ then are \cite{BBDM}
195: \begin{eqnarray}
196: \label{runalpha}
197:   \alpha_s(\mu) &=&\frac{4\pi}{\beta_0 \ln(\mu^2/\Lambda_{\overline{MS}}^2)}\left[1-\frac{\beta_1}{\beta_0^2}\frac{\ln\ln(\mu^2/\Lambda_{\overline{MS}}^2)}{\ln(\mu^2/\Lambda_{\overline{MS}}^2)}\right]\\
198: \label{runm}
199:   m(\mu) &=& m(\mu_0)\left[\frac{\alpha_s(\mu)}{\alpha_s(\mu_0)}\right]^{\frac{\gamma_m^{(0)}}{2\beta_0}}\left[1+\left(\frac{\gamma_m^{(1)}}{2\beta_0}-\frac{\beta_1\gamma_m^{(0)}}{2\beta_0^2}\right)\frac{\alpha_s(\mu)-\alpha_s(\mu_0)}{4\pi}\right]
200: \end{eqnarray}
201: Here $\Lambda_{\overline{MS}}$ is a characteristic scale both for QCD and the used $\overline{MS}$ scheme and depends also on the number of effective flavours present in $\beta_{0}$ and $\beta_1$. An $\alpha_s^{(5)}(M_Z)=0.118\pm0.005$ corresponds to $\Lambda^{(5)}_{\overline{MS}}=225^{+70}_{-57}$ MeV in NLO. It is interesting to note that such a mass scale $\Lambda$ emerges without making reference to any dimensional quantity and would be present also in a theory with completely massless particles. In QCD with three colours, even for six active flavours, both $\beta_0=7$ and $\gamma_m^{(0)}/2\beta_0=4/7$ are positive. This leads to asymptotic freedom as the coupling tends to zero with increasing $\mu$. The pole at $\Lambda_{\overline{MS}}$ signals the breakdown of perturbation theory but gives a plausible argument for confinement. Similarly, $m(\mu)$ decreases with $\mu$ getting larger.
202: 
203: A particularly useful application of the renormalization group is the summation of large logarithms. To see this we reexpress $\alpha_s$ of (\ref{runalpha}) as
204: \begin{equation}
205: \label{runalphamu0}
206:   \alpha_s(\mu)=\frac{\alpha_s(\mu_0)}{v(\mu)}\left[1-\frac{\beta_1}{\beta_0}\frac{\alpha_s(\mu_0)}{4\pi}\frac{\ln v(\mu)}{v(\mu)}\right]
207: \end{equation}
208: with
209: \begin{equation}
210:   v(\mu)=1-\beta_0 \frac{\alpha_s}{4\pi} \ln\frac{\mu_0^2}{\mu^2}
211: \end{equation}
212: If we expand the leading order term of (\ref{runalphamu0}) in $\alpha_s(\mu_0)$ we get
213: \begin{equation}
214: \label{RGIalpha}
215:   \alpha_s(\mu)=\alpha_s(\mu_0)\sum_{m=0}^\infty\left(\beta_0\frac{\alpha_s(\mu_0)}{4\pi}\ln\frac{\mu_0^2}{\mu^2}\right)^m
216: \end{equation}
217: Thus the solution of the RGE automatically sums the logarithms $\ln(\mu_0^2/\mu^2)$ which get large for $\mu\ll\mu_0$. Generally, solving the RGE to order $n$ sums in $\alpha_s(\mu)$ all terms of the form
218: \begin{equation}
219:   \alpha_s(\mu_0)^{m+1}\left(\alpha_s(\mu_0)\ln\frac{\mu_0^2}{\mu^2}\right)^k,\quad 0\le m\le n, \quad k \in \mathbb{N}_{\;\!0}
220: \end{equation}
221: This is particularly useful if, though $\alpha_s(\mu_0)$ is smaller than one, the combination $\alpha_s(\mu_0)\ln(\mu_0^2/\mu^2)$ is close to or even larger than one. Then the large logarithms would spoil the convergence of the perturbation series.
222: 
223: 
224: 
225: % =========================================
226: % =      Operator Product Expansion       =
227: % =========================================
228: 
229: \section{Operator Product Expansion}
230: \label{sec:OPE}
231: 
232: Up to now we treated processes of strong and electroweak interaction separately. But all weak processes involving hadrons receive QCD corrections, which can be substantial especially for non-leptonic and rare decays. The underlying quark level decay of a hadron is governed by the electroweak scale given by $M_{W,Z}={\cal O}(100 \,{\rm GeV})$. On the other hand, the available energy inherent in a $B$ meson decay is of ${\cal O}(m_B)$. In dimensional regularization for example we encounter logarithms of the ratio of either of these scales with the renormalization scale $\mu$. If the scales involved are widely separated it is not possible to make all the logarithms small by a suitable choice of the renormalization scale. As we have seen in the last section these large logarithms can be summed systematically using renormalization group techniques.
233: 
234: But we have yet another energy scale in the problem. A priori we cannot consider the decay of free quarks. Due to confinement quarks appear in colourless bound systems only. The binding of the quarks inside the hadron via strong interaction is characterized by a typical hadronic scale of ${\cal O}(1\,{\rm GeV})$. Here, even without large logarithms the strong coupling $\alpha_s$ is too large for perturbation theory to make sense. Unfortunately, in many cases the non-perturbative methods we have at hand nowadays are not yet developed enough to give accurate results.
235: 
236: Coming back to the typical energy $m_B$ in a $B$ decay. Do we have to know at all what is {\em really} going on at energies of ${\cal O}(100 \,{\rm GeV})$ or the corresponding extremely short distances? In fact we do not. We also do not bother general relativity to calculate the trajectory of an apple falling from a tree or QED and QCD to learn something about the properties of condensed matter. Instead, we employ Newtonian mechanics or the laws of chemistry and solid state physics, respectively. They are nonrelativistic approximations or {\em effective theories} appropriate for the low energy scale under consideration. This is exactly what we want to achieve for the weak interaction of quarks as well. The theoretical tool for this purpose is {\em operator product expansion} (OPE) \cite{OPE} which we shall introduce in the following.
237: 
238: For small separations, the product of two field operators $A(x)$ and $B(y)$ can be expanded in local operators $Q_i$ with potentially singular coefficient functions $c_i$ as
239: \begin{equation}
240: \label{OPE}
241:   A(x)B(y)=\sum_i c_i(x-y)Q_i(x)
242: \end{equation}
243: The {\em Wilson coefficient} $c_1(z)$ accompanying the operator with lowest energy dimension is the most singular one for $z\to 0$ and the degree of divergence of the $c_i$ decreases for increasing operator dimension. Furthermore, for dimensional reasons, contributions of operators with higher dimension are suppressed by inverse powers of the heavy mass (small distance) scale. In principle, we have to consider all operators compatible with the global symmetries of the operator product $AB$. The physical picture is that a product of local operators should appear as one local operator if their distance is small compared to the characteristic length of the system. One can systematically approximate the behaviour of an operator product at short distances with a finite set of local operators. This is exactly what is done for the theory of weak decays.
244: 
245: Here, the mass of the $W$ boson $M_W\approx 80\,{\rm GeV}$ is very large compared to a typical hadronic scale. Therefore, the $W$ propagator $\Delta^{\mu\nu}(x,y)$ is of very short range only. In the amplitude of weak decays it connects two charged currents $J_\mu^-(x)$ and $J_\nu^+(y)$, which hence interact almost locally so that we can perform an OPE. Let us consider the quark level decay $b\to cs\bar u$ for definiteness. The tree-level $W$-exchange amplitude for this decay is given by
246: \begin{eqnarray}
247: \label{Abtocsu}
248:   A(b \to cs\bar u)&=& -\frac{G_F}{\sqrt{2}} V^{}_{cb}
249:   V^*_{us}\frac{M_W^2}{k^2-M_W^2}(\bar s u)_{V-A}(\bar c b)_{V-A} 
250:   \nonumber \\
251:   &=&\frac{G_F}{\sqrt{2}} V^{}_{cb}V^*_{us}\underbrace{(\bar s u)_{V-A}
252:   (\bar c b)_{V-A}}_{\mbox{local operator}} +
253:   {\cal O}\!\left(\frac{k^2}{M_W^2}\right) 
254: \end{eqnarray}
255: Since the momentum transfer $k$ through the $W$ propagater is small as compared to $M_W$, we can safely neglect the terms ${\cal O}(k^2/M_W^2)$. The $W$ propagator then quasi shrinks to a point (see fig. \ref{fig:effvertex})
256: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
257: \begin{figure}
258:   \begin{center}
259:     \psfig{figure=effvertex.ps}
260: %    \input{effvertex.pstex_t}
261:   \end{center}
262:   \caption[The effective vertex]{Replacing a $W$ propagator with an effective four-fermion vertex.\label{fig:effvertex}}
263: \end{figure}
264: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
265: and we obtain an effective four-fermion interaction. This is the modern formulation of the classical Fermi theory of weak interaction with $G_F=1.166\cdot 10^{-5}\,{\rm GeV}^{-2}$ the Fermi constant. The Wilson coefficient in this example is simply one. The notation $(\overline{q}_1 q_2)_{V-A}$ in (\ref{Abtocsu}) is a practical shorthand for a left-handed charged quark current with the chiral vector minus axialvector structure
266: \begin{equation}
267:   (\overline{q}_1 q_2)_{V-A}:=\overline{q}_1\gamma_\mu (1-\gamma_5)q_2
268: \end{equation}
269: 
270: 
271: 
272: % =========================================
273: % =      Effective Theories               =
274: % =========================================
275: 
276: \section{Effective Theories}
277: \label{sec:ET}
278: 
279: The result (\ref{Abtocsu}) can also be derived from an effective Hamiltonian
280: \begin{equation}
281:   {\cal H}_{\rm eff}=\frac{G_F}{\sqrt{2}} V_{cb}V^*_{us}
282:   (\bar s u)_{V-A}(\bar c b)_{V-A}+\mbox{operators of higher dimension}
283: \end{equation}
284: where the operators of higher dimensions correspond to the terms ${\cal O}(k^2/M_W^2)$ in (\ref{Abtocsu}) and can likewise be neglected.  In the effective theory the $W$ boson is removed as an explicit, dynamical degree of freedom. It is ``integrated out'' or ``contracted out'' using the language of the path integral or canonical operator formalism, respectively. One can proceed in a completely analougous way with the heavy quarks. This leads to {\em effective $f$ quark theories} where $f$ denotes the ``active'' quarks, i.e. those that have not been integrated out.
285: 
286: If we include also short distance QCD or electroweak corrections more operators have to be added to the effective Hamiltonian which we generalize to
287: \begin{equation}
288:   \label{Heffgen}
289:   {\cal H}_\mathrm{eff}=\frac{G_F}{\sqrt{2}}\sum_i V_\mathrm{CKM}^i C_i(\mu)Q_i(\mu)
290: \end{equation}
291: Here the factor $V^i_\mathrm{CKM}$ denotes the CKM structure of the particular operator. If we want to calculate the amplitude for the decay of a meson $M=K,\,D,\,B,\,\ldots$ into a final state $F$ we just have to project the Hamilton operator onto the external states
292: \begin{eqnarray}
293:   A(M\to F)&=&\langle F| {\cal H}_\mathrm{eff} |M\rangle
294:   \nonumber \\
295:   &=&\frac{G_F}{\sqrt{2}}\sum_i V_\mathrm{CKM}^i C_i(\mu) \langle
296:   F|Q_i(\mu) |M\rangle\,.
297: \end{eqnarray}
298: The Wilson coefficients $C_i(\mu)$ can be interpreted as the coupling constants for the effective interaction terms $Q_i(\mu)$. They are calculable functions of $\alpha_s$, $M_W$, and the renormalization scale $\mu$. To any order in perturbation theory the Wilson coefficients can be obtained by {\em matching} the full theory onto the effective one. This simply is the requirement that the amplitude in the effective theory should reproduce the corresponding amplitude in the full theory. Hence, we first have to calculate the amplitude in the full theory and then the matrix elements $\langle Q_i\rangle$. In this second step the resulting expressions may, even after quark field renormalization, be still divergent. Consequently we have to perform an {\em operator renormalization}
299: \begin{equation}
300:   Q_i^{(0)}=Z_{ij} Q_j
301: \end{equation}
302: where $Q_i^{(0)}$ denotes the unrenormalized operator. This notation is somewhat sloppy and misleading. What actually is renormalized is not the operator but the operator matrix elements, or, even more exactly, the amputated Green functions $\langle Q_i\rangle$. Then we have to include the renormalization constant $Z_q^{1/2}$ for each of the four external quark fields:
303: \begin{equation}\label{opren}
304:   \langle Q_i\rangle^{(0)} =Z_q^{-2} Z_{ij} \langle Q_j\rangle
305: \end{equation}
306: In general, the renormalization constant $Z_{ij}$ is a matrix so that operators carrying the same quantum numbers can {\em mix under renormalization}. Operators of a given dimension mix only into operators of the same or of lower dimension. Again, the divergent parts of the renormalization constant are determined from the requirement that the amplitude in the effective theory is finite. The finite part in $Z_{ij}$ on the other hand defines a specific renormalization scheme. In a third step we extract the Wilson coefficients by comparing the full and the effective theory amplitude. These are the Wilson coefficients at some fixed scale $\mu_0$. A caveat here is that the external states in the full and the effective theory have to be treated in the same manner. Especially the same regularization and renormalization schemes have to be used on both sides. 
307: 
308: As the Wilson coefficients appear already at the level of the effective Hamiltonian, they are independent of the external states this Hamiltonian is projected onto to obtain the complete amplitude. When determining the Wilson coefficients, any external, even unphysical, state can be used. The coefficient functions represent the short-distance structure of the theory. Because they depend for example on the masses of the particles that were integrated out, they contain all information about the physics at the high energy scale. The long-distance contribution, on the other hand, is parametrized by the process-dependent matrix elements of the local operators. This {\em factorization} of SD and LD dynamics is one of the salient features of OPE. We can calculate the Wilson coefficients in perturbation theory and the hadronic matrix elements by means of some non-perturbative technique like $1/N$ expansion, sum rules, or lattice gauge theory. Especially to use the latter one, a separation of the SD part is essential for today's lattice sizes. The factorization can be visualized with large logarithms $\ln(M_W^2/m_q^2)$ being split into $\ln(M_W^2/\mu^2)+\ln(\mu^2/m_q^2)$. In doing so, the first logarithm will be retrieved in the Wilson coefficients and the second one in the matrix elements. From this point of view the renormalization scale $\mu$ can be interpreted as the {\em factorization scale} at which the full contribution is separated into a low energy and a high energy part.
309: 
310: A typical scale at which to calculate the hadronic matrix elements of local operators is low compared to $M_W$. For $B$ decays we would choose $\mu={\cal O}(m_B)$. Therefore, the logarithm $\ln(M_W^2/\mu^2)$ contained in the Wilson coefficient is large. So why not use the powerful technique of summing large logarithms developped in section \ref{sec:ren}? In order to do so we have to find the renormalization group equations for the Wilson coefficients and solve them. But so far the Wilson coefficients were not renormalized at all. If we remember, however, that in the effective Hamiltonian the operators, which have to be renormalized, are accompanied always by the appropriate Wilson coefficent we can shuffle the renormalization as well to the Wilson coefficients. Let us start with the Hamiltonian of the effective theory with fields and coupling constants as bare quantities, which are renormalized according to
311: \begin{eqnarray}
312:   q^{(0)} &=& Z_q^{1/2}q\\
313:   C_i^{(0)} &=& Z_{ij}^c C_j
314: \end{eqnarray}
315: Then the Hamiltonian (\ref{Heffgen}) is in essence
316: \begin{eqnarray}
317:   {\cal H}_\mathrm{eff} &\propto& C_i^{(0)} Q_i(q^{(0)})\nonumber\\
318:     &\equiv& Z_{ij}^c C_j Z_q^2 Q_i\nonumber\\ \label{Heffcounter}
319:     &\equiv& C_i Q_i +(Z_q^2 Z_{ij}^c -\delta_{ij}) C_j Q_i
320: \end{eqnarray}
321: i.e. it can be written in terms of the renormalized couplings $C_i$ and fields $Q_i$ plus counterterms. The $q^{(0)}$ indicates that the interaction term $Q_i$ is composed of unrenormalized fields. If we calculate the amplitude with the Hamiltonian (\ref{Heffcounter}) including the counterterms, we get the finite renormalized result
322: \begin{equation}
323:   Z_q^2 Z_{ij}^c C_j \langle Q_i\rangle^{(0)} = C_j\langle Q_j\rangle
324: \end{equation}
325: Comparing with (\ref{opren}) we read off the renormalization constant for the Wilson coefficients
326: \begin{equation}
327:   Z_{ij}^c = Z_{ji}^{-1}
328: \end{equation}
329: So we can think of the operator renormalization in terms of the completely equivalent renormalization of the coupling constants $C_i$, as in any field theory. If we again demand the unrenormalized Wilson coefficients not to depend on $\mu$ we obtain the renormalization group equation
330: \begin{equation}
331:   \label{RGEC}
332:   \frac{dC_i(\mu)}{d\ln\mu}=\gamma_{ji}(\mu)C_j(\mu)\,,
333: \end{equation}
334: with the anomalous dimension matrix for the operators
335: \begin{equation}
336:   \label{admop}
337:   \gamma_{ij}(\mu)=Z^{-1}_{ik}\frac{dZ_{kj}}{d\ln\mu}
338: \end{equation}
339: Let us simply state here that the numerical values for the $\gamma_{ij}$ can be determined directly from the divergent parts of the renormalization constants $Z_{ij}$. In (\ref{RGEC}) the transposed of this anomalous dimension matrix appears. It is only the sign and the fact that the anomalous dimension is a matrix instead of a single number that distinguishes the RGE for the Wilson coefficients from that of the running mass in (\ref{RGEm}). Therefore, we could use the solution (\ref{runm}) with the appropriate changes. To leading order this is in fact possible. But if we want to go to next-to-leading-order accuracy we run into problems, because the matrices $\gamma^{(0)}_{ij}$ and $\gamma^{(1)}_{ij}$ in the perturbative expansion
340: \begin{equation}
341: \label{adm}
342:   \gamma_{ij} = \gamma^{(0)}_{ij}\frac{\alpha_s}{4\pi} +\gamma^{(1)}_{ij}\left(\frac{\alpha_s}{4\pi}\right)^2 +{\cal O}(\alpha_s^3)
343: \end{equation}
344: do not commute with each other. Let us instead formally write the solution for the Wilson coefficients with an evolution matrix $U(\mu,\mu_0)$
345: \begin{equation}\label{cevolve}
346:   C_i(\mu)=U_{ij}(\mu,\mu_0) C_j(\mu_0)
347: \end{equation}
348: The leading order evolution matrix can be read off from (\ref{runm})
349: \begin{eqnarray}
350:   U^{(0)}(\mu,\mu_0) &=& \left[\frac{\alpha(\mu)}{\alpha(\mu_0)}\right]^{-\frac{{\gamma^{(0)}}^T}{2\beta_0}}\\ \nonumber
351:   \label{ulo}
352:   &=& V\left(\left[\frac{\alpha(\mu_0)}{\alpha(\mu)}\right]^\frac{\vec\gamma^{(0)}}{2\beta_0}\right)_{\!\!\!\!D} V^{-1}
353: \end{eqnarray}
354: where $V$ is the matrix that diagonalizes $\gamma^{(0)T}$
355: \begin{equation}
356:   \gamma^{(0)}_D=V^{-1} \gamma^{(0)T} V
357: \end{equation}
358: and $\vec\gamma^{(0)}$ is the vector containing the eigenvalues of $\gamma^{(0)}$. For the next-to-leading order solution we make the clever ansatz
359: \begin{equation}
360: \label{unlo}
361:   U(\mu,\mu_0)=\left[1+\frac{\alpha_s(\mu)}{4\pi}J\right] U^{(0)}(\mu,\mu_0)\left[1-\frac{\alpha_s(\mu_0)}{4\pi} J\right]
362: \end{equation}
363: which proves to solve (\ref{RGEC}) if \cite{RGEC}
364: \begin{equation}
365:   J=V H V^{-1}
366: \end{equation}
367: where the elements of $H$ are
368: \begin{equation}
369:   H_{ij}=\delta_{ij} \gamma_i^{(0)} \frac{\beta_1}{2\beta_0^2} -\frac{G_{ij}}{2\beta_0+\gamma_i^{(0)}-\gamma_j^{(0)}}
370: \end{equation}
371: with
372: \begin{equation}
373:   G=V^{-1} \gamma^{(1)T} V
374: \end{equation}
375: As we have mentioned in section \ref{sec:ren}, the procedure of renormalization allows to subtract arbitrary finite parts along with the ultraviolet singularities. Whereas physical quantities must clearly be independent of the renormalization scheme chosen, at NLO unphysical quantities, like the Wilson coefficients and the anomalous dimensions, depend on the choice of the renormalization scheme. To ensure a proper cancellation of this scheme dependence in the product of Wilson coefficients and matrix elements the same scheme has to be used for both. In order to uniquely define a renormalization scheme it is not sufficient to quote only the regularization and renormalization procedure but one also has to choose a specific form for the so-called {\em evanescent operators}. These are operators which exist in $D\neq 4$ dimensions but vanish in $D=4$ \cite{BW,DG,HN}.
376: 
377: So what do we have achieved so far? We have determined the Wilson coefficients at a scale $\mu_0$ via a matching procedure. These are the initial conditions for the evolution from $\mu_0$ down to an appropriate low energy scale $\mu$ via $U(\mu,\mu_0)$ which sums large logarithms. Herefore, we had to determine the anomalous dimensions of the operators and solve the renormalization group equation for the Wilson coefficients. We thus arrive at a RG improved perturbation theory and officially don't speak any more of ``leading'' (LO) and ``next-to-leading order'' (NLO) but rather of ``leading'' (LL) and ``next-to-leading-logarithmic order'' (NLL). Yet, we might carelessly use the terms synonymously. In our task to evaluate weak decay amplitudes involving hadrons in the framwork of a low energy effective theory we then only lack the calculation of the hadronic matrix elements $\langle Q_i(\mu)\rangle$. This, however, is a highly non-trivial problem which this and many other works are devoted to.
378: 
379: Looking at the complicated NLO formulas one might ask why at all going to next-to-leading order accuracy. After all these calculations imply the evaluation of two or even more loop diagrams which are technically very challenging. But they are very important. First of all we can test the validity of the renormalization group improved perturbation theory. Then, of course, we hope that the theoretical uncertainties get reduced. One particular issue is the residual renormalization scale dependence of the result. The scale $\mu$ enters for example in $\alpha_s(\mu)$ or the running quark masses, in particular $m_t(\mu)$, $m_b(\mu)$, and $m_c(\mu)$. In principle, a physical quantity cannot depend on the renormalization scale. But as we have to truncate the perturbative series at some fixed order, this property is broken. The renormalization scale dependence of Wilson coefficients and operator matrix elements cancels only to the order of perturbation theory included in the calculation. Therefore, one can use the remaining scale ambiguity as an estimate for the neglected higher order corrections. Usually one varies $\mu$ between half and twice the typical scale of the problem, i.e. $m_b/2<\mu<2m_b$ for $B$ decays. Going to NLO significantly reduces these scale ambiguities. Furthermore, the renormalization scheme dependence of the Wilson coefficients appears at NLO for the first time. Only if we properly match the long distance matrix elements, obtained for example from lattice calculations, to the short distance contributions, these unphysical scheme dependences will cancel. Another issue is that the QCD scale $\Lambda_{\overline{\mathrm{MS}}}$, which can be extracted from various high energy processes, cannot be used meaningfully in weak decays without going to NLO.
380: 
381: 
382: 
383: % =========================================
384: % =      The Effective Hamiltonian        =
385: % =========================================
386: 
387: \section{The Effective $b\to s\gamma$ Hamiltonian}
388: \label{sec:EH}
389: 
390: In this section we want to discuss the effective Hamiltonian necessary for the calculations to follow. For $b\to s\gamma$ transitions it reads
391: \begin{equation}\label{heff}
392: {\cal H}_\mathrm{eff}=\frac{G_F}{\sqrt{2}}\sum_{p=u,c}\lambda_p^{(s)}
393: \left[ C_1 Q^p_1 + C_2 Q^p_2 +\sum_{i=3,\ldots ,8} C_i Q_i\right]
394: \end{equation}
395: where
396: \begin{equation}\label{lamps}
397: \lambda_p^{(s)}=V^*_{ps}V_{pb}
398: \end{equation}
399: The operators originate from the diagrams in Fig.~\ref{fig:opdiag}
400: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
401: \begin{figure}
402:   \begin{center}
403:     \psfig{figure=opdiag.ps}
404:   \end{center}
405:   \caption{The diagrams where the operator basis for $b\to s\gamma$ originates from: Current-current diagram (a) with QCD corrections (b), (c), (d); gluon penguin diagram (e), magnetic photon penguin diagrams (f), (g), and magnetic gluon penguin diagram (h). The cross in (f), (g), and (h) denotes that the mass of the external $b$ quark has to be kept.\label{fig:opdiag}}
406: \end{figure}
407: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
408: and are given by
409: \begin{eqnarray}
410:   \label{q1def}
411:     Q^p_1 &=& (\bar sp)_{V-A}(\bar pb)_{V-A} \\
412:   \label{q2def}
413:     Q^p_2 &=& (\bar s_i p_j)_{V-A}(\bar p_j b_i)_{V-A} \\
414:   \label{q3def}
415:     Q_3 &=& (\bar sb)_{V-A} \sum_q (\bar qq)_{V-A} \\
416:   \label{q4def}
417:     Q_4 &=& (\bar s_i b_j)_{V-A} \sum_q (\bar q_j q_i)_{V-A} \\
418:   \label{q5def}
419:     Q_5 &=& (\bar sb)_{V-A} \sum_q (\bar qq)_{V+A} \\
420:   \label{q6def}
421:     Q_6 &=& (\bar s_i b_j)_{V-A} \sum_q (\bar q_j q_i)_{V+A} \\
422:   \label{q7def}
423:     Q_7 &=& \frac{e}{8\pi^2}m_b\,\bar s_i\sigma^{\mu\nu}(1+\gamma_5)b_i\, F_{\mu\nu}\\
424:   \label{q8def}
425:     Q_8 &=& \frac{g_s}{8\pi^2}m_b\,\bar s_i\sigma^{\mu\nu}(1+\gamma_5)T^a_{ij} b_j\, G^a_{\mu\nu}
426: \end{eqnarray}
427: with $e$ and $g_s$ the coupling constants of electromagnetic and strong interaction and $F_{\mu\nu}$ and $G_{\mu\nu}$ the photonic and gluonic field strength tensors, respectively. In $Q_7$ and $Q_8$ we neglected the very small $m_s(1-\gamma_5)$ contribution. The $i,j$ are colour indices. If no colour index is given the two operators are assumed to be in a colour singlet state. The operator basis (\ref{q1def}--\ref{q8def}) consists of all possible gauge invariant operators with energy dimension six with the following properties: they have the correct quantum numbers to contribute to $b\to s\gamma$, they are compatible with the symmetries of electroweak interaction, and they cannot be transformed into each other by applying equations of motion. As a consequence, our operator basis is only the correct one if all external states are taken on-shell \cite{HP}. Additional operators have to be considered if an off-shell calculation is performed. There is yet another operator basis used in the literature. It was introduced by Chetyrkin, Misiak, and M\"unz (CMM) \cite{CMM,CMM2} because then no Dirac traces containing $\gamma_5$ arise in effective theory calculations, wich allows to use fully anticommuting $\gamma_5$ in dimensional regularization. It reads
428: \begin{eqnarray}
429:   \label{p1def}
430:     P^p_1 &=& (\bar s_L \gamma_\mu T^a p_L)(\bar p_L \gamma^\mu T^a b_L) \\
431:   \label{p2def}
432:     P^p_2 &=& (\bar s_L \gamma_\mu p_L)(\bar p_L \gamma^\mu b_L) \\
433:   \label{p3def}
434:     P_3 &=& (\bar s_L \gamma_\mu b_L) \sum_q (\bar q\gamma^\mu q) \\
435:   \label{p4def}
436:     P_4 &=& (\bar s_L \gamma_\mu T^a b_L) \sum_q (\bar q\gamma^\mu T^a q) \\
437:   \label{p5def}
438:     P_5 &=& (\bar s_L \gamma_{\mu_1}\gamma_{\mu_2}\gamma_{\mu_3} b_L) \sum_q (\bar q \gamma^{\mu_1}\gamma^{\mu_2}\gamma^{\mu_3} q) \\
439:   \label{p6def}
440:     P_6 &=& (\bar s_L \gamma_{\mu_1}\gamma_{\mu_2}\gamma_{\mu_3} T^a b_L) \sum_q (\bar q \gamma^{\mu_1}\gamma^{\mu_2}\gamma^{\mu_3} T^a q) \\
441:   \label{p7def}
442:     P_7 &=& \frac{e}{16\pi^2}m_b(\bar s_L\sigma^{\mu\nu}b_R)F_{\mu\nu}\\
443:   \label{p8def}
444:     P_8 &=& \frac{g_s}{16\pi^2}m_b(\bar s_L\sigma^{\mu\nu}T^a b_R)G^a_{\mu\nu}
445: \end{eqnarray}
446: where $T^a$ stand for $SU(3)_\mathrm{color}$ generators and $L=(1-\gamma_5)/2$ and $R=(1+\gamma_5)/2$ for the left and right-handed projection operators. We denote the corresponding Wilson coefficients with $Z_i$. The operator basis (\ref{p1def}--\ref{p8def}) is in principle the more natural one as the operators appear exactly in this form when calculating the diagrams of Fig.~\ref{fig:opdiag}. In our normal operator basis (\ref{q1def}--\ref{q8def}), the following four-dimensional identity was used to ``simplify'' $P_5$ and $P_6$
447: \begin{equation}
448:   \gamma_\mu \gamma_\nu \gamma_\rho = g_{\mu\nu} \gamma_\rho + g_{\nu\rho} \gamma_\mu -g_{\mu\rho} \gamma_\nu +i\epsilon_{\sigma\mu\nu\rho} \gamma^\sigma \gamma_5 \qquad\mbox{in }D=4
449: \end{equation}
450: This step, however, requires to introduce several more evanescent operators and leads to problematic traces with $\gamma_5$ in two-loop calculations. The choice of the operator's colour structure is more natural in the CMM basis, too, as it is the one emerging from the diagrams in Fig.~\ref{fig:opdiag}.
451: 
452: In the common nomenclature $Q_1$ and $Q_2$ are called {\em current-current operators}, $Q_{3\ldots 6}$ {\em QCD penguin operators} and $Q_7$ and $Q_8$ {\em electromagnetic} and {\em chromomagnetic penguin operator}, respectively. For historical reasons the numbering of $Q^p_{1,2}$ is sometimes reversed in the literature \cite{BBL}. The sign conventions for the electromagnetic and strong couplings correspond to the covariant derivative $D_\mu=\partial_\mu +ie Q_f A_\mu + i g T^a A^a_\mu$. The coefficients $C_{7}$ and $C_8$ then are negative in the Standard Model, which is the choice generally adopted in the literature. The effective Hamiltonian for $b\to d\gamma$ is obtained from equations (\ref{heff}--\ref{p8def}) by the replacement $s\to d$.
453: 
454: Let us summarize the status quo in determining the effective Hamiltonian for $b\to s\gamma$. At leading order only $Q_7$ contributes to $b\to s\gamma$. The corresponding Wilson coefficient was first calculated by Inami and Lim in 1980 \cite{IL}. For the leading logarithmic RG improved calculation the $8\times 8$ anomalous dimension matrix was obtained bit by bit. Mixing of the four-quark operators $Q_{1\ldots 6}$ among each other was considered already earlier for the analysis of nonleptonic decays \cite{6x6ADM}. The $2\times 2$ submatrix for the magnetic penguins was obtained by Grinstein, Springer, and Wise \cite{GSWbsg}. Because the mixing of $Q_{1\ldots 6}$ into $Q_7$ and $Q_8$ vanishes at the one-loop level, one has to perform two-loop calculations if the LL result for $C_7(\mu_b)$ is wanted. This technical complication delayed the first completely correct result for the $8\times 8$ anomalous dimension matrix until 1993 \cite{CFMRS} followed by independent confirmations \cite{confLL}.
455: 
456: Going to next-to-leading order accuracy was highly desirable, because the leading logarithmic expression for the branching ratio $B(B\to X_s\gamma)$ suffers from sizable renormalization scale uncertainties at the $\pm 25\%$ level. The step from leading order to leading logarithmic order already increased the branching ratio $B(B\to X_s\gamma)$ by almost a factor of three \citer{BBM,GOSN}. The high-flying NLL enterprise was a joint effort of many groups. QCD corrections affect both Wilson coefficients and operator matrix elements. The ${\cal O}(\alpha_s)$ matching for the current-current operators was calculated already a long time ago \cite{ACMP}, the one for the QCD penguins more than ten years later \cite{BJLW}. Two-loop matching is necessary for $Q_7$ and $Q_8$, which was achieved first by Adel and Yao \cite{AY} and subsequently checked thoroughly \cite{checkAY,CDGG}. The ${\cal O}(\alpha_s^2)$ contributions to the $6\times 6$ anomalous dimension matrix were calculated in \cite{BW,ACMP,BJLW,CFMR}, the submatrix for $Q_7$ and $Q_8$ in \cite{MM}. Chetyrkin, Misiak, and M\"unz finally succeeded in determining the three-loop mixing of four-quark operators into magnetic penguin operators \cite{CMM}. The explicit formulas for the Wilson coefficients needed subsequently can be found in Appendix~\ref{app:WC}, and in Appendix~\ref{app:opbase} we comment on the transformation properties of the two operator bases used in this work.
457: 
458: For the {\em inclusive} decay $B\to X_s\gamma$, i.e. the radiative decay of a $B$ meson into the sum of final states with strangeness $S=-1$, the operator matrix elements can be computed perturbatively employing the {\em heavy-quark expansion} (HQE) \cite{HQE}. This method consists of an OPE in inverse powers of the large dynamical scale of energy release $\sim m_b$ followed by a nonrelativistic expansion for the $b$ field. Using the optical theorem, the inclusive $B$-decay rate can be written in terms of the absorptive part of the forward scattering amplitude $\langle B|{\cal T}|B\rangle$. The transition operator ${\cal T}$ is the absorptive part of the time-ordered product of ${\cal H}_\mathrm{eff}(x){\cal H}_\mathrm{eff}(0)$
459: \begin{equation}\label{Tdef}
460:   {\cal T} = {\rm Im}\!\left[ i\!\int\! d^4x \,T{\cal H}_\mathrm{eff}(x){\cal H}_\mathrm{eff}(0)\right]
461: \end{equation}
462: If we insert a complete set of states inside the time-ordered product we see that this was just a fancy way of writing the standard expression for the decay rate
463: \begin{equation}
464:   \Gamma(B\to X)=\frac{1}{2m_B}\sum_X (2\pi)^4 \delta^4(p_B-p_X)\left|\langle X|{\cal H}_\mathrm{eff}|B\rangle\right|^2
465: \end{equation}
466: However, the formulation in terms of the $T$ product allows for a direct evaluation using Feynman diagrams. Because of the large mass of the $b$ quark, we can construct an OPE in which ${\cal T}$ is represented as a series of local operators containing the heavy-quark fields. The operator of lowest dimension is $\bar b b$. Its matrix element is simplified by a nonrelativistic expansion in powers of $1/m_b$ starting with unity. In the limit of an infinitely heavy $b$ quark the $B$ meson decay rate is therefore given by the $b$ quark decay rate. This {\em quark-hadron duality} is of great use for many inclusive calculations and justifies for example the spectator model. Corrections to this relation appear only at ${\cal O}(1/m_b^2)$ because the potential dimension 4 operator $\bar b D\hspace{-0.65em}/\hspace{0.15em} b$ can be reduced to $m_b \bar b b$. 
467: However, the OPE only converges for sufficiently inclusive observables. But for $B\to X_s\gamma$ decays experimental cuts are necessary to reduce the background from charm production. This restricts the available phase space considerably leading to complications of the OPE based analysis.
468: 
469: The virtual corrections to the matrix elements $\langle s\gamma|Q_{1,7,8}|b\rangle$ were first calculated by Greub, Hurth, and Wyler \cite{GHW} and checked by Buras, Czarnecki, Misiak, and Urban using another method \cite{BCMU}. The latter authors presented recently also the last missing item in the NLO analysis of $B\to X_s\gamma$, namely the two-loop matrix elements of the QCD-penguin operators \cite{BCMU2}. Yet, their contribution is numerically very small because the corresponding Wilson coefficients almost vanish. The Bremsstrahlung corrections, i.e. the process $b\to s\gamma g$, influence especially the photon energy spectrum. At leading order this spectrum is a $\delta$ function, smeared out by the Fermi motion of the $b$ quark inside the $B$ meson, whereas it is broadened substantially at NLO \cite{Brems}.
470: 
471: Even higher electroweak corrections have been calculated for $B\to X_s\gamma$ \cite{EW}. Finally, there are non-perturbative contributions which can be singled out in the framework of HQE. The $1/m_b^2$ corrections mainly account for the fact that in reality a $B$ meson and not a $b$ quark is decaying. Additionally, one has to consider long distance contributions originating in the photon coupling to a virtual $\bar c c$ loop. This {\em Voloshin effect} is proportional to $1/m_c^2$ and enhances the decay rate by 3\% \citer{Voloshin,BIR}.
472: 
473: All this effort was taken to reduce the theoretical error and account for the ever increasing experimental precision. The current experimental world average for the inclusive branching fraction is
474: \begin{equation}
475: \label{bsgamexp}
476:   {\cal B}(B\to X_s\gamma)^\mathrm{exp}=(3.23\pm 0.42)\cdot 10^{-4}
477: \end{equation}
478: combining the results of \citer{CHEN,ABE}. For the theoretical prediction \cite{BCMU2}
479: \begin{equation}
480: \label{bsgamth}
481:   {\cal B}(B\to X_s\gamma)_{E_\gamma >1.6\,\mathrm{GeV}}^\mathrm{th}=(3.57\pm0.30)\cdot 10^{-4}
482: \end{equation}
483: there is an ongoing discussion about quark mass effects. The matrix elements of $Q_1$ depend at two-loop level on the mass ratio $m_c^2/ m_b^2$. Gambino and Misiak \cite{GM} argue that the running charm quark mass should be used instead of the pole mass, because the charm quarks in the loop are dominantly off-shell. Strictly speaking, this is a NNLO issue and could be used to estimate the sensitivity to NNLO corrections. Numerically, however, it increased the branching ratio by 11\%. A more conservative error estimate would rather add this shift to the theoretical error.
484: 
485: As $b\to s\gamma$ decays appear at one-loop level for the first time, they play an important role in indirect searches for physics beyond the Standard Model. Effects from new particles in the loops could easily be of the same order of magnitude than the Standard Model contributions. The excellent agreement of experimental measurement (\ref{bsgamexp}) and theoretical prediction (\ref{bsgamth}) therefore places severe bounds on the parameter space of New Physics scenarios, like multi-Higgs models \cite{CDGG,2HDM}, Technicolor \cite{techcol}, or the MSSM \cite{MSSM}.
486: 
487: The treatment of the matrix elements for {\em exclusive} $b\to s\gamma$ decays as for example $B\to K^*\gamma$ is in general more complicated. In this case, bound-state effects are essential and need to be described by non-perturbative hadronic quantities like form factors. Exactly these exclusive radiative decays of $B$ mesons are the subject of this work and will be dealt with in detail in part II and III.
488: