hep-ph0310094/lsz.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Uses LaTeX with ReVTeX 3.0.
3: %%%%%\documentstyle[12pt,preprint,aps,floats]{revtex}
4: %\documentstyle[prl,aps,twocolumn,floats,epsf]{revtex}
5: \documentstyle[prd,aps,floats,epsf,fleqn]{revtex}
6: 
7: %\topmargin=1cm
8: 
9: %%%%%%%%%%%%%
10: %\tighten
11: %%%%%%%%%%%%%
12: \newcommand{\postbb}[3]
13: {\setlength{\epsfxsize}{#3\hsize}
14:  \centerline{\epsfbox[#1]{#2}}}
15: 
16: %\newcommand{\rmp}[2]{{\em Rev. Mod. Phys.}          {\bf #1},   #2}
17: \newcommand{\plb}[2]{{\em Phys. Lett.}              {\bf #1B}, #2 }
18: \newcommand{\npb}[2]{{\em Nucl. Phys.}              {\bf B#1}, #2 }
19: \newcommand{\npp}[2]{{\em Nucl. Phys. Proc. Suppl.} {\bf #1C}, #2 }
20: \newcommand{\pr }[2]{{\em Phys. Rep.}               {\bf  #1}, #2 }
21: \newcommand{\prt}[2]{{\em Phys. Rev.}               {\bf D#1}, #2 }
22: \newcommand{\pru}[2]{{\em Phys. Rev. Lett.}         {\bf  #1}, #2 }
23: \newcommand{\zpc}[2]{{\em Z. Phys.}                 {\bf C#1}, #2 }
24: \newcommand{\jp }[2]{{\em J. Phys.}                 {\bf C#1}, #2 }
25: \newcommand{\sci}[2]{{\em Science}                  {\bf  #1}, #2 }
26: \newcommand{\app}[2]{{\em Acta Phys. Polon.}        {\bf B#1}, #2 }
27: \newcommand{\mpl}[2]{{\em Mod. Phys. Lett.}         {\bf A#1}, #2 }
28: \newcommand{\cpc}[2]{{\em Comput. Phys. Commun}     {\bf  #1}, #2 }
29: \newcommand{\jpl}[2]{{\em JETP Lett.}               {\bf  #1}, #2 }
30: \newcommand{\epc}[2]{{\em Eur. Phys. J.}            {\bf C#1}, #2 }
31: \newcommand{\con}[2]{                               {\bf  #1}, #2 }
32: %
33: \newcommand{\etal}{{\em et al.}}
34: \newcommand{\ibid}{{\em ibid.}}
35: \newcommand{\col}{Collaboration}
36: \newcommand{\as}{\hat\alpha_s}
37: \newcommand{\aspi}{{\hat\alpha_s\over \pi}}
38: %
39: %\newcommand{\arns}[2]{{\em Ann. Rev. Nucl. Sci.} {\bf #1}, #2 (19#3)}
40: %\newcommand{\ns}[2]{{\em Nucl. Sci.} {\bf #1}, #2 (19#3)}
41: %
42: \newcommand{\ovl}{\overline}
43: \newcommand{\be}{\begin{equation}}
44: \newcommand{\ee}{\end{equation}}
45: \newcommand{\ba}{\begin{eqnarray}}
46: \newcommand{\ea}{\end{eqnarray}}
47: \newcommand{\acht}{\\[8pt]}
48: \def\fiv{${\bf 5}+{\bf\ol5}$}
49: \def\ten{${\bf 10}+{\bf\ol{10}}$}
50: \def\dr{\mbox{{\footnotesize{$\overline{\rm DR}$}}} }
51: \def\ms{\mbox{{\footnotesize{$\overline{\rm MS}$}}} }
52: \def\lg{{\rm lg}}
53: \newcommand{\lsim}{\buildrel < \over {_\sim}}
54: \newcommand{\gsim}{\buildrel > \over {_\sim}}
55: \newcommand{\lnc}{\ln {\mu_0^2\over \hat{m}_c^2}}
56: \def\mmut{\frac{m_{\mu}^2} {\mu_0^2}}
57: \def\mmuq{\frac{m_{\mu}^4} {\mu_0^4}}
58: \def\mmuqt{\frac{m_\mu^2}{\hat{m}_q^2}}
59: \def\mmuqq{\frac{m_\mu^4}{\hat{m}_q^4}}
60: \def\mmuqh{\frac{m_\mu^6}{\hat{m}_q^6}}
61: \def\mmuct{\frac{m_\mu^2}{\hat{m}_c^2}}
62: \def\mmucq{\frac{m_\mu^4}{\hat{m}_c^4}}
63: \def\mmubt{\frac{m_\mu^2}{\hat{m}_b^2}}
64: \def\mmubq{\frac{m_\mu^4}{\hat{m}_b^4}}
65: \def\mst{\frac{\hat{m}_s^2}{\mu_0^2}}
66: \def\msq{\frac{\hat{m}_s^4}{\mu_0^4}}
67: \def\al2{\frac{\alpha^2}{\pi^2}}
68: \def\order#1{{\cal O}\left(#1\right)}
69: \def\eq#1{(\ref{#1})}
70: 
71: 
72: \begin{document}
73: 
74: \preprint{
75: \noindent
76: \hfill
77: \begin{minipage}[t]{3in}
78: \begin{flushright}
79: %hep-ph \\
80: \vspace*{2cm}
81: \end{flushright}
82: \end{minipage}
83: }
84: 
85: \draft
86: 
87: \title{$S$-Matrix Elements in General Renormalization Schemes}
88: \author{Mingxing Luo}
89: \address{
90: Zhejiang Institute of Modern Physics, Department of Physics,
91: Zhejiang University, Hangzhou, Zhejiang 310027, P R China\\
92: Max Plank Institut fur Physik, Werner Heisenberg Institut, Theorie,
93: Fohringer Ring 6, D-80805 Munchen
94: }
95: 
96: \date{\today}
97: 
98: \maketitle
99: 
100: \begin{abstract}
101: Starting from the Lehmann-Symanzik-Zimmermann reduction theorem,
102: we provide a general procedure to extract $S$-matrix elements from Green functions
103: in arbitrary renormalization schemes.
104: \end{abstract}
105: 
106: \pacs{PACS numbers: 11.55.-m, 11.10.Gh, 11.10.-z}
107: 
108: In the framework of quantum field theories, renormalization is a necessity.
109: The bare parameters in the Lagrangian are infinite 
110: and cannot be used conveniently in physical predictions.
111: A re-parametrization of the theory in terms of finite variables is required.
112: To obtain finite Green functions, quantum fields themselves should be renormalized also.
113: Different re-parameterizations constitute different renormalization schemes.
114: For some physical processes, one scheme could be more convenient than others
115: though all schemes are equivalent.
116: It is hard to argue that there is a universally best scheme for all purposes.
117: To discuss different physical processes consistently,
118: such as in a global analysis of high precision electroweak experiments \cite{LLM},
119: all calculations need to be performed in one scheme.
120: 
121: One of the most frequently used schemes in the electroweak theory
122: is the so-called on-shell scheme \cite{sirlin},
123: in which physical masses of particles are used to parametrize the theory, 
124: coupling constants are defined in terms of certain scattering cross-sections at given energy scales,
125: and quantum fields are renormalized to give the two-point functions a residue of unity on the mass-poles. 
126: One big advantage of this scheme is that
127: $S$-matrix elements can be trivially obtained from the corresponding Green functions.
128: However, the Green functions are themselves complicated by 
129: the implementation of renormalization conditions.
130: More so, in theories such as softly-broken supersymmetric ones, 
131: the on-shell scheme cannot be realized for all
132: fields, due to over-constraints from symmetries.
133: 
134: It is thus expedient and sometimes necessary to introduce more general renormalization schemes.
135: In a general renormalization scheme,
136: the theory is parametrized by intermediate quantities that are not necessary physical observables
137: and renormalized fields are not required to give any particular value of residues on the mass-poles.
138: For example, in the modified minimal subtraction ($\overline{MS}$) scheme, 
139: one introduces the so-called $\overline{MS}$ parameters and renormalized fields,
140: which are obtained by subtracting the infinities and related $\log(4\pi)-\gamma$ terms
141: from the corresponding bare quantities \cite{msbar}.
142: In these schemes, Green functions assume simpler forms. 
143: However, care should be taken to obtain $S$-matrix elements.
144: In this note, 
145: we outline a general procedure to extract $S$-matrix elements from Green functions
146: in arbitrary renormalization schemes,
147: based upon the Lehmann-Symanzik-Zimmermann reduction theorem \cite{lsz}.
148: 
149: In some effective theories,
150: it could be convenient to keep non-canonical kinetic terms in the Lagrangian,
151: as the conversion to canonical forms may complicate other parts of the Lagrangian greatly.
152: Our procedure can be trivially extended to accommodate these cases.
153: On the other hand,
154: some calculations start by defining auxiliary quantities such as mixing angles between different fields.
155: These quantities are only well-defined at tree-level and not gauge invariant in general.
156: In our procedure, these quantities do not show up explicitly.
157: So they can be avoided in principle,
158: though it might be convenient for them to be introduced for phenomenological purposes.
159: 
160: To define $S$-matrix elements properly\cite{BD,Weinberg}, we separate the full Hamiltonian $H$ into two parts,
161: a free Hamiltonian $H_0$ and an interaction $H_{int}$,
162: $ % \be
163: H = H_0 + H_{int},
164: $ %\ee
165: in such a way that $H$ and $H_0$ have the same eigenvalue spectrum.
166: For each eigenstate $\left|\Phi_\alpha^{(0)}\right\rangle$ of $H_0$ with eigenvalue $E_\alpha$,
167: % \be
168: % H_0 \left|\Phi_\alpha^{(0)}\right\rangle =  E_\alpha \left|\Phi_\alpha^{(0)}\right\rangle,
169: % \ee
170: one defines corresponding ``in" and ``out" states as eigenstates of $H$
171: \ba
172: H \left|\Phi_\alpha^{\pm}\right\rangle =  E_\alpha \left|\Phi_\alpha^{\pm}\right\rangle,
173: \ea
174: which satisfy the asymptotic condition
175: \ba
176: exp(-iH t) \int d\alpha g(\alpha) \left|\Phi_\alpha^{\pm}\right\rangle 
177: \rightarrow 
178: exp(-iH_0 t) \int d\alpha g(\alpha) \left|\Phi_\alpha^{(0)}\right\rangle, \nonumber
179: \ea
180: for $t \rightarrow - \infty$ and $t \rightarrow + \infty$, respectively.
181: Here $g(\alpha)$ is an arbitrary function but smoothly varying and non-zero over some finite range
182: $\Delta E$ of energy.
183: An $S$-matrix element is defined to be the transition probability amplitude from
184: an in-state $|\Phi_\alpha^{+}\rangle$ to an out-state $|\Phi_\beta^{-}\rangle$
185: \be
186: S_{\beta\alpha} = \left\langle \Phi_\beta^{-} \right.\left| \Phi_\alpha^{+} \right\rangle.
187: \ee
188: 
189: Following \cite{Weinberg}, we define an arbitrary Green function in momentum-space
190: \be
191: G(q_1 q_2 \cdots) =% \int d^4x_1 d^4x_2 \cdots e^{-iq_1 \cdot x_1}  e^{-iq_2 \cdot x_2} \cdots
192: \int_{FT}
193: \left\langle\Phi_0| T\left\{ A_1(x_1) A_2(x_2) \cdots A_n(x_n) \right\} |\Phi_0\right\rangle ,
194: \ee
195: where the $A$'s are Heisenberg-picture operators of arbitrary Lorentz type, $\Phi_0$ is the true vacuum, and
196: $\int_{FT}$ denotes integrations for Fourier transformations
197: \ba
198: \int_{FT} = \int d^4x_1  \cdots  d^4x_n e^{-iq_1 \cdot x_1}  \cdots  e^{-iq_n \cdot x_n}. \nonumber
199: \ea
200: If the $A$'s are ordinary fields appearing in the Lagrangian, then $G$ is a sum of terms calculated using
201: the ordinary Feynman rules, for all graphs with external lines corresponding to the fields $A$'s,
202: carrying off-shell four-momenta $q$'s into the graph.
203: We assume that the theory can be properly regularized and renormalized, so $G$ is well-defined.
204: It can be shown that $G$ has a pole at $\bar{s}=m^2-i \epsilon$, where $m$ is the mass of any one-particle state
205: $\Psi_{\vec{q}\sigma}$ that has non-vanishing matrix elements with states
206: $A_1^\dagger \Phi_0$ and $A_2 A_3 \cdots\Phi_0$, $\epsilon$ is a positive infinitesimal,
207: and the residue is given by\footnote
208: {Our normalization condition for one-particle states is
209: $
210: \sum_\sigma \int {d^3p \over (2\pi)^3} {1 \over \sqrt{2 p^0}} 
211: \left. |\Phi_{\vec{p},\sigma}\right\rangle \left\langle\Phi_{\vec{p},\sigma}| \right. = 1,
212: $
213: ($p^0=\sqrt{\vec{p}^2+m^2}$),
214: instead of 
215: $
216: \sum_\sigma \int d^3p \left. |\Phi_{\vec{p},\sigma}\right\rangle \left\langle\Phi_{\vec{p},\sigma}| \right. = 1,
217: $
218: as that in \cite{Weinberg}.}
219: \be
220: G \rightarrow {i \over q^2 - \bar{s}} \sum_\sigma
221: \left\langle\Phi_0 | A_1(0) |\Phi_{\vec{q}\sigma}\right\rangle  
222: %\int d^4x_2 \cdots e^{-iq_2 \cdot x_2} \cdots 
223: \int_{FT}
224: \left\langle\Phi_{\vec{q}\sigma}| T\left\{ A_2(x_2) \cdots \right\} |\Phi_0\right\rangle,
225: \label{pole-eq}
226: \ee
227: where the sum is over all spin states of the particle of mass $m$.
228: Depending upon the physics problem, $\Psi_{\vec{q}\sigma}$ can correspond to
229: either an in-state or an out-state.
230: If the particle is unstable and 
231: can decay into lighter particles of total decay width $\Gamma$,
232: the pole is displaced from the real axis by a finite amount, $\bar{s}=m^2- i m \Gamma$.
233: 
234: If $A_1$ has the Lorentz transformation properties of 
235: free field $\Psi_l$ belonging to an irreducible representation of the homogeneous Lorentz
236: group, as labeled by the subscript $l$, we can use Lorentz invariance to write
237: \be
238: \left\langle\Phi_0| A_1(0) |\Phi_{\vec{q}\sigma}\right\rangle  = N u_l(q,\sigma)
239: \ee
240: where $u_l(q,\sigma)$ is the coefficient function appearing in the free field $\psi_l$ 
241: and $N$ is a constant.
242: Define a truncated matrix element $M_l$ by
243: \be
244: %\int d^4x_2 \cdots e^{-iq_2 \cdot x_2} \cdots 
245: \int_{FT} 
246: \left\langle\Phi_{\vec{q}\sigma}| T\left\{ A_2(x_2) \cdots \right\} |\Phi_0\right\rangle =
247: \sum_\sigma u^*_l(q,\sigma) M_l(q_2\cdots).
248: \ee
249: Then as $q^2 \rightarrow \bar{s}$
250: \be
251: G \rightarrow N {i \over q^2 - \bar{s} } \sum_{\sigma, m}
252: u_l(q,\sigma)  u^*_m(q,\sigma) M_m(q_2\cdots). \label{eqM}
253: \ee
254: The quantity multiplying $M_m$ in Eq. (\ref{eqM}) is the momentum space matrix propagator $\Delta_{l,m}(q)$
255: for the free field with the Lorentz transformation properties of $A_1$, so $M_{l}$ is the sum of all
256: graphs with external lines carrying momenta $q_1, q_2, \cdots$, corresponding to the operators
257: $A_1, A_2, \cdots$, but with the final propagator for the $A_1$ line stripped away.
258: %
259: Thus, to calculate the matrix for emission of a
260: particle from the sum of Feynman diagrams, one strips away the particle propagator and contracts
261: with the usual external line factor $u^*$.
262: This provides an alternative proof of the Lehmann-Symanzik-Zimmermann reduction theorem\cite{Weinberg},
263: which is applicable to cases of arbitrary spin.
264: 
265: If a theory has $N$ fields $\Psi^i_l$, which have the same Lorentz transformation properties
266: and other conserved quantum numbers as $\Psi_l$,
267: the two point functions between these fields are in general non-vanishing,
268: \be
269: (2\pi)^4 \delta^4(q_i+q_j) G^{ij}_{lm} (q_i) = % \int d^4x_1 d^4x_2  e^{-iq_1 \cdot x_1} e^{-iq_2 \cdot x_2}
270: \int_{FT}
271: \left\langle\Phi_0| T \left\{ \Psi^i_l(x_i) \Psi^{j\dagger}_m(x_j)  \right\} |\Phi_0\right\rangle.
272: \ee
273: Use the Lorentz invariance to write
274: \be
275: \left\langle\Phi_0| \Psi^i_l(0) |\Phi_{\vec{q}\sigma}\right\rangle  = N_i u_l(q,\sigma).
276: \ee
277: As $q^2 \rightarrow \bar{s}$, Eq. (\ref{pole-eq}) gives\footnote
278: {We assume that there is no degeneracy in mass for the same type of particles.
279: In case there is a degeneracy in mass, the particles will be distinguished by
280: their other quantum numbers and can be easily separated.
281: If there are particles with the same mass and other quantum numbers,
282: they can obviously be described by one quantum field.}
283: \be
284: G^{ij}_{lm} (q)
285: \rightarrow
286: N_i N_j^* {i \over q^2-\bar{s}} \sum_\sigma u_l(q,\sigma)  u^*_m(q,\sigma).
287: \label{Ndef}
288: \ee
289: %
290: \begin{figure}
291: \begin{center}
292: {\epsfxsize=2.500truein \epsfbox{dia.ps}} 
293: %\includegraphics[width=3.5in]%{fig1.ps}
294: \caption{\label{feynmanD} General structure of Feynman diagrams.}
295: \end{center}
296: \end{figure} 
297: On the other hand, an inspection of the structure of Feynman diagrams yields (Figure 1),
298: \be
299: G = \sum_{jm} G^{ij}_{lm} (q) \Gamma^j_m(q\cdots),
300: \ee
301: where $\Gamma^j_m$ is the sum of all Feynman diagrams
302: by stripping away the particle propagator $G^{ij}_{lm} (q)$.
303: As an intermediate step, we define a new truncated matrix element
304: \be
305: \sum_{jm} N_j^* u^*_m(q,\sigma) \Gamma^j_m(q\cdots).
306: \label{Sdef}
307: \ee
308: By repeating the process for all operators in $G$
309: and defining corresponding truncated matrix elements at each step, one finally gets the $S$-matrix element
310: up to an (irrelevant) overall phase factor.
311: This procedure has the salient feature which is independent of renormalization schemes.
312: In the on-shell scheme, one effectively defines a linear combination of $\Psi^i_l$ 
313: such that only one of the $N_i$'s is unity while all others vanish on each mass-pole.
314: Obviously this can only be performed recursively and is potentially tedious in practice.
315: Note that mixing angles between different fields
316: do not show up in Eq. (\ref{Sdef}) explicitly and can be avoided in principle.
317: 
318: We now express the $N$'s in terms of one-particle-irreducible (1PI) two-point functions.
319: To proceed, we need to be specific.
320: The 1PI two-point functions for scalar fields can be expressed as
321: \be
322:   \Gamma^B_{ij}(p^2) = p^2 Z_{ij}  - m^2_{ij} - \Sigma_{ij}(p^2),
323: \ee
324: where $Z_{ij}$ and $m^2_{ij}$ are coefficients of kinetic terms in the Lagrangian
325: and $\Sigma_{ij}$ are due to quantum loop contributions.
326: When normalized canonically, $Z$ is a unit matrix and $m^2$ is diagonal,
327: which are not required in our formalism.
328: Define $\gamma^B_{ij}$ as the residual matrix
329: of $\Gamma^B$ by crossing out its $j$-th row and the $i$-th column.
330: Denote $\Delta^B(p^2)=Det\left[\Gamma^B(p^2)\right]$
331: and $\Delta^B_{ij}(p^2)=(-)^{i+j}Det\left[\gamma^B_{ij}(p^2)\right]$.
332: The inverse of $\Gamma^B$ yields two-point functions of scalar fields up to a factor of $i$, 
333: \be
334: S_{ij}^B = {i \Delta^B_{ij}(p^2) \over \Delta^B(p^2)}.
335: \ee
336: $\bar{s}$ are determined from the equation $ \Delta^B(p^2) = 0$, its real part gives the mass-square of
337: the particle and the imaginary part the total decay width. The residue of $S_{ij}^B$ on the pole gives
338: \be
339: N_i N_j^* =  \lim_{p^2 \rightarrow \bar{s}} \left[ {d \Delta^B(p^2) \over dp^2} \right]^{-1}
340:  \Delta^B_{ij}(p^2),
341: \label{Nscalar}
342: \ee
343: from which the $N_i$'s are determined up to an (irrelevant) overall phase factor.
344: An application of Eqs. (\ref{Sdef}) and (\ref{Nscalar}) to the Higgs sector in the
345: Minimal Supersymmetric Standard Model readily yields 
346: results in the literature  \cite{susy}.
347: 
348: For vector fields, only gauge theories are known to be consistent.
349: Gauge fields are either massless or acquire masses via spontaneously symmetry breaking,
350: which can be realized in one way by introducing non-vanishing vacuum expectation values of scalars.
351: The 1PI two-point functions for the transverse components of vector fields are in general
352: \be
353:   \Gamma_{ij}^{\mu\nu}(p^2) = - g^{\mu\nu} \Gamma^V_{ij}(p^2) + \cdots,
354: \ee
355: from which the $N$'s can be obtained in the same manner as that of scalar fields.
356: For their longitudinal components $V^{i\mu}_L$, the situation is more subtle.
357: They have non-physical poles which are gauge dependent.
358: Fortunately, these non-physical poles are canceled exactly by
359: the same ones from the would-be Goldstone bosons due to BRST invariance \cite{BRST},
360: so they decouple from the $S$-matrix elements.
361: Accordingly, the Higgs boson masses can be selected from the whole set of scalar field poles
362: by excluding the would-be Goldstone bosons,
363: which are identical to the ones of $V^{i\mu}_L$'s.
364: However,
365: the correlation functions between $V^{i\mu}_L$'s and scalar fields need to be included
366: in calculating the $S$-matrix elements of Higgs bosons.
367: % The 1PI two-point functions of the theory assumes the general form
368: % \be
369: % \pmatrix{ \left(-g_{\mu\nu}+{p_\mu p_\nu \over p^2} \right) \Pi_T + {p_\mu p_\nu \over p^2} \Pi_L, &
370: %   p_\mu \Pi_{Lj} \cr
371: %   p_nu \Pi_{i,L},&  \Gamma_{ij}}
372: %\ee
373: %Define $\hat{\Gamma}_{ij} = \Gamma_{ij} - p^2 \Pi_{i,L} (\Pi_L)^{-1} \Pi_{Lj}$
374: %and its residual matrix $\hat{\gamma}_{ij}$  by crossing out its $j$-th row and $i$-th column.
375: %Denote $\hat{\Delta}(p^2)=Det\left[\bar{\Gamma}(p^2)\right]$
376: %and $\hat{\Delta}_{ij}(p^2)=(-)^{i+j} Det\left[\bar{\gamma}_{ij}(p^2)\right]$.
377: %$\hat{\Delta}(p^2)=0$ gives $n$ poles, $n-1$ of which are physical Higgs bosons and one of which coincides 
378: %the pole of the longitudial part of the vector field, 
379: %$\Pi_L - p^2 p_\mu \Pi_{Lj} \Gamma^{-1}_{ji} \Pi_{i,L}$.
380: 
381: For Dirac fields without parity violation, 
382: %\be
383: %\Gamma_{ij}(p) = \not{p} Z_{ij}  - m_{ij} - \tilde{\Sigma}_{ij}(\not{p})
384: %\ee
385: one gets by Lorentz invariance,
386: \ba
387: \Gamma^D_{ij}(p) &=& \not{p} Z_{ij}  - m_{ij} - \not{p}  \Sigma^{(1)}_{ij}(p^2) - \Sigma^{(2)}_{ij}(p^2)
388: % \nonumber \\ &=& 
389: = \not{p} \Pi_{ij}(p^2)   - \Sigma_{ij}(p^2)
390: \ea
391: Define $\bar{\Gamma} = p^2  \Pi - \Sigma \Pi^{-1} \Sigma$ and its
392: residual matrix $\bar{\gamma}_{ij}$  by crossing out its $j$-th row and $i$-th column.
393: Denote $\bar{\Delta}(p^2)=Det\left[\bar{\Gamma}(p^2)\right]$
394: and $\bar{\Delta}_{ij}(p^2)=(-)^{i+j} Det\left[\bar{\gamma}_{ij}(p^2)\right]$.
395: The inverse of $\Gamma^D$ yields two-point functions of Dirac fields up to a factor of $i$, 
396: \be
397: S_{ij}^D = {i \over \bar{\Delta}} 
398:  \left[ \not{p} \bar{\Delta}_{ij} + (\Pi^{-1}\Sigma)_{ik} \bar{\Delta}_{kj} \right] 
399: \ee     
400: The pole position $\bar{s}$ is determined from the equation $\bar{\Delta}(p^2) = 0$
401: and the residue of $S_{ij}^D$ on the pole gives
402: \be
403: N_i N_j^* = \lim_{p^2 \rightarrow \bar{s}} \left[ {d \bar{\Delta} \over dp^2} \right]^{-1}
404:  \bar{\Delta}_{ij}.
405: \ee
406: 
407: For Dirac fields with parity violation, 
408: \ba
409: \Gamma^F_{ij}(p) & = &
410:  \left[ \not{p} Z^L_{ij}  - m^L_{ij} - \not{p}  \Sigma^{L(1)}_{ij}(p^2) - \Sigma^{L(2)}_{ij}(p^2)
411:    \right] P_L  % \nonumber \\ &&
412: +\left[ \not{p} Z^R_{ij}  - m^R_{ij} - \not{p}  \Sigma^{R(1)}_{ij}(p^2) - \Sigma^{R(2)}_{ij}(p^2)
413:    \right] P_R \nonumber \\
414: &=& \left[ \not{p} \Pi_{ij}^L(p^2) -  \Sigma_{ij}^L(p^2) \right] P_L 
415:   + \left[ \not{p} \Pi_{ij}^R(p^2) -  \Sigma_{ij}^R(p^2) \right] P_R 
416: \ea
417: where $P_{L,R}=(1\mp\gamma_5)/2$ are the projection matrices.
418: Define $\Gamma^{R,L} = p^2  \Pi^{R,L} - \Sigma^{L,R} (\Pi^{L,R})^{-1} \Sigma^{R,L} $ and their
419: residual matrices $\gamma_{ij}^{R,L}$ by crossing out their $j$-th rows and $i$-th columns.
420: Denote $\Delta^{L,R}(p^2)=Det\left[\Gamma^{L,R}(p^2)\right]$
421: and $\Delta^{L,R}_{ij}(p^2)=(-)^{i+j} Det\left[\gamma^{L,R}_{ij}(p^2)\right]$.
422: The two-point functions are again the inverse of $\Gamma^F$ up to a factor $i$,
423: \ba
424: S^F_{ij} & = & {i \over \Delta^R} 
425:  \left\{ \not{p} \Delta^R_{ij} + \left[(\Pi^L)^{-1}\Sigma^R \right]_{ik} \Delta^R_{kj} \right\} P_L
426: % \nonumber \\ & & 
427: +{i \over \Delta^L}
428:  \left\{ \not{p} \Delta^L_{ij} + \left[(\Pi^R)^{-1}\Sigma^L \right]_{ik} \Delta^L_{kj} \right\} P_R 
429: \ea
430: From $\Delta^R(p^2) = 0$ and $\Delta^L(p^2) = 0$, one gets the same set of pole positions.
431: From the correspondence of pole positions,
432: one identifies the left- and right- handed components of Dirac fermions.
433: Define $N_{iL,R}$
434: \be
435: \left\langle\Phi_0| \psi_{iL,R}(0) |\Phi_{\vec{q}L,R}\right\rangle  = N_{iL,R} u_{L,R}(q)
436: \ee  
437: The residue of $S_{ij}^F$ on the pole gives
438: \ba
439: N_{iL,R} N_{jL,R}^* &=& \lim_{p^2 \rightarrow \bar{s}} 
440: \left[  {d \Delta^{L,R}(p^2) \over dp^2} \right]^{-1} \Delta^{L,R}_{ij}(p^2),
441: % \nonumber \\
442: %N_{iL,R} N_{jR,L}^* &=& {1\over m}
443: % \lim_{p^2 \rightarrow \bar{s}} \left[ {d \Delta^{L,R}(p^2) \over dp^2} \right]^{-1}
444: % \left[ (\Pi^{L,R})^{-1}\Sigma^{L,R} \right]_{ik}  \Delta^{R,L}_{kj}(p^2) 
445: \ea
446: from which we determine the $N$'s up to an overall phase factor 
447: %The matrix elements are 
448: %\be
449: %\sum_{jm} N_{jL,R}^* u^*_{L,R}(q,\sigma) \Gamma^{jL,R}_m(q\cdots),
450: % \ \sum_{jm} N_{jR}^* u^*_m(q,\sigma) \Gamma^{jR}_m(q\cdots),
451: %\ee                 
452: for left- and right-handed fermions, respectively.  
453: These results also apply for Majorana fields, for which 
454: $\Pi^L = \left( \Pi^R \right)^*$ and $ \Sigma^L =\left( \Sigma^R \right)^*$, 
455: so  $N_{jL}=N_{jR}^*$. 
456: 
457: In summary, we have in Eqs. (\ref{Ndef}) and (\ref{Sdef}) provided a 
458: general procedure to extract $S$-matrix elements from Green functions
459: in arbitrary renormalization schemes. 
460: Furthermore, we have determined the normalization coefficients $N$
461: for scalar, vector, and various types of spin-1/2 fermion fields.
462: The analysis can be readily extend to cases of arbitrary spin
463: and our formalism is applicable to calculations in arbitrary quantum field theories.
464: It should be, in particular, useful for sophisticated theories such as
465: softly-broken supersymmetric ones,
466: where a full realization of the on-shell scheme is complicated if not impossible.
467: 
468: {\bf Acknowledgments:}
469: We would like to thank W.~Hollik for valuable discussions 
470: and to acknowledge the hospitality of Max Plank Institut fur Physik (Werner Heisenberg Institut).
471: This work was supported in part by CNSF-90103009
472: and a fund for Trans-Century Talents.
473: %<<<<<
474: 
475: %\vspace{-0.5cm}
476: \begin{thebibliography}{10}
477: \bibitem{LLM}
478:    P. Langacker, M. Luo, and A. K. Mann, {\em Rev. Mod. Phys.} {\bf 64}, 87 (1992) and references within.
479: \bibitem{sirlin}
480:    A. Sirlin, \prt{22}{971} (1980).
481: \bibitem{msbar}
482:    G. 't Hooft and M. Veltman, \npb{44}{189} (1972);
483:    W. Bardeen, A. Buras, D. Duke, and T. Muta, \prt{18}{3998} (1978).
484: \bibitem{lsz}
485:    H. Lehmann, K. Symanzik, and W. Zimmermann, {\em Nuovo Cimento} {\bf 1}, 1425 (1955).
486: \bibitem{BD}
487:    J. Bjorken and S. Drell, {\sl Relativisitic Quantum Fields}, McGraw-Hill Book Company, New York (1965).
488: \bibitem{Weinberg}
489:    S. Weinberg, {\sl The Quantum Theory of Fields}, Cambridge University Press, Cambridge (1995). 
490: \bibitem{susy}
491:    See, for example, 
492:    P. Chankowski, S. Pokorski, and Janusz Rosiek, \npb{423}{437} (1994);
493:    S. Heinemeyer, W. Hollik, and G. Weiglein, {\em Eur. Phys. J.} {\bf C16}, 139 (2000) and references within. 
494: \bibitem{BRST}
495:    C. Becchi, A. Rouet, and R. Stora, {\em Comm. Math. Phys.} {\bf 42}, 127 (1975);
496:    I. V. Tyutin, Lebedev Institute preprint N39 (1975).
497: \end{thebibliography}
498: \end{document}
499: