1: %% LyX 1.3 created this file. For more info, see http://www.lyx.org/.
2: %% Do not edit unless you really know what you are doing.
3: \documentclass[12pt,english,a4]{article}
4: \usepackage[T1]{fontenc}
5: \usepackage[latin1]{inputenc}
6: \usepackage{geometry}
7: \geometry{verbose,a4paper,tmargin=2cm,bmargin=2cm,lmargin=2cm,rmargin=2cm}
8:
9: \makeatletter
10:
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% LyX specific LaTeX commands.
12: %% Bold symbol macro for standard LaTeX users
13: \newcommand{\boldsymbol}[1]{\mbox{\boldmath $#1$}}
14:
15:
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% User specified LaTeX commands.
17: \usepackage{color}
18: \usepackage{graphicx}
19: \usepackage{setspace}
20:
21:
22: \@addtoreset{equation}{section}
23: \renewcommand{\theequation}{\thesection.\arabic{equation}}
24: \makeatletter
25: \renewcommand\@makefnmark{\@textsuperscript{\normalfont(\@thefnmark)}}
26: \makeatother
27:
28: \usepackage{babel}
29: \makeatother
30: \begin{document}
31: \vspace*{0.8cm}
32:
33: \thispagestyle{empty}
34:
35: \begin{center}\textbf{\huge Radiative Corrections, New Physics Fits
36: and $e^{+}e^{-}\rightarrow?\rightarrow f\bar{f}$.}\end{center}{\huge \par}
37:
38: \begin{center}{\huge \vspace*{1cm}}\end{center}{\huge \par}
39:
40: \begin{center}{\large K.M.Hamilton}%
41: \footnote{email: k.hamilton1@physics.ox.ac.uk%
42: }\end{center}{\large \par}
43:
44: \begin{center}\emph{\large $\textrm{Denys Wilkinson Building, Keble Road, Oxford OX1 3RH, U.K.}$}\end{center}{\large \par}
45:
46: \begin{center}\vspace*{0.5cm}\end{center}
47:
48: \begin{center}{\large November 2003}\end{center}{\large \par}
49:
50: {\large \par{}}{\large \par}
51:
52: \begin{center}\vspace*{1cm}\end{center}
53:
54: \begin{center}\textbf{\large Abstract}\end{center}{\large \par}
55:
56: {\large \par{}}{\large \par}
57:
58: \noindent This document reviews the general approach to correcting
59: the process $e^{+}e^{-}\rightarrow X^{0}\rightarrow f\bar{f}$ for
60: radiative effects, where $X^{0}$ represents an exchanged boson arising
61: from some new physics. The validity of current methods is discussed
62: in the context of the differential cross section. To this end the
63: universality of the dominant QED radiative corrections to such processes
64: is discussed and an attempt is made to quantify it. The paper aims
65: to justify, as much as possible, the general approach taken by $e^{+}e^{-}$
66: collider experiments to the issue of how to treat the dominant radiative
67: corrections in fitting models of new physics using inclusive and exclusive
68: cross section measurements. We conclude that in all but the most pathological
69: of new physics models the dominant radiative corrections (QED) to
70: the tree level processes of the standard model can be expected to
71: hold well for new physics. This argument follows from the fact that
72: the phase space of indirect new physics searches is generally restrictive
73: (high $\sqrt{s^{\prime}}$ events) in such a way that the factorization
74: of radiative corrections is expected to hold well and generally universal
75: infrared corrections should be prevalent.
76:
77: \newpage
78:
79:
80: \section{Introduction.}
81:
82: In this document we discuss how radiative corrections which significantly
83: affect the standard model difermion process maybe exploited such that
84: they can be used with analogous difermion processes containing new
85: physics components, \emph{i.e.} we discuss the universality of radiative
86: corrections to the difermion process. To this end we must derive again
87: the significant QED corrections to the difermion process. The results
88: of our derivations agree with those in the literature. It has also
89: long been known that QED radiative corrections can exhibit soft behavior
90: which is independent of the underlying process, the so called \emph{hard
91: process} (see \emph{e.g.} \cite{Renton:td,Peskin:ev}), it is this behaviour
92: that we wish to find and study. The known radiative corrections generally
93: manifest themselves as long and complicated analytic expressions and
94: one generally has no clue as to whether they are significant or universal
95: or, ideally, both. In reviewing and {}``dismantling'' known results
96: we are able to thoroughly check the dependence of radiative corrections
97: on the hard scattering process thus highlighting, as much as possible,
98: universal parts of radiative corrections which are relevant to more
99: general difermion processes.
100:
101:
102: \section{Fitting Beyond the Standard Model Physics.}
103:
104: Currently there are a large number of models involving physics beyond
105: the standard model which may be constrained from measurements of the
106: difermion process $e^{+}e^{-}\rightarrow f\bar{f}$. New physics models
107: appear frequently in the literature which propose modifications to
108: these observables. It is common for these models to involve the exchange
109: of a new gauge boson mimicking the exchange of a $Z^{0}$ or photon
110: in difermion processes. Let us generically denote these particles
111: $X^{0}$. The result of such exchanges is to increase the measured
112: cross sections above that which one would expect for difermion processes
113: as defined above, the differential cross section will also be modified.
114:
115: The new physics contribution to the measured difermion cross section
116: is always given as a function of some parameters of the theory, typically
117: the energy scale of the new physics. The total new physics contribution
118: due to exchange of $X^{0}$ bosons comprises of a term due to interference
119: of $X^{0}$ exchange amplitudes with each of the $Z^{0}$ and $\gamma$
120: exchange amplitudes, as well as a pure $X^{0}$ exchange term. Given
121: that new physics effects are weak the largest contribution from them
122: comes from the interference term. By adding this new physics prediction
123: to the prediction for genuine difermion events and fitting the result
124: to the data for various values of the new physics parameters one can
125: obtain $\chi^{2}$ as a function of the parameters. From this $\chi^{2}$
126: function one can obtain confidence limits on the parameters.
127:
128: Due to the fact that the standard model generally agrees well with
129: experimental measurements any new physics effect which may be occurring
130: in current $e^{+}e^{-}$ collider experiments must be small. Generally
131: one is trying to set the most stringent constraints on the new physics
132: model one can. It is for these reasons that when selecting events
133: with which to set limits one should not include events for which $\sqrt{s^{\prime}}\ll\sqrt{s}$%
134: \footnote{$\sqrt{s^{\prime}}$ is the so-called \emph{reduced invariant mass}.
135: $\sqrt{s^{\prime}}<\sqrt{s}$ due to bremsstrahlung emitted from the
136: initial state particles. Generally $\sqrt{s^{\prime}}$ is the invariant
137: mass of the exchanged boson, it is defined in appendix \ref{sec:Appendix-A}.%
138: } as previous runs of the experiment will have directly probed these
139: energies (and found nothing) so including such effects will essentially
140: \emph{dilute} the effects of a new physics contribution resulting
141: in a more relaxed limit. Obviously there is a trade off between the
142: number of events in the sample and the amount by which new physics
143: would be diluted. At LEP2 limits are set on the various models of
144: new physics using only samples of events for which $\sqrt{s^{\prime}}\geq0.85\sqrt{s}$.
145:
146: Given that radiative corrections have striking consequences for difermion
147: events it would be naive to think that they are inconsequential for
148: new physics processes. New physics cross sections and differential
149: cross sections that appear in the literature are only ever given to
150: tree level, to obtain realistic limits for the parameters of the model
151: one should try to correct these predictions for radiative effects.
152: Effective theories are generally non-renormalizable, so the idea of
153: computing radiative corrections to processes derived from them may
154: not seem sensible. Nevertheless the idea of fitting the born level
155: predictions is certainly not correct. Finally, in the event that explicit
156: calculations of radiative corrections were possible, to do this would
157: be unfeasible given the current rate at which new physics scenarios
158: are conceived.
159:
160: The conventional method employed by experiments to improve the tree
161: level new physics predictions of an observable $O_{NP}$, to include
162: radiative effects $\left.O^{\prime}\right|_{NP}$, is to multiply
163: the tree level prediction by a simple factor of the standard model
164: prediction including radiative corrections $\left.O^{\prime}\right|_{SM}$
165: divided by the standard model prediction at tree level $\left.O\right|_{SM}$,\begin{equation}
166: \left.O^{\prime}\right|_{NP}=\left.O\right|_{NP}\times\frac{\left.O^{\prime}\right|_{SM}}{\left.O\right|_{SM}}.\label{eq:5.1.1}\end{equation}
167: This correction factor embodies the aggregated effects of all of
168: the \emph{standard model} radiative corrections. This approach amounts
169: to claiming that all standard model radiative corrections \emph{factorize}
170: trivially from the born level quantity, that radiative corrections
171: are independent of the born level process. In the rest of the paper
172: we investigate whether the approach of equation \ref{eq:5.1.1} is
173: at all valid.
174:
175:
176: \section{Radiator Functions: A {}``Black Box'' Empirical Study\label{sec:Radiator-Functions:-A}.}
177:
178: The effect of bremsstrahlung on $e^{+}e^{-}$ collisions is highly
179: significant, the data show that the cross-sections for events with
180: $\sqrt{s^{\prime}}\geq75\textrm{ GeV}$ is larger than those events
181: with $\sqrt{\frac{s^{\prime}}{s}}\geq0.85$ by a factor of two or
182: more. If one imagines a single bremsstrahlung correction to a difermion
183: process, in which one of the colliding particles emits a photon, one
184: effect will be to lower the invariant mass of the basic difermion
185: reaction we are trying to measure from $\sqrt{s}$ to $\sqrt{s^{\prime}}$
186: (or $\sqrt{t}$ to $\sqrt{t^{\prime}}$ for $t$-channel processes).
187: Assuming that the processes of emitting a photon and undergoing a
188: difermion reaction are otherwise independent, that is to say they
189: \emph{factorize}, one can think of the cross section $\sigma\left(s\right)$measured
190: at some value of $\sqrt{s}$, in some range $s_{min}\leq s^{\prime}\leq s$
191: as being comprised of a weighted sum over all of the tree level cross
192: sections $\sigma_{Tree}\left(s^{\prime}\right)$ in that $\sqrt{s^{\prime}}$
193: range,\begin{equation}
194: \sigma\left(s\right)=\int_{s_{min}}^{s}\textrm{d}s^{\prime}\textrm{ }R\left(s,s^{\prime}\right)\sigma_{Tree}\left(s^{\prime}\right).\label{eq:5.2.1}\end{equation}
195: $R\left(s,s^{\prime}\right)$ is the weighting function, it is known
196: in the literature as the \emph{radiator} \emph{function}. Given that
197: the cross section is proportional to the number of events in the sample
198: we could write\begin{equation}
199: N\left(s\right)=\int_{s_{min}}^{s}\textrm{d}s^{\prime}\textrm{ }R\left(s,s^{\prime}\right)N_{Tree}\left(s^{\prime}\right),\label{eq:5.2.2}\end{equation}
200: which tells us that $\textrm{d}s^{\prime}\textrm{ }R\left(s,s^{\prime}\right)N_{Tree}\left(s^{\prime}\right)$
201: is the number of tree level difermion events in the range $s^{\prime}\rightarrow s^{\prime}+\textrm{d}s'$
202: entering the measured sample of $N\left(s\right)$. Therefore $\textrm{d}s^{\prime}\textrm{ }R\left(s,s^{\prime}\right)$
203: is the fraction of the tree level cross section $\sigma_{Tree}\left(s^{\prime}\right)$
204: that goes into making $\sigma\left(s\right)$. What is being described
205: here is essentially the \emph{structure} \emph{function} approach,
206: this is the idea that the electron and positron are not to be thought
207: of as fundamental but rather they should be thought of as objects
208: \emph{containing} truly fundamental electrons, positrons and photons.
209: With this in mind one can define a structure function for the electron
210: $D_{e}\left(x\right)$. $D_{e}\left(x\right)$ is a probability density
211: function of $x$ which is the \emph{fraction} of the longitudinal
212: component of the electrons momentum that it has when it undergoes
213: the tree level hard scattering process. $D_{e}\left(x\right)\textrm{d}x$
214: is therefore the probability that the electron collides with the positron
215: with a fraction $x$ of its original longitudinal four momentum. The
216: positron structure function is the same as the electron structure
217: function. Neglecting transverse momenta for the time being we have\begin{equation}
218: s^{\prime}=x_{1}x_{2}s.\label{eq:5.2.3}\end{equation}
219: With these definitions the cross section for an electron and positron
220: to undergo a difermion interaction such that $s^{\prime}\geq s_{min}^{\prime}$
221: is\begin{equation}
222: \sigma\left(s\right)=\int_{\frac{s_{min}}{s}}^{1}\textrm{d}x_{1}\int_{\frac{s_{min}}{x_{1}s}}^{1}\textrm{d}x_{2}\textrm{ }D_{e}\left(x_{1}\right)D_{e}\left(x_{2}\right)\sigma_{Tree}\left(x_{1}x_{2}s\right).\label{eq:5.2.4}\end{equation}
223: Defining $1-x=x_{1}x_{2}$ and demanding it be greater than $\frac{s_{min}}{s}$
224: this can be written\begin{equation}
225: \begin{array}{rcl}
226: \sigma\left(s\right) & = & \int_{0}^{1-\frac{s_{min}}{s}}\textrm{d}x\left[\int_{1-x}^{1}\textrm{d}x_{2}\textrm{ }\frac{1}{x_{2}}\textrm{ }D_{e}\left(\frac{1-x}{x_{2}}\right)D_{e}\left(x_{2}\right)\right]\sigma_{Tree}\left(\left(1-x\right)s\right)\\
227: & = & \int_{0}^{1-\frac{s_{min}}{s}}\textrm{d}x\textrm{ }H\left(x,s\right)\sigma_{Tree}\left(\left(1-x\right)s\right)\end{array},\label{eq:5.2.5}\end{equation}
228: allowing an identification between the radiator function and the
229: structure functions. Clearly $H\left(x,s\right)$ and $R\left(s,s^{\prime}\right)$
230: are related by a factor.
231:
232: The most general new physics distribution to fit is the differential
233: cross section. In this case the quantity $\left.O\right|_{NP}$ to
234: be corrected via equation \ref{eq:5.1.1} is the value of the differential
235: cross section in a given angular bin. Conventionally one would then
236: use a number of correction factors determined using the aforementioned
237: standard model cross sections in that bin to correct the bins of the
238: new physics tree level angular distribution. It is however hard to
239: attach a physical meaning to these correction factors, they do not
240: lend themselves to a description in terms of a radiator function or
241: structure functions. Ideally one would like to employ a radiator function
242: approach to correcting new physics observables and so correct them
243: by folding the given tree level quantities with the radiator function.
244: This raises two questions, does a radiator function approach exist
245: for the differential cross section and if it does exist how should
246: it be implemented?
247:
248: Radiator functions for the differential cross section do exist and
249: are present inside $e^{+}e^{-}\rightarrow f\bar{f}$ simulation programs.
250: There are a variety of such programs which are freely available. There
251: are two kinds of $e^{+}e^{-}\rightarrow f\bar{f}$ simulation packages.
252: Semi-analytic simulations such as \texttt{TOPAZ0} \cite{Montagna:1998kp}
253: and \texttt{ZFITTER} \cite{Bardin:1999yd} produce predictions of
254: observables like cross-sections and differential cross sections. They
255: contain analytic expressions for these quantities corrected for a
256: myriad of radiative corrections all in the context of the standard
257: model. Some outputs of these programs involve numerical integrations
258: hence {}``Semi-analytical''. In addition to the semi-analytic packages
259: there are difermion Monte Carlo programs such as ${\cal {KK}}\textrm{2f}$
260: \cite{Ward:2002qq} which simulate actual difermion events \emph{i.e.}
261: there output includes four vectors for the various final state particles.
262: At the heart of the Monte Carlo is a random number generator, as the
263: number of generated events tends to infinity the cross sections \emph{etc}
264: that are calculable with the generated events tend to those of the
265: standard model \emph{i.e.} random fluctuations die away. Naturally
266: the semi-analytic packages have the advantage of requiring much less
267: computational time than the Monte Carlo programs and the \texttt{ZFITTER}
268: package appears to be the most popular of these in the experimental
269: community.
270:
271: In \cite{Bardin:ak} (and the \texttt{ZFITTER} manual \cite{Bardin:1999yd})
272: the differential cross section for difermion production in the presence
273: of ISR is given in the form%
274: \footnote{The form of equation \ref{eq:5.2.6} in \cite{Bardin:1999yd} differs
275: slightly from what we quote which is a simplified version. The only
276: difference of note is that \cite{Bardin:1999yd} split the born level
277: cross section into a part which is symmetric in $\cos\theta$ and
278: an asymmetric part, each of which is convoluted with a slightly different
279: radiator function. Equation \ref{eq:5.2.6} is appropriate for the
280: time being. %
281: }\begin{equation}
282: \frac{{\rm {d}}\sigma\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}=\int_{s_{min}^{\prime}}^{s}{\rm {d}}s^{\prime}\textrm{ }R\left(s,s^{\prime},\cos\theta\right)\sigma_{Tree}\left(s^{\prime}\right).\label{eq:5.2.6}\end{equation}
283: Note crucially that \emph{all} angular dependence is contained within
284: the radiator function. Let us denote the cross section for a new physics
285: process incorporating radiative corrections by $\sigma_{NP}$ and
286: the corresponding tree level cross section $\sigma_{Tree,NP}$. Assuming
287: that it is possible to extract $R\left(s,s^{\prime},\cos\theta\right)$
288: from the \texttt{ZFITTER} package\begin{equation}
289: \frac{{\rm {d}}\sigma_{NP}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}=\int_{s_{min}^{\prime}}^{s}{\rm {d}}s^{\prime}\textrm{ }R\left(s,s^{\prime},\cos\theta\right)\sigma_{Tree,NP}\left(s^{\prime}\right),\label{eq:5.2.7}\end{equation}
290: will not be the correct differential cross section for a new physics
291: process, obviously there is absolutely no reference to the angular
292: distribution of new physics in the above formula, only to that of
293: the standard model. Studying references \cite{Bardin:ak} and \cite{Bardin:1999yd}
294: in a more detail one finds that in fact the radiator function which
295: exactly factorizes at the level of the integrated cross section (by
296: integrated we mean integrated over the full solid angle), as shown
297: in \ref{eq:5.2.6}, also \emph{largely} factorizes from the differential
298: cross section \emph{i.e.} \ref{eq:5.2.6} is of the form\begin{equation}
299: \begin{array}{rcl}
300: \frac{{\textrm{{d}}}\sigma\left(s,\cos\theta\right)}{{\textrm{{d}}}\cos\theta} & = & \int_{s_{min}^{\prime}}^{s}{\textrm{{d}}}s^{\prime}\textrm{ }\tilde{R}_{1}\left(s,s^{\prime}\right)\frac{{\rm {d}}\sigma_{Tree}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}+\int_{s_{min}^{\prime}}^{s}{\textrm{{d}}}s^{\prime}\textrm{ }\tilde{R}_{2}\left(s,s^{\prime},\cos\theta\right)\sigma_{Tree}\left(s^{\prime}\right)\\
301: & = & \int_{s_{min}^{\prime}}^{s}{\textrm{{d}}}s^{\prime}\textrm{ }\tilde{R}_{1}\left(s,s^{\prime}\right)\frac{{\rm {d}}\sigma_{Tree}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}+\textrm{Non-factorizable.}\end{array}.\label{eq:5.2.8}\end{equation}
302: In \ref{eq:5.2.8} $\textrm{ }R\left(s,s^{\prime},\cos\theta\right)$
303: has been split into a factorizable part%
304: \footnote{Henceforth we use the word factorizable to mean that the radiative
305: corrections can be represented by a radiator function convoluted with
306: the \emph{differential cross section} $\textrm{d}\sigma$.%
307: } $\tilde{R}_{1}\left(s,s^{\prime}\right)\frac{{\rm {d}}\sigma_{0}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}$
308: and a non-factorizable part $\tilde{R}_{2}\left(s,s^{\prime},\cos\theta\right)$.
309: If the parts of the above which do not factorize into the differential
310: cross section are negligible then it appears that the radiator function
311: does not depend on the tree level differential cross section, in fact
312: it does not depend on angles at all. If this is true then $\tilde{R}_{1}\left(s,s^{\prime}\right)$
313: would appear to be independent of the tree level physics analogous
314: to a structure function (the electron structure function is discussed
315: in the next section), then one would appear to have a recipe for making
316: new physics angular distribution corrected for radiative effects\begin{equation}
317: \frac{{\rm {d}}\sigma_{NP}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}=\int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }\tilde{R}_{1}\left(s,s^{\prime}\right)\frac{{\rm {d}}\sigma_{T,NP}\left(s,\cos\theta\right)}{{\rm {d}}\cos\theta}.\label{eq:5.2.9}\end{equation}
318: Assuming it is true that the non-factorizable parts of $R\left(s,s^{\prime},\cos\theta\right)$
319: are negligible we then wish to obtain $\tilde{R}_{1}\left(s,s^{\prime}\right)$.
320:
321: The \texttt{ZFITTER} package contains just this information though
322: as much as the function is convoluted with the standard model in the
323: above equations it is convoluted to an greater extent with the standard
324: model in terms of the actual coding of the package. A different, more
325: practical means of deconvolving the radiator function is required.
326: Denoting $\sigma\left(s,c_{b}\right)$ as the cross section for the
327: difermion process occurring such that the final state fermion goes
328: into the $\cos\theta$ bin in the range $\cos\theta_{b}\leq\cos\theta\leq\cos\theta_{b+1}$
329: denoted by $c_{b}$ we have\[
330: \begin{array}{rcl}
331: \sigma\left(s,c_{b}\right) & = & \int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }\left\{ \tilde{R}_{1}\left(s,s^{\prime}\right)\sigma_{T}\left(s,c_{b}\right)+\int_{\cos\theta_{b}}^{\cos\theta_{b+1}}{\rm {d}}\cos\theta\textrm{ }\int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }\tilde{R}_{2}\left(s,s^{\prime},\cos\theta\right)\sigma_{T}\left(s^{\prime}\right)\right\} \\
332: & = & \int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }\left\{ \tilde{R}_{1}\left(s,s^{\prime}\right)\sigma_{T}\left(s,c_{b}\right)+\int_{\cos\theta_{b}}^{\cos\theta_{b+1}}{\rm {d}}\cos\theta\textrm{ Non-factorizable}\right\} \end{array},\]
333: where here {}``Non-factorizable'' is the same as in \ref{eq:5.2.8}.
334: The method by which we extract the radiator function assumes that
335: the non-factorizable part is zero. If this assumption is true one
336: should get the same function $\tilde{R}\left(s,s^{\prime}\right)$
337: regardless of what binning one uses in the equation above. The fact
338: that the non-factorizable part has a complex angular form (this can
339: be checked by reference to \cite{Bardin:1999yd}) will mean that the
340: $\tilde{R}_{1}\left(s,s^{\prime}\right)$ which one extracts from
341: the program could look very different depending on which angular bin
342: is used and, importantly, the size of the non-factorizable contributions.
343: If the non-factorizable component really is small then the $\tilde{R}_{1}\left(s,s^{\prime}\right)$
344: which is extracted from the program should not be a function of the
345: bin used to extract it, it should be flat when plotted against the
346: bin used to extract it. The author has devised a means of extracting
347: a numerical version of $\tilde{R}_{1}\left(s,s^{\prime}\right)$ from
348: the \texttt{ZFITTER} package by making finite approximations to the
349: integral over $s^{'}$. The resulting numerical radiator function
350: is shown in figure \ref{cap:nointf}, it is, as hoped, flat across
351: the angular region shown.%
352: \begin{figure}
353: \begin{center}\includegraphics[%
354: width=0.30\paperwidth,
355: height=0.35\paperwidth]{200MMrad_no_intf.eps}\end{center}
356:
357:
358: \caption{\label{cap:nointf}The radiator function $\tilde{R}_{1}\left(s,s^{\prime}\right)$
359: with $s=\left(200\textrm{ GeV}\right)^{2}$, obtained from the \texttt{ZFITTER}
360: package by extracting it separately for each angular bin for the $\mu^{+}\mu^{-}$
361: final state.}
362: \end{figure}
363:
364:
365:
366: \section{Structure Functions, Leading Log Approximation\label{sec:Structure-Functions,-Leading}.}
367:
368: The purpose of this section is to add weight to the empirical observations
369: of the last section by deriving the electron structure function. The
370: idea for a structure function for the electron originates from the
371: paper of Gribov and Lipatov \cite{Gribov:rt}. In the papers of Kuraev,
372: Fadin, Jadach and Skrzypek \cite{Skrzypek:1990qs,Kuraev:hb} the authors
373: consider the process of $e^{+}e^{-}$ annihilation as being a Drell-Yan
374: process and exploit the perturbative nature of QED to calculate the
375: electron structure function analytically using the Gribov-Lipatov
376: Altarelli-Parisi equations. This structure function approach represents
377: a \emph{leading logarithmic} \emph{approximation} (\emph{LLA}), this
378: corresponds to assuming near collinear emission of photons from the
379: incoming $e^{+}e^{-}$. The radiator function in the literature is
380: by no means unique. The structure function approach / the leading
381: log approximation, is one way to take into account the effects of
382: ISR. The structure function approach is intuitive, simple and it produces
383: a cross section to better than $0.015$\% \cite{Skrzypek:1992vk}%
384: \footnote{This number is quoted in relation to the accuracy of the LLA around
385: the $Z^{0}$ peak at LEP1 however in LEP2 studies \cite{Boudjema:1996qg}
386: the LEP1 and higher order LEP2 structure functions have been shown
387: to produce cross sections agreeing to better than 0.1\% for cuts like
388: those used in obtaining our non-radiative class of events. %
389: }, this alone is a good reason to discuss it. For clarity and completeness
390: we review the results of Kuraev and Fadin (KF) as well as Jadach and
391: Skrzypek (JS) with the help of a combination of papers \cite{Skrzypek:1990qs,Skrzypek:1992vk,Kuraev:hb,Gribov:ri,Nicrosini:1986sm}.
392: In doing this we explicitly confirm (as noted in the introduction
393: of \cite{Skrzypek:1992vk}) that no element of the underlying hard
394: scattering process enters this description of ISR%
395: \footnote{It had been a cause of concern that, contrary to QCD, the perturbative,
396: calculable nature of this structure function may somehow allow for
397: some specifically process dependent corrections to enter it, as is
398: suggested by the explicit angular dependence in the radiator functions
399: of \cite{Bardin:1999yd}.%
400: } and check whether or not it may be applied to the differential cross
401: section as well as the integrated cross section.
402:
403: The \emph{non-singlet} structure function $D^{NS}\left(x,s\right)$
404: corresponds to the probability of finding an electron with longitudinal
405: momentum fraction $x$ `inside' the original on-shell electron at
406: energy scale $\sqrt{s}$. The generic starting point for determining
407: the structure function is Lipatov's equation (see \emph{e.g.} \cite{Peskin:ev})\begin{equation}
408: D^{NS}\left(x,s\right)=\delta\left(1-x\right)+\int_{m_{e}^{2}}^{s}\frac{{\rm {d}}s_{1}}{s_{1}}\frac{\alpha\left(s_{1}\right)}{2\pi}\int_{x}^{1}\frac{{\rm {d}}x_{1}}{x_{1}}\textrm{ }P\left(x_{1}\right)D^{NS}\left(\frac{x}{x_{1}},s_{1}\right).\label{eq:5.3.1}\end{equation}
409: where $P\left(z\right)$ is the regularized splitting function\begin{equation}
410: P\left(x_{1}\right)=\frac{1+x_{1}^{2}}{1-x_{1}}-\delta\left(1-x_{1}\right)\int_{0}^{1}\textrm{d}z\textrm{ }\frac{1+z^{2}}{1-z}.\label{eq:5.3.2}\end{equation}
411: For our purposes a more useful form of the regularized splitting
412: function involves introducing a cut off at $z=1-\epsilon$ where $\epsilon\ll1$
413: is very small\begin{equation}
414: -\int_{0}^{1}\textrm{d}z\textrm{ }\frac{1+z^{2}}{1-z}=\lim_{\epsilon\rightarrow0}2\left(\frac{3}{4}+\log\epsilon\right)-2\epsilon+\frac{1}{2}\epsilon^{2}.\label{eq:5.3.3}\end{equation}
415: In this way we can define a different regularized splitting function\begin{equation}
416: \textrm{ }P\left(x_{1}\right)=\textrm{ }2\delta\left(1-x_{1}\right)\left(\frac{3}{4}+\log\epsilon\right)+\theta\left(1-\epsilon-x_{1}\right)\frac{1+x_{1}^{2}}{1-x_{1}}.\label{eq:5.3.4}\end{equation}
417: which is finite when integrated over in the limit $\epsilon\rightarrow0$.
418: Clearly this form is the natural extension of \ref{eq:5.3.2}, in
419: the limit $\epsilon\rightarrow0$ \ref{eq:5.3.2} and \ref{eq:5.3.3}
420: are identical. Henceforth we omit writing $\lim_{\epsilon\rightarrow0}$
421: and drop all terms from our calculations which are vanishing in the
422: limit $\epsilon\rightarrow0$.
423:
424: Lipatov's equation \ref{eq:5.3.1} has an obvious iterative solution,
425: to see this it is convenient to first rewrite the second integral
426: using\begin{equation}
427: \int_{0}^{1}\textrm{d}x_{1}\textrm{d}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)D^{NS}\left(x_{2},s_{1}\right)=\int_{x}^{1}\frac{\textrm{d}x_{1}}{x_{1}}\textrm{ }P\left(x_{1}\right)D^{NS}\left(\frac{x}{x_{1}},s_{1}\right).\label{eq:5.3.6}\end{equation}
428: In addition we approximate the running coupling constant to a constant
429: $\alpha\left(s\right)=\alpha$ and define $\eta\equiv\eta\left(s\right)=\frac{2\alpha}{\pi}\log\frac{s}{m_{e}^{2}}$,
430: $\eta_{i}\equiv\eta\left(s_{i}\right)$ therefore\begin{equation}
431: \textrm{d}\eta_{i}=\frac{\textrm{d}s_{i}}{s_{i}}\frac{2\alpha}{\pi}.\label{eq:5.3.6}\end{equation}
432: With these substitutions Lipatov's equation \ref{eq:5.3.1} becomes,\begin{equation}
433: D^{NS}\left(x,\eta\right)=\delta\left(1-x\right)+\frac{1}{4}\int_{0}^{\eta}\textrm{d}\eta_{1}\int_{0}^{1}\textrm{d}x_{1}\textrm{d}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)D^{NS}\left(x_{2},\eta_{1}\right)\label{eq:5.3.7}\end{equation}
434: Substituting relation \ref{eq:5.3.6} into Lipatov's equation \ref{eq:5.3.1}
435: and iterating it \emph{i.e.} substituting $D\left(x_{i},\eta_{i}\right)$as
436: defined by \ref{eq:5.3.1} back into \ref{eq:5.3.1} we have an expansion
437: in $\eta$;\begin{equation}
438: \begin{array}{rcl}
439: D^{NS}\left(x,\eta\right) & = & \delta\left(1-x\right)\\
440: & + & \frac{1}{4}\int_{0}^{\eta}{\rm {d}}\eta_{1}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)\delta\left(1-x_{2}\right)\\
441: & + & \frac{1}{4}\int_{0}^{\eta}{\rm {d}}\eta_{1}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)\frac{1}{4}\int_{0}^{\eta_{1}}{\rm {d}}\eta_{2}\int_{0}^{1}{\rm {d}}x_{3}{\rm {d}}x_{4}\textrm{ }\delta\left(x_{2}-x_{3}x_{4}\right)P\left(x_{3}\right).....\\
442: & + & {\mathcal{{O}}}\left(\eta^{3}\right)\end{array}.\label{eq:5.3.8}\end{equation}
443: The fact that the integrals $\int_{0}^{\eta_{i}}\textrm{d}\eta_{i}$
444: are \emph{nested}, that is to say the upper limit of one integrand
445: is the dummy variable of the one preceding it, results in a sequence
446: of exponential factors $\frac{1}{i!}$ multiplying the terms above.
447: \ref{eq:5.3.7} is condensed by simple delta function manipulations\begin{equation}
448: \begin{array}{rl}
449: & D^{NS}\left(x,\eta\right)\\
450: = & \delta\left(1-x\right)\\
451: + & 1\times\left(\frac{\eta}{4}\right)^{1}P\left(x\right)\\
452: + & \frac{1}{2}\times\left(\frac{\eta}{4}\right)^{2}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)P\left(x_{2}\right)\\
453: + & \frac{1}{6}\times\left(\frac{\eta}{4}\right)^{3}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)\int_{0}^{1}{\rm {d}}x_{3}{\rm {d}}x_{4}\textrm{ }\delta\left(x_{2}-x_{3}x_{4}\right)P\left(x_{3}\right)P\left(x_{4}\right)\\
454: + & {\mathcal{{O}}}\left(\eta^{4}\right)\end{array}.\label{eq:5.3.9}\end{equation}
455: Denoting the convolution integrals above by\begin{equation}
456: \int_{0}^{1}\textrm{d}x_{i}\textrm{d}x_{j}\textrm{ }\delta\left(x-x_{i}x_{j}\right)f\left(x_{i}\right)g\left(x_{j}\right)=f\left(.\right)\otimes g\left(.\right)\left(x\right),\label{eq:5.3.10}\end{equation}
457: equation \ref{eq:5.3.8} becomes\begin{equation}
458: D^{NS}\left(x,s\right)=\delta\left(1-x\right)+\sum_{i=1}^{\infty}\frac{1}{i!}\left(\frac{\eta}{4}\right)^{i}P^{\otimes i}\left(x\right),\label{eq:5.3.11}\end{equation}
459: where\begin{equation}
460: P^{\otimes3}\left(x\right)=\int_{0}^{1}\textrm{d}x_{1}\textrm{d}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)\int_{0}^{1}\textrm{d}x_{3}\textrm{d}x_{4}\textrm{ }\delta\left(x_{2}-x_{3}x_{4}\right)P\left(x_{3}\right)P\left(x_{4}\right)\label{eq:5.3.12}\end{equation}
461: \emph{etc}.
462:
463: The ${\mathcal{{O}}}\left(\eta\right)$ term is simply the regularized
464: splitting function, $\frac{1}{4}\eta P\left(x\right)$. In calculating
465: the ${\mathcal{{O}}}\left(\eta^{2}\right)$ term first we calculate
466: the \emph{theta} term, we proceed essentially in an identical way
467: to that in which we calculated the regularized splitting function.
468: We calculate a regularized version of this term by taking $x$ to
469: be less than a cut off $x<1-\epsilon$.\begin{equation}
470: \begin{array}{rl}
471: & \theta\left(1-x-\epsilon\right)\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)P\left(x_{2}\right)\\
472: = & \theta\left(1-x-\epsilon\right)4\left(\frac{3}{4}+\log\epsilon\right)^{2}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)\delta\left(1-x_{1}\right)\delta\left(1-x_{2}\right)\\
473: + & \theta\left(1-x-\epsilon\right)2\left(\frac{3}{4}+\log\epsilon\right)\int_{0}^{1-\epsilon}{\rm {d}}x_{1}\int_{0}^{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)\delta\left(1-x_{2}\right)\frac{1+x_{1}^{2}}{1-x_{1}}\\
474: + & \theta\left(1-x-\epsilon\right)2\left(\frac{3}{4}+\log\epsilon\right)\int_{0}^{1}{\rm {d}}x_{1}\int_{0}^{1-\epsilon}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)\textrm{ }\delta\left(1-x_{1}\right)\frac{1+x_{2}^{2}}{1-x_{2}}\\
475: + & \theta\left(1-x-\epsilon\right)\int_{0}^{1}{\rm {d}}x_{1}\int_{0}^{1}{\rm {d}}x_{2}\delta\left(x-x_{1}x_{2}\right)\textrm{ }\theta\left(1-\epsilon-x_{1}\right)\theta\left(1-\epsilon-x_{2}\right)\frac{1+x_{1}^{2}}{1-x_{1}}\frac{1+x_{2}^{2}}{1-x_{2}}\end{array}\label{eq:5.3.14}\end{equation}
476: The first term is zero due to the cut off we impose on $x$, the
477: other terms are also straightforward to evaluate. As stated earlier
478: small terms ${\mathcal{{O}}}\left(\epsilon\right)$ and above, vanishing
479: in the limit $\epsilon\rightarrow0$, are dropped. The total theta
480: term obtained with the cut off $x<1-\epsilon$ is \begin{equation}
481: \begin{array}{rl}
482: & \theta\left(1-\epsilon-x\right)\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)P\left(x_{2}\right)\\
483: = & \theta\left(1-x-\epsilon\right)2\left(\frac{1+x^{2}}{1-x}\left(2\log\left(1-x\right)-\log x+\frac{3}{2}\right)+\frac{1}{2}\left(1+x\right)\log x-1+x\right)\end{array}\label{eq:5.3.17}\end{equation}
484: The \emph{delta} term, analogous to the delta function term in the
485: regularized splitting function \ref{eq:5.3.4}, {}``soaks up'' the
486: divergent part of the term above arising from the limit $x\rightarrow1$
487: $\left(\epsilon\rightarrow0\right)$\begin{equation}
488: \begin{array}{rl}
489: & -\lim_{\epsilon\rightarrow0}\int_{0}^{1-\epsilon}{\rm {d}}x\textrm{ }2\left(\frac{1+x^{2}}{1-x}\left(2\log\left(1-x\right)-\log x+\frac{3}{2}\right)+\frac{1}{2}\left(1+x\right)\log x-1+x\right)\\
490: = & 4\left(\left(\frac{3}{4}+\log\epsilon\right)^{2}-\zeta\left(2\right)\right)\end{array}.\label{eq:5.3.18}\end{equation}
491: Finally the full regularized ${\mathcal{{O}}}\left(\eta^{2}\right)$
492: term is \begin{equation}
493: \begin{array}{rl}
494: & \frac{1}{2}\left(\frac{\eta}{4}\right)^{2}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)P\left(x_{2}\right)\\
495: = & \left(\frac{\eta}{4}\right)^{2}\delta\left(1-x\right)\left(2\left(\frac{3}{4}+\log\epsilon\right)^{2}-2\zeta\left(2\right)\right)\\
496: + & \left(\frac{\eta}{4}\right)^{2}\theta\left(1-x-\epsilon\right)\left(\frac{1+x^{2}}{1-x}\left(2\log\left(1-x\right)-\log x+\frac{3}{2}\right)+\frac{1}{2}\left(1+x\right)\log x-1+x\right)\end{array}.\label{eq:5.3.19}\end{equation}
497: The third order term, ${\mathcal{{O}}}\left(\eta^{3}\right)$, is
498: simply\begin{equation}
499: \frac{1}{6}\left(\frac{\beta}{4}\right)^{3}\int_{0}^{1}{\rm {d}}x_{1}{\rm {d}}x_{2}\textrm{ }\delta\left(x-x_{1}x_{2}\right)P\left(x_{1}\right)\int_{0}^{1}{\rm {d}}x_{3}{\rm {d}}x_{4}\textrm{ }\delta\left(x_{2}-x_{3}x_{4}\right)P\left(x_{3}\right)P\left(x_{4}\right).\label{eq:5.3.20}\end{equation}
500: Clearly the $\int_{0}^{1}\textrm{d}x_{3}\textrm{d}x_{4}$ part of
501: this term is given by simply substituting in the ${\mathcal{{O}}}\left(\eta^{2}\right)$
502: term with the replacement, $x\rightarrow x_{2}$. Noting this simplification
503: the resulting calculation is significantly simplified and a long but
504: straightforward sequence of integrations gives the ${\mathcal{{O}}}\left(\eta^{3}\right)$
505: theta and delta terms%
506: \footnote{The following dilogarithm identity is required, $\textrm{Li}_{2}\left(x\right)=\frac{\pi^{2}}{6}-\log\left(x\right)\log\left(1-x\right)-\textrm{Li}_{2}\left(1-x\right)$
507: to obtain the term in the form of \cite{Skrzypek:1990qs}.%
508: } of \cite{Skrzypek:1990qs}.
509:
510: The Lipatov equation can also be solved \emph{analytically} in the
511: soft limit $x\rightarrow0$ by means of a Mellin transformation. The
512: Mellin transform is defined as,\begin{equation}
513: \tilde{D}^{NS\left(z\right)}\left(\eta\right)=\int_{0}^{1}{\rm {d}}x\textrm{ }x^{z-1}D^{NS}\left(x,\eta\right)\label{eq:5.3.25}\end{equation}
514: for which the inverse is,\begin{equation}
515: D^{NS}\left(x,\eta\right)=\frac{1}{2\pi i}\int_{c-i\infty}^{c+i\infty}{\rm {d}}z\textrm{ }x^{-z}\tilde{D}^{\left(z\right)}\left(\eta\right).\label{eq:5.3.26}\end{equation}
516: The transform exists if $\int_{0}^{1}\textrm{d}x\textrm{ }x^{z-1}\left|D^{NS}\left(x,\eta\right)\right|$
517: is bounded for some $z>0$ in which case the inverse exists for $c>z$.
518:
519: Taking the Mellin transform of Lipatov's equation (\ref{eq:5.3.7})
520: gives\begin{equation}
521: \tilde{D}^{NS\left(z\right)}\left(\eta\right)=1+\frac{1}{4}\int_{0}^{\eta}\textrm{d}\eta_{1}\tilde{P}^{\left(z\right)}\tilde{D}^{NS\left(z\right)}\left(\eta_{1}\right)\label{eq:5.2.27}\end{equation}
522: Differentiating the above with respect to $\eta$ gives, \begin{equation}
523: \frac{\textrm{d}}{\textrm{d}\eta}\tilde{D}^{NS\left(z\right)}\left(\eta\right)=\frac{1}{4}\tilde{P}^{\left(z\right)}\tilde{D}^{NS\left(z\right)}\left(\eta\right).\label{eq:5.3.28}\end{equation}
524: Integrating this equation one finds a simple expression for the Mellin
525: moments of the structure function in terms of those of the splitting
526: function\begin{equation}
527: \tilde{D}^{NS\left(z\right)}\left(\eta\right)=\exp\left(\frac{1}{4}\eta\tilde{P}^{\left(z\right)}\right)\label{eq:5.3.29}\end{equation}
528: where we have used as an initial condition $\tilde{D}^{NS\left(z\right)}\left(m_{e}^{2}\right)=1$.
529: The Mellin moments for the splitting function $\tilde{P}^{\left(z\right)}$
530: are\begin{equation}
531: \tilde{P}^{\left(z\right)}=2\left(\frac{3}{4}+\log\epsilon\right)-\int_{0}^{1-\epsilon}\textrm{d}x\textrm{ }\frac{1-x^{z-1}}{1-x}+\frac{1-x^{z+1}}{1-x}-\frac{2}{1-x}\label{eq:5.3.30}\end{equation}
532: Using the Taylor expansion $\frac{x^{z}}{1-x}=\sum_{n=z}^{\infty}x^{n}$
533: the integral in \ref{eq:5.3.30} is trivial, as before we drop terms
534: vanishing in the limit $\epsilon\rightarrow0$. With a little manipulation
535: of the summations $\tilde{P}^{\left(z\right)}$ becomes,\begin{equation}
536: \begin{array}{rcl}
537: \tilde{P}^{\left(z\right)} & = & 2\left(\frac{3}{4}+\log\epsilon\right)+\int_{0}^{1-\epsilon}\textrm{d}x\textrm{ }x^{z-1}+x^{z}+\frac{2}{1-x}-2\sum_{n=0}^{z}x^{n}\\
538: & = & -\sum_{n=3}^{z+1}\textrm{ }\frac{1}{n}-\sum_{n=1}^{z-1}\textrm{ }\frac{1}{n}\end{array}.\label{eq:5.3.31}\end{equation}
539: It is worth pointing out that in dropping the terms vanishing in
540: the limit $\epsilon\rightarrow0$ the cut off $\epsilon$ has now
541: disappeared altogether from the calculation. The Euler function is
542: defined as\begin{equation}
543: \psi\left(x\right)=\frac{\textrm{d}}{\textrm{d}x}\log\Gamma\left(x\right)\label{eq:5.3.32}\end{equation}
544: hence,\begin{equation}
545: \psi\left(x+1\right)-\psi\left(x\right)=\frac{1}{x}\label{eq:5.3.33}\end{equation}
546: ($\Gamma\left(x\right)$ is the Gamma function). In terms of the
547: Euler function,\begin{equation}
548: \tilde{P}^{\left(z\right)}=-\psi\left(z+2\right)-\psi\left(z\right)+\psi\left(3\right)+\psi\left(1\right).\label{eq:5.3.34}\end{equation}
549: Using the identity \ref{eq:5.3.33} again and expanding $\Gamma\left(x\right)$
550: around $x=1$ we find\begin{equation}
551: \tilde{P}^{\left(z\right)}=-\psi\left(z+2\right)-\psi\left(z\right)+\frac{3}{2}-2\gamma.\label{eq:5.3.35}\end{equation}
552: Finally, substituting \ref{eq:5.3.35} into \ref{eq:5.3.29}, we
553: have an explicit expression for the structure function in terms of
554: its Mellin transform (\ref{eq:5.3.26})\begin{equation}
555: D^{NS}\left(x,\eta\right)=\frac{1}{2\pi i}\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}-\gamma\right)\right)\int_{c-i\infty}^{c+i\infty}\textrm{d}z\textrm{ }x^{-z}\exp\left(\frac{1}{4}\eta\left(-\psi\left(z+2\right)-\psi\left(z\right)\right)\right).\label{eq:5.3.37}\end{equation}
556: An analytic expression for the integral above is not known. The integral
557: may be performed in the \emph{soft} \emph{limit} $x\rightarrow1$,
558: this is known as the \emph{Gribov} \emph{approximation}. The factor
559: $x^{-z}$ shows that the integral is dominated by large values of
560: $z$. The Euler Gamma function is defined,\begin{equation}
561: \Gamma\left(z\right)=\left(z-1\right)!\label{eq:5.3.38}\end{equation}
562: in the large $z$ limit one can approximate the factorial operation
563: by Stirling's formula $n!\approx\sqrt{2\pi}z^{z+\frac{1}{2}}\exp\left(-z\right)$.
564: In the large $z$ limit we have $\psi\left(z\right)\approx\log z$\begin{equation}
565: \Rightarrow D^{NS}\left(x,\eta\right)=\frac{1}{2\pi i}\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}-\gamma\right)\right)\int_{c-i\infty}^{c+i\infty}\textrm{d}z\textrm{ }x^{-z}z^{-\frac{1}{2}\eta}.\label{eq:5.3.41}\end{equation}
566: By substituting $y=z\log x$,\begin{equation}
567: D^{NS}\left(x,\eta\right)=\frac{1}{2\pi i}\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}-\gamma\right)\right)\left(\log x\right)^{\frac{1}{2}\eta-1}\int_{c^{\prime}+i\infty}^{c^{\prime}-i\infty}\textrm{d}y\textrm{ }y^{-\frac{1}{2}\eta}\exp\left(-y\right).\label{eq:5.3.42}\end{equation}
568: After the transformation the limits have changed $c+i\infty\rightarrow c\log x+i\infty\log x=c^{\prime}-i\infty$.
569: Importantly $c^{\prime}$ is negative as $\log x<0$ and $c>z>0$
570: is required for the Mellin transform to exist. The integrand has two
571: obvious singularities, one at $y=0$ and another at $\textrm{Re}\left(y\right)\rightarrow-\infty$
572: $\left(\textrm{Re}\left(z\right)\rightarrow\infty\right)$. For $\left|y\right|\rightarrow\infty$
573: and $\textrm{Re}\left(y\right)\ne-\infty$ the integrand is zero,
574: this being the case we can \emph{close} the integration contour by
575: joining up the two ends at $c^{\prime}+i\infty$ and $c^{\prime}-i\infty$
576: such that it becomes a hemisphere of radius $\left|y\right|=\infty$
577: in the $\textrm{Re}\left(y\right)>c^{\prime}$ region of the complex
578: $y-$plane, as the integrand is zero along this addition to the contour.
579: Given the contour is closed in the complex plane we can deform it
580: as we please so long has we keep all poles inside it as its value
581: is given by $2\pi i$ times the residue at the pole, in this case
582: the only pole is at $y=0$. The \emph{Hankel} \emph{contour} ${\mathcal{{C_{H}}}}$
583: in the complex plane is an open contour which comes in along just
584: under the real axis from $+\infty$ goes around the origin and back
585: out to $-\infty$ just above the real axis. We can clearly deform
586: our contour to Hankel's contour as it contains the origin and no other
587: poles and because our integrand is zero at $\left|y\right|=\infty$
588: and $\textrm{Re}\left(y\right)\ne-\infty$ we can open the contour
589: again at $\left|y\right|=\infty$ making it exactly the Hankel contour.
590: The integral can then be brought into the form of Hankel's integral
591: representation of the Gamma function\begin{equation}
592: \Gamma\left(\frac{1}{2}\eta\right)^{-1}=-\frac{1}{2\pi i}\int_{{\mathcal{{C_{H}}}}}\textrm{d}y\textrm{ }\left(-y\right)^{-\frac{1}{2}\eta}\exp\left(-y\right).\label{eq:5.3.43}\end{equation}
593: Inserting $1=\left(-1\right)^{-\frac{1}{2}\eta}\left(-1\right)^{+\frac{1}{2}\eta-1}\left(-1\right)^{+1}$
594: and \ref{eq:5.3.43} into \ref{eq:5.3.42} gives\begin{equation}
595: D_{Gribov}^{NS}\left(x,\eta\right)=\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}-\gamma\right)\right)\frac{\left(1-x\right)^{\frac{1}{2}\eta-1}}{\Gamma\left(\frac{1}{2}\eta\right)}.\label{eq:5.3.44}\end{equation}
596: (we have used $\lim_{x\rightarrow1}-\log\left(x\right)=\lim_{x\rightarrow1}1-x$).
597:
598: The result of the Gribov approximation constitutes the perturbative
599: result derived earlier extended to \emph{all} \emph{orders} with the
600: caveat that it is only valid for the limit $x\rightarrow1$. To quote
601: Gribov and Lipatov \cite{Gribov:ri} the approximation above {}``can
602: be considered as a generalization of the Sudakov form factor''. The
603: perturbative structure function calculated earlier has no such constraint
604: on it but is nonetheless a finite order perturbative result. Ideally
605: one wants a single expression for the structure function which tends
606: to the Gribov result at in the soft limit and to the perturbative
607: result away from $x\rightarrow1$, the desired expression will interpolate
608: between the perturbative structure function and the Gribov approximation.
609: Such interpolation constitutes what is known as \emph{exponentiation}.
610: Integrating \ref{eq:5.3.44} over the soft phase space gives,\begin{equation}
611: \int_{1-\epsilon}^{1}\textrm{d}x\textrm{ }D_{Gribov}^{NS}\left(x,\eta\right)=\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}-\gamma\right)\right)\frac{\epsilon^{\frac{1}{2}\eta}}{\Gamma\left(1+\frac{1}{2}\eta\right)}.\label{eq:5.3.45}\end{equation}
612: We denote this $D_{Soft}^{NS}\left(\epsilon,\eta\right)$ and rewrite
613: the $\epsilon$ term $\exp\left(\frac{1}{2}\log\epsilon\right)$,\begin{equation}
614: D_{Soft}^{NS}\left(\epsilon,\eta\right)=\frac{\exp\left(-\frac{1}{2}\eta\gamma\right)}{\Gamma\left(1+\frac{1}{2}\eta\right)}\exp\left(\frac{1}{2}\eta\left(\frac{3}{4}+\log\epsilon\right)\right).\label{eq:5.3.46}\end{equation}
615: The soft part of the structure function $D_{Soft}^{NS}\left(\epsilon,\eta\right)$
616: gives an expansion in powers of $\eta$ with coefficients exactly
617: the same as the coefficients of $\eta$ in the delta term obtained
618: earlier by iterating the Lipatov equation, so justifying the earlier
619: quote of Gribov and Lipatov. We have checked this explicitly to ${\mathcal{{O}}}\left(\beta^{3}\right)$.
620: Had we calculated the ${\mathcal{{O}}}\left(\beta^{4}\right)$ delta
621: term with Lipatov's equation we would see its coefficient is the same
622: as the coefficient of the ${\mathcal{{O}}}\left(\beta^{4}\right)$
623: term in the expansion of $D_{Soft}^{NS}\left(\epsilon,\eta\right)$.
624: In the perturbative solution the delta terms give the structure function
625: in the $1-\epsilon<x\leq1$ region. If we were to integrate the delta
626: terms over $1-\epsilon<x\leq1$ and expand in $\eta$ we would get
627: the same result as we would just expanding $D_{Soft}^{NS}\left(\epsilon,\eta\right)$
628: up to the same power in $\eta$. Denoting the delta term in the perturbative
629: expansion calculated to ${\mathcal{{O}}}\left(\eta^{i}\right)$ as
630: $\delta\left(1-x\right)\Delta^{\left(i\right)}\left(\epsilon,\eta\right)$
631: we have,\begin{equation}
632: \int_{1-\epsilon}^{1}\textrm{d}x\textrm{ }\delta\left(1-x\right)\Delta^{\left(i\right)}\left(\epsilon,\eta\right)-\int_{1-\epsilon}^{1}\textrm{d}x\textrm{ }D_{Gribov}^{NS}\left(x,\eta\right)\sim{\mathcal{{O}}}\left(\eta^{i+1}\right)\label{eq:5.3.48}\end{equation}
633: this suggests that we modify the structure function by hand to include
634: higher order soft effects by replacing the delta term altogether with
635: $D_{Gribov}^{NS}\left(x,\eta\right)$,\begin{equation}
636: \delta\left(1-x\right)\Delta^{\left(i\right)}\left(\epsilon,\eta\right)\rightarrow D_{Gribov}^{NS}\left(x,\eta\right).\label{eq:5.3.49}\end{equation}
637: How does this affect the theta term \emph{etc}? With this replacement
638: we can safely take the already implied limit $\epsilon\rightarrow0$,
639: the theta function for the theta term is then replaced by $1$. Then
640: the most obvious thing to do is to demand that, up to the same order
641: in $\eta$, the new structure function is the same as the old purely
642: perturbative one was in the region $0<x<1-\epsilon$ and not care
643: about the terms that are higher order in $\eta$. Expanding in $\eta$
644: we have\begin{equation}
645: \begin{array}{rcl}
646: D_{Gribov}^{NS}\left(x,\eta\right) & = & \frac{2}{1-x}\left(\frac{\eta}{4}\right)^{1}\\
647: & + & \frac{3+4\log\left(1-x\right)}{1-x}\left(\frac{\eta}{4}\right)^{2}\\
648: & + & \frac{\frac{9}{4}-\frac{3}{4}\pi^{2}+6\log\left(1-x\right)+4\log^{2}\left(1-x\right)}{1-x}\left(\frac{\eta}{4}\right)^{3}\\
649: & + & {\mathcal{{O}}}\left(\eta^{4}\right)\end{array}.\label{eq:5.3.49b}\end{equation}
650: Recall that the Gribov solution is valid in the large $x$ limit.
651: The perturbative solution obtained order by order in $\eta$ by iterating
652: solutions through Lipatov's equation requires no such approximation.
653: Consequently one expects that the perturbative solution and the Gribov
654: approximation should agree in the limit $x\rightarrow1$. This is
655: indeed the case, it is easy to verify that the $x\rightarrow1$ limit
656: of the theta terms obtained in the perturbative case are equal to
657: those obtained in the non-perturbative case. If we denote the $ith$
658: order in $\eta$ of a quantity by appending superscript $\left(i\right)$
659: to it we can define an improved structure function\begin{equation}
660: D^{\prime NS\left(i\right)}\left(x,\eta\right)=\left\{ \begin{array}{ll}
661: D_{Gribov}^{NS\left(i\right)}\left(x,\eta\right)+\left(\left.D^{NS\left(i\right)}\left(x,\eta\right)\right|_{\theta-term}-D_{Gribov}^{NS\left(i\right)}\left(x,\eta\right)\right) & 0<i\leq3\\
662: D_{Gribov}^{NS\left(i\right)}\left(x,\eta\right) & i>3\end{array}\right.\label{eq:5.3.50}\end{equation}
663: \emph{i.e.}\begin{equation}
664: D^{\prime NS}\left(x,\eta\right)=D_{Gribov}^{NS}\left(x,\eta\right)+\sum_{i=0}^{3}\left(\left.D^{NS\left(i\right)}\left(x,\eta\right)\right|_{\theta-term}-D_{Gribov}^{NS\left(i\right)}\left(x,\eta\right)\right).\label{eq:5.3.50b}\end{equation}
665: This is the exponentiation prescription of Kuraev and Fadin \cite{Kuraev:hb}.
666: The condition on $i$ relates to the order to which the perturbative
667: solution was found, in our case $3$. Applying prescription \ref{eq:5.3.50}
668: to the perturbative theta terms obtained so far is trivial albeit
669: tedious subtraction, we find agreement with the results in \cite{Skrzypek:1990qs,Skrzypek:1992vk}.
670: It is of note that the integral of this function over the range $1-\epsilon\rightarrow\epsilon$
671: is finite in the limit $\epsilon\rightarrow0$. The integral is finite
672: because, as stated above, in the divergent $x\rightarrow1$ limit
673: the theta terms are equal to the corresponding Gribov terms inside
674: the summation, in addition the integral over the all orders Gribov
675: approximation to the left of the sum is finite (see equation \ref{eq:5.3.46}).
676:
677: An alternative exponentiation prescription is that of Jadach and Ward
678: called YFS exponentiation. Here the only difference is that the Gribov
679: term is extracted from the iterative solution as a \emph{factor} rather
680: than subtracted as in the Kuraev Fadin exponentiation \emph{viz}\begin{equation}
681: D^{\prime NS\left(i\right)}\left(x,\eta\right)=\left\{ \begin{array}{ll}
682: \sum_{j+k=i}D_{Gribov}^{NS\left(j\right)}\left(x,\eta\right)\Delta^{\left(k\right)}\left(x,\eta\right) & 0<i\leq3\\
683: D_{Gribov}^{NS\left(i\right)}\left(x,\eta\right)\Delta^{\left(k\right)}\left(x,\eta\right) & i>3\end{array}\right.,\label{eq:5.3.51}\end{equation}
684: where $\Delta^{\left(k\right)}\left(x,\eta\right)$ is defined by\begin{equation}
685: \left.D^{NS\left(i\right)}\left(x,\eta\right)\right|_{\theta-term}=\sum_{j+k=i}D_{Gribov}^{NS\left(j\right)}\left(x,\eta\right)\Delta^{\left(k\right)}\left(x,\eta\right)\textrm{ }\left(1\leq j+k\leq3,\textrm{ }0\leq k\leq2\right).\label{eq:5.3.52}\end{equation}
686: This is a system of linear equations that is easily solved for the
687: $\Delta^{\left(k\right)}\left(x,\eta\right)$,\begin{equation}
688: \begin{array}{rcl}
689: \Delta^{\left(0\right)}\left(x,\eta\right) & = & \frac{\left.D^{NS\left(1\right)}\left(x,\eta\right)\right|_{\theta-term}}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}\\
690: \Delta^{\left(1\right)}\left(x,\eta\right) & = & \frac{\left.D^{NS\left(2\right)}\left(x,\eta\right)\right|_{\theta-term}}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}-\frac{D_{Gribov}^{NS\left(2\right)}\left(x,\eta\right)}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}\Delta^{\left(0\right)}\left(x,\eta\right)\\
691: \Delta^{\left(2\right)}\left(x,\eta\right) & = & \frac{\left.D^{NS\left(3\right)}\left(x,\eta\right)\right|_{\theta-term}}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}-\frac{D_{Gribov}^{NS\left(3\right)}\left(x,\eta\right)}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}\Delta^{\left(0\right)}\left(x,\eta\right)-\frac{D_{Gribov}^{NS\left(2\right)}\left(x,\eta\right)}{D_{Gribov}^{NS\left(1\right)}\left(x,\eta\right)}\Delta^{\left(1\right)}\left(x,\eta\right)\end{array}\label{eq:5.3.53}\end{equation}
692: giving\begin{equation}
693: \begin{array}{rl}
694: & D^{\prime NS}\left(x,\eta\right)\\
695: = & D_{Gribov}^{NS}\left(x,\eta\right)\times\left(\frac{1}{2}\left(1+x\right)^{2}+\frac{1}{16}\left(-2\left(1-x\right)^{2}-\left(1+3x^{2}\right)\log x\right)\eta\right.\\
696: + & \left.\frac{1}{384}\left(12\left(1-x\right)^{2}+\left(6\left(1-4x+3x^{2}\right)+\left(1+7x^{2}\right)\log x\right)\log x-12\left(1-x\right)^{2}\textrm{Li}_{2}\left(1-x\right)\right)\eta^{2}\right)\end{array}.\label{eq:5.3.54}\end{equation}
697: This is the result given in \cite{Skrzypek:1990qs,Skrzypek:1992vk}.
698:
699: To summarize, we have re-derived in this section again the structure
700: functions of Kuraev and Fadin \ref{eq:5.3.50} and Jadach and Skrzypek.
701: At no point is the underlying hard scattering process referred to,
702: with this approach folding the radiator function with New Physics
703: is just as valid as folding it with the standard model. In addition,
704: within the context of Lipatov's equation (\emph{i.e.} for the case
705: of near collinear emission of radiation) the calculations suggest
706: that the factorization of the corrections holds at the level of the
707: differential cross section not just the total cross section. We are
708: seeing this effect in the flatness of the radiator functions we have
709: derived in figure \ref{cap:nointf}. In fact the third order Kuraev-Fadin
710: structure function \ref{eq:5.3.50} is folded with itself \ref{eq:5.2.5}
711: and present within \texttt{ZFITTER} as the default setting for the
712: flag higher \texttt{FOT2} which governs the implementation of higher
713: order radiative corrections. Crucially at this level of approximation
714: the hard scattering interaction is not referred to, this is clearly
715: good news for fits to new physics. Approximations are made in the
716: leading log approximations \emph{e.g.} the finite order perturbative
717: result nevertheless they have been found to agree very well (KF better
718: than 0.2\%, YFS better than \textasciitilde{}0.01\%) with \emph{exact}
719: numerical solutions of the Lipatov equation. This would seem to vindicate
720: the use of ISR corrections, the numerical radiator function, derived
721: from \texttt{ZFITTER} and other such packages.
722:
723:
724: \section{The Breakdown of the Simple LLA Structure Function Approach.}
725:
726: In general the standard model predictions obtained from \texttt{ZFITTER}
727: and the other standard model difermion packages allow many more sophisticated
728: radiative corrections to be applied to the process in question besides
729: just ISR. In particular initial state-final state interference (ISR{*}FSR)
730: effects corresponding to interference between diagrams with bremsstrahlung
731: emitted from initial state particles and diagrams with bremsstrahlung
732: emitted from final state particles, as well as box diagrams resulting
733: from emission and re-absorption of photons, have a significant (\textasciitilde{}10\%)
734: effect at the level of the differential cross section despite having
735: only a negligible (<1\%) effect on the total cross section. The radiator
736: functions derived from \texttt{ZFITTER} with these corrections included
737: are shown in figure \ref{cap:interference} where they have acquired
738: a significant angular dependence that was not seen in figure \ref{cap:nointf}.
739:
740: %
741: \begin{figure}
742: \begin{center}\includegraphics[%
743: width=0.30\paperwidth,
744: height=0.35\paperwidth]{200MMrad_intf.eps}\end{center}
745:
746:
747: \caption{\label{cap:interference}The plot above shows the radiator functions
748: $\tilde{R}_{1}\left(s,s^{\prime}\right)$ with $s=\left(200\textrm{ GeV}\right)^{2}$,
749: obtained from the \texttt{ZFITTER} package by extracting it separately
750: for each angular bin just as in figure \ref{cap:nointf} but this
751: time including initial state-final state interference effects.}
752: \end{figure}
753:
754:
755: This angular dependence would seem to ruin the argument put forward
756: in \ref{sec:Radiator-Functions:-A}, which says that these corrections
757: should be independent of the polar angle if in fact they are independent
758: of the hard physics process. It also highlights the weak point in
759: the leading log approximation, the leading log approximation starts
760: to break down for exclusive observables \emph{i.e.} highly differential
761: quantities and measurements obtained with strong cuts \cite{Skrzypek:1990qs}.
762: ISR{*}FSR interference corrections clearly fall outside the mandate
763: of the structure function / leading log approach which describes ISR
764: effects so well.
765:
766: In fact the cross section obtained with angular cuts requires \emph{two}
767: radiator functions. The $\cos\theta$ symmetric and $\cos\theta$
768: asymmetric parts of the tree level cross section are convoluted separately
769: with different radiator functions and the result added \cite{Bardin:1989cw,Bardin:1990fu,Bardin:1999yd}.
770: The cross section in an angular bin $i$ with edges at $c=\cos\theta=c_{i},c_{i+1}$
771: $\sigma_{i}\left(s,c_{i},c_{i+1}\right)$ is\begin{equation}
772: \sigma\left(s,c_{i},c_{i+1}\right)=\int_{c_{i}}^{c_{i+1}}\textrm{d}c\textrm{ }\frac{{\rm {d}}\sigma}{{\rm {d}}c}.\label{eq:5.3.55}\end{equation}
773: Defining the symmetric and asymmetric cross sections $\sigma_{T}\left(s,-c_{i},c_{i}\right)$
774: and $\sigma_{FB}\left(s,-c_{i},c_{i}\right)$ respectively in the
775: region $-c_{i}<c\leq c_{i}$ as\begin{equation}
776: \begin{array}{rcl}
777: \sigma_{T}\left(s,-c_{i},c_{i}\right) & = & \frac{1}{2}\left[\int_{0}^{c_{i}}\textrm{d}c\textrm{ }\frac{{\rm {d}}\sigma}{{\rm {d}}c}+\int_{-c_{i}}^{0}\textrm{d}c\textrm{ }\frac{{\rm {d}}\sigma}{{\rm {d}}c}\right]\\
778: \sigma_{FB}\left(s,-c_{i},c_{i}\right) & = & \frac{1}{2}\left[\int_{0}^{c_{i}}\textrm{d}c\textrm{ }\frac{{\rm {d}}\sigma}{{\rm {d}}}-\int_{-c_{i}}^{0}\textrm{d}c\textrm{ }\frac{{\rm {d}}\sigma}{{\rm {d}}c}\right]\end{array},\label{eq:5.3.56}\end{equation}
779: the cross section in the angular bin is\begin{equation}
780: \sigma\left(s,c_{i},c_{i+1}\right)=\sigma_{T}\left(s,-c_{i+1},c_{i+1}\right)+\sigma_{FB}\left(s,-c_{i+1},c_{i+1}\right)-\sigma_{T}\left(s,-c_{i},c_{i}\right)-\sigma_{FB}\left(s,-c_{i},c_{i}\right).\label{eq:5.3.57}\end{equation}
781: In terms of a cross section corrected for radiative effects we have\begin{equation}
782: \begin{array}{rcl}
783: \sigma_{T}\left(s,-c_{i},c_{i}\right) & = & \int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }R_{T}\left(s,s^{\prime},c_{i}\right)\sigma_{Tree}\left(s^{\prime}\right)\\
784: \sigma_{FB}\left(s,-c_{i},c_{i}\right) & = & \int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }R_{FB}\left(s,s^{\prime},c_{i}\right)\sigma_{Tree}\left(s^{\prime}\right)\end{array}.\label{eq:5.3.58}\end{equation}
785:
786:
787: So far we have assumed \begin{equation}
788: R_{T}\left(s,s^{\prime},c_{i}\right)=R_{FB}\left(s,s^{\prime},c_{i}\right),\label{eq:5.3.59}\end{equation}
789: and that to good approximation (\ref{eq:5.2.8}),\begin{equation}
790: \sigma\left(s,c_{i},c_{i+1}\right)=\int_{s_{min}^{\prime}}^{s}\textrm{d}s^{\prime}\textrm{ }\tilde{R}_{1}\left(s,s^{\prime}\right)\sigma_{Tree}\left(s^{\prime},c_{i},c_{i+1}\right)\label{eq:5.3.60}\end{equation}
791: where, recalling section \ref{sec:Radiator-Functions:-A}, $\tilde{R}\left(s,s^{\prime}\right)$
792: is a universally applicable radiator function. According to Bardin
793: \emph{et al} \cite{Bardin:1990fu} {}``Strictly speaking an ansatz
794: like \ref{eq:5.3.59} is wrong.'' In \cite{Bardin:1990fu} the authors
795: go on to say that near the Z peak the ansatz \ref{eq:5.3.59} nonetheless
796: gives {}``excellent agreement with the correct result''. The authors
797: say that this effect is due to the dominant corrections being from
798: soft photons. They go on to say that if an $s^{\prime}$ cut is in
799: use the agreement is even better. We appeal to this soft photon dominance
800: ansatz due to the restrictive $s^{\prime}$ cuts used in obtaining
801: samples to fit with, $\sqrt{s^{\prime}}\geq0.85\sqrt{s}$, the degree
802: to which it holds is born out by the flatness in $\cos\theta$ of
803: the corrections we derived, $\tilde{R}\left(s,s^{\prime}\right)$
804: in figure \ref{cap:nointf}. If the corrections were significantly
805: different for the $\cos\theta$ symmetric and asymmetric parts of
806: the differential cross section $\tilde{R}\left(s,s^{\prime}\right)$would
807: not be flat in $\cos\theta$ as we essentially are just dividing the
808: corrected cross section in each angular bin by the born cross section
809: in each angular bin, bin by bin in $s^{\prime}$, the born cross section
810: could not possibly \emph{factor out}. Prior to considering initial
811: state-final state interference effects there seems to be nothing wrong
812: in using the same radiator function for $c$ even and $c$ odd parts
813: of the cross section, the structure functions we derived describe
814: the radiator function very well.
815:
816: As you can see in figure \ref{cap:interference} the addition of initial
817: state-final state interference effects seem to ruin all our previous
818: hypotheses. Ideally one would hope that just because the radiative
819: corrections have acquired this angular dependence they are still nonetheless
820: essentially universal like the other (ISR) corrections or at least
821: largely process independent in our phase space. Given that we are
822: dealing with box diagrams such factorization seems like wishful thinking.
823:
824:
825: \section{ISR{*}FSR Universality.}
826:
827: In this section we describe how the ISR{*}FSR interference corrections,
828: despite giving the naive radiator function an important angular dependence,
829: is nonetheless mostly a universal correction. The ISR{*}FSR interference
830: corrections for some underlying hard scattering involving the exchange
831: of an unknown particle (???) may be represented at ${\mathcal{{O}}}\left(\alpha^{3}\right)$
832: by the Feynman diagrams in figures \ref{cap:ISR*FSR-brem} and \ref{cap:ISR*FSR-box}.%
833: \begin{figure}
834: \begin{center}\includegraphics[%
835: width=0.30\textwidth,
836: height=0.20\textwidth]{bremf+.eps}\includegraphics[%
837: clip,
838: width=0.9cm,
839: height=0.20\textwidth]{redcross.eps}\includegraphics[%
840: width=0.30\textwidth,
841: height=0.20\textwidth]{breme-.eps}\end{center}
842:
843:
844: \caption{\label{cap:ISR*FSR-brem}ISR{*}FSR terms in the cross section originate
845: from diagrams with photons emitted from the initial legs and final
846: state legs interfering. }
847: \end{figure}
848:
849:
850: %
851: \begin{figure}
852: \begin{center}\includegraphics[%
853: width=0.30\textwidth,
854: height=0.20\textwidth]{brembox.eps}\includegraphics[%
855: clip,
856: width=0.9cm,
857: height=0.20\textwidth]{redcross.eps}\includegraphics[%
858: width=0.30\textwidth,
859: height=0.20\textwidth]{bremborn.eps}\end{center}
860:
861:
862: \caption{\label{cap:ISR*FSR-box}In addition to the bremsstrahlung diagrams
863: in figure \ref{cap:ISR*FSR-brem}, at the same order in $\alpha$,
864: ISR{*}FSR terms in the cross section also originate from diagrams
865: with photons box diagrams interfering with the lowest order diagram. }
866: \end{figure}
867:
868:
869: In subsections \ref{sub:Oa3brem} and \ref{sub:Oa3Virtual} we will
870: use the Feynman rules of \cite{Peskin:ev} and the kinematic notations
871: of appendix A. In these subsections we will try to maintain an explicit
872: notation. The matrix elements of the two fermion events under discussion
873: at this order have four fermions and hence four spinors in them, $\bar{\nu}_{e}\left(p_{+}\right),\textrm{ }u_{e}\left(p_{-}\right)$
874: represent the incoming $e^{+}$ and $e^{-}$, $\nu_{f}\left(q_{+}\right),\textrm{ }\bar{u}_{f}\left(q_{-}\right)$
875: represent the outgoing antifermion and fermion respectively. Given
876: that we are trying to discuss radiative corrections in such a way
877: as to find and highlight any universal features of them it will therefore
878: be useful to denote the born level process generically as\begin{equation}
879: {\mathcal{{A}}}_{Born}=\bar{\nu}_{e}\left(p_{+}\right)...Hard...\nu_{f}\left(q_{+}\right)\label{eq:5.5.1}\end{equation}
880: where $...Hard...$ is defined to contain the electron and fermion
881: spinors and more importantly, the propagator of the exchanged particle
882: and its couplings to the incoming and outgoing particles (see also
883: equation \ref{eq:5.5.1}). For example in the case of QED we have
884: simply\begin{equation}
885: ...Hard...=iQ_{e}e\gamma^{\mu}u_{e}\left(p_{-}\right)\frac{-ig_{\mu\nu}}{\left(p_{-}+p_{+}\right)^{2}}\bar{u}_{f}\left(q_{-}\right)iQ_{f}e\gamma^{\nu}.\label{eq:5.5.1b}\end{equation}
886:
887:
888:
889: \subsection{${\mathcal{{O}}}\left(\alpha^{3}\right)$ Soft Bremsstrahlung\label{sub:Oa3brem}.}
890:
891: The contribution of soft bremsstrahlung corrections to 2-fermion processes
892: are greatly important. By soft bremsstrahlung we mean bremsstrahlung
893: which carries off an amount of energy smaller than the energy resolution
894: of the detector, soft photons carry off a negligible amount of energy.
895: Given that the detector cannot therefore tell the difference between
896: a genuine ${\mathcal{{O}}}\left(\alpha\right)$ process and an ${\mathcal{{O}}}\left(\alpha^{3}\right)$
897: soft bremsstrahlung correction to it, we must add such contributions
898: to our predicted cross section. The amplitude for a \emph{general
899: difermion process} with bremsstrahlung emitted from the external \emph{positron}
900: leg of the diagram (figure \ref{cap:Some brem diags}) will be of
901: the form%
902: \begin{figure}
903: \begin{center}\includegraphics[%
904: width=0.30\textwidth,
905: height=0.20\textwidth]{se+sbrem.eps}\includegraphics[%
906: width=0.30\textwidth,
907: height=0.20\textwidth]{te+sbrem.eps}\end{center}
908:
909:
910: \caption{\label{cap:Some brem diags}Some ${\mathcal{{O}}}\left(\alpha^{3}\right)$
911: bremsstrahlung correction diagrams.}
912: \end{figure}
913:
914:
915: \noindent \textcolor{black}{\begin{equation}
916: {\mathcal{{A}}}_{Brem}=\epsilon_{\lambda}^{*}\left(k\right)\bar{\nu}_{e}\left(p_{+}\right)\left\{ iQ_{e}e\gamma^{\lambda}\frac{i\left(-\left(\not p_{+}-\not k\right)+m_{e}\right)}{\left(p_{+}-k\right)^{2}-m_{e}^{2}}\right\} ...Hard...\nu_{f}\left(q_{+}\right),\label{eq:5.5.1}\end{equation}
917: with $Q_{e}=-1$ for an electron (generally $Q_{f}$ is the charge
918: on the fermion/antifermion $f$ in units of $e$). If we now work
919: in the limit that $\left|k\right|\rightarrow0$, $s^{\prime}\rightarrow s$,
920: the} \textcolor{black}{\emph{soft photon approximation}}\textcolor{black}{,
921: we see that\begin{equation}
922: {\mathcal{{A}}}_{Soft}=-Q_{e}e\epsilon_{\lambda}^{*}\left(k\right)\bar{\nu}_{e}\left(p_{+}\right)\gamma^{\lambda}\frac{\not p_{+}-m}{2k.p_{+}}...Hard...\nu_{f}\left(q_{+}\right).\label{eq:5.5.2}\end{equation}
923: Simple $\gamma$ matrix manipulation and the application of the Dirac
924: equation to \ref{eq:5.5.2} gives\begin{equation}
925: \begin{array}{rcl}
926: {\mathcal{{A}}}_{soft} & = & -\frac{1}{2k.p_{+}}Q_{e}e\epsilon_{\lambda}^{*}\left(k\right)\bar{\nu}_{e}\left(p_{+}\right)\left(2p_{+}^{\lambda}\right)...Hard...\nu_{f}\left(q_{+}\right)\\
927: & = & -Q_{e}e\frac{\epsilon^{*}.p_{+}}{k.p_{+}}{\mathcal{{A}}}_{born}\end{array}.\label{eq:5.5.5}\end{equation}
928: From this we see that in the soft photon approximation the bremsstrahlung
929: corrections to the born level process are independent of what that
930: born process is, we have extracted the bremsstrahlung correction as
931: an over all constant factor while staying blissfully ignorant of the
932: underlying process. Hopefully it is clear that these tricks can be
933: employed with the other external legs in considering bremsstrahlung
934: emitted from them in the soft photon approximation giving for bremsstrahlung
935: emitted from the $i$th external leg of the diagram\begin{equation}
936: {\mathcal{{A}}}_{Soft}^{\left(i\right)}=-Q_{i}e\frac{\epsilon^{*}.p_{i}}{k.p_{i}}{\mathcal{{A}}}_{Born}\label{eq:5.5.6}\end{equation}
937: where $p_{i}$ denotes the momentum of the external leg. Consequently,
938: taking into account the bremsstrahlung emitted from all different
939: legs we find the matrix element\begin{equation}
940: {\mathcal{{A}}}_{Soft}^{\dagger}{\mathcal{{A}}}_{Soft}=\sum_{i,j}Q_{i}Q_{j}e^{2}\frac{\left(\epsilon.p_{i}\right)\left(\epsilon^{*}.p_{j}\right)}{\left(k.p_{i}\right)\left(k.p_{j}\right)}{\mathcal{{A}}}_{Born}^{\dagger}{\mathcal{{A}}}_{Born}.\label{eq:5.5.7}\end{equation}
941: Summing fermion spins and polarizations of the photon $\left(\sum_{pols}\epsilon^{\mu}\epsilon^{*\nu}=-g^{\mu\nu}\right)$
942: and inserting the phase space and flux factors factors gives in the
943: $k\approx0$} \textcolor{black}{\emph{soft photon approximation}}\textcolor{black}{\begin{equation}
944: \textrm{d}\sigma=-\frac{\alpha}{4\pi^{2}}\sum_{i,j}Q_{i}Q_{j}p_{i}.p_{j}\frac{{\rm {d}}^{3}k}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\textrm{d}\sigma_{Born}.\label{eq:5.5.11}\end{equation}
945: The soft photon cross section is the contribution to the cross section
946: from soft photons where soft photons were defined earlier as being
947: those low enough in energy to render the bremsstrahlung process indistinguishable
948: from the tree level process, we denote this energy $\omega$, the} \textcolor{black}{\emph{soft
949: photon cut-off}}\textcolor{black}{. Integrating over the final state
950: phase space of the soft photon we have\begin{equation}
951: \textrm{d}\sigma_{Soft}=-\frac{\alpha}{\pi}\sum_{i,j}Q_{i}Q_{j}p_{i}.p_{j}\left(\frac{1}{4\pi}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\right)\textrm{d}\sigma_{Born}.\label{eq:5.5.12}\end{equation}
952: Note that this differential cross section contains contributions
953: from pairs of diagrams which constitute initial state-final state
954: interference. This would appear to be the desired result, the radiative
955: corrections are factorized from the hard scattering process at the
956: level of the differential cross section. }
957:
958: On its own the result \ref{eq:5.5.12} arguably does not make much
959: sense as the phase space integral\begin{equation}
960: \frac{1}{4\pi}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\label{eq:5.5.13}\end{equation}
961: is infrared divergent, as $k\rightarrow0$ $\textrm{d}\sigma\rightarrow\infty$.
962: The infrared divergence of the ISR{*}FSR bremsstrahlung contribution
963: actually cancels an infrared divergence from the ISR{*}FSR box diagrams.
964: The phase space integral must be expressed in a form whereby we can
965: add the bremsstrahlung and box contributions to cancel the divergences
966: before we can fully assess the universality of the ISR{*}FSR interference.
967: We have treated the divergence with dimensional regularization. We
968: replace all four dimensional dependence with the usual prescription
969: $4\rightarrow n=4+\epsilon^{\prime}$
970:
971: \begin{equation}
972: \frac{1}{4\pi}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\rightarrow\frac{1}{4\pi}\mu^{N}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{n-1}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}.\label{eq:5.5.14}\end{equation}
973: The factor $\mu^{N}$ is introduced to keep the overall mass dimensions
974: of the integral the same as they were before extending to $N$ dimensions.
975: $\mu^{N}$ has mass dimensions $\left[\mu^{N}\right]=N$ \emph{i.e.}
976: $N=-\epsilon^{\prime}$ \emph{i.e.} \ref{eq:5.5.14} is\begin{equation}
977: \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}.\label{eq:5.5.15}\end{equation}
978: This photon phase space integral is technical and not so illuminating
979: for our present discussion so we shall confine it to appendix B. After
980: completing the integral the contribution to the differential cross
981: section from ${\mathcal{{O}}}\left(\alpha^{3}\right)$ soft bremsstrahlung
982: diagrams is found to be\begin{equation}
983: \textrm{d}\sigma_{Soft,Int}=\frac{2\alpha}{\pi}\left[\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{4\omega^{2}}{\mu^{2}}\right)\right)\log\left(\frac{t}{u}\right)-\textrm{Li}_{2}\left(1+\frac{s}{t}\right)+\textrm{Li}_{2}\left(1+\frac{s}{u}\right)\right]\textrm{d}\sigma_{Born}.\label{eq:5.5.45}\end{equation}
984:
985:
986: \textcolor{black}{This is the} \textcolor{black}{\emph{universal}} \textcolor{black}{soft
987: photon bremsstrahlung correction to the differential cross section
988: from ISR{*}FSR interference. Left as it is above, the correction to
989: the differential cross section is infrared divergent, there are terms
990: ${\mathcal{{O}}}\left(\epsilon^{\prime-1}\right)$ and the limit $\epsilon^{\prime}\rightarrow0$
991: must be taken to return to four dimensional Minkowski space. For our
992: purposes the essential feature of this result is that the radiative
993: correction is factorized from / independent of the new physics exchange
994: process. }
995:
996:
997: \subsection{${\mathcal{{O}}}\left(\alpha^{3}\right)$ Virtual Diagrams\label{sub:Oa3Virtual}.}
998:
999: \textcolor{black}{The infrared divergences $\left(\frac{1}{\hat{\epsilon}}\right)$
1000: in the soft bremsstrahlung result are well known to be } \textcolor{black}{\emph{physical}}\textcolor{black}{,
1001: that is to say that they are a real part of the theory as opposed
1002: to the ultraviolet divergences encountered in renormalization which
1003: are a consequence of our ignorance of short distance physics, likewise
1004: the infrared divergences cannot be subtracted} \textcolor{black}{\emph{ad
1005: hoc}}\textcolor{black}{. The infrared divergences due to the emission
1006: of soft photons do however cancel against the corresponding virtual
1007: diagrams, it is a well known fact that the soft photon contribution
1008: to the cross section from initial state bremsstrahlung cancels against
1009: a corresponding infrared divergence from the corresponding vertex
1010: diagram which involves emitting a photon from one initial state leg
1011: and absorbing it on the other. Likewise the initial state-final state
1012: interference cross section arising from the interference of initial
1013: and final state bremsstrahlung diagrams does not make much sense on
1014: its own, it must be added to the corresponding virtual graph. The
1015: virtual graph for ISR{*}FSR bremsstrahlung is by analogy to the case
1016: of initial state bremsstrahlung the one formed by taking the photon
1017: emitted from an initial state fermion and joining it to a final state
1018: fermion. This gives the two diagrams of figure \ref{cap:box diagrams}.}
1019:
1020: \textcolor{black}{}%
1021: \begin{figure}
1022: \begin{center}\textcolor{black}{\includegraphics[%
1023: width=0.49\textwidth,
1024: height=0.40\textwidth]{newdrboxa.eps}\includegraphics[%
1025: width=0.49\textwidth,
1026: height=0.40\textwidth]{newcrboxa.eps}}\end{center}
1027:
1028: \begin{center}\textcolor{black}{\includegraphics[%
1029: width=0.49\textwidth,
1030: height=0.40\textwidth]{newdrboxb.eps}\includegraphics[%
1031: width=0.49\textwidth,
1032: height=0.40\textwidth]{newcrboxb.eps}}\end{center}
1033:
1034:
1035: \caption{\textcolor{black}{\label{cap:box diagrams}The box diagrams contributing
1036: to the ISR{*}FSR interference cross section. On the left hand side
1037: we have} \textcolor{black}{\emph{direct box diagrams}}\textcolor{black}{,
1038: from top to bottom these are represented by amplitudes ${\mathcal{{A}}}_{D1}$
1039: and ${\mathcal{{A}}}_{D2}$ respectively. On the right hand side we
1040: have} \textcolor{black}{\emph{crossed box diagrams}}\textcolor{black}{,
1041: from top to bottom these are represented by amplitudes ${\mathcal{{A}}}_{C1}$
1042: and ${\mathcal{{A}}}_{C2}$ respectively. }}
1043: \end{figure}
1044:
1045:
1046: \noindent \textcolor{black}{At ${\mathcal{{O}}}\left(\alpha^{3}\right)$
1047: these diagrams only contribute by their interference with the tree
1048: level (hard) process. Unlike the soft bremsstrahlung corrections the
1049: ISR{*}FSR box diagrams do not simply represent multiplicative corrections
1050: to the born level amplitude. These diagrams were studied in the the
1051: hope that, though they do not fall under the banner of universal radiative
1052: corrections, they may have some component which is universal. It is
1053: quite easy to show that the box diagrams have universal, factorizable
1054: corrections. The tricks previously applied to the soft bremsstrahlung
1055: diagrams also work, to some degree, with box diagrams (provided there
1056: is an internal photon). Let us write down the amplitude for the first
1057: direct box diagram in figure \ref{cap:box diagrams}, were we feign
1058: ignorance of the couplings and propagator of the $X^{0}$ boson}%
1059: \footnote{\textcolor{black}{For now let us assume that $X^{0}$ has the generic
1060: $\frac{1}{k^{2}}$ high energy boson behavior} \textcolor{black}{\emph{i.e.}} \textcolor{black}{the
1061: diagram is UV finite. We discuss this point more later. }%
1062: }\textcolor{black}{. We denote the couplings associated of $X^{0}$
1063: some particle $A$ as ${\mathcal{{C}}}_{A}$ and propagator contributions
1064: to the numerator as ${\mathcal{{P}}}_{X^{0}}$. We assume the denominator
1065: of the propagator is the usual $\left(k^{2}-m_{X}^{2}\right)$ form
1066: (for an $X^{0}$ particle four momentum $k$ and mass $m_{X}$). The
1067: couplings can be considered to have Dirac and Lorentz indices, the
1068: Dirac indices are contracted with the rest of the Dirac algebra in
1069: the numerator and the Lorentz indices are contracted with those in
1070: ${\mathcal{{P}}}_{X^{0}}$. With these notations ${\mathcal{{P}}}_{X^{0}}...{\mathcal{{C}}}_{e}u\left(p_{-}\right)\bar{u}\left(p_{+}\right){\mathcal{{C}}}_{f}.../\left(s^{2}-m_{X}^{2}\right)$
1071: corresponds to $...Hard...$ in the case of the soft bremsstrahlung.
1072: Finally we shall assume the interaction is charge conjugation invariant} \textcolor{black}{\emph{i.e.}} \textcolor{black}{${\mathcal{{C}}}_{A}=-{\mathcal{{C}}}_{\bar{A}}$.
1073: The amplitudes for the diagrams in figure \ref{cap:box diagrams}
1074: are then }
1075:
1076: \begin{equation}
1077: \begin{array}{rcl}
1078: {\mathcal{{A}}}_{D1} & = & -ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right)\gamma^{\mu}\left(-{\not k}-\not p_{+}+m_{e}\right){\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}\left(-{\not k}-\not q_{+}+m_{f}\right)\gamma_{\mu}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{+}\right)\left(k^{2}+2k.q_{+}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1079: {\mathcal{{A}}}_{D2} & = & -ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right){\mathcal{{C}}}_{e^{+}}\left(\not k+\not p_{-}+m_{e}\right)\gamma^{\mu}u\left(p_{-}\right)\bar{u}\left(q_{-}\right)\gamma_{\mu}\left(\not k+\not q_{-}+m_{f}\right){\mathcal{{C}}}_{\bar{f}}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{-}\right)\left(k^{2}+2k.q_{-}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1080: {\mathcal{{A}}}_{C1} & = & +ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right){\mathcal{{C}}}_{e^{+}}\left(\not k+\not p_{-}+m_{e}\right)\gamma^{\mu}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}\left(-{\not k}-\not q_{+}+m_{f}\right)\gamma_{\mu}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{-}\right)\left(k^{2}+2k.q_{+}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1081: {\mathcal{{A}}}_{C2} & = & +ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right)\gamma^{\mu}\left(-{\not k}-\not p_{+}+m_{e}\right){\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right)\gamma_{\mu}\left(\not k+\not q_{-}+m_{f}\right){\mathcal{{C}}}_{\bar{f}}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{+}\right)\left(k^{2}+2k.q_{-}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\end{array},\label{eq:5.5.2.1}\end{equation}
1082: where we have defined the loop momentum in each case such that it
1083: flows counter-clockwise with the $X^{0}$ propagator carrying momentum
1084: $k+p_{+}+p_{-}$\emph{i.e.} the propagator is the same in each diagram.
1085: We have defined the loop momentum $k$ such that in each diagram it
1086: is the photon four momentum flowing into the incoming antifermion
1087: line. Firstly condense the numerator as much as possible. The ${\not p}$
1088: terms in the numerator may be quickly got rid of by anticommuting
1089: them so that the on-shell Dirac equation may be used. Considering
1090: the left hand side of the numerator of ${\mathcal{{A}}}_{D1}$\begin{equation}
1091: \begin{array}{rl}
1092: & \bar{\nu}\left(p_{+}\right)\gamma^{\mu}\left(-{\not k}-{\not p_{+}}+m_{e}\right)...\\
1093: = & e\bar{\nu}\left(p_{+}\right)\left(-\gamma^{\mu}{\not k}+2p_{+}^{\mu}\right)...\end{array}.\label{eq:5.5.2.2}\end{equation}
1094: Using the same techniques on the right hand side of the various numerators
1095: the amplitude can be simplified to
1096:
1097: \begin{equation}
1098: \begin{array}{rcl}
1099: {\mathcal{{A}}}_{D1} & = & -ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right)\left(\gamma^{\mu}\not k+2p_{+}^{\mu}\right){\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}\left(\not k\gamma_{\mu}+2q_{+\mu}\right)\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{+}\right)\left(k^{2}+2k.q_{+}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1100: {\mathcal{{A}}}_{D1} & = & -ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right){\mathcal{{C}}}_{e^{-}}\left(\not k\gamma^{\mu}+2p_{-}^{\mu}\right)u\left(p_{-}\right)\bar{u}\left(q_{-}\right)\left(\gamma_{\mu}\not k+2q_{-\mu}\right){\mathcal{{C}}}_{f}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{-}\right)\left(k^{2}+2k.q_{-}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1101: {\mathcal{{A}}}_{C1} & = & +ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right){\mathcal{{C}}}_{e^{-}}\left(\not k\gamma^{\mu}+2p_{-}^{\mu}\right)u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}\left(\not k\gamma_{\mu}+2q_{+\mu}\right)\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{-}\right)\left(k^{2}+2k.q_{+}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\\
1102: {\mathcal{{A}}}_{C2} & = & +ie^{2}\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\textrm{ }{\mathcal{{P}}}_{X^{0}}\frac{\bar{\nu}\left(p_{+}\right)\left(\gamma^{\mu}\not k+2p_{+}^{\mu}\right){\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right)\left(\gamma_{\mu}\not k+2q_{-\mu}\right){\mathcal{{C}}}_{f}\nu\left(q_{+}\right)}{k^{2}\left(k^{2}+2k.p_{+}\right)\left(k^{2}+2k.q_{-}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\end{array}.\label{eq:5.5.2.3}\end{equation}
1103: Assuming that the couplings $\left({\mathcal{{C}}}_{e/f}\right)$and
1104: numerator of the $X^{0}$ propagator ${\mathcal{{P}}}_{X^{0}}$ do
1105: not depend on the loop momentum we can decompose the amplitudes into
1106: \emph{tensor} $\left(T_{\alpha\beta}\right)$, \emph{vector} $\left(V_{\alpha}\right)$
1107: and \emph{scalar} $\left(S\right)$ loop integrals according to the
1108: number of powers of $k$ in the numerator
1109:
1110: \begin{equation}
1111: \begin{array}{rcl}
1112: {\mathcal{{A}}}_{D1} & = & -ie^{2}\bar{\nu}\left(p_{+}\right)\gamma^{\mu}\gamma^{\alpha}...\gamma^{\beta}\gamma_{\mu}\nu\left(q_{+}\right)\times T_{\alpha\beta}\\
1113: & & -ie^{2}\bar{\nu}\left(p_{+}\right)\left(\gamma^{\mu}\gamma^{\alpha}...2q_{+\mu}+2p_{+}^{\mu}...\gamma^{\alpha}\gamma_{\mu}\right)\nu\left(q_{+}\right)\times V_{\alpha}\\
1114: & & -4ie^{2}\left(p_{+}.q_{+}\right)\bar{\nu}\left(p_{+}\right)...\nu\left(q_{+}\right)\times S\end{array},\label{eq:5.5.2.4}\end{equation}
1115: where\begin{equation}
1116: \begin{array}{lcl}
1117: T_{\alpha\beta} & = & \int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{k_{\alpha}k_{\beta}}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(\left(k+p_{+}+p_{-}\right)^{2}-m_{X}^{2}\right)}\\
1118: V_{\alpha} & = & \int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{k_{\alpha}}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(\left(k+p_{+}+p_{-}\right)^{2}-m_{X}^{2}\right)}\\
1119: S & = & \int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(\left(k+p_{+}+p_{-}\right)^{2}-m_{X}^{2}\right)}\end{array}\label{eq:5.5.2.5}\end{equation}
1120: and ${\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}$
1121: has been replaced with {}``$...$'' for brevity. We can see that
1122: the scalar term in the amplitude of \ref{eq:5.5.2.4} is a factor
1123: multiplied by the born scattering amplitude of the $X^{0}$particle\begin{equation}
1124: 2it\times e^{2}\times S\times\left(s-m_{X}^{2}\right)\times\left(\frac{\bar{\nu}\left(p_{+}\right){\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}\nu\left(q_{+}\right)}{s-m_{X}^{2}}\right)=2it\times e^{2}\times S\times\left(s-m_{X}^{2}\right)\times{\mathcal{{A}}}_{Born}.\label{eq:5.5.2.6}\end{equation}
1125: The numerator algebra has given a universal%
1126: \footnote{Universal in the sense that we have not referred to the specific details
1127: of the $X^{0}$ exchange process. The amplitudes corresponding to
1128: the other diagrams $\left({\mathcal{{A}}}_{D2},{\mathcal{{A}}}_{C1},{\mathcal{{A}}}_{C2}\right)$
1129: also give such factors the only difference being the replacement $t\leftrightarrow u$
1130: for crossed box diagrams \emph{i.e.} the differences in the universal
1131: factors are due to the topology of the diagrams rather than their
1132: physics. %
1133: } factor which multiplies the tree level numerator algebra $4\left(p_{+}.q_{+}\right)=-2t$
1134: however the same is not true of the denominator. The factor $S\times\left(s-m_{X}^{2}\right)$
1135: depends on the underlying hard scattering process through the dependence
1136: of $S$ on $m_{X}$, the mass of the exchanged particle. Using the
1137: on mass shell relations we can rewrite the denominator of the integrals
1138: in the $\left|k\right|\rightarrow0$ limit as\begin{equation}
1139: \begin{array}{rl}
1140: & \lim_{\left|k\right|\rightarrow0}\frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(\left(k+p_{+}+p_{-}\right)^{2}-m_{X}^{2}\right)}\\
1141: = & \frac{1}{k^{2}\left(2k.p_{+}\right)\left(2k.q_{+}\right)\left(s-m_{X}^{2}\right)}\end{array}.\label{eq:5.5.2.7}\end{equation}
1142: By trivially counting powers of $\left|k\right|$ we see that in
1143: limit $\left|k\right|\rightarrow0$ the tensor integral goes as $\int\textrm{d}\left|k\right|\left|k\right|$,
1144: the vector integral goes as $\int\textrm{d}\left|k\right|$ and the
1145: scalar integral as $\int\textrm{d}\left|k\right|\left|k\right|^{-1}$.
1146: Consequently the scalar integral is infrared divergent and the other
1147: integrals are infrared finite. Simple power counting also shows that
1148: all of the integrals are finite in the ultraviolet limit $\left|k\right|\rightarrow\infty$.
1149: Expanding the last bracket in the denominator we can write,\begin{equation}
1150: S=\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}.\label{eq:5.5.2.8}\end{equation}
1151: In the infrared divergent $\left|k\right|\rightarrow0$ limit we
1152: can rewrite the last part of the denominator\begin{equation}
1153: \left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)=\left(s-m_{X}^{2}\right),\label{eq:5.5.2.9}\end{equation}
1154: therefore the radiative correction factor in the scalar integral
1155: term \begin{equation}
1156: 2it\times e^{2}\times S\times\left(s-m_{X}^{2}\right),\label{eq:5.5.2.10}\end{equation}
1157: has a divergent piece which is universal because the constant factor
1158: $\left(s-m_{X}^{2}\right)$ in \ref{eq:5.5.2.10} cancels that which
1159: appears in the divergent limit of \ref{eq:5.5.2.8}. We can perhaps
1160: express this result in a better way writing\begin{equation}
1161: \frac{\left(s-m_{X}^{2}\right)}{\left(k^{2}+2k.\left(p_{+}+p_{-}\right)+s-m_{X}^{2}\right)}=1-\frac{k^{2}+2k.\left(p_{+}+p_{-}\right)}{\left(k^{2}+2k.\left(p_{+}+p_{-}\right)+s-m_{X}^{2}\right)},\label{eq:5.5.2.11}\end{equation}
1162: which makes the scalar integral term equal to\begin{equation}
1163: \begin{array}{rcl}
1164: 2it\times e^{2}\times S\times\left(s-m_{X}^{2}\right) & = & \int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{2it\times e^{2}}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)}\\
1165: & - & \int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{2it\times e^{2}\times\left(k^{2}+2k.\left(p_{+}+p_{-}\right)\right)}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(k^{2}+s+2k.\left(p_{+}+p_{-}\right)-m_{X}^{2}\right)}\end{array}.\label{eq:5.5.2.12}\end{equation}
1166: The first term on the right hand side of \ref{eq:5.5.2.12} is universal,
1167: the terms in the denominator correspond to the photon propagator and
1168: two fermion propagators and the $-2t$ in the numerator came from
1169: the universal part of the numerator algebra. The second term on the
1170: right hand side of \ref{eq:5.5.2.12} clearly depends on the hard
1171: scattering process (note the $m_{X}$ terms), it is actually a vector
1172: and a tensor loop integral. The universal term goes as $\int\textrm{d}\left|k\right|\left|k\right|^{-1}$
1173: in the $\left|k\right|\rightarrow0$ limit while the other term goes
1174: as $\int\textrm{d}\left|k\right|$ \emph{i.e.} only the universal
1175: part of the scalar integral is infrared divergent. The degree to which
1176: the box diagram corrections are universal depends on how big the first
1177: integral is in \ref{eq:5.5.2.12} relative to the other terms. If
1178: we denote the universal part of the scalar integral coming from the
1179: denominator of the amplitude \begin{equation}
1180: S_{U}\left(p_{+},q_{+}\right)=\int\frac{{\rm {d}}^{4}k}{\left(2\pi\right)^{4}}\frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)},\label{eq:5.5.2.13}\end{equation}
1181: we can then rewrite the amplitude \ref{eq:5.5.2.4}
1182:
1183: \begin{equation}
1184: \begin{array}{rcl}
1185: {\mathcal{{A}}}_{D1} & = & -ie^{2}\bar{\nu}\left(p_{+}\right)\gamma^{\mu}\gamma^{\alpha}...\gamma^{\beta}\gamma_{\mu}\nu\left(q_{+}\right)T_{\alpha\beta}-2ie^{2}tT_{\alpha}^{\alpha}{\mathcal{{A}}}_{Born}\\
1186: & & -ie^{2}\bar{\nu}\left(p_{+}\right)\left(\gamma^{\mu}\gamma^{\alpha}...2q_{+\mu}+2p_{+}^{\mu}...\gamma^{\alpha}\gamma_{\mu}\right)\nu\left(q_{+}\right)V_{\alpha}-4ie^{2}tV.\left(p_{+}+p_{-}\right){\mathcal{{A}}}_{Born}\\
1187: & & +2ie^{2}tS_{U}\left(p_{+},q_{+}\right){\mathcal{{A}}}_{Born}\end{array}.\label{eq:5.5.2.14}\end{equation}
1188: where {}``$...$'' represents ${\mathcal{{C}}}_{e^{-}}u\left(p_{-}\right)\bar{u}\left(q_{-}\right){\mathcal{{C}}}_{f}$.
1189: Hopefully it is clear that the decomposition of such scalar integrals
1190: into a universal part and non-universal part is as general as propagators
1191: with denominators of the form $p^{2}-m^{2}$, that is to say no matter
1192: what $X^{0}$is we will always find a term $S_{U}$ \ref{eq:5.5.2.13}.
1193:
1194: We can repeat this process with the other diagrams (figure \ref{cap:box diagrams})
1195: in exactly the same way, in each case we have the generic result that
1196: the radiative correction from ISR{*}FSR box diagrams factorizes from
1197: the tree level process in the infrared divergent part of the scalar
1198: loop integral. Referring back to \ref{eq:5.5.2.3} we see that the
1199: other topologies of the process give the following scalar terms\begin{equation}
1200: \begin{array}{rcl}
1201: \left.{\mathcal{{A}}}_{D1}\right|_{Scalar} & = & +2ie^{2}tS_{U}\left(p_{+},q_{+}\right){\mathcal{{A}}}_{Born}\\
1202: \left.{\mathcal{{A}}}_{D2}\right|_{Scalar} & = & +2ie^{2}tS_{U}\left(p_{-},q_{-}\right){\mathcal{{A}}}_{Born}\\
1203: \left.{\mathcal{{A}}}_{C1}\right|_{Scalar} & = & -2ie^{2}uS_{U}\left(p_{-},q_{+}\right){\mathcal{{A}}}_{Born}\\
1204: \left.{\mathcal{{A}}}_{C2}\right|_{Scalar} & = & -2ie^{2}uS_{U}\left(p_{+},q_{-}\right){\mathcal{{A}}}_{Born}\end{array}.\label{eq:5.5.2.14a}\end{equation}
1205:
1206:
1207: Again we have found what we are looking for, factorization of the
1208: radiative corrections. The ideal result, namely that the differential
1209: cross section due to ISR{*}FSR box diagrams interfering with the tree
1210: level process is of the form\begin{equation}
1211: \textrm{d}\sigma_{Box}=\textrm{Universal Factor}\times\textrm{d}\sigma_{Born},\label{eq:5.5.2.15}\end{equation}
1212: \emph{i.e.} \emph{exactly} \emph{factorizable} requires that the
1213: vector and tensor integral terms as well as the non-universal parts
1214: of $S$ are negligible relative to the infrared divergent, universal
1215: piece of $S$. Clearly an exact factorization is impossible, factorization
1216: will only ever occur to some degree. The question then becomes, is
1217: a good degree of factorization possible and generic? Requiring a universally
1218: good degree of factorization means that the factorizable part of the
1219: amplitude must somehow dominate all of the other parts. It does not
1220: seem improbable that this be the case as the universal correction
1221: we are interested in \emph{always} corresponds to an infrared divergence,
1222: in fact it corresponds to the only divergence, infrared or otherwise.
1223: We shall briefly postpone the discussion regarding the relative sizes
1224: of the various contributions to $\textrm{d}\sigma_{Box}$ to discuss
1225: an initial simplifying assumption.
1226:
1227: It is important to note that some of the new physics $X^{0}$ particles
1228: which we hope to apply this analysis to will have momentum dependent
1229: couplings \emph{e.g.} gravitons (gravitons couple to the energy momentum
1230: tensor). Another, more pertinent point is that bosonic propagators
1231: generically contribute polynomials in the momentum of the particle
1232: they represent to the numerator as well as the denominator of the
1233: amplitude. This does not change our current analysis except at the
1234: point of decomposition into scalar vector and tensor integrals \ref{eq:5.5.2.4}
1235: as gravitons will give ${\mathcal{{C}}}_{e^{-}},\textrm{ }{\mathcal{{C}}}_{f}$
1236: and ${\mathcal{{P}}}_{X^{0}}$ then have a dependence on the loop
1237: momentum. In this case the same decomposition can be done into scalar,
1238: vector and tensor integrals and crucially one finds that the scalar
1239: term is the same as in equation \ref{eq:5.5.2.14}. To see this all
1240: one has to do is simply multiply out any $k$ dependence in ${\mathcal{{P}}}_{X^{0}},\textrm{ }{\mathcal{{C}}}_{e^{-}},\textrm{ }{\mathcal{{C}}}_{f}$
1241: \emph{i.e.} decompose ${\mathcal{{P}}}_{X^{0}},\textrm{ }{\mathcal{{C}}}_{e^{-}},\textrm{ }{\mathcal{{C}}}_{f}$
1242: into scalar, vector bits \emph{etc}. Another way to understand this
1243: is to recall that the scalar part is infrared divergent and the momentum
1244: in the photon in our diagrams is just $k$ the loop momentum, so naively
1245: one can think of the photon line in the diagrams as vanishing in the
1246: infrared divergent limit. Thus even in the case where there is a non-trivial
1247: dependence of the couplings and propagator of $X^{0}$ on the loop
1248: momentum the generic infrared term still occurs, essentially one just
1249: has to expand in the loop momentum. In the case of momentum dependence
1250: of the couplings \emph{etc} the tensor integrals will naturally involve
1251: tensors of higher rank than just two \emph{i.e.} we will have tensor
1252: integrals of the form\begin{equation}
1253: \int\frac{\textrm{d}^{4}k}{\left(2\pi\right)^{4}}\frac{k_{\alpha}k_{\beta}k_{\gamma}...}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)\left(\left(k+p_{+}+p_{-}\right)^{2}-m_{X}^{2}\right)}.\label{eq:5.5.2.25}\end{equation}
1254: This causes us to rethink the divergent structure of the amplitude,
1255: up to now only the term involving the scalar integral, the term that
1256: represents the universal radiative correction was (infrared) divergent,
1257: all other terms were ultraviolet and infrared finite. This was encouraging
1258: from the point of view that this could be expected to make the universal
1259: correction the dominant one. In the most general scenario momentum
1260: dependent couplings and the numerator of the propagator could give
1261: ultraviolet terms. Power counting shows that one requires four powers
1262: of the loop momentum in the numerator to have an ultraviolet divergence.
1263: In Giudice \emph{et al} \cite{Giudice:1998ck} the Feynman rules are
1264: derived in the unitary gauge in which the graviton has a propagator
1265: similar to that of the standard model gauge bosons in unitary gauge,
1266: that is to say the high energy behavior goes as $\frac{1}{m^{2}}$.
1267: We shall suppose that in this model (or at least in a fully consistent
1268: theory of gravity) it is possible to work in a gauge analogous to
1269: the the $R_{\xi}$ gauges in which the graviton propagator goes as
1270: $\frac{1}{p^{2}}$ at high energy. With this assumption then we have
1271: at most four powers of the loop momentum in the numerator of any loop
1272: integral, two from the internal fermions and one from each coupling
1273: of the graviton to the fermions. Such integrals diverge logarithmically.
1274: Assuming we cut the integral off at $\left|k\right|=\Lambda$ they
1275: should contribute logarithms $\sim\log\frac{\Lambda^{2}}{s}$ ($\Lambda$
1276: and $\sqrt{s}$ are essentially the only two energy scales in the
1277: diagram). The model of \cite{Giudice:1998ck} has a cut off of order
1278: $1\textrm{ TeV}$. For $\Lambda=10\textrm{ TeV}$, $s=200\textrm{ GeV}$
1279: we have $\log\frac{\Lambda^{2}}{s}=5$, this is something we should
1280: be mindful of when considering the size of any infrared divergences.
1281:
1282: The infrared divergent universal term must cancel the corresponding
1283: infrared divergence from the bremsstrahlung calculated in the last
1284: section. To see this we have to regularize the integral. We extend
1285: the dimensionality of the integral from $4$ to $n=4+\epsilon^{\prime}$
1286: dimensions as in section \ref{sub:Oa3brem}\begin{equation}
1287: S_{U}\left(p_{+},q_{+}\right)=\frac{1}{\left(2\pi\right)^{4}}\lim_{\epsilon^{\prime}\rightarrow0}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int\textrm{d}^{4+\epsilon^{\prime}}k\textrm{ }\frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)}.\label{eq:5.5.2.26}\end{equation}
1288: The denominators can be combined through the usual method of introducing
1289: Feynman parameters\begin{equation}
1290: \frac{1}{k^{2}\left(\left(k+p_{+}\right)^{2}-m_{e}^{2}\right)\left(\left(k+q_{+}\right)^{2}-m_{f}^{2}\right)}=2\int_{0}^{1}\textrm{d}y\textrm{ }y\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\left(k^{2}+2yk.\left(xp_{+}+\left(1-x\right)q_{+}\right)\right)^{3}}.\label{eq:5.5.2.27}\end{equation}
1291: To use the standard integrals we complete the square in the denominator
1292: and shift the variable of integration by a constant amount, $k^{\mu}\rightarrow\tilde{k}^{\mu}=k^{\mu}+y\left(xp_{+}^{\mu}+\left(1-x\right)q_{+}^{\mu}\right)$
1293: to give\begin{equation}
1294: S_{U}\left(p_{+},q_{+}\right)=\frac{-i\pi^{2}}{\left(2\pi\right)^{4}}\lim_{\epsilon^{\prime}\rightarrow0}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{0}^{1}\textrm{d}x\textrm{ }\int_{0}^{1}\textrm{d}y\textrm{ }y\textrm{ }\pi^{\frac{1}{2}\epsilon^{\prime}}\Gamma\left(1-\frac{1}{2}\epsilon^{\prime}\right)\left(\frac{1}{y^{2}\left(xp_{+}+\left(1-x\right)q_{+}\right)^{2}}\right)^{1-\frac{1}{2}\epsilon^{\prime}}.\label{eq:5.5.2.29}\end{equation}
1295: From now on we denote $\Delta\left(p_{+},q_{+}\right)=\left(xp_{+}+\left(1-x\right)q_{+}\right)^{2}$.
1296: In the current form we can perform the integral over $y$,\begin{equation}
1297: \int_{0}^{1}\textrm{d}y\textrm{ }y\left(\frac{1}{y^{2}}\right)^{1-\frac{1}{2}\epsilon^{\prime}}=\left[\frac{1}{\epsilon^{\prime}}y^{\epsilon^{\prime}}\right]_{0}^{1}=\frac{1}{\epsilon^{\prime}}.\label{eq:5.5.2.30}\end{equation}
1298: Substituting this into $S_{U}\left(p_{+},q_{+}\right)$ and expanding
1299: the term in brackets about $\epsilon^{\prime}=0$ gives\begin{equation}
1300: S_{U}\left(p_{+},q_{+}\right)=-\frac{i}{32\pi^{2}}\lim_{\epsilon^{\prime}\rightarrow0}\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta}\left(\frac{1}{\hat{\epsilon}}-\log\left(\frac{\mu^{2}}{s}\right)+\log\left(\frac{\Delta}{s}\right)\right).\label{eq:5.5.2.32}\end{equation}
1301: Expanding $\Delta\left(p_{+},q_{+}\right)$ (keeping all masses)\begin{equation}
1302: \Delta\left(p_{+},q_{+}\right)=\left(x^{2}t-x\left(t-m_{e}^{2}+m_{f}^{2}\right)+m_{f}^{2}\right),\label{eq:5.5.2.33}\end{equation}
1303: we see that $S_{U}\left(p_{+},q_{+}\right)$ is really a function
1304: of the Mandelstam variable $t$. Likewise $\Delta\left(p_{+},q_{+}\right)$
1305: is really a function of $t$, $\Delta\left(t\right)$. We can perform
1306: the integral by completing the square in $\Delta\left(t\right)$\begin{equation}
1307: \Delta\left(p_{+},q_{+}\right)=-t\left(k^{2}-y^{2}\right),\label{eq:5.5.2.34}\end{equation}
1308: where \begin{equation}
1309: \begin{array}{rcl}
1310: y & = & x-\frac{1}{2}\left(1+\delta\right)\\
1311: k^{2} & = & \frac{1}{4}\left(1+\delta\right)^{2}-\frac{m_{f}^{2}}{t}\\
1312: \delta & = & \frac{m_{f}^{2}-m_{e}^{2}}{t}\end{array}.\label{eq:5.5.2.35}\end{equation}
1313: In this case we have\begin{equation}
1314: \int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta\left(t\right)}\log\left(\frac{\Delta\left(t\right)}{s}\right)=-\frac{1}{t}\int_{-\frac{1}{2}\left(1+\delta\right)}^{\frac{1}{2}\left(1-\delta\right)}\textrm{d}y\textrm{ }\frac{1}{k^{2}-y^{2}}\log\frac{-t}{s}\left(k^{2}-y^{2}\right).\label{eq:5.5.2.34}\end{equation}
1315: The integral is awkward and may be made easier by using the following
1316: relation\begin{equation}
1317: \frac{1}{k-y}\log\left(k+y\right)=\frac{\textrm{d}}{\textrm{d}y}\textrm{Li}_{2}\left(\frac{1}{2k}\left(k-y\right)\right)+\frac{1}{k-y}\log\left(2k\right).\label{eq:5.5.2.36}\end{equation}
1318: Substituting back in for $k^{2}$ $\left(k=\frac{1}{2}+\frac{1}{2}\delta-\frac{m_{f}^{2}}{t}\right)$
1319: \emph{etc} and dropping terms ${O}\left(\delta^{2}\right)$ the integral
1320: is found to be \begin{equation}
1321: \begin{array}{rcl}
1322: \int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta\left(t\right)}\log\left(\frac{\Delta\left(t\right)}{s}\right) & = & \frac{1}{4kt}\left[\log^{2}\left(\frac{m_{e}^{2}}{s}\right)+\log^{2}\left(\frac{m_{f}^{2}}{s}\right)-2\log^{2}\left(\frac{-t}{s}\right)\right]\\
1323: & + & \frac{1}{2kt}\left[\textrm{Li}_{2}\left(1+\frac{m_{f}^{2}}{t}\right)-\textrm{Li}_{2}\left(-\frac{m_{e}^{2}}{t}\right)\right]\\
1324: & + & \frac{1}{2kt}\left[\textrm{Li}_{2}\left(1+\frac{m_{e}^{2}}{t}\right)-\textrm{Li}_{2}\left(-\frac{m_{f}^{2}}{t}\right)\right]\end{array}.\label{eq:5.5.2.39}\end{equation}
1325: This simple step is illustrated to condense the structure of divergences
1326: that are present in the limit $m_{e},m_{f}\rightarrow0$, the \emph{collinear
1327: divergences}. A similar decomposition to that in equation \ref{eq:5.5.2.35}
1328: gives the first two terms in \ref{eq:5.5.2.32} as \begin{equation}
1329: \int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta}\left(\frac{1}{\hat{\epsilon}}-\log\left(\frac{\mu^{2}}{s}\right)\right)=\frac{1}{2kt}\left(\frac{1}{\hat{\epsilon}}-\log\left(\frac{\mu^{2}}{s}\right)\right)\left(\log\left(-\frac{m_{e}^{2}}{t}\right)+\log\left(-\frac{m_{f}^{2}}{t}\right)\right).\label{eq:5.5.2.40}\end{equation}
1330: Finally, taking $2k\approx1$ \emph{outside} logarithms gives {\small \begin{equation}
1331: S_{U}\left(t\right)=-\frac{i}{32\pi^{2}}\lim_{\epsilon^{\prime}\rightarrow0}\left[\frac{1}{t}\left(\log\left(-\frac{m_{e}^{2}}{t}\right)+\log\left(-\frac{m_{f}^{2}}{t}\right)\right)\left(\frac{1}{\hat{\epsilon}}-\log\left(\frac{\mu^{2}}{s}\right)\right)+\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta\left(t\right)}\log\left(\frac{\Delta\left(t\right)}{s}\right)\right].\label{eq:5.5.2.41}\end{equation}
1332: }{\small \par}
1333:
1334: In summary we have that in the most general case the box diagrams
1335: under consideration have an amplitude which consists of a scalar IR
1336: divergent term $2ie^{2}tS_{U}\left(t\right){\mathcal{{A}}}_{Born}$
1337: ($-2ie^{2}uS_{U}\left(u\right){\mathcal{{A}}}_{Born}$ for the crossed
1338: box diagrams) along with other infrared finite terms the exact nature
1339: of which depends on the physics of the exchanged $X^{0}$ boson. In
1340: addition it is possible that ultraviolet divergences are present among
1341: the other non-universal terms, in the case of the quantum gravity
1342: model of Giudice \emph{et al} the amplitude is at most logarithmically
1343: divergent and we expect that divergence enhances such a term by a
1344: small factor. Adding the \emph{universal} \emph{contributions} of
1345: the four diagrams \ref{cap:box diagrams} together we have\begin{equation}
1346: \begin{array}{rl}
1347: & \left.{\mathcal{{A}}}_{Box}\right|_{Universal}\\
1348: = & {\mathcal{{A}}}_{D1}+{\mathcal{{A}}}_{D2}+{\mathcal{{A}}}_{C1}+{\mathcal{{A}}}_{C2}\\
1349: = & \frac{\alpha}{\pi}\lim_{\epsilon^{\prime}\rightarrow0}\left[\left(-\frac{1}{\hat{\epsilon}}+\log\left(\frac{\mu^{2}}{s}\right)\right)\log\left(\frac{t}{u}\right)+\frac{t}{2}\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\Delta\left(t\right)}\log\left(\frac{\Delta\left(t\right)}{s}\right)-\frac{u}{2}\int_{0}^{1}{\rm {d}}x\textrm{ }\frac{1}{\Delta\left(u\right)}\log\left(\frac{\Delta\left(u\right)}{s}\right)\right]{\mathcal{{A}}}_{Born}\end{array}.\label{eq:5.5.2.42}\end{equation}
1350:
1351:
1352: \noindent The integral final integral in \ref{eq:5.5.2.42} is the
1353: same as the one preceding it (shown in \ref{eq:5.5.2.39}) with the
1354: replacement $t\rightarrow u$. These two integrals add to give terms
1355: which are finite as $m_{e},m_{f}\rightarrow0$. It is known in the
1356: literature that the initial state-final state interference corrections
1357: contain no collinear divergences, here we see explicitly that this
1358: result is true also for the universal correction as one would expect.
1359: We can safely take the limit $m_{e},m_{f}\rightarrow0$ leaving
1360:
1361: \noindent \begin{equation}
1362: \left.{\mathcal{{A}}}_{Box}\right|_{Universal}=\frac{\alpha}{\pi}\lim_{\epsilon^{\prime}\rightarrow0}\left[\left(-\frac{1}{\hat{\epsilon}}+\log\left(\frac{\mu^{2}}{s}\right)\right)\log\left(\frac{t}{u}\right)+\frac{1}{2}\log^{2}\left(\frac{-u}{s}\right)-\frac{1}{2}\log^{2}\left(\frac{-t}{s}\right)\right].\label{eq:5.5.2.44}\end{equation}
1363: Denoting $\lim_{\hat{\epsilon}\rightarrow0}\left[...\right]$ as
1364: ${\mathcal{{F}}}$, the interference of the box diagrams with the
1365: born amplitude may be written\begin{equation}
1366: \Sigma_{spins}\left(A_{Born}A_{Box}^{\dagger}+A_{Box}A_{Born}^{\dagger}\right)=\frac{2\alpha}{\pi}{\mathcal{{F}}}\textrm{ }\Sigma_{spins}\textrm{ }A_{Born}^{\dagger}A_{Born},\label{eq:5.5.2.45}\end{equation}
1367: hence \begin{equation}
1368: \left.\textrm{d}\sigma_{Box}\right|_{Universal}=\frac{2\alpha}{\pi}{\mathcal{{F}}}\textrm{d}\sigma_{Born}.\label{eq:5.5.2.46}\end{equation}
1369: If we now add the differential cross section due to box diagrams
1370: interfering with the Born level amplitude to the differential cross
1371: section due to soft bremsstrahlung diagrams we obtain the total differential
1372: cross section for the \emph{universal part of} \emph{initial state-final
1373: state interference} as\begin{equation}
1374: \begin{array}{rcl}
1375: \textrm{d}\sigma_{Universal} & = & \textrm{d}\sigma_{Soft,Int}+\left.\textrm{d}\sigma_{Box}\right|_{Universal}\\
1376: & = & \frac{2\alpha}{\pi}\left(\log\left(\frac{s}{4\omega^{2}}\right)\log\left(\frac{u}{t}\right)-\textrm{Li}_{2}\left(1+\frac{s}{t}\right)+\textrm{Li}_{2}\left(1+\frac{s}{u}\right)\right.\\
1377: & & \left.+\frac{1}{2}\log^{2}\left(\frac{-u}{s}\right)-\frac{1}{2}\log^{2}\left(\frac{-t}{s}\right)\right)\textrm{d}\sigma_{Born}\end{array}.\label{eq:5.5.2.47}\end{equation}
1378: This is the universal contribution to the differential cross section
1379: from initial state-final state interference contributions to some
1380: new physics exchange process $\left(X^{0}\right)$ with tree level
1381: differential cross section $\textrm{d}\sigma_{Born}$. For $\sqrt{s}\sim200\textrm{ GeV}$
1382: while one may take the soft photon cut-off to be $\omega\sim50\textrm{ MeV}$
1383: (\emph{i.e.} typical LEP experiments) this gives $\log\left(\frac{s}{4\omega^{2}}\right)=15$,
1384: this constitutes a so-called \emph{large logarithm}. This large logarithm
1385: $\log\left(\frac{s}{4\omega^{2}}\right)$ arises from the cancellation
1386: of the infrared divergences in the box diagram and initial and final
1387: state bremsstrahlung contributions. As noted earlier only the universal
1388: scalar integral terms are infrared divergent thus \emph{no other terms
1389: will undergo this large logarithmic enhancement}. These universal
1390: factorizable corrections maybe resummed (exponentiated) as in the
1391: conventional treatment of the Sudakov effect. Considering only terms
1392: ${\mathcal{{O}}}\left(\alpha\log\left(\frac{s}{4\omega^{2}}\right)\right)$
1393: corresponds to the leading log approximation for initial state-final
1394: state interference. We stress that terms which are not universal are
1395: not infrared divergent and so they will not have the large logarithm
1396: $\log\left(\frac{s}{4\omega^{2}}\right)$, they correspond to \emph{sub-leading
1397: terms} ${\mathcal{{O}}}\left(\alpha\right)$. We conclude that initial
1398: state-final state interference corrections are universal within the
1399: leading log approximation and we anticipate, conservatively, that
1400: the leading log (universal) terms dominate non-universal sub-leading
1401: terms by a factor of 10.
1402:
1403: It is of note that the theoretical study of the ISR{*}FSR interference
1404: by the author began with the calculation of the QED box diagram (\emph{i.e.}
1405: for the case that $X^{0}$ is simply a photon). The \emph{FeynCalc}
1406: \cite{Mertig:an} package was used to assist Dirac traces and reduction
1407: of the \emph{Passarino Veltman} \emph{functions} \cite{Passarino:1978jh}.
1408: The universal leading log term was compared to the non-universal terms
1409: in this case and was found to dominate the sub-leading terms by a
1410: factor of at least $13$ across the $\cos\theta$ region under study.
1411:
1412:
1413: \section{Summary and Conclusions}
1414:
1415: We have seen that the radiator functions describing ISR, the dominant
1416: correction, are merely the result of an intrinsically factorized DGLAP
1417: type analysis with no additional non-factorizable components and that
1418: these describe the total cross section to better than the 1\% level.
1419: In addition we have established that even ISR{*}FSR interference contributions
1420: to the differential cross section are universal to better than the
1421: leading log approximation (note that equation \ref{eq:5.5.2.47} contains
1422: sub-leading terms which are nonetheless universal). With this in mind
1423: we feel justified in using the numerical radiator functions obtained
1424: from the \texttt{ZFITTER} program or its relatives to correct the
1425: tree-level differential cross sections of processes involving the
1426: exchange of some new particle. This method is valid within the context
1427: of the leading log approximation (and also somewhat beyond).
1428:
1429: We would like to stress however that though we conclude that to good
1430: approximation the radiative corrections we have discussed are independent
1431: of the new physics in the diagrams they are not independent of the
1432: topology of the diagrams. The \texttt{ZFITTER} package and this analysis
1433: is only strictly valid for the case of $s$-channel difermion processes.
1434: \texttt{ZFITTER} treats the $t$-channel processes \emph{i.e.} $e^{+}e^{-}\rightarrow e^{+}e^{-}$
1435: differently to the other difermion final states. Our analysis of the
1436: DGLAP electron structure function approach to ISR will still hold
1437: for $t$-channel processes, these corrections pertain only to the
1438: incoming particles. On the other hand we expect that our analysis
1439: of the universality of ISR{*}FSR interference as discussed in section
1440: \ref{sub:Oa3Virtual} will be modified, certainly one should expect
1441: the angular dependence of the radiative corrections to be different.
1442: Nevertheless a large logarithmic, universal, factorizable term will
1443: result. This term will be, by analogy to the $s$-channel result,
1444: of the form $\frac{2\alpha}{\pi}\log\left(\frac{t}{4\omega^{2}}\right)\log\left(...\right)$
1445: \emph{i.e.} the $t$-channel result should still involve a logarithm
1446: of the ratio of the hard scale physics scale to the soft scale $\left(\omega^{2}\right)$.
1447: Provided that the cuts on the $\cos\theta$ region are not too loose
1448: \emph{i.e.} provided all the events in the sample are such that $t\gg\omega^{2}$,
1449: the resulting enhancement for the universal term $\left(\log\left(\frac{t}{4\omega^{2}}\right)\right)$
1450: will still be large thus making the use of radiative corrections to
1451: $e^{+}e^{-}\rightarrow e^{+}e^{-}$ in the standard model valid, to
1452: good approximation, for new physics $e^{+}e^{-}\rightarrow e^{+}e^{-}$
1453: processes.
1454:
1455:
1456: \section{Acknowledgments}
1457:
1458: Thanks to P.Renton, P.J.Holt and O.Vives for helpful discussions.
1459:
1460: \appendix
1461: \renewcommand{\theequation}{A.\arabic{equation}}
1462:
1463: \setcounter{equation}{0}
1464:
1465:
1466: \section*{Appendix A\label{sec:Appendix-A}}
1467:
1468: A difermion event is basically an event where an electron and positron
1469: in the initial state exchange a $Z^{0*}$ or a photon $\gamma^{*}$
1470: and leaves a fermion and an anti-fermion in the final state. This
1471: can happen via $s$-channel $t$-channel exchange, these are depicted
1472: in figure \ref{cap:tree level difermion} at tree level. $t$-channel
1473: processes are only possible for a dielectron final state. %
1474: \begin{figure}
1475: \begin{center}\includegraphics[%
1476: width=0.27\paperwidth,
1477: height=0.18\paperwidth]{stree.eps}\includegraphics[%
1478: width=0.27\paperwidth,
1479: height=0.18\paperwidth]{ttree.eps}\end{center}
1480:
1481:
1482: \caption{\label{cap:tree level difermion}The diagrams for tree level difermion
1483: processes. From left to right these are $s$-channel and $t$-channel
1484: processes respectively. Only electron-positron final states may result
1485: from $t$-channel processes. }
1486: \end{figure}
1487:
1488:
1489: Radiative corrections to these processes are significant in the experiment
1490: and the definition of a difermion process is modified. More generally
1491: we define a difermion event as an event where an electron and positron
1492: emit photons in the initial state then undergo an interaction which
1493: produces a fermion anti-fermion pair and photons, finally the fermion
1494: and anti-fermion may emit additional bremsstrahlung photons. Generally
1495: this can be reduced to three types of diagram, the tree level diagrams
1496: with bremsstrahlung from the external legs and vertices replaced by
1497: effective vertices (one-particle-irreducible diagrams with possible
1498: photons emitted from them) and one with photons radiated off the external
1499: legs and a single 4-point effective vertex, to lowest order this is
1500: a simple \emph{box diagram}. In addition it is possible that photons
1501: are radiated off the one-particle-irreducible diagrams representing
1502: the vertices.
1503:
1504: %
1505: \begin{figure}
1506: \begin{center}\includegraphics[%
1507: width=0.27\paperwidth,
1508: height=0.18\paperwidth]{spr.eps}\includegraphics[%
1509: width=0.27\paperwidth,
1510: height=0.18\paperwidth]{4pt.eps}\end{center}
1511:
1512:
1513: \caption{\label{cap:s-chan boxes}Diagrams corresponding to difermion events
1514: involving bremsstrahlung. The black circles denote all possible one-particle-irreducible
1515: diagrams. The arrows denote the flow of fermion number, momenta $p_{+}$,
1516: $p_{-}$ are flowing inward, $q_{+}$,$q_{-}$ are flowing outward. }
1517: \end{figure}
1518:
1519:
1520: We define the Mandelstam variables $s$, $t$ , $u$ as%
1521: \footnote{Momenta $p_{i}$ are flowing inward, $q_{i}$ flow outward.%
1522: }\begin{equation}
1523: \begin{array}{lclcl}
1524: s & = & \left(p_{+}+p_{-}\right)^{2} & = & \left(q_{+}+q_{-}\right)^{2}\\
1525: t & = & \left(p_{-}-q_{-}\right)^{2} & = & \left(p_{+}-q_{+}\right)^{2}\\
1526: u & = & \left(p_{-}-q_{+}\right)^{2} & = & \left(p_{+}-q_{-}\right)^{2}\end{array}.\label{eq:A.1}\end{equation}
1527: Neglecting the mass of the electron it has four momentum $p_{-}=\left(E_{b},0,0,E_{b}\right)$
1528: prior to any bremsstrahlung, this defines the $z$-axis for the experiment,
1529: the positron has four momentum $p_{-}=\left(E_{b},0,0,-E_{b}\right)$,
1530: consequently \begin{equation}
1531: E_{b}=\frac{1}{2}\sqrt{s}.\label{eq:A.2}\end{equation}
1532: The \emph{initial state radiation} (\emph{ISR}) is defined as that
1533: which is emitted from the external legs of the diagrams. We denote
1534: the combined \emph{outgoing} momentum of the radiation emitted from
1535: the external legs of momentum $p_{i}$ by $k_{j}$ $\left(i=+,-,\textrm{ }j=1,2\right)$
1536: and from the external legs of momentum $q_{i}$ by $k_{j}$ $\left(i=+,-,\textrm{ }j=3,4\right)$.
1537: Primed variables are defined to enable us to discuss the difermion
1538: process in the presence of radiative corrections where the invariant
1539: mass is reduced\begin{equation}
1540: \begin{array}{rcl}
1541: p_{-}^{\prime} & = & p_{-}-k_{1}\\
1542: p_{+}^{\prime} & = & p_{+}-k_{2}\\
1543: q_{-}^{\prime} & = & q_{-}+k_{3}\\
1544: q_{+}^{\prime} & = & q_{+}+k_{4}\end{array}.\label{eq:A.3}\end{equation}
1545: In the presence of such radiative corrections there is some ambiguity
1546: in the definition of $s$, $t$, $u$ as \begin{equation}
1547: \left(p_{-}+p_{+}\right)^{2}\ne\left(q_{-}+q_{+}\right)^{2}\label{eq:A.4}\end{equation}
1548: \emph{etc}, hence for the case of radiative corrections we use \begin{equation}
1549: \begin{array}{lclcl}
1550: s & = & \left(p_{-}+p_{+}\right)^{2}\\
1551: t & = & \left(p_{-}-q_{-}\right)^{2}\\
1552: u & = & \left(p_{+}-q_{-}\right)^{2}\\
1553: s^{\prime} & = & \left(p_{-}^{\prime}+p_{+}^{\prime}\right)^{2} & = & \left(q_{-}^{\prime}+q_{+}^{\prime}\right)^{2}\\
1554: t^{\prime} & = & \left(p_{-}^{\prime}-q_{-}^{\prime}\right)^{2} & = & \left(p_{+}^{\prime}-q_{+}^{\prime}\right)^{2}\\
1555: u^{\prime} & = & \left(p_{-}^{\prime}-q_{+}^{\prime}\right)^{2} & = & \left(p_{+}^{\prime}-q_{-}^{\prime}\right)^{2}\end{array}.\label{eq:A.5}\end{equation}
1556:
1557:
1558: \appendix
1559: \renewcommand{\theequation}{B.\arabic{equation}}\setcounter{equation}{0}
1560:
1561:
1562: \section*{Appendix B\label{sec:Appendix-B}}
1563:
1564: In this appendix we show briefly the integration of the soft photon
1565: phase space factor associated with the soft bremsstrahlung discussed
1566: in subsection \ref{sub:Oa3brem}. The integral to be performed is
1567:
1568: \begin{equation}
1569: \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{\textrm{d}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}.\label{eq:5.5.15b}\end{equation}
1570: Following 't Hooft and Veltman \cite{'tHooft:1978xw} we introduce
1571: a parameter $\kappa$ and define\begin{equation}
1572: \begin{array}{rcl}
1573: p & = & \kappa p_{i}\\
1574: q & = & p_{j}\end{array},\label{eq:5.5.16}\end{equation}
1575: where $\kappa$ is defined to be the solution of $\left(p-q\right)^{2}=0$
1576: which gives $q_{0}$ with the same sign as $p_{0}-q_{0}$. With these
1577: definitions \ref{eq:5.5.15b} becomes,\begin{equation}
1578: \frac{\kappa}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{\textrm{d}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p\right)\left(k.q\right)}.\label{eq:5.5.17}\end{equation}
1579: now combine the denominators with a Feynman parameter $\left(ab\right)^{-1}=\int_{0}^{1}\textrm{d}x\textrm{ }\left[ax+b\left(1-x\right)\right]^{-2}$.
1580: This easily gives,\begin{equation}
1581: \begin{array}{rl}
1582: & \frac{\kappa}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{\left(k.P\right)^{2}}\\
1583: = & \frac{\kappa}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{0}^{1}\textrm{d}x\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\textrm{ }\frac{1}{k_{0}^{2}\left(P_{0}-\left|\vec{P}\right|\cos\tilde{\theta}\right)^{2}}\end{array}\label{eq:5.5.18}\end{equation}
1584: where we have defined $P=q+x\left(p-q\right)$.
1585:
1586: The integration measure can be written explicitly in terms of its
1587: angular components and momentum component in the usual way. The angular
1588: integrations are simplified by essentially redefining the photon momentum
1589: space axes such that $\tilde{\theta}$ is the polar angle, in this
1590: case all but the polar angle integration is trivial. This is analogous
1591: to how integration over the azimuthal angle in three dimensional problems
1592: gives a factor $2\pi$ when the integrand has rotational symmetry
1593: about one axis, the $d$-dimensional analogy is a well known result
1594: see \emph{e.g.} \cite{Bardin:ak}\begin{equation}
1595: \int\textrm{d}\Omega_{d}=\frac{2\pi^{\frac{d}{2}}}{\Gamma\left(\frac{d}{2}\right)}\int_{0}^{\pi}\textrm{d}\tilde{\theta}\textrm{ }\sin^{d-1}\tilde{\theta}.\label{eq:5.5.19}\end{equation}
1596: Equation \ref{eq:5.5.16} gives the integration measure $\textrm{d}\Omega_{d}$
1597: on a $d$-sphere, in our case the phase space is a $d=n-2$ dimensional
1598: sphere of radius $\left|\vec{k}\right|$ embedded in $n-1$ dimensions.
1599: Hence we decompose $\textrm{d}^{\epsilon^{\prime}+3}\vec{k}=\left|\vec{k}\right|^{\epsilon^{\prime}+2}\textrm{d}\left|\vec{k}\right|\textrm{d}\Omega_{\epsilon^{\prime}+2}$
1600: giving \ref{eq:5.5.15b} as\begin{equation}
1601: \frac{\kappa}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\frac{2\pi^{1+\frac{\epsilon^{\prime}}{2}}}{\Gamma\left(1+\frac{\epsilon^{\prime}}{2}\right)}\int_{0}^{1}\textrm{d}x\int_{0}^{\pi}\textrm{d}\tilde{\theta}\textrm{ }\frac{\sin^{\epsilon^{\prime}+1}\tilde{\theta}}{\left(P_{0}-\left|\vec{P}\right|\cos\tilde{\theta}\right)^{2}}\int_{\left|\vec{k}\right|<\omega}\textrm{d}\left|\vec{k}\right|k_{0}^{\epsilon^{\prime}-1},\label{eq:5.5.20}\end{equation}
1602: making the replacement $\sin\tilde{\theta}\textrm{d}\tilde{\theta}\rightarrow-\textrm{d}\cos\tilde{\theta}$
1603: this becomes,\begin{equation}
1604: \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{\textrm{d}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}=\frac{\kappa}{2}\frac{\pi^{\frac{\epsilon^{\prime}}{2}}}{\Gamma\left(1+\frac{\epsilon^{\prime}}{2}\right)}\frac{1}{\epsilon^{\prime}}\left(\frac{\omega}{\mu}\right)^{\epsilon^{\prime}}\int_{0}^{1}\textrm{d}x\int_{-1}^{+1}\textrm{d}\cos\tilde{\theta}\textrm{ }\frac{\left(1-\cos^{2}\tilde{\theta}\right)^{\frac{\epsilon^{\prime}}{2}}}{\left(P_{0}-\left|\vec{P}\right|\cos\tilde{\theta}\right)^{2}}.\label{eq:5.5.21}\end{equation}
1605: Expanding in $\epsilon^{\prime}$, dropping terms ${\mathcal{{O}}}\left(\epsilon^{\prime}\right)$
1606: and above this becomes,\begin{equation}
1607: \frac{\kappa}{2}\int_{0}^{1}\textrm{d}x\int_{-1}^{+1}\textrm{d}C\textrm{ }\frac{1}{\left(P_{0}-\left|\vec{P}\right|C\right)^{2}}\left(\frac{1}{\epsilon^{\prime}}+\log\frac{\omega}{\mu}+\frac{\gamma}{2}+\frac{1}{2}\log\pi+\frac{1}{2}\log\left(1-C^{2}\right)\right)\label{eq:5.5.22}\end{equation}
1608: with $\cos\theta$ relabeled as $C$. The angular integrations maybe
1609: safely carried out using a symbolic computer algebra package giving
1610: for integral \ref{eq:5.5.22}\begin{equation}
1611: \frac{\kappa}{2}\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{P^{2}}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{\omega}{\mu}\right)^{2}+2\log2+\frac{P_{0}}{\left|\vec{P}\right|}\log\frac{P_{0}-\left|\vec{P}\right|}{P_{0}+\left|\vec{P}\right|}\right),\label{eq:5.5.23}\end{equation}
1612: (recall $P=q+x\left(p-q\right)$) where we have introduced the infrared
1613: regulator\begin{equation}
1614: \frac{1}{\hat{\epsilon}}=\frac{2}{\epsilon^{\prime}}+\gamma+\log\pi.\label{eq:5.5.24}\end{equation}
1615: Again by analogy to \cite{'tHooft:1978xw}, the integration is transformed
1616: such that it is over $m=P_{0}-\left|\vec{P}\right|$ using\begin{equation}
1617: \begin{array}{lcl}
1618: v & = & \frac{1}{2}\frac{p^{2}-q^{2}}{p_{0}-q_{0}}\\
1619: P^{2} & = & q^{2}+\frac{P_{0}-q_{0}}{p_{0}-q_{0}}\left(p^{2}-q^{2}\right)\\
1620: & = & q^{2}+2vP_{0}-2vq_{0}\\
1621: \left|\vec{P}\right|^{2} & = & P_{0}^{2}-P^{2}\\
1622: & = & P_{0}^{2}-q^{2}-2vP_{0}+2vq_{0}\end{array},\label{eq:5.5.25}\end{equation}
1623: where we have used the fact that $\kappa$ was defined so that $\left(p-q\right)^{2}=0$
1624: in determining $P^{2}$ and $\left|\vec{P}\right|^{2}$ above. In
1625: terms of the new variables we have\begin{equation}
1626: \log\left(\frac{P_{0}-\left|\vec{P}\right|}{P_{0}+\left|\vec{P}\right|}\right)=\log\left(\frac{m-v}{v}.\frac{m}{m-\frac{1}{v}\left(2vq_{0}-q^{2}\right)}\right).\label{eq:5.5.26}\end{equation}
1627: The rest of the integrand and the integration measure transforms
1628: as\begin{equation}
1629: \textrm{d}x\textrm{ }\frac{1}{P^{2}}\frac{P_{0}}{\left|\vec{P}\right|}=\frac{\textrm{d}m}{2vl}\textrm{ }\left(\frac{2}{P^{2}}.\frac{vP_{0}}{\left|\vec{P}\right|}\frac{\textrm{d}P_{0}}{\textrm{d}m}\textrm{ }\right).\label{eq:5.5.27}\end{equation}
1630: Differentiating \ref{eq:5.5.26} with respect to $m$ we find\begin{equation}
1631: \frac{\textrm{d}}{\textrm{d}m}\log\left(\frac{P_{0}-\left|\vec{P}\right|}{P_{0}+\left|\vec{P}\right|}\right)=\frac{2}{P^{2}}\left(\frac{\left|\vec{P}\right|^{2}-P_{0}^{2}}{\left|\vec{P}\right|}\right)\frac{\textrm{d}P_{0}}{\textrm{d}m}+\frac{2}{P^{2}}\frac{vP_{0}}{\left|\vec{P}\right|}\frac{\textrm{d}P_{0}}{\textrm{d}m},\label{eq:5.5.28}\end{equation}
1632: this enables us to rewrite \ref{eq:5.5.27} as\begin{equation}
1633: \textrm{d}x\textrm{ }\frac{1}{P^{2}}\frac{P_{0}}{\left|\vec{P}\right|}=\frac{\textrm{d}m}{2vl}\textrm{ }\left(\frac{\textrm{d}}{\textrm{d}m}\log\left(\frac{P_{0}-\left|\vec{P}\right|}{P_{0}+\left|\vec{P}\right|}\right)-\frac{2}{m-v}\right).\label{eq:5.5.29}\end{equation}
1634: Substituting these transformed quantities \ref{eq:5.5.27} and \ref{eq:5.5.29}
1635: into the phase space integral \ref{eq:5.5.21} gives\begin{equation}
1636: \begin{array}{rl}
1637: & \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\\
1638: = & \frac{\kappa}{2}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{\omega}{\mu}\right)^{2}+2\log2\right)\int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{P^{2}}\\
1639: + & \frac{\kappa}{2}\int_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\frac{{\rm {d}}m}{2vl}\textrm{ }\left(\frac{{\rm {d}}}{{\rm {d}}m}\log\left(\frac{m-v}{v}.\frac{m}{m-\frac{1}{v}\left(2vq_{0}-q^{2}\right)}\right)-\frac{2}{m-v}\right)\log\left(\frac{m-v}{v}.\frac{m}{m-\frac{1}{v}\left(2vq_{0}-q^{2}\right)}\right)\end{array}.\label{eq:5.5.30}\end{equation}
1640: The term $\log\left(...\right)\frac{\textrm{d}}{\textrm{d}m}\log\left(...\right)$
1641: can be rewritten as $\frac{1}{2}\frac{\textrm{d}}{\textrm{d}m}\log^{2}\left(...\right)$
1642: in which case the integration is trivial. The first integral is also
1643: easy, \begin{equation}
1644: \int_{0}^{1}\textrm{d}x\textrm{ }\frac{1}{P^{2}}=\frac{1}{2vl}\log\left(1+\frac{2vl}{q^{2}}\right).\label{eq:5.5.31}\end{equation}
1645: This leaves one integral\begin{equation}
1646: \begin{array}{rl}
1647: & \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\\
1648: = & \frac{\kappa}{4vl}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{\omega}{\mu}\right)^{2}+2\log2\right)\log\left(1+\frac{2vl}{q^{2}}\right)\\
1649: + & \frac{\kappa}{2vl}\textrm{ }\left[\frac{1}{4}\log^{2}\left(\frac{m\left(m-v\right)}{q^{2}+mv-2vq_{0}}\right)\right]_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}-\frac{\kappa}{2vl}\int_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\textrm{d}m\textrm{ }\frac{1}{m-v}\log\left(\frac{m-v}{v}.\frac{m}{m-\frac{1}{v}\left(2vq_{0}-q^{2}\right)}\right)\end{array}.\label{eq:5.5.30}\end{equation}
1650: The final integral is awkward and is done with the help of the following
1651: two identities\begin{equation}
1652: \begin{array}{lcl}
1653: \log\left(m-a\right) & = & \log\left(1-\frac{m-v}{a-v}\right)+\log\left(v-a\right)\\
1654: \frac{\textrm{d}}{\textrm{d}m}\textrm{Li}_{2}\left(A\left(m-v\right)\right) & = & -\frac{1}{m-v}\log\left(1-A\left(m-v\right)\right)\end{array},\label{eq:5.5.31}\end{equation}
1655: {\small \begin{equation}
1656: \begin{array}{rl}
1657: & \int_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}{\rm {d}}m\textrm{ }\frac{1}{m-v}\log\left(\frac{m-v}{v}.\frac{m}{m-\frac{1}{v}\left(2vq_{0}-q^{2}\right)}\right)\\
1658: = & \left[-\textrm{Li}_{2}\left(1-\frac{m}{v}\right)+\textrm{Li}_{2}\left(\frac{v\left(m-v\right)}{\left(2vq_{0}-q^{2}\right)-v^{2}}\right)\right]_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}+\int_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}{\rm {d}}m\textrm{ }\frac{1}{m-v}\log\left(\frac{\left(m-v\right)v}{v^{2}-2vq_{0}+q^{2}}\right)\end{array}.\label{eq:5.5.32}\end{equation}
1659: } Finally we abbreviate $v\left(v^{2}-2vq_{0}+q^{2}\right)^{-1}=c$
1660: and perform a transformation of variables $x=\left(m-v\right)c$,
1661: $\textrm{d}m=\frac{1}{c}\textrm{d}x$ hence\begin{equation}
1662: \begin{array}{rcl}
1663: \int_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\textrm{d}m\textrm{ }\frac{c}{\left(m-v\right)c}\log\left(\left(m-v\right)c\right) & = & \int_{x_{1}}^{x_{2}}\textrm{d}x\textrm{ }\frac{1}{x}\log\left(x\right)\\
1664: & = & \left[\frac{1}{2}\log^{2}\left(\frac{v\left(m-v\right)}{v^{2}-2vq_{0}+q^{2}}\right)\right]_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\end{array}.\label{eq:5.5.33}\end{equation}
1665: For the phase space integral \ref{eq:5.5.14} we now have\begin{equation}
1666: \begin{array}{rl}
1667: \Rightarrow & \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\\
1668: = & \frac{\kappa}{4vl}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{4\omega^{2}}{\mu^{2}}\right)\right)\log\left(1+\frac{2vl}{q^{2}}\right)\\
1669: + & \frac{\kappa}{2vl}\textrm{ }\left[\frac{1}{4}\log^{2}\left(\frac{m\left(m-v\right)}{q^{2}+mv-2vq_{0}}\right)-\frac{1}{2}\log^{2}\left(\frac{v\left(m-v\right)}{q^{2}+v^{2}-2vq_{0}}\right)+\textrm{Li}_{2}\left(1-\frac{m}{v}\right)-\textrm{Li}_{2}\left(\frac{v\left(m-v\right)}{2vq_{0}-q^{2}-v^{2}}\right)\right]_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\end{array}.\label{eq:5.5.34}\end{equation}
1670: Using the identity for the dilogarithm identity $\textrm{Li}_{2}\left(\frac{1}{z}\right)=-\textrm{Li}_{2}\left(z\right)-\frac{1}{2}\log^{2}\left(-z\right)-\zeta\left(2\right)$
1671: this becomes\begin{equation}
1672: \begin{array}{rl}
1673: \Rightarrow & \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\\
1674: = & \frac{\kappa}{4vl}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{4\omega^{2}}{\mu^{2}}\right)\right)\log\left(1+\frac{2vl}{q^{2}}\right)\\
1675: + & \frac{\kappa}{2vl}\textrm{ }\left[\frac{1}{4}\log^{2}\left(\frac{m\left(m-v\right)}{q^{2}+mv-2vq_{0}}\right)+\textrm{Li}_{2}\left(1-\frac{m}{v}\right)+\textrm{Li}_{2}\left(\frac{2vq_{0}-q^{2}-v^{2}}{v\left(m-v\right)}\right)\right]_{q_{0}-\left|\vec{q}\right|}^{p_{0}-\left|\vec{p}\right|}\end{array}.\label{eq:5.5.34}\end{equation}
1676: Retracing the algebra of \ref{eq:5.5.26} backward and using the
1677: relation $p^{2}-2vp_{0}=q^{2}-2vq_{0}$ we can write\begin{equation}
1678: \begin{array}{rcl}
1679: \frac{\left(q_{0}-\left|\vec{q}\right|\right)\left(q_{0}-\left|\vec{q}\right|-v\right)}{q^{2}+\left(q_{0}-\left|\vec{q}\right|\right)v-2vq_{0}} & = & \frac{q_{0}-\left|\vec{q}\right|}{q_{0}+\left|\vec{q}\right|}\\
1680: \frac{\left(p_{0}-\left|\vec{p}\right|\right)\left(p_{0}-\left|\vec{p}\right|-v\right)}{p^{2}+\left(p_{0}-\left|\vec{p}\right|\right)v-2vp_{0}} & = & \frac{p_{0}-\left|\vec{p}\right|}{p_{0}+\left|\vec{p}\right|}\\
1681: \frac{2vq_{0}-q^{2}-v^{2}}{v\left(q_{0}-\left|\vec{q}\right|-v\right)} & = & \frac{v-q_{0}-\left|\vec{q}\right|}{v}\\
1682: \frac{2vp_{0}-p^{2}-v^{2}}{v\left(p_{0}-\left|\vec{p}\right|-v\right)} & = & \frac{v-p_{0}-\left|\vec{p}\right|}{v}\end{array}.\label{eq:5.5.35}\end{equation}
1683: Also for any momentum four vector $V$, in the limit of small masses,
1684: $V^{2}\ll V_{0}^{2}$ and we can approximate, \begin{equation}
1685: \begin{array}{lclcl}
1686: V_{0}-\left|\vec{V}\right| & = & V_{0}-V_{0}\left(1-\frac{V^{2}}{V_{0}^{2}}\right)^{\frac{1}{2}} & = & \frac{V^{2}}{2V_{0}}\\
1687: V_{0}+\left|\vec{V}\right| & = & V_{0}+V_{0}\left(1-\frac{V^{2}}{V_{0}^{2}}\right)^{\frac{1}{2}} & = & 2V_{0}\end{array}.\label{eq:5.5.36}\end{equation}
1688: Specifically in the case of initial state-final state interference
1689: contributions to $\textrm{d}\sigma_{Soft}$ we require four phase
1690: space integrals corresponding to all possible permutations of a bremsstrahlung
1691: photon from an initial state leg and a bremsstrahlung photon from
1692: a final state leg \emph{i.e.} all combinations of $p_{i}$ and $p_{j}$
1693: where $p_{i}=p_{-},\textrm{ }p_{+}$ and $p_{j}=q_{-},\textrm{ }q_{+}$.
1694: Working also in the limit that the photon carries off no energy $p_{i,0}=p_{j,0}$,
1695: the phase space integral can now be written\begin{equation}
1696: \begin{array}{rl}
1697: \Rightarrow & \frac{1}{4\pi}\left(\frac{1}{\mu}\right)^{\epsilon^{\prime}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{\epsilon^{\prime}+3}\vec{k}}{k_{0}}\frac{1}{\left(k.p_{i}\right)\left(k.p_{j}\right)}\\
1698: = & \frac{\kappa}{4vl}\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{4\omega^{2}}{\mu^{2}}\right)\right)\log\left(1+\frac{2vl}{q^{2}}\right)+\frac{\kappa}{8vl}\log^{2}\left(\frac{m_{e}^{2}}{s}\right)-\frac{\kappa}{8vl}\log^{2}\left(\frac{m_{f}^{2}}{s}\right)\\
1699: + & \frac{\kappa}{2vl}\textrm{ }\left[\textrm{Li}_{2}\left(\frac{v-m_{0}+\left|\vec{m}\right|}{v}\right)+\textrm{Li}_{2}\left(\frac{v-m_{0}-\left|\vec{m}\right|}{v}\right)\right]_{m=p_{j}}^{m=\kappa p_{i}}\end{array}.\label{eq:5.5.37}\end{equation}
1700: \textcolor{black}{Returning to equation \ref{eq:5.5.12} we see
1701: that the soft photon contribution to ISR{*}FSR bremsstrahlung is{\small \begin{equation}
1702: {\textrm{{d}}}\sigma_{Soft,Int}=-\frac{\alpha}{4\pi^{2}}\int_{\left|\vec{k}\right|<\omega}\frac{{\rm {d}}^{3}\vec{k}}{k_{0}}\left(\frac{t}{\left(k.p_{+}\right)\left(k.q_{+}\right)}+\frac{t}{\left(k.p_{-}\right)\left(k.q_{-}\right)}-\frac{u}{\left(k.p_{+}\right)\left(k.q_{-}\right)}-\frac{u}{\left(k.p_{-}\right)\left(k.q_{+}\right)}\right){\textrm{{d}}}\sigma_{Born}.\label{eq:5.5.38}\end{equation}
1703: } }Finally we must determine $\kappa$ for each phase space integral
1704: where $\kappa$ was defined earlier as the solution to \begin{equation}
1705: \left(\kappa p_{i}-p_{j}\right)^{2}=0,\label{eq:5.5.39}\end{equation}
1706: for which $\kappa p_{i,0}-p_{j,0}$ and $p_{j,0}$ have the same
1707: sign. We shall demonstrate how to obtain $\kappa$ for the case $p_{i}=p_{+}$
1708: $p_{j}=q_{+}$, the results generalize easily to the other combinations
1709: of momenta. In this case expanding \ref{eq:5.5.39} gives a simple
1710: quadratic equation for $\kappa$. Working in the limit $t\gg m_{f}^{2}$
1711: gives the two solutions $\left(\kappa_{+},\kappa_{-}\right)$ for
1712: $\kappa$ to first order as\begin{equation}
1713: \begin{array}{lcl}
1714: \kappa_{+} & = & -\frac{t}{m_{e}^{2}}\\
1715: \kappa_{-} & = & -\frac{m_{f}^{2}}{t}\end{array}.\label{eq:5.5.40}\end{equation}
1716: In the massless limit the Mandelstam variable $t$ is\begin{equation}
1717: t=-\frac{s}{2}\left(1-\cos\theta_{++}\right),\label{eq:5.5.41}\end{equation}
1718: where $\theta_{++}$ is the angle between $\left|\vec{p_{+}}\right|$
1719: and $\left|\vec{q_{+}}\right|$, therefore $t$ is a negative quantity.
1720: We require the solution for which $\textrm{sign}\left(\kappa p_{+,0}-q_{+,0}\right)=\textrm{sign}\left(q_{+,0}\right)$
1721: \emph{i.e.}\begin{equation}
1722: \textrm{sign}\left(\frac{1}{2}\sqrt{s}\left(\kappa-1\right)\right)=\textrm{sign}\left(\frac{1}{2}\sqrt{s}\right)\label{eq:5.5.42}\end{equation}
1723: so we require $\kappa>1$. Again working in the limit $t\gg m_{f}^{2}$
1724: this means taking solution $\kappa_{+}=-\frac{t}{m_{e}^{2}}$. Exactly
1725: the same mathematics and the same $\kappa$ are obtained using $p_{i}=p_{-}$,
1726: $p_{j}=q_{-}$ and only marginal differences in working give $\kappa_{+}=-\frac{u}{m_{e}^{2}}$
1727: for the other two combinations of momenta. Noting that $\kappa$ is
1728: large we have that to lowest order in small things \begin{equation}
1729: \begin{array}{rcl}
1730: v & = & \frac{1}{2\kappa p_{0}}\kappa^{2}p^{2}\left(1-\frac{q^{2}}{\kappa^{2}p^{2}}\right)\left(1-\frac{q_{0}}{\kappa p_{0}}\right)^{-1}\\
1731: & \approx & \frac{\kappa m_{e}^{2}}{\sqrt{s}}\end{array}.\label{eq:5.5.43}\end{equation}
1732: Now considering solely the dilogarithm terms in the small mass approximation
1733: \ref{eq:5.5.36}, we have for $\kappa=-\frac{t}{m_{e}^{2}}$\begin{equation}
1734: \left[\textrm{Li}_{2}\left(1-\frac{m_{0}-\left|\vec{m}\right|}{v}\right)+\textrm{Li}_{2}\left(1-\frac{m_{0}+\left|\vec{m}\right|}{v}\right)\right]_{m=p_{j}}^{m=\kappa p_{i}}=-\frac{1}{2}\log\left(\frac{m_{e}^{2}}{s}\right)-\textrm{Li}_{2}\left(1+\frac{s}{t}\right)-\frac{\pi^{2}}{3},\label{eq:5.5.44}\end{equation}
1735: where we have used $\textrm{Li}_{2}\left(-\frac{s}{m_{e}^{2}}\right)=-\textrm{Li}_{2}\left(-\frac{m_{e}^{2}}{s}\right)-\frac{1}{2}\log\left(\frac{m_{e}^{2}}{s}\right)-\zeta\left(2\right)$
1736: and $\textrm{Li}_{2}\left(1\right)=\zeta\left(2\right)$. The same
1737: term for the second solution $\kappa=-\frac{u}{m_{e}^{2}}$ takes
1738: the same form as above but with the replacement $t\rightarrow u$.
1739: Substituting all of this into \ref{eq:5.5.38} gives finally\begin{equation}
1740: \textrm{d}\sigma_{Soft,Int}=\frac{2\alpha}{\pi}\left[\left(\frac{1}{\hat{\epsilon}}+\log\left(\frac{4\omega^{2}}{\mu^{2}}\right)\right)\log\left(\frac{t}{u}\right)-\textrm{Li}_{2}\left(1+\frac{s}{t}\right)+\textrm{Li}_{2}\left(1+\frac{s}{u}\right)\right]\textrm{d}\sigma_{Born}.\label{eq:5.5.45}\end{equation}
1741:
1742:
1743: \begin{thebibliography}{10}
1744: %\cite{Renton:td}
1745: \bibitem{Renton:td}
1746: P.~Renton, ``Electroweak Interactions: An Introduction To The Physics Of Quarks And Leptons,'' Cambridge, UK: Univ. Pr. (1990).
1747: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=2192926}{SPIRES entry}
1748: %\cite{Peskin:ev}
1749: \bibitem{Peskin:ev}
1750: M.~E.~Peskin and D.~V.~Schroeder, ``An Introduction To Quantum Field Theory,'' Reading, USA: Addison-Wesley (1995).
1751: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=3485960}{SPIRES entry}
1752: %\cite{Bardin:ak}
1753: \bibitem{Bardin:ak}
1754: D.~Y.~Bardin and G.~Passarino, ``The Standard Model In The Making: Precision Study Of The Electroweak Interactions,'' Oxford, UK: Clarendon (1999).
1755: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=4320115}{SPIRES entry}
1756: %\cite{Field:uq}
1757: \bibitem{Field:uq}
1758: R.~D.~Field, ``Applications Of Perturbative QCD,'' Redwood City, USA: Addison-Wesley (1989).
1759: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=2170230}{SPIRES entry}
1760: %\cite{Hamilton:gy}
1761: \bibitem{Hamilton:gy}
1762: K.~M.~Hamilton and J.~F.~Wheater,
1763: %``Symmetries In Qft,''
1764: arXiv:hep-ph/0310065.
1765: %%CITATION = HEP-PH 0310065;%%
1766: %\cite{Skrzypek:1990qs}
1767: \bibitem{Skrzypek:1990qs}
1768: M.~Skrzypek and S.~Jadach,
1769: %``Exact And Approximate Solutions For The Electron Nonsinglet Structure Function In QED,''
1770: Z.\ Phys.\ C {\bf 49} (1991) 577.
1771: %%CITATION = ZEPYA,C49,577;%%
1772: %\cite{Skrzypek:1992vk}
1773: \bibitem{Skrzypek:1992vk}
1774: M.~Skrzypek,
1775: %``Leading logarithmic calculations of QED corrections at LEP,''
1776: Acta Phys.\ Polon.\ B {\bf 23} (1992) 135.
1777: %%CITATION = APPOA,B23,135;%%
1778: \bibitem{AH:1996}Heitmann, A. (1996), \emph{Approximation der Elektron-Strukturfunktion
1779: im Monte Carlo} \texttt{\emph{CLOV}} \emph{(In German)}, Diplom Thesis,
1780: Technische Hochschule, Darmstadt (\texttt{http://www.physik.uni-kassel.de/\textasciitilde{}aheit/}).
1781: %\cite{Kuraev:hb}
1782: \bibitem{Kuraev:hb}
1783: E.~A.~Kuraev and V.~S.~Fadin,
1784: %``On Radiative Corrections To E+ E- Single Photon Annihilation At High-Energy,''
1785: Sov.\ J.\ Nucl.\ Phys.\ {\bf 41} (1985) 466
1786: [Yad.\ Fiz.\ {\bf 41} (1985) 733].
1787: %%CITATION = SJNCA,41,466;%%
1788: %\cite{Gribov:ri}
1789: \bibitem{Gribov:ri}
1790: V.~N.~Gribov and L.~N.~Lipatov,
1791: %``Deep Inelastic E P Scattering In Perturbation Theory,''
1792: Yad.\ Fiz.\ {\bf 15} (1972) 781
1793: [Sov.\ J.\ Nucl.\ Phys.\ {\bf 15} (1972) 438].
1794: %%CITATION = YAFIA,15,781;%%
1795: %\cite{Gribov:rt}
1796: \bibitem{Gribov:rt}
1797: V.~N.~Gribov and L.~N.~Lipatov,
1798: %``E+ E- Pair Annihilation And Deep Inelastic E P Scattering In Perturbation Theory,''
1799: Yad.\ Fiz.\ {\bf 15} (1972) 1218
1800: [Sov.\ J.\ Nucl.\ Phys.\ {\bf 15} (1972) 675].
1801: %%CITATION = YAFIA,15,1218;%%
1802: %\cite{Nicrosini:1986sm}
1803: \bibitem{Nicrosini:1986sm}
1804: O.~Nicrosini and L.~Trentadue,
1805: %``Soft Photons And Second Order Radiative Corrections To E+ E- $\to$ Z0,''
1806: Phys.\ Lett.\ B {\bf 196} (1987) 551.
1807: %%CITATION = PHLTA,B196,551;%%
1808: %\cite{Boudjema:1996qg}
1809: \bibitem{Boudjema:1996qg}
1810: F.~Boudjema {\it et al.},
1811: %``Standard Model Processes at LEP-2,''
1812: arXiv:hep-ph/9601224.
1813: %%CITATION = HEP-PH 9601224;%%
1814: %\cite{Berends:1987ab}
1815: \bibitem{Berends:1987ab}
1816: F.~A.~Berends, W.~L.~van Neerven and G.~J.~H.~Burgers,
1817: %``Higher Order Radiative Corrections At Lep Energies,''
1818: Nucl.\ Phys.\ B {\bf 297} (1988) 429
1819: [Erratum-ibid.\ B {\bf 304} (1988) 921].
1820: %%CITATION = NUPHA,B297,429;%%
1821: %\cite{Montagna:1998kp}
1822: \bibitem{Montagna:1998kp}
1823: G.~Montagna, O.~Nicrosini, F.~Piccinini and G.~Passarino,
1824: %``TOPAZ0 4.0: A new version of a computer program for evaluation of de-convoluted and realistic observables at LEP 1 and LEP 2,''
1825: Comput.\ Phys.\ Commun.\ {\bf 117} (1999) 278
1826: [arXiv:hep-ph/9804211].
1827: %%CITATION = HEP-PH 9804211;%%
1828: %\cite{Bardin:1999yd}
1829: \bibitem{Bardin:1999yd}
1830: D.~Y.~Bardin, P.~Christova, M.~Jack, L.~Kalinovskaya, A.~Olchevski, S.~Riemann and T.~Riemann,
1831: %``ZFITTER v.6.21: A semi-analytical program for fermion pair production in e+ e- annihilation,''
1832: Comput.\ Phys.\ Commun.\ {\bf 133} (2001) 229
1833: [arXiv:hep-ph/9908433].
1834: %%CITATION = HEP-PH 9908433;%%
1835: %\cite{Ward:2002qq}
1836: \bibitem{Ward:2002qq}
1837: B.~F.~L.~Ward, S.~Jadach and Z.~Was,
1838: %``Precision calculation for e+ e- $\to$ 2f: The K K MC project,''
1839: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 116} (2003) 73
1840: [arXiv:hep-ph/0211132].
1841: %%CITATION = HEP-PH 0211132;%%
1842: %\cite{Bardin:1989cw}
1843: \bibitem{Bardin:1989cw}
1844: D.~Y.~Bardin {\it et al.},
1845: %``The Convolution Integral For The Forward - Backward Asymmetry In E+ E- Annihilation,''
1846: Phys.\ Lett.\ B {\bf 229} (1989) 405.
1847: %%CITATION = PHLTA,B229,405;%%
1848: %\cite{Bardin:1990fu}
1849: \bibitem{Bardin:1990fu}
1850: D.~Y.~Bardin {\it et al.},
1851: %``Analytic Approach To The Complete Set Of QED Corrections To Fermion Pair Production In E+ E- Annihilation,''
1852: Nucl.\ Phys.\ B {\bf 351} (1991) 1
1853: [arXiv:hep-ph/9801208].
1854: %%CITATION = HEP-PH 9801208;%%
1855: %\cite{Bardin:1999kz}
1856: \bibitem{Bardin:1999kz}
1857: D.~Y.~Bardin,
1858: %``Field Theory and the Standard Model,''
1859: CERN-OPEN-2000-292
1860: %\href{http://www.slac.stanford.edu/spires/find/hep/www?r=cern-open-2000-292}{SPIRES entry}
1861: {\it 7th European School of High-Energy Physics, Casta-Papiernicka, Slovak Republic, 22 Aug - 4 Sep 1999}
1862: %\cite{Berends:ie}
1863: \bibitem{Berends:ie}
1864: F.~A.~Berends, R.~Kleiss and S.~Jadach,
1865: %``Radiative Corrections To Muon Pair And Quark Pair Production In Electron - Positron Collisions In The Z(0) Region,''
1866: Nucl.\ Phys.\ B {\bf 202} (1982) 63.
1867: %%CITATION = NUPHA,B202,63;%%
1868: %\cite{Nicrosini:1987sw}
1869: \bibitem{Nicrosini:1987sw}
1870: O.~Nicrosini and L.~Trentadue,
1871: %``Second Order Electromagnetic Radiative Corrections To E+ E- $\to$ Gamma*, Z0 $\to$ Mu+ Mu-,''
1872: Z.\ Phys.\ C {\bf 39} (1988) 479.
1873: %%CITATION = ZEPYA,C39,479;%%
1874: %\cite{Passarino:1978jh}
1875: \bibitem{Passarino:1978jh}
1876: G.~Passarino and M.~J.~G.~Veltman,
1877: %``One Loop Corrections For E+ E- Annihilation Into Mu+ Mu- In The Weinberg Model,''
1878: Nucl.\ Phys.\ B {\bf 160} (1979) 151.
1879: %%CITATION = NUPHA,B160,151;%%
1880: %\cite{Krammer:1975ed}
1881: \bibitem{Krammer:1975ed}
1882: A.~B.~Krammer and B.~Lautrup,
1883: %``Radiative Asymmetry Around The Resonance In E+ E- $\to$ Mu+ Mu-,''
1884: Nucl.\ Phys.\ B {\bf 95} (1975) 380.
1885: %%CITATION = NUPHA,B95,380;%%
1886: %\cite{Veltman:1988au}
1887: \bibitem{Veltman:1988au}
1888: M.~J.~G.~Veltman,
1889: %``Gammatrica,''
1890: Nucl.\ Phys.\ B {\bf 319} (1989) 253.
1891: %%CITATION = NUPHA,B319,253;%%
1892: %\cite{Was:1994kg}
1893: \bibitem{Was:1994kg}
1894: Z.~Was,
1895: %``Radiative corrections,''
1896: CERN-TH-7154-94
1897: %\href{http://www.slac.stanford.edu/spires/find/hep/www?r=cern-th-7154-94}{SPIRES entry}
1898: {\it Written on the basis of lectures given at the 1993 European School of High Energy Physics, Zakopane, Poland, 12-25 Sep 1993}
1899: %\cite{'tHooft:1978xw}
1900: \bibitem{'tHooft:1978xw}
1901: G.~'t Hooft and M.~J.~G.~Veltman,
1902: %``Scalar One Loop Integrals,''
1903: Nucl.\ Phys.\ B {\bf 153} (1979) 365.
1904: %%CITATION = NUPHA,B153,365;%%
1905: %\cite{Giudice:1998ck}
1906: \bibitem{Giudice:1998ck}
1907: G.~F.~Giudice, R.~Rattazzi and J.~D.~Wells,
1908: %``Quantum gravity and extra dimensions at high-energy colliders,''
1909: Nucl.\ Phys.\ B {\bf 544} (1999) 3
1910: [arXiv:hep-ph/9811291].
1911: %%CITATION = HEP-PH 9811291;%%
1912: %\cite{Mertig:an}
1913: \bibitem{Mertig:an}
1914: R.~Mertig, M.~Bohm and A.~Denner,
1915: %``Feyn Calc: Computer Algebraic Calculation Of Feynman Amplitudes,''
1916: Comput.\ Phys.\ Commun.\ {\bf 64} (1991) 345.
1917: %%CITATION = CPHCB,64,345;%%
1918: \end{thebibliography}
1919:
1920: \end{document}
1921: