hep-th0404085/gmn.tex
1: \documentclass[12pt,letterpaper]{article}
2: 
3: \usepackage{epsfig}
4: \usepackage{amsmath}
5: \usepackage[psamsfonts]{amssymb}
6: \usepackage{amsthm}
7: \usepackage{fullpage}
8: \usepackage{indentfirst}
9: 
10: \numberwithin{equation}{section}
11: 
12: % MACROS FROM ANDY 
13: \newcommand{\nid}{\noindent}
14: \newcommand{\for}{\quad \text{for }}
15: \newcommand{\deltabar}{\overline{\delta}}
16: 
17: \newcommand{\R}{\ensuremath{\mathbb R}}
18: \newcommand{\C}{\ensuremath{\mathbb C}}
19: \newcommand{\PP}{\ensuremath{\mathbb P}}
20: \newcommand{\Z}{\ensuremath{\mathbb Z}}
21: \newcommand{\Q}{\ensuremath{\mathbb Q}}
22: \newcommand{\half}{\ensuremath{\frac{1}{2}}}
23: \newcommand{\qtr}{\ensuremath{\frac{1}{4}}}
24: \newcommand{\N}{{\mathcal N}}
25: \newcommand{\dirac}{\!\!\not\!\partial}
26: \newcommand{\Dirac}{\!\!\not\!\!D}
27: \newcommand{\M}{{\mathcal M}}
28: \newcommand{\Mstab}[2]{{\overline{\M_{0,#1,#2}}}}
29: \newcommand{\Mint}{{\M_{{\rm int}}}}
30: \newcommand{\Mintgamma}{{\M^\Gamma_{{\rm int}}}}
31: \newcommand{\Mintlambda}{{\M^\Lambda_{{\rm int}}}}
32: \newcommand{\Mintone}{{\M^1_{{\rm int}}}}
33: \newcommand{\Mintgammaprime}{{\M^{\Gamma'}_{{\rm int}}}}
34: \newcommand{\MintK}{{\M^{K}_{{\rm int}}}}
35: \newcommand{\Mlines}{{\M_{{\rm lines}}}}
36: \newcommand{\Mlinesgamma}{{\M^\Gamma_{{\rm lines}}}}
37: \newcommand{\mulines}{{\mu_{{\rm lines}}}}
38: \newcommand{\muint}{{\mu_{{\rm int}}}}
39: 
40: \newcommand{\abs}[1]{\lvert#1\rvert}
41: \newcommand{\norm}[1]{\lVert#1\rVert}
42: \newcommand{\IP}[1]{\langle#1\rangle}
43: \newcommand{\dwrt}[1]{\frac{\partial}{\partial#1}}
44: \newcommand{\eps}{\epsilon}
45: \newcommand{\ci}{\oint}
46: 
47: \newcommand{\la}{\label}
48: \newcommand{\ti}[1]{\textit{#1}}
49: 
50: \newcommand{\mfn}[1]{%
51:    \footnote{#1}
52:    \marginpar{\hfill {\sf\thefootnote}}%
53: }
54: 
55: \newcommand{\CPP}{{\C\PP^{3|4}}}
56: 
57: 
58: \newcommand{\prodprime}[2]{{\prod_{#1}^{#2} \!\!\!\!\begin{array}{c}\prime\\ \,\end{array}}}
59: 
60: 
61: 
62: % MACROS FROM LUBOS
63: \def\IZ{{\mathbb Z}}           % \mathfrak - gothic, \mathcal - script
64: \def\IR{{\mathbb R}}
65: \def\IC{{\mathbb C}}
66: \def\A#1{{\cal A}_{[#1]}}        % Amplitude - for example \A{++++---}
67: \def\Beta{B}
68: \def\calD{{\cal D}}
69: \def\calH{{\cal H}}
70: \def\calN{{\cal N}}
71: \def\calO{{\cal O}}
72: \def\Box{\square}
73: \def\IA{{\mathbb{A}}}
74: \def\IB{{\mathbb{B}}}
75: \def\IP{{\mathbb{P}}}
76: 
77: 
78: \def \tilde{\widetilde}
79: \def \tb#1{\left(\begin{array}#1\end{array}\right)}
80: \def \cpt{\mathbb{CP}^{(3|4)}}
81: \def \abs#1{\left|#1\right|}
82: \def \lr{\leftrightarrow}
83: \def \eqref#1{(\ref{#1})}
84: \def \bra#1{\left\langle #1\right\vert}
85: \def \ket#1{\left\vert #1\right\rangle}
86: \def \dslash{{\not\!\partial}}
87: \def \exp{\mbox{exp}}
88: \def \tp{\mbox{tp\,}}
89: \def \ignoruj#1{}
90: \def \eqn#1#2{\begin{equation}#2\label{#1}\end{equation}}
91: \def \ha{{1\over 2}}
92: \def \Tr{\mbox{Tr\,}}
93: \def \tr{\mbox{Tr\,}}
94: \def \ham{\mathcal H}
95: \def \Lag{\mathcal L}
96: \def \gs{g_{\mathrm{string}}}
97: \def \inte{{\mathrm{interaction}}}
98: \def \ls{l_{string}}
99: \def \lpl{l_{\mathrm{Planck}}}
100: \def \action{\mathcal A}
101: \def \Ga{\Gamma}
102: \def \a{\alpha}
103: \def \half{{1\over 2}}
104: \def \s{\sigma}
105: \def \l{\lambda}
106: \def \g{\gamma}
107: \def \d{\delta}
108: \def \de{{\rm d}}
109: \def \e{\epsilon}
110: \def \ad{{\dot a}}
111: \def \bd{{\dot b}}
112: \def \cd{{\dot c}}
113: \def \dd{{\dot d}}
114: \def \CH{\mathfrak {CH}}
115: \def \b{\beta}
116: \def \vt{\vartheta}
117: \def \ttheta{{\tilde \theta}}
118: \def \tSigma{{\tilde \Sigma}}
119: %
120: \def\ignorethis#1{}
121: \def\be{\begin{equation}}
122: \def\ee{\end{equation}}
123: \def\ar#1#2{\begin{array}{#1}#2\end{array}}
124: \def\bear{\begin{eqnarray}}
125: \def\eear{\end{eqnarray}}
126: \def\p{\partial }
127: \def\Rads{R_{\mathrm{AdS}}}
128: \def\AdS{\mathrm{AdS}}
129: 
130: 
131: 
132: %---------------------------------------------------------------------------
133: 
134: 
135: \begin{document}
136: 
137: \bibliographystyle{utphys}
138: 
139: \setcounter{page}{1}
140: \pagestyle{plain}
141: 
142: 
143: \begin{titlepage}
144: 
145: \begin{center}
146: \hfill HUTP-04/A016\\
147: \hfill HEP-UK-0021\\
148: \hfill hep-th/0404085
149: 
150: \vskip 1.5 cm
151: {\huge \bf Equivalence of twistor prescriptions\\
152: {\small~\newline} for super Yang-Mills}
153: \vskip 1.3 cm
154: {\large Sergei Gukov, Lubo\v{s} Motl, and Andrew Neitzke}\\
155: \vskip 0.5 cm
156: {Jefferson Physical Laboratory,
157: Harvard University,\\
158: Cambridge, MA 02138, USA}
159: \vskip 0.3cm
160: {
161: {\tt gukov@feynman.harvard.edu}\\
162: {\tt motl@feynman.harvard.edu}\\
163: {\tt neitzke@fas.harvard.edu}}
164: \end{center}
165: 
166: 
167: \vskip 0.5 cm
168: \begin{abstract}
169: There is evidence that one can compute tree level super Yang-Mills amplitudes
170: using either connected or completely disconnected curves in twistor space.
171: We argue that the two computations 
172: are equivalent, if the integration contours are chosen in a specific way, 
173: by showing that they can both be reduced to the same integral over a moduli space 
174: of singular curves.  
175: We also formulate a class of new ``intermediate'' 
176: prescriptions to calculate the same amplitudes.
177: \end{abstract}
178: 
179: \end{titlepage}
180: 
181: \renewcommand{\baselinestretch}{1.4}
182: \small\normalsize
183: 
184: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185: 
186: \tableofcontents
187: 
188: \line(1,0){450}
189: 
190: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
191: 
192: \section{Introduction}
193: 
194: Recently in \cite{Witten:2003nn} Witten proposed a new approach to
195: perturbative gauge theories in four dimensions which, among other things,
196: implies remarkable regularities in the perturbative scattering amplitudes
197: of $\N=4$ super Yang-Mills and leads to new ways of computing them.
198: The scattering amplitudes in question depend on the momentum
199: and polarization vectors of the external gluons,
200: and are devilishly difficult to compute using the standard Feynman diagram
201: techniques.
202: For example, even computing a tree level amplitude with
203: 4 external gluons of positive helicity and 3 gluons of negative helicity
204: (such an amplitude will be denoted $\A{++++---}$)
205: requires summing over hundreds of different diagrams!
206: 
207: According to the conjecture of \cite{Witten:2003nn},
208: perturbative $\N=4$ super Yang-Mills theory can be described
209: as a string theory in twistor space $\CPP$.
210: In this reformulation, the Yang-Mills scattering amplitudes
211: are given by certain integrals over moduli spaces
212: of holomorphic curves in $\CPP$, which can be interpreted as D1-brane instantons.
213: More precisely, for a tree level process involving $q$ negative
214: helicity gluons, the amplitude is given by an integral over
215: moduli of curves of total degree $d$, where
216: %
217: \eqn{dviaq}{ d=q-1. }
218: %
219: For example, the simplest non-vanishing amplitude with $q=2$
220: gluons of negative helicity\footnote{We follow the conventions
221: of \cite{Witten:2003nn} where a $n$-gluon scattering amplitude
222: is called MHV if $n\!-\!2$ external gluons have positive helicity,
223: and $\overline{{\rm MHV}}$ (or ``googly'') if $n\!-\!2$ gluons have
224: negative helicity.} --- the so-called maximally helicity
225: violating (MHV) amplitude \cite{Parke:1986gb,Mangano:1988xk} --- can be computed by integrating over
226: the moduli space of degree $1$ curves in $\CPP$ \cite{Witten:2003nn}.
227: 
228: However, when one considers the next simplest case, $q=3$, there is a puzzle.
229: In the prescription of \cite{Witten:2003nn} this amplitude seems to 
230: involve a sum over two distinct contributions:
231: one from an integral over connected degree $2$ curves,
232: and another from an integral over disconnected pairs
233: of degree $1$ curves; see Figure \ref{linesab}.
234: Surprisingly, in the case of $\A{++---}$, it was found
235: that the contribution from connected degree $2$ curves alone
236: gives the full Yang-Mills amplitude,
237: at least up to a multiplicative constant \cite{Roiban:2004vt}.
238: This computation was extended to all googly \cite{Roiban:2004ka}
239: and some non-MHV \cite{Roiban:2004yf} amplitudes,
240: again with the surprising result that connected degree $d$ curves already
241: account for the full Yang-Mills amplitude, without adding any disconnected curves.
242: 
243: \begin{figure}
244: \begin{center}
245: \epsfig{file=linesab.eps,width=130mm}
246: \end{center}
247: \caption{An instanton contribution: {\bf (a)} from a connected
248: curve of degree 2;\quad {\bf (b)} from a pair of degree 1 curves.
249: The dotted line represents a propagator in holomorphic
250: Chern-Simons theory.}
251: \label{linesab}
252: \end{figure}
253: 
254: On the other hand, there is some evidence that these tree level
255: amplitudes can also be computed by considering only the contribution
256: of curves which are ``maximally disconnected,'' namely,
257: they consist of $d$ distinct degree $1$ lines.
258: Since degree 1 curves are associated with MHV amplitudes,
259: this result suggests an alternative method of computing
260: generic tree amplitudes from graphs with MHV vertices \cite{Cachazo:2004kj}.
261: The number $v$ of vertices is determined by the number
262: of gluons with negative helicity; it is actually equal to the degree
263: \eqref{dviaq},
264: %
265: \eqn{vviaq}{ v=q-1. }
266: %
267: This approach leads to a spectacular simplification of the computations.
268: For example, the 7-gluon amplitude $\A{++++---}$ mentioned earlier can be
269: computed using only 8 diagrams with MHV vertices.
270: However, it also leads to a puzzle.
271: 
272: As we just discussed, the evidence so far in the literature
273: suggests that rather than one prescription
274: for Yang-Mills amplitudes there are at least two:
275: one involving connected curves only, another involving maximally disconnected ones.
276: We will refer to these as the ``connected prescription''
277: and the ``disconnected prescription'' respectively.  
278: These different prescriptions have so far not been related directly.
279: In a sense, they seem to have complementary virtues:  the connected prescription
280: expresses the whole amplitude as a single integral, and 
281: from this form it is easier to prove some properties of the amplitude,
282: such as the parity symmetry; on the other hand, the disconnected prescription
283: leads to concrete and immediately useful formulas for the tree level amplitudes.
284: 
285: The purpose of this note is to argue that the connected and disconnected prescriptions
286: are equivalent, at least for an appropriate choice of the integration contours,
287: and to give an \ti{a priori} explanation for this agreement.
288: The explanation is that, in both prescriptions, the integral
289: over the moduli space is localized to poles on a particular submoduli space. 
290: This submoduli space parameterizes configurations of intersecting degree $1$ curves.
291: 
292: Let us illustrate this explanation in
293: the simplest case of degree $2$ curves.
294: We have two different moduli spaces,  $\Mstab{n}{2}$ and $\Mlines$,
295: parameterizing respectively connected degree $2$ curves in $\CPP$
296: and disconnected pairs of lines in $\CPP$,
297: and integrands $\omega_{{\rm conn}}$
298: and $\omega_{{\rm disc}}$ on the two spaces (we will review the construction of these integrands
299: in Section \ref{sec-review}).  Our job is to explain the equality
300: \begin{equation}
301: \int_\Mstab{n}{2} \omega_{{\rm conn}} = \int_\Mlines \omega_{{\rm disc}}.
302: \end{equation}
303: The explanation begins by noting that both
304: $\Mlines$ and $\Mstab{n}{2}$ contain a codimension-one
305: ``degeneration locus'' $\Mint$ parameterizing the moduli of
306: pairs of intersecting lines in $\CPP$.
307: In the case of $\Mlines$ we get such a degenerate configuration
308: just by taking two lines in $\CPP$ which happen to intersect.
309: For $\Mstab{n}{2}$ we get such a degeneration by
310: considering a hyperbola $xy = C$ in the limit $C \to 0$,
311: appropriately embedded in $\CPP$.
312: The crucial point is that both
313: $\omega_{{\rm conn}}$ and $\omega_{{\rm disc}}$
314: turn out to have a simple pole along $\Mint$,
315: and furthermore the residue is the same in both cases.\footnote{We learned of
316: the possibility of such an explanation from Edward Witten.} Therefore, 
317: provided that the integration contours on
318: $\Mlines$ and $\Mstab{n}{2}$ are chosen compatibly
319: (so that they both encircle $\Mint$ and reduce
320: to the same contour along it), the desired agreement follows.
321: 
322: The argument for general degree $d$ proceeds along similar lines.  In the moduli space
323: $\Mstab{n}{d}$ we find a pole where a degree $d$ curve degenerates into two intersecting
324: curves of degrees $d_1$ and $d_2$; the integral over $\Mstab{n}{d}$ localizes to this sublocus;
325: then inside this sublocus there is a pole where one of the two curves degenerates further, and
326: so on until we reduce finally to the moduli space
327: $\Mint$ of connected trees built from degree $1$ curves.
328: On the other hand, the integral over $\Mlines$ also reduces to the same $\Mint$, 
329: because the propagators connecting the different lines have poles when the lines intersect.
330: Furthermore it turns out that the integrands on $\Mint$ coming from the two prescriptions are
331: proportional.  This establishes the agreement between these two prescriptions, again provided
332: that the contours are chosen appropriately, and up to an overall constant which we do not fix.
333: 
334: This iterative argument pays a surprising dividend:  
335: for any $K = 0, \dots, d-1$, we can define an ``intermediate prescription,''
336: in which we integrate over configurations of $K+1$ curves with total degree $d$.
337: We will show that all of these intermediate prescriptions agree with
338: the connected and disconnected prescriptions.
339: They can also be understood diagrammatically:
340: one sums over tree diagrams with $K+1$ vertices,
341: where each vertex is decorated with a degree.
342: In these notations,
343: vertices of degree 1 are the MHV vertices of \cite{Cachazo:2004kj},
344: whereas vertices with $d>1$ could be called ``non-MHV vertices''.
345: These intermediate prescriptions deserve further study.
346: 
347: For other recent work on the twistor string approach to Yang-Mills, see
348: \cite{Roiban:2004vt,Roiban:2004ka,Roiban:2004yf,Berkovits:2004hg,Berkovits:2004tx,Witten:2004cp} 
349: for the connected prescription, 
350: \cite{Cachazo:2004kj,Zhu:2004kr,tree-scalar-mhv} for the disconnected prescription,
351: and \cite{Aganagic:2004yh,Nekrasov:2004js,Neitzke:2004pf} for related
352: topics.
353: 
354: 
355: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
356: 
357: \subsection{Notation and moduli spaces}
358: 
359: We always consider scattering amplitudes of $n$ external gluons
360: associated with the particular trace factor $\tr (T_1 T_2 \ldots T_n)$.
361: 
362: We use a coordinate representation
363: for the super twistor space $\C^{4|4}$.
364: We unify the bosonic and fermionic indices into
365: a superspace index $\IA$ taking values in
366: \eqn{superindex}{\IA\in
367: \{1,\,2,\,3,\,4\, | 1',2',3',4'\}.}
368: The components of all objects with bosonic values of the 
369: superspace index are commuting, while components with fermionic
370: (primed) values of the superspace index are anticommuting.
371: The coordinates on the super twistor space will be denoted by $Z^\IA$,
372: which are related to the coordinates in the literature by
373: \eqn{twistorcomponents}{(Z^1,Z^2,Z^3,Z^4|Z^{1'},Z^{2'},Z^{3'},Z^{4'})
374: =(\lambda^1,\lambda^2,\mu^1,\mu^2 |
375: \psi^{1}, \psi^{2}, \psi^{3}, \psi^{4})
376: \in {\IC^{4|4}}.} 
377: 
378: We will also be considering various moduli spaces of
379: curves in $\CPP$  with marked points.  We use the standard notation 
380: \begin{equation}
381: \M_{0,n,d}(\CPP)
382: \end{equation}
383: for the moduli space of ``genus $0$, $n$-pointed curves of degree $d$ in $\CPP$.''
384: This moduli space has dimension $(4d+n)|(4d+4)$.
385: As in \cite{Witten:2003nn} we realize it as the space of
386: automorphism classes of maps \,$\C\PP^1 \to \CPP$, of degree $d$, with $n$ marked points on $\C\PP^1$.
387: Since the target space is always $\CPP$ in this paper, 
388: sometimes we abuse notation and write simply $\M_{0,n,d}$.
389: 
390: \vspace{3mm}
391: 
392: \begin{figure}
393: \begin{center}
394: \epsfig{file=deg-lines.eps,width=120mm}
395: \end{center}
396: \caption{A curve of degree 2 can degenerate into a pair
397: of intersecting lines.}
398: \label{deg-lines}
399: \end{figure}
400: 
401: We will be interested in integrating 
402: over $\M_{0,n,d}(\CPP)$, so we need to understand the
403: properties of this moduli space.
404: First, $\M_{0,n,d}(\CPP)$ is non-compact, due
405: to certain degenerations that a degree $d$ curve with $n$ marked
406: points can have which are not simply described by a map
407: $\C\PP^1 \to\CPP$. 
408: One type of degeneration that will be important
409: below is when a curve develops a node, {\it i.e.}\ splits into two components.  
410: There is a standard way of incorporating these degenerate curves into 
411: our moduli space of maps; one then obtains a larger compact space 
412: $\overline{\M_{0,n,d}}(\CPP)$, called the ``moduli space of
413: stable maps.''  This moduli space is a smooth algebraic variety, except for
414: certain orbifold points which will not play an important role in this paper.\footnote{Strictly
415: speaking this theorem has been proven when the
416: target space is $\C\PP^3$ \cite{MR98m:14025}, not for the supermanifold $\CPP$, 
417: but we do not expect any important differences.}
418: 
419: In particular, the ``boundary'' of this moduli space,
420: \begin{equation}
421: \overline{\M_{0,n,d}} (\CPP) \setminus \M_{0,n,d}(\CPP),
422: \end{equation}
423: contains a codimension $1$ divisor
424: which parameterizes curves which have split into two components.
425: Similarly, for any $K$ there is a subspace $\MintK$
426: of codimension $K$ that parameterizes reducible curves
427: with $K$ nodes, {\it i.e.}\ curves which have split up into $K\!+\!1$ intersecting 
428: components which intersect in a tree.
429: This $\MintK$ can be further decomposed into irreducible pieces,
430: \begin{equation}
431: \Mint = 
432: \bigcup_\Gamma \Mintgamma,
433: \end{equation}
434: where the different $\Gamma$ label 
435: different shapes of the tree, together with different decompositions
436: of $d$ into individual degrees $\{d_i\}$, ~
437: $i=1,2,\dots K\!+\!1$, ~
438: $d_i \geq d_{i+1}$,\quad and
439: different ways in which the $n$
440: marked points can be distributed over the 
441: $K\!+\!1$
442: components.
443: Some of these $\Mintgamma$ will play an important role in our discussion below.
444: 
445: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
446: 
447: \section{Review of connected and disconnected prescriptions} \label{sec-review}
448: 
449: Suppose we want to use the twistor prescription of \cite{Witten:2003nn}
450: to evaluate a Yang-Mills amplitude with $q=d+1$ negative helicity gluons.
451: All contributions to this amplitude are expected to involve holomorphic
452: curves of total degree $d$, but \ti{a priori} these can be either
453: connected or disconnected. In this section we review the contributions
454: which would be expected from the two most extreme cases: connected degree $d$
455: curves and completely disconnected families of $d$ degree $1$ curves.
456: 
457: In both cases we will consider the Yang-Mills amplitude with arbitrary 
458: external scattering states.
459: Via the Penrose transform these scattering states are described by
460: twistor space wavefunctions,\footnote{Actually,
461: the wavefunctions are not defined on all of $\CPP$, but this distinction 
462: will not be important for us.} which are
463: $\overline{\partial}$-closed $(0,1)$ forms $\phi_i$ ($i=1, \dots, n$) on $\CPP$.  We always treat these $\phi_i$ as generic.
464: In our computation, we will be focusing on 
465: poles which arise in integrals over moduli spaces of curves; 
466: we emphasize that the poles in question never come from the $\phi_i$.
467: 
468: The prescriptions as we write them below are not gauge invariant.
469: To make the amplitudes gauge invariant we would
470: probably have to include additional diagrams in both
471: prescriptions, involving cubic Chern-Simons interaction vertices.
472: Nevertheless, both prescriptions make sense provided 
473: we choose a specific gauge for
474: the gauge field, such as an axial gauge.
475: In this gauge one expects
476: that the cubic vertices do not contribute \cite{Witten:2003nn}.\footnote{We
477: thank Peter Svr\v{c}ek for reminding us of this point.}
478: 
479: \subsection{Connected prescription} \label{review-connected}
480: 
481: We first review the connected prescription for computation
482: of $n$-point Yang-Mills amplitudes.
483: The amplitude is obtained as an integral over degree $d$ maps 
484: \begin{equation}
485: P: \C\PP^1 \to \CPP.
486: \end{equation}
487: Such a map $P$ can be written explicitly, in terms of the inhomogeneous
488: coordinate $\sigma$ on $\C\PP^1$, as
489: \eqn{polyno}{
490: P^\IA(\sigma)=Z^\IA = \sum_{k=0}^d \beta^\IA_k \sigma^k
491: }
492: The supermoduli of the degree $d$ map $P$
493: are $\beta^\IA_k$; these span a space $\C^{4d+4|4d+4}$, which comes
494: equipped with the natural measure 
495: \begin{equation}
496: \mu_d = \prod_{k,\IA} \de \beta^\IA_k.
497: \label{natural}
498: \end{equation}
499: We also have a holomorphic $n$-form on $(\C\PP^1)^n$ given by the free-fermion correlator, 
500: \begin{equation} \label{free-fermions}
501: \omega(\sigma_1, \dots, \sigma_n) = 
502: \prod_{i=1}^n \frac{\de\sigma_i}{\sigma_i - \sigma_{i+1}},\qquad
503: \sigma_{n+1}\equiv \sigma_1.
504: \end{equation}
505: 
506: Note that both $\mu$ and $\omega$ are invariant under the group
507: $GL(2,\C)$ that acts linearly on the homogeneous coordinates on $\C\PP^1$. 
508: Its action on $\sigma$ is given by the usual expression
509: \eqn{modular-sigma}{ \sigma \mapsto \sigma' = \frac{a\sigma + b}{c\sigma+ d},
510: \qquad
511: ad-bc\neq 0}
512: while its action on $\beta^\IA_k$ is dictated by the invariance of $Z^\IA$
513: in \eqref{polyno}: the coefficients $\beta^\IA_k$ may be reorganized 
514: (up to some combinatorial factors suppressed for simplicity)
515: into a rank $d$ tensor under $GL(2,\IC)$,
516: \eqn{ranktensor}{ \{\beta^\IA_k\} = \{\beta^\IA_{I_1 I_2 \dots I_d} \},
517: \qquad
518: I_l = 1,2,}
519: where the number of indices $I_l=2$ equals $k$,
520: so that the action of $GL(2,\C)$ on $\beta^\IA_k$ becomes
521: \eqn{action-on-beta}{ \beta^\IA_{I_1 I_2 \dots I_d}
522: \mapsto {\beta'}^\IA_{I_1 I_2 \dots I_d} =
523: M_{I_1}^{\,I'_1}
524: M_{I_2}^{\,I'_2}\dots
525: M_{I_d}^{\,I'_d}
526: \beta^\IA_{I'_1 I'_2 \dots I'_d},\quad
527: M_{I}^{\,I'}=\left(\!\!
528: \begin{array}{rr}d&-b\\ -c&a\end{array}
529: \right)
530: .}
531: Along with $\mu$ and $\omega$ we also have to include the external wavefunctions,\footnote{We write
532: $\phi(P(\sigma_i))$ for the pullback of $\phi$ to moduli space via
533: the evaluation map sending $P$ to $P(\sigma_i)$.}
534: \begin{equation}
535: \Phi = \prod_{i=1}^n \phi_i(P(\sigma_i)).
536: \end{equation}
537: Putting everything together,
538: the Yang-Mills amplitude is formally\footnote{Here and below,
539: by $\mathrm{vol}(GL(2,\C))$ we really mean the volume form
540: on that group; this is just the standard quotient, when written
541: in terms of an integral over the quotient space.}
542: \begin{equation} \label{formal-yma}
543: \int_{\Mstab{n}{d}} 
544: \frac{\mu_d \wedge \omega(\sigma_1, \dots, \sigma_n)}{\mathrm{vol}(GL(2,\C))}
545: \wedge \Phi.
546: \end{equation}
547: 
548: The expression \eqref{formal-yma} is formal for several reasons.
549: The first and most serious reason is that we have to choose a contour
550: for the integral over the coordinates $\beta^\IA_k$ in 
551: $\Mstab{n}{d}$, and the proper choice of contour
552: is not yet well understood. (We do not have
553: to choose a contour for the integrals over $\sigma$,
554: because the integrand includes both $\de \sigma$ from $\omega$
555: and $\de \bar{\sigma}$ from the external wavefunctions.)
556: We will have more to say about the contour below; to match the
557: disconnected prescription we will 
558: essentially use a contour around infinity (suitably defined) so that {\it all}
559: residues are counted.
560: 
561: Second, we have to divide out by the action of $GL(2,\C)$.
562: A convenient gauge-fixing will be chosen below, but of course the amplitude
563: is independent of the choice of gauge.
564: We should perhaps mention that we consider $GL(2,\C)$ over $\C$, {\it i.e.}\  
565: we divide by the ``holomorphic'' volume form.  This means that
566: \begin{itemize}
567: \item this symmetry will always be fixed by a set of holomorphic
568: conditions; 
569: \item we will sum over all inequivalent solutions;
570: \item only the holomorphic Jacobian will be included in the integrals.
571: \end{itemize}
572: These rules are compatible with the computations of
573: \cite{Roiban:2004vt,Roiban:2004ka,Roiban:2004yf}.
574: 
575: 
576: 
577: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
578: 
579: 
580: \vspace{3mm}
581: 
582: \begin{figure}[t]
583: \begin{center}
584: \epsfig{file=twistor1.eps,width=140mm}
585: \end{center}
586: \caption{A contribution to Yang-Mills amplitudes with $5$ positive
587: and $5$ negative helicity gluons, represented
588: {\bf (a)} as four disconnected lines in twistor space,
589: {\bf (b)} as a graph $\Gamma$ with four MHV vertices.
590: }
591: \label{fig-mhv-tree}
592: \end{figure}
593: 
594: 
595: 
596: \subsection{Disconnected prescription} \label{review-disconnected}
597: 
598: Now we describe the disconnected prescription for the same amplitudes, 
599: formulated in twistor space along the lines of
600: the derivation given in \cite{Cachazo:2004kj}.
601: In this prescription a tree level amplitude involving $d+1$ negative helicity
602: gluons, with a particular cyclic ordering, is obtained as a sum over
603: various tree diagrams with $d$ vertices.
604: In Figure \ref{fig-mhv-tree} we show a representative example of a 
605: diagram $\Gamma$ which contributes to
606: amplitudes with $5$ positive and $5$ negative helicity gluons.
607: The $10$ external gluons are arranged cyclically around the index loop,
608: and since there are $5$ negative helicity gluons there are $5-1=4$ vertices.
609: The vertices have arbitrary
610: valence.\footnote{Ultimately, it turns out that any diagram containing
611: a vertex of valence $\le 2$ does not contribute to the amplitude \cite{Cachazo:2004kj}.}
612: We have not specified which gluons have which helicities;
613: the twistor space computation yields superspace expressions
614: which generate the answers for all possible choices
615: when suitably expanded in the fermionic coordinates.
616: 
617: \vspace{3mm}
618: 
619: \begin{figure}
620: \begin{center}
621: \epsfig{file=tree-circles.eps,width=70mm}
622: \end{center}
623: \caption{A different version of Figure \ref{fig-mhv-tree},
624: representing the same single-trace amplitude
625: with the index line made manifest. The circles represent
626: degree 1 curves in twistor space.}
627: \label{fig-mhv-tree-index-line}
628: \end{figure}
629: 
630: Each vertex of $\Gamma$ corresponds to a $\C\PP^1$ in $\CPP$, equipped with
631: marked points corresponding to internal or external lines attached to the vertex.
632: To compute the contribution of $\Gamma$ to the amplitude we have
633: to integrate over the moduli of these curves,
634: given by $d$ degree $1$ maps
635: \begin{equation}
636: Q_i: \C\PP^1 \to \CPP.
637: \end{equation}
638: Each such map can be written
639: \begin{equation}
640: Q_i^\IA(\sigma) = \sum_{k=0}^1 \beta^\IA_{k,i} \sigma^k
641: \end{equation}
642: so there are a total of $8d|8d$ supermoduli $\beta^\IA_{k,i}$ for these $d$ maps,
643: reduced to $4d|8d$ by the $GL(2,\C)^d$ symmetry.  We also have to integrate over the
644: moduli for the marked points; if in the diagram $\Gamma$ there
645: are $n_i$ marked points on the $i$-th $\C\PP^1$, 
646: then the full moduli space is
647: \begin{equation}
648: \Mlinesgamma = \prod_{i=1}^d \M_{0,n_i,1}(\CPP). 
649: \end{equation}
650: 
651: As in the connected case there is a natural measure
652: for the moduli of the curves,
653: \begin{equation}
654: \mulines = \prod_{k,\IA,i} \de \beta^\IA_{k,i}.
655: \end{equation}
656: There are several factors in the integrand which depend on the marked points.
657: First, there is a free-fermion correlator for each curve; the points on the $i$-th
658: $\C\PP^1$ come with a cyclic ordering as indicated in Figure \ref{fig-mhv-tree}, and if
659: we label them $\sigma_1, \dots, \sigma_{n_i}$, they contribute
660: \begin{equation}
661: \omega_i = \omega(\sigma_1, \dots, \sigma_{n_i})
662: \end{equation}
663: with $\omega$ defined in \eqref{free-fermions}.
664: These free-fermion correlators contain $\de \sigma$ for each
665: marked point.
666: 
667: Next we have to include the external wavefunctions:
668: each external wavefunction $\phi_j$ is connected
669: to a marked point $\sigma$ on the $i$-th $\C\PP^1$,
670: for some $i$, and the integrand includes the factor 
671: \begin{equation} \label{external}
672: \phi_j(Q_i(\sigma))
673: \end{equation}
674: just as in the connected prescription.
675: But unlike the connected prescription, here we also have some marked points
676: which are connected to internal propagators.
677: Let us write $D(\cdot, \cdot)$ for the twistor space propagator,
678: which is a $(0,2)$-form on $\CPP \times \CPP$.
679: Each internal propagator is connected to two
680: marked points $\sigma$, $\sigma'$ on the $i$-th
681: and $i'$-th $\C\PP^1$'s respectively,
682: for some $i$, $i'$, and contributes to the integrand a factor
683: \begin{equation} \label{internal}
684: D(Q_i(\sigma), Q_{i'}(\sigma')).
685: \end{equation}
686: Let us write $\Phi \wedge D$ for the product of all the wavefunctions
687: and propagators from \eqref{external}, \eqref{internal}.
688: Since every marked point is attached either
689: to a propagator or to an external wavefunction, this $\Phi \wedge D$
690: includes one factor $\de \bar{\sigma}$ for each marked point.
691: 
692: Then the amplitude in the disconnected prescription is
693: given by the sum over tree diagrams,
694: \begin{equation}
695: \sum_{\Gamma} \int_{\Mlinesgamma} 
696: \frac{\mulines
697: \wedge \left(\prod_{i=1}^d \omega_i \right) 
698: \wedge \Phi\wedge D}{\mathrm{vol}(GL(2,\IC))^d}
699: \label{discpr}.
700: \end{equation}
701: 
702: As with the connected prescription, to make this integral
703: concrete we have to do two more things.
704: First, we must gauge-fix the symmetry $GL(2,\C)^d$ which
705: acts separately on each $\C\PP^1$.  Second, we must choose
706: a contour for the integrals over the moduli $\beta^\IA_{k,i}$.
707: 
708: In \cite{Cachazo:2004kj} it was argued that if one makes a particular choice
709: of contour, and chooses external wavefunctions corresponding to gluons of fixed helicity
710: and momentum, then the integral over $\Mlinesgamma$ in \eqref{discpr} can be evaluated by a simple rule.
711: Namely, one first assigns $(+)$ and 
712: $(-)$ helicities to the endpoints of each propagator,
713: consistent with the rule that each vertex should have exactly two $(-)$ helicities on it;
714: for given $\Gamma$, 
715: there is at most one way to do this.  (If there is no way to do it, then the diagram $\Gamma$
716: just contributes zero.)  Then each vertex gives a copy of the MHV 
717: amplitude --- continued off-shell in a specific way to accommodate the internal lines ---
718: while each propagator carrying momentum $q$ gives $1/q^2$.
719: 
720: For future use in section \ref{csw-generalized} we also mention a natural
721: generalization of the disconnected prescription:  instead of using $d$ degree $1$ curves
722: we could use $K+1$ curves for some $K$, with total degree $d$, 
723: connected into a tree by $K$ propagators.  The integrand is then defined in a way
724: precisely analogous to \eqref{discpr}, except that the sum over $\Gamma$ includes 
725: all choices for the degrees of the curves in addition to distributions of the marked points.
726: 
727: 
728: 
729: 
730: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
731: 
732: \section{Matching the prescriptions in degree $2$ case}
733: 
734: 
735: \subsection{The argument in degree 2 case}
736: 
737: How can the disconnected and connected prescriptions give the same result?
738: Let us consider next-to-maximally helicity violating amplitudes, $q=3$,
739: which come from degree $2$ curves.  We postpone the discussion
740: of curves of higher degree to section 
741: \ref{higherdegree}.
742: 
743: The contribution of disconnected
744: instantons comes from pairs of degree $1$ curves connected by a
745: single propagator, with $n$ marked points distributed over the pair of
746: curves.  This moduli space has dimension $(8+n)|16$ (which includes
747: $4|8$ for each degree $1$ curve
748: plus $n$ for the marked points.)
749: Different distributions of the marked points correspond to
750: different MHV diagrams $\Gamma$.\footnote{There 
751: are $n(n+1)/2$ such diagrams, although once
752: we fix the external wavefunctions not every diagram gives a nonzero contribution
753: to the sum \eqref{discpr}; 
754: if the helicities are $---+++ \cdots ++$, then there are $2(n-3)$
755: diagrams which contribute.}
756: 
757: \begin{figure}[t]
758: \begin{center}
759: \epsfig{file=branches2.eps,width=145mm}
760: \end{center}
761: \caption{A degenerate configuration of two intersecting lines
762: in $\CPP$ can be deformed into a smooth connected curve
763: of degree 2 or into two disconnected lines.
764: The transition between the two branches of
765: moduli space is reminiscent of
766: a conifold transition.
767: }
768: \label{branches}
769: \end{figure}
770: 
771: It was shown in \cite{Cachazo:2004kj} that for each $\Gamma$ the
772: integrand in \eqref{discpr} has a simple pole on the submoduli space $\Mintgamma$,
773: parameterizing degenerate configurations of intersecting lines of degree 1.  
774: This submoduli space has dimension $(7+n)|12$, because
775: the condition that there exists an intersection in the bosonic space 
776: removes one bosonic modulus, and the condition that all four fermionic 
777: coordinates of the two lines coincide at this point removes four fermionic 
778: moduli.\footnote{This fermionic delta-function guarantees the opposite helicity of
779: the two endpoints of the propagators when one expands in fermions to evaluate a 
780: particular amplitude.}
781: 
782: After contour-integrating to localize to $\Mintgamma$, the sum \eqref{discpr} can be written as
783: %
784: \eqn{dtwoi}{
785: \sum_{\Gamma} \int_{\Mintgamma} 
786: \frac{1}{\mathrm{vol}(GL(2,\C))^{2}}
787: \left( \muint \wedge
788: \left( \prod_{i=1}^{n_1} \frac{\de\sigma_i}{\sigma_i - \sigma_{i+1}} \right)
789: \wedge
790: \left( \prod_{j=1}^{n_2} \frac{\de\sigma_j'}{\sigma_j' - \sigma_{j+1}'} \right) \right)
791: \wedge \Phi.
792: }
793: %
794: Here $i$ and $j$ run over the marked points on each $\C\PP^1$,
795: including the point of intersection;
796: so for a diagram with $m$ external wavefunctions
797: attached to the first line, $n_1 = m + 1$
798: and $n_2 = n - m + 1$.  
799: Also, $\sigma_{n_1+1} \equiv \sigma_1$ and $\sigma'_{n_2+1} \equiv \sigma'_1$.
800: The measure $\muint$ is completely determined by the symmetries
801: of $\CPP$.
802: 
803: On the other hand, from the connected
804: prescription \eqref{formal-yma} we find
805: %
806: \eqn{dtwoc}{
807: \int_{\Mstab{n}{2}}
808: \frac{1}{\mathrm{vol}(GL(2,\C))}
809: \left( \mu_2 \wedge
810: \left( \prod_{i=1}^{n} \frac{\de\sigma_i}{\sigma_i - \sigma_{i+1}} \right) \right)
811: \wedge 
812: \Phi.
813: }
814: %
815: We will reorganize the integral \eqref{dtwoc} over the $(8+n)|12$-dimensional
816: space $\Mstab{n}{2}$ of conics in the following way:
817: Locally, to any conic we will associate a pair of intersecting lines which 
818: are its ``asymptotes.''  The moduli space of pairs of intersecting lines 
819: with $n$ marked points is  
820: the $\Mint$ which occurred in the disconnected prescription.  
821: This $\Mint$ has 
822: dimension $(7+n)|12$, so in $\Mstab{n}{2}$ there is one more coordinate, which we call $C$;
823: $C=0$ corresponds to the singular conics, which coincide with their
824: asymptotes. 
825: This $C$ can be thought of as a ``deformation parameter'' which resolves the singularity.
826: We will find that the integrand has a pole at $C=0$, {\it i.e.}\ along
827: $\Mint$.
828: 
829: More precisely, $\Mint$ includes only those degenerations in which the marked
830: points are distributed 
831: in a way corresponding to some MHV tree graph $\Gamma$.  This just means the points are broken
832: into two groups which are cyclically ordered --- so e.g.\ if $n=6$, there is a component of $\Mint$ with points $1,2,3$ 
833: on one line and $4,5,6$ on the other, but we do not include the degeneration which has $1,2,4$ on one line and $3,5,6$ on the other.  Indeed, we will see that the latter degeneration 
834: does \ti{not} give a pole.
835: We will find poles only along $n(n+1)/2$ distinct components $\Mintgamma$, 
836: which are in one-to-one correspondence with the diagrams $\Gamma$ contributing to \eqref{dtwoi}.
837: 
838: Moreover, we will show that the residue along $\Mintgamma$ is precisely
839: such that the integral \eqref{dtwoc} agrees with \eqref{dtwoi} after localizing.
840: This will complete the argument for the equivalence in the degree $2$ case.
841: 
842: 
843: \subsection{Computing the residue in degree 2 case\label{lower-residue}}
844: 
845: In this section we show that the integral \eqref{dtwoc} over the moduli
846: space $\Mstab{n}{2}$ of genus zero, degree 2 curves in $\CPP$ with $n$ marked points 
847: has a pole at the subspace $\Mint$ describing
848: pairs of intersecting lines, and that
849: it has the desired residue as discussed in the last section.
850: 
851: Let us start by fixing part of the $GL(2,\C)$ symmetry reviewed in section
852: \ref{review-connected}.
853: We use three generators of $GL(2,\C)$ to impose
854: the constraints
855: \eqn{gaugechoice}{
856: P^4(\sigma) = \sigma
857: \mbox{\qquad {\it i.e.}\qquad}
858: (\beta^4_0,\beta^4_1,\beta^4_2) = (0,1,0).
859: }
860: In other words, we are imposing the conditions that the two intersections 
861: of the hyperplane $Z^4\!=\!0$\, with the curve have
862: coordinates\footnote{The point $\sigma=\infty$ can be written
863: as $(1:0)$ in homogeneous coordinates, and therefore is 
864: completely nonsingular.}
865: $\sigma\!=\!0$ and $\sigma\!=\!\infty$, and normalizing
866: the coefficients $\beta^4_{0,1,2}$.
867: There is one more generator of $GL(2,\C)$
868: to be fixed, the matrix
869: \begin{equation}
870: M = \begin{pmatrix}
871: \lambda & 0 \\
872: 0 & \lambda^{-1}
873: \end{pmatrix},
874: \end{equation}
875: which acts as
876: \eqn{unfixed}{\beta^\IA_k \to \lambda^{2-2k} \beta^\IA_k,
877: \qquad
878: \sigma\to \lambda^2 \sigma.
879: }
880: This transformation preserves the gauge choice \eqref{gaugechoice}.
881: 
882: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
883: 
884: \subsubsection{Factors from the measure on the moduli space}
885: 
886: Using the freedom to divide all twistor coordinates $Z^\IA$ by $\sigma$,
887: we can write \eqref{polyno} as
888: \eqn{polynoo}{
889: P^\IA(\sigma)=Z^\IA = \sum_{k=0}^{2} \beta^\IA_{k} \sigma^{k-1} =
890: \beta^\IA_0\sigma^{-1} + \beta^\IA_1 + \beta^\IA_2 \sigma,
891: }
892: which using \eqref{gaugechoice} implies $P^4(\sigma)=1$.  As $\sigma \to \infty$ or $\sigma \to 0$,
893: we can neglect the first or the last term in \eqref{polynoo}, respectively.
894: So \eqref{polynoo} describes a hyperbola that approaches
895: two asymptotic lines in the superspace $\IC^{3|4}$:
896: \eqn{asymplines}{Z^\IA=
897: \beta^\IA_0\sigma^{-1} + \beta^\IA_1,
898: \qquad
899: Z^\IA = \beta^\IA_1 + \beta^\IA_2 \sigma.
900: }
901: These two lines intersect at the point $Z^\IA=\beta^\IA_1$, while
902: $\beta^\IA_0$ and $\beta^\IA_2$ with $\IA\neq 4$
903: are the tangent vectors along 
904: these lines.  It is important that for every conic
905: $\Sigma := P_* (\C\PP^1) \subset \CPP$ we can find a singular
906: conic $\Sigma'$ (a pair of intersecting lines) in $\Mint$
907: defining the asymptotes of $\Sigma$.
908: This rule is not canonical; 
909: it depended on our choice to single out 
910: the points at infinity, {\it i.e.}\ the hyperplane $Z^4\!=\!0$.
911: 
912: We want to express $\Mstab{n}{2}$ locally
913: as a product of $\Mint$ and $\C$, with the extra $\C$ parameterized
914: by the deformation parameter $C$.  What are the appropriate coordinates?
915: The $3|4$ parameters 
916: \eqn{firstmoduli}{\beta^\IA_1,\qquad \IA\neq 4,} 
917: describing the position of the intersection of the asymptotes,  
918: give coordinates on $\Mint$.  The remaining $4|8$ 
919: coordinates on $\Mint$ are the directions of the two asymptotes; 
920: each asymptote gives us $2|4$ moduli.  
921: We want to describe these directions by ``unit vectors'' in a suitable sense.
922: As we approach a generic point of $\Mint$, 
923: $\beta^3_0$ and $\beta^3_2$ are nonzero, and we 
924: may use them to normalize the direction vectors. 
925: In other words, the remaining $2|4$ plus $2|4$ coordinates on $\M_{{\rm int}}$ may be 
926: chosen as
927: \eqn{remainingm}{
928: \frac{\beta^\IA_0}{\beta^3_0}
929: \mbox{\quad and \quad}
930: \frac{\beta^\IA_2}{\beta^3_2},\qquad
931: \IA\in\{1,\,2\,|1',2',3',4'\}.
932: }
933: (Choosing different coordinates on $\Mint$ instead of \eqref{firstmoduli}
934: and \eqref{remainingm} would not change the result below; the only change 
935: would be a $C$-independent Jacobian.)
936: 
937: Looking at our original coordinates on $\Mstab{n}{2}$, we still have
938: two more bosonic components of $\beta$ which are independent of our
939: coordinates on $\Mint$, namely $\beta^3_0$ and $\beta^3_2$ themselves.
940: We also have one unfixed generator of $GL(2,\C)$ given in
941: \eqref{unfixed}.  This generator simply multiplies the ratio 
942: $\beta^3_0 / \beta^3_2$ by $\lambda^4$, so we can use it to fix that
943: ratio to a constant, such as
944: \eqn{gaugechoicee}{\frac{\beta^3_0 }{\beta^3_2} = 1.} 
945: Now having fixed the full $GL(2,\C)$ symmetry we can write the measure $\mu_2$
946: from \eqref{natural} as
947: \begin{equation} \label{measure-deltas}
948: (J/4)\ \prod_{k, \IA} \de \beta_k^\IA\ \delta(\beta_0^3 / \beta_2^3 -1)\ \delta(\beta_0^4)\ \delta(\beta_1^4 - 1)\ \delta(\beta_2^4).
949: \end{equation}
950: Here $J$ is the determinant of the Jacobian matrix of variations of the constraints
951: with respect to the $GL(2,\C)$ generators.
952: If we parameterize the generators of $GL(2,\C)$ by
953: \begin{equation}
954: M = \begin{pmatrix}1+a & b \\ c & 1+d\end{pmatrix}
955: \end{equation}
956: then this matrix is
957: \eqn{fixingmatrix}{
958: \delta
959: \left(
960: \begin{array}{c}
961: \beta^4_0\\
962: \beta^4_1\\
963: \beta^4_2\\
964: \beta^3_0 / \beta^3_2
965: \end{array}
966: \right)
967: =
968: \left(
969: \begin{array}{rrrr}
970: \!\!1&0&0&0\\
971: 0&1&0&0\\
972: 0&0&1&1\\
973: *&\phantom{-}*&\phantom{-}2&-2
974: \end{array}
975: \right)
976: \left(
977: \begin{array}{c}
978: b\\
979: c\\
980: a\\
981: d
982: \end{array}
983: \right)
984: }
985: and hence we get simply
986: \begin{equation}
987: J = -4.
988: \end{equation}
989: The factor $J/4$ in \eqref{measure-deltas} represents
990: $1/\mathrm{vol}(GL(2,\C))$; we had to divide by $4$
991: because the $\Z_4 \subset GL(2,\C)$ generated by
992: \begin{equation} \label{unfixed-discrete-matrix}
993: M = 
994: \left(\begin{array}{rr}
995:   i    & 0 \\
996:   0    & -i
997: \end{array}\right)
998: \end{equation}
999: is left unfixed by our gauge condition.
1000: 
1001: The three delta functions in \eqref{measure-deltas} involving
1002: $\beta_k^4$ just eliminate the integrals over those variables,
1003: imposing \eqref{gaugechoice}.  Let us
1004: also use $\delta(\beta_0^3 / \beta_2^3 -1)$ to eliminate $\beta_0^3$,
1005: imposing \eqref{gaugechoicee}.
1006: Integrating over $\beta_0^3$ gives
1007: a factor $\beta_2^3$, so the measure becomes
1008: \eqn{natura}{
1009: - \beta_2^3\,\de \beta_2^3
1010: \prod_{\IA\neq 4} \de \beta_1^\IA
1011: \prod_{k\in\{0,2\}}  \prod_{\IA\neq 3,4}
1012: \de
1013: \beta_k^\IA.
1014: }
1015: We rewrite this as a measure for
1016: the single transverse coordinate $\beta_2^3$, times
1017: a measure on $\Mint$, for which a full set of $7|12$ coordinates were given 
1018: in \eqref{firstmoduli}, \eqref{remainingm}:
1019: \eqn{naturab}{
1020: \left(- (\beta_2^3)^{1-4} \de \beta_2^3\right)
1021: \times \left( \prod_{\IA\neq 4} \de \beta_1^\IA
1022: \prod_{k\in\{0,2\}}  \prod_{\IA\neq 3,4}
1023: \de
1024: \left(
1025: \frac{\beta_k^\IA}{\beta_k^3}
1026: \right) \right).
1027: }
1028: The extra power $(-4)$ in $(\beta_2^3)^{-4}$ was calculated as
1029: $2_{k=0,2}\times (2_B-4_F)$; the terms $2_B$ and $-4_F$ arise from the
1030: redefined bosonic {\it and} 
1031: fermionic measures involving $\beta_k^\IA$, respectively.
1032: 
1033: The coordinate $\beta_2^3$ is related to 
1034: the deformation parameter $C$ --- we will see
1035: that the natural definition of $C$ is $(\beta_2^3)^{2}$.
1036: The measure
1037: $(\beta_2^3)^{-3} \de \beta_2^3$ occurring in \eqref{naturab} 
1038: will be corrected to
1039: $\de \beta_2^3 / \beta_2^3$ --- the desired pole --- once we include an extra factor
1040: $(\beta_2^3)^{2}$ which comes from the free-fermion correlator $\omega$.
1041: We now turn to the analysis of this factor.
1042: 
1043: 
1044: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1045: 
1046: 
1047: \subsubsection{Factors from the fermion correlator}
1048: 
1049: The integrand \eqref{dtwoc} contains the factor
1050: \begin{equation} \label{free-fermions-copy}
1051: \omega(\sigma_1, \dots, \sigma_n) =
1052: \prod_{i=1}^n \frac{\de\sigma_i}{\sigma_i - \sigma_{i+1}},\qquad
1053: \sigma_{n+1}\equiv \sigma_1.
1054: \end{equation}
1055: We would like to investigate how this form behaves on
1056: conics that are degenerating into a pair of lines ({\it i.e.}\ near $\Mint$.)
1057: The result will be that along $\Mint$, $\omega$ factorizes into a product
1058: of two copies of $\omega$ defined on the two lines separately (with an extra
1059: marked $\sigma$ on each line at the point of intersection), while transverse
1060: to $\Mint$, $\omega$ vanishes like $(\beta_2^3)^{2}$.
1061:  
1062: As the curve degenerates to
1063: a pair of lines, some of the $n$ insertions approach one line and some approach the
1064: other.
1065: We consider the case where
1066: \eqn{firstgroup}{\sigma_1, \dots,\sigma_m}
1067: approach one asymptote while the remaining
1068: $(n-m)$ insertions 
1069: \eqn{secondgroup}{\sigma_{m+1}, \dots, \sigma_n}
1070: approach the other.  This is not the most general choice, since the $\sigma_i$ come
1071: with a fixed cyclic ordering which is built into \eqref{free-fermions-copy}; our
1072: choice is characterized by the fact that as we run through the cyclic ordering we jump from the first
1073: line to the second and back only once.
1074: We will comment on other possibilities at the end.
1075: 
1076: With the $GL(2,\C)$ gauge-fixing we chose above, 
1077: as we approach some point of $\Mint$, the coordinates $\sigma_i$ do not remain finite;
1078: one of the lines is $\sigma \to 0$ while the other line is $\sigma \to \infty$.
1079: So we need to rescale the $\sigma_i$ to get new coordinates $\hat \sigma_i$ on $\Mint$
1080: which label the positions of the marked points;
1081: we define $\hat \sigma_i$ so that 
1082: $Z^\IA$ defined in \eqref{asymplines} remains constant
1083: as $\hat\sigma_i$ is kept fixed and $\beta_0^3,\beta_2^3 \to 0$. 
1084: The correct redefinition is
1085: \eqn{newsigma}{\sigma_i
1086: =
1087: \left\{
1088: \begin{array}{lcl}
1089: (\beta_2^3)^{-1} \hat\sigma_i &\mbox{~for~}& i\in\{1,2,\dots m\}\\
1090: \beta_0^3 (\hat\sigma'_i)^{-1} &\mbox{~for~}& i\in\{m+1, 
1091: m+2,\dots n\}\\
1092: \end{array}
1093: \right\}.
1094: }
1095: (We use two different symbols $\hat\sigma_i$ and $\hat\sigma'_i$
1096: to distinguish the coordinates on the two different lines.)
1097: Rewriting $\omega$ from \eqref{free-fermions-copy} in terms of
1098: $\hat\sigma_i$ and $\hat\sigma'_i$, we obtain
1099: \eqn{newsigmaomega}{
1100: \omega(\hat\sigma_1, \dots \hat \sigma'_n)
1101: =
1102: \beta_0^3\beta_2^3
1103: \left(\prod_{i=1}^{m-1}
1104: \frac{\de \hat\sigma_i}{\hat\sigma_i-\hat\sigma_{i+1}}\right)
1105: \frac{\de \hat\sigma_m}{\hat\sigma_1\hat\sigma_m}
1106: \left(\prod_{i=m+1}^{n-1}
1107: \frac{\de \hat\sigma'_i}{\hat\sigma'_i-\hat\sigma'_{i+1}}
1108: \right)
1109: \frac{\de\hat\sigma'_n}{\hat\sigma'_{m+1}\hat\sigma'_n}+ \dots
1110: }
1111: where the intersection was defined to be at $\hat\sigma=\hat\sigma'=0$.
1112: The dots in \eqref{newsigmaomega}
1113: indicate terms suppressed by powers of $\beta_0^3\beta_2^3$.
1114: 
1115: Most of the powers of $\beta^3_0$ and $\beta^3_2$ have canceled,
1116: but there is an extra factor of $\beta_0^3\beta_2^3$,
1117: which equals $(\beta_2^3)^2$ because of our gauge
1118: choice \eqref{gaugechoicee}.  Also, we obtained the expected free fermion contractions, 
1119: including the $2+2$ 
1120: contractions 
1121: involving the intersection of the
1122: two lines at
1123: $\hat\sigma=\hat\sigma'=0$.
1124: 
1125: Note that $\beta_0^3$ and $\beta_2^3$ always appeared in the combination
1126: \eqn{ccombi}{C=\beta_0^3\beta_2^3}
1127: that is invariant under 
1128: \eqref{unfixed}.
1129: This is the same $C$ that we used 
1130: in Figure \ref{branches};
1131: in fact, one can rewrite our curve in the form
1132: \eqn{xyc}{xy=C}
1133: where $x$,$y$ are coordinates on a plane in $\CPP$.
1134: The limit $C\to 0$ describes the singular conics.
1135: Note that it is $C$ rather than $\beta_2^3$ that is a good coordinate ---
1136: this is because a simultaneous sign flip on $\beta_0^3$ and $\beta_2^3$ 
1137: is the gauge transformation 
1138: \eqref{unfixed} with $\lambda = i$,
1139: which preserves our gauge choices \eqref{gaugechoicee}.
1140: 
1141: Finally, it is easy to check that if we choose a different distribution of
1142: the marked points, the result comes out suppressed by additional powers of $C$.
1143: We are only interested in the leading terms, which are linear in $C$ and will give 
1144: the coefficient of \,$\de C / C$.
1145: 
1146: 
1147: \subsection{Finishing the proof in degree 2 case} \label{finishing}
1148: 
1149: Now we can collect the results from the previous two 
1150: subsections.  
1151: The powers of $\beta_2^3$ from 
1152: \eqref{naturab} and \eqref{newsigmaomega} combine to give
1153: $\smallint \de \beta_2^3 / \beta_2^3$, which is proportional to $\smallint
1154: \de C / C$.  So as advertised, the integral \eqref{dtwoc} localizes
1155: after contour integration to an integral over $\Mint$.  The symmetries of 
1156: $\CPP$ determine the measure for the moduli of the two lines in $\Mint$, which therefore agrees
1157: with the measure $\muint$ in \eqref{dtwoi} up to an overall constant; as for
1158: the integral over the marked points, comparing \eqref{newsigmaomega} and
1159: \eqref{dtwoi} we see that these measures are also identical.  This completes
1160: the argument for equivalence in the $d=2$ case.
1161: 
1162: Incidentally, one can also compare the measures on $\Mint$ directly, without
1163: recourse to a symmetry argument.  
1164: We have already computed
1165: the measure which arises from the connected prescription, in \eqref{naturab},
1166: so the job is to compute the measure $\muint$ which arises from the disconnected
1167: prescription.  This computation is given (in greater generality) in section \ref{finishing-higher}.
1168: 
1169: 
1170: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1171: 
1172: \section{Higher degree \label{higherdegree}}
1173: 
1174: Now let us consider the connected prescription for general degree $d$.
1175: We will see that the fully disconnected description and the fully connected
1176: prescription are not only equivalent, they are
1177: just two extreme cases of a more general class of
1178: rules to calculate the amplitude.
1179: We will find $d$ \ti{a priori} different expressions for
1180: the scattering amplitude with $d\!+\!1$ negative-helicity
1181: gluons,
1182: \eqn{ways}{{\cal A}_{[K]},\qquad K=0,1,2,\dots , d-1,}
1183: where $K\!+\!1$ denotes the total number of curves involved in the 
1184: prescription.\footnote{Later we will see that $K$ also represents the codimension in moduli space on which
1185: the prescription is localized, or equivalently the number of internal propagators
1186: which appear in the prescription.}
1187: 
1188: The organization of this section is as follows:
1189: 
1190: \begin{itemize}
1191: 
1192: 
1193: \item subsection \ref{plethora} outlines the argument that
1194: the completely connected and completely disconnected prescriptions
1195: agree;
1196: 
1197: \item subsection \ref{csw-generalized} discusses the intermediate prescriptions
1198: with arbitrary $K$ and their diagrammatic interpretation;
1199: 
1200: \item subsection \ref{higher-residue} generalizes the residue
1201: calculation of subsection \ref{lower-residue} to the case of a degree $d$
1202: curve splitting into two curves of degrees $d_1$ and $d_2$;
1203: 
1204: \item subsection \ref{finishing-higher} shows that the residues occurring
1205: for any degeneration are actually independent of the chosen prescription, 
1206: completing the argument.
1207: 
1208: \end{itemize}
1209: 
1210: 
1211: \subsection{The proof in higher degree case\label{plethora}}
1212: 
1213: Rather than showing directly that the connected prescription arising from a single
1214: connected degree $d$ curve is equivalent to the disconnected
1215: prescription involving $d$ lines, we will first show that it is
1216: equivalent to a computation 
1217: involving two disconnected components of degrees $d_1$, $d_2$, such that 
1218: \begin{equation} \label{dd}
1219: d_1 + d_2 = d.
1220: \end{equation}
1221: The proof is a 
1222: generalization of the computation we did in section \ref{lower-residue}:  
1223: namely, in subsection \ref{higher-residue} we will find a pole 
1224: on each boundary divisor $\Mintgamma$, corresponding to a degeneration into
1225: intersecting curves,
1226: \begin{equation}
1227: \Sigma_d \longrightarrow \Sigma_{d_1} \cup \Sigma_{d_2},
1228: \quad \quad \quad d_1 + d_2 = d,
1229: \label{dddcurves}
1230: \end{equation}
1231: with a particular distribution of the marked points.  
1232: 
1233: \begin{figure}[t]
1234: \begin{center}
1235: \epsfig{file=d3curves.eps,width=120mm}
1236: \end{center}
1237: \caption{A degeneration of a degree 3 curve into three intersecting
1238: lines can be viewed as a two-step process.
1239: The moduli space of degree 3 maps with 5 marked points, $\Mstab{5}{3}$,
1240: contains divisors, $\M^{\Lambda_1}_{{\rm int}}$ and $\M^{\Lambda_2}_{{\rm int}}$,
1241: associated with degenerations into a degree 2 curve and a line,
1242: shown at the intermediate stages. The moduli space $\Mintgamma$
1243: of three intersecting lines (shown in the lower right corner) can be
1244: identified with the intersection $\M^{\Lambda_1}_{{\rm int}} \cap \M^{\Lambda_2}_{{\rm int}}$.}
1245: \label{d3curves}
1246: \end{figure}
1247: 
1248: Next we want to show iteratively that
1249: this integral over curves with $2$ irreducible components is equivalent to one over curves with $3$ components, and so on until eventually we reach $d$ components (all of which must
1250: have degree $1$.)  The idea which makes this iteration possible is the following:  
1251: consider some locus
1252: $\Mintgamma$, corresponding to a particular degeneration of $\Sigma$ into
1253: $K\!+\!1$ components, with a particular distribution of the marked points.
1254: This locus
1255: can be obtained as an intersection of $K$ boundary divisors, $\M^{\Lambda_j}_{{\rm int}}$,
1256: each of which is associated with a degeneration of $\Sigma_d$
1257: into two irreducible components,\footnote{We use $\Gamma$ to denote a general 
1258: degeneration into $K\!+\!1$ components, and $\Lambda$ to denote a degeneration into just two components.}
1259: \begin{equation}
1260: \Mintgamma = \M^{\Lambda_1}_{{\rm int}} \cap \cdots \cap \M^{\Lambda_{K}}_{{\rm int}}.
1261: \label{mintascap}
1262: \end{equation}
1263: An example is shown in Figure \ref{d3curves}.  
1264: In this sense, the problem 
1265: of studying a general degeneration boils down to
1266: understanding the basic process \eqref{dddcurves}.
1267: 
1268: So let's start with the integral over 
1269: $K$-component curves and try to prove it agrees
1270: with an integral over $(K+1)$-component curves.  
1271: In the $K$-component case we
1272: have to integrate over various loci $\Mintgamma$ as in
1273: \eqref{mintascap}.
1274: Since the various divisors $\Mintlambda$ meet transversally \cite{MR98m:14025},
1275: in integrating over each such $\Mintgamma$ we will encounter poles
1276: wherever $\Mintgamma$ intersects another divisor $\Mintlambda$.\footnote{One
1277: way to understand this is to note that if we start with
1278: the full $\Mstab{n}{d}$ and look near such an intersection of $K$ divisors,
1279: the integrand looks like
1280: \begin{equation}
1281: \frac{\de C_1}{C_1} \wedge \cdots \wedge \frac{\de C_{K}}{C_{K}}
1282: \wedge ({{\rm regular}}).
1283: \end{equation}
1284: We have already contour-integrated 
1285: over $C_1, \dots, C_{K-1}$ and thus restricted to $C_1 = \cdots = C_{K-1} = 0$, \ti{i.e.}\ 
1286: to $\Mintgamma$; after doing this we get simply $\de C_{K}/C_{K}$, with a pole at $C_K = 0$, 
1287: \ti{i.e.}\ at $\Mintgamma \cap \Mintlambda$.}
1288: We choose our contour so that
1289: it picks up the residues at these poles.
1290: In this way we reduce the integral over $\Mintgamma$ to the sum
1291: of integrals over all intersections $\Mintgamma \cap \Mintlambda$.
1292: Then we have to sum over all $\Gamma$ describing $K$-component degenerations.
1293: What is the result of all this summation? From the perspective
1294: of the $(K\!+\!1)$-component degenerations --- which we label by $\Gamma'$ ---
1295: the answer is clear: given some 
1296: \begin{equation}
1297: \Mintgammaprime = \M^{\Lambda_1}_{{\rm int}} \cap \ldots \cap \M^{\Lambda_{K}}_{{\rm int}},
1298: \label{mintascap2}
1299: \end{equation}
1300: there are $K$ ways to make it by intersecting
1301: some $\Mintgamma$ with some $\M^{\Lambda_i}_{{\rm int}}$.
1302: Therefore we get a sum over all $(K\!+\!1)$-component
1303: degenerations, with an \ti{overall} multiplicative factor $K$.
1304: 
1305: Finally, after repeating this process $d-1$ times, we arrive at an 
1306: integral over the moduli space of connected trees consisting of $d$ lines,
1307: with all possible shapes for the tree and all allowed distributions of
1308: marked points.  But the arguments of \cite{Cachazo:2004kj} show that the disconnected prescription also reduces to such an integral, by a similar process of localization
1309: to poles.  Furthermore, in section \ref{higher-residue}
1310: we will see that the residues in these two computations agree;
1311: this will complete the proof.
1312: 
1313: 
1314: 
1315: \subsection{Intermediate prescriptions} \label{csw-generalized}
1316: 
1317: In subsection \ref{plethora} we encountered $d-1$ different moduli spaces $\MintK$ of singular curves, characterized by the number $K+1$ of components, 
1318: which interpolated between the nonsingular degree $d$ curve ($K=0$) and the tree of degree $1$
1319: curves ($K=d-1$).  Furthermore we obtained integrals over each $\MintK$ by starting with the connected
1320: prescription ($K=0$) and successively localizing to poles.  As a result of this localization
1321: all these integrals are equal; now we want to argue that the intermediate cases $K=1, \dots, 
1322: d-2$ can be naturally interpreted as coming from ``intermediate prescriptions,'' involving
1323: integrals over the moduli of $K+1$ disconnected curves with $K$ propagators connecting them.  
1324: We defined these prescriptions at the end of section \ref{review-disconnected}.
1325: 
1326: The argument is a generalization of the ``heuristic'' derivation of
1327: the computational rules for the disconnected prescription, given 
1328: in \cite{Cachazo:2004kj}.
1329: Namely, starting from the intermediate prescription,
1330: note that the propagator $D(\cdot, \cdot)$ by definition satisfies
1331: \begin{equation} \label{prop-def}
1332: \bar{\partial} D = \Delta.
1333: \end{equation}
1334: Here $\Delta$ is a $(0,3)$-form on $(\CPP)^2$
1335: which is concentrated on the diagonal $\CPP$:
1336: in inhomogeneous coordinates with $Z^4 = Z'^4 = 1$ it may be written
1337: \begin{equation}
1338: \Delta = \deltabar(Z^1 - Z'^1)\,\deltabar(Z^2 - Z'^2)\,
1339: \deltabar(Z^3 - Z'^3)\,\delta^{4}(\psi - \psi'),
1340: \qquad
1341: \deltabar(f) := \delta^2(f) \de\bar{f}.
1342: \end{equation}
1343: The equation \eqref{prop-def} means that $D(\cdot, \cdot)$
1344: is meromorphic with a pole along the diagonal.  
1345: The integral over $\MintK$ in the disconnected prescription contains $K$
1346: propagators \eqref{internal}; these factors therefore
1347: have poles when $Q_i(\sigma) = Q_{i'}(\sigma')$.  
1348: 
1349: As in \cite{Cachazo:2004kj}, we assume that $K$ of the integrals over
1350: moduli of the disconnected curves are evaluated on contours
1351: which encircle these poles, in a suitable sense.
1352: Using \eqref{prop-def}, performing these contour
1353: integrals is equivalent to filling in the contour and 
1354: replacing $D$ by $\Delta$.
1355: This localizes the integral to the sublocus of moduli
1356: space where all propagators have shrunk to zero length,
1357: which is exactly $\MintK$.
1358: 
1359: So finally we have $d$ different prescriptions, involving summing 
1360: over configurations with $1$ curve (connected case), $2$, $3$, \dots, $d$ curves (maximally disconnected case); and we have argued that each of these prescriptions is equivalent, up to an overall rescaling.  In this sense any of them can be used to calculate the Yang-Mills
1361: amplitudes.
1362: 
1363: Of course, another possibility is that the correct amplitudes are obtained by summing
1364: different contributions from various sorts of diagrams with various numbers of curves.  We have argued that all such contributions are proportional to one 
1365: another, so such a modified rule would only change the overall prefactor.
1366: Although we will not try to make the final verdict in this paper, we believe that a more
1367: detailed analysis of the prescriptions (including the coefficients) should be able
1368: to resolve this uncertainty.
1369: 
1370: 
1371: 
1372: \subsubsection{Diagrammatic interpretation and an example}
1373: 
1374: Now let us discuss the diagrammatic interpretation of the
1375: intermediate prescriptions.
1376: We have seen that the $K$-th intermediate prescription is naturally localized
1377: on $\MintK$, which is a union of various $\Mintgamma$.  Here $\Gamma$
1378: describes the decomposition of the curve $\Sigma_d$ into $K+1$ components
1379: and the distribution of marked points along these components.  
1380: Equivalently, we could say that
1381: $\Gamma$ describes a slight generalization of an MHV tree diagram:
1382: namely, it is a tree diagram with $K+1$ vertices, where each vertex now carries
1383: an internal index $d_i$, subject to the rule that $\sum d_i = d$.  The MHV
1384: diagrams are the case where all $d_i = 1$.
1385: 
1386: \begin{figure}
1387: \begin{center}
1388: \epsfig{file=twistor3.eps,width=100mm}
1389: \end{center}
1390: \caption{An MHV tree diagram contributing to $\A{+-+---}$.}
1391: \label{twistor3}
1392: \end{figure}
1393: 
1394: It would be very useful if we could give a compact formula for the contribution
1395: of a general vertex with arbitrary $d_i$, analogous to the off-shell
1396: continuation of the MHV amplitude given in \cite{Cachazo:2004kj}.
1397: At the moment we do not possess such a formula, so we can only define the diagram
1398: $\Gamma$ to be the integral over $\Mintgamma$ which we considered above.  In this
1399: language, our localization argument relating different prescriptions 
1400: becomes the statement that the contribution from a diagram $\Gamma$ agrees with
1401: the sum over all $\Gamma'$ obtained by ``splitting a vertex'' in $\Gamma$.  In
1402: other words, $\Gamma'$ should be obtained by replacing a 
1403: vertex with index $d$ by a pair of vertices with indices
1404: $d_1$, $d_2$, such that $d_1 + d_2 = d$, with a propagator connecting them.  This is
1405: the diagrammatic analog of a degree $d$ curve which degenerates into two curves
1406: with degrees $d_1$, $d_2$.
1407: 
1408: We can also repeat the combinatorics from
1409: subsection \ref{plethora} in this language.  Start with a diagram with $K\!+\!1$
1410: vertices.  This diagram contains $K$ propagators.
1411: Therefore there are $K$ ways to shrink a single propagator
1412: and obtain a ``parent'' diagram with $K$ vertices.
1413: Because a diagram with $K\!+\!1$ vertices has $K$ parents, 
1414: the sum over the daughters 
1415: with $K\!+\!1$ vertices equals $K$ times the sum over
1416: the parents with $K$ vertices. 
1417: 
1418: \begin{figure}[t]
1419: \begin{center}
1420: \epsfig{file=twistor2.eps,width=110mm}
1421: \end{center}
1422: \caption{Two types of tree diagram with one MHV and one
1423: non-MHV (degree 2) vertex that contribute to the $\A{+-+---}$ amplitude.
1424: In total, there are six diagrams of each kind.
1425: The number attached to each vertex represents the degree
1426: of the corresponding curve in twistor space.}
1427: \label{twistor2}
1428: \end{figure}
1429: 
1430: To illustrate how all this works when external wavefunctions of fixed
1431: helicity are included, let us consider a 6-gluon amplitude $\A{+-+---}$.
1432: If we were to use the connected prescription, we would have
1433: to integrate over the moduli space $\Mstab{6}{3}$ of degree 3 curves.
1434: On the other hand, in the disconnected prescription one has
1435: to consider three degree 1 curves, which can be interpreted
1436: as MHV vertices in Yang-Mills theory \cite{Cachazo:2004kj}.
1437: Therefore, in this case one has to sum over all tree graphs with
1438: three MHV vertices connected by Yang-Mills propagators --- see Figure \ref{twistor3}.
1439: In total, there are 19 such graphs contributing to $\A{+-+---}$.
1440: 
1441: Now let us consider the intermediate prescription with $K=1$. 
1442: This prescription leads
1443: to a sum over tree graphs with two vertices, one MHV
1444: and one non-MHV (the non-MHV vertex involves three insertions of negative helicity).
1445: Examples of such graphs with non-MHV vertices are shown in Figure \ref{twistor2}.
1446: There are 12 such diagrams which contribute to $\A{+-+---}$.
1447: Since each non-MHV vertex itself can be represented
1448: as a sum over tree diagrams with two MHV vertices, we should be
1449: able to reproduce the disconnected prescription if we split
1450: all non-MHV vertices into MHV ones.
1451: More precisely, in this decomposition we should encounter each
1452: MHV diagram twice (since in the disconnected prescription $K=2$).
1453: Indeed, it is straightforward to check that the 12 non-MHV diagrams
1454: lead to 38 MHV graphs, in agreement with the general rule.
1455: 
1456: 
1457: 
1458: \subsection{Computing the residue in higher degree case\label{higher-residue}}
1459: 
1460: Returning from our digression to discuss the intermediate prescriptions, 
1461: in this section we show that the integral \eqref{formal-yma} over the
1462: moduli space $\Mstab{n}{d}$ which arises in the connected prescription
1463: has a pole along the codimension 1 divisor $\Mintone$ describing
1464: curves that are degenerated into 2 components.  We further verify
1465: that the residue is the same as that which arises after localization of the $K=1$ 
1466: prescription on $\Mintone$, thus establishing the equivalence between connected
1467: and $K=1$ prescriptions.
1468: 
1469: We want to study a degeneration in which 
1470: the curve $\Sigma_d$ degenerates into a pair of intersecting curves,
1471: $\Sigma_{d_1}$ and $\Sigma_{d_2}$, of degree $d_1$ and $d_2$,
1472: as in \eqref{dddcurves}.
1473: Using the projective symmetry to divide by $\sigma^{d_1}$,
1474: we can write the degree $d$ map \eqref{polyno} as
1475: \eqn{polynood}{
1476: Z^\IA(\sigma) = \sum_{k= -d_1}^{d_2} \beta^\IA_{d_1+k} \sigma^k.
1477: }
1478: We fix the $GL(2,\C)$ symmetry similarly to 
1479: the degree 2 case,
1480: namely by conditions based on \eqref{gaugechoice} and \eqref{gaugechoicee}:
1481: \eqn{fixing-d}{
1482: (\beta^4_{d_1-1},\beta^4_{d_1},\beta^4_{d_1+1}) = (0,1,0),\qquad
1483: \frac{\beta^3_{d_1-1}}{\beta^3_{d_1+1}} = 1,
1484: }
1485: and define the deformation parameter $C:= \beta_{d_1-1}^3 \beta_{d_1+1}^3$.
1486: As in degree $2$, the singular limit will be $C \to 0$, or equivalently
1487: $\beta_{d_1+1}^3 \to 0$, and the question is
1488: how the other coefficients should scale in this limit.
1489: 
1490: \begin{figure}
1491: \begin{center}
1492: \epsfig{file=table-beta.eps,width=140mm}
1493: \end{center}
1494: \caption{The organization of the coefficients $\beta^\IA_k$
1495: for a degree $d$ curve degenerating into curves of degrees $d_1$ and $d_2$.
1496: The symmetry $GL(2,\C)$ is fixed by setting three bosonic coefficients to the values
1497: $(0,1,0)$ and two others to $\sqrt{C}$; 
1498: this $C$ is the deformation parameter, which approaches zero in the degeneration
1499: limit.
1500: }
1501: \label{table-beta}
1502: \end{figure}
1503: 
1504: The correct scaling is as follows:  we take $\beta_{d_1-1}^3 = \beta_{d_1+1}^3 \to 0$ while holding
1505: finite the quantities
1506: \eqn{coef-done-dtwo}{
1507: \alpha_k^\IA := \frac{\beta^\IA_{d_1-k}}{(\beta^3_{d_1-1})^k}, \quad 0\leq k \leq d_1; \qquad
1508: \alpha'^\IA_k := \frac{\beta^\IA_{d_1+k}}{(\beta^3_{d_1+1})^k}, \quad 0\leq k \leq d_2.}
1509: In that limit we obtain two curves,
1510: \begin{eqnarray} \label{curves-new-coords}
1511: \Sigma_{d_1}~: && Z^\IA(\hat\sigma) = \sum_{~k= 0~}^{d_1} 
1512: \alpha_k^\IA \hat\sigma^{k}, \nonumber \\
1513: \Sigma_{d_2}~: && Z^\IA(\hat\sigma') = \sum_{~k= 0~}^{d_2} 
1514: \alpha'^\IA_k \hat\sigma^{\prime k}.\label{two-curves-b}
1515: \end{eqnarray}
1516: Namely, we obtain the points $Z^\IA(\hat\sigma)$ on $\Sigma_{d_1}$ by holding fixed
1517: $\hat\sigma = \sigma / \beta^3_{d_1-1}$ in the limit, 
1518: and we obtain the points $Z^\IA(\hat\sigma')$ on $\Sigma_{d_2}$ by holding fixed
1519: $\hat\sigma' = \sigma \beta^3_{d_1+1}$ in the same limit.
1520: See Figure \ref{table-beta}.
1521: 
1522: Therefore the parameters $\alpha_k^\IA, \alpha'^\IA_k$ give coordinates on $\Mintone$,
1523: specifying the moduli of the degenerated curve.
1524: (Note
1525: that $\alpha_0^\IA = \alpha'^\IA_0$; these shared coordinates 
1526: specify the intersection
1527: point of $\Sigma_{d_1}$ and $\Sigma_{d_2}$.)
1528: 
1529: Now we want to study how our integral \eqref{formal-yma} behaves near $\Mintone$.
1530: As in section \ref{lower-residue}, we have to compute the Jacobian $J$ from the
1531: gauge-fixing of $GL(2,\C)$.  The matrix of variations generalizing \eqref{fixingmatrix}
1532: is
1533: \eqn{fixingmatrix-2}{
1534: \delta
1535: \left(
1536: \begin{array}{c}
1537: \beta^4_{d_1-1}\\
1538: \beta^4_{d_1}\\
1539: \beta^4_{d_1+1}\\
1540: \beta^3_{d_1-1} / \beta^3_{d_1+1}
1541: \end{array}
1542: \right)
1543: =
1544: \left(
1545: \begin{array}{ccrr}
1546: d_2&\!\!\!\!(d_1+2)\beta^4_{d_1+2}&0&0\\
1547: \!\!\!\!(d_2+2)\beta^4_{d_1-2}\!\!&d_1 &0&0\\
1548: 0&0&d_1\!&d_2\! \\
1549: *&*&2&-2
1550: \end{array}
1551: \right)
1552: \left(
1553: \begin{array}{c}
1554: b\\
1555: c\\
1556: a\\
1557: d
1558: \end{array}
1559: \right).
1560: }
1561: In the singular limit, the $\beta^4_{d_1 \pm 2}$ terms in \eqref{fixingmatrix-2} 
1562: vanish, and we get
1563: \begin{equation} \label{jac-1}
1564: J \to -2 d_1 d_2 d.
1565: \end{equation}
1566: The gauge-fixed integral includes the factor $J / 2d$; the $2d$ comes from
1567: an unfixed subgroup of $GL(2,\C)$, analogous to \eqref{unfixed-discrete-matrix}, 
1568: which is $\Z_2 \times \Z_d$ if both
1569: $d_1$ and $d_2$ are even and $\Z_{2d}$ otherwise.
1570: Next we have to rewrite the integrand in terms of the new coordinates \eqref{coef-done-dtwo}.
1571: One might be worried that switching to these coordinates
1572: will generate extra powers of $C$ beyond what we had in the degree 2 case,
1573: spoiling the conclusion that there is a pole along $\Mintone$.
1574: But this does not occur; if we increase $d_1$ by 
1575: $1$, for example, the integrand just acquires an extra integral over $4|4$ variables:
1576: \eqn{measure-new}{ \mu \to \mu \wedge \prod_\IA \de \beta^\IA_{0} =
1577: \mu \wedge
1578: \prod_\IA \de \alpha^\IA_{d_1}
1579: }
1580: The powers of $\beta^3_{d_1+1}$ simply cancel between the 4 bosons and 4 fermions!
1581: Unlike the coefficients $\beta^\IA_{d_1}$ and
1582: $\beta^\IA_{d_1\pm 1}$, among which $5$ special bosonic components
1583: have been used to gauge-fix the $GL(2,\C)$ symmetry or to describe the parameter $C$,
1584: the additional moduli $\beta^\IA_{d_1\pm k}$ with $k>1$ come in full ``supermultiplets''
1585: containing 4 bosons and 4 fermions.  Therefore no new powers of $C$ are generated in
1586: rescaling $\beta$'s to $\alpha$'s,
1587: so the measure for the moduli of the degenerating curve still 
1588: behaves as $\de C / C^2$ near $C=0$.  Similarly, the free
1589: fermion correlator $\omega$ factorizes,
1590: \begin{equation}
1591: \omega(\sigma) \to C\,\omega_1(\hat\sigma) \wedge \omega_2(\hat\sigma'), 
1592: \end{equation}
1593: just as in degree $2$.
1594: 
1595: So we have a pole along $\Mintone$, as in the degree $2$ case, and after integrating
1596: around this pole the fully gauge-fixed measure
1597: for the moduli of the degenerate curve is
1598: \begin{equation} \label{measure-degenerate}
1599: -d_1 d_2 \prod_{\IA} \left( {\prodprime{k=0}{d_1}}\! \de \alpha_k^\IA\,{\prodprime{k=1}{d_2}}\! \de \alpha'^\IA_k \right).
1600: \end{equation}
1601: Here the symbol $\Pi'$ indicates that we omit the $5$ factors
1602: \begin{equation} \label{fixed-components}
1603: \de \alpha^4_1,\, \de \alpha'^4_1,\, \de \alpha^4_0,\, \de \alpha^3_1,\, \de \alpha'^3_1;
1604: \end{equation}
1605: there are no such $\alpha$'s among the coordinates on $\Mintone$, because their corresponding $\beta$'s
1606: were used up in the gauge-fixing and in the transverse coordinate $C$, as shown in Figure \ref{table-beta}.  
1607: 
1608: 
1609: 
1610: \subsection{Finishing the proof in higher degree case} \label{finishing-higher}
1611: 
1612: Finally we
1613: have to check that the measure \eqref{measure-degenerate} agrees with the one coming from localization
1614: of the $K=1$ prescription.  From section \ref{csw-generalized} we know that the latter measure
1615: is obtained as follows:  start with two curves of degree $d_1$, $d_2$,
1616: \begin{align} 
1617: Q^\IA(\sigma) &= \sum_{k=0}^{d_1} \alpha^\IA_{k} \sigma^k, \nonumber \\
1618: Q'^\IA(\sigma') &= \sum_{k=0}^{d_2} \alpha'^\IA_{k} \sigma'^k. \label{dcurves}
1619: \end{align} 
1620: (The notation $\alpha^\IA_{k}$, $\alpha'^\IA_k$ we use here
1621: agrees with the notation we
1622: used above for the moduli of the curves obtained by a degeneration; compare \eqref{dcurves} with \eqref{curves-new-coords}, \eqref{coef-done-dtwo}.  The only difference is that here we do not necessarily have $\alpha^\IA_0 = \alpha'^\IA_0$.)
1623: Then we have the standard measure \eqref{natural} on the two curves separately, which before gauge-fixing is
1624: \begin{equation} \label{standard}
1625: \mu_{d_1} \wedge \mu_{d_2} = \prod_{\IA} \left( \prod_{k=0}^{d_1} \de \alpha^\IA_k \prod_{k=0}^{d_2} \de \alpha'^\IA_k \right).
1626: \end{equation}
1627: As explained in section \ref{csw-generalized}, the requirement that the two curves actually intersect is enforced by a delta function which is coupled to one marked point on each curve,
1628: \begin{equation} \label{deltafunc}
1629: \Delta(Q(\sigma), Q'(\sigma')).
1630: \end{equation}
1631: To compare the measures (including this delta function) we have to gauge-fix
1632: the $GL(2,\C)^2$ symmetry acting on the coefficients $\alpha^\IA_k, \alpha'^\IA_k$.
1633: There are many ways to do this;
1634: we choose a way that is as similar as possible to the gauge-fixing we used for the degenerating
1635: degree $d$ curve, so that the unfixed moduli will match directly.
1636: Namely, we take
1637: \begin{align} \label{gauge-conditions1}
1638: \alpha^i_0 &= \alpha'^i_0 \quad \textrm{for}\ i \in \{2,3\}, \\
1639: \alpha^4_0 &= \alpha'^4_0 = 1, \label{gauge-conditions2} \\
1640: \alpha^4_1 &= \alpha'^4_1 = 0, \\
1641: \alpha^3_1 &= \alpha'^3_1 = 1.
1642: \end{align}
1643: The matrix of variations from this gauge-fixing is similar to \eqref{fixingmatrix-2},
1644: but since it is an $8 \times 8$ matrix we just write the answer here:
1645: \begin{equation} \label{jac-2}
1646: J = (d_1 d_2)^2 (\alpha^2_1 - \alpha'^2_1).
1647: \end{equation}
1648: The gauge-fixing factor is $J / {d_1 d_2}$, because of the subgroup $\Z_{d_1} \times \Z_{d_2} \subset
1649: GL(2,\C) \times GL(2,\C)$,
1650: roots of unity acting on each curve separately; since this subgroup
1651: acts trivially it is unfixed by our gauge condition. 
1652: Next we must include the integral over the delta function \eqref{deltafunc}, which we write as
1653: \begin{equation}
1654: \int \de \sigma\ \de \sigma'\ \delta^{(3|4)} \left( \frac{Q^\IA(\sigma)}{Q^4(\sigma)} - \frac{Q'^\IA(\sigma')}{Q'^4(\sigma')} \right).
1655: \end{equation}
1656: With our gauge choice, it is easy to study the behavior of this delta function in the vicinity
1657: of $\sigma = \sigma' = 0$.\footnote{Although our gauge choice was rigged so that studying 
1658: $\sigma = \sigma' = 0$ would recover the desired moduli space of intersecting curves, 
1659: it is not clear \ti{a priori} 
1660: from our arguments why one should consider only this region; this is related to the issue of the
1661: exact contour choice in the intermediate prescription, which we will not settle here.  We are also integrating over the delta function as if it were real instead of holomorphic;
1662: similar manipulations were used in \cite{Witten:2004cp}.}  
1663: One uses the $Z^2$ and $Z^3$ components of the delta function to set $\sigma = \sigma' = 0$, 
1664: obtaining
1665: \begin{equation} \label{deltas}
1666: \frac{1}{(\alpha^2_1 - \alpha'^2_1)}\ \delta(\alpha^1_0 - \alpha'^1_0) \prod_{\IA=1'}^{4'} \delta(\alpha^\IA_0 - \alpha'^\IA_0).
1667: \end{equation}
1668: Note that the $1|4$ delta functions in \eqref{deltas}, combined with the gauge conditions \eqref{gauge-conditions1}, \eqref{gauge-conditions2}, 
1669: are enough to set all $\alpha'^\IA_0 = \alpha^\IA_0$.  This was
1670: the main motivation for this gauge-fixing; the point $\alpha^\IA_0$ represents the intersection 
1671: of the two curves, and the remaining moduli are precisely the ones we had for the degenerating degree
1672: $d$ curve in \eqref{measure-degenerate}.  Therefore we easily see that the measures agree, including
1673: the prefactor $d_1 d_2$.
1674: (Although we have not been careful about overall constant factors,
1675: the absence of a relative factor here is important --- it corresponds to the absence of prefactors
1676: weighing different diagrams in the intermediate prescriptions.)
1677: 
1678: This completes the argument for the equivalence between the connected and $K=1$ prescriptions.
1679: It also sets up the iteration we described in section \ref{plethora} to prove
1680: the equivalence of all prescriptions, by successive localization to poles in higher
1681: and higher codimension, corresponding to more and more degenerate curves.
1682: 
1683: One detail remains:
1684: we have to check that the residues we obtain are always independent of which
1685: prescription we started with.  
1686: In other words, we have to prove that
1687: the measure for the integral over $K+1$-component trees obtained by some degeneration process
1688: always agrees with the measure coming from the disconnected prescription.  
1689: As we know from section \ref{csw-generalized}, the latter measure
1690: can be written as a product of measures for the individual curves, with delta-functions
1691: that guarantee the curves intersect.  
1692: We just proved the agreement for $K=1$.  For general $K$ we can work 
1693: inductively; given a $K+1$-component tree
1694: on which some curve is further degenerating, just focus on the measure for that curve, and 
1695: note that the delta-functions from the other curves are well behaved on moduli space
1696: near the degeneration we are studying.  In this sense the degenerating curve can
1697: be isolated from the rest of the tree.  The computation done above in the $K=1$ case then 
1698: shows that the measure after this degeneration agrees with that from the disconnected 
1699: prescription.  This then completes the argument for the equivalence of all prescriptions.
1700: 
1701: 
1702: 
1703: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1704: \section{Conclusions and open questions}
1705: 
1706: We have argued for the equivalence of the connected and disconnected
1707: twistorial formulae for the tree level scattering amplitudes of $\N=4$ super Yang-Mills,
1708: provided that the contours are appropriately chosen.
1709: Using this equivalence we can now exploit the complementary virtues of the two
1710: prescriptions simultaneously.  As we remarked in the introduction, the connected prescription
1711: minimizes the number of diagrams one has to sum, namely, there is only one; the amplitude 
1712: is expressed as a single integral, which was the starting point for several theoretical developments
1713: \cite{Berkovits:2004hg,Berkovits:2004tx,Witten:2004cp}.  The disconnected
1714: prescription involves more diagrams, but still a manageable number for some interesting
1715: amplitudes, and the contribution from each diagram can be immediately written down.  
1716: 
1717: To conclude, we summarize some of the many remaining open problems in this area:
1718: 
1719: \begin{itemize}
1720: 
1721: \item {\bf Contours I.}
1722: Is there a rigorous justification of the choice of contours in all these calculations?
1723: In our argument for the equivalence between connected and disconnected
1724: prescriptions we identified specific poles in the integral
1725: over moduli, and we roughly wanted a contour which encircles all of these poles.
1726: We believe it should be possible to show by a deformation argument that 
1727: our choice of contour is equivalent to the one used in \cite{Roiban:2004yf},
1728: thus completing the proof of equivalence,
1729: but this seems to be nontrivial; the 
1730: computations in \cite{Roiban:2004yf} depend on a particular method of evaluating the integral
1731: in the connected prescription by saturating delta-functions, and it is difficult to see which contour 
1732: it corresponds to.
1733: 
1734: \item {\bf Contours II.} Once the residues are isolated in both prescriptions,
1735: we must still integrate over the degeneration locus 
1736: $\Mint$, which requires yet another choice of
1737: contour; for example, the integration over $t$ from $0$ to $\infty$ in section
1738: 6 of \cite{Cachazo:2004kj} should have some \ti{a priori} justification. 
1739: This paper has not addressed this question.
1740: Our argument for the equivalence requires that the contours on $\Mint$ are chosen
1741: to be equivalent in all prescriptions.
1742: 
1743: \item {\bf Explicit external wavefunctions.}
1744: Our derivation was rather formal. It did not depend on the particular form of the wavefunctions.
1745: Of course, it would be interesting to verify the picture by calculating the amplitudes 
1746: involving particles with well-defined momenta {\it i.e.}\ $(\lambda,\tilde\lambda,\psi)$
1747: using our generalized prescriptions.  Unlike the MHV vertices in \cite{Cachazo:2004kj},
1748: one might expect that the $d>1$ vertices will be ratios of polynomials involving both $\lambda$ and $\tilde\lambda$.  (Of course, it is also possible that one will not 
1749: obtain any compact formula for the $d>1$ vertices in this way.)
1750: 
1751: \item {\bf Derivation from the B-model.}
1752: Both connected and disconnected contributions seem to arise in the topological
1753: B-model of \cite{Witten:2003nn} as long as we use not only the degree $d$ 
1754: D1-instantons but also the propagators (and vertices) of the holomorphic Chern-Simons theory.
1755: Does our equivalence suggest that the D1-instantons are not independent
1756: of the Chern-Simons degrees of freedom?
1757: 
1758: \item {\bf Real versions.}
1759: The framework first proposed by Berkovits \cite{Berkovits:2004hg} 
1760: and the topological A-model of \cite{Neitzke:2004pf} seem to prefer
1761: the real version of the twistor space, $\mathbb{RP}^{3|4}$, and correspondingly
1762: real values of the moduli. Is there a real variation of our procedures?
1763: One can imagine that the disconnected rules for the amplitudes might be derived
1764: from the cubic twistorial string field theory of \cite{Berkovits:2004tx} if $K$
1765: stringy propagators are expanded in component fields, so that the different parts
1766: of the worldsheet become effectively disconnected.
1767: 
1768: \item {\bf Choice of prescriptions.}
1769: According to our analysis, there is significant freedom to choose a 
1770: twistor prescription for tree diagrams; we gave $d$ different rules,
1771: all of which agree up to overall prefactors.  Is this all one can say, or
1772: would a more sensitive study give more information
1773: about which is the ``correct'' prescription?  Does this proliferation of
1774: prescriptions persist beyond tree level?
1775: 
1776: \item {\bf Loops and higher genus.} 
1777: We only studied tree diagrams, corresponding to genus zero curves.  
1778: What are the exact rules and equivalences between various formulae for 
1779: loop and nonplanar amplitudes?
1780: Our analysis suggests that an investigation of possible degenerations of genus $g$ curves 
1781: should be relevant for the understanding of loop diagrams in the twistor string.
1782: 
1783: \end{itemize}
1784: 
1785: 
1786: 
1787: \section*{Acknowledgements}
1788: We are grateful to Michal Fabinger, Peter Svr\v{c}ek, Cumrun Vafa, Anastasia Volovich and 
1789: Edward Witten for very useful discussions.
1790: This work was conducted during the period S.G.\ served
1791: as a Clay Mathematics Institute Long-Term Prize Fellow.
1792: S.G. is also supported in part by RFBR grant 01-02-17488.
1793: The work of L.M.\ was supported in part by Harvard DOE grant
1794: DE-FG01-91ER40654 and the Harvard Society of Fellows.  The
1795: work of A.N.\ was supported by NSF grants PHY-0255841 
1796: and DMS-0244464.
1797: 
1798: \renewcommand{\baselinestretch}{1}
1799: \small\normalsize
1800: 
1801: \bibliography{physics}
1802: 
1803: \end{document}
1804: