1: %% draft 4/24/07 12:18am
2:
3: \documentclass[12pt,letter]{article}
4: \usepackage{amssymb,amsmath,epsf,graphicx,cite}
5: \input epsf.sty
6:
7:
8: \textheight=9in
9: \textwidth=6.5in
10: \headheight=0in
11: \headsep=0in
12: \topmargin=0in
13: \oddsidemargin=0in
14:
15:
16:
17: \numberwithin{equation}{section}
18: \numberwithin{table}{section}\setlength{\multlinegap}{25pt}
19:
20:
21:
22: \def \BB {{\widehat{\cal B}}}
23: \def \ad {\text{ad}}
24:
25:
26: \begin{document}
27: {}~ \hfill\vbox{\hbox{UK-08/02} }\break
28: \vskip 4.1cm
29:
30: \centerline{\Large \bf The closed string tadpole in open string field theory}
31: \vspace*{10.0ex}
32:
33: \centerline{\large \rm Ian Ellwood}
34:
35: \vspace*{8.0ex}
36:
37: \centerline{\large \it Department of Physics and Astronomy, }
38: \centerline{\large \it University of Kentucky, Lexington, KY 40506, USA}
39: \vspace*{2.0ex}
40: \centerline{E-mail: {\tt iellwood@pa.uky.edu}}
41:
42: \vspace*{6.0ex}
43:
44:
45: \vspace*{6.0ex}
46:
47: \centerline{\bf Abstract}
48: \bigskip
49:
50: We compute a class of gauge invariant observables for marginal
51: solutions and the tachyon vacuum. In each case we find that the
52: observables are related in a simple way to the closed-string tadpole
53: on a disk with appropriate boundary conditions. We give a sketch of an
54: argument that this result should hold in general using the BRST
55: invariance of the closed string two-point function. Finally, we
56: discuss the analogous set of invariants in the Berkovits superstring
57: field theory.
58:
59:
60: \baselineskip=16pt
61: \parskip = 3pt
62:
63:
64: \newpage
65: \tableofcontents
66:
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: \section{Introduction}
69: \label{s:intro}
70:
71: Open string field theory originated as an attempt to find a classical
72: theory which, upon quantization, would reproduce the complete
73: perturbation expansion of open string scattering diagrams
74: \cite{Witten:1985cc,Giddings:1986wp,Giddings:1986bp,Thorn:1988hm,Zwiebach:1990az}.
75: Recently, it has become clear that even before quantization,
76: classical string field theory contains a rich amount of
77: information about D-brane physics and, more generally, boundary
78: conformal field theories
79: \cite{Sen:1999nx,Zwiebach:2001nj,Taylor:2003gn,Schnabl:2005gv,Okawa:2006vm,Fuchs:2006hw,Ellwood:2006ba,Schnabl:2007az,Kiermaier:2007ba,Okawa:2007ri,Erler:2007rh,Okawa:2007it,Fuchs:2007yy,Fuchs:2007gw,Ellwood:2007xr,Kiermaier:2007vu,Kiermaier:2007ki,Erler:2007xt,Kwon:2008ap,Hellerman:2008wp}.
80:
81: Indeed, there is a assumption among many string field theory
82: practitioners that, given a solution of the classical equations of
83: motion of string field theory $\Psi$, there is a corresponding
84: boundary $\text{CFT}_\Psi$. Furthermore, given a boundary
85: $\text{CFT}_\Psi$ (which is in some unspecified sense ``not too far
86: away'' from the boundary $\text{CFT}_0$ around which the string field
87: theory was defined), there is a classical solution $\Psi$ which
88: shifts us from $\text{CFT}_0$ to $\text{CFT}_\Psi$.
89:
90: This would-be duality between string fields and boundary CFTs is
91: obfuscated by the large amount of gauge symmetry in open string field
92: theory. For example, if we are working in bosonic cubic string field
93: theory, and the string field $\Psi$ represents some boundary CFT,
94: then the string field,
95: \begin{equation}
96: \Psi' = e^{\Lambda} (\Psi + Q_B) e^{-\Lambda} \ ,
97: \end{equation}
98: should represent the same boundary CFT for any ghost number 0 string
99: field, $\Lambda$. This gauge symmetry has no analogue in boundary
100: conformal field theory so, if we wish to compare the two sides of the
101: duality, it is useful to consider gauge-invariant quantities.
102:
103: The list of known gauge-invariant objects is very short. For the
104: bosonic string, one has the classical action \cite{Witten:1985cc},
105: \begin{equation} \label{HIinvariant}
106: S(\Psi) = \frac{1}{2} \int \Psi * Q_B \Psi + \frac{1}{3} \int \Psi*\Psi*\Psi \ ,
107: \end{equation}
108: and the quantities discovered independently by Hashimoto and Itzhaki \cite{Hashimoto:2001sm} and Gaiotto, Rastelli, Sen, and Zwiebach \cite{Gaiotto:2001ji}, which take the
109: form\footnote{These invariants were first
110: introduced in a different context by Shapiro and Thorn in
111: \cite{Shapiro:1987gq,Shapiro:1987ac}.},
112: \begin{equation} \label{InvariantIntro}
113: W(\Psi,\mathcal{V}) = \langle \mathcal{I}| \mathcal{V}(i) |\Psi\rangle \ ,
114: \end{equation}
115: where $\mathcal{I}$ is the identity string field, and $\mathcal{V} = c
116: \bar c \mathcal{O}^{\text{m}}$ is an on-shell closed-string vertex
117: operator inserted at the midpoint of the string (which is at the point
118: $z=i$ in the standard UHP coordinates).
119:
120: While the classical action has a straightforward interpretation, it is
121: less clear what the invariants (\ref{InvariantIntro}) compute. In fact,
122: since $W(\Psi,\mathcal{V})$ involves the identity field, one might
123: worry that it would be singular, but, as we'll see in explicit
124: computations, it is well-defined for the known solutions.
125:
126: Since $W(\Psi,\mathcal{V})$ is gauge-invariant, it should correspond
127: to some definite quantity in the CFT associated with $\Psi$. In this paper we
128: motivate the following proposal:
129:
130: \bigskip
131: \noindent
132: Let the string field theory of interest be defined
133: around a boundary CFT${}_0$. Let $\Psi$ be a string-field associated
134: to the boundary $\text{CFT}_\Psi$. Then
135: \begin{equation} \label{MainRelation}
136: W(\Psi, \mathcal{V}) = \mathcal{A}^{\text{disk}}_\Psi(\mathcal{V})- \mathcal{A}^{\text{disk}}_0(\mathcal{V}) \ ,
137: \end{equation}
138: where $\mathcal{A}^{\text{disk}}_\Phi(\mathcal{V})$ is the disk
139: amplitude with one closed string vertex operator $\mathcal{V}$ and
140: boundary conditions given by CFT${}_\Phi$.
141:
142: As we will show, this relationship can be derived from the BRST
143: invariance of the closed string two-point function. This derivation
144: is very delicate both in its use of BRST invariance and its implicit reliance
145: on certain assumptions about
146: the nature of the string fields used in the computation of the
147: invariants. As such, our derivation is non-rigorous, and we consider the
148: fact that (\ref{MainRelation}) holds in explicit examples as important
149: evidence that it is correct.
150:
151: The relation (\ref{MainRelation}) can be viewed in two ways: First,
152: given a $\Psi$, we may compute the left hand side for all possible
153: $\mathcal{V}$ to determine the complete {\em physical} part of the boundary
154: state of $\text{CFT}_\Psi$. Second, given a boundary state of some
155: boundary CFT for which we don't know the associated $\Psi$, we can use
156: (\ref{MainRelation}) to find a number of linear constraints on
157: $\Psi$. These may aid in the search for new solutions to the string
158: field theory equations of motion, though it does not seem that they
159: are enough information to derive a string field theory solution given
160: a CFT since the on-shell condition on the closed string field puts tight
161: restrictions on its form in most cases.
162:
163: Having given an interpretation for $W(\Psi,\mathcal{V})$, it is
164: natural to extend the construction to the Berkovits open superstring
165: field theory \cite{Berkovits:1995ab,Berkovits:1998bt,Berkovits:2000zj}.
166: The string field in this case has a different gauge invariance,
167: \begin{equation}
168: e^{\Phi} \to e^{Q_B \Lambda} e^{\Phi} e^{\eta_0 \Lambda'}
169: \end{equation}
170: where $\Lambda$ and $\Lambda'$ are independent gauge parameters and
171: $\eta_0$ is the zero mode of $\eta$ in the $\eta$, $\xi$, $\phi$
172: superconformal ghost system. Nonetheless, a set of invariants, which are very similar to
173: the bosonic invariants was written down in \cite{Michishita:2004rx}.
174:
175: We use a slightly different, but equivalent, form of these invariants: As has held true
176: in a number of examples \cite{Erler:2007rh,Okawa:2007ri,Okawa:2007it,Kiermaier:2007ki}, the analogue of the bosonic string field
177: $\Psi$ in the superstring is $e^{-\Phi} Q_B e^{\Phi}$. This leads to
178: a set of invariants in superstring field theory,
179: \begin{equation}
180: \widehat{W}(\Phi,\mathcal{V}) = \langle \mathcal{I}| \mathcal{V}(i)| e^{-\Phi}
181: Q_B e^{\Phi}\rangle \ ,
182: \end{equation}
183: where $\mathcal{V}$ is a weight zero primary field inserted at the
184: midpoint which satisfies
185: \begin{equation}
186: Q_B \eta_0 \mathcal{V} = 0 \ .
187: \end{equation}
188: The operator $\mathcal{V}$ lives in the big Hilbert space which
189: includes the zero-mode of $\xi$ and should be thought of as $(\xi +
190: \tilde{\xi}) \mathcal{O}$ where $\mathcal{O}$ is in the small Hilbert
191: space.
192:
193: We will see in an example that this quantity appears to compute the change in
194: the closed string one-point function, just as is it does in the bosonic case.
195: However, because of the complexity of perturbation theory in the Berkovits superstring,
196: we do not have a general derivation of this result.
197:
198: The organization of this paper is as follows: In section \ref{Review},
199: we review the construction of the invariants $W(\Psi,\mathcal{V})$, the
200: $\arctan(z)$ coordinate system and the closed string tadpole. In
201: section \ref{ExplicitComputations}, we compute $W(\Psi,\mathcal{V})$
202: for marginal deformations and the tachyon vacuum. In section
203: \ref{Derivation}, we show how the relation between the closed string
204: one-point function and $W(\Psi,\mathcal{V})$ can be derived from BRST
205: invariance of the closed string two-point function. Finally, in
206: section \ref{susyversion} we discuss an extension to the
207: Berkovits superstring field theory.
208:
209:
210: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
211: \section{Review} \label{Review}
212: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
213:
214: In this section we review the invariants $W(\Psi,\mathcal{V})$ introduced in
215: \cite{Hashimoto:2001sm,Gaiotto:2001ji} and discuss how they are computed in the
216: $\arctan(z)$ coordinates. We then discuss some aspects of the closed
217: string tadpole diagram which will be useful later.
218:
219: \subsection{The invariants $W(\Psi,\mathcal{V})$}
220:
221: Consider a string field $\Psi$, defined as the state $|\Psi\rangle =
222: \mathcal{O}_\Psi(0)|0\rangle$, where $\mathcal{O}_\Psi$ is a ghost number 1
223: boundary operator and $|0\rangle$ is the $SL_2(\mathbb{R})$ vacuum.
224: In the upper half plane, we may think of the state $|\Psi\rangle$ as living
225: on the unit semi-circle as in figure \ref{MapToIdentity}a.
226:
227:
228: \begin{figure}
229: \centerline{
230: \begin{picture}(380,122)(0,-10)
231: \includegraphics{MapToIdentity}
232: \end{picture}
233: \begin{picture}(0,0)(380,-10)
234: \put(68,-12){$\mathcal{O}_{\Psi}$}
235: \put(249,-12){ $w\circ \mathcal{O}_{\Psi}$}
236: \put(272,53){ $\mathcal{V}(i)$}
237: \put(69,65){$i$}
238: \put(117,38){L}
239: \put(16,38){R}
240: \put(272,83){L}
241: \put(255,83){R}
242: \put(-10,100){\bf a)}
243: \put(155,100){\bf b)}
244: \end{picture}
245: }
246: \caption{The construction of the invariants $W(\Psi,\mathcal{V})$ is
247: shown. In a), we begin with a state $\Psi$ formed by inserting the
248: vertex operator $\mathcal{O}_{\Psi}$ into the UHP at the origin. The
249: wavefunction for the state $\Psi$ is to be thought of as living on the
250: unit semi-circle. The left and right halves of the string as seen
251: from infinity are labeled L and R. The string midpoint is at $z = i$.
252: To contract the state with the identity, glue the semicircles L and R
253: together and map the resulting geometry to the plane using $z \to
254: w(z)$ as shown in b). To saturate the ghostnumber, insert a closed
255: string field $\mathcal{V}$ at the midpoint, $w(i) =
256: i$. \label{MapToIdentity}}
257: \end{figure}
258: To define the invariants $W(\Psi,\mathcal{V})$, first map the upper half disk to the
259: entire upper half plane using the map,
260: \begin{equation}
261: w(z) = \frac{2 z}{1-z^2} \ .
262: \end{equation}
263: This is shown in figure \ref{MapToIdentity}b. Next, to saturate the
264: ghost number on the UHP, add a ghostnumber 2 vertex operator
265: $\mathcal{V}(i)$ at the midpoint. Finally, compute the correlator,
266: \begin{equation}
267: W(\Psi,\mathcal{V}) = \langle \mathcal{V}(i) \, w\circ
268: \mathcal{O}_\Psi \rangle_{\text{UHP}} \ .
269: \end{equation}
270: The key property of $W(\Psi,\mathcal{V})$ is that if $\mathcal{V}$ is
271: a weight $(0,0)$ primary, satisfying $\{Q_B, \mathcal{V}\} = 0$, then $W$
272: is invariant under the open string field theory gauge group,
273: \begin{equation}
274: W(\Psi + Q_B \Lambda + [\Psi,\Lambda],\mathcal{V}) = W(\Psi,\mathcal{V}) \ .
275: \end{equation}
276: Since $W(\Psi,\mathcal{V})$ is linear in $\Psi$, this follows from the identities,
277: \begin{align}
278: W(Q_B \Lambda,\mathcal{V}) &=0 \label{I1}\ ,
279: \\ W([\Psi,\Lambda],\mathcal{V}) &=0 \label{I2} \ .
280: \end{align}
281: To show (\ref{I1}), suppose $|\Lambda\rangle =
282: \mathcal{O}_\Lambda(0) |0\rangle$. Then,
283: \begin{equation}
284: W(Q_B \Lambda,\mathcal{V}) = \langle \mathcal{V}(i) \, w\circ
285: \{Q_B,\mathcal{O}_\Lambda \} \rangle_{\text{UHP}}
286: = -\langle [Q_B, \mathcal{V}(i)] \, w\circ
287: \mathcal{O}_\Lambda \rangle_{\text{UHP}} = 0 \ ,
288: \end{equation}
289: where the second equality uses the BRST invariance of the boundary
290: conditions on the UHP, $\langle \{Q_B, \ldots \} \rangle = 0$.
291:
292: The second identity (\ref{I2}) follows from essentially the same
293: arguments that show
294: \begin{equation}
295: \int \Psi_1*\Psi_2 = \int \Psi_2*\Psi_1 \ .
296: \end{equation}
297: Assuming that $\mathcal{V}$ is a weight (0,0)
298: primary,
299: \begin{align}
300: W(\Psi*\Lambda,\mathcal{V}) = \langle \mathcal{V}(i)
301: \mathcal{O}_{\Psi*\Lambda} \rangle_{\text{UHP}} =\langle \mathcal{V}(i)\, f_1
302: \circ \mathcal{O}_\Psi \, f_2 \circ \mathcal{O}_\Lambda \rangle_{\text{UHP}} \ ,
303: \label{PL}
304: \\
305: W(\Lambda*\Psi,\mathcal{V}) = \langle \mathcal{V}(i)
306: \mathcal{O}_{\Lambda*\Psi} \rangle_{\text{UHP}} =\langle \mathcal{V}(i)\, f_1
307: \circ \mathcal{O}_\Lambda \, f_2 \circ \mathcal{O}_\Psi \rangle_{\text{UHP}} \ ,
308: \label{LP}
309: \end{align}
310: where
311: \begin{equation}
312: f_1(z) = \frac{1+z}{1-z} \ , \qquad f_2(z) = -\frac{1-z}{1+z} \ .
313: \end{equation}
314: Noting that $f_1 = I\circ f_2$, where $I(z) = -1/z$ is the BPZ dual,
315: it follows that (\ref{PL}) and (\ref{LP}) are related by an $SL_2(\mathbb{Z})$
316: transformation and, hence, equal. This implies
317: \begin{equation}
318: W(\Psi*\Lambda - \Lambda*\Psi,\mathcal{V}) = 0 \ .
319: \end{equation}
320:
321: \subsection{The $\arctan(z)$ frame}
322:
323: It will be useful in the discussion to follow to know how to compute $W(\Psi,\mathcal{V})$ when
324: the state $\Psi$ is given in the $\arctan(z)$ coordinate system that has played a prominent
325: role in recent developments. Define,
326: \begin{equation}\label{fDef}
327: \tilde z = f(z) = \tfrac{2}{\pi} \arctan(z) \ ,
328: \end{equation}
329: which takes the upper half plane to a semi-infinite cylinder of
330: circumference $2$. A correlator on a semi-infinite cylinder of
331: circumference $n$ is defined by first rescaling $\tilde z \to \frac{2}{n} \tilde z$
332: to get back to a cylinder of width $2$ and then mapping $\tilde z \to
333: f^{-1}(\tilde z)$ to get back to the upper half plane. We will often
334: follow the notation of \cite{Kiermaier:2007ba} and consider the fundamental region of
335: the cylinder to be the region $-\frac{1}{2} < \Re(\tilde z) < n-\frac{1}{2}$.
336: This unusual choice happens to be convenient for the form
337: of some string field solutions.
338:
339: A prototypical state $|\Sigma\rangle$ defined in cylinder coordinates is
340: shown pictorially in \ref{WOfCylindarState}a. Algebraically, we
341: define $|\Sigma\rangle$ through its overlap with an arbitrary test
342: state $\langle \phi|$,
343: \begin{equation}
344: \langle \phi | \Sigma \rangle
345: = \langle f \circ \phi(0) \, \,\mathcal{O}(\tilde z_1) \ldots \mathcal{O}(\tilde z_n) \rangle_{C_n}
346: \end{equation}
347: where the $\mathcal{O}$'s
348: are some local operators and the subscript $C_n$ indicates that the
349: correlator is to be evaluated on a cylinder of circumference $n$. In
350: order for this to be a non-singular definition, we must require that
351: none of the $\tilde z_i$ are contained in the image of the unit
352: half-disk under the map $f(z)$. This region is given by $-\frac{1}{2} \le
353: \Re(\tilde z) \le \frac{1}{2}$ (and its images under $\tilde z \to
354: \tilde z + n$).
355:
356: Given a state $|\Sigma\rangle$ defined in this way, we would like to
357: compute $W(\Sigma,\mathcal{V})$. The first step is to glue the left
358: and right halves of $\Sigma$ together. In the $\tilde z$ coordinates,
359: the left and right halves of the string live at $\Re(\tilde z) = n
360: +\frac{1}{2}$ and $\Re(\tilde z) = \frac{1}{2}$ respectively as shown
361: in the figure. To glue them together, we remove the coordinate patch
362: $-\frac{1}{2} < \Re(\tilde z)<\frac{1}{2}$, leaving us with a strip of
363: worldsheet of width $n-1$ and then glue the two sides of the
364: worldsheet together, giving us back a cylinder of circumference $n-1$. This
365: is shown in figure \ref{WOfCylindarState}b. Finally, the operator
366: $\mathcal{V}$ should be inserted at $i\infty,$ which is the string
367: midpoint in the $\tilde z$ coordinates. In total,
368: \begin{equation}
369: W(\Sigma,\mathcal{V}) =
370: \langle \mathcal{V}(i\infty) \, \mathcal{O}(\tilde z_1) \ldots \mathcal{O}(\tilde z_n)\rangle_{C_{n-1}} \ .
371: \end{equation}
372: \begin{figure}
373: \centerline{
374: \begin{picture}(416,148)(0,-10)
375: \includegraphics{WOfCylindarState}
376: \end{picture}
377: \begin{picture}(0,0)(416,-10)
378: \put(24,15){$f\circ \phi$}
379: \put(-10,-12){$-\frac{1}{2}$}
380: \put(71,-12){$\frac{1}{2}$}
381: \put(91,-12){$z_1$}
382: \put(120,-12){$z_2$}
383: \put(147,-12){$\cdots$}
384: \put(178,-12){$z_n$}
385: \put(89,15){$\mathcal{O}$}
386: \put(118,15){$\mathcal{O}$}
387: \put(176,15){$\mathcal{O}$}
388: \put(205,-12){$n-\frac{1}{2}$}
389: %
390: \put(261,-12){$\frac{1}{2}$}
391: \put(281,-12){$z_1$}
392: \put(310,-12){$z_2$}
393: \put(337,-12){$\cdots$}
394: \put(368,-12){$z_n$}
395: \put(279,15){$\mathcal{O}$}
396: \put(308,15){$\mathcal{O}$}
397: \put(366,15){$\mathcal{O}$}
398: \put(395,-12){$n-\frac{1}{2}$}
399: \put(-25,130){\bf a)}
400: \put(241,130){\bf b)}
401: \put(82,110){R}
402: \put(204,110){L}
403: \end{picture}
404: }
405: \caption{In a) a typical state $|\Sigma\rangle$ is shown in cylinder
406: coordinates. The shaded region represents the coordinate patch, or,
407: in other words, the image of the unit half disk under $f(z)$. The
408: left and right halves of the state $|\Sigma\rangle$ are labeled L and
409: R. In b) $W(\Sigma,\mathcal{V})$ is shown. This is obtained by
410: removing the coordinate patch and gluing the lines labeled by L and R
411: together. As a final step the operator $\mathcal{V}$ should be
412: inserted at $i\infty$. \label{WOfCylindarState}}
413: \end{figure}
414:
415: \subsection{The closed string one-point function}
416:
417: Since we wish to relate $W(\Psi,\mathcal{V})$ to the tree-level closed string
418: one-point function, it is useful to review how this diagram is
419: computed. The closed-string one-point function is the amplitude with
420: one vertex operator $\mathcal{V}$ inserted on the disk. Since there
421: are 3 CKVs on the disk, we may fix the position of the one vertex operator to the center of the disk, $z
422: = 0$. Hence, $\mathcal{V}$ should be a fixed vertex operator of the form $c\tilde c
423: \mathcal{O}^{\text{m}}$ where $\mathcal{O}^{\text{matter}}$ is a
424: weight $(1,1)$ matter operator. Note, however, that
425: \begin{equation}
426: \langle \mathcal{V}(0) \rangle_{\text{disk}} = 0 \ ,
427: \end{equation}
428: since, to get a non-vanishing answer, we need soak up three ghost
429: zero-modes and we have only soaked up two. The problem is that fixing
430: the position of $\mathcal{V}$ only removes two out of the three CKV's
431: and the third, which generates rotations of the disk, has an
432: associated ghost-zeromode. Typically, if we have CKV's left over, a
433: diagram will vanish because the volume of the associated group of
434: symmetries is infinite. In this case, the volume of the group of
435: rotations of the disk is just $2\pi$ so the amplitude is finite.
436:
437: To soak up the remaining zero-mode, we add the ghost-measure corresponding to fixing one of the points $z = e^{i\theta}$ on the boundary of the disk. Given an infinitesimal coordinate shift $\delta \sigma^a$,
438: its component along the boundary is given by
439: \begin{equation}
440: \sin \theta \,\delta \sigma^1 -\cos \theta\,\delta \sigma^2 = -\Im(e^{-i\theta} \delta \sigma^z) \ ,
441: \end{equation}
442: at the point $z = e^{i \theta}$. To get the correct measure, we
443: should then add\footnote{We are not attempting to determine the
444: overall sign of the ghost measure. It has been picked to give
445: (\ref{MainRelation}) rather than
446: $\mathcal{A}_0^{\text{disk}}(\mathcal{V})
447: -\mathcal{A}_\Psi^{\text{disk}}(\mathcal{V}) $.}
448: \begin{equation}
449: -\Im(e^{-i \theta} c(e^{i\theta})) = i e^{-i\theta} c(e^{i\theta})
450: \end{equation}
451: to ghost path integral. The complete one-point function is given
452: by
453: \begin{equation}\label{OnePointFunction}
454: \mathcal{A}^{\text{disk}}(\mathcal{V}) = -\frac{e^{-i\theta}}{2\pi i} \langle \mathcal{V}(0) \,c(e^{i\theta})\rangle_{\text{disk}} \ .
455: \end{equation}
456: Note that we have included an extra factor of $(2\pi)^{-1}$ to account
457: for the volume of the CKV group. One can check that
458: (\ref{OnePointFunction}) is independent of $\theta$ as it should be.
459: In general, we will pick $\theta = 0$.
460:
461: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
462: \section{Computation of $W(\Psi,\mathcal{V})$ for known solutions}
463: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
464: \label{ExplicitComputations}
465:
466: In this section, the invariants $W(\Psi,\mathcal{V})$ are computed for
467: various known solutions. In each case, the result is found to be
468: consistent with the change in the one-point function of the closed
469: string under the shift from the original boundary conditions to the
470: new boundary conditions associated with the string field solution.
471:
472:
473: \subsection{Invariants of marginal deformations with trivial OPEs}
474: \label{TrivialOPE}
475:
476: There are currently two (presumably) gauge-equivalent solutions to
477: the OSFT equations of motion that describe marginal deformations with
478: trivial OPE. The first \cite{Schnabl:2007az,Kiermaier:2007ba}, which
479: is in Schnabl-gauge \cite{Schnabl:2005gv}, turns out to be impractical
480: for computing $W(\Psi,\mathcal{V})$. The second state, discovered by Fuchs, Kroyter and Potting \cite{Fuchs:2007yy} and
481: Kiermaier and Okawa \cite{Kiermaier:2007vu}, appears to be more closely
482: related to the boundary conformal field theory and is better suited for our
483: computation. Their solution also has a natural extension to the
484: non-trivial OPE case, which we will take up in the next subsection.
485:
486: The complete solution takes the form \cite{Kiermaier:2007vu},
487: \begin{equation}\label{KOsolution}
488: \Psi^{\text{KO}} = \frac{1}{\sqrt{U}}(\Psi_L + Q_B) \sqrt{U} \ ,
489: \end{equation}
490: where $\Psi_L$ is a state to be introduced shortly and $U$ is a string
491: field whose form we will not need. The state (\ref{KOsolution})
492: appears to be a gauge-transformation of the state $\Psi_L$; however,
493: neither $\Psi_L$ nor $U$ are real string fields so (\ref{KOsolution})
494: is not a proper gauge transformation. Nevertheless, since,
495: $W(\Psi,\mathcal{V})$ has no knowledge of the reality condition, we
496: can work with the simpler state $\Psi_L$.
497:
498: The state $\Psi_L$ is given by\footnote{In \cite{Kiermaier:2007vu}, this
499: would be written $\sum_{n = 1}^\infty \lambda^n \, \Psi^{(n)}_L$ as
500: they pick the opposite convention for the left and right halves of the
501: string wave function. This affects the overall sign of the invariant as well as the sign of the deformation.},
502: \begin{equation}
503: \Psi_L =- \sum_{n = 1}^\infty (-\lambda)^n \, \Psi^{(n)}_L \ ,
504: \end{equation}
505: where, following \cite{Kiermaier:2007vu}, we define the states $\Psi^{(n)}$ on a
506: cylinder of circumference $n+1$,
507: \begin{equation}
508: \langle \phi | \Psi^{(n)}_L\rangle =
509: \left\langle
510: f\circ \phi(0) c J(1) \int_1^2 dt_1 \int_{t_1}^3 dt_2 \int_{t_2}^4 dt_3
511: \ldots
512: \int_{t_{n-2}}^n \, J(t_1) J(t_2) J(t_3) \ldots J(t_{n-1}) \right\rangle_{C_{n+1}} \ .
513: \end{equation}
514: As defined in (\ref{fDef}), the map $f$ is given by $f(z) = \frac{2}{\pi} \arctan(z)$. The field
515: $J$ is assumed to be a weight $1$ primary boundary matter operator
516: with trivial OPE: $J(z) J(0) \sim \mathcal{O}(1)$.
517:
518: To compute $W(\Psi^{(n)},\mathcal{V})$, remove the coordinate patch
519: $-1/2 < \Re(\tilde z) <1/2$ and re-glue to form a cylinder of width $n$. Then insert $\mathcal{V}(i\infty)$:
520: \begin{multline}
521: W(\Psi^{(n)},\mathcal{V}) =
522: \\
523: \left\langle
524: \mathcal{V}(i\infty) c J(0) \int_0^1 dt_1 \int_{t_1}^2 dt_2 \int_{t_2}^3 dt_3
525: \ldots
526: \int_{t_{n-2}}^{n-1} dt_{n-1}\, J(t_1) J(t_2) J(t_3) \ldots J(t_{n-1}) \right\rangle_{C_{n}} \ .
527: \end{multline}
528: Mapping this geometry to the disk using
529: \begin{equation}
530: g(\tilde z) = e^{2\pi i \tilde z/n} \ ,
531: \end{equation}
532: yields\footnote{Note that under $z \to \chi(z)$, a weight $h$ boundary
533: operator transforms as $\mathcal{O}(z) \to |\frac{\partial\chi}{\partial z}|^h
534: \mathcal{O}(\chi(z))$.}
535: \begin{equation}\label{KOintegral}
536: -i\left\langle \mathcal{V}(0) c J(1)
537: \int_0^{\omega} dt_1 \int_{t_1}^{2\omega} dt_2
538: \ldots \int_{t_{n-2}}^{(n-1)\omega}dt_{n-1} \, J(e^{i t_1})
539: J(e^{i t_2}) \ldots J(e^{i t_{n-1}}) \right\rangle_{\text{disk}} \ ,
540: \end{equation}
541: where $\omega = 2\pi/n$.
542:
543: Remarkably, as we will now demonstrate, this complicated integral is equal to the simpler integral,
544:
545: \begin{equation} \label{freeintegral}
546: -\frac{i}{2\pi n!}\left\langle \mathcal{V}(0)\, c(1)
547: \int_0^{2\pi} dt_1 \int_{0}^{2\pi} dt_2
548: \ldots \int_{0}^{2\pi} dt_{n}\, J(e^{i t_1})
549: J(e^{i t_2}) \ldots J(e^{i t_{n}}) \right\rangle^{\text{m}}_{\text{disk}} \ .
550: \end{equation}
551: Notice that the difference between the two integration regions is that
552: in (\ref{KOintegral}) we have the constraints that $t_k \le \omega k$.
553: These inequalities are explained by the following
554: lemma:
555:
556: \noindent
557: {\bf Lemma}: Given $n$ points on the unit circle, we may always label
558: them in counter clockwise order, $z_i = e^{i\theta_i}$, $i\in \{1,\ldots,n\}$ with increasing
559: $\theta_i$ such that
560: \begin{equation} \label{ineq}
561: \theta_j-\theta_1 \le \tfrac{2\pi}{n} (j-1) \ .
562: \end{equation}
563:
564: \noindent
565: {\em Proof}: We use proof by contradiction. Begin by extending the
566: definition of $\theta_i$ to include $i\in \mathbb{Z}$, by defining
567: $\theta_{i+n} = \theta_i + 2\pi$. Assuming the lemma is false, we have that
568: for every $\theta_i$, there exists a $\theta_j$ with $j>i$ such that
569: \begin{equation}
570: \theta_j - \theta_i > \tfrac{2\pi}{n} (j-i) \ .
571: \end{equation}
572: Hence, there exists a sequence $\{ \theta_{i_m}\}$ such that
573: \begin{equation}
574: \theta_{i_{m}} - \theta_{i_{m-1}} > \tfrac{2 \pi}{n} (i_m - i_{m-1}) \ ,
575: \end{equation}
576: from which it follows that
577: \begin{equation} \label{thetaineq}
578: \theta_{i_m} - \theta_{i_p} > \tfrac{2\pi}{n}(i_m - i_p) \ .
579: \end{equation}
580: Since there are only a finite number of points on the circle, there
581: must be two points in the sequence such that $i_a - i_b = k n$ for
582: some $k\in \mathbb{Z}$. Since these represent the same point on the
583: circle, we learn that
584: \begin{equation}
585: \theta_{i_a} - \theta_{i_b} = 2\pi k \ ,
586: \end{equation}
587: which is in contradiction with (\ref{thetaineq}) for $m= a$ and $p = b$. $\square$
588:
589: The choice of $z_1$ is generically unique. If there are two
590: possible points which may be chosen as the first point, it follows
591: from (\ref{ineq}) that they must be separated by an integer multiple of $2\pi/n$.
592:
593: Now, consider the integral (\ref{freeintegral}). Ignoring special points
594: in the integration region (which are measure zero), we can divide the
595: integral up into $n$ integrals in which one of the $n$ points is
596: picked to be $z_1$ and the rest of the points satisfy (\ref{ineq}).
597: We can fix the order of the remaining points at the expense of
598: introducing a factor of $(n-1)!$ and we may fix $z_1 = 1$ by a
599: rotation if we multiply the integral by $2\pi$ (which cancels the
600: $2\pi$ in (\ref{freeintegral})). Finally, all of these $n$ integrals
601: are identical giving a factor of $n$ which combines with the $(n-1)!$
602: to cancel the $n!$ in (\ref{freeintegral}) giving (\ref{KOintegral}).
603:
604: Summing up the terms in $W(\Psi,\mathcal{V})$ using (\ref{freeintegral}) gives
605:
606: \begin{equation}
607: W(\Psi,\mathcal{V}) = -\frac{1}{2\pi i}\left\langle \mathcal{V}(0) \,c(1)
608: \left[\exp\left(-\int_0^{2\pi} dt \,\lambda J(e^{it})\right) -1\right]\right\rangle_{\text{disk}} \ ,
609: \end{equation}
610: which, using (\ref{OnePointFunction}), is equivalent to
611: \begin{equation}
612: W(\Psi,\mathcal{V}) = \mathcal{A}^{\text{disk}}_{\Psi}(\mathcal{V}) - \mathcal{A}^\text{disk}_{0}(\mathcal{V}) \ .
613: \end{equation}
614: As defined in the introduction,
615: $\mathcal{A}^{\text{disk}}_{\Psi}(\mathcal{V})$ is the one-point
616: function with boundary conditions deformed by $\lambda J$.
617:
618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
619: \subsection{Invariants of marginal deformations with non-trivial OPE}
620: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
621:
622: The preceding argument can be extended to the case with non-trivial
623: OPE in the case when the OPE takes the form,
624: \begin{equation}\label{marginalOPE}
625: J(z) J(w) \sim \frac{1}{(z-w)^2} \ .
626: \end{equation}
627: The main change to the previous discussion is that the operators $J$
628: must be renormalized. There are, however, some subtleties which we
629: dwell on here that more general readers may not be interested in and
630: we encourage them to skip to the next subsection.
631:
632: In the non-trivial OPE case, the solution is again given in \cite{Fuchs:2007yy,Kiermaier:2007vu}.
633: We follow the notation of Kiermaier-Okawa \cite{Kiermaier:2007vu}. Before we can introduce their
634: state, we need to describe their renormalization scheme. This requires
635: a number of definitions which we now repeat:
636:
637: Let the Green's function on the cylinder be
638: denoted
639: \begin{equation}
640: G(y_1,y_2) = \langle J(y_1)J(y_2) \rangle \ ,
641: \end{equation}
642: and, following \cite{Kiermaier:2007vu}, define the normal ordered
643: product,
644: \begin{equation}\label{normalorder}
645: :\prod_{i = 1}^n J(y_i): = e^{-\frac{1}{2} \int dx_1 \,dx_2\, G(x_1,x_2) \frac{\delta}{\delta J(x_1)} \frac{\delta}{\delta J(x_2)} }
646: \prod_{i = 1}^n J(y_i) \ .
647: \end{equation}
648: The object $\int_a^b dy\, J(y)$ appears often enough that it is useful
649: to define \cite{Kiermaier:2007vu}
650: \begin{equation}
651: J(a,b) \equiv \int_a^b dy\, J(y) \ .
652: \end{equation}
653: To write down the marginal solution, we need to specify two
654: renormalized operators
655: \begin{equation}
656: \left[ e^{-\lambda \, J(a,b)} \right]_r \ , \qquad \left[ J(a) e^{-\lambda J(a,b)} \right]_r \ .
657: \end{equation}
658: To do this, we need the renormalized correlators
659: \cite{Kiermaier:2007vu},
660: \begin{align}
661: \label{V2ren}
662: \langle J(a,b)^2 \rangle_r &\equiv 2 \lim_{\epsilon \to 0} \left(
663: \int_a^{b-\epsilon} dy_1 \int_{t_1+\epsilon}^b dy_2\, G(y_1,y_2) - \frac{b-\epsilon-a}{\epsilon} - \log \epsilon
664: \right) \ ,
665: \\
666: \langle J(a) J(a,b)\rangle_r &\equiv \lim_{\epsilon \to 0}\left( \int_{a+\epsilon}^b dy\, G(a,y) - \frac{1}{\epsilon}\right) \ .
667: \end{align}
668: The full renormalized operators are given by
669: \cite{Kiermaier:2007vu}
670: \begin{align}
671: \left[ e^{-\lambda J(a,b)} \right]_r &\equiv e^{\frac{1}{2} \lambda^2 \langle J(a,b)^2 \rangle_r}
672: :e^{-\lambda J(a,b)}: \ ,
673: \\
674: \left[ J(a) e^{-\lambda J(a,b)} \right]_r &\equiv e^{\frac{1}{2} \lambda^2 \langle J(a,b)^2 \rangle_r}
675: :(J(a) - \lambda \langle J(a) J(a,b) \rangle_r )e^{-\lambda J(a,b)}: \ .
676: \end{align}
677: Note that these can be rewritten as
678: \begin{align} \label{renorm}
679: \left[ e^{\lambda J(a,b)} \right]_r = \lim_{\epsilon \to 0}
680: R_{\epsilon} \exp \left(- \lambda^2 (\log \epsilon-1)+
681: \int_a^b dy \, (-\lambda J(y) - \frac{1}{\epsilon} \lambda^2)
682: \right) \ ,
683: \\
684: \left[ J(a) e^{\lambda J(a,b)} \right]_r = \lim_{\epsilon \to 0}
685: R_{\epsilon} (J(a) + \frac{1}{\epsilon} \lambda)\exp \left(-\lambda^2 (\log \epsilon-1)+
686: \int_a^b dy \, (-\lambda J(y) - \frac{1}{\epsilon}\, \lambda^2)
687: \right) \ ,
688: \end{align}
689: where the operator $R_\epsilon$ removes all terms in which two $J$'s
690: are within $\epsilon$ of each other. A few comments may help clarify
691: these choices. Essentially, we are renormalizing $-\lambda J \to -\lambda J -
692: \frac{1}{\epsilon} \lambda^2$.
693: However, note the first term in the exponential, $\chi =
694: -\lambda^2(\log \epsilon -1)$, which comes from
695: $\log\epsilon$ and finite piece subtracted off in (\ref{V2ren}).
696:
697: The $e^\chi$ prefactor is unexpected from the point of view of the
698: renormalization of the boundary operator $J$ since only the
699: counterterm $\frac{1}{\epsilon} \lambda^2$ is needed in boundary
700: perturbation theory \cite{Recknagel:1998ih}. Fortunately, all
701: dependence on $\chi$ will drop out when the full solution is
702: assembled.
703:
704: We define the powers $J^{(n)}(a,b)$ through the expansions (absorbing, as
705: in \cite{Kiermaier:2007vu}, the factors of $n!$),
706: \begin{equation}
707: [e^{-\lambda J(a,b)}]_r = \sum_{n = 0}^\infty (-\lambda)^n [J^{(n)} (a,b)]_r \ , \qquad
708: [J(a) e^{-\lambda J(a,b)}]_r = \sum_{n = 0}^\infty (-\lambda)^n [J(a) J^{(n)} (a,b)]_r \ .
709: \end{equation}
710:
711: Define the states\footnote{To compare with \cite{Kiermaier:2007vu},
712: note that $A_L = A_0+\tilde A_0$.}
713: \begin{equation}
714: U_{\alpha} \equiv \sum_{n = 0}^\infty (-\lambda)^n U_{\alpha}^{(n)} \ , \qquad
715: A_{\alpha} = \sum_{n = 1}^\infty (-\lambda)^n A_{\alpha}^{(n)} \ , \qquad \tilde A_{\alpha} = \sum_{n= 2}^\infty (-\lambda)^{n}\tilde A_{\alpha}^{(n)} \ ,
716: \end{equation}
717: where
718: \begin{align}
719: \langle \phi| U_\alpha^{(n)} \rangle&= \langle f \circ \phi(0) \, [J^{(n)}(1,n+\alpha)]_r \rangle_{C_{n+\alpha+1}} \ ,
720: \\
721: \langle \phi| A_\alpha^{(n)} \rangle&= \langle f \circ \phi(0) \, [cJ(1) J^{(n-1)}(1,n+\alpha)]_r \rangle_{C_{n+\alpha+1}} \ ,
722: \\
723: \langle \phi| \tilde A_\alpha^{(n)} \rangle&= \tfrac{1}{2} \langle f \circ \phi(0) \, \partial c \,[ J^{(n-2)}(1,n+\alpha)]_r \rangle_{C_{n+\alpha+1}}
724: \ .
725: \end{align}
726: The complete marginal solution is given by\footnote{As in the trivial
727: OPE case, this solution does not satisfy the reality condition.
728: However, the real solution is once again gauge equivalent if we allow
729: complex gauge transformations.}
730: \begin{equation}
731: \Psi = -(A_{0}+\tilde A_{0}) U_0^{-1} \ .
732: \end{equation}
733: Conveniently, if one computes the contribution of $\tilde A_0
734: U_0^{-1}$ to $W(\Psi,\mathcal{V})$, it is proportional to the ghost
735: correlator,
736: \begin{equation}
737: \langle c\tilde c (i) \partial c(0) \rangle_{\text{UHP}} = 0 \ .
738: \end{equation}
739: Hence, we can ignore $\tilde{A}$ in our discussion and we need only compute
740: \begin{equation}
741: W(\Psi,\mathcal{V}) = W(-A_0 U^{-1}_0,\mathcal{V}) \ .
742: \end{equation}
743:
744: We now want to show that $A_0 U^{-1}_0$ contains only subtractions of
745: inverse powers of $\epsilon$ and that the contribution from $\chi =
746: -\lambda^2(\log \epsilon -1)$ does not enter. To do this, define a
747: new renormalization $[\,]'_{r}$ in which the $\log \epsilon$ and
748: finite piece in (\ref{V2ren}) are not subtracted,
749: \begin{align}
750: \left[ e^{-\lambda J(a,b)} \right]'_r &=
751: R_{\epsilon} \exp \left(
752: \int_a^b dy \, (-\lambda J(y) - \frac{1}{\epsilon}\, \lambda^2)
753: \right) \ ,
754: \\
755: \left[ J(a) e^{-\lambda J(a,b)} \right]'_r &=
756: R_{\epsilon} (J(a) + \frac{1}{\epsilon} \lambda)\exp \left( \int_a^b dy \, (-\lambda J(y) - \frac{1}{\epsilon}\, \lambda^2)
757: \right) \ .
758: \end{align}
759: Note that we can no longer take $\epsilon \to 0$ since these operators are not finite in that limit.
760: Next, define $U'_{\alpha}$ and $A'_{\alpha}$ to be the same as $U_{\alpha}$ and $A_{\alpha}$ except using $[\,]'_r $ instead of $[\,]_r$. We can express one in terms of the other as follows:
761: \begin{equation}
762: U = \sum_{n =0}^\infty \chi^{n} U_{2n}' \ ,\qquad A_0 = \sum_{n = 0}^{\infty} \chi^{n} A'_{2n} \ .
763: \end{equation}
764: We then have
765: \begin{multline}
766: A_0 U^{-1}_0 = \sum_{n = 0}^\infty \chi^n A_{2n}' (\sum_{m = 0}^\infty \chi^n U_{2m}')^{-1}
767: \\
768: = \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N\left( \prod_{i = 1}^N \sum_{k_i = 1}^\infty \right)
769: \chi^{n+k_1+\ldots +k_N}
770: A'_{2n} (U'_0)^{-1} \prod_{i = 1}^N U_{2k_i}'(U'_0)^{-1} \ .
771: \end{multline}
772: Using the identity \cite{Kiermaier:2007vu},
773: \begin{equation}
774: A'_{\alpha} (U'_0)^{-1} U'_{\beta} = A'_{\alpha+\beta} \ ,
775: \end{equation}
776: We find
777: \begin{equation}
778: \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N\left( \prod_{i = 1}^N \sum_{k_i = 1}^\infty \right)
779: \chi^{n+k_1+\ldots +k_N}
780: A'_{2n+2k_1+\ldots 2k_N} (U'_0)^{-1} \ .
781: \end{equation}
782: Note that the coefficient of $\chi^K A'_{2K}$ is
783: \begin{equation} \label{chicoeff}
784: \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N\left( \prod_{i = 1}^N \sum_{k_i = 1}^\infty \right)
785: \delta_{K,n+k_1+\ldots +k_N} \ .
786: \end{equation}
787: Replacing the Kronicker delta with a Dirac delta-function, we can write this as
788: \begin{multline}
789: \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N\left( \prod_{i = 1}^N \sum_{k_i = 1}^\infty \right)
790: \delta(K-(n+k_1+\ldots +k_N))
791: \\
792: = \int_{-\infty}^\infty dy
793: \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N\left( \prod_{i = 1}^N \sum_{k_i = 1}^\infty \right)
794: e^{i y(K-(n+k_1+\ldots +k_N))} \ .
795: \end{multline}
796: Performing the sums over $n$ and $k_i$ gives
797: \begin{equation}
798: \int_{-\infty}^\infty dy
799: \sum_{n = 0}^\infty\sum_{N = 0}^{\infty} (-1)^N e^{iy(1+K)}\left( \frac{1}{e^{iy} - 1}\right)^{N+1}
800: = \int_{-\infty}^\infty dy\, e^{i y K} = \delta(K) \ ,
801: \end{equation}
802: from which we learn (dividing by $\delta(0)$ if you will), that (\ref{chicoeff}) is just $\delta_{K,0}$.
803: We have found that
804: \begin{equation}
805: A_0 U_0^{-1} = A'_0 (U'_0)^{-1} \ ,
806: \end{equation}
807: so that all $\chi$ dependence has dropped out as promised. Note that,
808: since the left hand side is finite, the right hand side must be
809: finite. This useful fact, which can be verified at low orders, tells
810: us that no $\log \epsilon$ terms ever arise in the full form of
811: $\Psi$. This also implies that as far as $\Psi$ is concerned, we can
812: use $[ \, ]'_r$, which is the expected renormalization of $J$. We can
813: now write
814: \begin{multline}
815: \langle \phi|A_0' (U'_0)^{-1} \rangle
816: \\
817: =
818: \sum_{n= 1}^\infty (-\lambda)^n
819: \left\langle
820: f\circ \phi\, \int_1^2 dy_1 \int_{y_1}^2 dy_1 \, \ldots \int_{y_{n-2}}^n dy_{n-1}
821: \left[ cJ(0) J(y_1) J(y_2)\ldots J(y_{n-1}) \right]'_r \right\rangle_{C_{n+1}} \ .
822: \end{multline}
823: Inserting this state into $W$, the argument proceeds in the same manner as in the trivial OPE case. We find, simply
824: \begin{equation}
825: W(\Psi,\mathcal{V}) = -\frac{1}{2\pi i}\left\langle \mathcal{V}(0) \,c(1)
826: \left[\exp\left(-\int_0^{2\pi} dt \,\lambda J(e^{it})\right) -1\right]'_r\right\rangle_{\text{disk}} \ ,
827: \end{equation}
828: which, using (\ref{OnePointFunction}), gives
829: \begin{equation}
830: W(\Psi,\mathcal{V}) = \mathcal{A}^{\text{disk}}_{\Psi}(\mathcal{V}) - \mathcal{A}^\text{disk}_{0}(\mathcal{V}) \ .
831: \end{equation}
832: The only new feature here is that the boundary deformation generated by
833: $J$ has been renormalized using the appropriate counter term as
834: discussed in \cite{Recknagel:1998ih}.
835:
836: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
837: \subsection{Invariants of the tachyon vacuum}
838: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
839:
840: We can also compute the invariants for the tachyon vacuum solution. The tachyon vacuum state is given by \cite{Schnabl:2005gv}
841: \begin{equation}
842: \lim_{N\to \infty} \left( \psi_N - \sum_{n = 0}^N \partial_n \psi_n\right) \ ,
843: \end{equation}
844: where
845: \begin{equation}
846: \langle \phi | \psi_k \rangle = \left \langle [f\circ \phi](0)
847: \,c(-1) \left(\int_{-i\infty}^{i\infty} \frac{d\tilde z}{2\pi i} \,
848: b(\tilde z)\right) c(1)
849: \right\rangle_{C_{n+2}} \ .
850: \end{equation}
851:
852: The invariant $W(\psi_n,c \bar c \mathcal{O}^{\text{m}})$ is given by
853: \begin{equation}
854: \left\langle c(i\infty)c(-i\infty) \mathcal{O}^\text{m}(i\infty) \,\,
855: c(n/2) \left(\int_{-i\infty}^{i\infty} \frac{d\tilde z}{2\pi i}\, b(\tilde z)\right) c(-n/2) \right\rangle_{C_{n+1}} \ .
856: \end{equation}
857: Applying
858: \begin{equation}
859: g(\tilde z) = \tan\left( \frac{\pi \tilde z}{n+1} \right) \ ,
860: \end{equation}
861: we get
862: \begin{equation}
863: \frac{n+1}{\pi} \frac{1}{(1+x^2)^2}
864: \left\langle
865: c(i)c(-i) \mathcal{O}^\text{m}(i) \,\,
866: c(x)
867: \left(
868: \int_{-i\infty}^{i\infty} \frac{dz}{2\pi i}\,(1+z^2) b(z)
869: \right)
870: c(-x)
871: \right\rangle_{\text{UHP}} \ ,
872: \end{equation}
873: where $x = \tan(\frac{\pi}{2} \frac{n}{n+1})$.
874: Evaluating the ghost correlator, this reduces to
875: \begin{equation}
876: W(\psi_n,c \tilde c\mathcal{O}^\text{m})= \frac{2i}{\pi}\langle \mathcal{O}^\text{m}(i) \rangle^\text{m}_{\text{UHP}} \ .
877: \end{equation}
878: Remarkably, this is independent of $n$. It follows that
879: \begin{equation}
880: W (\Psi,\mathcal{O}^{\text{m}}) = \lim_{N\to \infty} W (\psi_N - \sum_n \partial_n\psi_n,\mathcal{O}^{\text{m}}) = \lim_{N\to \infty} W (\psi_N,\mathcal{O}^{\text{m}}) \ ,
881: \end{equation}
882: which we can write as
883: \begin{equation} \label{tachyonvacinvariant}
884: \frac{2 i }{\pi} \langle \mathcal{O}^{\text{m}}(i) \rangle^{\text{m}}_{\text{UHP}} = \frac{1}{\pi} \langle c\tilde c \mathcal{O}(i) c(0)\rangle_{\text{UHP}} = \frac{1}{2\pi i} \langle \mathcal{V}(0) c(1) \rangle_{\text{disk}} = -\mathcal{A}^{\text{disk}}_0(\mathcal{V}) \ .
885: \end{equation}
886: It might seem surprising that the terms $\partial_n \psi_n$ would make
887: no contribution. The reason for this simplification is that the sum,
888: $-\sum \lambda^n \partial_n \psi_n$ is a pure gauge state for $\lambda
889: <1$. Since $W$ is gauge invariant, it follows that $W(\partial_n
890: \psi_n,\mathcal{O}^{\text{m}})$ must vanish for every $n$.
891:
892: The result (\ref{tachyonvacinvariant}) should be interpreted as
893: \begin{equation}
894: W (\Psi,\mathcal{O}^{\text{m}})
895: = \mathcal{A}_{\Psi}^\text{disk}(\mathcal{V})- \mathcal{A}_{0}^\text{disk}(\mathcal{V})\ ,
896: \end{equation}
897: where $\mathcal{A}_{\Psi}^\text{disk}(\mathcal{V})=0$ since there is
898: no source for closed strings in the tachyon vacuum.
899:
900:
901: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
902: \section{Derivation of the invariants from BRST invariance}\label{Derivation}
903: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
904:
905: Having seen in two examples that $W(\Psi,\mathcal{V})$ computes the
906: closed string tadpole, it is desirable to find a general derivation of
907: this result.
908:
909: Naively, one should begin with the usual method for finding a string
910: field theory diagram for a given amplitude: Open string field theory
911: diagrams are given by picking a minimal metric on the worldsheet
912: subject to the condition that any non-contractible Jordan open curves
913: have length at least $\pi$ \cite{Zwiebach:1990az}. For a disk with one closed string
914: insertion and no open string insertions, however, there are no
915: non-contractible curves and the minimal metric surface has zero size.
916: Furthermore, including a background string field, representing a
917: change in the disk boundary conditions, it is not clear how to find
918: the appropriate minimal metric.
919:
920: Although this direct approach fails, one can still try to use an
921: argument from BRST invariance: Consider a disk with {\em two} closed
922: string insertions and take the limit as the two insertions become
923: close together. In this limit, the diagram is conformally equivalent
924: to a diagram in which the two closed string insertions are connected
925: to the boundary of the disk by a long tube. If we pick the momenta of
926: the two closed string insertions such that the intermediate closed
927: string state is on-shell, this long tube will lead to a divergence
928: when we integrate over its length. Conveniently, this divergence
929: gives rise to a BRST anomaly\footnote{Note that the diagram is neither
930: divergent, nor anomalous for generic momenta \cite{Klebanov:1995ni,Hashimoto:1996bf}. See
931: \cite{Callan:1987px,Polchinski:1987tu} for a general discussion of how tadpoles can arise as surface terms in moduli space.} which is proportional
932: to the closed string tadpole diagram.
933:
934:
935: \begin{figure}
936: \centerline{
937: \begin{picture}(413,109)
938: \includegraphics{CSTwoPoint}
939: \end{picture}
940: \begin{picture}(0,0)(413,0)
941: \put(-18,49){$\mathcal{V}_1$}
942: \put(189,49){$\mathcal{V}_2$}
943: \put(210,72){$\mathcal{V}_1$}
944: \put(412,72){$\mathcal{V}_2$}
945: \put(87,-1){$T$}
946: \put(313,8){$T$}
947: \put(40,46){$b$}
948: \put(129,49){$\pi$}
949: \put(-30,95){\bf a)}
950: \put(200,95){\bf b)}
951: \end{picture}
952: }
953: \caption{ The closed string two-point function in the string field
954: theory conformal frame. There is one modulus, $T$, which is integrated
955: from $0$ to $\infty$. There is a single ghost insertion given by an
956: integral of the $b$-ghost over the red line. In a), the geometry is
957: shown as a flat strip with two identifications given by the hatches on
958: the right and left. In b), the same geometry is shown in after the
959: identifications are performed. Note the conical singularities at the
960: closed string insertions. For consistency, $\mathcal{V}_{1,2}$ must
961: be weight $(0,0)$.\label{CSTwoPoint}}
962: \end{figure}
963:
964: The closed string two-point function on the disk in the conformal
965: frame appropriate to string field theory is shown in figure
966: \ref{CSTwoPoint}
967: \cite{Shapiro:1987gq,Shapiro:1987ac,Freedman:1987fr,Zwiebach:1990az,Hashimoto:2001sm,Takahashi:2003kq,Garousi:2004pi}. The
968: amplitude is given by\footnote{Computations of the closed string
969: two-point function in open string field theory include
970: \cite{Takahashi:2003kq,Garousi:2004pi}.}
971: \begin{equation}\label{TwoPointFunction}
972: \mathcal{A}(\mathcal{V}_1,\mathcal{V}_2) =
973: \langle \mathcal{I}|
974: \mathcal{V}_1(i)\,
975: b_0
976: \int_{\epsilon/2}^{\infty} dT\,e^{-L_0 T}
977: \mathcal{V}_2(i)
978: |\mathcal{I}\rangle \ ,
979: \end{equation}
980: where $\epsilon$ is a UV cutoff on the worldsheet, but an IR cutoff in
981: spacetime. That this diagram is given by a propagator sandwiched
982: between two states will be very convenient when we repeat this
983: computation with a background open string field.
984:
985: To see the origin of the BRST anomaly, consider the case
986: when $\mathcal{V}_1 = Q_B \mathcal{O}$. We then find,
987: \begin{multline}
988: \langle \mathcal{I}|
989: [Q_B,\mathcal{O}(i)]\,
990: b_0
991: \int_{\epsilon/2}^{\infty} dT\,e^{-L_0 T}
992: \mathcal{V}_2(i)
993: |\mathcal{I}\rangle
994: \\
995: = -\langle \mathcal{I}|
996: \mathcal{O}(i)\,
997: \{Q_B, b_0
998: \int_{\epsilon/2}^{\infty} dT\,e^{-L_0 T} \}
999: \mathcal{V}_2(i)
1000: |\mathcal{I}\rangle
1001: = -\langle \mathcal{I}|
1002: \mathcal{O}(i)\,
1003: \,e^{-L_0 T}\biggr|_{T = \epsilon/2}^\infty \mathcal{V}_2(i)
1004: |\mathcal{I}\rangle \ ,
1005: \end{multline}
1006: where we have used the properties $\{Q_B,b_0\} = L_0$ and the on-shell
1007: condition $\{Q_B,\mathcal{V}_2\} = 0$ as well as $Q_B
1008: |\mathcal{I}\rangle = 0$. The contributions at $T \to \infty$ are not
1009: relevant for the current discussion. Dropping them gives
1010: \begin{equation} \label{SurfaceTerm}
1011: \langle \mathcal{I}|
1012: \mathcal{O}(i)\,
1013: \,e^{-L_0 \epsilon/2} \mathcal{V}_2(i)
1014: |\mathcal{I}\rangle \ .
1015: \end{equation}
1016: This amplitude is shown in figure \ref{OPEReplace}a. Since $\epsilon$
1017: is assumed to be very small, we may replace the two insertions of
1018: $\mathcal{O}$ and $\mathcal{V}_2$ with their OPE, giving the geometry
1019: in figure \ref{OPEReplace}b. The geometry is considerably simplified.
1020: We now have a closed string state, $|\Omega \rangle$ coming in
1021: from in infinity and ending on a boundary. Note that the OPE could
1022: have singular terms since we are in a theory with tachyons. Such terms
1023: correspond to propagation of the tachyon over long distances and
1024: should be removed either by analytic continuation or explicit
1025: subtraction. In the absence of singularities, it follows that
1026: $(L_0+\tilde{L}_0) |\Omega \rangle = 0$. Note that if the OPE contains
1027: no finite piece, the surface term vanishes. This is why
1028: $\mathcal{O}$ and $\mathcal{V}_2$ must be tuned so that the intermediate
1029: closed string state is on-shell.
1030:
1031:
1032: \begin{figure}
1033: \centerline{
1034: \begin{picture}(265,114)
1035: \includegraphics{TtoZeroLim}
1036: \end{picture}
1037: \begin{picture}(0,0)(265,0)
1038: \put(-20,73){$\mathcal{O}$}
1039: \put(-20,108){$\mathcal{V}_2$}
1040: \put(104,60){$\pi$}
1041: \put(248,91){$\epsilon$}
1042: \put(248,18){$\epsilon$}
1043: \put(-27,19){$\langle \Omega |$}
1044: \put(-60,98){\bf a)}
1045: \put(-60,28){\bf b)}
1046: \end{picture}
1047: }
1048: \caption{The surface term from replacing $\mathcal{V}_2 =
1049: [Q_B,\mathcal{O}]$. In a), the amplitude (\ref{SurfaceTerm}) is
1050: shown. In b) the two closed string insertions are replaced with their
1051: OPE. \label{OPEReplace}}
1052: \end{figure}
1053:
1054: Since $\Omega$ is overlapped with the $L_0+\tilde{L}_0 = 0$ part of
1055: the boundary state, which is in the cohomology of $Q_B$, we may drop
1056: the parts of $\Omega$ which are not physical; Hence, we may take\footnote{We are assuming that it is possible to divide the closed string fock space into two orthogonal pieces $\mathcal{H}_{\text{CFT}} = \mathcal{H}_{\text{coh}} \oplus \mathcal{H}_{\text{rest}}$ with the weight zero piece of the boundary state in the $Q_B$-cohomology, $\mathcal{H}_{\text{coh}}$. Note that we have not shown that an {\em arbitrary} element of $\mathcal{H}_{\text{coh}}$ can be created from the OPE of the states $\mathcal{O}$ and $\mathcal{V}_2$, which would be required for a complete derivation.}
1057: $\{Q_B,\Omega \} = 0$. This is the closed string one-point
1058: function which we wished to compute. The point of this exercise is
1059: that when we turn on an open string vev, we can repeat the same
1060: computation to find the one-point function in the presence of an open
1061: string field background.
1062:
1063: When we shift the vacuum $\Psi \to \Psi + \Psi_{cl}$, the only change
1064: in the open string field theory action is a shift in the BRST operator,
1065: \begin{equation}
1066: Q_B \to Q_B + [\Psi_{\text{cl}},\quad] \ .
1067: \end{equation}
1068: This introduces a term,
1069: \begin{equation}
1070: \int \Psi * \Psi*\Psi_{\text{cl}} \ ,
1071: \end{equation}
1072: in the action which shifts the propagator. The new propagator is given by
1073: summing over all the ways of inserting $\Psi_{\text{cl}}$ into the old
1074: propagator together with the appropriate ghost insertions. This is illustrated in figure \ref{Propagator}.
1075:
1076:
1077: \begin{figure}
1078: \centerline{
1079: \begin{picture}(448,148)
1080: \includegraphics{Propagator}
1081: \end{picture}
1082: \begin{picture}(0,0)(448,0)
1083: \put(-10,130){\bf a)}
1084: \put(235,130){\bf b)}
1085: \put(5,16){$T_1$}
1086: \put(38,16){$T_2$}
1087: \put(85,16){$T_3$}
1088: \put(145,16){$T_4$}
1089: \put(195,16){$T_5$}
1090: \end{picture}
1091: }
1092: \caption{The propagator in the presence of a open string field vev shown with four insertions of $\Psi_{\text{cl}}$. In
1093: a), the insertions of $\Psi_{\text{cl}}$ are represented by cuts in
1094: the worldsheet. As shown in b), to get the full worldsheet geometry,
1095: one must glue an infinitely long strip into each cut.
1096: Each insertion of $\Psi_{\text{cl}}$ introduces one extra moduli in
1097: addition to the modulus of the overall length of the propagator. With
1098: each modulus, one must add an integral of $b$ -- as shown in red in a) --
1099: in order to get the right measure on moduli space.
1100: \label{Propagator}}
1101: \end{figure}
1102:
1103: Algebraically, the propagator between states $|A\rangle$ and $|B\rangle$
1104: can be written as follows. Define the adjoint action of $\Psi_{\text{cl}}$ by
1105: \begin{equation}
1106: \ad_{\Psi_{\text{cl}}} \Phi = \Psi_{\text{cl}}*\Phi - (-1)^{\text{gh}(\Phi)} \Phi*\Psi_{\text{cl}} \ ,
1107: \end{equation}
1108: and
1109: \begin{equation}
1110: D = \int_0^\infty dT \, e^{-T L_0} \ .
1111: \end{equation}
1112: Then the full propagator is given by
1113: \begin{equation}
1114: \sum_{n = 0}^{\infty}
1115: \langle A| b_0 D \left(\ad_{\Psi_{\text{cl}}}\, b_0 D\right)^n |B\rangle \ .
1116: \end{equation}
1117:
1118:
1119:
1120:
1121: Given the propagator in the presence of $\Psi_{\text{cl}}$ one can
1122: compute the modified closed-string two point function by replacing
1123: the old propagator in (\ref{TwoPointFunction}) with the new one,
1124: \begin{equation} \label{TwoPointFull}
1125: \mathcal{A}_\Psi(\mathcal{V}_1,\mathcal{V}_2) =
1126: \sum_{n = 0}^\infty\langle \mathcal{I}|\, \mathcal{V}_1(i)
1127: \,\, b_0 D \left( \ad_{\Psi_{\text{cl}}} b_0 D \right)^n \mathcal{V}_2 (i)|\mathcal{I}\rangle
1128: \end{equation}
1129: To extract the one-point function, again replace $\mathcal{V}_1 =
1130: \{Q_B,\mathcal{O}\}$. After some algebra and using the equations of motion for
1131: $\Psi_{\text{cl}}$ one finds (see appendix \ref{SurfaceTermComp} for details):
1132: \begin{equation} \label{TwoPointFullSurface}
1133: -\int_{\epsilon/2}^\infty dT\,\frac{\partial}{\partial T}
1134: \sum_{n = 0}^\infty\left( \prod_{i = 0}^n \int_0^\infty dT_{n}\right)\, \delta (T-\sum_{i = 0}^n T_i)
1135: \langle \mathcal{I}|\, \mathcal{O}(i)
1136: \,\, D_{T_0} \left( \prod_{i = 1}^n \{b_0, \ad_{\Psi_{\text{cl}}}\} D_{T_i} \right) \mathcal{V}_2(i)|\mathcal{I}\rangle \ ,
1137: \end{equation}
1138: where
1139: \begin{figure}
1140: \centerline{
1141: \begin{picture}(426,206)(0,-10)
1142: \includegraphics{SurfaceTermWithPsi}
1143: \end{picture}
1144: \begin{picture}(0,0)(426,-10)
1145: \put(-20,153){$\mathcal{O}$}
1146: \put(-20,45){$\mathcal{O}$}
1147: \put(228,153){$\mathcal{V}_2$}
1148: \put(228,45){$\mathcal{V}_2$}
1149: \put(10,95){$T_1$}
1150: \put(43,95){$T_2$}
1151: \put(90,95){$T_3$}
1152: \put(150,95){$T_4$}
1153: \put(200,95){$T_5$}
1154: \put(10,-12){$T_1$}
1155: \put(43,-12){$T_2$}
1156: \put(90,-12){$T_3$}
1157: \put(150,-12){$T_4$}
1158: \put(200,-12){$T_5$}
1159: \put(270,180){$\mathcal{O}$}
1160: \put(310,180){$\mathcal{V}_2$}
1161: \put(381,190){$\langle \Omega |$}
1162: \put(243,100){$\pi/2$}
1163: \put(-25,192){\bf a)}
1164: \put(-25,86){\bf b)}
1165: \put(230,192){\bf c)}
1166: \put(336,192){\bf d)}
1167: \end{picture}
1168: }
1169: \caption{Various representations of the surface term are shown for the
1170: case of four insertions. In a), a representation of
1171: (\ref{surfaceTerm}) is given. It is assumed that $\sum T_i =
1172: \epsilon/2$. In this form the $\epsilon\to 0$ limit is difficult because
1173: the operators $\mathcal{O}$ and $\mathcal{V}_2$ collide with the ends
1174: of the cuts and the $b$-ghost insertions. In b) a reparametrization for
1175: the classical solution $\Psi_{\text{cl}}$ is used so that the cuts
1176: do no reach the midpoint of the string. Performing the identifications
1177: in b) produces the diagram c) which now has a long tube separating the
1178: operators $\mathcal{O}$ and $\mathcal{V}_2$ from the cuts. As shown
1179: in d), when $\epsilon$ is small we can replace the top of the diagram with a
1180: single closed string state, $|\Omega\rangle$.
1181: \label{SurfaceTermWithPsi}}
1182: \end{figure}
1183: \begin{equation}
1184: D_{T_i} = b_0 e^{-T_i L_0} \ .
1185: \end{equation}
1186: This leads to the surface term,
1187: \begin{equation} \label{surfaceTerm}
1188: \sum_{n = 0}^\infty\left( \prod_{i = 0}^n \int_0^\infty dT_{n}\right)\, \delta (\epsilon/2-\sum_{i = 0}^n T_i)
1189: \langle \mathcal{I}|\, \mathcal{O}(i)
1190: \,\, D_{T_0} \left( \prod_{i = 1}^n \{b_0, \ad_{\Psi_{\text{cl}}}\} D_{T_i} \right) \mathcal{V}_2(i)|\mathcal{I}\rangle \ .
1191: \end{equation}
1192: Geometrically, this amplitude is given by figure
1193: \ref{SurfaceTermWithPsi}a. It is important to point out that that the
1194: cutoff $\epsilon$ is not conformally/BRST invariant so the expression
1195: (\ref{surfaceTerm}) is not invariant under gauge transformations of
1196: $\Psi_{\text{cl}}$ except in the limit $\epsilon \to 0$.
1197:
1198:
1199:
1200:
1201: Unlike in the case without an open string background, it is not clear that, when $\epsilon$ is very small in
1202: (\ref{surfaceTerm}), one can replace $\mathcal{O}$ and
1203: $\mathcal{V}_2$ with their OPE. The problem is that the two closed
1204: string operators are not separated from the rest of the geometry by a
1205: long tube. Instead, the midpoints of the $\Psi_{\text{cl}}$ insertions and
1206: integrals of the $b$-ghost all remain close to the closed string
1207: insertions.
1208:
1209: To fix this problem, one can perform a gauge transformation of
1210: $\Psi_{\text{cl}}$ which reduces its height. This reparametrizaion,
1211: which is discussed in appendix \ref{reparamdiscussion}, allows one to
1212: make a cut in the propagator which is some height $h<\pi/2$ and insert strip representing $\Psi_{\text{cl}}$ which has
1213: been shrunk by a factor of $2h/\pi$. Since, as mentioned above, the
1214: amplitude is not invariant under gauge transformations for finite
1215: $\epsilon$, this step may seem suspicious. However, as will be seen
1216: in a moment, gauge invariance will be restored in the small $\epsilon$
1217: limit and the dependence on $h$ will drop out.
1218:
1219:
1220: The amplitude with the gauge transformed $\Psi_{\text{cl}}$'s is shown
1221: in figure \ref{SurfaceTermWithPsi}b. Performing the identifications
1222: leads to a geometry shown in figure \ref{SurfaceTermWithPsi}c. As can
1223: be seen from the figure, there is now a long tube separating the
1224: closed string insertions from the rest of the geometry so one may
1225: replace them with their OPE as shown in figure
1226: \ref{SurfaceTermWithPsi}d. One can then check that, assuming we can
1227: drop the non-physical parts of $\Omega$, so that $Q_B |\Omega \rangle
1228: = 0$, the gauge invariance $\Psi_{\text{cl}}\to \Psi_{\text{cl}} + Q_B
1229: \Lambda + [\Psi_{\text{cl}},\Lambda]$ is restored\footnote{It is nice
1230: to have an independent check that this amplitude is the closed string
1231: one-point function. Here is a sketch of an alternate argument: since
1232: gauge invariance is restored, we can reparametrize the width of the
1233: state $\Psi_{\text{cl}}$ to limit it to an identity state with a
1234: single operator $c \mathcal{O}$ inserted on the boundary. Using the
1235: $b$-integrals to remove the $c$ ghost, we are left with a disk with
1236: the boundary deformation $\exp(\int \mathcal{O})$. As one can check
1237: in simple cases, this typically generates the renormalized boundary
1238: deformation associated with the state $\Psi_{\text{cl}}$ so that the
1239: diagram reduces to $\mathcal{A}_{\Psi}^{\text{disk}} (\mathcal{V})$.}.
1240:
1241: By unitarity, the amplitude pictured in figure
1242: \ref{SurfaceTermWithPsi}d should be the closed string one-point
1243: function on a disk with boundary conditions CFT${}_{\Psi_{\text{cl}}}$.
1244: We may suppose, without loss of generality, that
1245: \begin{equation}
1246: \Omega = (\partial c - \bar \partial \tilde c)c \tilde c \mathcal{O}^{\text{m}} \ ,
1247: \end{equation}
1248: where $\mathcal{O}^{\text{m}}$ is a weight $(1,1)$ primary. Set
1249: $c\tilde c \mathcal{O}^{\text{m}} = \mathcal{V}$. The vertex operator
1250: $\Omega$ is ghost number 3. The extra ghostnumber corresponds to
1251: fixing the CKV corresponding to the rotation of the cylinder. To write
1252: the amplitude in terms the standard ghostnumber 2 operator
1253: $\mathcal{V}$, pull one of the $b$-ghost integrals off of the bottom
1254: of the cylinder and push it up till it encircles the state $|\Omega\rangle$.
1255: Next, let the $b$-ghost integral act on $|\Omega\rangle$ giving
1256: \begin{equation}
1257: \frac{2\pi}{\epsilon }(b_0 - \tilde{b}_0) |\Omega \rangle =
1258: \frac{2\pi}{\epsilon} |\mathcal{V}\rangle \ .
1259: \end{equation}
1260: The $\epsilon^{-1}$ can be used to fix the location of the cut
1261: whose $b$-ghost integral we removed since, by rotational invariance,
1262: the integral over its position just gives a factor of $\epsilon$.
1263:
1264:
1265:
1266:
1267: At this point, the amplitude still bears little resemblance to the
1268: invariants $W(\Psi,\mathcal{V})$. However, it turns out that by
1269: simultaneously increasing the height $h$ of the insertions {\em and}
1270: rescaling the wedge width on which the state $\Psi_{\text{cl}}$ is
1271: defined, the amplitude dramatically simplifies. To see why, consider
1272: the state $\Psi_{cl}$ to be defined in the $\arctan(z)$ coordinates.
1273: To map $\Psi_{cl}$ to the strip coordinates appropriate for gluing
1274: $\Psi_{\text{cl}}$ to the cylinder, we should use
1275: \begin{equation}
1276: \xi(z) =\frac{2h}{\pi}\log (\tan (\frac{\pi \tilde z}{2})),
1277: \end{equation}
1278: where the factor of $h$ accounts for the change in height of the insertion. Suppose
1279: that, in addition to changing the height of the solution, we also reparametrize it
1280: by changing its {\em width}. This can be accomplished by rescaling the state using $\tilde z\to \rho \tilde z$
1281: while leaving the coordinate patch alone. This is the standard reparametrization of the
1282: wedge width discussed, for example in \cite{Schnabl:2002gg,Schnabl:2002ff,Rastelli:2006ap,Okawa:2006sn}.
1283:
1284: \begin{figure}
1285: \centerline{
1286: \begin{picture}(353,122)(0,-10)
1287: \includegraphics{widthReparam}
1288: \end{picture}
1289: \begin{picture}(0,0)(353,-10)
1290: \put(60,-10){$f\circ \phi$}
1291: \put(142,-10){$1$}
1292: \put(-10,-10){$-1$}
1293: \put(277,-10){$f\circ \phi$}
1294: \put(337,-10){$\frac{1+\rho}{2}$}
1295: \put(220,-10){$-\frac{1+\rho}{2}$}
1296: \put(-25,100){\bf a)}
1297: \put(210,100){\bf b)}
1298: \end{picture}
1299: }
1300: \caption{Reparametrization of the wedge width. In a), a standard
1301: state is given in the $\arctan(z)$ coordinates. In b), the state is
1302: shrunk by a factor of $\rho$ while the coordinate patch is left alone
1303: giving a cylinder of width $1+\rho$.
1304: \label{widthReparam}}
1305: \end{figure}
1306:
1307:
1308: In detail, suppose we take the original state to be defined on a
1309: cylinder of circumference $2$ as shown in figure \ref{widthReparam}a.
1310: Shrinking the wedge width by taking $\tilde z\to \rho \tilde z$ while leaving the
1311: coordinate patch alone defines a new state $\Psi_{\text{cl}}'$ which
1312: is shown in figure \ref{widthReparam}b. The full map from the
1313: original state $\Psi_{\text{cl}}$ to the coordinates we are using for gluing is
1314: then given by
1315: \begin{equation}
1316: \xi'(z) = \frac{2h}{\pi}\log \left[\tan \left(\tfrac{\pi}{2} ((\tilde z-\tfrac{1}{2}) \rho + \tfrac{1}{2})\right)\right] \ .
1317: \end{equation}
1318: The limit we are interested in is taking $h \to \infty$ with $\rho =
1319: 1/2h$. Focusing on the region of worldsheet near $\tilde z = 1/2$ (to avoid the branchcut of the $\log$), one can verify that
1320: \begin{equation}
1321: \lim_{h\to \infty} \frac{2h}{\pi}\log \left[\tan \left(\tfrac{\pi}{2} ((\tilde z-\tfrac{1}{2})\tfrac{1}{2h} + \tfrac{1}{2})\right)\right] = \tilde z-\frac{1}{2} \ ,
1322: \end{equation}
1323: which is just a simple translation. In other words, in the limit $h
1324: \to \infty$ $\rho \to 0$, $h\rho =1/2$, a state $\Psi_{\text{cl}}$ as defined
1325: in the $\arctan(z)$ coordinates should be inserted into the cylinder
1326: geometry by cutting a infinite vertical strip in the cylinder and
1327: gluing in $\Psi_{\text{cl}}$ {\em without any conformal
1328: transformations}. The general picture is shown in figure \ref{FlattenedGeometry}\footnote{This representation of a string field theory amplitude is reminiscent of \cite{Fuji:2006me,Rastelli:2007gg,Kiermaier:2007jg}.}
1329:
1330: In the resulting geometry, the integrals over the $b$-ghost just become
1331: the operator $B_1 = b_{-1}+b_{1}$, which, in cylinder coordinates, is
1332: \begin{equation}
1333: \arctan \circ B_1 = \oint \frac{d\tilde z}{2\pi i} b(\tilde z) \ .
1334: \end{equation}
1335: The important point to note is that using the double gauge
1336: transformation, we have flattened out the conical singularities that
1337: arose from inserting $\Psi_{\text{cl}}$ into the cylinder geometry.
1338: This allows one to act with $B_1$ on $\Psi_{\text{cl}}$ in the
1339: obvious way.
1340:
1341: One might worry about two problems in this limit: First, although the
1342: curvature singularities are disappearing as we increase the height and
1343: decrease wedge width, we are nonetheless bringing a curvature
1344: singularity near the insertion of $\mathcal{V}$. We believe that,
1345: because $\mathcal{V}$ is a weight zero primary, there should be no
1346: divergences from this limit. Second, increasing the height of the
1347: insertions pushes the contour integrals of $b$ close to $\mathcal{V}$.
1348: Here again we believe there should be no singularity since the
1349: $b$-integral contours can be made to go through $\mathcal{V}$ without
1350: any divergence as can be checked by mapping the geometry to a disk.
1351: (Note that this would not have been true before we removed a
1352: $b$-integral from one of the $\Psi_{\text{cl}}$ insertions and let it
1353: act on the closed string state). We fully admit, however that this
1354: double reparametrizaion is delicate and additional operators inside the state
1355: $\Psi_{\text{cl}}$ could also create potential divergences.
1356:
1357: \begin{figure}
1358: \centerline{
1359: \begin{picture}(263,177)(-30,-10)
1360: \includegraphics{FlattenedGeometry}
1361: \end{picture}
1362: \begin{picture}(0,0)(233,-10)
1363: \put(12,20){$\Psi_{\text{cl}}$}
1364: \put(60,-12){$T_1$}
1365: \put(93,20){$B_1 \Psi_{\text{cl}}$}
1366: \put(144,-12){$T_2$}
1367: \put(175,20){$B_1 \Psi_{\text{cl}}$}
1368: \put(227,-12){$T_3$}
1369: \put(120,170){$|\mathcal{V}\rangle$}
1370: \put(-120,86){\parbox{1in}{\begin{equation*}\int_0^\infty dT_i\,\,\delta(\sum_{i} T_i - \epsilon)\end{equation*}}}
1371: \end{picture}
1372: }
1373: \caption{The resulting geometry for the case of three
1374: $\Psi_{\text{cl}}$ insertions after flattening the insertions of
1375: $\Psi_{\text{cl}}$ using a double gauge transformation. The field
1376: $\Psi_{\text{cl}}$ is now inserted into the geometry in the $\arctan$
1377: coordinates. The $b$-ghost integrals have become $B_1$'s acting on
1378: all but one of the $\Psi_{\text{cl}}$'s.
1379: \label{FlattenedGeometry}}
1380: \end{figure}
1381:
1382: With these caveats in mind, consider taking the $\epsilon \to 0$
1383: limit. First, note that the worldsheet does not become singular
1384: anywhere in this limit since the $\Psi_{\text{cl}}$ insertions can be
1385: assumed to have a finite minimum thickness. Furthermore, there are no
1386: singularties when $\Psi_{\text{cl}}$ insertions become close as $B_1
1387: \Psi_{\text{cl}} * \Psi_{\text{cl}}$ and $ \Psi_{\text{cl}} * B_1
1388: \Psi_{\text{cl}}$ are finite\footnote{This is true at least for the
1389: known solutions. Since there is, at present, no general ``regularity
1390: condition'' on the string field, we cannot say if this assumption is
1391: always true, even if it seems reasonable.}. However, the
1392: integration regions go to zero size in this limit, so each term with
1393: more than one $\Psi_{\text{cl}}$ will vanish.
1394:
1395: The only terms that remain, are the case with one $\Psi_{\text{cl}}$ which
1396: we recognize as the invariant $W(\Psi_{\text{cl}},\mathcal{V})$ and the case with no $\Psi_{\text{cl}}$'s
1397: which is just the one-point function with $\Psi_{\text{cl}} =0 $.
1398: Hence, we have found
1399: \begin{equation}
1400: \mathcal{A}_{\Psi}^{\text{disk}}(\mathcal{V}) = \mathcal{A}_{0}^{\text{disk}}(\mathcal{V}) + W(\Psi,\mathcal{V}) \ ,
1401: \end{equation}
1402: which reproduces (\ref{MainRelation}).
1403:
1404:
1405: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1406: \section{Extension to Berkovits' open superstring field theory} \label{susyversion}
1407: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1408:
1409: In this section, the extension to the Berkovits
1410: superstring field theory of the invariants $W(\Psi,\mathcal{V})$ is discussed. The invariants are computed for
1411: the case of marginal deformations with trivial OPE, yielding a
1412: formula for the invariants in terms of the closed string one-point
1413: function analogous to the bosonic case.
1414:
1415: \subsection{A gauge-invariant observable for the superstring}
1416:
1417: To extend to the superstring case, one needs an object which is
1418: invariant under the modified gauge trasformation,
1419: \begin{equation} \label{supergauge}
1420: e^{\Phi} \to e^{Q_B \Lambda} e^{\Phi} e^{\eta_0 \Lambda'} \ ,
1421: \end{equation}
1422: where $\Lambda$ and $\Lambda'$ are two gauge parameters. Such an
1423: invariant was written down in \cite{Michishita:2004rx}. Here we take
1424: a slightly different, but equivalent, approach\footnote{The invariant
1425: written down in \cite{Michishita:2004rx} is simply $\langle
1426: \mathcal{I}|\mathcal{V}(i) |\Phi\rangle$, with $Q_B \mathcal{V} =
1427: \eta_0 \mathcal{V} = 0$. Our invariant gives $\langle
1428: \mathcal{I}|\mathcal{V}(i) |e^{-\Phi}Q_B e^{\Phi}\rangle = \langle
1429: \mathcal{I}|\mathcal{V}(i) |Q_B \Phi\rangle = \langle
1430: \mathcal{I}|\{Q_B,\mathcal{V}(i)\} |\Phi\rangle$, which, given our
1431: assumptions on $\mathcal{V}$, reduces to the same thing. The
1432: advantage of our form comes from the fact that many superstring
1433: solutions are found by guessing $\Omega$ and then later finding
1434: $\Phi$, which is often much more complicated.}.
1435:
1436: Define
1437: \begin{equation} \label{omegadef}
1438: \Omega = e^{-\Phi} Q_B e^{\Phi}.
1439: \end{equation}
1440: The field $\Omega$ transforms under (\ref{supergauge}) as
1441: \begin{equation} \label{omegatransformation}
1442: \Omega \to e^{-\eta_0\Lambda'} \left(\Omega + Q_B\right) e^{\eta_0 \Lambda'} \ .
1443: \end{equation}
1444: Notice that it is invariant under the transformations generated by
1445: $\Lambda$. Consider the object,
1446: \begin{equation}\label{susyinvariant}
1447: \widehat{W}(\Phi,\mathcal{V}) = \langle \mathcal{I} | \mathcal{V}(i)
1448: | \Omega(\Phi)\rangle \ ,
1449: \end{equation}
1450: where $\mathcal{V}$ is a weight $(0,0)$ primary. If $\mathcal{V}$
1451: satisfied $Q_B \mathcal{V} = 0$ then we would find that $\widehat{W} =
1452: 0$ since by (\ref{omegadef}) $\Omega$ is pure-gauge in the bosonic sense.
1453: We instead assume that
1454: \begin{equation}
1455: Q_B (\eta_0+\tilde{\eta}_0) \mathcal{V}
1456: = (\eta_0+\tilde{\eta}_0) Q_B \mathcal{V}= 0 \ , \qquad Q_B \mathcal{V} \ne 0 \ .
1457: \end{equation}
1458: We can now check that (\ref{susyinvariant}) is invariant under
1459: (\ref{omegatransformation}). To see this, note that under the gauge
1460: transformation (\ref{supergauge}),
1461: \begin{equation} \label{Wshift}
1462: \widehat{W}(\Omega,\mathcal{V}) \to \widehat{W}(\Omega,\mathcal{V}) +
1463: \widehat{W}(e^{-\eta_0 \Lambda'} Q_B e^{\eta_0 \Lambda'},\mathcal{V})
1464: \end{equation}
1465: To show that the second term vanishes, define
1466: \begin{equation}
1467: \Sigma_{\tau} = e^{-\tau \eta_0 \Lambda} Q_B e^{\tau \eta_0 \Lambda} \ ,
1468: \end{equation}
1469: and consider
1470: \begin{multline}
1471: \partial_\tau \langle \mathcal{I} |\mathcal{V}(i) |\Sigma_\tau\rangle
1472: = \langle \mathcal{I} |\mathcal{V}(i) | \left(Q_B \eta_0 \Lambda
1473: + [\Sigma_\tau,\eta_0 \Lambda]\right)\rangle
1474: \\
1475: = \langle \mathcal{I} |\mathcal{V}(i) |Q_B \eta_0 \Lambda \rangle
1476: = \langle \mathcal{I} |Q_B (\eta_0 + \tilde{\eta}_0)\mathcal{V}(i)
1477: | \Lambda \rangle = 0 \ .
1478: \end{multline}
1479: Since $\Sigma_0 = 0$, it follows that
1480: \begin{equation}
1481: \langle \mathcal{I} |\mathcal{V}(i) |\Sigma_\tau\rangle = 0 \ .
1482: \end{equation}
1483: Since $\Sigma_1$ is the shift term in the gauge transformation
1484: (\ref{Wshift}), $\widehat{W}(\Phi,\mathcal{V})$ is gauge
1485: invariant under (\ref{supergauge}).
1486:
1487: \subsection{Computation of $\widehat{W}$ for marginal solutions with trivial OPE}
1488:
1489: For marginal solutions with trivial OPE there are two known solutions for the Berkovits superstring field theory.
1490: The first, found by Erler and Okawa \cite{Erler:2007rh,Okawa:2007ri},
1491: is similar to the Schnabl gauge solution in the bosonic theory and
1492: does not appear to be simple to work with in this context. The
1493: second, found by Fuchs and Kroyter \cite{Fuchs:2007gw} and Kiermaier and Okawa \cite{Kiermaier:2007ki}, which is
1494: analogous to their bosonic solutions, is, once again, more practical for our considerations.
1495:
1496: Following the notation of Kiermaier and Okawa \cite{Kiermaier:2007ki}, let $\widehat{V}_{1/2}$ be a
1497: superconformal primary with weight $1/2$ and define $\widehat{V}_1 =
1498: G_{-1/2} \widehat{V}_{1/2}$. Putting
1499: \begin{equation} \label{susyO}
1500: \mathcal{O}_L = c \widehat{V}_1 + \eta e^\phi \widehat{V}_{1/2} \ ,
1501: \end{equation}
1502: an exact solution for $\Psi_L = e^{-\Phi} Q_B e^{\Phi}$ can be written as
1503: \begin{equation}
1504: \Psi_{L} = -\sum_{n = 1}^\infty (-\lambda)^n \Psi_{L}^{(n)} \ ,
1505: \end{equation}
1506: where
1507: \begin{equation}
1508: \langle \phi |\Psi_L^{(n)}\rangle = \left\langle f\circ \mathcal{\phi}(0)
1509: \mathcal{O}_L(1) \prod_{m = 2}^n \int_{t_{m-1}}^m dt_m \,\widehat{V}_1(t_m) \right\rangle_{C_{n+1}} \ ,
1510: \end{equation}
1511: and $t_1 \equiv 1$. One can now compute the invariant
1512: $\widehat{W}(\Psi_L,\mathcal{V})$ in a similar fashion to the bosonic case.
1513: For an NS-NS closed string field, we can represent $\mathcal{V}$ by
1514: \begin{equation}
1515: \mathcal{V} = (\xi + \tilde \xi) c \tilde c e^{-\phi-\tilde \phi} \mathcal{O}^{(\frac{1}{2},\frac{1}{2})} \ ,
1516: \end{equation}
1517: where $\mathcal{O}^{(\frac{1}{2},\frac{1}{2})}$ is a weight
1518: $(\frac{1}{2},\frac{1}{2})$ matter primary. On the disk
1519: \begin{equation}
1520: \widehat{W}(\Psi_L,\mathcal{V})
1521: = i\sum_{n = 1}^\infty (-\lambda)^n\left\langle
1522: \mathcal{V}(0) \mathcal{O}_L (1) \prod_{m = 2}^n
1523: \int_{\theta_{m-1}}^{2\pi \frac{m-1}{n}} d\theta \,\widehat{V}_1 (e^{i\theta_m})
1524: \right \rangle_{\text{disk}} \ .
1525: \end{equation}
1526: Examining the $\xi \eta$
1527: ghost system reveals that we can replace $\mathcal{O}_L$ with just its
1528: first term $c \widehat{V}_1$ since the second term will make no
1529: contribution. The $\eta\xi$ part of the amplitude becomes simply
1530: $\langle \xi(z) + \tilde \xi(\bar z)\rangle =2$, saturating the $\xi$ zeromode.
1531: We thus find,
1532: \begin{equation}
1533: \widehat{W}(\Psi_L,\mathcal{V})
1534: = i\sum_{n = 1}^\infty (-\lambda)^n\left\langle
1535: \mathcal{V}(0) c\widehat{V}_1 (1) \prod_{m = 2}^n \int_{\theta_{m-1}}^{2\pi \frac{m-1}{n}} d\theta \,\widehat{V}_1 (e^{i\theta_m})
1536: \right \rangle_{\text{disk}} \ .
1537: \end{equation}
1538: This integral can be rewritten as
1539: \begin{equation}
1540: \widehat{W}(\Psi_L,\mathcal{V})
1541: = -\frac{1}{2\pi i}\sum_{n = 1}^\infty \left\langle
1542: \mathcal{V}(0)\, c(1) \left\{ \exp\left( -\int_0^{2\pi} d\theta\, \widehat{V}_1(e^{i\theta}) \right) - 1\right\}
1543: \right \rangle_{\text{disk}} \ .
1544: \end{equation}
1545: Hence, at least for this particular $\Phi$, we find a similar result to the bosonic case,
1546: \begin{equation}
1547: \widehat{W}(\Phi,\mathcal{V}) = \mathcal{A}_{\Phi}^{\text{disk}}(\mathcal{V})
1548: - \mathcal{A}_{0}^\text{disk}(\mathcal{V}) \ .
1549: \end{equation}
1550:
1551: Inserting R-R-vertex operators on the disk is somewhat more subtle as
1552: one has to pick the vertex operators in an asymmetric picture
1553: \cite{Bianchi:1991eu,Polchinski:1987tu,DiVecchia:1997pr,Billo:1998vr,DiVecchia:1999rh}.
1554: Moreover, to preserve the arguments made above, it is necessary to
1555: pick a representation of the vertex operator which has total
1556: $\phi$-momentum $-2$ and doesn't have any additional insertions of the
1557: $\xi$-ghost zero-mode besides the factor of $(\xi +\tilde{\xi})$ that
1558: will be inserted by hand. The advantage of such a representation is
1559: that it allows us to drop the second term in $\mathcal{O}_L$ as we did
1560: in the NS-NS case. Such representations exist, but contain an infinite
1561: number of terms \cite{Billo:1998vr}:
1562: \begin{equation}
1563: \mathcal{V} = (\xi+\tilde{\xi}) \sum_{M = 0}^\infty \mathcal{V}^{(M)} (k,z,\tilde{z}) \ ,
1564: \end{equation}
1565: where
1566: \begin{equation}
1567: \mathcal{V}^{(M)}(z,\bar z) = a_M \Omega_{AB}
1568: \mathbb{V}^A_{-1/2 + M}
1569: \tilde{\mathbb{V}}^B_{-3/2 -M} (\bar z) \ ,
1570: \end{equation}
1571: and the $a_M$ are constants, $\Omega_{AB}$ is a spinor
1572: representation of the R-R-field of interest and
1573: \begin{align}
1574: \mathbb{V}^A_{-1/2+M}(z) &= \partial^{M-1} \eta(z) \ldots \eta(z)
1575: c(z) S^A(z) e^{(-\frac{1}{2} +M) \phi(z)} e^{i k X(z)/2} \ ,
1576: \\
1577: \mathbb{V}^A_{-1/2+M}(z) &= \bar{\partial}^{M} \tilde{\xi}(\bar z) \ldots \bar \partial \tilde{\xi}(\bar z)
1578: \tilde{c}(\bar{z}) \tilde{S}^A(\bar{z}) e^{(-\frac{3}{2} -M) \tilde{\phi}(\bar{z})} e^{i k \tilde{X}(\bar{z})/2} \ .
1579: \end{align}
1580: Noting that each term has one more $\xi$ than $\eta$ and a factor of
1581: $e^{(-\frac{1}{2} +M)\phi + (-\frac{3}{2}-M)\tilde\phi}$, which
1582: saturates the $\phi$-momentum of the disk, we can, as in the NS-NS
1583: case, drop the second term in $\mathcal{O}_L$ given in (\ref{susyO})
1584: from the computation and the same results follow. Note that we are
1585: free to pick other representations of the NS-NS vertex. This choice
1586: is convenient only in that it simplifies the relationship between
1587: $\widehat{W}(\Phi,\mathcal{V})$ and the closed string one-point
1588: function. See also \cite{Michishita:2004rx} for a computation of the R-R invariants without
1589: using this more complicated vertex operator.
1590:
1591: Given that one can compute the R-R one-point function, the reader will
1592: immediately wonder if it is possible to compute the R-R charges of a
1593: given background. Here we offer a few general remarks. We leave a detailed analysis to future work.
1594: In general, computing the R-R charges using $\widehat W(\Phi,\mathcal{V})$ is difficult because of the on-shell
1595: constraint on the R-R vertex operator. The on-shell constraint
1596: typically allows one only to compute the coupling of the zero-mode of
1597: the R-R field to the brane, which gives something proportional to the
1598: integral of the R-R charge over the brane world volume (including the
1599: infinite volume factor for the brane world-volume). For the special
1600: case of the D-instanton, there are no volume factors and the zero-mode
1601: of the R-R tadpole is proportional to the number of D-instantons.
1602:
1603: Even in the D-instanton case, however, this is not a manifestly topological quantity. It
1604: is only for classical solutions $\Phi$ that we can interpret
1605: $\widehat{W}(\Phi,\mathcal{V})$ as being a closed string one-point function.
1606: For example, since $\widehat W(\Phi,\mathcal{V})$ is linear in $\Phi$, if we allow $\Phi$ to
1607: be an arbitrary state, there is no way that $\widehat W(\Phi,\mathcal{V})$
1608: could always be an integer. It appears, then, that
1609: $\widehat{W}(\Phi,\mathcal{V})$ cannot be used to classify different $\Phi$'s as
1610: having different charges off-shell.
1611:
1612:
1613: \section*{Acknowledgments}
1614: We would like to that A. Awad, S. Das, W. Merrel, Y. Okawa, and
1615: B. Zwiebach for useful discussions and A. Hashimoto and M. Schnabl for comments on the draft. We would also like to thank the participants of the {\em String field theory and related aspects} workshop for many useful comments. This work was supported by
1616: Department of Energy Grant No. DE-FG01-00ER45832.
1617:
1618: \appendix
1619:
1620: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1621: \section{Computation of the surface term}\label{SurfaceTermComp}
1622: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1623: In this appendix, we explain the steps between (\ref{TwoPointFull}) and (\ref{TwoPointFullSurface}).
1624:
1625: Define the adjoint action of $\Psi$ by
1626: \begin{equation}
1627: \text{ad}_\Psi A = \Psi* A - (-1)^{\text{gh}(A)} A*\Psi \ .
1628: \end{equation}
1629: Note that because of the grading,
1630: \begin{equation}
1631: \left(\ad_\Psi\right)^2 A = \ad_{\Psi^2} A \ .
1632: \end{equation}
1633: We also have
1634: \begin{equation}
1635: \{Q_B,\ad_\Psi\} = \ad_{Q_B \Psi} = -\ad_{\Psi^2} \ ,
1636: \end{equation}
1637: where in the last step we use that $\Psi$ satisfies the classical equations of motion.
1638: Now, consider the two-point function with $\mathcal{V}_1 = \{Q_B,\mathcal{O}(i)\}$,
1639: \begin{equation}
1640: \mathcal{A} = \sum_{n = 0}^\infty \left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1641: \langle \mathcal{I}| \{Q_B,\mathcal{O}(i)\}
1642: b_0 D_{T_1} \left(\prod_{i = 2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle \ .
1643: \end{equation}
1644: Impose a short distance cutoff on the length of the propagator,
1645: \begin{equation}
1646: \mathcal{A}_\epsilon = \sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1647: \delta(\mbox{$\sum_i$} T_i - T)
1648: \langle \mathcal{I}| \{Q_B,\mathcal{O}(i)\}
1649: b_0 D_{T_1} \left( \prod_{i = 2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle \ .
1650: \end{equation}
1651: Now, push the $Q_B$ to the right:
1652: \begin{multline}
1653: \mathcal{A}_\epsilon = \sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1654: \delta(\mbox{$\sum_i$} T_i - T)
1655: \\
1656: \biggl\{
1657: - \langle \mathcal{I}| \mathcal{O}(i)
1658: \partial_{T_1} D_{T_1} \left(\prod_{i = 2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1659: \\
1660: -
1661: \sum_{m = 1}^n \langle \mathcal{I}| \mathcal{O}(i)
1662: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_{i}}\right) \left( \ad_{\Psi^2} b_0 D_{T_{m+1}}\right) \left( \prod_{i = m+2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1663: \\
1664: -
1665: \sum_{m = 1}^n\langle \mathcal{I}| \mathcal{O}(i)
1666: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_{i}} \right)\left( \ad_{\Psi} \partial_{T_i} D_{T_{m+1}}\right) \left( \prod_{i = m+2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1667: \biggr\} \ .
1668: \end{multline}
1669: Note that some of the terms have derivatives on the moduli. Integrating by parts, these derivatives can be made to act on the delta-function and interpreted as derivatives with respect to $T$. We write
1670: \begin{equation}
1671: \mathcal{A}_\epsilon = \mathcal{A}_1+\mathcal{A}_2 \ ,
1672: \end{equation}
1673: with $\mathcal{A}_1$ given by the terms where the derivatives hit the delta-function,
1674: \begin{multline}
1675: \mathcal{A}_1
1676: =
1677: -\sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \frac{\partial}{\partial T}\left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1678: \delta(\mbox{$\sum_i$} T_i - T)
1679: \\
1680: \biggl\{
1681: \langle \mathcal{I}| \mathcal{O}(i) D_{T_1} \left( \ad_{\Psi} b_0 D_{T_i}\right)^n \mathcal{V}_2 (i) |\mathcal{I}\rangle
1682: \\
1683: +
1684: \sum_{m = 1}^n\langle \mathcal{I}| \mathcal{O}(i)
1685: b_0 D_{T_1} \left( \prod_{i = 2}^m \ad_{\Psi} b_0 D_{T_i}\right)\left( \ad_{\Psi} D_{T_{m+1}}\right) \left( \prod_{i = m+2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1686: \biggr\} \ .
1687: \\
1688: =
1689: -\sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \frac{\partial}{\partial T}\left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1690: \delta(\mbox{$\sum_i$} T_i - T)
1691: \biggl\{
1692: \langle \mathcal{I}| \mathcal{O}(i) D_{T_1} \left(\prod_{i = 2}^{n+1} \{b_0,\ad_{\Psi}\} D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle \biggr\} \ ,
1693: \end{multline}
1694: and $\mathcal{A}_2$ the rest,
1695: \begin{multline}
1696: \mathcal{A}_2 = \sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1697: \delta(\mbox{$\sum_i$} T_i - T)
1698: \\
1699: \biggl\{
1700: - \langle \mathcal{I}| \mathcal{O}(i) \left(\prod_{i = 2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle \qquad \qquad \qquad \qquad \qquad \qquad \nonumber
1701: \end{multline}
1702: \begin{multline}
1703: -
1704: \sum_{m = 1}^n \langle \mathcal{I}| \mathcal{O}(i)
1705: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_i}\right) \left( \ad_{\Psi^2} b_0 D_{T_{m+1}}\right) \left( \prod_{i = m+2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1706: \\
1707: +
1708: \delta(T_{n+1}) \sum_{m = 1}^n\langle \mathcal{I}| \mathcal{O}(i)
1709: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_{i}} \right)\left( \ad_{\Psi}\right) \left( \prod_{i = m+1}^{n} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1710: \biggr\} \ .
1711: \end{multline}
1712: To simplify this, note that the first term in the $\{ \ \}$'s vanishes since
1713: \begin{equation}
1714: \langle \mathcal{I} |\mathcal{O}(i) \ad_{\Psi} = \ad_{\Psi} \mathcal{V}_2(i) |\mathcal{I}\rangle = 0 \ .
1715: \end{equation}
1716: This also kills the third term when $m = n$. We are left with
1717: \begin{multline}
1718: \mathcal{A}_2 = \sum_{n = 0}^\infty \int_{\epsilon/2}^{\infty} dT \left(\prod_{i = 1}^{n+1} \int_0^{\infty} dT_i \right)
1719: \delta(\mbox{$\sum_i$} T_i - T) \sum_{m = 1}^n
1720: \\
1721: \biggl\{
1722: -
1723: \langle \mathcal{I}| \mathcal{O}(i)
1724: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_i}\right) \left( \ad_{\Psi^2} b_0 D_{T_{m+1}}\right) \left( \prod_{i = m+2}^{n+1} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1725: \\
1726: +
1727: \delta(T_{n+1})\langle \mathcal{I}| \mathcal{O}(i)
1728: b_0 D_{T_1} \left(\prod_{i = 2}^{m} \ad_{\Psi} b_0 D_{T_i} \right)\left( \ad_{\Psi^2} b_0 D_{T_{m+1}}\right)
1729: \left( \prod_{i = m+2}^{n} \ad_{\Psi} b_0 D_{T_i}\right) \mathcal{V}_2 (i) |\mathcal{I}\rangle
1730: \biggr\} \ .
1731: \end{multline}
1732: This vanishes since the second term in the $\{ \ \}$'s is zero for $n<2$, while the first term is zero for $n<1$. If follows that
1733: \begin{equation}
1734: \mathcal{A}_{\epsilon} = \mathcal{A}_1 \ ,
1735: \end{equation}
1736: from which (\ref{TwoPointFullSurface}) follows.
1737:
1738:
1739:
1740:
1741: \section{Changing the height of a state by reparametrization}\label{reparamdiscussion}
1742:
1743: In this appendix, we briefly discuss why the height of an insertion
1744: $\Psi_{\text{cl}}$ may be changed by a reparametrization and, hence, a
1745: gauge transformation. In figure \ref{Reparam}a a state is shown in
1746: strip coordinates. To decrease the height of the insertion, we
1747: replace the region of the state near the midpoint with the identity
1748: state so that it has no effect when inserted into the propagator. The
1749: rest of the state is shrunk to a width $h$. This is shown in figure
1750: \ref{Reparam}b
1751:
1752: \begin{figure}
1753: \centerline{
1754: \scalebox{.5}{\begin{picture}(685,141)(-30,-10)
1755: \includegraphics{Reparam}
1756: \end{picture}}
1757: \begin{picture}(0,0)(342.5,0)
1758: \put(160,39){$\pi$}
1759: \put(216,39){$a$}
1760: \put(215,63){$b$}
1761: \put(215,14){$c$}
1762: \put(319,67){$h$}
1763: \put(-15,67){\bf a)}
1764: \put(185,67){\bf b)}
1765: \end{picture}
1766: }
1767: \caption{In a a typical state is shown with width $\pi/2$. In b, a
1768: modified state is shown which reduces to the identity state near the
1769: midpoint. The lines $ab$ and $ac$ are to be identified as well as the
1770: lines extending to the right of $b$ and $c$ as shown with the hatches.
1771: In the actual geometry of interest the thin vertical strip of
1772: worldsheet to the left of $cab$ would be of zero thickness.
1773: \label{Reparam}}
1774: \end{figure}
1775:
1776: The important point to recognize is that the height $h$ can be
1777: adjusted by simply rescaling the identity and strip segments of the
1778: state in a way that keeps the whole length of the state fixed. For
1779: example, if $0<\theta<\pi$ is a coordinate on the unit circle, we can
1780: perform the reparametrization
1781: \begin{equation}
1782: \tilde{\theta}(\theta) = \left\{ \begin{matrix}
1783: \rho \theta & \theta<h
1784: \\
1785: \vphantom{a} &
1786: \\
1787: \frac{\pi}{2} - \frac{\pi - 2\rho h}{\pi-2h}(\pi/2- \theta) & h<\theta<\pi/2
1788: \end{matrix} \right. \ ,
1789: \end{equation}
1790: where we also define $\tilde\theta(\pi - \theta) = \pi - \tilde
1791: \theta(\theta)$. This map scales $h \to \rho h$. Note that because the
1792: identity state is invariant under symmetric reparametrizations which
1793: preserve the midpoint and endpoints, there is considerable flexibility
1794: in the choice of $\tilde \theta(\theta)$ in the region $h<
1795: \theta<\pi - h$.
1796:
1797: Note also that picking $\rho = \pi/2h$ leads to a singular
1798: reparametrization; the entire region $h<\theta<\pi - h$ is mapped to
1799: the midpoint. However, as long as the state is inserted into a larger
1800: worldsheet geometry, this transformation remains smooth. One may also
1801: worry that $\tilde \theta(\theta)$ could create problems if there are
1802: operators near the midpoint (points $b$ and $c$ in figure
1803: \ref{Reparam}b). Though we have no basis for doing so (as we do not
1804: have a regularity condition on our string field), we assume that
1805: operators insertions near the midpoint are sufficiently mild that this
1806: will not be a problem.
1807:
1808:
1809:
1810:
1811:
1812:
1813:
1814:
1815:
1816:
1817:
1818:
1819: \bibliography{c}\bibliographystyle{utphys}
1820:
1821: \end{document}
1822: