1: \documentclass[11pt]{article}
2: \usepackage{amssymb,amsmath,epsf,graphicx}
3: \input epsf.sty
4: \topmargin -.5cm
5: \textheight 21cm
6: \oddsidemargin -.125cm
7: \textwidth 16cm
8:
9:
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\bea}{\begin{eqnarray}\displaystyle}
13: \newcommand{\eea}{\end{eqnarray}}
14: \newcommand{\bdm}{\begin{displaymath}}
15: \newcommand{\edm}{\end{displaymath}}
16: \newcommand{\sectiono}[1]{\section{#1}\setcounter{equation}{0}}
17: \renewcommand{\theequation}{\thesection.\arabic{equation}}
18:
19: \newcommand{\Tr}{\mathop{\rm Tr}\nolimits}
20: \newcommand{\re}{\mathop{\rm Re}\nolimits}
21: \newcommand{\im}{\mathop{\rm Im}\nolimits}
22: %\newcommand{\bpz}{\mathop{\rm bpz}\nolimits}
23: \newcommand{\ad}{\mathop{\rm ad}\nolimits}
24: \newcommand{\logit}{\mathop{\rm logit}\nolimits}
25: \newcommand{\arccot}{\mathop{\rm arccot}\nolimits}
26: \newcommand{\asympt}{\mathop{\sim}}
27: \newcommand{\leftpartial}{\overleftarrow{\partial}}
28: \newcommand{\rightpartial}{\overrightarrow{\partial}}
29: %\newcommand{\leftpartial}{\mathop{\!\stackrel{\leftarrow}{\partial}}\nolimits}
30: %\newcommand{\rightpartial}{\mathop{\!\stackrel{\rightarrow}{\partial}}\nolimits}
31: \newcommand{\leftD}{\mathop{\stackrel{\leftarrow}{D}}\nolimits}
32: \newcommand{\rightD}{\mathop{\stackrel{\rightarrow}{D}}\nolimits}
33:
34: %This defines the star product symbol
35: \newcommand{\starp}{*}
36:
37: %This defines the BPZ conjugation symbol
38: \newcommand{\bpz}{\star}
39:
40: \def\bra#1{\langle #1 |}
41: \def\ket#1{|#1 \rangle}
42: \def\kket#1{||#1 \rangle\!\rangle}
43: \def\aver#1{\langle\, #1 \,\rangle}
44: \def\slash#1{\not\!#1}
45: \def\rdslash{\not\!\partial}
46: \def\ldslash{\not\!\leftpartial}
47: \def\ov{\overline}
48: \def\l{\left}
49: \def\r{\right}
50: \def\res#1{\oint\! \frac{d#1}{2\pi i} \,}
51:
52: \let\eps = \varepsilon
53: \def \ot{\otimes}
54:
55: \def \nn {{\mathbb N}}
56: \def \zz {{\mathbb Z}}
57: \def \cc {{\mathbb C}}
58: \def \rr {{\mathbb R}}
59: \def \TT {{\mathbb T}}
60:
61: \def \ii {{\cal I}}
62: \def \ll {{\cal L}}
63: \def \dd {{\cal D}}
64: \def \kk {{\cal K}}
65: \def \hh {{\cal H}}
66: \def \aa {{\cal A}}
67: \def \bb {{\cal B}}
68: \def \mm {{\cal M}}
69: \def \oo {{\cal O}}
70: \def \pp {{\cal P}}
71: \def \QQ {{\cal Q}}
72: \def \nnn {{\cal N}}
73: \def \eee {{\cal E}}
74: \def \ccc {{\cal C}}
75: \def \ss {{\cal S}}
76: \def \uu {{\cal U}}
77: \def \nc {noncommutative }
78: \def \ncg {noncommutative geometry }
79: \def \sf {string field }
80: \def \sft {string field theory }
81: \def \da {\dagger}
82:
83:
84: \def \lll {{\widehat{\cal L}}}
85: \def \bbb {{\widehat{\cal B}}}
86:
87:
88: \begin{document}
89: {}~ \hfill\vbox{\hbox{hep-th/0606142}\hbox{MAD-TH-06-6}\hbox{CERN-PH-TH/2006-114} }\break
90: \vskip 2.1cm
91:
92: \centerline{\Large \bf Proof of vanishing cohomology at the tachyon vacuum } \vspace*{2.0ex}
93: %\centerline{\Large \bf in open string field theory}
94: \vspace*{8.0ex}
95:
96: \centerline{\large \rm Ian Ellwood$^a$ and Martin Schnabl$^b$}
97:
98: \vspace*{8.0ex}
99:
100: \centerline{\large \it $^a$Department of Physics, }
101: \centerline{\large \it University of Wisconsin, Madison, WI 53706, USA} \vspace*{2.0ex}
102: \centerline{E-mail: {\tt iellwood@physics.wisc.edu}}
103:
104: \vspace*{6.0ex}
105:
106: \centerline{\large \it $^b$Department of Physics, Theory Division,}
107: \centerline{\large \it CERN, CH-1211, Geneva 23, Switzerland} \vspace*{2.0ex}
108: \centerline{E-mail: {\tt martin.schnabl@cern.ch}}
109:
110: \vspace*{6.0ex}
111:
112: \centerline{\bf Abstract}
113: \bigskip
114:
115: We prove Sen's third conjecture that there are no on-shell
116: perturbative excitations of the tachyon vacuum in open bosonic string
117: field theory. The proof relies on the existence of a special state
118: $A$, which, when acted on by the BRST operator at the tachyon vacuum,
119: gives the identity. While this state was found numerically in
120: Feynman-Siegel gauge, here we give a simple analytic expression.
121:
122: \vfill \eject
123:
124: \baselineskip=16pt
125:
126: %\tableofcontents
127: \newpage
128:
129: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
130: %%%%%%%%%%%%%%%%
131: \sectiono{Introduction}
132: \label{s_intro}
133:
134: Following Sen's famous three conjectures \cite{Sen:1999mh,Sen:1999xm},
135: there has been an intensive effort to study the physics of tachyon
136: condensation in Witten's cubic open string field theory
137: \cite{Witten:1985cc}. The power of open string field theory (OSFT)
138: over conventional CFT methods is that OSFT is an off-shell formulation
139: of open string interactions. Many questions about open string vacua,
140: which must be understood using indirect arguments in CFT, can be
141: rephrased in OSFT as questions about the classical solutions of the
142: OSFT equations of motion.
143:
144: Unfortunately, finding solutions to the OSFT equations of motion is
145: non-trivial. Indeed, in the standard oscillator basis, these equations
146: become an infinite number of coupled non-linear differential equations
147: and, until recently, much of the work in OSFT has been numerical.
148:
149: In spite of the approximate nature of the analysis, it has been found that
150: OSFT has a rich structure. Starting from perturbative vacuum on the
151: D25-brane, one can find classical solutions to the equations of motion
152: representing lower-dimensional branes
153: \cite{Moeller:2000jy,deMelloKoch:2000xf,Moeller:2000hy,Beccaria:2005js} as well as the
154: tachyon vacuum
155: \cite{Kostelecky:1989nt,Sen:1999nx,Moeller:2000xv,Gaiotto:2002wy}, in
156: which there are no branes present. In each case, the energy of these
157: solutions precisely matches the energy of the relevant brane
158: configuration, beautifully demonstrating Sen's first and second
159: conjectures.
160:
161: Having found solutions representing various vacua, one can attempt to
162: find the spectrum of perturbative states around each solution. In
163: particular, Sen's third conjecture states that around the tachyon
164: vacuum, which represents the absence of any brane at all, there should
165: be no physical states. This conjecture has been checked in two
166: complementary ways. First, the kinetic terms and gauge
167: transformations of certain low-mass excitations were computed to verify
168: that, indeed, there were no on-shell states
169: \cite{Ellwood:2001py,Giusto:2003wc}. Second, it was argued that the
170: full spectrum of states was empty using a trick, which we now describe
171: \cite{Ellwood:2001ig}.
172:
173: The physical states around a given vacuum are given by the cohomology
174: of a BRST operator $Q_\Psi$. It turns out that the cohomology of
175: $Q_\Psi$ vanishes -- meaning that there are no physical states -- if
176: and only if there exists a state $A$ such that $Q_\Psi A =
177: \mathcal{I}$, where $\mathcal{I}$ is the identity of the star algebra.
178: Hence, the problem of showing that $Q_\Psi$ has vanishing cohomology
179: reduces to determining whether there is a solution to a single linear
180: equation. This makes the problem amenable to numerical analysis and
181: it was found in \cite{Ellwood:2001ig} that, within the
182: level-truncation approximation, one could find such a state $A$.
183:
184: Recently, one of us found an analytic solution to the OSFT equations
185: of motion representing the tachyon vacuum \cite{Schnabl:2005gv}. This
186: solution has now been checked to satisfy the equations of motion, even
187: when contracted with itself \cite{Okawa:2006vm,Fuchs:2006hw}, and has
188: the correct energy \cite{Schnabl:2005gv}, giving an analytic proof of
189: Sen's first conjecture. This solution opens up the possibility that
190: other questions in OSFT, which previously had only been understood
191: numerically, may have nice analytic solutions.
192:
193: Indeed, in this paper we give a simple proof of Sen's third
194: conjecture. We do this following the method described above: Given
195: the analytic solution $\Psi$, we find an analytic expression for a
196: state $A$ that satisfies $Q_\Psi A = \mathcal{I}$.
197:
198: The organization of this paper is as follows: In section
199: \ref{s:ReviewOfOSFT}, we review the relevant aspects of OSFT. In
200: section \ref{s:TachyonVacuum} we present the recently found analytic
201: solution to the equations of motion, $\Psi$. Next, in section
202: \ref{s:VanishingCohomologyProof}, we define a new string field $A$,
203: which we then prove satisfies $Q_\Psi A = \mathcal{I}$. Finally, in
204: section \ref{s:pg}, we discuss the fact that the tachyon vacuum is a limit
205: of a family of pure-gauge solutions and show how this does not spoil
206: our cohomology arguments.
207:
208: % We conclude with a brief discussion in section \ref{s:Discussion}.
209:
210:
211:
212:
213:
214:
215:
216:
217:
218:
219:
220:
221: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
222: %%%%%%%%%%%%%%%%
223:
224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
225: %%%%%%%%%%%%%%%%
226: \sectiono{Review of OSFT}
227: \label{s:ReviewOfOSFT}
228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
229: %%%%%%%%%%%%%%%%
230:
231: We begin with a review of some basic aspects of Witten's cubic open
232: string field theory. Since there are many excellent reviews of OSFT
233: \cite{Thorn:1988hm,Sen:1999nx,Zwiebach:2001nj,Taylor:2003gn}, we will
234: only touch on some of the more relevant points. The action is given
235: by \cite{Witten:1985cc}
236: \begin{equation} \label{WittenAction}
237: S = \frac{1}{2} \int \Phi \starp Q_B \Phi + \frac{1}{3} \int \Phi
238: \starp \Phi \starp \Phi.
239: \end{equation}
240: The classical field, $\Phi$, is an element of the free string Fock
241: space. For example, for OSFT on a D25-brane background, it has an
242: expansion,
243: \begin{equation}
244: \Phi = \int dp \left\{ t(p) + A_\mu (p) \alpha_{-1}^\mu + \psi(p)
245: c_0 + \ldots \right\} c_1 |p\rangle,
246: \end{equation}
247: where $t(p)$ is the tachyon, $A_\mu(p)$ is the gauge field and $\psi(p)$ is a
248: ghost field.
249:
250: The action (\ref{WittenAction}) has a large gauge invariance, which
251: makes solving the equations of motion in the non-gauge-fixed theory
252: difficult\footnote{The commutator is taken using the star product and
253: is graded by ghost number. Explicitly,
254: \[
255: [\Phi_1 , \Phi_2] = \Phi_1
256: \starp \Phi_2 - (-1)^{\text{gh}(\Phi_1) \text{gh}(\Phi_2)} \Phi_2 \starp
257: \Phi_1
258: \]
259: };
260: \begin{equation}
261: \Phi \to \Phi + Q_B \Lambda + [\Phi, \Lambda].
262: \end{equation}
263: Globally, fixing a gauge is a subtle issue
264: \cite{Ellwood:2001ne}. Around the perturbative vacuum, however, a
265: suitable choice is Feynman-Siegel gauge;
266: \begin{equation}
267: b_0 \Phi = 0.
268: \end{equation}
269: Most of the numerical work in OSFT was performed in this gauge.
270: However, as we will discuss shortly, there is a different gauge which
271: is more suitable for analytic analysis.
272:
273: The equations of motion of (\ref{WittenAction}) are given by
274: \begin{equation}
275: Q_B \Psi + \Psi \starp \Psi = 0.
276: \end{equation}
277: Given a solution, $\Psi$, one can re-expand the action around the new vacuum;
278: \begin{equation}
279: S(\Psi + \Phi) = \frac{1}{2} \int \Phi \starp Q_\Psi \Phi
280: + \frac{1}{3} \int \Phi \starp \Phi \starp \Phi + \text{constant}.
281: \end{equation}
282: The new action takes the same form as the old action: the cubic term
283: is left completely invariant, while the kinetic term is only modified
284: by a change in the BRST operator, $Q_B \to Q_\Psi$, where
285: \begin{equation}
286: Q_\Psi \Lambda = Q_B \Lambda + [\Psi,\Lambda].
287: \end{equation}
288: It is straightforward to check that $Q_\Psi^2 = 0$ using the equations
289: of motion of $\Psi$. Just as the spectrum around the perturbative
290: vacuum was given by the cohomology of $Q_B$, the spectrum around the
291: new vacuum is given by the cohomology of $Q_\Psi$.
292:
293: \subsection{OSFT in the $\arctan(z)$ coordinate system}
294:
295: \begin{figure}
296: \begin{center}
297: \setlength{\unitlength}{1pt}
298: \begin{picture}(380,132)(0,0)%
299: \includegraphics{arctanmap}%
300: \end{picture}
301: \begin{picture}(0,0)(380,0)%
302: \put(10,75){\makebox(0,0)[lb]{$z$}}
303: \put(112,37){\makebox(0,0)[lb]{$L$}}
304: \put(35,37){\makebox(0,0)[lb]{$R$}}
305: \put(72,57){\makebox(0,0)[lb]{$M$}}
306: \put(232,113){\makebox(0,0)[lb]{$\tilde{z}$}}
307: \put(74,0){\makebox(0,0)[lb]{$\mathcal{V}$}}
308: \put(282,0){\makebox(0,0)[lb]{$f\circ\mathcal{V}$}}
309: \put(244,70){\makebox(0,0)[lb]{$R$}}
310: \put(335,70){\makebox(0,0)[lb]{$L$}}
311: \put(30,0){\makebox(0,0)[lb]{$-1$}}
312: \put(113,0){\makebox(0,0)[lb]{$1$}}
313: \put(210,0){\makebox(0,0)[lb]{$-\frac{\pi}{2}$}}
314: \put(246,0){\makebox(0,0)[lb]{$-\frac{\pi}{4}$}}
315: \put(327,0){\makebox(0,0)[lb]{$\frac{\pi}{4}$}}
316: \put(362,0){\makebox(0,0)[lb]{$\frac{\pi}{2}$}}
317: \put(0,108){\makebox(0,0)[lb]{{\bf a)}}}
318: \put(200,108){\makebox(0,0)[lb]{{\bf{b)}}}}
319: \put(289,131){\makebox(0,0)[lb]{$M$}}
320: \end{picture}
321: \end{center}
322: \caption{ \small The string field as seen by two coordinate systems.
323: In a) the standard description on upper half plane is illustrated. A
324: vertex operator $\mathcal{V}$ generates a state on the unit
325: circle. The right half, left half and midpoint of the string are
326: labeled as viewed from infinity. Diagram b) gives the same state in
327: the $\tilde{z} = \arctan(z)$ coordinate. The left and right sides of
328: the figure are identified to give a cylinder. The left/right half of
329: the string now lies along the line $\Re(\tilde{z}) = \pm \pi/4$. The
330: midpoint of the string is mapped to infinity.}
331: \label{f:arctanmap}
332: \end{figure}
333:
334: Most of the difficulty in working with OSFT arises from the complexity
335: of the star product. It was one of the key realizations of
336: \cite{Rastelli:2001vb,Schnabl:2005gv}, however, that the star product
337: simplifies when written in a different coordinate frame.
338:
339: The standard method for specifying states in open string theory is by
340: putting a vertex operator, $\mathcal{V}$, on the boundary of the upper
341: half plane at the point $z= 0$. By the operator-state correspondence
342: we can associate with $\mathcal{V}$ a state $|\mathcal{V}\rangle$ in
343: the string Fock space that lives on the unit circle.
344:
345: However, there was no reason why we had to choose the upper half plane
346: to define our states. It turns out to be useful to work instead in
347: the coordinate $\tilde{z} = f(z) = \arctan(z)$. Under $z \to f(z)$,
348: the upper half plane is mapped to an infinitely tall cylinder as
349: illustrated in figure~\ref{f:arctanmap}. In this frame, the star
350: product can be described purely geometrically; one simply glues the
351: strips of world-sheet together that correspond to the two string
352: states. This is illustrated in figure~\ref{f:starproduct}.
353:
354: \begin{figure}
355: \begin{center}
356: \setlength{\unitlength}{1pt}
357: \begin{picture}(235,132)(0,0)%
358: \includegraphics{starproduct}%
359: \end{picture}
360: \begin{picture}(0,0)(235,0)%
361: \put(14,112){\makebox(0,0)[lb]{$\tilde{z}$}}
362: \put(-9,-2){\makebox(0,0)[lb]{$-\frac{3\pi}{4}$}}
363: \put(28,-2){\makebox(0,0)[lb]{$-\frac{\pi}{2}$}}
364: \put(112,0){\makebox(0,0)[lb]{$0$}}
365: \put(183,-2){\makebox(0,0)[lb]{$\frac{\pi}{2}$}}
366: \put(216,-2){\makebox(0,0)[lb]{$\frac{3\pi}{4}$}}
367: \put(74,-2){\makebox(0,0)[lb]{$\widetilde{\mathcal{V}}_2$}}
368: \put(146,-2){\makebox(0,0)[lb]{$\widetilde{\mathcal{V}}_1$}}
369: \put(119,76){\makebox(0,0)[lb]{$R$}}
370: \put(46,76){\makebox(0,0)[lb]{$R$}}
371: \put(102,76){\makebox(0,0)[lb]{$L$}}
372: \put(173,76){\makebox(0,0)[lb]{$L$}}
373: \end{picture}
374: \end{center}
375: \caption{ \small A pictorial description of the star product. Given
376: two states $|\tilde{\mathcal{V}}_1\rangle $ and
377: $|\tilde{\mathcal{V}}_2\rangle$ generated by inserting vertex
378: operators $\tilde{\mathcal{V}}_1$ and $\tilde{\mathcal{V}}_2$ in the
379: $\tilde{z}$ coordinate, the star product,
380: $|\tilde{\mathcal{V}}_1\rangle \starp |\tilde{\mathcal{V}}_2\rangle$,
381: is computed by gluing the right side of the
382: $|\tilde{\mathcal{V}}_1\rangle $ state to the left side of the
383: $|\tilde{\mathcal{V}}_2\rangle$ state. This gives a cylinder of width
384: $3 \pi/2$. }
385: \label{f:starproduct}
386: \end{figure}
387: Multiplying $n$ strips of width $\pi/2$ will produce a strip of width
388: $n \pi/2$ and it is useful to consider the class of all such states.
389: When there are no operator insertions, a state described by a strip of
390: width $n \pi/ 2$ is called a wedge state and is denoted $|n+1\rangle$.
391: These states were first introduced in \cite{Rastelli:2000iu}, and obey
392: the algebra,
393: \begin{equation}
394: |n\rangle \starp |m\rangle = |m+n -1\rangle.
395: \end{equation}
396: The state $|2\rangle$ is just the original strip of width $\pi/2$ with
397: no vertex operator inserted at the origin and is, thus, the
398: $SL(2,\mathbb{R})$ invariant vacuum $|0\rangle$.
399:
400: It turns out that taking the limit as the width of the strip tends to
401: infinity leads to a finite state; $|\infty\rangle = \lim_{n\to \infty}
402: |n\rangle$. This state is known as the sliver \cite{Rastelli:2000iu} and as is a projector
403: under star multiplication;
404: \begin{equation}
405: |\infty \rangle \starp |\infty \rangle = |\infty \rangle.
406: \end{equation}
407: Notice that multiplying a state $\Lambda$ by the wedge state of zero width,
408: $|1\rangle$, leaves $\Lambda$ invariant. Hence, $\mathcal{I} =
409: |1\rangle$ is an identity of the star algebra;
410: \begin{equation}
411: \Lambda \starp \mathcal{I} = \mathcal{I} \starp \Lambda = \Lambda.
412: \end{equation}
413: A useful property of $\mathcal{I}$ is that, at least formally, for any operator $\mathcal{O}$ it
414: obeys \cite{Gross:1986ia,Gross:1986fk}
415: \begin{equation}\label{eq:OIequalsOIdagger}
416: \mathcal{O} |\mathcal{I}\rangle = \mathcal{O}^\bpz |\mathcal{I}\rangle
417: = \tfrac{1}{2} ( \mathcal{O}+ \mathcal{O}^\bpz) |\mathcal{I}\rangle ,
418: \end{equation}
419: where, in the notation of \cite{RZ2006}, $\mathcal{O}^\bpz$ denotes
420: BPZ conjugation; $\mathcal{O}^\bpz = I\circ \mathcal{O}$, where $I(z)
421: = -1/z$.
422:
423:
424: %%%%%%%%%%%%%%%%%%%
425: \subsection{Some important operators}
426: %%%%%%%%%%%%%%%%%%%
427:
428:
429:
430: In general, each of the familiar operators in the $\tilde{z}$
431: coordinate can be pulled back into the $z$ coordinate using
432: $f^{-1}(\tilde{z}) = \tan(\tilde{z})$. We will occasionally denote
433: such an operator using a tilde; e.g. $\tilde{c}(\tilde{z}) = f^{-1}
434: \circ c(z)$. It is also useful to make the following definitions:
435: \begin{equation}
436: \qquad
437: \mathcal{L}_0 = f^{-1} \circ L_0,
438: \qquad
439: \mathcal{B}_0 = f^{-1} \circ b_0,
440: \qquad
441: K_1 = f^{-1} \circ L_{-1},
442: \qquad
443: B_1 = f^{-1} \circ b_{-1}.
444: \end{equation}
445: Just as $L_0$ gave the mass level of fields in the $z$ coordinate,
446: $\mathcal{L}_0$ is the analogous level in the $\tilde{z}$ coordinates.
447: Similarly, while the standard gauge fixing condition in the
448: $z$-coordinate was $b_0 \Phi = 0$, in the $\tilde{z}$-coordinate, one
449: uses $\mathcal{B}_0 \Phi = 0$.
450:
451:
452:
453: Explicit mode expansions of $\mathcal{L}_0$ and $\mathcal{B}_0$ are given by
454: \begin{align}
455: \mathcal{L}_0 &= L_0 + \tfrac{2}{3} L_2 - \tfrac{2}{15} L_4 +\cdots,
456: \\
457: \mathcal{B}_0 &= b_0 + \tfrac{2}{3} b_2 - \tfrac{2}{15} b_4 +\cdots.
458: \end{align}
459: Note that while $L_0$ and $b_0$ are BPZ dual to themselves, their
460: script cousins are not and we also have operators $\mathcal{L}_0^\bpz$
461: and $\mathcal{B}_0^\bpz$, which are given by $L_{n}^\bpz = (-1)^n
462: L_{-n}$ and $b_n^\bpz = (-1)^n b_{-n}$. These obey the commutation
463: relations\footnote{These commutation relations are an important
464: property of the conformal frame of the sliver. Recently
465: \cite{RZ2006}, it has been shown that the conformal frames of other
466: projectors, known as special projectors, lead to similar algebras;
467: $[\mathcal{L}_0,\mathcal{L}_0^\bpz] =
468: s(\mathcal{L}_0+\mathcal{L}_0^\bpz)$. These special projectors have
469: many similarities with the sliver and can be used to solve the
470: ghostnumber zero equations of motion \cite{Gaiotto:2002uk,RZ2006}.},
471: \begin{equation}\label{eq:LLandBBcommutators}
472: [\mathcal{L}_0,\mathcal{L}_0^\bpz] = \mathcal{L}_0+\mathcal{L}_0^\bpz,
473: \end{equation}
474: as well as
475: \begin{equation} \label{eq:BLcommutators}
476: [\mathcal{L}_0, \mathcal{B}_0] = [\mathcal{L}_0^\bpz, \mathcal{B}_0^\bpz]= 0,
477: \qquad
478: [\mathcal{L}_0^\bpz, \mathcal{B}_0] = -\mathcal{B}_0-\mathcal{B}_0^\bpz,
479: \qquad
480: [\mathcal{L}_0, \mathcal{B}_0^\bpz] = \mathcal{B}_0+\mathcal{B}_0^\bpz.
481: \end{equation}
482: An important property of $\mathcal{L}_0$ is that the wedge states can be represented in the form \cite{LeClair:1988sp,LeClair:1988sj,Rastelli:2000iu,Schnabl:2002gg}
483: \begin{equation} \label{eq:wedgeInTermsOfU}
484: |r\rangle = U_r^\bpz |0\rangle,
485: \end{equation}
486: where $U_r = (2/r)^{\mathcal{L}_0}$. The operators $U_r$ and $U_s^\bpz$ obey the
487: important relation \cite{Schnabl:2002gg},
488: \begin{equation} \label{eq:UUdaggerrelation}
489: U_r U_s^\bpz = U_{2+ \frac{2}{r}(s-2)}^\bpz U_{2+\frac{2}{s} (r-2)},
490: \end{equation}
491: which can be used to derive (\ref{eq:LLandBBcommutators}).
492:
493: The operators $K_1$ and $B_1$ take a very simply form,
494: \begin{equation}
495: K_1 = L_1+L_{-1}, \qquad B_1 = b_1 + b_{-1}.
496: \end{equation}
497: These operators were first studied in \cite{Rastelli:2000iu}, where it
498: was shown that they are derivations of the star algebra;
499: \begin{align}\label{eq:K1B1asderivations}
500: K_1 (\Phi_1 \starp \Phi_2) &=
501: (K_1 \Phi_1) \starp \Phi_2 + \Phi_1 \starp (K_1 \Phi_2).
502: \\
503: B_1 (\Phi_1 \starp \Phi_2) &=
504: (B_1 \Phi_1) \starp \Phi_2 + (-1)^{\text{gh}(\Phi_1)} \Phi_1 \starp (B_1 \Phi_2).
505: \end{align}
506: They also annihilate the wedge states;
507: \begin{equation} \label{eq:LBLKonwedge}
508: K_1|r\rangle =
509: B_1|r\rangle =0.
510: \end{equation}
511: In the $\tilde{z}$ coordinates they take the form,
512: \begin{equation}
513: K_1 = \oint d\tilde{z} \,T(\tilde{z}),
514: \qquad
515: B_1 = \oint d\tilde{z} \,b(\tilde{z}).
516: \end{equation}
517: It is also useful to define the ``left'' and ``right'' parts of these
518: operators, which are given by taking only the left or right parts --
519: as viewed from infinity -- of the contour integral;
520: \begin{equation} \label{eq:KLRBLR}
521: K_1^{L,R} = \oint_{\gamma^{L,R}} d\tilde{z}\, \,T(\tilde{z}),
522: \qquad
523: B_1^{L,R} = \oint_{\gamma^{L,R}} d\tilde{z}\, \,b(\tilde{z}).
524: \end{equation}
525: In the $\tilde{z}$ coordinates, the contours, $\gamma^{L,R}$, are
526: given by the vertical lines on the right and left of the strip. Note
527: that, $K_1^L + K_1^R = K_1$ and $B_1^L + B_1^R = B_1$. Also,
528: \begin{align}\label{eq:KLRonstarproduct}
529: K_1^L (\Phi_1 \starp \Phi_2) &= (K_1^L\Phi_1 ) \starp \Phi_2,
530: &
531: B_1^L (\Phi_1 \starp \Phi_2) &= (B_1^L\Phi_1 ) \starp \Phi_2,
532: \\
533: K_1^R (\Phi_1 \starp \Phi_2) &= \Phi_1 \starp (K_1^R\Phi_2 ),
534: &
535: B_1^R (\Phi_1 \starp \Phi_2) &= (-1)^{\text{gh}(\Phi_1)}\Phi_1 \starp (B_1^R\Phi_2 ).
536: \end{align}
537:
538:
539: An important property of the operators $K_1^{L,R}$ is that they act as
540: a derivative with respect to the width of the state. This follows
541: from (\ref{eq:KLRBLR}). Since the $K^{L,R}$ are just integrals of $T$
542: in the $\tilde{z}$ coordinate and $\int T(\tilde{z})$ is the
543: world-sheet Hamiltonian, $\epsilon K^{R,L}$ can be thought of as
544: adding/subtracting an infinitesimal strip of with $\epsilon$ from the
545: right/left of the world-sheet. This gives the useful identity,
546: \begin{equation} \label{eq:KLRonwedge}
547: \partial_n |n\rangle = \pm \tfrac{\pi}{2} K_1^{R,L} |n\rangle,
548: \end{equation}
549: which can be integrated to give
550: \begin{equation} \label{eq:wedgefromKs}
551: |n\rangle = e^{\pm \frac{\pi}{2} (n-2) K_1^{R,L}} |0\rangle.
552: \end{equation}
553:
554: The operators $K_1^{L,R}$, $\mathcal{L}_0$ and $\mathcal{L}_0^\bpz$,
555: as well as $B_1^{L,R}$, $\mathcal{B}_0$ and $\mathcal{B}_0^\bpz$ are
556: related through the identities,
557: \begin{equation} \label{eq:KLRandLLdagger}
558: K_1^L - K_1^R = \tfrac{2}{\pi} (\mathcal{L}_0 + \mathcal{L}_0^\bpz),
559: \qquad
560: B_1^L - B_1^R = \tfrac{2}{\pi} (\mathcal{B}_0 + \mathcal{B}_0^\bpz),
561: \end{equation}
562: which follow from the definitions of these operators. Using (\ref{eq:KLRandLLdagger}), we can rewrite (\ref{eq:wedgefromKs})
563: as
564: \begin{equation} \label{eq:wedgeasexpofLs}
565: |n\rangle = e^{\frac{(2-n)}{2}
566: (\mathcal{L}_0+\mathcal{L}_0^\bpz)} |0\rangle.
567: \end{equation}
568: This expression can be related to (\ref{eq:wedgeInTermsOfU}) using the identity,
569: \begin{equation}\label{eq:UUdagger}
570: e^{\frac{(2-n)}{2}
571: (\mathcal{L}_0+\mathcal{L}_0^\bpz)} = U_n^\bpz U_n.
572: \end{equation}
573: A more general collection of such identities can be found in
574: \cite{Schnabl:2002gg,Schnabl:2002ff,Schnabl:2002ef,Schnabl:2005gv}.
575:
576:
577: \section{The exact tachyon vacuum solution}
578: \label{s:TachyonVacuum}
579:
580: In this section, we review the exact tachyon vacuum state found in
581: \cite{Schnabl:2005gv}. Define
582: \begin{equation}
583: \psi_n = \tfrac{2}{\pi} c_1 |0\rangle \starp B_1^L |n\rangle \starp
584: c_1| 0\rangle.
585: \end{equation}
586: Then the tachyon vacuum is given by\footnote{The term $-\partial_n
587: \psi_n$ for $n = 0$ can be defined by carefully taking the limit.
588: Explicitly, one finds $Q_B B_1^L c_1|0\rangle$.}
589: \begin{equation}
590: \Psi =
591: \lim_{N\to \infty}
592: \left( \psi_N - \sum_{n=0}^N \partial_n \psi_n \right).
593: \end{equation}
594: Formally, the $\psi_N$ piece vanishes in level truncation as $N \to
595: \infty$, but it gives finite contributions to the energy and is
596: required for $\Psi$ to satisfy the equations of motion when contracted
597: with itself \cite{Okawa:2006vm,Fuchs:2006hw};
598: \begin{equation}
599: \langle \Psi | Q_B \Psi\rangle +\langle \Psi |\Psi \starp \Psi \rangle = 0.
600: \end{equation}
601: We will see that this term is also required to give a complete proof
602: that the cohomology of $Q_\Psi$ vanishes.
603:
604: The solution satisfies the gauge fixing condition,
605: \begin{equation}
606: \mathcal{B}_0 \Psi = 0,
607: \end{equation}
608: which as alluded to earlier, is the analogue of Feynman-Siegel gauge
609: fixing in the $\tilde{z}$-coordinate.
610:
611: The states $-\partial_n \psi_n$ can be written using
612: (\ref{eq:KLRonwedge}) as
613: \begin{equation}\label{eq:dpsidn}
614: -\partial_n \psi_n =
615: c_1|0\rangle \starp B_1^L K_1^L |n\rangle \starp c_1 |0\rangle.
616: \end{equation}
617: These states take a simple form in the $\tilde{z}$ coordinate, as illustrated
618: in figure \ref{f:dpsi}.
619:
620: \begin{figure}
621: \begin{center}
622: \setlength{\unitlength}{1pt}
623: \begin{picture}(380,129)(0,0)%
624: \includegraphics{dpsin}%
625: \end{picture}
626: \begin{picture}(0,0)(380,0)%
627: \put(14,112){\makebox(0,0)[lb]{$\tilde{z}$}}
628: \put(76,4){\makebox(0,0)[lb]{$c$}}
629: \put(292.5,4){\makebox(0,0)[lb]{$c$}}
630: \put(130,78){\makebox(0,0)[lb]{$\gamma_1$}}
631: \put(158,78){\makebox(0,0)[lb]{$\gamma_2$}}
632: \put(39,2){\makebox(0,0)[lb]{$0$}}
633: \put(109,-1){\makebox(0,0)[lb]{$\frac{\pi}{2}$}}
634: \put(184,2){\makebox(0,0)[lb]{$\pi$}}
635: \put(254,-1){\makebox(0,0)[lb]{$\frac{n \pi}{2}$}}
636: \put(319,-3){\makebox(0,0)[lb]{$\frac{(n+1) \pi}{2}$}}
637: \end{picture}
638: \end{center}
639: \caption{ \small The state $-\partial_n \psi_n$ is given by a strip of
640: width $\frac{\pi}{2} (n+1)$ with two insertions of $c(\tilde{z})$ as
641: well as two contour integrals of $T(\tilde{z})$ and $b(\tilde{z})$
642: along the curves $\gamma_1$ and $\gamma_2$. }
643: \label{f:dpsi}
644: \end{figure}
645:
646: \section{Proof that $Q_\Psi$ has no cohomology}
647: \label{s:VanishingCohomologyProof}
648:
649: Having defined $\Psi$, we can now turn to the main aim of this paper,
650: to prove that $Q_\Psi$ has vanishing cohomology so that there are no
651: on-shell perturbative states around the tachyon vacuum. As discussed
652: in the introduction, we can do this using a trick, which we state as a
653: simple lemma:
654:
655: \bigskip
656:
657: \noindent
658: {\em Lemma:} The cohomology of a BRST operator $Q_\Psi$ vanishes if
659: and only if there exists a string field $A$ such that $Q_\Psi A =
660: \mathcal{I}$.
661:
662: \bigskip
663:
664: \noindent
665: {\em Proof:} First, suppose that $Q_\Psi$ has no cohomology. Consider
666: $Q_\Psi \mathcal{I} = Q_B \mathcal{I} + \Psi \starp \mathcal{I}
667: -\mathcal{I}\starp \Psi = Q_B \mathcal{I}$. Since, as was first shown
668: in \cite{Gross:1986ia,Gross:1986fk}, $Q_B \mathcal{I} =0$, it follows
669: that $Q_\Psi \mathcal{I} = 0$. Since $\mathcal{I}$ is $Q_\Psi$-closed
670: and $Q_\Psi$ has no cohomology, there must exist some $A$ such that
671: $\mathcal{I} = Q_\Psi A$.
672:
673: Now suppose, instead, that we have a state $A$ such that $Q_\Psi A =
674: \mathcal{I}$. Suppose we also have some $Q_\Psi$-closed state
675: $\Lambda$ such that $Q_\Psi \Lambda = 0$. Then
676: \begin{equation}
677: Q_\Psi (A\starp \Lambda) = (Q_\Psi A)\starp \Lambda = \mathcal{I} \starp \Lambda
678: = \Lambda,
679: \end{equation}
680: so that $\Lambda$ is $Q_\Psi$-exact. Since any $Q_\Psi$-closed state
681: is also $Q_\Psi$-exact, it follows that $Q_\Psi$ has no cohomology.
682:
683:
684: Such an operator $A$ is known in the math literature as a homotopy
685: operator. Note that the existence of $A$ proves that the cohomology
686: of $Q_\Psi$ vanishes at all ghost numbers, not just ghost number zero
687: as required by Sen's conjectures.\footnote{This seems to contradict
688: the numerical results of \cite{Giusto:2003wc}. Nonzero cohomology at
689: other ghost numbers has also been found for the so-called universal
690: solution \cite{Takahashi:2002ez} in \cite{Kishimoto:2002xi}.}
691:
692: \subsection{Finding the state $A$}
693:
694: We now describe how to find an $A$ satisfying,
695: \begin{equation}\label{eq:QPsiA}
696: Q_\Psi A = Q_B A + \Psi \starp A + A \starp \Psi = \mathcal{I}.
697: \end{equation}
698: Although (\ref{eq:QPsiA}) is a linear equation for $A$, a blind search
699: for a solution could be very difficult. Fortunately, for the
700: Feynman Siegel gauge solution, (\ref{eq:QPsiA}) was solved numerically
701: in \cite{Ellwood:2001ig} and we can use the results found there to guess a solution.
702:
703: Surprisingly, it was found in \cite{Ellwood:2001ig} that, in Feynman-Siegel gauge,
704: $A$ takes the approximate form,
705: \begin{equation}
706: A_{\text{FS}} \sim \frac{1}{L_0} b_0\mathcal{I}.
707: \end{equation}
708: Curiously, this form of $A_{\text{FS}}$ is the state one would write
709: down if one was trying to show that, in the {\em perturbative} vacuum,
710: $Q_B$ had vanishing cohomology. Indeed one has
711: \begin{equation}\label{QBAFS}
712: Q_B A_{\text{FS}} = \mathcal{I} - |0\rangle,
713: \end{equation}
714: so that one finds the identity state minus the one piece of the
715: identity that is in the cohomology of $Q_B$.
716:
717: A natural guess for the $\mathcal{B}_0$-gauge solution is to take the
718: same form for $A$, but with $b_0$ and $L_0$ replaced by their
719: counterparts in the $\tilde{z}$ coordinate, $\mathcal{B}_0$ and
720: $\mathcal{L}_0$;
721: \begin{equation}
722: A = \frac{1}{\mathcal{L}_0} \mathcal{B}_0 \mathcal{I}.
723: \end{equation}
724: It turns out that this $A$ can be written in a nicer form, as an
725: integral over wedge states with insertions. Using
726: (\ref{eq:OIequalsOIdagger}), we have
727: \begin{equation} \label{eq:BtoBBdagger}
728: A = \frac{1}{2\mathcal{L}_0} (\mathcal{B}_0 +\mathcal{B}^\bpz_0)
729: \mathcal{I}.
730: \end{equation}
731: Since $(\mathcal{B}_0 + \mathcal{B}_0^\bpz)$ raises the
732: $\mathcal{L}_0$-level by one, we may rewrite (\ref{eq:BtoBBdagger}) as
733: \begin{equation}
734: A = \tfrac{1}{2} (\mathcal{B}_0 +\mathcal{B}^\bpz_0)
735: \frac{1}{\mathcal{L}_0+1} \mathcal{I}.
736: \end{equation}
737: This can be further simplified by writing
738: \begin{equation}
739: \frac{1}{\mathcal{L}_0+1}
740: = \int_0^1 z^{\mathcal{L}_0} dz
741: = \int_0^1 dz\, U_{2/z} .
742: \end{equation}
743: Using (\ref{eq:UUdaggerrelation}), we have
744: \begin{equation}
745: U_{2/z} \mathcal{I} = U_{2/z} U_1^\bpz |0\rangle
746: = U_{2-z}^\bpz |0\rangle = |2-z\rangle,
747: \end{equation}
748: which yields\footnote{This result can also be found directly in the
749: $\mathcal{L}_0$-level expansion;
750: \begin{multline}
751: A = \frac{1}{2\mathcal{L}_0} (\mathcal{B}_0
752: + \mathcal{B}_0^\bpz)
753: \sum_{n = 0}^\infty \frac{1}{2^n n!} (\mathcal{L}_0
754: + \mathcal{L}_0^\bpz)^n |0\rangle
755: = \frac{1}{2} \sum_{n = 0}^\infty (\mathcal{B}_0
756: + \mathcal{B}_0^\bpz)
757: \frac{1}{2^n (n+1)!} (\mathcal{L}_0
758: + \mathcal{L}_0^\bpz)^n |0\rangle
759: \\
760: = \tfrac{1}{2} (\mathcal{B}_0
761: + \mathcal{B}_0^\bpz)
762: \frac{
763: e^{\frac{1}{2}
764: (\mathcal{L}_0 + \mathcal{L}_0^\bpz)} - 1}{(\mathcal{L}_0
765: + \mathcal{L}_0^\bpz)/2} |0\rangle
766: = \tfrac{1}{2} (\mathcal{B}_0
767: + \mathcal{B}_0^\bpz)
768: \int_1^2 dr\, e^{\frac{2-r}{2} (\mathcal{L}_0 + \mathcal{L}_0^\bpz)}
769: |0\rangle = \tfrac{1}{2} (\mathcal{B}_0
770: + \mathcal{B}_0^\bpz) \int_1^2 dr\, |r\rangle.
771: \end{multline}
772: }
773: \begin{equation}
774: A = \tfrac{1}{2} (\mathcal{B}_0+\mathcal{B}_0^\bpz) \int_0^1 dz
775: \,|2-z\rangle = \tfrac{1}{2} (\mathcal{B}_0+\mathcal{B}_0^\bpz)
776: \int_1^2 dr\, |r\rangle.
777: \end{equation}
778: Using (\ref{eq:KLRandLLdagger}) and (\ref{eq:LBLKonwedge}) this becomes
779: \begin{equation}
780: A = \tfrac{\pi}{2} B^{L}_1 \int_1^2 dr\, |r\rangle.
781: \end{equation}
782: This state has a simple geometric interpretation, as shown in
783: figure~\ref{f:Astate}.
784:
785: \begin{figure}
786: \begin{center}
787: \setlength{\unitlength}{1pt}
788: \begin{picture}(136,140)(0,0)%
789: \includegraphics{Astate}%
790: \end{picture}
791: \begin{picture}(0,0)(136,0)%
792: \put(12,118){\makebox(0,0)[lb]{$\tilde{z}$}}
793: \put(57,86){\makebox(0,0)[lb]{$\gamma$}}
794: \put(42,1){\makebox(0,0)[lb]{$\underbrace{\hspace{1.5cm}}_{\pi(r-1)/2}$}}
795: \put(-70,56){\makebox(0,0)[lb]{\parbox{1cm}{$$A =
796: \tfrac{\pi}{2} \int_1^2 dr$$}}}
797: \end{picture}
798: \end{center}
799: \caption{ \small The state $A$ can be represented as a sum over wedge
800: states $|r\rangle$. The only operator insertion is a single contour
801: integral of $b(\tilde{z})$ along the curve $\gamma$. }
802: \label{f:Astate}
803: \end{figure}
804:
805: \subsection{Computation of $Q_\Psi A$}
806:
807: The first term in $Q_\Psi A$ is just $Q_B A$. This is given by
808: \begin{equation}\label{eq:QBA}
809: Q_B A = \tfrac{\pi}{2} K_1^L \int_1^2 |r\rangle
810: = - \int_1^2 dr\, \partial_r |r\rangle
811: = \mathcal{I} - |0\rangle,
812: \end{equation}
813: which reproduces the Feynman-Siegel gauge result, (\ref{QBAFS}).
814:
815: Next we must compute the star-products $\Psi\starp A$ and $A\starp
816: \Psi$. Because the tachyon vacuum solution is twist invariant, these
817: two computations are related to each other by a twist. Hence, we need
818: to compute just one of them, $\Psi\starp A$.
819:
820: Since $\Psi = \psi_N - \sum_{m = 0}^n \partial_n \psi_n$, we begin by
821: evaluating $ \psi_n\starp A$. Using (\ref{eq:dpsidn}), we have
822: \begin{equation}
823: \psi_n\starp A = \int_1^2 dr \, c_1|0\rangle \starp B_1^L |n\rangle
824: \starp c_1 |0\rangle
825: \starp B_1^L |r\rangle,
826: \end{equation}
827: which we can rewrite using (\ref{eq:KLRonstarproduct}) as
828: \begin{equation}
829: \int_1^2 dr \, c_1|0\rangle \starp B_1^L B_1^R(|n\rangle
830: \starp c_1 |0\rangle
831: \starp |r\rangle).
832: \end{equation}
833: Now, using $B_1^L B_1^R = B_1^L (B_1-B_1^L) = B_1^L B_1$ this becomes
834: \begin{equation}
835: \int_1^2 dr \, c_1|0\rangle \starp B_1^L B_1(|n\rangle
836: \starp c_1 |0\rangle
837: \starp |r\rangle)
838: =\int_1^2 dr \, c_1|0\rangle \starp B_1^L (|n\rangle
839: \starp |0\rangle
840: \starp |r\rangle),
841: \end{equation}
842: where we have used the derivation property
843: (\ref{eq:K1B1asderivations}) of $B_1$ as well as
844: (\ref{eq:LBLKonwedge}). It follows that
845: \begin{equation} \label{eq:psistarA}
846: \psi_n \starp A = \int_1^2 dr \, B_1^R(c_1|0\rangle \starp |n+r\rangle).
847: \end{equation}
848: Similarly, one can compute $A\starp \psi_n$ either by repeating the
849: above computation or by exploiting twist symmetry. Either way, one
850: finds
851: \begin{equation}\label{eq:Astarpsi}
852: A\starp \psi_n = \int_1^2 dr\, B_1^L (|r+n\rangle \starp c_1 |0\rangle).
853: \end{equation}
854: Now consider $- \sum_{n=0}^N \partial_n \psi_n\starp A$. Since $n$
855: and $r$ appear only in the combination $n+r$, we can replace the
856: derivative $\partial_n$ with $\partial_r$. This gives
857: \begin{equation}
858: - \sum_{n=0}^N \partial_n \psi_n\starp A = -\sum_{n=0}^N \int_1^2 dr
859: \, B_1^R(c_1|0\rangle \starp \partial_r |n+r\rangle)
860: = \sum_{n=0}^N
861: \, B_1^R(c_1|0\rangle \starp \{|n+1\rangle - |n+2\rangle\}).
862: \end{equation}
863: Notice that the sum can now be trivially performed since
864: \begin{equation}
865: \sum_{n = 0}^N |n+1\rangle - |n+2\rangle
866: = \mathcal{I} - |N+2\rangle.
867: \end{equation}
868: Hence, we find
869: \begin{equation}\label{eq:dpsistarA}
870: - \sum_{n=0}^N \partial_n \psi_n\starp A
871: = B_1^R c_1|0\rangle
872: - B_1^R (c_1|0\rangle \starp |N+2\rangle ).
873: \end{equation}
874: Similarly, one can compute
875: \begin{equation}\label{eq:Astardpsi}
876: - A \starp \sum_{n=0}^N \partial_n \psi_n
877: =
878: B_1^L c_1|0\rangle
879: - B_1^L (|N+2\rangle \starp c_1|0\rangle ).
880: \end{equation}
881: Using (\ref{eq:psistarA}), (\ref{eq:Astarpsi}), (\ref{eq:dpsistarA})
882: and (\ref{eq:Astardpsi}), we find, in total, that
883: \begin{equation}
884: \Psi\starp A+ A\starp \Psi
885: = |0\rangle -\Sigma,
886: \end{equation}
887: where the state $\Sigma$ is given by
888: \begin{equation}
889: \Sigma=
890: B_1^R\left\{c_1|0\rangle \starp
891: \left(|N+2\rangle - \int_1^2 dr \,|N+r\rangle\right)\right\}
892: \\
893: +B_1^L\left\{ \left(|N+2\rangle - \int_1^2 dr \,|N+r\rangle\right)
894: \starp c_1|0\rangle\right\}.
895: \end{equation}
896: Now, as $N \to \infty$, the state $|N\rangle$ limits to the sliver so
897: that $|N+2\rangle - \int_1^2 dr\, |N+r\rangle \to 0$ as $N \to
898: \infty$. In fact, it is straightforward to check that, in the level
899: expansion, it goes to zero as $\mathcal{O}(N^{-3})$. Hence, when we
900: remove the regulator we find
901: \begin{equation} \label{psistarAplusAstarpsi}
902: \Psi \starp A + A \starp \Psi = |0\rangle.
903: \end{equation}
904: Note that it was important to include the $\psi_N$ piece in $\Psi$ to
905: cancel out the surface terms in the sums (\ref{eq:dpsistarA}) and
906: (\ref{eq:Astardpsi}). Combining (\ref{psistarAplusAstarpsi}) with
907: (\ref{eq:QBA}), we find the desired result;
908: \begin{equation}
909: Q_\Psi A = Q_B A + \Psi \starp A + A\starp \Psi = \mathcal{I}.
910: \end{equation}
911: This proves that the cohomology of $Q_\Psi$ is empty.
912:
913: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
914: %%%%%%%%%%%%%%%%
915: \subsection{Comparison with vacuum string field theory}
916: \label{s:Aproperties}
917: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
918: %%%%%%%%%%%%%%%%
919: It is interesting to compare our results with the results of vacuum
920: string field theory (VSFT) \cite{Rastelli:2000hv,Rastelli:2001uv}. In
921: VSFT, the BRST operator around the tachyon vacuum is taken, by ansatz,
922: to be a simple pure ghost operator. For example, one of the early
923: choices was the zero mode of the $c$-ghost, $c_0$. To show that $c_0$
924: has empty cohomology, one notes that $\{c_0,b_0\} = 1$, so that $b_0$
925: plays the role of our string field, $A$.
926:
927: This analogy can be made a little closer. Just as $b_0^2 = 0$, it
928: happens that $A\starp A =0$. This property is easy to see from the
929: geometric form of $A$ in figure \ref{f:Astate}. Moreover, just as
930: $b_0$ is a Hermitian operator, one can also construct a Hermitian
931: operator $\hat{A}$ defined by
932: \begin{equation}
933: \hat{A} \Phi = A\starp \Phi + (-1)^{\text{gh}(\Phi)} \Phi \starp A,
934: \end{equation}
935: which satisfies $\hat{A}^2 = 0$ and $\{Q_\Psi ,\hat{A}\} = 1$, as well
936: as the Hermiticity property, $\langle \Phi_1 | \hat{A} \Phi_2\rangle =
937: \langle \hat{A} \Phi_1 | \Phi_2\rangle$. Since VSFT is thought to be
938: a singular limit of ordinary OSFT, in which the BRST operator becomes a
939: $c$-ghost operator inserted at the midpoint \cite{Gaiotto:2001ji}, it
940: would be interesting to see whether $\hat{A}$ becomes a simple
941: operator formed out of just the $b$-ghost in this limit.
942:
943: \subsection{Brane decay in the presence of other branes}
944:
945: In this subsection, we show that one can extend our cohomology
946: arguments to the case where we include other branes that have not
947: decayed. Consider OSFT around a 2 brane background, which we describe
948: by adding Chan-Paton indices to our string fields;
949: \begin{equation}
950: \phi =
951: \left(\begin{matrix}
952: \Phi_{11} & \Phi_{12}
953: \\
954: \Phi_{21} & \Phi_{22}
955: \end{matrix}
956: \right),
957: \end{equation}
958: where $\phi^\dagger = \phi$. To decay one of the branes, we may turn on
959: \begin{equation}
960: \psi =
961: \left(\begin{matrix}
962: \Psi & 0
963: \\
964: 0 & 0
965: \end{matrix}
966: \right).
967: \end{equation}
968: The BRST operator $Q_\psi$ acts as
969: \begin{equation} \label{multiBRST}
970: Q_B \phi + [\psi ,\phi]
971: =
972: \left(\begin{matrix}
973: Q_B \Phi_{11} + [\Psi,\Phi_{11}] & Q_B \Phi_{12} + \Psi \starp \Phi_{12}
974: \\
975: Q_B \Phi_{21} -
976: (-1)^{\text{gh}(\Phi_{21})} \Phi_{21} \starp \Psi & Q_B \Phi_{22}
977: \end{matrix}
978: \right).
979: \end{equation}
980: Since we have decayed the first brane, we expect that there are no
981: on-shell $11$, $12$ or $21$ strings. This implies that the three
982: BRST-operators,
983: \begin{equation}
984: Q_{11}\Phi = Q_B\Phi + [\Psi , \Phi], \qquad
985: Q_{12}\Phi = Q_B\Phi + \Psi\starp\Phi ,
986: \qquad \text{and} \qquad
987: Q_{21}\Phi = Q_B\Phi - (-1)^{\text{gh}(\Phi)}\Phi \starp \Psi,
988: \end{equation}
989: should all have vanishing cohomology. Since $Q_{11} = Q_\Psi$, there
990: is nothing new to show. For $Q_{12}$ and $Q_{21}$, our old argument
991: still works as long as we are careful about left multiplication versus
992: right multiplication. Suppose that $Q_{12} \Phi = 0$. Then it is easy
993: to check that
994: \begin{equation}
995: Q_{12} (A\starp \Phi) =(Q_{\Psi} A) \starp \Phi = \Phi.
996: \end{equation}
997: Thus, as we expect, every closed state is exact.
998: Similarly, if $Q_{21} \Phi = 0$, we have
999: \begin{equation}
1000: Q_{21} (-\Phi \starp A) = \Phi \starp (Q_\Psi A) = \Phi.
1001: \end{equation}
1002: Putting the $A$ on the left of $\Phi$ would not work. Hence, we have
1003: shown that the only open strings that remain in the spectrum are those
1004: that live on the undecayed brane. This argument generalizes to the
1005: case of $n$ decayed branes and $m$ undecayed branes in the expected
1006: way.
1007:
1008:
1009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1010: %%%%%%%%%%%%%%%%
1011: \sectiono{Pure-gauge-like form}
1012: \label{s:pg}
1013: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1014: %%%%%%%%%%%%%%%%
1015: One of the curious features of the analytic tachyon vacuum is that it
1016: is very close to being pure gauge. Indeed, it was found by Okawa
1017: \cite{Okawa:2006vm} that if one ignores the $\psi_N$ term -- which one
1018: can in the $L_0$ basis\footnote{Interestingly, in the $\ll_0$ level
1019: truncation we find $\Psi_\lambda = \frac{\lambda}{1-\lambda} Q\Phi +
1020: \cdots$, where the dots stand for terms of $\ll_0$-level higher than
1021: $0$. Hence, the $\lambda \to 1$ limit does not exist in this basis.}
1022: -- the full solution can be written as the limit, $\lambda \to 1$, of
1023: the state,
1024: \begin{equation}
1025: \Psi_\lambda = U_\lambda \starp Q_B V_\lambda,
1026: \end{equation}
1027: where\footnote{$V_\lambda$ is defined by the Taylor series,
1028: \begin{equation}
1029: \frac{1}{1-\lambda \Phi} = \sum_{n=0}^\infty \lambda^n \Phi^n;
1030: \qquad \Phi^n = \underbrace{\Phi \starp \Phi \starp \ldots \starp \Phi}_{n}.
1031: \end{equation}}
1032: \begin{equation}
1033: U_\lambda = 1-\lambda \Phi, \qquad V_\lambda = \frac{1}{1-\lambda \Phi}
1034: \end{equation}
1035: and
1036: \begin{equation}
1037: \Phi = B_1^L c_1 |0\rangle.
1038: \end{equation}
1039: When $\lambda <1$ the states $U_\lambda$ and $V_\lambda$ are well
1040: defined in the level-expansion and the state $\Psi_\lambda$ is a true
1041: pure-gauge solution with zero energy.
1042:
1043: Obviously, the tachyon solution itself, cannot be a pure-gauge
1044: solution related by a continuous deformation to the vacuum for two
1045: reasons. First, the energy of such a solution would have to be zero in
1046: contradiction with the now proven Sen's first conjecture. Second, it
1047: would imply that the cohomology of the kinetic operator at the true
1048: vacuum would be isomorphic to the cohomology of $Q_B$ in contradiction
1049: with Sen's third conjecture. It is therefore interesting to
1050: understand how the solution ceases to be a pure gauge at $\lambda=1$
1051: and how Sen's conjectures are rescued.
1052:
1053: The basic property of the pure-gauge solutions is that
1054: \begin{equation}\label{VUandUV}
1055: U_\lambda \starp V_\lambda = V_\lambda \starp U_\lambda = \mathcal{I}.
1056: \end{equation}
1057: This allows one to define an isomorphism between the states in the
1058: perturbative vacuum and their corresponding states in the pure-gauge
1059: vacuum;
1060: \begin{equation}
1061: \phi \to \mathcal{F}_\lambda[\phi]= U_\lambda \starp \phi \starp V_\lambda,
1062: \end{equation}
1063: which has inverse, $\mathcal{F}^{-1}_\lambda[\phi] = V_\lambda \starp \phi \starp U_\lambda$.
1064:
1065: This isomorphism relates the original BRST operator, $Q_B$, to the new BRST operator, $Q_\lambda$,
1066: around the pure-gauge vacuum;
1067: \begin{equation}\label{BRSTisomorphism}
1068: Q_\lambda (\mathcal{F}_\lambda[\phi])= \mathcal{F}_\lambda [Q_B \phi].
1069: \end{equation}
1070: It follows that the two operators have identical cohomology.
1071:
1072: We can now ask how (\ref{VUandUV})-(\ref{BRSTisomorphism}) break down
1073: when $\lambda \to 1$. Clearly, since the right hand side of
1074: (\ref{VUandUV}) is independent of $\lambda$, we will find
1075: $\lim_{\lambda \to 1} U_\lambda \starp V_\lambda = \lim_{\lambda \to
1076: 1} V_\lambda \starp U_\lambda = \mathcal{I}$. However, the state
1077: $V_\lambda$ by itself diverges in the $\ll_0$ level-expansion,
1078: although it appears to remain finite in the $L_0$ expansion.
1079:
1080: Similar divergences occur when we consider $\mathcal{F}_{\lambda}
1081: (\phi)$ and its inverse. For concreteness, take $\phi = c \mathcal{O}
1082: |0\rangle$, where $\mathcal{O}$ is a matter operator that satisfies
1083: \begin{equation}
1084: [\ll_0,\mathcal{O}] = h \mathcal{O}.
1085: \end{equation}
1086: Following the rules of \cite{Schnabl:2005gv} we find
1087: %
1088: \begin{multline}
1089: \mathcal{F}_\lambda[\phi] = c\mathcal{O}(0)\ket{0} + \sum_{m=1}^\infty \lambda^m U_{m+2}^\bpz
1090: U_{m+2} \left\lbrace \frac{1}{2} \tilde{\mathcal{O}}(x) (\tilde c(x)+ \tilde c(-x)) + \frac{1}{2}
1091: \tilde{\mathcal{O}}(y) (\tilde c(x) - \tilde c(y)) \right.
1092: \\
1093: \left. - \frac{1}{\pi} (\mathcal{B}_0+\mathcal{B}^\bpz) \left( \tilde{\mathcal{O}}(x)\tilde
1094: c(x)\tilde c(-x)+ \tilde{\mathcal{O}}(y)
1095: (\tilde c(x) - \tilde c(y))\tilde c(-x) \right)\right\rbrace \ket{0} ,
1096: \end{multline}
1097: %
1098: where, for brevity, we have introduced $x=\frac{\pi}{4} m$ and
1099: $y=\frac{\pi}{4} (m-2)$. Using this form, it is straightforward to work
1100: out the coefficients in the $\ll_0$ basis
1101: %
1102: \begin{align}
1103: \mathcal{F}_\lambda[\phi] = & \frac{1}{1-\lambda} c\mathcal{O}(0) \ket{0} + \nonumber \\ &
1104: +\frac{\lambda}{(1-\lambda)^2} \left[ -\frac{1}{2} \left(\ll_0 + \ll_0^\bpz\right) \tilde
1105: c\tilde{\mathcal{O}}(0) + \left(\bb_0 + \bb_0^\bpz\right) \tilde c\tilde\partial \tilde c
1106: \tilde{\mathcal{O}}(0) + \frac{\pi}{4} \bigl((1-\lambda)\tilde\partial \tilde c
1107: \tilde{\mathcal{O}}(0) + \tilde c \tilde\partial \tilde{\mathcal{O}}(0)\bigr) \right] \ket{0} +
1108: \nonumber\\ & + \cdots,
1109: \end{align}
1110: %
1111: where the dots stand for terms of higher $\ll_0$-level. We see that,
1112: due to the presence of poles at $\lambda=1$, the state
1113: $\mathcal{F}_{\lambda=1}[\phi]$ does not make sense in this
1114: basis. Note that one cannot rescale $\phi_\lambda$ by a positive power
1115: of $1-\lambda$ to get a finite representative of the cohomology, since
1116: the maximal order of the poles grows with level.
1117:
1118: We find similar behavior when we compute $\mathcal{F}[\phi]$ in the
1119: ordinary $L_0$ level truncation. Since such computations are more
1120: difficult, we have restricted ourselves to the case where
1121: $\mathcal{O}$ is a weight one primary. This case is of particular
1122: interest, as any cohomology class of $Q_B$ has a representative of
1123: this form.
1124:
1125: Computing the coefficient of $\mathcal{F}_\lambda[\phi]$ in front of $c\mathcal{O}(0)\ket{0}$, we find
1126: %
1127: \be
1128: 1+\sum_{m=1}^\infty \lambda^m \frac{m+2}{2} \left[\frac{1}{2} +
1129: \frac{\alpha}{\pi}\left(1-\left(\frac{\sin\alpha}{\sin
1130: 2\alpha}\right)^2\right) - \frac{1}{\pi} \sin 2\alpha + \frac{1}{2\pi}
1131: \left(\frac{\sin\alpha}{\sin 2\alpha}\right)^2 \left(\sin 4\alpha
1132: -\sin 2\alpha \right) \right],
1133: \ee
1134: %
1135: where $\alpha = \pi/(m+2)$. Since
1136: the summand behaves as \bdm \lambda^m \left[ \frac{m+2}{4}
1137: -\frac{1}{2} + \frac{\pi^2}{12} \frac{1}{(m+2)^2} + \cdots \right],
1138: \edm
1139: %
1140: we see that, apart from the mild polylogarithmic singularities at
1141: $\lambda=1$, which are present also for the solution $\Psi_\lambda$
1142: itself, $\mathcal{F}_\lambda[\phi]$ contains double and single poles
1143: and therefore the limit $\lim_{\lambda \to 1}
1144: \mathcal{F}_\lambda[\phi]$ does not exist.
1145:
1146: So far in this discussion we have tried to show that elements of the
1147: cohomology of $Q_B$ are not mapped via $\mathcal{F}$ to elements of
1148: the cohomology of $Q_\Psi$. However, it is also interesting to ask
1149: why $A$ cannot be pulled back to the perturbative vacuum to show that
1150: $Q_B$ has no cohomology. Hence, we compute
1151: \begin{multline}
1152: \mathcal{F}^{-1} (A)
1153: = \frac{1}{1-\Phi} \starp A \starp (1-\Phi)
1154: = \tfrac{\pi}{2} \int_1^2 dr\, (1+ \sum_{n=1}^\infty B_1^L |n\rangle \starp c_1 |0\rangle)
1155: \starp B_1^L |r\rangle \starp (1-B_1^L c_1 |0\rangle)
1156: \\
1157: = \tfrac{\pi}{2} \sum_{n=1}^\infty \int_1^2 dr\, B_1^L |n+r-1\rangle.
1158: \end{multline}
1159: This simplifies to
1160: \begin{equation}
1161: \mathcal{F}^{-1}(A) = \tfrac{\pi}{2} \int_1^\infty dr\, B_1^L |r\rangle,
1162: \end{equation}
1163: which should be thought of as the ``$A$" of the perturbative vacuum.
1164: We can now act on this state with $Q_B$;
1165: \begin{equation} \label{eq:QFA}
1166: Q_B(\mathcal{F}_{-1}(A)) = -\int_1^\infty dr\, \partial_r |r\rangle =
1167: \mathcal{I} - |\infty\rangle.
1168: \end{equation}
1169: Happily, we do not find just the identity on the right hand side, so
1170: the cohomology of $Q_B$ need not vanish.\footnote{Formally one could
1171: write $Q_B \mathcal{F}^{-1}(A)=\mathcal{F}^{-1}(Q_\Psi
1172: A)=\mathcal{F}^{-1}(\mathcal{I}) = V_{\lambda=1} \starp
1173: U_{\lambda=1}$. Using (\ref{eq:QFA}), this would imply $V \starp
1174: U=\mathcal{I} - |\infty\rangle$ suggesting that $V$ and $U$ are a
1175: nontrivial pair of partial isometries as first proposed in
1176: \cite{Schnabl:2000cp}. On the other hand a direct computation seems to
1177: yield $V \starp U=\mathcal{I}$ in the strict $\lambda \to 1$ limit, in
1178: both $L_0$ and $\mathcal{L}_0$ level truncation. It would be nice to
1179: understand this anomaly more deeply.} Equation (\ref{eq:QFA}) has a
1180: nice interpretation in terms of half strings. Consider a state $\phi$
1181: which is $Q_B$-closed, but whose left half has no overlap with the
1182: right half of $|\infty\rangle$. In other words, $|\infty\rangle \starp
1183: \phi = 0$. It follows that $\phi$ is $Q_B$-exact. To see this,
1184: consider
1185: \begin{equation}
1186: Q_B(\mathcal{F}^{-1}(A) \starp \phi ) = (\mathcal{I} -|\infty\rangle ) \starp\phi = \phi.
1187: \end{equation}
1188: A similar result holds for states whose right half has no overlap with the left half of
1189: $|\infty\rangle$. This implies that the entire cohomology of $Q_B$ should be found on states whose
1190: left and right halves are given by the left and right halves of $|\infty\rangle$. Such a set of
1191: states is easy to find. For example, at ghost number $0$, the cohomology of $Q_B$ is represented
1192: by just $|\infty\rangle$ itself. At ghost number 1, which is the interesting case, the cohomology
1193: of $Q_B$ has representatives given by weight $(0,0)$ primaries of the form $c J$, where $J$ is a
1194: weight one matter primary. Inserting these operators at the midpoint of $|\infty\rangle$ gives a
1195: set of ghost number $1$ states in the cohomology of $Q_B$ with left and right halves given by the
1196: left and right halves of $|\infty \rangle$.
1197:
1198:
1199: %\section{Discussion}
1200: %\label{s:Discussion}
1201: %
1202: %In this paper we gave a concise proof of Sen's third conjecture that there
1203: %are no open string states around the tachyon vacuum.
1204: %
1205: \section*{Acknowledgments} We would like to thank T.~Grimm, A.~Hashimoto, L.~Motl, Y.~Okawa, E.~Witten and B.~Zwiebach
1206: for useful conversations. I.E. is supported in part by the DOE grant DE-FG02-95ER40896 and by
1207: funds from the University of Wisconsin.
1208:
1209: \bibliography{c}\bibliographystyle{utphys}
1210:
1211: \end{document}
1212: \bye
1213: