hep-th0311270/24.tex
1: \documentclass[nohyper,11pt,letterpaper]{JHEP3}
2: \usepackage[dvips]{epsfig}
3: %\usepackage{showkeys} % for cool labelling!
4: \usepackage{epsf,amsfonts,amssymb}
5: %\usepackage[active]{srcltx}
6: 
7: 
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: %%%%%%%%%%%%%%%%%%%%%%%%%% Macros %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
10: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
11: 
12: %%%%%%%%%%%%%%%%%%%%%%%% Structure definitions %%%%%%%%%%%%%%%%%%%%%%%%%%%
13: 
14: \newcommand{\eqn}[1]{(\ref{#1})}
15: \newcommand{\be}{\begin{equation}}
16: \newcommand{\ee}{\end{equation}}
17: \newcommand{\ben}{\begin{displaymath}}
18: \newcommand{\een}{\end{displaymath}}
19: \newcommand{\bea}{\begin{eqnarray}}
20: \newcommand{\eea}{\end{eqnarray}}
21: \newcommand{\bean}{\begin{eqnarray*}}
22: \newcommand{\eean}{\end{eqnarray*}}
23: \newcommand{\nn}{\nonumber \\}
24: \newcommand{\ba}{\begin{array}}
25: \newcommand{\ea}{\end{array}}
26: \newcommand{\bi}{\begin{itemize}}
27: \newcommand{\ei}{\end{itemize}}
28: 
29: \newtheorem{theorem}{Theorem}
30: 
31: %%%%%%%%%%%%%%%%%%%%%%%% Greek Letters %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: 
33: \def\l {\lambda}
34: \def\a {\alpha}
35: \def\ap {\alpha'}
36: \def\b {\beta}
37: \def\g {\gamma}
38: \def\G {\Gamma}
39: \def\d {\delta}
40: \def\s {\sigma}
41: \def\e {\epsilon}
42: \def\vt {\vartheta}
43: \def\vp {\varphi}
44: \def\T {\Theta}
45: 
46: \renewcommand{\O}{\Omega}
47: \renewcommand{\L}{\Lambda}
48: \renewcommand{\t}{\theta}
49: 
50: %%%%%%%%%%%%%%%%%% Calligraphic Letters %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: 
52: \newcommand{\cala}{\mbox{${\cal A}$}}
53: \newcommand{\calb}{\mbox{${\cal B}$}}
54: \newcommand{\calc}{\mbox{${\cal C}$}}
55: \newcommand{\cald}{\mbox{${\cal D}$}}
56: \newcommand{\cale}{\mbox{${\cal E}$}}
57: \newcommand{\calf}{\mbox{${\cal F}$}}
58: \newcommand{\calg}{\mbox{${\cal G}$}}
59: \newcommand{\calh}{\mbox{${\cal H}$}}
60: \newcommand{\cali}{\mbox{${\cal I}$}}
61: \newcommand{\calj}{\mbox{${\cal J}$}}
62: \newcommand{\calk}{\mbox{${\cal K}$}}
63: \newcommand{\call}{\mbox{${\cal L}$}}
64: \newcommand{\calm}{\mbox{${\cal M}$}}
65: \newcommand{\caln}{\mbox{${\cal N}$}}
66: \newcommand{\calo}{\mbox{${\cal O}$}}
67: \newcommand{\calp}{\mbox{${\cal P}$}}
68: \newcommand{\calq}{\mbox{${\cal Q}$}}
69: \newcommand{\calr}{\mbox{${\cal R}$}}
70: \newcommand{\cals}{\mbox{${\cal S}$}}
71: \newcommand{\calt}{\mbox{${\cal T}$}}
72: \newcommand{\calu}{\mbox{${\cal U}$}}
73: \newcommand{\calv}{\mbox{${\cal V}$}}
74: \newcommand{\calw}{\mbox{${\cal W}$}}
75: \newcommand{\calx}{\mbox{${\cal X}$}}
76: \newcommand{\caly}{\mbox{${\cal Y}$}}
77: \newcommand{\calz}{\mbox{${\cal Z}$}}
78: 
79: \newcommand{\bfcalc}{\mbox{\boldmath ${\cal C}$}}
80: 
81: 
82: %%%%%%%%%%%%%%%%% Boldmath Letters %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
83: 
84: \newcommand{\bfe}{\mbox{\boldmath $E$}}
85: \newcommand{\bfb}{\mbox{\boldmath $B$}}
86: \newcommand{\bfpi}{\mbox{\boldmath $\Pi$}}
87: \newcommand{\bfp}{\mbox{\boldmath $P$}}
88: \newcommand{\bfna}{\mbox{\boldmath $\nabla$}}
89: \newcommand{\bfd}{\mbox{\boldmath $D$}}
90: \newcommand{\bfeh}{\mbox{\boldmath $\hat{E}$}}
91: \newcommand{\bfbh}{\mbox{\boldmath $\hat{B}$}}
92: \newcommand{\bfah}{\mbox{\boldmath $\hat{A}$}}
93: \newcommand{\bx}{\mbox{\boldmath $x$}}
94: \newcommand{\by}{\mbox{\boldmath $y$}}
95: \newcommand{\bX}{\mbox{\boldmath $X$}}
96: \newcommand{\bY}{\mbox{\boldmath $Y$}}
97: \newcommand{\bV}{\mbox{\boldmath $V$}}
98: \newcommand{\bU}{\mbox{\boldmath $U$}}
99: 
100: \newcommand{\bfK}{\mbox{\boldmath $K$}}
101: \newcommand{\bfz}{\mbox{\boldmath $\theta$}}
102: \newcommand{\bfZ}{\mbox{\boldmath $Z$}}
103: \newcommand{\bfr}{\mbox{\boldmath $\rho$}}
104: \newcommand{\bfxi}{\mbox{\boldmath $\xi$}}
105: \newcommand{\bfs}{\mbox{\boldmath $\sigma$}}
106: \newcommand{\bft}{\mbox{\boldmath $t$}}
107: \newcommand{\bfT}{\mbox{\boldmath $T$}}
108: \newcommand{\bfss}{\mbox{\boldmath $s$}}
109: \newcommand{\bfS}{\mbox{\boldmath $S$}}
110: \newcommand{\bfG}{\mbox{\boldmath $\Gamma$}}
111: \newcommand{\bfo}{\mbox{\boldmath $0$}}
112: 
113: \newcommand{\bbe}[1]{\mbox{${\mathbb E}^{#1}$}}
114: \newcommand{\bbr}[1]{\mbox{${\mathbb R}^{#1}$}}
115: %\newcommand{\bbz}[1]{\mbox{${\mathbb Z}^{#1}$}}
116: \newcommand{\bbz}[1]{\mbox{${\mathbb Z}_{#1}$}}
117: \newcommand{\bbi}[1]{\mbox{${\mathbb I}_{#1}$}}
118: \newcommand{\bbo}[1]{\mbox{${\mathbb O}_{#1}$}}
119: 
120: %%%%%%%%%%%%%%%%%%%%%% Miscellaneous  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
121: 
122: \newcommand{\nef}{{$\caln =4$} }
123: \newcommand{\mkk}{M_\mt{KK}}
124: 
125: \newcommand{\dd}[3]{\mbox{$( #1 | \mbox{D} #2 \perp \mbox{D} #3)$}}
126: \newcommand{\ds}[3]{\mbox{$( #1 | D #2 \perp \calt #3)$}}
127: \newcommand{\para}{\parallel}
128: \newcommand{\inter}{\, \cap \,}
129: \newcommand{\su}[1]{$\calt #1\,$}
130: 
131: \newcommand{\ads}[1]{\mbox{${AdS}_{#1}$}}
132: \newcommand{\adss}[2]{\mbox{$AdS_{#1}\times {S}^{#2}$}}
133: \newcommand{\rn}{Reissner-Nordstr\"{o}m }
134: 
135: \newcommand{\bra}[1]{\mbox{$\langle #1 |$}}
136: \newcommand{\ket}[1]{\mbox{$| #1 \rangle$}}
137: \newcommand{\braket}[2]{\mbox{$\langle #1  | #2 \rangle$}}
138: \newcommand{\proj}[1]{\ket{#1}\!\bra{#1}}
139: 
140: \newcommand{\Gu}[1]{\Gamma_{\underline{#1}}}
141: \newcommand{\Gn}{\Gamma_\natural}
142: 
143: \newcommand{\un}{\underline}
144: \newcommand{\pa}{\partial}
145: \newcommand{\fc}{\frac}
146: \newcommand{\w}{\wedge}
147: \newcommand{\trace}{\mbox{Tr}}
148: \newcommand{\sac}{\ , \qquad}
149: \newcommand{\eg}{{\it e.g.}}
150: \newcommand{\ie}{{\it i.e.}}
151: \newcommand{\sgn}[1]{\mbox{sgn}(#1)}
152: \newcommand{\com}[1]{[[[{\bf #1}]]]}
153: \newcommand{\ad}{\dot{a}}
154: \newcommand{\bd}{\dot{b}}
155: \newcommand{\ik}{{\it k}}
156: \newcommand{\ione}{{\it 1}}
157: \newcommand{\itwo}{{\it 2}}
158: \newcommand{\ifive}{{\it 5}}
159: 
160: \newcommand{\ph}{\hat{P}}
161: \newcommand{\xh}{\hat{X}}
162: \newcommand{\ve}{\varepsilon}
163: \newcommand{\tr}{\mbox{Tr}}
164: 
165: \newcommand{\ra}{\rightarrow}
166: 
167: \newcommand{\comment}[1]{{\bf [[[#1]]]}}
168: 
169: %\newcommand{\atmp}[3]{{\it Adv. Theor. Math. Phys.} {\bf #1} {(#2)} #3}
170: %\newcommand{\ijtp}[3]{{\it Int. J. Theor. Phys.} {\bf #1} {(#2)} #3}
171: 
172: 
173: %%%%%%%%%%%%%%%%% Equation Numbering %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
174: 
175: \newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
176: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
177: 
178: %%%%%%%%%%%%%%%%%%%% Page formatting %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
179: 
180: %\renewcommand{\baselinestretch}{1.4}
181: %\hoffset -0.5in % moves text horizontally
182: %\textwidth 165mm
183: %\textheight 220mm
184: %\evensidemargin 0mm
185: %\topmargin -15mm
186: %\headsep 1.3cm
187: %\footskip 20mm
188: %\headheight 0pt
189: 
190: 
191: %%%%%%%%%%%%%%%%%% For this paper %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
192: 
193: 
194: \newcommand{\beq}{\begin{equation}}
195: \newcommand{\eeq}{\end{equation}}
196: \newcommand{\beqr}{\begin{displaymath}}
197: \newcommand{\eeqr}{\end{displaymath}}
198: \newcommand{\beqa}{\begin{eqnarray}}
199: \newcommand{\eeqa}{\end{eqnarray}}
200: \newcommand{\beqar}{\begin{eqnarray*}}
201: \newcommand{\eeqar}{\end{eqnarray*}}
202: %\newcommand{\htx}{s} %\hat{t}}
203: %\renewcommand{\a}{\alpha}
204: %\newcommand{\ap}{\alpha'}
205: %\renewcommand{\b}{\beta}
206: %\renewcommand{\d}{\delta}
207: %\newcommand{\D}{\Delta}
208: %\newcommand{\e}{\epsilon}
209: %\newcommand{\g}{\gamma}
210: %\newcommand{\G}{\Gamma}
211: \renewcommand{\k}{\kappa}
212: %\renewcommand{\l}{\lambda}
213: %\renewcommand{\L}{\Lambda}
214: \newcommand{\m}{\mu}
215: \newcommand{\n}{\nu}
216: \newcommand{\p}{\phi}
217: %\renewcommand{\o}{\omega}
218: %\renewcommand{\O}{\Omega}
219: %\renewcommand{\P}{\Phi}
220: %\newcommand{\p}{\phi}
221: %\newcommand{\Ps}{\Psi}
222: \renewcommand{\r}{\rho}
223: \newcommand{\vr}{\varrho}
224: %\newcommand{\s}{\sigma}
225: %\renewcommand{\t}{\theta}
226: %\newcommand{\vt}{\vartheta}
227: %\newcommand{\ve}{\varepsilon}
228: %\newcommand{\z}{\zeta}
229: %\newcommand{\x}{\xi}
230: %\newcommand{\cM}{{\cal M}}
231: %\newcommand{\cA}{{\cal A}}
232: %\newcommand{\cH}{{\cal H}}
233: %\newcommand{\cG}{{\cal G}}
234: %\newcommand{\cS}{{\cal S}}
235: \newcommand{\cN}{{\cal N}}
236: %\newcommand{\cD}{{\cal D}}
237: %\newcommand{\cO}{{\cal O}}
238: %\newcommand{\cF}{{\cal F}}
239: \newcommand{\cL}{{\cal L}}
240: \newcommand{\ssc}{\scriptscriptstyle}
241: %%\newcommand{\eg}{{\it e.g.}}
242: %%\newcommand{\ie}{{\it i.e.}}
243: %\newcommand{\etal}{{\it et al.~}}
244: \newcommand{\labell}[1]{\label{#1}\qquad_{#1}} %{\label{#1}}
245: %\newcommand{\labels}[1]{\label{#1}} %{\vskip-.6cm${#1}$\label{#1}}
246: \newcommand{\reef}[1]{(\ref{#1})}
247: \newcommand{\non}{\nonumber}
248: \newcommand{\pf}{\partial}
249: %\renewcommand{\ni}{\vspace{0.2cm}\noindent}
250: %\newcommand{\dir}[1]{\overline{#1}}
251: %\newcommand{\ZZ}{\mathbb{Z}}
252: %\newcommand{\ket}[1]{|#1\rangle}
253: %\newcommand{\bra}[1]{\langle #1|}
254: %\newcommand{\sR}{\textrm{\tiny R}}
255: %\newcommand{\sNS}{\textrm{\tiny NS}}
256: %\newcommand{\tr}{\textrm{Tr}\;}
257: %\newcommand{\til}[1]{\widetilde{#1}}
258: %\newcommand{\ti}[1]{\tilde{#1}}
259: \newcommand{\df}{\textrm{d}}
260: %\newcommand{\sla}[1]{#1\hspace{-0.16cm}/}
261: \newcommand{\mt}[1]{\textrm{\tiny #1}}
262: %\newcommand{\ph}{\phantom{1}}
263: %\newcommand{\com}[1]{{\bf (#1)}}
264: \newcommand{\cF}{{\cal F}}
265: \newcommand{\cA}{{\cal A}}
266: \newcommand{\cB}{{\cal B}}
267: \newcommand{\ls}{\ell_s}
268: 
269: \newcommand{\zD}{\ensuremath{z_{D7}}}      % Position of D7 brane
270: \newcommand{\tzD}{\ensuremath{\zeta_m}}      % Position of D7 brane
271: \newcommand{\trho}{\ensuremath{\tilde{\rho}}}
272: % rescaled rho can also be \tilde{\rho}
273: \newcommand{\tz}{\ensuremath{\tilde{z}}}
274: % rescaled $z$, can also be \tilde{z}
275: \newcommand{\Rl}{\ensuremath{(R/l_s)^2}}% prefactor for Wilson loops
276: \newcommand{\tc}{\ensuremath{\sqrt{g_s \nc}}} %'t Hooft coupling
277: \newcommand{\stc}{\ensuremath{(g_sN)^{\frac{1}{4}}}}
278: % sqrt of 't Hooft coupling
279: \newcommand{\mq}{\ensuremath{m_\mt{q}}}      % Quark mass.
280: \newcommand{\eff}{\ensuremath{\mbox{\small eff.}}}
281: \newcommand{\calN}{\mbox{${\cal N}$}}
282: \newcommand{\cY}{\ensuremath{{\cal Y}}}
283: \newcommand{\vY}{\ensuremath{{\cal Y}^{\ell,\pm}}}
284: \newcommand{\ellh}{\ensuremath{\frac{\ell}{2}}}
285: \newcommand{\suLR}{\ensuremath{SU(2)_R\times SU(2)_L}}
286: \newcommand{\br}{\ensuremath{\bar{\rho}}}
287: \newcommand{\sun}{\ensuremath{{SU(N)}} }
288: \newcommand{\half}{\ensuremath{\frac{1}{2}}}
289: \newcommand{\mpi}{\ensuremath{M_{\pi}}}
290: \newcommand{\Lqcd}{\ensuremath{\Lambda_{\mt{QCD}}}}
291: \newcommand{\fphi}{\ensuremath{f_{\phi}}}
292: \newcommand{\cc}{\langle \bar{\psi} \psi \rangle}
293: \newcommand{\fpi}{f_\pi}
294: 
295: \def\lt {\lambda}
296: \def\rt {r}
297: \def\rhot {\rho}
298: \def \rvac{r_\mt{vac}}
299: \def\nc {N_\mt{c}}
300: \def\nf {N_\mt{f}}
301: \def\ua {U(1)_\mt{A}}
302: \def\t6 {T_\mt{D6}}
303: \def\ut {U_\mt{KK}}
304: \def\uh {U_\mt{T}}
305: \def\gym {g_\mt{YM}}
306: \newcommand{\nonsol}{D4-soliton}
307: \newcommand{\te}{t_\mt{E}}
308: \newcommand{\ct}{T_\mt{deconf}}
309: \newcommand{\tcc}{T_\mt{fund}}
310: \newcommand{\tb}{\bar{M}}
311: \newcommand{\msusy}{M_\mt{susy}}
312: \newcommand{\lm}{\l_\mt{match}}
313: 
314: 
315: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
316: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
317: %%%%%%%%%%%%%%%%%%%%%% TITLEPAGE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
318: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
319: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
320: 
321: 
322: \title{\LARGE Towards a holographic dual of large-$N_\mathrm{\large c}$ QCD}
323: 
324: \author{Mart\'\i n Kruczenski,$^{a}$
325:   David Mateos,$^{b}$ Robert C. Myers$\,^{b,c}$ and
326:   David J. Winters$\,^{b,d}$ \\
327:   $^a$ Department of Physics, Brandeis University \\
328:        Waltham, MA 02454, USA \\
329:   $^b$ Perimeter Institute for Theoretical Physics \\
330:        Waterloo, Ontario N2J 2W9, Canada \\
331:   $^c$ Department of Physics, University of Waterloo  \\
332:        Waterloo, Ontario N2L 3G1, Canada \\
333:   $^d$ Department of Physics, McGill University \\
334:        Montr\'eal, Qu\'ebec H3A 2T8, Canada
335: 
336: E-mail: \email{martink@brandeis.edu,
337:   dmateos@perimeterinstitute.ca, rmyers@perimeterinstitute.ca,
338:   winters@physics.mcgill.ca}}
339: 
340: \abstract{We study $\nf$ D6-brane probes in the supergravity
341:   background dual to $\nc$ D4-branes compactified on
342:   a circle with supersymmetry-breaking boundary conditions.
343:   In the limit in which the resulting Kaluza--Klein modes decouple,
344:   the gauge theory reduces to non-supersymmetric,
345:   four-dimensional QCD with $\nc$ colours and $\nf \ll \nc$ flavours.
346:   As expected, this decoupling is not fully realised within the
347:   supergravity/Born--Infeld approximation.
348:   For \mbox{$\nf=1$} and massless quarks, $\mq =0$,
349:   we exhibit spontaneous chiral symmetry breaking by a quark
350:   condensate, $\cc \neq 0$, and find the associated massless `pion' in the
351:   spectrum. The latter becomes massive for $\mq >0$, obeying the
352:   Gell-Mann--Oakes--Renner relation: $\mpi^2 = - \mq \cc/ \fpi^2$.
353:   In the case $\nf >1$ we provide a holographic version of the
354:   Vafa--Witten theorem, which states that the $U(\nf)$ flavour symmetry
355:   cannot be spontaneously broken. Further, we find $\nf^2-1$ unexpectedly light
356:   pseudo-scalar mesons in the spectrum. We argue that these are not
357:   (pseudo-)Goldstone bosons and speculate on the string mechanism
358:   responsible for their lightness. We then study the theory at finite
359:   temperature and exhibit a phase transition associated with a
360:   discontinuity in $\cc (T)$. D6/$\overline{\mbox{D6}}$ pairs are also
361:   briefly discussed.}
362: 
363: 
364: 
365: 
366: \keywords{D-branes, Supersymmetry and Duality}
367: 
368: %\preprint{BRX TH-258}
369: \preprint{BRX TH-258 \\ 
370: \tt{hep-th/0311270}}
371: 
372: \begin{document}
373: 
374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
375: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
376: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
377: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
379: \section{Introduction and summary of results} \label{intro}
380: 
381: The AdS$_5$/CFT$_4$ correspondence relates string theory on the
382: near-horizon background of $\nc$ D3-branes to the conformal field
383: theory living on their worldvolume \cite{Maldacena97}. Inspired by
384: this correspondence, Witten proposed a construction of the
385: holographic dual of four-dimensional, pure $SU(\nc)$ Yang--Mills
386: (YM) theory \cite{Witten98b}. One starts with $\nc$ D4-branes
387: compactified on a circle of radius $M_\mt{KK}^{-1}$, and further
388: imposes anti-periodic boundary conditions for the worldvolume
389: fermions on this circle. Before compactification, the D4-brane
390: theory is a five-dimensional, supersymmetric $SU(\nc)$ gauge
391: theory whose field content includes fermions and scalars in the 
392: adjoint representation of $SU(\nc)$, in addition to the gauge
393: fields. At energies much below the compactification scale,
394: $M_\mt{KK}$, the theory is effectively four-dimensional. The
395: anti-periodic boundary conditions break all of the supersymmetries
396: and give a tree-level mass to the fermions, while the scalars also
397: acquire a mass through one loop-effects. Thus at sufficiently low
398: energies the dynamics is that of four-dimensional, massless
399: gluons.
400: 
401: The D4-brane system above has a dual description in terms of
402: string theory in the near-horizon region of the associated (non-supersymmetric) 
403: supergravity background.
404: Unfortunately, as observed in \cite{Witten98b}, the Kaluza--Klein
405: (KK) modes on the D4-brane do not decouple within the supergravity
406: approximation. For example, the mass of the lightest glueball is
407: of the same order as the strong coupling scale. In this sense,
408: there is no energy region, $\Lambda_\mt{QCD} \ll E \ll M_\mt{KK}$,
409: in which a description in terms of weakly-coupled gluons is
410: appropriate. Nevertheless, the qualitative features of the
411: glueball spectrum agree with lattice calculations \cite{COOT98},
412: and the supergravity description suffices, for example, to exhibit
413: an area law for Wilson loops \cite{Witten98b}.
414: 
415: Since most of our phenomenological understanding of low-energy QCD
416: comes from the study of mesons and baryons, it is interesting to
417: construct a holographic dual of a gauge theory whose low-energy
418: degrees of freedom consist not only of gluons but also of
419: fundamental quarks. In the context of AdS/CFT, fundamental matter
420: can be added to the gauge theory by introducing D-brane probes in
421: the dual supergravity background
422: \cite{AFM98,KR01,BDFLM01,BDFM01,KK02, KKW02, WH03, Ouyang03}. This has
423: been recently exploited to study mesons holographically in several
424: examples of gauge/gravity duals \cite{KMMW03, SS03, BEEGK03, NPR03}. 
425: The goal of this paper is to study a gauge/gravity dual in which the
426: gauge theory reduces to non-supersymmetric, four-dimensional QCD
427: in the limit in which the KK modes decouple. The construction is
428: as follows.
429: 
430: Consider the D4/D6 system with the branes oriented as described
431: by the following array:
432: \be
433: \begin{array}{rccccccccccl}
434: \nc \,\, \mbox{D4:}\,\,\, & 0 & 1 & 2 & 3 & 4 & \_ & \_ & \_ & \_ & \_ & \, \\
435: \nf \,\, \mbox{D6:}\,\,\, & 0 & 1 & 2 & 3 & \_ & 5 & 6 & 7 & \_ & \_ & \, .
436: \ea
437: \label{intersection}
438: \ee
439: Note that the D4- and the D6-branes may be separated from each other
440: along the 89-directions. This system is T-dual to the D3/D5
441: intersection, and in the decoupling limit for the D4-branes
442: \cite{IMSY98} provides a non-conformal version of the AdS/dCFT
443: correspondence \cite{KR01, DFO01}. On the gauge theory side one has
444: a supersymmetric, five-dimensional $SU(\nc)$ gauge theory coupled to a
445: four-dimensional defect. The entire system is invariant under
446: eight supercharges, that is, $\caln =2$ supersymmetry in
447: four-dimensional language. The degrees of freedom localized on
448: the defect are $\nf$ hypermultiplets in the fundamental representation
449: of $SU(\nc)$, which arise from the open strings connecting the D4- and
450: the D6-branes. Each hypermultiplet consists of two Weyl fermions
451: of opposite chiralities, $\psi_\mt{L}$ and $\psi_\mt{R}$,
452: and two complex scalars.
453: 
454: As discussed above, identifying the 4-direction with period
455: $2\pi/M_\mt{KK}$, and with anti-periodic boundary conditions for
456: the D4-brane fermions, breaks all of the supersymmetries and
457: renders the theory effectively four-dimensional at energies $E \ll
458: M_\mt{KK}$. Further, the adjoint fermions and scalars become
459: massive. Now, the bare mass of each hypermultiplet, $\mq$, is
460: proportional to the distance between the corresponding D6-brane
461: and the D4-branes. Even if these bare masses are zero, we expect
462: loop effects to induce a mass for the scalars in the fundamental
463: representation. Generation of a mass for the fundamental fermions
464: is, however, forbidden by a chiral, $\ua$ symmetry that rotates
465: $\psi_\mt{L}$ and $\psi_\mt{R}$ with opposite
466: phases.\footnote{This symmetry is broken by instanton
467: effects, but these vanish in the large-$\nc$ limit in which
468: we work.} Therefore, at low energies, we expect to
469: be left with a four-dimensional $SU(\nc)$ gauge theory coupled to
470: $\nf$ flavours of fundamental quark.
471: 
472: In the dual string theory description, the D4-branes are again
473: replaced by their supergravity background. In the so-called `probe
474: limit', $\nf \ll \nc$, the backreaction of the D6-branes on this
475: background is negligible and hence they can be treated as probes.
476: The D6-brane worldvolume fields are dual to
477: gauge-invariant field theory operators constructed with at least two
478: hypermultiplet fields, that is, meson-like operators; of
479: particular importance here will be the quark bilinear operator,
480: $\bar{\psi} \psi \equiv \bar\psi_i \psi^i$, where 
481: $\psi^i = \psi_\mt{L}^i + \psi_\mt{R}^i$ and $i=1, \ldots , \nf$ 
482: is the flavour index. In
483: the string description the $\ua$ symmetry is nothing but the
484: rotation symmetry in the 89-plane. In addition to acting on the
485: fermions as explained above, this symmetry also acts on the
486: adjoint scalar $X=X^8 + i X^9$ by a phase rotation.
487: 
488: Having presented the general construction, we now summarise our main
489: results. 
490: 
491: We begin in section \ref{broken} by considering the case of a single 
492: D6-brane.  For $\mq=0$, we expect the chiral $\ua$ symmetry to be 
493: spontaneously broken by a chiral condensate, \mbox{$\cc \neq 0$}, and there 
494: to be an associated massless, pseudo-scalar Goldstone boson in the 
495: spectrum, as in QCD with one massless flavour. In QCD, this is the 
496: $\eta'$ (which becomes massless in the large-$\nc$ limit), but, in an 
497: abuse of language, we will refer to it as a `pion'. 
498: We are indeed able to show that the string description provides a
499: holographic realisation of this physics, by showing that the $\ua$ 
500: symmetry is spontaneously broken by the brane embedding, as in
501: \cite{BEEGK03}. Moreover, we numerically determine the chiral condensate 
502: for an arbitrary quark mass. We find \mbox{$\cc(\mq=0)\neq 0$}, as expected,
503: and $\cc(\mq) \propto 1/\mq$ for $\mq \ra \infty$, again in agreement 
504: with field theory expectations (which we briefly review).
505:  
506: In sections \ref{fluct} and \ref{pseudo} we study the meson spectrum. 
507: We first show analytically, in section \ref{fluct}, that at $\mq=0$ 
508: there is exactly one normalisable, massless pseudo-scalar 
509: (in the four-dimensional sense) fluctuation of the D6-brane. 
510: This open-string mode corresponds 
511: precisely to the Goldstone mode of the D6-brane embedding. If $\mq>0$, 
512: this pion becomes a pseudo-Goldstone boson. We are able to show, 
513: analytically, that its squared mass scales linearly with the quark
514: mass, $\mpi^2 \propto \mq$, in the limits of small (section
515: \ref{pseudo}) and large (section \ref{fluct}) quark mass. With 
516: numerical calculations we then confirm that, in fact, such a linear 
517: relation holds for all quark masses. 
518: For small $\mq$ this linear relation is in perfect agreement with the 
519: Gell-Mann--Oakes--Renner (GMOR) relation \cite{GMOR68},
520: \be 
521: \mpi^2 = - \fc{\mq \, \cc}{\fpi^2} \,,
522: \label{GMOR} 
523: \ee
524: which gives the first term in the expansion of the pion mass around 
525: $\mq=0$. In section \ref{pseudo}, we compute the chiral condensate and the 
526: pion decay constant analytically (at $\mq=0$), and with these results 
527: we are able to verify that, for small quark mass, the mass of our pion
528: precisely satisfies the GMOR relation. In the opposite limit, 
529: $\mq \ra \infty$, we show in section \ref{fluct} that 
530: \be
531: \mpi^2 = a \, \fc{\mq \mkk}{\gym^2 \nc} \,,
532: \label{linear} 
533: \ee
534: where $a$ is a pure number; for the lightest mesons, we find $a \sim 20$.
535: 
536: The remaining mesons (D6-brane excitations) are studied in section
537: \ref{fluct}. For $\mq \lesssim \msusy$, where 
538: \be
539: \msusy = \gym^2 \nc \mkk \,,
540: \label{msusy}
541: \ee 
542: they all have masses of order the compactification scale, $\mkk$, and 
543: so the (pseudo-)Goldstone boson dominates the infrared physics in the
544: regime of small $\mq$. Of course, this result signals the lack of
545: decoupling between the KK and the QCD scales within the 
546: supergravity/Born--Infeld approximation, as expected. For 
547: $\mq \gtrsim \msusy$, supersymmetry is approximately restored, 
548: the spectrum exhibits the corresponding degeneracy, and all meson
549: masses obey a formula like \eqn{linear} with appropriate values of
550: $a$. The fact that supersymmetry is restored at $\mq \sim \msusy$ 
551: suggests that the effective mass of some of the microscopic degrees 
552: of freedom must be at least of order $\msusy$. We close section 
553: \ref{fluct} with an examination of certain stability issues for the
554: D6-brane embeddings. 
555: 
556: In section \ref{vafa}, for the case of multiple flavours, $\nf >1$,
557: we provide a holographic version of the Vafa--Witten theorem
558: \cite{VW84}, which states that the $U(\nf)$ flavour symmetry
559: cannot be spontaneously broken if $\mq >0$. In the holographic
560: description this is realised by the fact that the $\nf$ D6-branes
561: must be coincident in order to minimise their energy.
562: 
563: In section \ref{fintem}, we examine the theory at finite
564: temperature. If $\mq> \msusy$, the theory exhibits two
565: phase transitions: the first one is the well-known
566: confinement/deconfinement transition at $T = \ct \sim \mkk$
567: \cite{Witten98b}, and the second one is a phase transition at 
568: \be
569: T = \tcc  \sim \sqrt{\fc{\mq \mkk}{\gym^2 \nc}} > \ct \,, 
570: \ee
571: characterised by a discontinuity in the chiral condensate, 
572: $\cc (T)$, and in the specific heat. If $\mq < \msusy$ 
573: the second transition does not occur as a separate phase transition. 
574: 
575: In section \ref{Dbar} we discuss six-brane embeddings 
576: that, from the brane-construction viewpoint, correspond not to
577: D4/D6 intersections but to intersections of D4-branes with
578: D6/$\overline{\mbox{D6}}$ pairs. From the gauge theory viewpoint these
579: configurations correspond to having a defect/anti-defect pair. 
580: 
581: We close with a brief discussion of our results in section
582: \ref{discus}. In particular, one result of the analysis of the
583: $\nf>1$ case in section \ref{vafa} is the existence of $\nf^2$
584: massless, pseudo-scalar mesons in the gauge theory spectrum if $\mq=0$.
585: (We expect the inclusion of sub-leading effects in the $1/\nc$
586: expansion to generate small masses for these particles.) Although
587: it is tempting to interpret these particles as Goldstone bosons of
588: a putative, spontaneously-broken $U(\nf)_\mt{A}$ chiral symmetry,
589: we argue that this interpretation cannot be correct: even at large
590: $\nc$, only one of them is a true Goldstone boson. 
591: Thus we are unable to find any obvious reasons, from the viewpoint of
592: the four- or five-dimensional gauge theory, for these
593: scalars to be light. We speculate that, from the string
594: viewpoint, ten-dimensional $U(\nf)$ gauge invariance (as opposed
595: to just seven-dimensional gauge invariance on the D6-branes) is
596: responsible for their lightness.
597: 
598: 
599: 
600: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
601: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
602: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
603: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
604: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
605: \section{Chiral symmetry breaking from D6-brane embeddings}
606: \label{broken}
607: 
608: In this section we will show how the spontaneous breaking of the
609: $\ua$ chiral symmetry expected from the field theory side is
610: realized in the string description.
611: 
612: 
613: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
614: \subsection{The D4-soliton background}
615: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
616: 
617: The type IIA supergravity background dual to $\nc$ D4-branes
618: compactified on a circle with anti-periodic boundary conditions
619: for the fermions takes the form
620: \begin{eqnarray}
621: ds^{2} &=& \left(\frac{U}{R}\right)^{3/2} \left( \eta_{\mu \nu} \,
622: dx^\mu dx^\nu + f(U) d\tau^{2} \right) + \left(
623: \frac{R}{U}\right)^{3/2} \frac{dU^{2}}{f(U)} +
624: R^{3/2} U^{1/2} \, d\Omega_{\it 4}^{2} \,, \label{metric} \\
625: e^{\phi} &=& g_s \left( \frac{U}{R}\right)^{3/4}
626: \sac F_{\it 4} = \frac{\nc}{V_{\it 4}} \, \varepsilon_{\it 4} \sac
627: f(U) = 1-\frac{\ut^{3}}{U^{3}} \,.
628: \label{metric1}
629: \end{eqnarray}
630: The coordinates $x^\mu=\{ x^0, \ldots , x^3\}$ parametrize the
631: four non-compact directions along the D4-branes, as in
632: \eqn{intersection}, whereas $\tau$ parametrizes the circular
633: 4-direction on which the branes are compactified. $d\Omega_{\it
634: 4}^2$ and $\varepsilon_{\it 4}$ are the $SO(5)$-invariant line
635: element and volume form on a unit four-sphere, respectively, and
636: $V_{\it 4}=8\pi^2/3$ is its volume. $U$ has dimensions of length
637: and may be thought of as a radial coordinate in the
638: 56789-directions transverse to the D4-branes. To avoid a conical
639: singularity at $U=\ut$, $\tau$ must be identified with period
640: \be 
641: \d \tau = \fc{4 \pi}{3} \, \fc{R^{3/2}}{\ut^{1/2}} \,.
642: \label{deltatau} 
643: \ee
644: Following the nomenclature of \cite{soliton}, we refer to this
645: solution as the \nonsol.
646: 
647: This supergravity solution above is regular everywhere and is
648: completely specified by the string coupling constant, $g_s$, the
649: Ramond--Ramond flux quantum (\ie, the number of D4-branes), $\nc$, and 
650: the constant $\ut$. The remaining
651: parameter, $R$, is given in terms of these quantities and the string
652: length, $\ell_s$, by
653: \be 
654: R^3 =  \pi g_s \nc\,\ell_s^3\,. \label{R} 
655: \ee
656: If $\ut$ is set to zero, the solution (\ref{metric},
657: \ref{metric1}) reduces to the extremal, 1/2-supersymmetric
658: D4-brane solution. Hence we may say that $\ut$ characterises the
659: deviation of the \nonsol\ from extremality.
660: 
661: The $SU(\nc)$ field theory dual to (\ref{metric}, \ref{metric1}) is
662: defined by the compactification scale, $\mkk$, below which the
663: theory is effectively four-dimensional, and the four-dimensional
664: coupling constant {\it at} the compactification scale, $\gym$. These are
665: related to the string parameters by
666: \be 
667: \mkk = \fc{3}{2}\fc{\ut^{1/2}}{R^{3/2}} =
668: \fc{3}{2 \sqrt{\pi}}\fc{\ut^{1/2}}{{( g_s \nc)}^{1/2} \ell_s^{3/2}} 
669: \sac \gym^2 =3 \sqrt{\pi} \left( \fc{g_s\ut}{\nc \ell_s} \right)^{1/2} \,.
670: \label{parameters} 
671: \ee
672: The first equation follows from the fact that $\tau$ is 
673: directly identified with the compact direction in the gauge theory,
674: so $\mkk \equiv {2\pi/\d\tau}$. The second equation follows from 
675: the fact that $g_{\it{5}}^2 = (2\pi)^2 g_s \ell_s$ and the relation between 
676: the four- and five-dimensional coupling 
677: constants, $\gym^2 = g_{\it{5}}^2 / \d \tau$, which can be seen  
678: by expanding the Born--Infeld action of the D4-brane and adopting 
679: the standard normalization for the gauge field kinetic term, 
680: $-F^2 / 4\,g_{\it{5}}^2$.
681: 
682: It is useful to invert these relations to express the string
683: parameters in terms of the gauge theory ones:
684: \be
685: R^3 = {1\over2} \fc{\gym^2 \nc \, \ell_s^2}{\mkk} \sac
686: g_s = {1\over2\pi} \fc{\gym^2}{\mkk \ell_s} \sac
687: \ut = {2\over9} \gym^2 \nc \, \mkk \ell_s^2 \,.
688: \label{inverse}
689: \ee
690: The string length will cancel in any calculation of a physical
691: quantity in the field theory. For example, the QCD string tension 
692: is\footnote{This is the tension of a string lying at $U=\ut$
693: and extending along $x$. It can also be computed by projecting such a
694: string onto the boundary and integrating $\langle F^2\rangle$ across
695: a transverse section of the string \cite{adsprop}.}
696: \be
697: \sigma =  \frac{1}{2\pi\ell_s^2} \left. \sqrt{-G_{tt} G_{xx}}
698: \right|_{U=\ut} = \frac{1}{2\pi \ell_s^2}
699: \left(\frac{\ut}{R}\right)^{3/2} = {2\over27\pi} \gym^2 \nc \, \mkk^2 \,.
700: \ee
701: Now we may ask in what situation the \nonsol\  provides a reliable
702: background in which one can study the dual field theory using
703: only classical supergravity. First, we must require that the curvature is
704: everywhere small compared to the fundamental string tension. This
705: ensures that higher derivative string corrections to the low
706: energy equations of motion are negligible. The maximum curvature
707: in the geometry given by \reef{metric} occurs at precisely $U=\ut$,
708: where the curvatures are of the order $(\ut R^3)^{-1/2}$. Hence we
709: require
710: \be
711: {\ut^{1/2} R^{3/2}\over\ell_s^2}  \simeq \gym^2 \nc  \gg 1\,,
712: \label{hoof}
713: \ee
714: where we have applied the results in eq.~\reef{inverse}. Therefore
715: the restriction to small curvatures corresponds to a
716: large 't Hooft coupling in the effective four-dimensional gauge
717: theory, precisely as in the conventional AdS/CFT correspondence
718: \cite{Maldacena97}. Further, to suppress string loop effects, we
719: must also require that the local string coupling, $e^\phi$, be small. Using
720: eqs.~\reef{metric1} and \eqn{inverse} we see that, for finite values 
721: of the gauge theory parameters, the inequality $e^\phi \ll 1$ can
722: only be satisfied up to some critical radius
723: \be
724: U_\mt{crit} \simeq \fc{\nc^{1/3} \mkk \ell_s^2}{\gym^2} \,.
725: \label{crit}
726: \ee
727: Beyond this radius, the M-theory circle opens up to
728: reveal an \adss{7}{4} background (with identifications).
729: Similarly, at the corresponding high-energy scales, the
730: five-dimensional Yang--Mills theory reveals a UV completion in
731: terms of the (2,0) theory compactified on a circle. In the present
732: context, we naturally demand that $U_\mt{crit}\gg\ut$, which,
733: using \eqn{crit} and \eqn{inverse}, reduces to
734: \be 
735: \gym^4\ll{1\over\gym^2\nc}\ll1\,, 
736: \label{small} 
737: \ee
738: where the second inequality follows from eq.~\eqn{hoof}. 
739: Eqs.~\eqn{hoof} and \eqn{small} imply that the supergravity analysis 
740: in the \nonsol\ background is reliable in precisely the
741: strong-coupling regime of the 't Hooft limit of the four-dimensional 
742: gauge theory: $\gym \ra 0$, $\nc \ra \infty$, $\gym^2 \nc$ fixed and
743: large.
744: 
745: 
746: 
747: 
748: 
749: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
750: \subsection{D6-brane embeddings}
751: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
752: 
753: We are now ready to study the embedding of a D6-brane probe in the
754: \nonsol\ geometry. Asymptotically (as $U\rightarrow \infty$), the
755: D6-brane is embedded as described by the array \eqn{intersection}.
756: The analysis is greatly simplified by introducing isotropic
757: coordinates in the 56789-directions. Towards this end, we first
758: define a new radial coordinate, $\rho$, related to $U$ by
759: \beq 
760: U(\rho) = \left(\rho^{3/2} +
761: \frac{\ut^3}{4\rho^{3/2}}\right)^{2/3} \,, 
762: \label{isomer}
763: \eeq
764: and then five coordinates $\vec{z}=(z^5, \ldots, z^9)$ such that
765: $\rho = |\vec{z}|$ and $d\vec{z} \cdot d\vec{z} = d\rho^2 + \rho^2
766: \, d\Omega_{\it 4}^2$. In terms of these coordinates the metric
767: \eqn{metric} becomes
768: \be
769: ds^{2} = \left(\frac{U}{R}\right)^{3/2}
770: \left( \eta_{\mu \nu} \, dx^\mu dx^\nu + f(U) d\tau^{2} \right) +
771: K(\rho) \, d\vec{z} \cdot d\vec{z} \,,
772: \ee
773: where
774: \be K(\rho) \equiv \fc{R^{3/2} U^{1/2}}{\rho^2}\,. 
775: \ee
776: Here $U$ is now thought of as a function of $\rho$. Finally, to
777: exploit the symmetries of the D6-brane embedding we seek, we
778: introduce spherical coordinates $\l, \Omega_{\it 2}$ for the
779: $z^{5,6,7}$-space and polar coordinates $r, \phi$ for the
780: $z^{8,9}$-space. The final form of the D4-brane metric is then
781: \be
782: ds^{2} = \left(\frac{U}{R}\right)^{3/2} \left( \eta_{\mu \nu} \,
783: dx^\mu dx^\nu + f(U) d\tau^{2} \right) + K(\rho) \, \left(
784: d\lambda^2 + \lambda^2 \, d\Omega_{\it 2}^2 + dr^2 + r^2 \,
785: d\phi^2 \right) \,, \label{isometric}
786: \ee
787: where $\rho^2 = \l^2 + r^2$.
788: 
789: In these coordinates the D6-brane embedding takes a particularly
790: simple form. We use $x^\mu$, $\l$ and $\Omega_{\it 2}$ (or $\s^a$,
791: $a=0,\ldots,6$, collectively) as worldvolume coordinates.  The
792: D6-brane's position in the 89-plane is specified as $r=r(\l)$, $\phi
793: = \phi_0$, where $\phi_0$ is a constant. Note that $\l$ is the
794: only variable on which $r$ is allowed to depend, by translational
795: and rotational symmetry in the 0123- and 567-directions,
796: respectively. We also set $\tau=\mbox{constant}$ in the following,
797: which corresponds to a single D6-brane localized in the circle
798: direction. We will study configurations in which this condition is
799: relaxed in section \ref{Dbar}.
800: 
801: With this ansatz for the embedding, the induced metric on the
802: D6-brane, $g_{ab}$, takes the form
803: \be 
804: ds^2(g) = \left(\frac{U}{R}\right)^{3/2} \, \eta_{\mu \nu} \,
805: dx^\mu dx^\nu + K(\rho) \, \left[ \left( 1+\dot{r}^2 \right)
806: d\lambda^2 + \lambda^2 \, d\Omega_{\it 2}^2 \right] \,,
807: \label{induced} 
808: \ee
809: where $\dot{r}\equiv\partial_\lambda r$. The D6-brane action
810: becomes
811: \be
812: S_{D6} = -\fc{1}{(2\pi)^6 \ell_s^7} 
813: \int \df^7\s\, e^{-\phi} \sqrt{-\det g} =
814: -\t6 \int \df^7\s\, \sqrt{h} 
815: \left( 1 + \fc{\ut^3}{4\rho^3} \right)^2 \,
816: \l^2 \sqrt{1 + \dot{r}^2} \,,
817: \label{eq:D6action}
818: \ee
819: where $T_\mt{D6}= 2\pi/g_s(2\pi \ell_s)^7$ is the six-brane
820: tension and $h$ is the determinant of the metric on the round unit
821: two-sphere. Recall that $\rho$ is a function of $\l$ both
822: explicitly and through its dependence on $r(\l)$. Hence the
823: equation of motion for $r(\l)$ is
824: \beq
825: \frac{d}{d\lambda} \left[\left(1+\frac{\ut^3}{4\rho^3}\right)^2
826:   \lambda^2 \frac{\dot{r}}{\sqrt{1+\dot{r}^2}} \right]
827: = -\frac{3}{2} \frac{\ut^3}{\rho^5}
828: \left(1+\frac{\ut^3}{4\rho^3}\right)
829: \lambda^2 \, r \, \sqrt{1+\dot{r}^2} \,.
830: \label{embedeq}
831: \eeq
832: Note that  $r(\l)=r_0$, where $r_0$ is a constant, is a solution in 
833: the supersymmetric limit ($\ut=0$), as in
834: \cite{KK02,KMMW03}. This reflects the BPS nature of the system,
835: which implies that there is no force on the D6-brane regardless
836: of its position in the 89-plane. In particular, then, the solution
837: with $r_0 =0$ preserves the $\ua$ rotational symmetry in the
838: 89-directions. If $\ut\neq 0$ the force on the D6-brane no longer
839: vanishes and causes it to bend as dictated by the equation of
840: motion above. We will see below that in this case there are no
841: (physical) solutions that preserve the $U(1)_\mt{A}$ symmetry.
842: 
843: If $\ut\neq 0$, the analysis is facilitated by rescaling to {\it
844: dimensionless} variables as follows:
845: \beq
846: \l \to \ut\l \sac r \to \ut r \sac \rho \to
847: \ut\rho \,, 
848: \label{tildes}
849: \eeq
850: in terms of which equation \eqn{embedeq} becomes
851: \beq
852: \frac{d}{d\lt} \left[\left(1+\frac{1}{4\rhot^3}\right)^2 \lt^2
853: \frac{\dot{\rt}}{\sqrt{1+\dot{\rt}^2}} \right] = -\frac{3}{2}
854: \frac{1}{\rhot^5} \left(1+\frac{1}{4\rhot^3}\right) \lt^2 \, \rt
855: \, \sqrt{1+\dot{\rt}^2} \,. 
856: \label{resc}
857: \eeq
858: We seek solutions such that the asymptotic separation of the
859: D6-brane and the D4-branes, $L$, is finite. That is, as
860: $\lt\rightarrow\infty$, $\rt(\lt) \rightarrow \rt_\infty$ with
861: \be
862: \rt_\infty = \fc{L}{\ut} \,.
863: \label{L}
864: \ee
865: In the region $\lt\rightarrow\infty$ we have $\dot{\rt}\rightarrow 0$
866: and $\rhot \simeq \lt$. Under these conditions we can linearize equation
867: \eqn{resc} with the result
868: \beq
869: \frac{d}{d\lt} \left[ \left. \lt \right.^2 \dot{\rt} \right]
870: \simeq -\frac{3}{2} \frac{1}{\left.\lt \right.^3}  \rt \,,
871: \eeq
872: the solution to which is
873: \beq
874: \rt(\lt) = A \, \frac{1}{\sqrt{\lt}} J_{-1/3} \left(
875: \frac{2^{1/2}3^{-1/2}}{\left. \lt \right.^{3/2}} \right) + B \,
876: \frac{1}{\sqrt{\lt}} J_{1/3} \left( \frac{2^{1/2}3^{-1/2}}{\left.
877: \lt \right.^{3/2}} \right) \,,
878: \eeq
879: where $J_{\nu}$ are Bessel functions and $A$ and $B$ are arbitrary
880: constants. For large $\lt$, the function multiplied by $A$ tends
881: to a constant, whereas that multiplied by $B$ behaves as
882: $1/\lt$. This means that for large $\l$ we have
883: \be
884: \rt(\l) \simeq \rt_\infty + \fc{c}{\lt} \,,
885: \label{rt}
886: \ee
887: where the constants $\rt_\infty$ and $c$ are related to the quark mass
888: and the chiral condensate, as we now show.
889: 
890: The bare quark mass is given by the string tension times the
891: {\it asymptotic} distance between the D4- and the 
892: D6-brane.\footnote{This mass is trivially derived for the 
893: brane array \reef{intersection} in asymptotically flat space. 
894: With supersymmetry ($\ut =0$), this bare mass persists in the 
895: decoupling limit and is then inherited by the compactified theory 
896: with supersymmetry-breaking boundary conditions, since setting 
897: $\ut \neq 0$ does not alter the asymptotic properties of the model
898: that, as usual in AdS/CFT-like dualities, determine the gauge theory
899: parameters.} 
900: Taking \eqn{L} into account, we have
901: \beq
902: \mq = \frac{L}{2\pi\ell_s^2} = \frac{\ut \, \rt_\infty}{2\pi\ell_s^2} \,.
903: \label{quarkmass}
904: \eeq
905: The quark condensate can now be computed using a simple argument.
906: The Hamiltonian density of the theory can be written as
907: \beq
908: \calh = \calh_0 + \mq \int \df^2\theta \, \tilde{Q}Q \,,
909: \eeq
910: where $\calh_0$ is $\mq$-independent, and $\tilde{Q}$, $Q$ represent
911: the hypermultiplet superfields in $\cN=1$ notation. It follows that
912: \beq
913: \frac{\delta \cale}{\delta \mq} =
914: %\frac{\delta}{\delta \mq} \langle \calh \rangle =
915: \langle  \int \df^2\theta \, \tilde{Q}Q \, \rangle = \cc \,,
916: \eeq
917: where $\cale = \langle \calh \rangle$ is the vacuum energy density. 
918: In the last equality, we have assumed that the vacuum expectation 
919: value of the fundamental scalars vanishes, since they are massive 
920: (even with $\mq=0$ after supersymmetry breaking), that is, the energy 
921: density increases if they acquire a non-zero expectation value.
922: 
923: Since in the string description the quark mass is given by the
924: asymptotic position of the D6-brane, we need to evaluate the change in
925: energy of the D6-brane associated to a change of the boundary
926: condition $\rt_\infty$. In view of (\ref{eq:D6action}) this is given by
927: \beq
928: \delta \cale = - \int \df\lambda \df\Omega_{\it 2} \, \delta \cL =
929: \t6 \ut^3 \int \df\lambda \df\Omega_{\it 2} \, \delta
930: \left[\sqrt{h} \left( 1 + \fc{1}{4\rho^3} \right)^2 \, \l^2 \sqrt{1 +
931: \dot{r}^2} \right] \,,
932: \eeq
933: where $\cL$ is the D6-brane Lagrangian density rescaled in accordance
934: with \reef{tildes}. Using the equations of motion this becomes
935: \beq
936: \delta \cale = -4\pi \left.\left(\d r\frac{\partial}{\partial\dot{r}}
937: \frac{\cL}{\sqrt{h}}\right)\right|^{\l=\infty}_{\l=0} =
938: 4\pi \t6 \ut^3\left. \left(1+\frac{1}{4\rho^3}\right)^2
939: \lambda^2\frac{\dot{r}\, \d r}{\sqrt{1+\dot{r}^2}}
940: \right|^{\l=\infty}_{\l=0} =
941: - 4\pi \t6 \ut^3 \, c\, \delta r_\infty  \,,
942: \eeq
943: where we used \eqn{rt} and the fact that $\dot{r}|_{\l =0}=0$,
944: by rotational symmetry in the 567-directions. In view of
945: \eqn{quarkmass} we finally obtain
946: \beq
947: \cc = \frac{\delta \cale}{\delta m_\mt{q}} =
948: -8 \pi^2 \ell_s^2 \t6 \ut^2 \, c \,.
949: \label{quarkcond}
950: \eeq
951: As anticipated, the quark condensate is directly related to
952: $c$; we will see below that $c$ is determined by $\rt_\infty$.
953: Note that, in terms of gauge theory quantities,
954: \be \fc{1}{\nc} \cc \simeq \gym^2 \nc \mkk^3 c\,. 
955: \ee
956: The normalization on the left-hand side is that expected in the 't
957: Hooft limit.
958: 
959: We now return to the full equation of motion (\ref{resc}), which we
960: will solve numerically. In order to understand the nature of the 
961: solutions better, however, it is
962: helpful to first consider the limit $\rt_\infty \gg 1$, in which
963: (approximate) analytical solutions can be found.
964: 
965: If $\rt_\infty \gg 1$, the entire D6-brane should lie far
966: away from the `bolt' at the center of the \nonsol, hence we 
967: expect the solution to be a small perturbation around the solution
968: for $\ut=0$. In other words, we set $\rt(\lt)=\rt_\infty + \d \rt
969: (\lt)$ and assume that $\d \rt \ll 1$. Substituting this into
970: \eqn{resc} and expanding to linear order in $\d \rt$ we find
971: \beq
972: \frac{d}{d\lt}\left(\left. \lt \right. ^2 \delta\dot{\rt}\right) \simeq
973: -\frac{3}{2} \lt^2 \frac{\rt_\infty}{\left( \left. \lt \right.^2+ \left.
974: \rt \right._\infty^2 \right)^{5/2}} \,,
975: \eeq
976: which is easily integrated with the result
977: \beq
978: \d \dot{\rt} \simeq -\frac{3}{2}\frac{\rt_\infty}{\left. \lt
979: \right.^2} \int_0^{\lt} \df x\, \frac{x^2}
980: {\left( x^2+ \left. \rt \right._\infty^2 \right)^{5/2}} =
981: - \frac{1}{2 \rt_\infty} \, \frac{\lt}
982: {\left( \left. \lt \right. ^2 +
983: \left. \rt \right._\infty^2 \right)^{3/2}} \,.
984: \eeq
985: Here we have imposed the boundary condition $\dot{\rt} = 0$ at
986: $\lt=0$, as required by regularity of the solution at the origin of
987: the 567-space. Integrating once more, with the boundary condition
988: $\d \rt|_{\lt=\infty}=0$,  we find
989: \beq
990: \rt(\lt) \simeq \rt_\infty + \frac{1}{2 \rt_\infty} \,
991: \frac{1}{\sqrt{\left. \lt \right. ^2 + \left. \rt \right._\infty^2}} \,.
992: \label{asymptotic}
993: \eeq
994: From this large-$\rt_\infty$ solution, we see  that the
995: D6-brane bends `outwards', that is, it is `repelled' by the
996: D4-branes. This repulsion persists for arbitrary values
997: of $\rt_\infty$, as is confirmed by numerical
998: analysis. Eq.~\eqn{resc} can be integrated numerically for
999: any value of $\rt_\infty$, and we have plotted solutions for several
1000: values of $\rt_\infty$ in Figure \ref{fig:D6embedding}.
1001: 
1002: \FIGURE{\epsfig{file=profile.eps, height=10cm}
1003: \caption{The D6-brane embedding for several values of the quark
1004:   mass $\mq \propto \rt_\infty$. The interior of the `circle' in the center is
1005:   the region $U<\ut$, which is actually not part of the space.
1006:   The `2' and `3' embeddings
1007:   correspond to the same value of $\rt_\infty$ but have opposite-sign
1008:   $c$'s; the `3' embedding has lower energy.}
1009: \label{fig:D6embedding} }
1010: 
1011: Figure \ref{condensate} displays the function $c(\rt_\infty)$
1012: found by numerical integrations. As mentioned above, the value of
1013: $\rt_\infty$ together with the requirement of regularity at
1014: $\lt=0$ (\ie, $\dot{\rt}|_{\lt=0}=0$) determines the solution
1015: completely. Hence choosing $\rt_\infty$ fixes $c=c(\rt_\infty)$.
1016: Recalling eqs.~\reef{quarkmass} and \reef{quarkcond}, which relate
1017: $\rt_\infty$ and $c$ to the quark mass and condensate,
1018: respectively, we see that this corresponds precisely to what is
1019: expected on field-theoretic grounds: once the quark mass is specified, 
1020: the infrared dynamics determines the chiral condensate.  In the field
1021: theory we expect the $\ua$ symmetry to be spontaneously broken by a
1022: non-zero condensate in the limit $\mq \ra 0$. This is indeed confirmed
1023: by the numerical results for the D6-brane embedding, since we
1024: see that, in figure \ref{condensate}, $c(\rt_\infty)$ approaches a
1025: finite constant in the limit $\rt_\infty \ra 0$. In this limit,
1026: the asymptotic boundary condition on the D6-brane respects the
1027: rotational symmetry in the 89-plane, but it is energetically favourable
1028: for the D6-brane to bend out in the 89-plane. 
1029: Hence the embedding spontaneously breaks this rotational symmetry, 
1030: which matches the $\ua$ symmetry-breaking in the
1031: field theory.
1032: 
1033: We see from \eqn{asymptotic} that $c(\rt_\infty) \sim 1/ (2\rt_\infty)$
1034: for large $\rt_\infty$, which is confirmed by the numerical
1035: results. This implies that the chiral condensate scales as $1/\mq$ for
1036: large $\mq$, as expected on field theory grounds \cite{SVZ79}.
1037: A quick, somewhat heuristic argument for this in QCD is as
1038: follows.\footnote{We thank D.\ T.\ Son for explaining this point to us.}   
1039: The trace of the energy-momentum tensor reads 
1040: \be
1041: T^\mu_{\,\,\,\, \mu} = \mq \bar{\psi} \psi -
1042: \fc{\alpha_s (11 \nc -2 \nf)}{24 \pi} \, \mbox{Tr} F^2 \,,
1043: \ee
1044: where we have explicitly written down the classical contribution, as
1045: well as the anomalous one coming from the one-loop beta-function. 
1046: In the limit $\mq \ra \infty$ the theory resembles QCD with no quarks,
1047: for which 
1048: \be
1049: T^\mu_{\,\,\,\, \mu} = -\fc{11 \nc \alpha_s}{24 \pi} \,\mbox{Tr} F^2  \,.
1050: \ee
1051: Taking vacuum expectation values and equating the two expressions we find 
1052: \be
1053: \cc = \fc{\alpha_s \nf}{12 \pi \mq} \, \langle \mbox{Tr} F^2 \rangle \,.
1054: \ee
1055: Assuming that $\langle \mbox{Tr} F^2 \rangle \neq 0$ we conclude that 
1056: $\cc \propto 1/\mq$.
1057: 
1058: For each value of $\rt_\infty>0$, the numerical analysis actually
1059: finds two regular solutions, which have $c$'s of opposite sign. 
1060: The curves labelled `2' and `3' in figure
1061: \ref{fig:D6embedding} are a representative example of such a pair
1062: of solutions.\footnote{ To describe the second solution (curve 2),
1063: we allow $\rt(\lt)<0$ and restrict $\phi$ to the range $[0,\pi)$.}
1064: We have confirmed that the solutions with positive $c$, for which
1065: $\rt(\lt)>0$ for all $\lt$ (\eg, curve 3), have the lower energy
1066: density. The negative-$c$ solutions bend around the opposite side of
1067: the bolt at the center of the \nonsol\ geometry. Naively, these
1068: solutions should be unstable, since the D6-brane is free to `slide
1069: off the bolt' by moving out of the $\phi=\phi_0$ plane, and relax to the 
1070: lower-energy, positive-$c$ solution. We will
1071: return to this issue in the next section, where we study
1072: fluctuations of the D6-brane worldvolume fields around embedding
1073: solutions of each type --- in fact, we will find tachyonic
1074: fluctuations for the negative-$c$ solutions, confirming the
1075: intuition given above.
1076: 
1077: \FIGURE{\epsfig{file=condensate2.eps, height=10cm}
1078:  \caption{The quark condensate
1079:    $\langle \bar{\psi} \psi \rangle \propto c$ as a function of the
1080:    quark mass $\mq \propto \rt_\infty$.}
1081: \label{condensate}}
1082: 
1083: Note that $r(\l)=0$ is an exact mathematical solution of
1084: eq.~\eqn{embedeq} even if $\ut\neq 0$. Therefore one might be
1085: tempted to conclude that it provides an alternative,
1086: $\ua$-preserving solution whose energy should be shown to be
1087: higher than that of the $\ua$-breaking solution above in order to
1088: establish spontaneous chiral symmetry breaking. However, the
1089: solution $r(\l)=0$ is not physically acceptable if $\ut \neq 0$
1090: because it corresponds to an open D6-brane. That is, this solution
1091: yields a D6-brane with a boundary, which is forbidden by charge
1092: conservation. The boundary is easily seen as follows. With
1093: $r(\l)=0$ we have $\rho =\l$, so this solution terminates at the origin
1094: of the $U\tau$-plane, that is, at $U=\ut$. Since, at this
1095: point, the $S^2$ in the metric \eqn{induced} still has a finite
1096: radius, $R^{3/4}\ut^{1/4}$, the D6-brane would have an 
1097: $\bbr{1,3} \times S^2$-boundary if it terminated at $U=\ut$. Note that if
1098: $\ut=0$ then the two-sphere radius shrinks to zero-size and the
1099: six-brane closes off at $U=0$. Hence the $\ua$-preserving
1100: solution $r(\l)=0$ is physically sensible in the supersymmetric
1101: case. This result is in agreement with the fact that unbroken
1102: supersymmetry forbids a non-zero chiral condensate. Note, however,
1103: that this is only a formal argument, since in the limit $\ut \ra 0$ the
1104: curvature at $U=\ut$ diverges, meaning that the supergravity description
1105: cannot be trusted in this region \cite{IMSY98}.
1106: 
1107: In section \ref{Dbar}, we will see that the solution $r(\l)=0$ can
1108: play a role in the case $\ut \neq 0$ if it is simply extended past
1109: the origin of the $U\tau$-plane. The physical interpretation of
1110: such a solution is not that of a D4/D6 intersection, but rather
1111: that of a D4/D6/$\overline{\mbox{D6}}$ intersection.
1112: 
1113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1115: %%%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1118: \section{Meson spectroscopy from D6-brane fluctuations ($\nf =1$)} 
1119: \label{fluct}
1120: 
1121: Here, we consider a certain class of fluctuations of the
1122: D6-brane around the embeddings described in the previous section.
1123: We consider first the positive-$c$ (negative-condensate)
1124: embeddings, which were argued to be stable and so should correspond
1125: to the true `vacuum state' for a given quark mass. For simplicity
1126: we restrict ourselves to fluctuations of the fields $r$ and
1127: $\phi$, as these are sufficient to illustrate the physics we wish
1128: to exhibit. We are therefore considering embeddings of the D6-brane 
1129: of the form
1130: \beq 
1131: \phi=0+\d\phi \sac r=\rvac(\l)+\d r \sac \tau=\mbox{constant}
1132: \,, \label{fluctansatz} 
1133: \eeq
1134: where the fluctuations $\d \phi$ and $\d r$ are functions of 
1135: {\it all} of the worldvolume coordinates and $\rvac(\l)$ is the
1136: numerically-determined vacuum embedding of the previous section.
1137: 
1138: In corroboration of our previous argument for the stability of
1139: these embeddings, we find, in particular, that the spectra of
1140: these fluctuations are non-tachyonic. We also briefly consider
1141: fluctuations around the negative-$c$ (positive-condensate)
1142: embeddings and show their spectra do contain tachyonic modes,
1143: making manifest the instability of these configurations.
1144: 
1145: In the dual gauge theory, the $r$- and the $\phi$-fluctuations
1146: correspond to a class of scalar and pseudo-scalar mesons, 
1147: respectively. To see this, consider the complex scalar
1148: field parametrising the 89-plane, $X = X_8 + i X_9 = r e^{i\phi}$. 
1149: Gauge theory Lorentz transformations act only on the
1150: 0123-directions, under which $X$ is inert, so $X$-fluctuations 
1151: clearly correspond to spin-zero mesons. However, a parity 
1152: transformation in the gauge theory corresponds to a ten-dimensional
1153: transformation that not only reverses the sign of (say) $X_3$, 
1154: but also replaces $X$ by its complex conjugate, $\bar{X}$; this
1155: leaves $r$ invariant but reverses the sign of $\phi$. 
1156: From the gauge theory viewpoint, this is clear from the
1157: presence in the microscopic Lagrangian of a coupling of the form 
1158: $\psi_\mt{R}^\dagger X \psi_\mt{L} + 
1159: \psi_\mt{L}^\dagger \bar{X} \psi_\mt{R}$: the transformation 
1160: $X_3 \ra -X_3$ exchanges $\psi_\mt{R}$ and 
1161: $\psi_\mt{L}$, so this term is only invariant if, simultaneously,
1162: $X \ra \bar{X}$. From the string theory viewpoint, the fact that only
1163: the combination of both transformations is a symmetry can be seen 
1164: by examining the fermionic spectrum of the D4-D6 open strings.
1165: This arises from the zero modes of the world-sheet fermions 
1166: in the Ramond sector, that is, in the NN and DD directions: 
1167: the 23-  and 89-directions (in the light-cone gauge) in 
1168: \eqn{intersection}. Each of these modes is labelled by its weight
1169: under rotations in each of the 23- and 89-planes, $(s_1, s_2), s_i=\pm 1/2$. 
1170: The GSO projection requires $s_1=-s_2$. Since under $X_3 \ra -X_3$ we
1171: have $(s_1, s_2) \ra (-s_1, s_2)$, and under $X_9 \ra -X_9$ we have 
1172: $(s_1, s_2) \ra (s_1, -s_2)$, we see that the spectrum is only
1173: invariant under the combination of both transformations.
1174: 
1175: 
1176: 
1177: 
1178: 
1179: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1180: \subsection{Analysis of the spectrum}
1181: \label{anal}
1182: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1183: 
1184: Using the background metric \eqn{isometric} with the rescalings
1185: \eqn{tildes}, the pullback of \eqn{isometric} to the embedding
1186: given in \eqn{fluctansatz} is
1187: %
1188: \beqa
1189: ds^2&=&\left(\frac{U}{R}\right)^{3/2}\eta_{\mu\nu}dx^\mu dx^\nu+
1190: K[(1+\dot{r}_\mt{vac}^2)d\l^2 + \l^2d\O_{\it 2}^2+
1191: 2\dot{r}_\mt{vac}\pf_a(\d r) d\l dx^a]\non\\
1192: %
1193: &+&K[\pf_a(\d r) \pf_b(\d r) dx^a dx^b+ (r_\mt{vac}+\d
1194: r)^2\pf_a(\d \phi) \pf_b(\d \phi) dx^a dx^b]\ , \label{indfluc}
1195: \eeqa
1196: %
1197: where $a$ and $b$ run over {\it all} of the worldvolume
1198: directions. To quadratic order in fluctuations, the D6-brane
1199: Lagrangian density is then
1200: %
1201: \beqa \call =\call_0&-&\t6 \ut^3 \l^2 \sqrt{h}
1202: \sqrt{1+\dot{r}_\mt{vac}^2}
1203: \Bigg\{\left(\frac{3(7\r_\mt{vac}^2-\l^2)}{16\rho_\mt{vac}^{10}}+
1204: \frac{3(4r_\mt{vac}^2-\l^2)}{4\rho_\mt{vac}^7}\right)(\d r)^2\non\\
1205: %
1206: &&\hspace{2cm}-\frac{3r_\mt{vac}}{4\rho_\mt{vac}^5}\left(1+
1207: \frac{1}{4\rho_\mt{vac}^3}\right) \frac{\dot{r}_\mt{vac} \pf_\l(\d
1208: r^2)}{1+\dot{r}_\mt{vac}^2}\non\\
1209: &&+\left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\, \sum_a
1210: \frac{K}{2g_{aa}} \left(\frac{[\pf_a(\d
1211: r)]^2}{1+\dot{r}_\mt{vac}^2}+ r_\mt{vac}^2[\pf_a(\d
1212: \phi)]^2\right)\Bigg\}\ , \label{expandlag} \eeqa
1213: %
1214: where $\call_0$ is the Lagrangian density evaluated for the vacuum
1215: embedding, as in eq.~\eqn{eq:D6action}. Here,
1216: $\rho_\mt{vac}^2=\l^2+r_\mt{vac}^2$ and the $g_{aa}$ in the last
1217: line correspond to the metric coefficients from eq.~\eqn{induced}.
1218: In particular then, they contain no dependence on the
1219: fluctuations. Of course, integration by parts and the equation of
1220: motion \eqn{embedeq} for $r_\mt{vac}$ allowed the terms linear in
1221: $\d r$ to be eliminated.
1222: 
1223: The linearised equations of motion that follow are, for $\d \phi$:
1224: \beqa
1225: &&\fc{9}{4\mkk^2} \frac{\rvac^2}{\rho_\mt{vac}^3}
1226: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^{4/3}\pf_\mu \pf^\mu (\d\phi)
1227: +\frac{\rvac^2}{\l^2}
1228: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\nabla^2 (\d\phi) \non\\
1229: &&\hspace{2cm}+\frac{1}{\l^2\sqrt{1+\dot{r}_\mt{vac}^2}}
1230: \frac{d}{d\l}\left[\frac{\l^2 \rvac^2}{\sqrt{1+\dot{r}_\mt{vac}^2}}
1231: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\pf_\l (\d\phi)\right]=0 \,,
1232: \label{phi}
1233: \eeqa
1234: and, for $\d r$:
1235: \beqa
1236: &&\frac{\l^2}{\sqrt{1+\dot{r}_\mt{vac}^2}}
1237: \left[\fc{9}{4\mkk^2} \frac{1}{\rho_\mt{vac}^3}
1238: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^{4/3}\pf_\mu \pf^\mu (\d r)
1239: +\frac{1}{\l^2}\left(
1240: 1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\nabla^2 (\d r)\right] \non\\
1241: && \hspace{1cm}+\frac{d}{d\l}\left[\frac{\l^2}{(1+\dot{r}_\mt{vac}^2)^{3/2}}
1242: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\pf_\l(\d r)\right]
1243: -\frac{d}{d\l}\left[\frac{3\l^2\rvac\dot{r}_\mt{vac}}
1244: {2\rho_\mt{vac}^5\sqrt{1+\dot{r}_\mt{vac}^2}}
1245: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)\right] \d r \non\\
1246: &&\hspace{4cm}-\l^2\sqrt{1+\dot{r}_\mt{vac}^2}\left(\frac{3(7\rvac^2-\l^2)}
1247: {8\rho_\mt{vac}^{10}}+\frac{3(4\rvac^2-\l^2)}{2\rho_\mt{vac}^7}\right)\d r =0 \,.
1248: \label{r}
1249: \eeqa
1250: %
1251: In these expressions, $\nabla^2$ is the Laplacian on the
1252: two-sphere.
1253: 
1254: From the form of the equations of motion, it is clear that we may
1255: separate variables to write
1256: \beq 
1257: \d \phi=\calp(\l) e^{i k_\phi \cdot x}\, Y_{\ell_\phi
1258: m_\phi}(S^2) \sac \d r=\calr(\l) e^{i k_r \cdot x}\, Y_{\ell_r
1259: m_r}(S^2) \,, 
1260: \eeq
1261: where $Y_{\ell m}$ are the conventional spherical harmonics. We
1262: then look for normalizable solutions with definite $S^2$-angular
1263: momentum $\ell_{\phi,r}$ and four-dimensional mass
1264: $M_{\phi,r}^2=-k_{\phi,r}^2$. For the sake of simplicity in the
1265: following, we only focus on the case $\ell_{\phi,r}=0$ and find
1266: the lowest-mass modes for each fluctuation.
1267: 
1268: We employ the shooting method to solve the linearized equations
1269: of motion. First we use the asymptotic form of the equations  for
1270: $\l\to\infty$ and the requirement of normalizability to determine
1271: the asymptotic boundary conditions for the fluctuations. Then we
1272: vary the four-dimensional mass parameters until, by numerically
1273: integrating the equation of motion, we find solutions that are
1274: also regular at the origin. Recall\footnote{Also recall that we
1275: are working with dimensionless variables $r$ and $\l$, rescaled as
1276: in eq.~\eqn{tildes}.} that the asymptotic behaviour of the vacuum
1277: profile is $\rvac\simeq r_\infty + c/\l$, where $r_\infty$ is
1278: related to the quark mass by equation \eqn{quarkmass}. In this
1279: asymptotic region we have that $\rho_\mt{vac}(\l) \simeq\l$, so
1280: the equation of motion for $\d\phi$ becomes, approximately,
1281: \beq
1282: \frac{9M_\phi^2}{4\mkk^2}\frac{\rvac^2}{\l}(\d\phi)+
1283: \frac{d}{d\l}\left[ \l^2\rvac^2\pf_\l(\d\phi)\right] = 0\,,
1284: \eeq
1285: and similarly that for $\d r$ becomes
1286: \beq
1287: \frac{9M_r^2}{4\mkk^2}\frac{1}{\l}(\d r)+\frac{d}{d\l}
1288: \left[ \l^2 \pf_\l(\d r)\right] = 0\,.
1289: \eeq
1290: It follows
1291: that the asymptotic behaviour of $\d r(\l)$ does not depend on
1292: that of $\rvac(\l)$, whereas that of $\d \phi(\l)$ depends
1293: crucially on whether or not $r_\infty=0$, that is, whether or not
1294: the quarks are massless. We will now show that if $r_\infty=0$
1295: then there exists a normalizable $\d \phi$ mode with $M_\phi=0$,
1296: whereas if $r_\infty \neq 0$ there is a mass gap in the spectrum.
1297: 
1298: \FIGURE{\epsfig{file=spectrum2.eps, height=10cm} 
1299: \caption{Squared masses for the four lowest-lying $\d\phi$ modes 
1300: (dashed lines), dual to pseudo-scalar mesons, and the four 
1301: lowest-lying $\d r$ modes (solid lines), dual to scalar mesons, 
1302: as functions of the quark mass $\mq \propto r_\infty$.}  
1303: \label{squaredscalarmasses} }
1304: 
1305: If we set $r_\infty=0$, we find that the asymptotic equation for
1306: $\d \phi$ becomes
1307: \beq
1308: \frac{9M_\phi^2}{4\mkk^2}\frac{1}{\l^3}(\d\phi)+\pf_\l^2(\d\phi) = 0 \,.
1309: \eeq
1310: The first term is sub-leading for large $\l$, so we may drop it to
1311: find that the asymptotic form of $\d \phi$, regardless of
1312: the value of $M_\phi$, is
1313: \beq
1314: \d\phi\simeq b + a \l \,,
1315: \label{m=0}
1316: \eeq
1317: where $a$ and $b$ are arbitrary constants. For $r_\infty\ne 0$,
1318: however, we find that
1319: \beq
1320: \frac{9M_\phi^2}{4\mkk^2}\frac{1}{\l^3}(\d\phi)+\frac{2}{\l} \pf_\l (\d\phi) +
1321: \pf_\l^2(\d\phi) = 0\,.
1322: \eeq
1323: Again the first term is sub-leading for large $\l$, so the solution is
1324: \beq
1325: \d\phi\simeq \tilde{a}+\tilde{b}/\l \,.
1326: \label{mne0}
1327: \eeq
1328: Normalizable solutions correspond to the boundary conditions
1329: $a=0$ and $\tilde{a}=0$ in the cases $r_\infty=0$ and $r_\infty \neq 0$,
1330: respectively, since the terms with coefficients $b$ and $\tilde{b}$
1331: are the ones that
1332: fall off more rapidly at infinity. Since the differential equations we
1333: are solving are linear, we may choose $b=\tilde{b}=1$ without loss of
1334: generality, so we are left with the boundary conditions
1335: $\{\d\phi=1 \ ,\ \d\phi'=0\}|_{\l\to\infty}$ if $r_\infty=0$, and
1336: $\{\d\phi=1/\l\ ,\  \d\phi'=-1/\l^2\}|_{\l\to\infty}$ if $r_\infty\neq 0$.
1337: A similar analysis for the $\d r$ fluctuations shows that they obey
1338: the same asymptotic equation of motion as the $\d\phi$ fluctuations
1339: in the $r_\infty\ne 0$ case, so the appropriate boundary conditions are
1340: $\{\d r=1/\l\ ,\ \d r'=-1/\l^2\}|_{\l\to \infty}$.
1341: 
1342: \FIGURE{\epsfig{file=linearanddegenerate.eps, height=7cm}
1343: \caption{(a) Linear fit of the lowest-lying $\d\phi$ mode 
1344:   for different values of $\mq \sim \rt_\infty$.
1345:   (b) Degeneracy of the lowest-lying $\d\phi$ (dashed) and $\d r$
1346:   (solid) modes in the supersymmetric limit $r_\infty\gg 1$.}
1347: \label{linearanddegenerate} }
1348: 
1349: Now note that $\d \phi = e^{i k_\phi \cdot x}$, with $k_\phi^2=0$,
1350: is an exact, regular solution of \reef{phi} with zero
1351: four-dimensional mass (and $\nabla^2(\d\phi)=0$), which is
1352: normalisable {\it only} if $r_\infty=0$. This means that with
1353: massless quarks there is a massless, pseudo-scalar meson in the 
1354: spectrum of the
1355: gauge theory, whose dual is the zero-mode fluctuation of the
1356: D6-brane field $\phi$. The mathematical form of these
1357: fluctuations shows that they correspond to rotations of the vacuum
1358: embedding in the 89-plane. Of course, this form is precisely as
1359: expected for the Goldstone mode associated with the spontaneous
1360: breaking of this rotational symmetry.
1361: 
1362: The shooting technique allows us to verify this numerically, and
1363: to find the values of $M_{\phi,r}$ for which other
1364: normalizable, regular solutions exist. The results are the
1365: low-lying (pseudo)scalar meson spectra, which are displayed in Figure
1366: \ref{squaredscalarmasses} for different values of the quark mass.
1367: As anticipated, there is no mass gap in the $\phi$-meson spectrum
1368: when the quark is massless. There is a smooth transition to a
1369: theory with a mass gap as the quark mass is increased and, in
1370: fact, a best-fit (shown in figure \ref{linearanddegenerate}a) 
1371: yields a linear relationship,\footnote{ 
1372: An analogous relationship was observed in \cite{BEEGK03}.} 
1373: $M_\phi^2 \simeq 0.73\, \mkk^2\, r_\infty$, for
1374: small $r_\infty\propto m_\mt{q}$, as expected from the GMOR
1375: relationship \eqn{GMOR}, which we will reproduce analytically 
1376: in the next section.
1377: 
1378: As shown in figure \ref{squaredscalarmasses}, there are no
1379: normalisable, regular solutions of equation \eqn{r} with
1380: $M_{r} =0$, so there is a mass gap in the $r$-meson spectrum
1381: regardless of the value of the quark mass. For small quark mass
1382: the mass scale of all of the mesons, except the lowest-lying $\d
1383: \phi$ mode, is\footnote{This also sets the scale of the glueball
1384: mass spectrum \cite{COOT98}.}
1385: \be
1386: M^2 \sim \mkk^2 \sim \frac{U_\mt{KK}}{R^3}\,.
1387: \label{region1}
1388: \ee
1389: This is another reflection of the lack of decoupling between the QCD
1390: scale and the compactification scale.
1391: 
1392: As suggested by Figure \ref{squaredscalarmasses}, the $r$- and
1393: $\phi$-mesons occur in pairs that become degenerate for large enough
1394: a quark mass. The reason is that, in the supersymmetric case
1395: ($\ut=0$), these mesons belong to the same supermultiplet and hence have
1396: equal masses, as in \cite{KMMW03}. If $\ut \neq 0$ this degeneracy is
1397: only approximate for quark masses much larger than the
1398: supersymmetry-breaking scale, $\msusy$, becoming exact in the limit
1399: $\mq \ra \infty$. Geometrically, this corresponds to the fact
1400: that the distortion of the D6-brane away from
1401: the flat, supersymmetric profile decreases the farther away from the
1402: D4-branes we embed it. Since this distance scale is measured by
1403: $L \propto m_\mt{q}$, and $\ut$ characterises the supersymmetry
1404: breaking, we expect that, for $L \gg \ut$, the characteristic
1405: mass scale of the mesons should become independent of $\ut$ and
1406: depend, instead, on some ratio of $L$ and $R$. To see that this is
1407: indeed the case, we approximate the vacuum profile by the flat
1408: solution in equations \reef{phi} and \reef{r}, that is, we set
1409: $\rvac(\l) = r_\infty$, and we expand these equations assuming that
1410: $r_\infty \gg 1$. In terms of a new variable $y= \l / L$ we obtain
1411: (for zero $S^2$-angular momentum)
1412: \beq
1413: \frac{R^3 M_\phi^2}{L (y^2+1)^{3/2}} \d \phi +
1414: \frac{1}{y^2}\frac{d}{dy}[y^2\pf_y\d\phi] = 0 \ ,
1415: \label{phitrans}
1416: \eeq
1417: and
1418: \beq
1419: \frac{R^3 M_r^2}{L (y^2+1)^{3/2}} \d r +
1420: \frac{1}{y^2}\frac{d}{dy}[y^2\pf_y\d r] = 0 \,.
1421: \label{rtrans}
1422: \eeq
1423: The fact that the two types of fluctuations satisfy the same equation 
1424: confirms that the spectrum becomes degenerate in this limit. It is
1425: also clear that the mass scale for the mesons in this limit is now
1426: \be
1427: M^2 \sim \fc{L}{R^3} \sim \fc{\mq \mkk}{\gym^2 \nc} \,,
1428: \label{region2}
1429: \ee
1430: since this is the only scale that appears in eq. \eqn{phitrans},
1431: \eqn{rtrans}.
1432: 
1433: The transition between the regimes \eqn{region1} and \eqn{region2}
1434: occurs at $L \sim \ut$, and is illustrated for the
1435: lightest $r$- and $\phi$-mesons in figure \ref{linearanddegenerate}b, 
1436: which was generated by solving equations \reef{phi} and \reef{r} 
1437: numerically. The figure clearly illustrates the degeneracy of the
1438: modes in the expected regime. The condition $L \sim \ut$
1439: translates into $\mq \sim \msusy$, with $\msusy$ as defined in
1440: \eqn{msusy}, so $\msusy$ is indeed the scale at which supersymmetry
1441: is restored, as suggested by our choice of notation. Note that at the 
1442: transition point, at which of course eqs. \eqn{region1} and \eqn{region2}
1443: coincide, we have $M \sim \mq / (\gym^2 \nc)$. As we are working in 
1444: the 't Hooft limit of the gauge theory, these mesons are very deeply bound, 
1445: \ie, $M \ll 2\mq$. The suppression of these meson masses by the 
1446: 't Hooft coupling is even stronger than that found in a similar 
1447: context in \cite{KMMW03}, where $M \sim \mq / \sqrt{\gym^2 \nc}$.
1448: For large quark mass there is an extra suppression implicit in
1449: eq. \eqn{region2} due to the fact that $M^2$ scales linearly in $\mq$,
1450: as opposed to quadratically. 
1451: 
1452: As we are working in a range where we expect the gauge theory is
1453: five-dimensional, it is perhaps more appropriate to re-express
1454: the transition scale in terms of the five-dimensional gauge coupling;
1455: it then reads $\mq \sim g_{\it 5}^2\nc \mkk^2$. Similarly, the mass scale
1456: typical of the mesons becomes $M^2 \sim \mq/(g_{\it 5}^2\nc)$. In any
1457: event, the transition scale suggests that the effective mass squared of
1458: some of the microscopic degrees of freedom must be at least of the
1459: order $\msusy \gg \mkk$, rather than just 
1460: $\mkk$. Note that this mass scale is a factor of $\sqrt{\gym^2 \nc}$
1461: larger than that coming from a simple one-loop calculation for the
1462: scalar masses, which yields $\sqrt{\gym^2 \nc} \mkk$. 
1463: This factor seems to be ubiquitous in extrapolations from weak to
1464: strong 't Hooft coupling in the context of AdS/CFT.
1465: 
1466: As we saw above, the squared masses of all
1467: mesons scale linearly in the quark mass as $\mq \ra \infty$. 
1468: For the lightest meson, the results displayed in figure 
1469: \ref{linearanddegenerate}b seem to indicate that, in fact, this linear 
1470: relationship is valid not only in the limits of small and large $\mq$,
1471: but for all quark masses. This result has been 
1472: confirmed to within the best accuracy of our numerical analysis. 
1473: Unfortunately, eq.~\eqn{phitrans} has four regular singular points 
1474: and so no closed-form analytic solution is available, but its
1475: numerical solution gives, for the lowest-mass modes,
1476: \beq
1477: M^2 \simeq 1.66\, \frac{L}{R^3} \simeq 
1478: 20.91\, \frac{\mkk}{\gym^2\nc} \mq \simeq 0.74\, \mkk^2 r_\infty\ .
1479: \eeq 
1480: The final form of this expression agrees with a best-fit of the data plotted
1481: in the region of figure \ref{linearanddegenerate}b with $r_\infty\gg 1$ and, 
1482: in principle, describes the extension of this figure to arbitrarily high values
1483: of $r_\infty$. We therefore conclude that, in our model, the GMOR
1484: linear relation between $M_\pi^2$ and $\mq$ extrapolates to all quark masses.
1485:  
1486: 
1487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1488: \subsection{Stability Issues}
1489: \label{stable}
1490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1491: 
1492: Recall that, at the end of section \ref{broken}, we remarked that
1493: our numerical integrations had revealed two classes of six-brane
1494: embeddings, those with positive $c$ and those with negative $c$.
1495: From the numerical analysis, we could show that, for a given
1496: $\rt_\infty$, the positive-$c$ or negative-condensate solutions
1497: had the lower energy density. Further we argued that the
1498: negative-$c$ solutions should be unstable.
1499: 
1500: The stability of the negative-condensate (positive-$c$) vacuum embedding is
1501: reflected in the fact that the spectrum of fluctuations does not
1502: contain tachyonic modes. While numerical searches did not reveal
1503: any tachyonic fluctuations, one can formulate a general argument
1504: that no such modes can exist in the spectrum of
1505: $\d\phi$.\footnote{We expect that these arguments extend to $\d r$
1506: fluctuations but we did not explicitly consider these modes a
1507: possible decay channel between the two D6-brane embeddings, due to
1508: the geometric obstruction provided by the `bolt' at the centre of
1509: the \nonsol.} Consider the equation of motion \eqn{phi} (with zero
1510: angular momentum on the two-sphere, for simplicity). Multiplying
1511: by $\d\phi$, integrating over $\l$ and integrating by parts,
1512: yields
1513: %
1514: \beqa &&\int_0^\infty\df \l\,
1515: \left\{\frac{9M_\phi^2}{4\mkk^2}\frac{\l^2\rvac^2}{\rho_\mt{vac}^3}
1516: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^{4/3}
1517: \sqrt{1+\dot{r}_\mt{vac}^2}\, \d\phi^2
1518: -\frac{\l^2\rvac^2}{\sqrt{1+\dot{r}_\mt{vac}^2}}
1519: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2
1520: [\pf_\l(\d\phi)]^2\right\}\non\\
1521: &&\hspace{4cm}=\left[\frac{\l^2\rvac^2}{\sqrt{1+\dot{r}_\mt{vac}^2}}
1522: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2\d\phi\,
1523: \pf_\l(\d\phi)\right]_0^\infty=0\ . \label{integrostability} \eeqa
1524: %
1525: Note that we were able to set the term on the right-hand side to
1526: zero using the behaviour of the normalizable $\d\phi$ modes at
1527: $\l=0$ and $\l=\infty$. On the left-hand side, the second term in
1528: the integrand is manifestly negative or zero, while the sign of
1529: the first term depends on that of $M_\phi^2$. Therefore, for
1530: tachyonic modes it is clear that the integral will be negative-definite,  
1531: which is inconsistent with the vanishing of the
1532: right-hand side. The argument extends to include
1533: nonvanishing $\ell_\phi$, which simply introduces an additional
1534: negative-definite term under the integral. Hence we conclude that
1535: the negative-condensate embeddings are free from tachyonic
1536: instabilities in this sector. As this was the sector with the
1537: lowest lying mode, we might have expected it to be the most
1538: potentially problematic with regards to stability. Hence we are
1539: confident that these positive-$c$ solutions are stable.
1540: 
1541: Now we turn our attention to the negative-$c$ or
1542: positive-condensate embeddings. Recall that our intuition was that
1543: these solutions should be unstable to `sliding off the bolt.'
1544: Since the embedding coordinate ($r$, here) always vanishes at some
1545: value of $\l$ for this case, we cannot address the issue of such
1546: an instability using the $\d\phi$ equation of motion, since $\phi$
1547: is not everywhere well-defined. 
1548: Hence, we change to Cartesian
1549: coordinates $X$ and $Y$ in the $r\phi$-plane. With the choice that
1550: the D6-brane lies at $\phi_\mt{vac}(\l)=\phi_0=0$ (\ie,
1551: $Y_\mt{vac}(\l)=0$), it follows that the embedding profile in $X$
1552: satisfies the equation of motion \reef{resc} and we may set
1553: $X_\mt{vac}(\l)=\rvac(\l)$. Fluctuations in $Y(\l)$ are now the
1554: modes relevant for stability and have the benefit of being
1555: well-defined for all $\l$. One finds the following equation of
1556: motion for these fluctuations:
1557: \beqa
1558: &&\frac{9}{4\mkk^2}\frac{1}{\rho_\mt{vac}^3}
1559: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^{4/3}
1560: \pf_\mu\pf^\mu(\d Y)
1561: +\frac{1}{\l^2}\left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2
1562: \nabla^2(\d Y)\non \\
1563: &&+\frac{1}{\l^2\sqrt{1+\dot{X}_\mt{vac}^2}}\frac{d}{d\l}
1564: \left[\frac{\l^2}{\sqrt{1+\dot{X}_\mt{vac}^2}}
1565: \left(1+\frac{1}{4\rho_\mt{vac}^2}\right)^2\pf_\l(\d Y)\right]+
1566: \frac{3(1+4\rho_\mt{vac}^3)}{8\rho_\mt{vac}^8}\d Y=0\ ,
1567: \label{Y}
1568: \eeqa
1569: where, now, $\rho_\mt{vac}^2=\l^2+X_\mt{vac}^2$. Making a
1570: stability argument along the lines of that given above for
1571: $\d\phi$ is no longer possible, due to the presence of the final
1572: term in \reef{Y}. This term introduces into the integrand
1573: analogous to that in eq.~\reef{integrostability} a contribution that is
1574: manifestly positive. Hence this approach does not yield a definite
1575: conclusion in this case.
1576: \FIGURE{{\epsfig{file=lowestmodes2.eps, height=7cm}}
1577: \caption{The lowest-lying $\d Y$ mode for some negative-condensate
1578: embeddings (upper line), and the lowest-lying $\d Y$ mode for the
1579: corresponding positive-condensate embeddings (lower line) --- the
1580: $\d Y$ modes are tachyonic.}
1581: \label{tachfig}}
1582: 
1583: However, it is still possible to look directly for tachyonic modes
1584: in the spectrum of fluctuations around the negative-$c$ embeddings
1585: using our numerical techniques. From eq.~\reef{Y} one can deduce
1586: the asymptotic behaviour of the $\d Y$ modes to be
1587: \beq 
1588: \d Y\simeq \tilde{a}+\tilde{b}/\l\ , 
1589: \eeq
1590: independent of the vacuum embedding. Using the shooting method as
1591: before, the spectrum of each positive-condensate embedding
1592: considered numerically is found to contain a tachyonic mode and,
1593: furthermore, these modes become `more tachyonic' as the quark mass
1594: increases --- see figure \ref{tachfig}. As is suggested in the
1595: figure, the squared mass of the lowest-lying $\d Y$ mode has a
1596: linear dependence on the quark mass. In fact, a best-fit of the
1597: data gives $M_Y^2\simeq -0.72\, \mkk^2 r_\infty$.
1598: 
1599: An examination of the next-to-lowest-lying $\d Y$ mode (not shown
1600: in the figure) indicates that the mass of this mode, which is real
1601: and positive for $m_\mt{q}=0$, also decreases as a function of
1602: $r_\infty$. So for sufficiently large quark mass the spectrum may
1603: contain more than one tachyon. However, we could not confirm this
1604: directly because for large $r_\infty$ the negative-$c$ embeddings
1605: approach extremely close to the bolt, and our computer code
1606: behaved erratically in this situation.
1607: 
1608: 
1609: 
1610: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1612: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1613: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1614: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1615: \section{The pseudo-Goldstone boson} \label{pseudo}
1616: 
1617: As discussed in the previous section, if $\mq=0$ the spectrum contains
1618: a massless, pseudo-scalar mode that can be understood as the Goldstone boson
1619: of the spontaneously broken $\ua$ symmetry. This is akin to the
1620: $\eta'$ meson in QCD, which is massless in the large-$\nc$ limit. In an
1621: abuse of language, however, we will refer to this Goldstone boson as a
1622: `pion'. For non-zero quark mass, the $\ua$ symmetry is explicitly broken
1623: and the pion becomes a pseudo-Goldstone boson with a mass $\mpi$
1624: that, according to our numerical analysis, grows as $\mpi^2 \sim \mq$.
1625: This means that at low energies, $E\ll \Lqcd$, the dynamics is
1626: dominated by this particle. In this section we will provide an
1627: analytic proof, using the string description, that the pion's mass
1628: actually obeys the well-known Gell-Mann--Oakes--Renner relation
1629: \eqn{GMOR}.
1630: 
1631: Recall that the gauge theory pion is dual to a $\phi$-fluctuation
1632: of the D6-brane of the form $\delta \phi = e^{ik \cdot x} \vp (\l)$
1633: that solves equation \eqn{phi} with $\nabla^2 (\d \phi)=0$, that is,
1634: \beq
1635: \frac{d}{d\l} \left[ p(\lambda)\dot{\vp} \right] =
1636: -M^2 \mu(\lambda) \vp\ ,
1637: \label{eq:eqvp}
1638: \eeq
1639: where $M^2 = -k^2$ is the four-dimensional mass,
1640: \beqa
1641: p(\lambda) &=& \lambda^2 \rvac^2
1642: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^2
1643: \frac{1}{\sqrt{1+\dot{r}_{\mt{vac}}^2}} \,,
1644: \label{pdef}
1645: \\ \mu(\lambda) &=& \lambda^2 \frac{9}{4\mkk^2}\frac{\rvac^2}{\rho_\mt{vac}^3}
1646: \left(1+\frac{1}{4\rho_\mt{vac}^3}\right)^{4/3}
1647: \sqrt{1+\dot{r}_{\mt{vac}}^2} \,,
1648: \label{pmu}
1649: \eeqa
1650: and $\rho_\mt{vac}^2 = \lambda^2 +\rvac^2$. 
1651: The allowed solutions
1652: are those that both satisfy $\dot{\vp}|_{\l =0} = 0$
1653: and are normalizable with respect to the scalar product
1654: \beq
1655: \langle \vp_1 |  \vp_2 \rangle = \int_0^\infty \df\l\, \mu(\lambda)
1656: \vp^*_1(\lambda) \vp_2(\lambda) \,.
1657: \eeq
1658: This defines an eigenvalue problem whose solution is
1659: a complete set of orthogonal
1660: eigenfunctions $\vp_n$ with eigenvalues $M^2_n$. Notice that
1661: if $\vp(\l)= \mbox{constant}$ then the eigenvalue equation
1662: is satisfied with $M^2=0$, but the mode is normalizable only if
1663: $\rvac(\l) \ra 0$ as $\lambda\rightarrow\infty$.
1664: If $\vp(\l)$ is not constant, multiplying eq.(\ref{eq:eqvp}) by $\vp$,
1665: integrating both sides with respect to $\l$ and performing an integral by
1666: parts in the left-hand side, we arrive at
1667: \beq
1668: M^2 = \frac{\int_0^{\infty} \df\l\, p(\l) \dot{\vp}^2}
1669: {\int_0^{\infty} \df\l\, \mu(\l) \vp^2} >  0\ .
1670: \eeq
1671: Since from (\ref{pdef}) and (\ref{pmu}) we have $p(\l)>0$ and
1672: $\mu(\l)>0$, it follows that $M^2$ is positive, as indicated.\footnote{
1673: This is not valid when $\rvac$ vanishes at some point (which
1674: is not the case studied here) --- see the more detailed discussion
1675: after equation (\ref{integrostability}).}
1676: 
1677: Having said that, we now want to compute the mode with lowest
1678: eigenvalue $M^2>0$ when $m_\mt{q}$ is non-vanishing but small, that
1679: is, let us suppose that $\rvac\rightarrow r_\infty$ as
1680: $\lambda\rightarrow\infty$, with $r_\infty$ small but non-zero. In
1681: this case we expect that the normalizable eigenmode can be obtained
1682: from the massless-quark zero-mode, namely $\vp(\l)=\mbox{constant}$,
1683: by using perturbation theory. To this end, it is convenient to define
1684: \beq
1685: \psi(\lambda) = \sqrt{p(\lambda)} \, \vp (\l) \,,
1686: \eeq
1687: which satisfies the new eigenvalue equation
1688: \beq
1689: \ddot{\psi} - \frac{\ddot\Psi}{\Psi} \psi = -M^2\, \nu \, \psi \,,
1690: \label{psieqn}
1691: \eeq
1692: where $\Psi=\sqrt{p}$, and $\nu=\mu/p$ is the new integration
1693: measure. If $r_\infty=0$ then $\psi(\l)=\Psi(\l)$ is a normalizable
1694: solution of \eqn{psieqn} with zero eigenvalue. Making a small change in
1695: the boundary condition, $r_\infty\to\d r_\infty\gtrsim 0$, induces a small
1696: change in the vacuum profile $\rvac(\l)$,
1697: which in turn induces a small change in the functions $\Psi(\l)$ and
1698: $\nu(\l)$ appearing in equation \eqn{psieqn}. Standard
1699: quantum mechanics perturbation theory can then be applied to obtain the
1700: new lowest eigenvalue.
1701: 
1702: Let us denote with a `bar' the quantities corresponding to
1703: $r_\infty=0$, so that for $r_\infty\neq 0$ we have
1704: %
1705:  \beqa
1706: \Psi &=& \bar{\Psi} + \delta \Psi \,, \\
1707: \nu &=& \bar{\nu} + \delta \nu \,,
1708: \eeqa
1709: %
1710: where $\delta \Psi$ and $\delta \nu$ are the differences induced
1711: by the change in $\rvac(\l)$. The new lowest eigenvalue and its
1712: associated wave-function, $M^2 \gtrsim 0$ and
1713: $\psi =\bar{\Psi} + \delta \psi$, obey, to leading order,
1714: \beq
1715: \delta \ddot\psi -\frac{\ddot{\bar{\Psi}}}{\bar{\Psi}} \, \delta \psi
1716: -\delta\left(\frac{\ddot\Psi}{\Psi}\right) \bar{\Psi} =
1717: - M^2 \, \bar{\nu} \, \bar{\Psi} \,.
1718: \eeq
1719: Decomposing $\delta \psi $ as a linear combination
1720: of massless-quark eigenfunctions,
1721: $\delta\psi = \sum_{n=0}^\infty \alpha_n \bar{\psi}_n $, we obtain
1722: \beq
1723: \sum_{n=1}^\infty \alpha_n \bar{M}^2_n \bar{\nu} \bar{\psi}_n +
1724: \delta\left(\frac{\ddot\Psi}{\Psi}\right) \bar{\Psi} =
1725: M^2 \, \bar{\nu} \, \bar{\Psi} \,,
1726: \eeq
1727: %
1728: where we have dropped the first term in the series, using
1729: $\bar{M}_0=0$. Multiplying by $\bar{\Psi}$, integrating over
1730: $\lambda$ and using the orthogonality condition $\int_0^{\infty}
1731: \df\l\, \bar\nu \bar{\Psi} \bar{\psi}_n =0$ for $n>0$ (since
1732: $\bar{\Psi}$ is the zero-mode eigenfunction and is real) we deduce
1733: that
1734: %
1735: \beq
1736: M^2 \int_0^{\infty} \df\l \, \bar{\nu}\, \bar{\Psi}^2 =
1737: \int_0^\infty \df\l \, 
1738: \delta\left(\frac{\ddot\Psi}{\Psi}\right) \bar{\Psi}^2 =
1739: \int_0^\infty \df\l \, 
1740: (\bar{\Psi}\delta\ddot{\Psi} -\ddot{\bar{\Psi}}\delta\Psi) =
1741: \left[ \bar{\Psi} \delta \dot\Psi - \dot{\bar{\Psi}} \delta \Psi
1742: \right]_{\l=0}^{\l=\infty} \,.
1743: \label{m2}
1744: \eeq
1745: Recalling that
1746: \be
1747: \bar{\Psi} = \lambda \bar{r}_\mt{vac}
1748: \left(1+\frac{1}{4\bar{\rho}_\mt{vac}^3}\right)
1749: (1+\dot{\bar{r}}_\mt{vac}^2)^{-1/4} \,,
1750: \ee
1751: which implies
1752: \bea
1753: \delta \Psi &=& \lambda \, \delta r_\mt{vac}
1754: \left(1+\frac{1}{4\bar{\rho}_\mt{vac}^3}\right)
1755: (1+\dot{\bar{r}}_\mt{vac}^2)^{-1/4} -
1756: \frac{3\lambda \bar{r}^2_{\mt{vac}}}{4\bar{\rho}_\mt{vac}^5}
1757: (1+\dot{\bar{r}}_\mt{vac}^2)^{-1/4} \, \delta \rvac - \nonumber \\
1758: && - \l\, \bar{r}_\mt{vac} \left(1+\frac{1}{4\bar{\rho}_\mt{vac}^3}\right)
1759: \frac{1}{2}(1+\dot{\bar{r}}_\mt{vac}^2)^{-5/4} \dot{\bar{r}}_\mt{vac} \,
1760: \delta\dot{r}_\mt{vac} \,,
1761: \eea
1762: we see that the only contribution to the last term in equation
1763: \eqn{m2} comes from $\l=\infty$. This is easily evaluated, since for
1764: large $\l$ we have
1765: \be
1766: \bar{r}_\mt{vac} \simeq \frac{\bar{c}}{\lambda} +
1767: \calo \left( \frac{1}{\lambda^2}\right) \sac
1768: \bar{\Psi} \simeq \l\bar{r}_\mt{vac} \simeq \bar{c} +
1769: \calo \left( \frac{1}{\l}\right) \,.
1770: \ee
1771: It follows that the new eigenvalue is given by
1772: \beq
1773: M^2 = \frac{\bar{c}\, \d r_\infty}
1774: {\int_0^\infty \df\l\, \bar{\nu} \bar{\Psi}^2} =
1775: \frac{\bar{c} \, \d r_\infty}{\int_0^\infty \df\l\, \bar{\mu}} \,.
1776: \label{GMORsugra}
1777: \eeq
1778: To show that this formula is precisely the GMOR relation
1779: \eqn{GMOR}, we need to identify the pion decay constant in the
1780: string description. In this description, a shift of the chiral
1781: angle that parametrizes the space of vacua related to each other
1782: by the $\ua$ symmetry corresponds to a rigid rotation of the
1783: D6-brane field $\phi$ (both take values between 0 and $2\pi$).
1784: Similarly, the pion field is identified with the normalizable,
1785: massless mode $\d\phi$, described in section \ref{anal}. The pion
1786: decay constant can be read off from the normalization of the
1787: kinetic term in the four-dimensional low-energy effective
1788: Lagrangian for this mode. That is, we integrate the $\d\phi$-terms
1789: in eq.~\eqn{expandlag} over $\l$ and the coordinates on $S^2$ and
1790: as a result find a four-dimensional action of the form
1791: %
1792: \be S =  -\fc{\fpi^2}{2}\int \df^4x\, \pf_\mu (\d{\phi})
1793: \pf^\mu(\d{\phi})\ , \ee
1794: %
1795: where $f_\pi$ is the pion decay constant. From
1796: eq.~\reef{expandlag}, then we have
1797: %
1798:  \beqa
1799:  \fpi^2 &=& \t6 \ut^3
1800: \int\df\Omega_{\it 2}\, \sqrt{h} \int_0^\infty \df\l\,
1801: \l^2\frac{9}{4\mkk^2}\frac{\bar{r}^2_\mt{vac}}{\bar{\rho}^3_\mt{vac}}
1802: \left(1+\frac{1}{4\bar{\rho}^3_\mt{vac}}\right)^{4/3}
1803: \sqrt{1+\dot{\bar{r}}^2_\mt{vac}}\non\\
1804: & =& 4\pi \t6 \ut^3 \int_0^\infty
1805: \df\l \, \bar{\mu}(\l)\ . \label{fpi} 
1806: \eeqa
1807: %
1808: Combining this result with eqs.~\eqn{quarkmass} and
1809: \eqn{quarkcond}, we see that eq.~\eqn{GMORsugra} is precisely the
1810: GMOR relation \eqn{GMOR}, as anticipated.
1811: 
1812: We close this section by verifying the agreement between the
1813: analytical expression \eqn{GMORsugra} above and the results obtained 
1814: purely by numerical methods and displayed in figure 
1815: \ref{linearanddegenerate}a. The integral in equation \eqn{GMORsugra} 
1816: can be evaluated numerically with the result
1817: \beq
1818: \int_0^\infty \df\l \, \bar{\mu}(\l) = \frac{9}{4\mkk^2} \int_0^\infty \df\l \,
1819: \frac{\lambda^2\bar{r}_\mt{vac}^2}{\bar{\rho}_\mt{vac}^3}
1820: \left(1+\frac{1}{4\bar{\rho}_\mt{vac}^3}\right)^{4/3}
1821: \sqrt{1+\dot{\bar{r}}_\mt{vac}^2} \simeq \frac{0.97}{\mkk^2} \,.
1822: \label{muint}
1823: \eeq
1824: Similarly, the numerical result for the constant $c$ for zero
1825: quark mass (see Figure \ref{condensate}) is $\bar{c} \simeq 0.71$.
1826: Substituting these two values in equation \eqn{GMORsugra} we get
1827: $M^2 \simeq 0.73\, \mkk^2 \d\rt_\infty$, in exact agreement with a best-fit
1828: of the points in figure \ref{linearanddegenerate}a.
1829: 
1830: Recall from section \ref{broken} that for nonvanishing quark mass
1831: we found that there are two possible values of the quark
1832: condensate, differing in sign, but that only the negative condensate is
1833: stable. In section \ref{stable} the positive-condensate solution
1834: was explicitly shown to be unstable by showing that the
1835: pseudo-Goldstone mode $\d Y$ was tachyonic for this configuration.
1836: Since the field-theoretic arguments establishing the GMOR relation
1837: apply regardless of the sign of the quark condensate, the squared 
1838: mass of this mode should also obey the GMOR relation, at least
1839: approximately, with an appropriate sign change. Recall that 
1840: eq. \reef{GMOR} represents the first term in an expansion
1841: around $\mq=0$. Now in the string description at $\mq=0$, the 
1842: D6-brane configurations for positive and negative $c$ are
1843: identical, except for a rotation of $\phi$ by $\pi$. Hence up to a
1844: crucial sign, the numerical coefficients in eq.~\reef{GMORsugra}
1845: are unchanged and hence the tachyonic mode should satisfy
1846: $M_Y^2 \simeq -0.73\, \mkk^2 \d r_\infty$. From Figure
1847: \ref{tachfig}, it does indeed appear that the slopes of the lowest-lying 
1848: $\d\phi$ and $\d Y$ modes are equal, but opposite in sign. More precisely, 
1849: our best-fit of the numerical results for negative $c$ gave 
1850: $M_Y^2\simeq -0.72\, \mkk^2 r_\infty$, which gives agreement to within 
1851: about 1.4\%.
1852: 
1853: 
1854: 
1855: 
1856: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1857: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1858: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1859: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1860: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1861: \section{Multiple flavours ($\nf >1$) and a holographic Vafa--Witten theorem}
1862: \label{vafa}
1863: 
1864: Consider a vector-like gauge theory with vanishing $\theta$-angle
1865: (like QCD) and $\nf$ flavours, all with identical mass $\mq>0$.
1866: The Vafa--Witten theorem states that the vector-like $U(\nf)$
1867: flavour symmetry cannot be spontaneously broken \cite{VW84}. This
1868: theorem does not exclude the possibility that, at $\mq =0$, the
1869: $U(\nf)$-preserving vacuum becomes degenerate with one or more
1870: $U(\nf)$-breaking vacua. In this section we discuss the
1871: holographic realization of the Vafa--Witten theorem, within the
1872: approximations implied by the validity of the
1873: supergravity/Born--Infeld theory. In particular, we are
1874: considering $\nc \ra \infty$ and $\nf \ll \nc$. Note that,
1875: strictly speaking, the Vafa--Witten theorem need not apply to the
1876: gauge theory on the D4/D6-brane system because of the presence of
1877: Yukawa-like couplings between the fundamental fermions and the
1878: adjoint scalars. However, the latter scalar fields are massive and
1879: the theorem should certainly apply in a limit where the scalar
1880: masses are sent to infinity. Further, one expects the same result
1881: for large but finite masses since the vacuum should still not be
1882: sensitive to the scalars. In the present case, the scalar masses
1883: are likely of order $\mkk \sim \Lqcd$ and so these couplings would
1884: not be suppressed. Hence, from the gauge theory viewpoint, 
1885: it is not clear a priori whether the effect of these fields can 
1886: change the vacuum structure to the extent of violating the Vafa--Witten
1887: theorem.\footnote{We thank A.\ Nelson for a discussion on this
1888: point.} The results of this section answer this question
1889: in the negative.
1890: 
1891: The $U(\nf)$ global symmetry of the gauge theory corresponds, in
1892: the string description, to the $U(\nf)$ gauge symmetry on the
1893: worldvolume of $\nf$ D6-brane probes in the geometry \eqn{metric}.
1894: Their dynamics is described by the so-called non-Abelian
1895: Dirac--Born--Infeld (DBI) action \cite{Myers99}. This is a
1896: generalisation of the single-brane Abelian theory
1897: \eqn{eq:D6action}, in which both the gauge fields and the
1898: transverse scalars are promoted to Hermitian, $U(\nf)$ matrices.
1899: This means, in particular, that the positions of the D6-branes are
1900: no longer commuting quantities in general. In the particular cases
1901: in which the transverse scalar matrices can be simultaneously
1902: diagonalised, each of their $\nf$ eigenvalues can be interpreted
1903: as the position of one of the D6-branes along the corresponding
1904: transverse direction.
1905: 
1906: For the $U(\nf)$ global symmetry to be realized in the gauge
1907: theory, the mass matrix for the fundamental quarks must be
1908: diagonal with all eigenvalues equal to $\mq$. In our string
1909: picture, this translates into the boundary condition that,
1910: asymptotically, all the D6-branes should have the same asymptotic
1911: boundary conditions. That is, they are all aligned with each other
1912: in the ($z^5,\ldots,z^9$) subspace and all lie at a distance 
1913: $2\pi \ell_s^2 \mq$ from the origin. We would like to show that
1914: the configuration (which obviously satisfies the boundary
1915: condition above) in which all the D6-branes are exactly
1916: coincident, and are embedded exactly as in the case of a single
1917: D6-brane, is the minimum-energy solution of the equations of
1918: motion that follow from the non-Abelian DBI action.
1919: 
1920: To avoid problems with the coordinate singularity at $r=0$, we
1921: adopt the Cartesian coordinates $X$ and $Y$ introduced in section
1922: \ref{stable} to describe the transverse positions of the D6-branes
1923: in the $r\phi$-plane. Then the configuration for the $\nf$
1924: coincident six-branes may be written as:
1925: \be 
1926: X(\l) = \rvac (\l) \cdot \bbi{} \sac Y(\l) = 0 \sac \tau =
1927: \tau_0 \cdot \bbi{} \,, 
1928: \label{conf} 
1929: \ee
1930: where $\rvac (\l)$ is the profile for a single D6-brane determined
1931: in section 2 and $\bbi{}$ is the $\nf \times \nf$ identity matrix.
1932: Since all the transverse scalars commute, there is no ambiguity in the
1933: interpretation of this solution as $\nf$ overlapping D6-branes
1934: lying along the curve $X(\l) = \rvac (\l)$.
1935: 
1936: Let us first argue that this configuration is a stable solution of
1937: the non-Abelian DBI equations of motion. We start by expanding the
1938: non-Abelian DBI action to quadratic order in fluctuations around
1939: the above configuration. Since there is an overall single trace in
1940: front of the action, to linear order only variations proportional
1941: to the identity generator couple to the $U(1)$ background fields,
1942: as given above. Hence the resulting expression for these fields is
1943: exactly that obtained from the linear variation of the action for a
1944: single D6-brane. It thus follows that the configuration \eqn{conf}
1945: is a solution of the non-Abelian equations provided $\rvac (\l)$
1946: solves the corresponding Abelian equations.
1947: 
1948: We further observe that, to quadratic order in fluctuations, the
1949: non-Abelian DBI action reduces to the sum of $\nf^2$ identical
1950: copies of the Abelian action, one for each of the $\nf^2$
1951: independent fluctuations. The reasons for this are two-fold:
1952: First, given that the background fields in \eqn{conf} are
1953: proportional to the identity, the quadratic fluctuations must
1954: again contribute a term in the overall $U(1)$. Second, the
1955: difference between the non-Abelian and the Abelian DBI actions
1956: consists entirely of terms that involve commutators of fields.
1957: Hence, with the given background, the only possible nonvanishing
1958: commutators will involve two fluctuations, but such expressions
1959: are necessarily $SU(\nf)$-valued, \ie, traceless. It follows that 
1960: upon tracing over the gauge indices (and integrating by parts), the
1961: quadratic action will take the form
1962: \be 
1963: \int \df^7\s\,\sum_{\alpha=1}^{\nf^2} \d Z^{i\a}\, {\cal O}_{ij}\,
1964: \d Z^{j\a}\ , 
1965: \label{house} 
1966: \ee
1967: where $Z^i=X, Y$, $\a$ is the gauge index and ${\cal O}_{ij}$ 
1968: is the same operator as appears in the
1969: quadratic expansion of the Abelian action. Thus the linearized
1970: equations of motion, and hence the spectrum, are just $\nf^2$ copies
1971: of those in the Abelian theory describing a single D6-brane
1972: embedding. It follows that the solution \eqn{conf} is stable
1973: provided its Abelian analogue is stable, which we know to be the
1974: case from the analysis of section \ref{fluct}.
1975: 
1976: Note that these $\nf^2$ fluctuations of the D6-branes, which give
1977: rise to the $\nf^2$ copies of the spectrum, are distinct
1978: physical excitations and {\it not} gauge-equivalent: each of them
1979: is associated to a different generator of $U(N_f)$, so there is no
1980: small $U(N_f)$ gauge transformation (that is, one that tends to
1981: the identity at infinity in the 0123-directions) that takes one
1982: mode into another. In particular, this implies that in the limit
1983: $\mq \ra 0$ there are $\nf^2$ massless, pseudo-scalar particles! 
1984: We will come back to this point in the last section.
1985: 
1986: The arguments presented so far show that \eqn{conf} provides a stable
1987: extremum of the non-Abelian DBI action, and therefore a local minimum
1988: of the energy of the $\nf$ D6-branes. Strictly speaking, this
1989: does not exclude the possibility that there exist other minima. If this was
1990: the case, then the true vacuum (or vacua) would be that (or those)
1991: with the lowest energy. On physical grounds, however, the existence of
1992: these other vacua is unlikely. The reason is as follows. If we ignore the
1993: interactions of the D6-branes with each other, then it is clear that
1994: \eqn{conf} is the minimum-energy solution amongst those satisfying the
1995: boundary conditions described above, since in the absence of
1996: interactions each D6-brane must minimise its own energy. If we
1997: include open string tree-level interactions by describing the
1998: D6-branes' dynamics with the non-Abelian DBI action,
1999: then the only obvious configuration that satisfies the appropriate
2000: boundary conditions is \eqn{conf}, which we have shown to
2001: be a stable solution. Inclusion of one-loop (or higher) open string
2002: interactions between the D6-branes (such as exchange of closed
2003: strings) would presumably just strengthen the conclusion that
2004: \eqn{conf} is the preferred solution. To see this, consider the simple
2005: case of $\nf=2$ D6-branes that overlap with each
2006: other asymptotically, but that are slightly misaligned in some
2007: small region. Locally, this situation is analogous to that of two
2008: planar D6-branes in flat space that are almost but not exactly
2009: parallel, in which we know that there is a net force that tends to
2010: align the branes. It seems plausible that a similar force
2011: will act on the branes in the situation of interest here, and
2012: therefore that the preferred configuration will be \eqn{conf},
2013: in which all the branes are perfectly aligned with each other.
2014: 
2015: This conclusion provides a holographic realisation of the
2016: Vafa--Witten theorem, since the $U(\nf)$ gauge symmetry on the
2017: D6-branes is unbroken if and only if they are all exactly
2018: coincident.
2019: 
2020: By continuity, the arguments above imply that, in the limit $\mq
2021: \ra 0$, the $U(\nf)$-preserving configuration has an energy no
2022: greater than that of any other configuration. Just like in the
2023: Vafa--Witten theorem, however, this does not rule out the
2024: possibility that in this limit some other state becomes degenerate
2025: with that in eq.~\eqn{conf}. In fact, the appearance of
2026: $\nf^2$ massless modes in the linearized spectrum at $\mq=0$ suggests 
2027: that the D6-branes might be free to rotate independently in the
2028: 89-plane.\footnote{Note that this is not
2029:   guaranteed by the existence of massless modes, since this does not 
2030:   exclude the appearance of higher order potential terms for the 
2031:   fluctuations, \eg\ $\delta Z^4$.}
2032: To see that this is indeed the case, consider, for simplicity, the 
2033: $\nf=2$ configuration 
2034: \be 
2035: X(\l) = \rvac (\l) \cos\phi_0 \cdot \bbi{} \sac  
2036: Y(\l) = \rvac (\l)
2037: \sin\phi_0 \cdot \sigma_{\it 3} \sac \tau = \tau_0 \cdot \bbi{}
2038: \,, 
2039: \label{confbis} 
2040: \ee
2041: where $\sigma_{\it 3} = \mbox{diag} (1, -1)$ and $\phi_0$ is some 
2042: fixed angle. The radial vacuum profile above is that 
2043: corresponding to $r_\infty=0$. This configuration consists of two 
2044: D6-branes with the same radial profiles but separated an angle 
2045: $2\phi_0$ in the 89-plane, and obviously satisfies the boundary
2046: condition that the branes overlap at $\l \ra \infty$. To see that it solves
2047: the non-Abelain equations of motion, consider expanding the 
2048: non-Abelian DBI action to linear order in fluctuations around \eqn{confbis},
2049: as before. 
2050: Since commutator terms are irrelevant at this order, and $X$ and
2051: $Y$ are diagonal, the non-Abelian action again reduces to the sum of
2052: two identical Abelian actions. In fact, the same argument shows that not
2053: only is the configuration \eqn{confbis} a solution of the non-Abelian 
2054: equations, but also that its energy is, for all $\phi_0$, equal to
2055: twice that of a single D6-brane. It follows that \eqn{confbis} is
2056: stable for all $\phi_0$. To see this, suppose that it is
2057: not, \ie, that there is some fluctuation around it that is tachyonic.
2058: Presumably, this would mean that this solution can decay to another
2059: one of lower energy. However, the latter would also have lower energy
2060: than the solution with $\phi_0=0$, in contradiction with our
2061: continuity argument above. 
2062: 
2063: Note that the previous reasoning applies equally well to any
2064: distribution of $\nf$ D6-branes around the $\phi$-circle, since it
2065: only relies on $X$ and $Y$ being diagonal. Thus, at $\mq=0$, all these 
2066: configurations are degenerate, lowest-energy, stable solutions. 
2067: Although it may be an interesting exercise to verify their stability 
2068: explicitly by expanding the non-Abelian DBI action to quadratic order in
2069: fluctuations around \eqn{confbis}, one should keep in mind
2070: that this conclusion is only expected to hold at leading order in the
2071: $1/\nc \sim g_s$ expansion, since, as discussed above, we expect 
2072: higher-order effects to lead to an attractive force between the branes,
2073: which would tend to align them.
2074: 
2075: 
2076: 
2077: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2078: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2079: %%%%%%%%%%%%%%%%%%%%%%%%%% SECTION %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2080: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2081: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2082: \section{Finite temperature physics} \label{fintem}
2083: 
2084: The gauge theory we have so far been describing is at zero
2085: temperature. To study the theory at a finite temperature $T$, we
2086: may follow the standard prescription: analytically continue the
2087: time coordinate $t\to t_\mt{E}=it$, periodically identify
2088: $t_\mt{E}$ with period $\delta \te = 1/T$, and impose
2089: anti-periodic boundary conditions on the fermions around the
2090: $\te$-circle. This prescription applies equally well in the gauge
2091: theory and in the dual string description. In the latter, however,
2092: one must consider all possible spacetime solutions with the
2093: appropriate boundary conditions, that is, possible saddle points 
2094: of the Euclidean path integral over supergravity (or rather, string)
2095: configurations \cite{Witten98b}.
2096: 
2097: In the present case, there are two\footnote{In analogy with
2098: \cite{fairies}, there may be an infinite family of Euclidean
2099: solutions labelled by two integers specifying which cycle of the
2100: $\te\tau$-torus shrinks to zero-size in the bulk. However, given
2101: that the $\te$- and $\tau$-axes are orthogonal, here either the
2102: solutions in eq.~\eqn{d4nut} or \eqn{bhmetric} will dominate the
2103: path integral.} Euclidean supergravity solutions which must be
2104: considered. The first is simply the Euclideanised version of that
2105: appearing in eq.~\eqn{metric},
2106: \beq 
2107: ds^2=\left(\frac{U}{R}\right)^{3/2} \left(d\te^2 +
2108: \sum_{i=1}^3 dx^i dx^i + f(U) d\tau^2\right)+
2109: \left(\frac{R}{U}\right)^{3/2}\frac{dU^2}{f(U)}
2110: +R^{3/2}U^{1/2}d\Omega^2_{\it 4}\,,
2111: \label{d4nut} 
2112: \eeq
2113: which dominates at low temperatures. Recall that 
2114: $f = 1 -\ut^3 / U^3$. The second is the Euclidean black hole 
2115: coming from a nonextremal D4-brane throat
2116: %
2117: \beq
2118: ds^2=\left(\frac{U}{R}\right)^{3/2}\left(\tilde{f}(U)dt_\mt{E}^2+
2119: \sum_{i=1}^3 dx^i dx^i +d\tau^2\right)+
2120: \left(\frac{R}{U}\right)^{3/2}\frac{dU^2}{\tilde{f}(U)}
2121: +R^{3/2}U^{1/2}d\Omega^2_{\it 4} \,, \label{bhmetric} \eeq
2122: %
2123: which dominates at high temperatures. In this case, $\tilde{f}=1-
2124: \uh^3 / U^3$. In the Lorentzian-signature solution, the event
2125: horizon is located at $U=\uh$.
2126: 
2127: Of course, these Euclidean metrics are identical up to
2128: interchanging the role of $\te$ and $\tau$ (and replacing $\ut$
2129: with $\uh$). To match the boundary conditions set by the 
2130: finite-temperature gauge theory, we identify these coordinates with
2131: periods
2132: \be 
2133: \d\te  = \fc{1}{T} \sac \d\tau = \fc{2\pi}{\mkk} \,,
2134: \label{periods}
2135: \ee
2136: in both solutions. Since we wish to avoid conical
2137: singularities in either spacetime at $U=\ut, \uh$, the
2138: boundary conditions fix the metric parameters as
2139: \be 
2140: \ut =\left({4\pi\over3\,\d\tau}\right)^2R^3\sac \uh=
2141: \left({4\pi\over3\,\d\te}\right)^2R^3\,. 
2142: \label{parmu}
2143: \ee
2144: 
2145: As we commented above, the metric \reef{d4nut} dominates the path
2146: integral at low temperatures, whereas that given by eq.~\reef{bhmetric} 
2147: dominates at high temperatures. One determines the
2148: transition point by comparing the Euclidean action of these two
2149: solutions \cite{Witten98b}. As the metrics describe the same
2150: geometry, it is easy to see that the transition occurs precisely when 
2151: $\d \tau=\d \te$. That is, a phase transition occurs at the critical
2152: temperature
2153: \be 
2154: \ct=\mkk/2\pi\,.
2155: \label{phtemp}
2156: \ee
2157: From the gauge theory viewpoint, this is a confinement/deconfinement 
2158: phase transition. This can be seen by the fact that the temporal 
2159: Wilson loop (around $\te$) vanishes for the low-temperature 
2160: background \reef{d4nut}, whereas it is
2161: nonvanishing for the Euclidean black hole \reef{bhmetric}
2162: \cite{Witten98b}. Alternatively, on the Lorentzian sections, the
2163: Wilson loop as computed from the solution \eqn{metric} exhibits an
2164: area law behaviour, while that computed for the Lorentzian black
2165: hole does not because the string can break in two pieces, each
2166: having an endpoint at the horizon \cite{break}. One might also
2167: observe that, in the low-temperature phase, one has a discrete
2168: spectrum of glueballs \cite{COOT98}, but this gives way to a
2169: continuum of excitations in the high-temperature phase dual to the
2170: black hole background.\footnote{Note that the spectrum would again
2171: be calculated with Lorentzian signature.}
2172: The analogous phase transition in 3+1 dimensions was discussed 
2173: in ref.~\cite{phases}. Finally, note that the phase transition takes 
2174: place at a temperature, \reef{phtemp}, where the gauge theory is
2175: starting to behave five-dimensionally, rather than four-dimensionally.
2176: 
2177: The preceding discussion refers only to the physics of the
2178: (compactified) five-dimensional gauge theory and is independent of
2179: the presence of fundamental matter fields. In previous sections we
2180: have exhaustively analysed the physics of dynamical quarks in the
2181: confining phase. We can extend the discussion to the deconfining
2182: phase by considering probe D6-branes in the black hole background
2183: \eqn{bhmetric}. In reference \cite{BEEGK03} the study of D7-branes
2184: in an asymptotically \ads{5} black hole background revealed the
2185: existence of two sets of regular embeddings, parametrised by the
2186: quark mass and chiral condensate of the dual gauge theory, and
2187: distinguished by whether or not the D7-brane intersects the
2188: horizon. It was suggested that the gauge theory undergoes a phase
2189: transition, signalled by a discontinuity in $dc/dm_\mt{q}$, when
2190: the D7-brane undergoes its `geometric' transition. Here we
2191: consider similar physics in our four-dimensional gauge theory.
2192: 
2193: As before, the D6-branes span (Euclidean) time and the spatial
2194: 123-directions in common with the D4-brane worldvolume, and lie at
2195: a fixed value of $\tau$. In parallel with the steps described by
2196: eqs.~(\ref{isomer}--\ref{isometric}), we introduce isotropic
2197: coordinates $\{\l, \Omega_{\it 2}, r, \phi\}$ for the Euclidean
2198: black hole \reef{bhmetric}. Furthermore,  we rescale to dimensionless
2199: coordinates
2200: \beq 
2201: \l \to \uh\l\ ,\qquad r\to \uh r\ ,\qquad \r\to \uh\r\ , 
2202: \eeq
2203: in analogy with eq.~\reef{tildes}. Now the embedding equation for
2204: $r(\l)$ becomes
2205: \beq
2206: \frac{d}{d\l}\left[\left(1-\left(\frac{1}{4\r^3}\right)^2\right)\l^2
2207: \frac{\dot{r}}{\sqrt{1+\dot{r}^2}}\right]
2208: =\frac{3}{8}\frac{r\l^2}{\r^8}\sqrt{1+\dot{r}^2}\,.
2209: \label{bhembed}
2210: \eeq
2211: Comparing this result with eq.~\reef{resc}, we see that the sign
2212: of the right-hand side has changed. As we will see below, this
2213: change of sign indicates that the brane now bends towards $U=\uh$:
2214: as expected, the black hole exerts an attractive force on
2215: the D6-brane, in contrast to the repulsive force in the 
2216: \nonsol\ background.
2217: 
2218: If we look for asymptotically constant embeddings such that
2219: $r(\l)\to r_\infty=L/\uh$ as $\l\to\infty$, we find the same
2220: long-distance behaviour as before:
2221: %
2222: \beq r\simeq
2223: r_\infty+\frac{c}{\l}\ .\label{asymprof} \eeq
2224: %
2225: These asymptotic parameters are related to the quark mass and
2226: chiral condensate as before, except for the substitution of $\uh$
2227: for $\ut$:
2228: %
2229: \be \mq = \fc{\uh r_\infty}{2\pi\ell_s^2}\sac \cc =  -8 \pi^2
2230: \ell_s^2 \t6 \uh^2 \, c \,. \label{magain} \ee
2231: %
2232: Again, the embedding equation can be solved numerically for
2233: any value of $r_\infty$, with the asymptotic boundary conditions
2234: implicit in \reef{asymprof}. Inspection of the results reveals
2235: that, just as in reference \cite{BEEGK03}, two qualitatively
2236: different types of profile are possible: those for which the probe
2237: brane intersects the horizon, and those for which it does not. As
2238: we vary the boundary condition $r_\infty$, we find that there is
2239: a unique embedding for sufficiently large or small values. In the
2240: case of small $r_\infty$, the unique embedding is such that the
2241: D6-brane falls into the horizon, whereas in the large-$r_\infty$
2242: case it does not, despite, of course, still being attracted by the
2243: black hole. For an intermediate range with
2244: $r_\infty\sim 1$, we find that more than one embedding is
2245: possible. In particular, for a given $r_\infty$, we find both
2246: embeddings which intersect the horizon and those where the
2247: D6-brane smoothly closes off before reaching the horizon.
2248: Furthermore, we also find that there are typically
2249: multiple solutions of at least one of these varieties of
2250: embedding; figure \ref{bhprof} displays some representative
2251: examples. These results imply that in varying from large to
2252: small $r_\infty$, there will be a phase transition in the
2253: behaviour of the quarks. This transition will occur in the
2254: intermediate regime, where, for a fixed temperature and quark mass,
2255: the condensate in the gauge theory can have multiple values. Of
2256: course, the true ground-state condensate will be selected by
2257: minimizing the energy density.
2258: 
2259: \FIGURE{ \epsfig{file=bhprofiles.eps, height=6cm} 
2260: \caption{Some representative D6-brane embeddings from the region 
2261: in which $c$ is multi-valued. The three profiles in each figure 
2262: have the same asymptotic value of $r_\infty$; in (a) two of them 
2263: avoid the horizon whereas in (b) two fall into the horizon.}
2264: \label{bhprof} }
2265: 
2266: Combining various results --- see eqs.~\eqn{inverse}, \eqn{periods}, 
2267: \reef{parmu} and \eqn{magain} --- the relation between the temperature
2268: and $r_\infty$ is found to be
2269: \be 
2270: T = \fc{\tb}{\sqrt{r_\infty}} \sac 
2271: {\rm with}\ \tb^2 \equiv
2272: \frac{9}{4\pi}\frac{m_\mt{q}M_\mt{KK}}{g^2_\mt{YM}N_\mt{c}} \,.
2273: \label{Tfromr} 
2274: \ee
2275: Here $\tb$ can be interpreted, for a given quark mass, as the 
2276: typical meson mass scale in the regime where supersymmetry is 
2277: restored in the zero-temperature gauge theory --- see the 
2278: discussion below eq.~\reef{region2}. In the zero-temperature 
2279: analysis of previous sections we regarded the variation of 
2280: $r_\infty$ as a variation of the quark mass, while $\ut$ or, 
2281: equivalently, the KK scale, was fixed. In this section we 
2282: wish to study the phase structure of the gauge theory as the
2283: temperature is varied, keeping the microscopic gauge theory
2284: parameters, and in particular the quark mass, fixed.
2285: For this reason, since in this phase the physics will only 
2286: depend on the ratio $T/\tb$, as suggested by eq. \eqn{Tfromr}, 
2287: we will think of the variation of $r_\infty$ as a 
2288: variation of the temperature.
2289: 
2290: \FIGURE{ \epsfig{file=ecvsT.eps, height=15cm} \caption{The quark
2291: condensate $\cc \propto c$ (plots (a) and (b)),
2292: and the energy density $\tilde{\cale}$ (plots (c) and (d)) as
2293: functions of temperature. The solid (dashed) lines describe the
2294: family of embeddings that do (not) intersect the horizon. Plots
2295: (b) and (d) show in more detail the regions of plots (a) and (c)
2296: where the two branches intersect. The vertical line lies at
2297: $T=\tcc$, the temperature at which the phase transition in the
2298: behaviour of the quarks takes place.}
2299:  \label{cvsT} }
2300: 
2301: The behaviour of the condensate (represented by $c$) as a function
2302: of the temperature is shown in figures \ref{cvsT}a and \ref{cvsT}b. 
2303: These plots
2304: display some interesting features. First note that $c\ra0$ both as
2305: $T\ra0$ and as $T\ra\infty$. In the high-temperature regime, the
2306: unique embedding is such that the D6-brane falls into the
2307: horizon. By following the solid line in the figure, we see
2308: that these embeddings only exist down to some some minimum
2309: temperature, $T\simeq 0.9869 \tb$. Similarly there is a unique
2310: embedding at sufficiently low temperatures such that the D6-brane
2311: does not reach the horizon, and following the dashed line shows
2312: that embeddings of this form only exist up to $T\simeq1.0202\tb$.
2313: As shown most clearly in figure \ref{cvsT}b, multiple solutions
2314: exist in the intermediate temperature range between the two values
2315: given above.
2316: 
2317: Naturally, where more than one solution to the embedding equation
2318: exists, the ground-state profile will be that selected out by
2319: lowest energy density. Since we are considering static
2320: configurations, the energy density is given by
2321: \beq
2322: \cale=-\int \df\l\df\Omega_{\it 2} {\cal L}= 4\pi
2323: T_\mt{D6}\uh^3 \,\int_{\l_\mt{min}}^\infty\df\l\
2324: \left(1-\left(\frac{1}{4\r^3}\right)^2\right)
2325: \l^2\sqrt{1+\dot{r}^2}\ ,
2326: \label{bhenergy}
2327: \eeq
2328: where $\l_\mt{min}$ is either zero, or the point where the brane
2329: meets the horizon, depending on which variety of embedding we are
2330: considering. This integral is divergent, so to compare the energy
2331: densities of different embeddings we regularize $\cale$ by using as a 
2332: reference the
2333: embedding $r(\l)=0$, which runs from the horizon (at $\l=\l_\mt{H}$, say)
2334: to infinity. Since this is a numerical calculation, we
2335: also introduce a cut-off, $\lm$, in $\l$ at  the point where, by
2336: imposing the appropriate boundary conditions, we match the
2337: numerical profile to the asymptotic solution, $r\simeq r_\infty+c/\l$
2338: . 
2339: There is an
2340: infinite contribution to $\cale$ from the remainder of the profile, but
2341: when regularized the leading behaviour of this correction
2342: can be shown to be
2343: \beq
2344: \cale_\mt{corr}=\frac{c^2}{2\lm}+\calo(\lm^{-3})\ .
2345: \eeq
2346: Therefore, to be precise, we compare the energy densities by calculating
2347: \beqa
2348: \tilde{\cale}=\frac{\cale_\mt{reg}}{4\pi
2349: T_\mt{D6}\uh^3}&\equiv&\int_{\l_\mt{min}}^{\lm} \df\l\
2350: \left(1-\left(\frac{1}{4\r^3}\right)^2\right)\l^2\sqrt{1+\dot{r}^2}\non\\
2351: && \hspace{1cm} -\int_{\l_\mt{H}}^{\lm} \df\l\
2352: \left(\l^2-\frac{1}{16\l^4}\right)+
2353: \frac{c^2}{2\lm} \,.
2354: \eeqa
2355: We then trust that this provides a valid basis for the comparison of
2356: different energy densities when the differences in $\tilde{\cale}$ are
2357: greater than $\calo(\lm^{-3})$. This is the case here, as we use
2358: $\lm=100$.
2359: 
2360: The energy density $\tilde{\cale}$ is plotted as a function of temperature
2361: in figures \ref{cvsT}c and \ref{cvsT}d. Starting from low temperatures, figure
2362: \ref{cvsT}d shows that, as $T$ is increased to that in the
2363: intermediate region, the physical embedding remains that which
2364: does not reach the horizon. At $T\simeq 1.0202 \tb$, the energy
2365: density for this embedding and that which falls into the horizon
2366: match. For $T > 1.0202 \tb$, the energetically favoured embedding is
2367: one which falls into the horizon. Therefore, at
2368: \be 
2369: T\equiv\tcc\simeq 1.0202\tb\,, 
2370: \ee
2371: there is a phase transition in the behaviour of the fundamental
2372: quarks. This phase transition is signalled by a discontinuity in
2373: $\langle \bar{\psi} \psi \rangle(T)$, as shown in Figure
2374: \ref{cvsT}b, as well as a discontinuity in the specific heat
2375: $d\tilde{\cale}(T)/dT$, as can be seen in Figure \ref{cvsT}d. Note
2376: that, to within the precision of our numerical calculations, $\tcc$ 
2377: cannot be distinguished from the maximum temperature for which 
2378: a horizon-avoiding embedding is possible.  
2379: 
2380: From the string description with the D4 throat geometry, the basic
2381: physics behind this transition is straightforward. Increasing the
2382: temperature increases both the radial position of the horizon and
2383: energy density of the black hole. For  small enough a
2384: temperature, the D6-brane is pulled towards the horizon but
2385: its tension balances the gravitational attraction. However, above
2386: the critical temperature $\tcc$, the gravitational force has
2387: increased to the point where it overcomes the tension and the
2388: brane is drawn into the horizon.
2389: 
2390: Another feature that characterizes the present phase transition is
2391: that the meson spectrum changes at $T=\tcc$. The discrete spectrum of
2392: mesons of the low-temperature phase becomes a continuum 
2393: of excitations in the high-temperature phase where the D6-brane 
2394: embedding extends to the horizon. This behaviour is similar to that of 
2395: the glueball spectrum at the deconfinement phase transition in the 
2396: gauge theory, discussed previously. 
2397: Note, however, that for temperatures above $\ct$ we are, by 
2398: assumption, in the deconfined phase of the gauge theory, and so it is 
2399: possible to introduce free quarks; these correspond to open strings 
2400: extending from the D6-brane down to the horizon. If $\ct < \tcc$ there 
2401: is then a range of temperatures, $\ct < T < \tcc$, in which there are
2402: free quarks but also a finite mass gap, $m_\mt{eff}(T)>0$, 
2403: associated with the introduction of such quarks into the 
2404: system. In this regime,
2405: there still exists a (discrete) spectrum of bound states with 
2406: $M<2m_\mt{eff}$. Note also that the deconfined adjoint degrees of
2407: freedom still generate a finite screening length beyond which the force 
2408: between a quark and an anti-quark vanishes \cite{break}; in the 
2409: string description this happens because the string that joins the 
2410: quarks together snaps into two pieces, each of them going from the
2411: brane to the horizon, that can be separated with no cost in energy. 
2412: Above $\tcc$, however, $m_\mt{eff}(T)=0$ and the spectrum of
2413: fundamental matter field excitations becomes continuous. 
2414: 
2415: Up to this point, we have discussed two independent phase
2416: transitions: the deconfining phase transition at $T=\ct$, for
2417: which the holographic dual description involved a discrete change
2418: of the bulk spacetime geometry, and the quark phase transition at
2419: $T=\tcc$, for which the holographic description involved a
2420: discrete jump in the embedding geometry of the D6-brane. However,
2421: if the latter is to be realized as a separate phase transition,
2422: then it must be true that $\tcc>\ct$.\footnote{Note that, in this
2423: regime, the fundamental fields still `feel' the deconfining
2424: transition at $T=\ct$. For example, there is a discontinuity in
2425: the chiral condensate $\cc$, which jumps from a negative to a
2426: positive value at this phase transition.} 
2427: In terms of the microscopic parameters, this condition translates into
2428: \beq 
2429: m_\mt{q} > {0.98\over 9\pi}g^2_\mt{YM} N_\mt{c} M_\mt{KK}\ . 
2430: \label{phtemp2}
2431: \eeq
2432: In the zero-temperature theory, this is roughly the quark mass
2433: scale at which supersymmetry is restored --- see the discussion below
2434: eq.~\reef{region2}.
2435: 
2436: If $\tcc<\ct$, the relevant spacetime geometry at $T=\tcc$ is the
2437: \nonsol{} \reef{d4nut} and the ambiguity in the D6-brane embedding
2438: discussed above is not relevant. Alternatively, the gauge theory
2439: is still in a confined phase and so the chemical potential to
2440: introduce a free quark remains infinite; in the dual description this
2441: is because, since there is no horizon, an open string must have both
2442: ends on the D6-brane, and so represents a quark/anti-quark bound
2443: state. In this case, the two phase transitions are realized
2444: simultaneously at $T=\ct$. An important example of this behaviour is 
2445: when $\mq=0$. In this case, the low-temperature phase is characterized 
2446: by spontaneous chiral symmetry breaking, as in section \ref{broken}.
2447: In the high-temperature phase, $\mq \ra 0$ is equivalent to 
2448: $T \ra \infty$ (see eq. \eqn{Tfromr}), in which limit the chiral
2449: condensate vanishes (see figure \ref{cvsT}a). Hence the deconfining 
2450: phase transition is also accompanied by the restoration of chiral 
2451: symmetry in this model.
2452: 
2453: To close the section, we wish to stress to the reader that while
2454: these phase transitions are of inherent interest in understanding the
2455: model under study here, they take place in a temperature regime
2456: where the gauge theory is intrinsically five-dimensional, rather
2457: than behaving in a four-dimensional way, as can be seen by
2458: comparing the transition temperatures given in eqs.~\reef{phtemp}
2459: and \reef{phtemp2} to the compactification scale $\mkk$.
2460: 
2461: 
2462: 
2463: 
2464: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2465: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2466: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2467: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2468: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2469: \section{D4/D6/$\overline{\mbox{D6}}$-intersections}
2470: \label{Dbar}
2471: 
2472: In the discussion of embeddings in section \ref{broken}, we noted
2473: that the solution $\rt (\lt) =0$ of eq.~\eqn{resc} is not a
2474: physical one (for $\ut \neq 0$) if it terminates at the origin of the
2475: $U\tau$-plane because it would correspond to an open D6-brane.
2476: However, this solution can be extended through this point to
2477: construct a physically acceptable embedding. To see this, recall
2478: that the full D6-brane embedding is further specified by the
2479: conditions $\phi=\phi_0$ and $\tau=\tau_0$, where $\phi_0$ and
2480: $\tau_0$ are arbitrary constants. We can therefore join together a
2481: solution with $\tau=\tau_0$ to another one with $\tau=\tau_0+ \d
2482: \tau/2$ (see eq.~\eqn{deltatau}) to obtain a complete regular
2483: solution that describes a D6-brane whose projection on the
2484: $U\tau$-plane is a straight line that passes through the origin
2485: and intersects the boundary at two antipodal values of $\tau$.
2486: This means that this solution describes an intersection of $\nc$
2487: D4-branes with two D6-branes, or, more precisely, with one
2488: D6-brane and one anti-D6-brane.\footnote{Technically, the reason
2489: is that the orientations on the intersections of the D6-brane with
2490: the boundary at $\tau=\tau_0$ and $\tau=\tau_0+ \d \tau/2$,
2491: induced from a given orientation on the D6-brane, are opposite to
2492: each other.} Note that these embeddings bear a striking resemblance to
2493: those discussed in \cite{SS03}. The parallel includes the facts that
2494: both constructions involve two boundary defects and both preserve
2495: the chiral symmetry.
2496: 
2497: 
2498: To better understand the defect theory corresponding to these
2499: D4/D6/$\overline{\mbox{D6}}$-intersections, we consider a family
2500: of more general embeddings that describe D6-branes that lie at
2501: $\rt=0$ (or alternatively, $z^8=0=z^9$) and stretch along some
2502: curve $U=U(\tau)$ in the $U\tau$-plane, intersecting the boundary
2503: at two, not necessarily antipodal, values of $\tau$. Such
2504: D6-branes wrap a maximal $S^2$ within the $S^4$ in the metric
2505: \eqn{metric}, so the induced metric on their worldvolume is
2506: \be 
2507: ds^2(g) = \left( \fc{U}{R} \right)^{3/2} \, 
2508: \eta_{\mu\nu} dx^\mu dx^\nu +
2509: R^{3/2} U^{1/2} \, d\Omega_{\it 2}^2 + \left[ \left( \fc{U}{R}
2510: \right)^{3/2} \, f(U) + \fc{U'^2}{\left( \fc{U}{R} \right)^{3/2}
2511: \, f(U)} \right] \, d\tau^2 \,. 
2512: \ee
2513: Since the Lagrangian density on the D6-brane
2514: \be 
2515: \call_{\mt{D6}} = - \fc{1}{(2\pi)^6 \ell_s^7} \, 
2516: e^{-\phi} \, \sqrt{-\det g_{\mt{ind}}} = 
2517: \t6 \, U^2 \, \sqrt{\left( \fc{U}{R} \right)^3 f(U) + \fc{U'^2}{f(U)}} 
2518: \ee
2519: (where $U' = dU/d\tau$) does not depend explicitly on $\tau$, 
2520: it follows that
2521: \be 
2522: U' \, \fc{\pa \call}{\pa U'} - \call = \fc{U^2 \left( \fc{U}{R}
2523: \right)^3 f(U)} {\sqrt{\left( \fc{U}{R} \right)^3 f(U) +
2524: \fc{U'^2}{f(U)}}} 
2525: \ee
2526: is also $\tau$-independent. Let $U_0 \geq \ut$ be the minimum
2527: value of $U(\tau)$. By symmetry we can assume that this value is
2528: reached at $\tau=0$, that is, that we have $U(0)=U_0$ and
2529: $U'(0)=0$. Integrating the equation above we then find
2530: \be
2531: \tilde{\tau} (u) = \fc{3}{2} \, b^{7/2} \, g^{1/2}(b) \, \int_b^u
2532: dx \fc{1}{x^{3/2} g(x) \sqrt{x^7 g(x) - b^7 g(b)}} \,, \label{tau}
2533: \ee
2534: where we have introduced
2535: \be 
2536: \tilde{\tau} \equiv \fc{3}{2} \,
2537: \fc{\ut^{1/2}}{R^{3/2}} \, \tau \sac u \equiv \fc{U}{\ut} \sac b
2538: \equiv \fc{U_0}{\ut} \geq 1 \sac g(x) \equiv 1 - \fc{1}{x^3} \,.
2539: \ee
2540: Note that the rescaled angle $\tilde{\tau}$ has period $2\pi$.
2541: Eq.~\eqn{tau} specifies the D6-brane profile in the $U\tau$-plane.
2542: The D6-brane intersects the boundary at $\tilde{\tau}_\pm (b) =
2543: \pm \tilde{\tau}(u \ra \infty)$. These two points are implicitly
2544: fixed by $U_0$, the `lowest' point of the D6-brane in the
2545: $U\tau$-plane. In particular, their separation is a
2546: monotonically decreasing function of $b$. It is easy to show that
2547: $\tilde{\tau}_\pm (b \ra 1) = \pm \pi/2$, so the two points
2548: become antipodal as the lowest point of the D6-brane approaches
2549: the origin of the $U\tau$-plane. In this limit the projection of
2550: the D6-brane on this plane is just a straight line, as discussed
2551: above. In the opposite limit in which the (lowest point of)
2552: D6-brane approaches the boundary, the separation between the
2553: intersection points becomes smaller and smaller. In the strict
2554: limit $b\ra \infty$ the two coincide and the D6-brane disappears.
2555: 
2556: Now, these embeddings seem to portend fascinating physics for the
2557: dual five-dimensional gauge theory coupled to two defects or, more
2558: precisely, to a defect and an anti-defect. It appears, however, that 
2559: this physics is intrinsically five-dimensional. 
2560: Since the focus of the present
2561: paper is the four-dimensional QCD-like physics of the original
2562: model, we have not fully explored this system. However, we offer
2563: some preliminary remarks in the following.
2564: 
2565: The deflection of the D6-brane in the $\tau$-direction, which
2566: corresponds to a worldvolume direction of the D4-branes, is
2567: reminiscent of certain D5-brane embeddings in \ads{5} studied in
2568: ref.~\cite{ST02}. These authors found (non-supersymmetric)
2569: embeddings in which the D5-brane began at large radius running
2570: along $r$, but were deflected to run parallel to the \ads{5}
2571: horizon at some finite radius. At large $r$, these configurations can
2572: be understood as D3/D5 intersections that are T-dual to that
2573: represented in \eqn{intersection}. The dual gauge theory involved
2574: coupling $\caln=4$ SYM to a ($2+1$)-dimensional defect which broke 
2575: all supersymmetries. The deflection was interpreted as giving an 
2576: expectation value to an operator in the defect theory. 
2577: 
2578: In the present case, the \nonsol\ background breaks the
2579: supersymmetry but with both defects, supersymmetry would still not
2580: be restored in the limit $\ut\ra0$;\footnote{Note that the
2581: supersymmetry breaking of the background is not essential to
2582: constructing the general embeddings, \ie, one can simply replace
2583: $g(x)$ by 1 everywhere in eq.~\reef{tau}.} 
2584: one way to understand this in the field theory is to note that the half 
2585: of the supersymmetries of the ambient theory preserved (broken) 
2586: by the defect are precisely those that are broken (preserved) by the
2587: anti-defect.\footnote{Defect/anti-defect configurations, and more
2588:   general multiple defect configurations, were discussed in \cite{MNT02}.} 
2589: The embeddings described by eq.~\reef{tau} would correspond
2590: to a state where a symmetric pair of operators in each defect
2591: theory simultaneously acquire identical expectation values.
2592: Indirectly, the D6-brane embedding would also induce different
2593: expectation values of gauge theory operators, \eg, Tr $F^2$, in the
2594: two regions on the $\tau$-circle divided by the defects.
2595: 
2596: At the microscopic level, there are two sets of fundamental matter 
2597: fields in the field theory, one associated with each defect.
2598: This is realised in the string description by the fact that,
2599: despite there being a single D6-brane, its intersection with a
2600: constant-$U > U_0$ slice consists of two disconnected pieces, 
2601: corresponding to two independent sets of active degrees of freedom 
2602: at the corresponding energy scale. These degrees of freedom seem to
2603: disappear for energy scales associated to radial positions $U < U_0$, 
2604: as in \cite{ST02}. As discussed above, the D6-brane disappears in the limit 
2605: $\tilde{\tau}_+-\tilde{\tau}_- \ra 0$. Consistently, in this limit 
2606: $U_0 \ra \infty$ and the degrees of freedom above are completely
2607: removed. In the gauge theory, this corresponds to the limit in which
2608: the defect and the anti-defect are brought on top of each other.
2609: 
2610: This situation may be compared with that where both
2611: defects are associated with a `standard' D6-brane embedding as
2612: described in section \ref{broken}; one at 
2613: $\tilde{\tau}= \tilde{\tau}_+$  and the other at 
2614: $\tilde{\tau}= \tilde{\tau}_-$. Of course, in this case there are 
2615: also two sets of independent degrees of freedom associated with open 
2616: string excitations on both the D6- and 
2617: $\overline{\mbox{D6}}$-brane.\footnote{Note that in this case one 
2618: may also realise a defect/defect (as opposed to a defect/anti-defect) 
2619: theory by considering two D6-branes.} However, one difference is that
2620: the spectrum of these excitations will now be independent of the 
2621: defect/anti-defect separation,
2622: $\tilde{\tau}_+-\tilde{\tau}_-$.\footnote{In the approximation in
2623:   which interactions between the branes, which are suppressed in the
2624:   large-$\nc$ limit, are neglected.}  
2625: In addition, one would also find excitations corresponding to open 
2626: strings stretching between the two separate branes. 
2627: In contrast, in the case of a single D6-brane, the two defects can 
2628: interact through open string modes that propagate along the brane from 
2629: one defect to the other. 
2630: 
2631: Of course, one should ask the question which embedding minimizes
2632: the energy density of the system. That is, does energetics favour
2633: the smooth embedding which joins the two boundary defects or that
2634: in which there are two independent branes? As a preliminary step
2635: in this direction, we compared the energy density for the
2636: `standard' embeddings of the D6- and $\overline{\mbox{D6}}$-brane
2637: pair with $r_\infty=0$ to that of the smooth embedding above with
2638: $U_0=\ut$ (and hence $\tilde{\tau}_+-\tilde{\tau}_- =\pi$). Our
2639: numerical results indicate that the smooth embedding produces a
2640: lower energy density and hence describes the true vacuum
2641: configuration. Note then that with both defects corresponding to
2642: the D6- and $\overline{\mbox{D6}}$-branes, the vacuum preserves
2643: chiral symmetry.
2644: 
2645: Presumably if $U_0$ is increased above $\ut$, the smooth
2646: embeddings described above would continue to be energetically
2647: favored as it seems the energy density for these solutions must
2648: decrease as $\tilde{\tau}_+-\tilde{\tau}_-$ decreases while that
2649: for the standard embeddings remains unchanged. Note, however, that
2650: the above are just two particular cases of a more general family
2651: of embeddings in which the brane bends simultaneously in the 89-
2652: and the $U\tau$-planes, that is, embeddings with nontrivial
2653: functions $r=r(\l)$ and $\tau=\tau(\l)$ (using the notation of
2654: section 2). In general, we expect the minimum-energy two-defect
2655: embedding is of this type.
2656: 
2657: These defect theories might be explored in other ways as well. For
2658: example, one might also consider varying the parameter $\r_\infty$
2659: for the two defects, which may be done independently for each of
2660: the defects. Finally, of course, one might also consider the
2661: effects of a finite temperature, as in section 6.
2662: 
2663: 
2664: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2665: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2666: %%%%%%%%%%%%%%%%%%%%%% SECTION  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2667: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2668: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2669: \section{Discussion} \label{discus}
2670: 
2671: We have analyzed in detail the case of a single D6-brane, which
2672: corresponds to a single flavour in the gauge theory.
2673: In this case the gauge theory enjoys (at large $\nc$) an exact
2674: $U(1)_\mt{V} \times \ua$ symmetry. The first factor is just the flavour
2675: symmetry, which in the string description corresponds to the gauge
2676: symmetry on the worldvolume of the D6-brane. The second factor is
2677: a chiral symmetry that rotates the left- and right-handed
2678: fermions, $\psi_\mt{L}$ and $\psi_\mt{R}$, with opposite phases;
2679: it also rotates by a phase the adjoint scalar $X=X^8 + i X^9$,
2680: as is clear from the fact that in the string description the
2681: $\ua$ symmetry is just the rotational symmetry in the 89-plane.
2682: 
2683: Like in QCD, we expect the $\ua$ symmetry to be spontaneously
2684: broken by a chiral condensate, $\cc \neq 0$, and hence the
2685: existence of the corresponding Goldstone boson in the spectrum;
2686: this is the analog of the QCD $\eta'$ particle (which is massless in
2687: the $\nc \ra \infty$ limit), but, in an abuse of language, we have
2688: referred to it as a `pion'. The string description indeed confirms
2689: the field theory expectation: for zero quark mass, there is precisely
2690: one massless pseudo-scalar (in the four-dimensional sense) fluctuation of the
2691: D6-brane. If $\mq> 0$ this pion becomes a pseudo-Goldstone boson. Its
2692: mass, as well as the chiral condensate and the pion decay constant,
2693: can be computed in the string description, and we have shown that
2694: these three quantities and the quark mass obey the
2695: Gell-Mann--Oakes--Renner relation \eqn{GMOR} \cite{GMOR68}.
2696: Although this is a reassuring result, one should keep in mind that,
2697: just like in field theory, it is essentially a consequence of the
2698: {\it symmetry}-breaking pattern.
2699: 
2700: In the case $\nf >1$ we provided a holographic version of the
2701: Vafa--Witten theorem, which states that, in a vector-like theory
2702: with zero $\theta$-angle (like QCD) and $\nf$ quark flavours of
2703: identical masses $\mq >0$, the global, $U(\nf)$ flavour
2704: symmetry cannot be spontaneously broken. As we discussed,
2705: it is a priori unclear, on field theory grounds, that the Vafa--Witten
2706: theorem does apply to the actual gauge theory on the D4-branes, because
2707: of the presence of Yukawa-like couplings to scalar fields
2708: of masses $\mkk \sim \Lqcd$. The result from the string analysis
2709: therefore established that the presence of these fields does not alter
2710: the vacuum structure of the theory to the extent of violating the
2711: Vafa--Witten theorem.
2712: 
2713: The global $U(\nf)$ symmetry of the gauge theory becomes the gauge
2714: symmetry on the worldvolume of $\nf$ D6-branes in the string
2715: description. The holographic version
2716: of the Vafa--Witten theorem is the statement that the minimum-energy
2717: configuration for the $\nf$ D6-branes is a $U(\nf)$-preserving one in
2718: which all branes lie on top of each other. It should be emphasised
2719: that, unlike the Gell-Mann--Oakes--Renner relation, the Vafa--Witten
2720: theorem is not based on a symmetry(-breaking) argument, but depends on the
2721: {\it dynamics} of the theory. In the field theory, the proof relies
2722: on the reality and positivity of the fermion determinant that arises
2723: in the path integral formulation; in the holographic version, it
2724: relies on details of the dynamics of multiple D6-branes.
2725: 
2726: One result of our analysis of the $\nf>1$ case is the existence of
2727: $\nf^2$ independent, massless fluctuations of the D6-branes if $\mq=0$.
2728: It is tempting to interpret the associated $\nf^2$ massless pseudo-scalar
2729: particles in the spectrum of the gauge theory as the Goldstone bosons
2730: of a spontaneously broken $U(\nf)_\mt{A}$ symmetry, to which the
2731: $\ua$ symmetry would putatively be enhanced for $\nf >1$. However,
2732: only one of the $\nf^2$ states above is a true Goldstone boson,
2733: for two reasons. First, the five-dimensional theory contains a
2734: term of the form $\bar\psi_i \, X \, \psi^i$, where $i=1, \ldots, \nf$
2735: is the flavour index. Since $X$ only transforms under the $\ua$
2736: subgroup of $U(\nf)_\mt{A}$, the theory is only invariant under
2737: $\ua \times U(\nf)_\mt{V}$, which is spontaneously broken to
2738: $U(\nf)_\mt{V}$. Although $X$ is a massive field that can
2739: be integrated out, its mass $M_X \sim M_\mt{KK} \sim \Lambda\mt{QCD}$
2740: is of the same order as the strong coupling scale, so its integration
2741: should not lead to a real suppression of the symmetry-breaking operator
2742: above.
2743: 
2744: The second reason that makes implausible the interpretation of all
2745: but one of the massless modes above as Goldstone bosons is that
2746: they have non-derivative interactions. Indeed, the effective,
2747: interacting Lagrangian for these modes is obtained by expanding
2748: the non-Abelian Dirac--Born--Infeld action for the $\nf$ D6-branes
2749: \cite{Myers99} beyond quadratic order and then dimensionally
2750: reducing it to four dimensions, as we did in Section 4 in the
2751: $\nf=1$ case. Since the action for the D6-branes contains
2752: non-derivative interactions, such as commutator terms squared, so
2753: does the effective four-dimensional Lagrangian. These terms are
2754: only absent for the mode that multiplies the identity generator of
2755: $U(\nf)$. The caveat as to why this argument makes the Goldstone boson
2756: interpretation for the other modes implausible, but not strictly
2757: impossible, is that the coefficients in the effective Lagrangian
2758: might conspire so that, once all contributing diagrams are
2759: included, only momentum-dependent parts survive in any scattering
2760: amplitude (as in the linear sigma-model for the pion Lagrangian
2761: \cite{DGH92}). This possibility seems rather unlikely.
2762: 
2763: Thus we conclude that the string description predicts the existence
2764: of $\nf^2$ massless pseudo-scalar particles in the spectrum of the
2765: quantum gauge theory in the large-$\nc$ limit. Once sub-leading
2766: effects in the $1/\nc$ expansion are included, we expect these
2767: particles to acquire a small mass. The $\ua$ symmetry is anomalous at
2768: finite $\nc$, so its associated Goldstone boson is not exactly
2769: massless at finite $\nc$; presumably, this anomaly can be analysed in
2770: the supergravity description along the lines of 
2771: \cite{BDFLM01,KOW02,BDFLM02,GHP03}.
2772: The other $\nf^2 -1$ modes are likely to acquire a mass due to closed
2773: string interactions between the D6-branes, which are a
2774: $g_s \sim 1/\nc$ effect. From the gauge theory viewpoint, we have found
2775: no obvious reason for the masslessness (or lightness, at finite $\nc$)
2776: of these $\nf^2 -1$ pseudo-scalars, since we have argued that they are not
2777: Goldstone bosons.\footnote{Other examples with `unexpectedly' light scalars in
2778: the context of AdS/CFT have been discussed in \cite{Strassler03}.}
2779: Recall that, from the string viewpoint, the reason for this is that
2780: the only difference between the non-Abelian and the Abelian DBI
2781: actions consists exclusively of terms that involve commutators of
2782: fields. This can be regarded as a consequence of the fact that the
2783: seven-dimensional D6-brane action is T-dual to the ten-dimensional
2784: action of D9-branes; all commutators of scalar fields in the D6-brane
2785: action originate from commutators of gauge fields in the non-Abelian field
2786: strength on the D9-branes. Thus, it is {\it ten}-dimensional $U(\nf)$
2787: gauge invariance that seems to be responsible, from the string
2788: viewpoint, for the lightness of the gauge theory scalars.\footnote{
2789: This observation is due to M.\ Strassler.} Given the
2790: importance in high energy physics of mechanisms that naturally allow
2791: small scalar masses, it would be interesting to understand how to
2792: rephrase the constraints imposed by ten-dimensional $U(\nf)$
2793: gauge invariance purely in four-dimensional field theory terms.
2794: 
2795: 
2796: 
2797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2798: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2799: %%%%%%%%%%%%%%%%%%%%  SECTION   %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2801: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2802: 
2803: 
2804: %\section*{Acknowledgements}
2805: \acknowledgments
2806: 
2807: We are especially grateful to J.\ Brodie for helpful
2808: conversations. We also wish to thank J.\ Donoghue, J.\ Erdmenger,
2809: J.\ I.\ Latorre, A.\ E.\ Nelson, D.\ T.\ Son, M.\ J.\ Strassler,
2810: S.\ Thomas, D.\ Tong and A.\ Uranga for discussions and comments. 
2811: Research at
2812: the Perimeter Institute is supported in part by funds from NSERC
2813: of Canada. RCM is further supported by an NSERC Discovery grant,
2814: DJW by Fonds FCAR du Qu\'ebec and by a McGill
2815: Major Fellowship, while MK is supported in part by NSF grants
2816: PHY-0331516, PHY99-73935 and DOE grant DE-FG02-92ER40706. 
2817: DJW wishes to thank the University of Waterloo Physics
2818: Department for their ongoing hospitality and MK would like to thank
2819: the Perimeter Institute for Theoretical Physics for partial
2820: support and hospitality while this work was being performed.
2821: 
2822: 
2823: \begin{thebibliography}{99}
2824: 
2825: \bibitem{Maldacena97}
2826: J.\ M.\ Maldacena, {\it The large N limit of superconformal field
2827: theories and supergravity}, \atmp{2}{1998}{231}
2828: [\ijtp{38}{1998}{1113}], \hepth{9711200}.
2829: 
2830: \bibitem{Witten98b}
2831: E.\ Witten, {\it Anti-de Sitter space, thermal phase transition,
2832: and confinement in gauge theories}, \atmp{2}{1998}{505},
2833: \hepth{9803131}.
2834: 
2835: \bibitem{COOT98}
2836: C.\ Csaki, H.\ Oouguri, Y.\ Oz and J.\ Terning,
2837: {\it Glueball mass spectrum from supergravity},
2838: \jhep{01}{1999}{017}, \hepth{9806021};
2839: R.\ de Mello Koch, A.\ Jevicki, M.\ Mihailescu and J.\ P.\ Nunes,
2840: {\it Evaluation of glueball masses from supergravity},
2841: \prd{58}{1998}{105009}, \hepth{9806125};
2842: H.\ Ooguri, H.\ Robins and J.\ Tannenhauser,
2843: {\it Glueballs and their Kaluza-Klein cousins},
2844: \plb{437}{1998}{77}, \hepth{9806171};
2845: J.\ A.\ Minahan,
2846: {\it Glueball mass spectra and other issues for supergravity duals
2847: of QCD models}, \jhep{01}{1999}{020}, \hepth{9811156};
2848: N.\ R.\ Constable and R.\ C.\ Myers,
2849: {\it Spin-two glueballs, positive energy theorems and
2850: the AdS/CFT correspondence}, \jhep{10}{1999}{037},
2851: \hepth{9908175};
2852: R.\ C.\ Brower, S.\ D.\ Mathur and C.\ Tan,
2853: {\it Glueball spectrum for QCD from AdS supergravity duality},
2854: \npb{587}{2000}{249}, \hepth{0003115}.
2855: 
2856: \bibitem{AFM98}
2857: O.\ Aharony, A.\ Fayyazuddin and J.\ M.\ Maldacena,
2858: {\it The large-N limit of ${\cal N} =2,1$ field theories
2859: from threebranes in F-theory}, \jhep{07}{1998}{013},
2860: \hepth{9806159}.
2861: 
2862: \bibitem{KR01}
2863: A.\ Karch and L.\ Randall,
2864: {\it Open and closed string interpretation of SUSY CFT's on
2865: branes with boundaries}, \jhep{06}{2001}{063}, \hepth{0105132}.
2866: 
2867: \bibitem{BDFLM01}
2868: M.\ Bertolini, P.\ Di Vecchia, M.\ Frau, A.\ Lerda, and R.\ Marotta,
2869: {\it N=2 Gauge theories on systems of fractional D3/D7 branes},
2870: \npb{621}{2002}{157}, \hepth{0107057}.    
2871: 
2872: \bibitem{BDFM01}
2873: M.\ Bertolini, P.\ Di Vecchia, G.\ Ferretti, and R.\ Marotta,
2874: {\it Fractional Branes and N=1 Gauge Theories},
2875: \npb{630}{2002}{222}, \hepth{0112187}.
2876: 
2877: \bibitem{KK02}
2878: A.\ Karch and E.\ Katz, {\it Adding flavor to AdS/CFT},
2879: \jhep{06}{2002}{043}, \hepth{0205236}.
2880: 
2881: \bibitem{KKW02}
2882: A.~Karch, E.~Katz and N.~Weiner, {\it Hadron masses and screening
2883: from AdS Wilson loops}, \prl{90}{2003}{091601}, \hepth{0211107}.
2884: 
2885: \bibitem{WH03}
2886: X.~J.~Wang and S.~Hu, {\it Intersecting branes and adding flavors
2887: to the Maldacena-Nunez background}, \jhep{09}{2003}{017},
2888: \hepth{0307218}.
2889: 
2890: \bibitem{Ouyang03}
2891: P.~Ouyang, {\it Holomorphic D7-branes and flavored N = 1 gauge
2892: theories}, \hepth{0311084}.
2893: 
2894: \bibitem{KMMW03}
2895: M.\ Kruczenski, D.\ Mateos, R.\ C.\ Myers and D.\ J.\ Winters,
2896: {\it Meson spectroscopy in AdS/CFT with flavour},
2897: \jhep{07}{2003}{049}, \hepth{0304032}.
2898: 
2899: \bibitem{SS03}
2900: T.\ Sakai and J.\ Sonnenschein,
2901: {\it Probing flavored mesons of confining gauge theories by
2902: supergravity}, \jhep{09}{2003}{047}, \hepth{0305049}.
2903: 
2904: \bibitem{BEEGK03}
2905: J.\ Babington, J.\ Erdmenger, N.\ Evans, Z.\ Guralnik and I.\ Kirsch,
2906: {\it Chiral symmetry breaking and pions in non-supersymmetric
2907: gauge/gravity duals}, \hepth{0306018}.
2908: 
2909: \bibitem{NPR03}
2910: C.\ Nunez, A.\ Paredes and A.\ V.\ Ramallo,
2911: {\it Flavoring the gravity dual of $\caln=1$ Yang-Mills with probes},
2912: \jhep{12}{2003}{024}, \hepth{0311201}.
2913: 
2914: \bibitem{IMSY98}
2915: N.\ Itzhaki, J.\ M.\ Maldacena, J.\ Sonnenschein and S.\ Yankielowicz,
2916: {\it Supergravity and the large N limit of theories with sixteen
2917: supercharges}, \prd{58}{1998}{046004}, \hepth{9802042}.
2918: 
2919: \bibitem{DFO01}
2920: O.\ DeWolfe, D.\ Z.\ Freedman and H.\ Ooguri,
2921: {\it Holography and Defect Conformal Field Theories},
2922: \prd{66}{2002}{025009}, \hepth{0111135}.
2923: 
2924: \bibitem{GMOR68}
2925: M.\ Gell-Mann, R.\ J.\ Oakes and B.\ Renner,
2926: {\it Behavior of current divergences under $SU(3) \times SU(3)$},
2927: \pr{175}{1968}{2195}.
2928: 
2929: \bibitem{VW84}
2930: C.\ Vafa and E.\ Witten, {\it Restrictions on symmetry breaking
2931: in vector-like gauge theories}, \npb{234}{1984}{173}.
2932: 
2933: \bibitem{SVZ79}
2934: M.\ A.\ Shifman, A.\ I.\ Vainshtein and V.\ I.\ Zakharov,
2935: {\it QCD and resonance physics. Sum rules.},
2936: \npb{147}{1979}{385}.
2937: 
2938: \bibitem{soliton}
2939: G.~T.~Horowitz and R.~C.~Myers, {\it The AdS/CFT correspondence
2940: and a new positive energy conjecture for  general relativity},
2941: Phys.\ Rev.\ {\bf D 59} (1999) 026005, \hepth{9808079}.
2942: 
2943: \bibitem{adsprop}
2944: U.~H.~Danielsson, E.~Keski-Vakkuri and M.~Kruczenski, {\it Vacua,
2945: propagators, and holographic probes in AdS/CFT}, JHEP {\bf 9901}
2946: (1999) 002, \hepth{9812007}.
2947: 
2948: \bibitem{Myers99}
2949: R.\ C.\ Myers, {\it Dielectric-Branes}, \jhep{12}{1999}{022},
2950: \hepth{9910053}.
2951: 
2952: \bibitem{fairies}
2953: R.\ Dijkgraaf, J.\ M.\ Maldacena, G.\ W.\ Moore and E.\ Verlinde,
2954: {\it A black hole farey tail}, \hepth{0005003}.
2955: 
2956: \bibitem{phases}
2957: R.\ C.\ Myers, {\it Stress tensors and Casimir energies
2958: in the AdS/CFT correspondence}, \prd{60}{1999}{046002},
2959: \hepth{9903203}; 
2960: S.\ Surya, K.\ Schleich and D.\ M.\ Witt,
2961: {\it Phase transitions for flat adS black holes},
2962: \prl{86}{2001}{5231}, \hepth{0101134}.
2963: 
2964: \bibitem{break}
2965: S.\ J.\ Rey, S.\ Theisen and J.\ T.\ Yee,
2966: {\it Wilson-Polyakov loop at finite temperature in large 
2967: N gauge theory and anti-de Sitter supergravity}, 
2968: \npb{527}{1998}{171}, \hepth{9803135}.
2969: 
2970: \bibitem{ST02}
2971: K.\ Skenderis and M.\ Taylor,
2972: {\it Branes in AdS and pp-wave spacetimes},
2973: \jhep{06}{2002}{025}, \hepth{0204054}.
2974: 
2975: \bibitem{MNT02}
2976: D.\ Mateos, S.\ Ng and P.\ K.\ Townsend,
2977: {\it Supersymmetric Defect Expansion in CFT from AdS Supertubes},
2978: \jhep{07}{2002}{048}, \hepth{0207136}.
2979: 
2980: \bibitem{DGH92}
2981: J.\ F.\ Donoghue, E.\ Golowich and B.\ R.\ Holstein,
2982: {\it Dynamics of the standard model},
2983: Cambridge University Press, 1992.
2984: 
2985: \bibitem{KOW02}
2986: I.\ R.\ Klebanov, P.\ Ouyang and E.\ Witten,
2987: {\it A gravity dual of the chiral anomaly},
2988: \prd{65}{2002}{105007}, \hepth{0202056}.
2989: 
2990: \bibitem{BDFLM02}
2991: M.\ Bertolini, P.\ Di Vecchia, M.\ Frau, A.\ Lerda and R.\ Marotta,
2992: {\it More Anomalies from Fractional Branes}, \plb{540}{2002}{104},
2993: \hepth{0202195}.
2994: 
2995: \bibitem{GHP03}
2996: U.\ Gursoy, S.\ A.\ Hartnoll and R.\ Portugues,
2997: {\it The chiral anomaly from M theory},
2998: \hepth{0311088}.
2999: 
3000: \bibitem{Strassler03}
3001: M.\ J.\ Strassler, {\it Non-supersymmetric theories with light scalar
3002: fields and large hierarchies}, \hepth{0309122}.
3003: 
3004: 
3005: \end{thebibliography}
3006: \end{document}
3007: