1: % JHEP.cls available at http://jhep.cern.ch/JOURNAL/tex.html
2: \documentclass[12pt,nohyper,letterpaper]{JHEP3} % For preprint
3: %\documentclass[12pt,draft,nohyper]{JHEP3} % For draft
4: \usepackage{amssymb,amsfonts}
5: %\usepackage{draftfil}
6: %\usepackage{showkeys}
7: %\usepackage{epsfig}
8: %gives names of refs and cites
9: %If you do not have the msbm fonts, delete the following 10 lines
10: \font\mybb=msbm10 at 10pt
11: %\font\mybb=msbm12 at 12pt
12:
13: \renewcommand{\arraystretch}{1.2}
14:
15: \usepackage{epsfig}
16: %\usepackage{amssymb}
17:
18: %\newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
19: %\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
20: %\def\appendix{{\newpage\section*{Appendix}}\let\appendix\section%
21: % {\setcounter{section}{0}
22: % \gdef\thesection{\Alph{section}}}\section}
23:
24:
25: %%%%%%%%%%%%%%%%%%%%%%%%%% My Macros %%%%%%%%%%%%%%%%%%%%%%%%%
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27:
28: \def\be{\begin{eqnarray}}
29: \def\ee{\end{eqnarray}}
30: \newcommand{\nn}{\nonumber}
31: \newcommand\para{\paragraph{}}
32: \newcommand{\ft}[2]{{\textstyle\frac{#1}{#2}}}
33: \newcommand{\eqn}[1]{(\ref{#1})}
34: \newcommand{\pl}[1]{\frac{\partial L}{\partial{#1}}}
35: \newcommand{\ppp}[2]{\frac{\partial {#1}}{\partial {#2}}}
36: \newcommand{\ph}[1]{\frac{\partial H}{\partial{#1}}}
37:
38: \newcommand\balpha{\mbox{\boldmath $\alpha$}}
39: \newcommand\bbeta{\mbox{\boldmath $\beta$}}
40: \newcommand\bgamma{\mbox{\boldmath $\gamma$}}
41: \newcommand\bomega{\mbox{\boldmath $\omega$}}
42: \newcommand\blambda{\mbox{\boldmath $\lambda$}}
43: \newcommand\bmu{\mbox{\boldmath $\mu$}}
44: \newcommand\bphi{\mbox{\boldmath $\phi$}}
45: \newcommand\bzeta{\mbox{\boldmath $\zeta$}}
46: \newcommand\bsigma{\mbox{\boldmath $\sigma$}}
47: \newcommand\bepsilon{\mbox{\boldmath $\epsilon$}}
48: \newcommand\btau{\mbox{\boldmath $\tau$}}
49: \newcommand\beeta{\mbox{\boldmath $\eta$}}
50: \newcommand\btheta{\mbox{\boldmath $\theta$}}
51: \newcommand\valpha{\vec{\alpha}}
52: \newcommand\vg{\vec{g}}
53:
54: \def\Dslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu D}}
55: \def\Dbarslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu {\bar D}}}
56: \def\delslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu \partial}}
57: \def\delbarslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu {\bar\partial}}}
58: \def\pslash{\,\,{\raise.15ex\hbox{/}\mkern-9mu p}}
59: \def\calDslash{\,\,{\raise.15ex\hbox{/}\mkern-12mu {\cal D}}}
60: \newcommand\Bprime{B${}^\prime$}
61: \newcommand{\sign}{{\rm sign}}
62: \newcommand{\D}{{\cal D}}
63:
64: \newcommand\bx{{\bf x}}
65: \newcommand\br{{\bf r}}
66: \newcommand\bF{{\bf F}}
67: \newcommand\bp{{\bf p}}
68: \newcommand\bL{{\bf L}}
69: \newcommand\bR{{\bf R}}
70: \newcommand\bP{{\bf P}}
71: \newcommand\bE{{\bf E}}
72: \newcommand\bB{{\bf B}}
73: \newcommand\bA{{\bf A}}
74: \newcommand\bee{{\bf e}}
75: \newcommand\bte{\tilde{\bf e}}
76: \newcommand{\NN}{{\mathbb N}}
77: \newcommand{\Z}{{\bf Z}}
78: \newcommand{\ZZ}{{\mathbb Z}}
79: \newcommand{\Q}{{\bf Q}}
80: \newcommand{\QQ}{{\mathbb Q}}
81: \newcommand{\R}{{\bf R}}
82: \newcommand{\RR}{{\mathbb R}}
83: \newcommand{\C}{{\bf C}}
84: \newcommand{\CC}{{\mathbb C}}
85: \newcommand{\PP}{{\mathbb P}}
86: \newcommand{\e}{\,{\rm e}}
87: \newcommand{\CP}{{\bf CP}}
88: \newcommand{\bra}{\langle}
89: \newcommand{\ket}{\rangle}
90: \newcommand{\Tr}{{\rm Tr}}
91: \newcommand{\N}{{\cal N}}
92: \newcommand{\hW}{\hat{\cal W}}
93:
94:
95: \def\implies{\Rightarrow}
96: \def\To{\longrightarrow}
97: \def\longlongrightarrow{\relbar\joinrel\relbar\joinrel\rightarrow}
98: \def\ridiculousrightarrow{\relbar\joinrel\relbar\joinrel\relbar%
99: \joinrel\rightarrow}
100:
101: \def\underarrow#1{\mathrel{\mathop{\longrightarrow}\limits_{#1}}}
102: \def\onnearrow#1{\mathrel{\mathop{\nearrow}\limits^{#1}}}
103: \def\undernearrow#1{\mathrel{\mathop{\nearrow}\limits_{#1}}}
104: \def\onarrow#1{\mathrel{\mathop{\longrightarrow}\limits^{#1}}}
105: \def\onArrow#1{\mathrel{\mathop{\longlongrightarrow}\limits^{#1}}}
106: \def\OnArrow#1{\mathrel{\mathop{\ridiculousrightarrow}\limits^{#1}}}
107: \def\lae{\mathrel{\mathop{\smash{\lower .5 ex \hbox{$\stackrel<\sim$}}}}}
108: \def\lae{\mathrel{\mathop{\smash{\lower .5 ex \hbox{$\stackrel>\sim$}}}}}
109: \def\eqq{\stackrel?=}
110:
111: \def\theequation{\thesection.\arabic{equation}}
112:
113: %%%%%%%%%%%%%%%%%%%%% NOTES START %%%%%%%%%%%%%%%%%%%%%%%%%%%
114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
115:
116:
117:
118:
119:
120:
121:
122: \title{Heterotic Vortex Strings}
123: \author{Mohammad Edalati\\
124: Department of Physics, \\
125: University of Cincinnati, \\
126: P. O. Box 210011, \\
127: Cincinnati, OH 45221-0011, USA \\
128: {\tt edalati@physics.uc.edu}}
129:
130: \author{David Tong\\
131: Department of Applied Mathematics and Theoretical Physics, \\
132: University of Cambridge, \\
133: Cambridge, CB3 0WA, UK\\{\tt d.tong@damtp.cam.ac.uk}}
134:
135: \abstract{We determine the low-energy ${\cal N}=(0,2)$ worldsheet
136: dynamics of vortex strings in a large class of non-Abelian ${\cal
137: N}=1$ supersymmetric gauge theories.}
138:
139:
140: \begin{document}
141: \pagestyle{plain} \setcounter{page}{1}
142: \newcounter{bean}
143: \baselineskip16pt \setcounter{section}{0}
144:
145:
146:
147: \tableofcontents
148: \section{Introduction}
149:
150: Vortex strings provide a map between four-dimensional non-Abelian
151: gauge theories and two-dimensional sigma-models. The
152: four-dimensional theories in question have a $U(N_c)$ gauge group
153: and a sufficient number of scalar fields to allow complete gauge
154: symmetry breaking, so that the system lies in the Higgs phase.
155: Theories with this property admit vortex strings. The embedding of
156: the vortex within the non-Abelian gauge group endows the string
157: with a number of orientation modes which parameterize the complex
158: projective space ${\bf CP}^{N_c-1}$. Further bosonic and fermionic
159: zero modes of the vortex live in line bundles over ${\bf
160: CP}^{N_c-1}$. In this manner, the low-energy dynamics of a single,
161: straight, infinite vortex string is described by some variant of
162: the ${\bf CP}^{N_c-1}$ sigma-model living on the $d=1+1$
163: dimensional worldsheet \cite{vib,auzzi}.
164:
165: \para
166: When the four-dimensional gauge theory has ${\cal N}=2$
167: supersymmetry, a pleasing story emerges. The strings are
168: $1/2$-BPS, ensuring that the worldsheet dynamics inherits ${\cal
169: N}=(2,2)$ supersymmetry. It was shown in \cite{sy,vstring},
170: following earlier work of \cite{nick,dht}, that the quantum
171: dynamics of the worldsheet theory encodes quantitative information
172: about the quantum dynamics of the parent four-dimensional theory,
173: including the Seiberg-Witten curve and the exact BPS mass
174: spectrum. More recently, the correspondence was extended to
175: superconformal points, with a matching between the scaling
176: dimensions of chiral primary operators in the four-dimensional
177: bulk and on the worldsheet \cite{scvs}. For a review of the
178: classical and quantum dynamics of these strings, see \cite{tasi}.
179:
180: \para
181: The purpose of this paper is to present a detailed study of the
182: classical dynamics of vortex strings in ${\cal N}=1$
183: four-dimensional gauge theories. For certain choices of parameters
184: the strings once again preserve $1/2$ of supersymmetry, now
185: guaranteeing ${\cal N}=(0,2)$ supersymmetry on the worldsheet. For
186: this reason, we refer to vortices in ${\cal N}=1$ theories as
187: ``heterotic vortex strings". We will determine the explicit ${\cal
188: N}=(0,2)$ ${\bf CP}^{N_c-1}$ sigma-models, and their variations,
189: which describe the low-energy dynamics of vortex strings in a
190: large class of ${\cal N}=1$ gauge theories\footnote{Vortex strings
191: in various non-Abelian theories with less supersymmetry were
192: previously studied in [9-13]
193: %\cite{other,others,bolog,syss,gsy}
194: and in some cases qualitative agreement was found between the
195: dynamics of the worldsheet theory and the bulk. We will comment
196: more on the relationship of our work to some of these papers in
197: Section 4.}.
198:
199: \para
200: The paper is organized as follows: Section 2 contains a detailed
201: discussion of the ${\cal N}=(2,2)$ worldsheet dynamics of vortex
202: strings in ${\cal N}=2$ four-dimensional theories. This section is
203: mostly a review of previous work, although explicit expressions
204: for bosonic and fermionic zero modes are provided which generalize
205: results in the literature from $U(2)$ gauge theories to $U(N_c)$
206: gauge theories. Particular attention is paid to the chirality of
207: different fermionic zero modes since this will prove important in
208: later sections. Section 3 also contains review material,
209: describing the basics of the superfield formalism for ${\cal
210: N}=(0,2)$ supersymmetry in $d=1+1$ dimensions.
211:
212: \para
213: The meat of the paper is in Section 4. We consider two different
214: classes of deformations, each of which breaks the
215: four-dimensional supersymmetry from ${\cal N}=2$ to ${\cal N}=1$
216: through the introduction of a superpotential for the adjoint
217: chiral multiplet. In each case, we show that there is a unique
218: ${\cal N}=(0,2)$ worldsheet theory which correctly captures all
219: BPS properties of the vortex string and predicts the interaction
220: of fermionic zero modes. We also include an appendix which
221: collates the notation for bulk and worldsheet fields used
222: throughout the paper.
223:
224:
225:
226:
227:
228:
229: \section{The ${\cal N}=(2,2)$ Dynamics of Vortex Strings}
230:
231: In this section we review the dynamics of vortex strings in
232: four-dimensional gauge theories with ${\cal N}=2$ supersymmetry.
233: The vortices are $1/2$-BPS, ensuring that the $d=1+1$ dimensional
234: worldsheet dynamics of the string inherits ${\cal N}=(2,2)$
235: supersymmetry.
236:
237: \subsection{The Four-Dimensional Theory}
238:
239: Our starting point is the $d=3+1$, ${\cal N}=2$ supersymmetric
240: $U(N_c)$ gauge theory, with $N_f$ flavors transforming in the
241: fundamental representation\footnote{Conventions: We pick Hermitian
242: generators $T^m$ with Killing form $\Tr\, T^mT^n= \ft12
243: \delta^{mn}$. We write the gauge field as $A_\mu=A_\mu^mT^m$ and
244: $F_{\mu\nu}=\partial_\mu A_\nu-\partial_\nu A_\mu -
245: i[A_\mu,A_\nu]$. Fundamental covariant derivatives are ${\cal
246: D}_\mu Q=\partial_\mu Q-iA_\mu Q$; adjoint covariant derivatives
247: are ${\cal D}_\mu A =
248: \partial_\mu A - i [A_\mu,A]$. Our summation
249: conventions are inconsistent: a sum over repeated indices is
250: usually left implicit unless there is some ambiguity or a point
251: that requires emphasis.}. We describe the theory in the language
252: of four-dimensional ${\cal N}=1$ superfields. The ${\cal N}=2$
253: vector multiplet consists of an ${\cal N}=1$ vector multiplet $V$
254: and an ${\cal N}=1$ adjoint chiral multiplet $A$. Similarly, each
255: flavor hypermultiplet splits into two chiral multiplets, $Q_i$ and
256: $\tilde{Q}_i$ where $i=1,\ldots, N_f$ is the flavor index. Each
257: $Q_i$ transforms in the fundamental ${\bf N}_c$ of the gauge
258: group, while each $\tilde{Q}_i$ transforms in the anti-fundamental
259: $\bar{\bf N}_c$. We denote the complexified gauge coupling of the
260: theory as
261: %
262: \be \tau=\frac{2\pi i}{e^2} + \frac{\theta}{2\pi}\ .
263: \label{tau}\ee
264: %
265: The four
266: dimensional theory has the usual superpotential required for
267: ${\cal N}=2$ supersymmetry,
268: %
269: \be{\cal W}_{{\cal
270: N}=2}=\sqrt{2}\,\sum_{i=1}^{N_f}\tilde{Q}_iAQ_i\ .\label{4dn2}\ee
271: %
272: The scalar potential of the theory is dictated by the D-term and
273: the F-terms arising from this superpotential. In components it is
274: given by,
275: %
276: \be V_{4d}&=& \frac{e^2}{2}\Tr(\,\sum_{i=1}^{N_f}\,Q_iQ_i^\dagger
277: - \tilde{Q}_i\tilde{Q}_i^\dagger - v^2\,1_{N_c})^2 +e^2\Tr|\,
278: \sum_{i=1}^{N_f}\tilde{Q}_iQ_i|^2\nn\\
279: &&+\sum_{i=1}^{N_f}\left(Q_i^\dagger \{A,A^\dagger\}Q_i +
280: \tilde{Q}_i\{A,A^\dagger\}\tilde{Q}_i^\dagger\right) +
281: \frac{1}{2e^2}\Tr|[A,A^\dagger]|^2\label{v4dn2}\ee
282: %
283: where we have taken the liberty of denoting the component scalar
284: fields by the same Roman letter as the superfield in which they
285: reside. We have included a D-term Fayet-Iliopoulos (FI) parameter
286: $v^2$ for the central $U(1)\subset U(N_c)$. This is consistent
287: with ${\cal N}=2$ supersymmetry and forces the theory into the
288: Higgs phase, with $Q_i$ gaining a vacuum expectation value (vev).
289: For $N_f<N_c$, the rank condition ensures the D-term cannot vanish
290: and there is no supersymmetric ground state. We do not consider
291: this case. When $N_f>N_c$, the D-term and F-term conditions do not
292: fix the vevs of $Q_i$ and $\tilde{Q}_i$ completely and there is a
293: Higgs branch of vacua; we shall discuss this situation in Section
294: \ref{add}. For now we restrict attention to the case $N_f=N_c$ for
295: which there is a unique supersymmetric ground state in which the
296: gauge group is completely broken. Up to a gauge transformation the
297: ground state is given by
298: %
299: \be Q^a_{\ i}=v\delta^a_{\ i}\ \ \ ,\ \ \
300: \tilde{Q}_i=A=0\label{vac1}\ee
301: %
302: where $a=1,\ldots, N_c$ is the color index. The theory lies in the
303: color-flavor locked phase, with the vacuum expectation value
304: preserved by a simultaneous gauge and flavor rotation. The
305: symmetry breaking pattern is thus broken to the diagonal
306: combination of the two (recall that we are looking at the theory with $N_f=N_c$)
307: %
308: \be U(N_c)\times SU(N_f)\rightarrow SU(N_c)_{\rm
309: diag}\ .\label{breaking}\ee
310: %
311:
312:
313:
314: \subsection{The Vortex}
315:
316: The central $U(1)\subset U(N_c)$ does not survive the symmetry
317: breaking \eqn{breaking}, a fact which provides sufficient topology
318: to ensure the presence of vortex strings in the theory \cite{no}.
319: These vortices preserve $1/2$ of the supersymmetry, ensuring
320: ${\cal N}=(2,2)$ supersymmetric dynamics on their $d=1+1$
321: dimensional worldvolume. Infinite, straight strings oriented in
322: the $x^3$ direction satisfy the first-order equations,
323: %
324: \be F_{12}=e^2(\sum_{i=1}^{N_f}Q_iQ_i^\dagger - v^2\,1_{N_c})\nn\\
325: {\cal D}_{{z}}Q_i\equiv \frac{1}{2}({\cal D}_1Q_i-i{\cal D}_2Q_i)
326: =0 \label{vort}\ee
327: %
328: where $z=x^1+ix^2$ parameterizes the transverse plane.
329: %\footnote{Anti-vortices satisfies
330: %$F_{12}=-e^2(\sum_{i=1}^{N_f}Q_iQ_i^\dagger - v^2)$ and ${\cal
331: %D}_{\bar{z}}Q_i=0$.}.
332: These are the non-Abelian vortex equations.
333: %(For example, each
334: %side of the first equation is an $N_c\times N_c$ matrix).
335: Solutions to these equations have tension
336: %
337: \be T_k=2\pi k v^2 \ee
338: %
339: where $k =-\Tr\,\int (F_{12}/2\pi) \in {\bf Z}^+$ is the winding
340: number.
341:
342:
343: \para
344: \EPSFIGURE{profile.eps,height=110pt}{} Solutions to the vortex
345: equations with winding number $k$ have $2kN$ bosonic collective
346: coordinates. For a single $k=1$ vortex, they break down as
347: follows: there are $2$ collective coordinates corresponding to the
348: position of the string in the $z=x^1+ix^2$ plane. The remaining
349: $2(N-1)$ collective coordinates are Goldstone modes arising from
350: the action of the surviving symmetry \eqn{breaking} on the vortex
351: string. They parameterize $SU(N_c)/[SU(N_c-1)\times U(1)] \cong
352: {\bf CP}^{N_c-1}$ \cite{vib,auzzi}.
353:
354: \para
355: An explicit realization of the orientational modes is most simply
356: given in {\it singular gauge} in which $Q$ does not wind
357: asymptotically, with the flux instead arising from a singular
358: gauge potential \cite{auzzi}. Suppose that the Abelian $N_c=1$
359: vortex equations are solved by two profile functions $q(\rho)$ and
360: $a(\rho)$, where $\rho=\sqrt{(x^1)^2+(x^2)^2}$ is the radial
361: distance from the string
362: %
363: \be Q_{\rm Abelian} = v q(\rho)\ \ \ \ {\rm and}\ \ \ \ (A_z)_{\rm
364: Abelian}= -i\bar{z} a(\rho)\ . \ee
365: %
366: Here the complexified gauge connection is $A_z=\ft12 (A_1-iA_2)$.
367: Plugging this ansatz into the vortex equations gives two first
368: order ordinary differential equations,
369: %
370: \be q'=2\rho aq\ \ \ {\rm and} \ \ \ 4a+2\rho a'=e^2v^2
371: (q^2-1)\ee
372: %
373: with prime denoting the derivative with respect to $\rho$. These
374: equations are known to admit a unique solution satisfying the
375: appropriate boundary conditions,
376: %
377: \be q(\rho)\rightarrow \left\{\begin{array}{c} 1 \\ 0
378: \end{array}\right.\ \ \ \ ,\ \ \ \ \ a(\rho)\rightarrow
379: \left\{\begin{array}{lc} 0 \ \ \ \ \ & {\rm as}\ \rho\rightarrow
380: \infty
381: \\ 1/2\rho^2 & {\rm as}\ \rho\rightarrow 0\ .
382: \end{array}\right.\ee
383: %
384: However, the solution does not have a simple analytic form. A
385: sketch of the profile $q(\rho)$ is shown in figure 1.
386:
387: \para
388: With the $k=1$ Abelian vortex solution in hand, one may simply
389: construct a solution to the non-Abelian equations by embedding
390: thus,
391: %
392: \be Q^a_{\ i}= \left(\frac{\phi^a\bar{\phi}_i}{r}\right)\,
393: v[q(\rho)-1]+v\delta^a_{\ i} \ \ \ \ {\rm and}\ \ \ \ (A_z)^a_{\
394: b} = -i\bar{z}a(\rho)\,\left(\frac{\phi^a\bar{\phi}_b}{r}\right) \
395: .\label{solutions}\ee
396: %
397: The $\phi^a \in {\bf C}^{N_c}$ define the orientation of the
398: vortex in the gauge and flavor groups. In order that this reduce
399: to the Abelian solution, we require
400: %
401: \be \sum_{a=1}^{N_c} |\phi^a|^2=r \label{phiphi}\ee
402: %
403: with $r$ a constant that will be fixed shortly. The solutions
404: \eqn{solutions} are invariant under the simultaneous rotation,
405: %
406: \be \phi^a\rightarrow e^{i\alpha}\phi^a\ .\label{iden}\ee
407: %
408: The $\phi^a$, subject to the constraint \eqn{phiphi} and
409: identification \eqn{iden}, provide homogeneous coordinates on
410: ${\bf CP}^{N_c-1}$. The $SU(N_c)$ symmetry of four-dimensions
411: descends to the vortex string, with the $\phi^a$ transforming in
412: the fundamental representation. This ensures that the ${\bf
413: CP}^{N_c-1}$ is endowed with the symmetric Fubini-Study metric.
414: The K\"ahler class of this space is $r$.
415:
416: \para
417: A comment on notation: since $N_f=N_c$, both $Q^a_{\ i}$ and
418: $(A_z)^a_{\ b}$ are $N_c\times N_c$ matrices. In what
419: follows, we shall often neglect to write the indices on both. In
420: this notation, $Q$ is a matrix on which gauge rotations $U\in
421: U(N_c)$ act from the left, while flavor rotations $V\in SU(N_f)$
422: act from the right, so that $Q\rightarrow UQV^\dagger$.
423:
424: \subsubsection{Bosonic Zero Modes}
425:
426: For general winding number $k$, the vortex zero modes are defined
427: to be solutions to the linearized vortex equations,
428: %
429: \be {\cal D}_z\delta A_{\bar{z}}-{\cal D}_{\bar{z}}\delta A_z &=&
430: \frac{ie^2}{2}(\delta Q Q^\dagger + Q\delta Q^\dagger) \nn\\
431: {\cal D}_z \delta Q &=& i\delta A_z Q\ .\ee
432: %
433: These are to be supplemented with a suitable gauge fixing
434: condition which is derived from Gauss' law and reads
435: %
436: \be {\cal D}_z\delta A_{\bar{z}}+{\cal D}_{\bar{z}}\delta A_z =
437: -\frac{ie^2}{2}(\delta Q Q^\dagger - Q\delta Q^\dagger)\ .\ee
438: %
439: This gauge fixing condition combines with the first of the
440: linearized vortex equations to leave us with two, complex, first
441: order equations to be solved around the background of a fixed
442: vortex configuration,
443: %
444: \be 2{\cal D}_{\bar{z}}\delta A_z &=& -ie^2\delta Q Q^\dagger
445: \nn\\ {\cal D}_z \delta Q &=& i\delta A_z Q\label{bogzero}.\ee
446: %
447: We now derive the solutions to these equations that arise from the
448: symmetries of the system.
449:
450:
451: \subsubsection*{Translational Mode}
452:
453: For any winding number $k$, the two translational modes are always
454: given by
455: %
456: \be \delta A_z=F_{\bar{z}z}\ \ \ {\rm and}\ \ \ \delta Q={\cal
457: D}_{\bar{z}} Q \label{transzero}\ee
458: %
459: which can be checked to satisfy \eqn{bogzero} using the fact that
460: the background fields obey the second order equations of motion.
461:
462: \subsubsection*{Orientational Modes}
463:
464: The zero modes corresponding to orientation are only slightly more
465: complicated. In general they can be written as
466: %
467: \be \delta A_z &=& {\cal D}_z\Omega \nn\\
468: \delta Q &=& i(\Omega Q - Q\hat{\Omega}).\label{orientzero}\ee
469: %
470: Here $\Omega(x)$ is an infinitesimal gauge rotation, while
471: $\hat{\Omega}$ is an infinitesimal flavor rotation. Since only the
472: diagonal subgroup \eqn{breaking} of these is preserved in the
473: vacuum, we require that $\Omega(x)\rightarrow \hat{\Omega}$ as
474: $x\rightarrow \infty$. In terms of our orientation coordinates
475: $\phi^i$, this diagonal rotation can be written as,
476: %
477: \be \hat{\Omega}^i_{\ j}=-i\left[\delta\phi^i\bar{\phi}_j - \phi^i
478: \delta \bar{\phi}_j - 2iu \phi^i\bar{\phi}_j\right]\ee
479: %
480: which holds for any $u$. Requiring that $\hat{\Omega} \in su(N_c)$
481: fixes $u$ to be
482: %
483: \be u = -i\bar{\phi}_i\,\delta\phi^i\ .\ee
484: %
485: Later $u$ will become a gauge field on the worldsheet whose role
486: is to implement the identification \eqn{iden}. For now, we can
487: treat $u$ as a connection and introduce the covariant variation
488: $\nabla\phi^i=\delta\phi^i-iu\phi^i$ which satisfies
489: $\nabla\phi^i\cdot\bar{\phi}_i=0$. In this notation
490: %
491: \be \hat{\Omega}^i_{\ j} =
492: -i[(\nabla\phi^i)\bar{\phi}_j-\phi^i\nabla\bar{\phi}_j]. \ee
493: %
494: The zero mode equations \eqn{bogzero} translate to the requirement
495: that $\Omega(x)$ satisfy the second order differential equation
496: %
497: \be {\cal D}^2\Omega = e^2 \left[\{\Omega,
498: QQ^\dagger\}-2Q\hat{\Omega} Q^\dagger\right].\label{hard}\ee
499: %
500: Everything above holds for arbitrary winding number $k$. For a
501: single vortex, with $k=1$, the solution to \eqn{hard} was provided
502: in \cite{others} (see equation (28) of that paper) and depends
503: only on the profile function $q(\rho)$ of the
504: vortex\footnote{Equation \eqn{osol} solves \eqn{hard} by virtue
505: of the vortex profile obeying the second order equation
506: $4\partial_z\partial_{\bar{z}}q-4a^2\rho^2q=e^2v^2q(q^2-1)$.}
507: %
508: \be \Omega(\rho) = q(\rho)\,\hat{\Omega}.\label{osol}\ee
509: %
510: Using the solution \eqn{osol}, we can now be more explicit about
511: the orientation zero modes for a single vortex. Making use of the
512: vortex equations \eqn{vort}, we find
513: %
514: \be (\delta A_z)^a_{\ b} &=& -2i(\partial_z q)
515: \,(\nabla\phi^a)\bar{\phi}_b \nn\\ \delta Q^a_{\ i} &=&
516: v(q^2-1)\,(\nabla\phi^a)\bar{\phi}_i.\ee
517: %
518:
519:
520: \subsection{Fermions}
521:
522: We now turn to a study of the fermionic zero modes \cite{jackr}. We start by
523: describing the Dirac equations in four-dimensions and their
524: solutions for a single $k=1$ vortex. We will pay particular
525: attention to the correlation between the chirality of the
526: worldsheet and four-dimensional fermions.
527:
528: \para
529: In the following we use four-dimensional Weyl fermions
530: $\psi_\alpha$ and $\bar{\lambda}^{\dot{\alpha}}$ with
531: $\alpha,\dot{\alpha}=1,2$. The notation is standard Wess and
532: Bagger fare \cite{wb} with, for example,
533: $\psi\lambda=\psi^\alpha\lambda_\alpha=\lambda\psi$ and
534: $\bar{\psi}\bar{\lambda}=\bar{\psi}_{\dot{\alpha}}\bar{\lambda}^{\dot{\alpha}}
535: =\bar{\lambda}\bar{\psi}$. Indices are raised and lowered with
536: $\epsilon^{\alpha\beta}=\epsilon^{\dot{\alpha}\dot{\beta}}=i\sigma_2$.
537: Our signature is mostly minus and we define
538: $(\sigma^\mu)_{\alpha\dot{\alpha}}=(-1,\sigma^i)$ and
539: $(\bar{\sigma}^\mu)^{\dot{\alpha}\alpha}=(-1,-\sigma^i)$.
540:
541: \para
542: The ${\cal N}=2$ vector multiplet in four dimensions contains two
543: Weyl fermions, $\lambda$ and $\eta$, each transforming in the
544: adjoint representation of the $U(N_c)$ gauge group. The fermion
545: $\lambda$ lives in the ${\cal N}=1$ vector multiplet while $\eta$
546: lives in the adjoint chiral multiplet $A$. Each hypermultiplet
547: also contains two Weyl fermions, $\psi$ and $\tilde{\psi}$. These
548: live in $Q$ and $\tilde{Q}$, and transform in the ${\bf N}_c$ and
549: $\bar{\bf N}_c$ representations respectively. The Dirac equations
550: in the ${\cal N}=2$ theory are
551: %\footnote{Each of these comes with its complex
552: %conjugate equation. Just to get conventions right, let me spell
553: %out the first of these which is
554: %
555: %\be
556: %-i\Dslash\bar{\lambda}-i\sqrt{2}[{\eta},a^\dagger]-i\sqrt{2}{\psi}_iq_i^\dagger+
557: %i\sqrt{2}\tilde{q}_i^\dagger\tilde{\psi}_i=0\ee.}
558: %
559: \be
560: -\frac{i}{e^2}\Dbarslash\lambda-\frac{i\sqrt{2}}{e^2}[\bar{\eta},A]+i\sqrt{2}Q_i\bar{\psi}_i-
561: i\sqrt{2}\bar{\tilde{\psi}}_i\tilde{Q}_i&=&0 \nn\\
562: -\frac{i}{e^2}\Dbarslash\eta-\frac{i\sqrt{2}}{e^2}
563: [A,\bar{\lambda}]-\sqrt{2}\tilde{Q}^\dagger_i\bar{\psi}_i-
564: \sqrt{2}\bar{\tilde{\psi}}_i{Q}_i^{\dagger}&=&0 \label{diracs}\\
565: -i\Dbarslash\psi_i+i\sqrt{2}\bar{\lambda}Q_i-\sqrt{2}A^\dagger\bar{\tilde{\psi}}_i
566: -\sqrt{2}\bar{\eta}\tilde{Q}_i^\dagger &=&0\nn\\
567: -i\Dbarslash\tilde{\psi}_i-i\sqrt{2}\tilde{Q}_i\bar{\lambda}-\sqrt{2}\bar{\psi}_i
568: A^\dagger -\sqrt{2}Q_i^{\dagger}\bar{\eta}&=&0. \nn\ee
569: %
570: We wish to study these equations in the background of the vortex.
571: Here they simplify considerably since we have $A=\tilde{Q}_i=0$.
572: The equations decouple into two pairs: the first set of equations
573: are for $\lambda$ and $\bar{\psi}_i$
574: %
575: \be -\frac{i}{e^2}\Dbarslash\lambda+i\sqrt{2}Q_i\bar{\psi}_i=0\ \
576: \ &{\rm and}&\ \ \ \
577: -i\Dslash\bar{\psi}_i-i\sqrt{2}Q_i^{\dagger}\lambda=0.
578: \label{lambdap}\ee
579: %
580: The second set of equations are for $\eta$ and
581: $\bar{\tilde{\psi}}_i$,
582: %
583: \be
584: -\frac{i}{e^2}\Dbarslash\eta-\sqrt{2}\bar{\tilde{\psi}}_iQ_i^{\dagger}=0\
585: \ \ &{\rm and}&\ \ \ -i\Dslash\bar{\tilde{\psi}}_i-\sqrt{2}\eta
586: Q_i=0.\label{etap}\ee
587: %
588:
589: \subsubsection{Chirality}
590:
591: Each pair of four-dimensional fermions gives rise to a fermi zero
592: mode on the vortex string of a specific chirality. Since this will
593: be important in later sections, we dwell on the point a little
594: here. The first step is to see which components of the spinors can
595: turn on in the background of a vortex or anti-vortex. We will need
596: the following identities,
597: %
598: \be
599: \Dslash_{\alpha\dot{\alpha}}\equiv(\sigma^\mu)_{\alpha\dot{\alpha}}
600: {\cal D}_\mu=2\left(\begin{array}{cc} -{\cal D}_- & {\cal D}_z \\
601: {\cal D}_{\bar{z}} & -{\cal D}_+\end{array}\right) \ \ {\rm and}\
602: \ \Dbarslash^{\dot{\alpha}\alpha}
603: \equiv(\bar{\sigma}^\mu)^{\dot{\alpha}\alpha}
604: {\cal D}_\mu=-2\left(\begin{array}{cc} {\cal D}_+ & {\cal D}_z \\
605: {\cal D}_{\bar{z}} & {\cal D}_-\end{array}\right)\label{ds}\ \ \ \
606: \ \ \ \ee
607: %
608: where ${\cal D}_\pm=\ft12({\cal D}_0\pm{\cal D}_3)$ and ${\cal
609: D}_z=\ft12({\cal D}_1-i{\cal D}_2)$ and ${\cal
610: D}_{\bar{z}}=\ft12({\cal D}_1+i{\cal D}_2)$. Our strings are
611: static and oriented in the $x^3$ direction, so in searching for
612: zero modes of the Dirac equation in the presence of a vortex we
613: may initially set ${\cal D}_\pm=0$. We decompose the spinors as
614: $(\lambda_1,\lambda_2)=(\lambda_-,\lambda_+)$ and
615: $(\lambda^1,\lambda^2)=(\lambda^-,\lambda^+)$ so that, with our
616: raising and lowering conventions, $\lambda^+=-\lambda_-$ and
617: $\lambda^-=\lambda_+$. To see which components turn on in the
618: background of the vortex, we act on the first and second equations
619: in \eqn{lambdap} with $\Dslash$ and $\Dbarslash$ respectively.
620: Making use of the vortex equation ${\cal D}_{{z}}Q_i=0$, we find
621: %
622: \be &&(-\frac{4}{e^2}{\cal D}_z{\cal
623: D}_{\bar{z}}+2Q_iQ_i^{\dagger})\lambda_- =0\nn\\&&
624: (-\frac{4}{e^2}{\cal D}_{\bar{z}}{\cal
625: D}_z+2Q_iQ_i^{\dagger})\lambda_+ - \sqrt{2}({\cal
626: D}_{\bar{z}}Q_i)\,\bar{\psi}_{+i}=0 \ee
627: %
628: and
629: %
630: \be &&(-{\cal D}_z{\cal D}_{\bar{z}}\delta_{ij}
631: +2Q_iQ_j^{\dagger})\bar{\psi}_{+j} - \sqrt{2}({\cal D}_z
632: Q_i^{\dagger})\, \lambda_+ =0\nn\\
633: &&(-{\cal D}_{\bar{z}}{\cal D}_z\delta_{ij}+2Q_i^{\dagger}
634: Q_j)\bar{\psi}_{-j} =0. \ee
635: %
636: %
637: %\be (-{\cal D}_z{\cal D}_{\bar{z}}+2Q_iQ^{i\dagger})\lambda_-
638: %-\sqrt{2}{\cal D}_zQ_i\,\bar{\psi}^{i+}
639: %&=&0\nn\\
640: %(-{\cal D}_{\bar{z}}{\cal D}_z+2Q_iQ^{i\dagger})\lambda_+&=&0 \ee
641: %
642: %and
643: %%
644: %\be (-{\cal D}_z{\cal D}_{\bar{z}}\delta^j_{\
645: %i}+2Q_iQ^{j\dagger})\bar{\psi}^{j-}&=&0\nn\\
646: %(-{\cal D}_{\bar{z}}{\cal D}_z\delta_j^{\ i}+2Q^{i\dagger}
647: %Q_j)\bar{\psi}^{j+} - \sqrt{2}{\cal D}_{\bar{z}}Q^{i\dagger}
648: %\lambda_- &=&0 \ee
649: %
650: The operators appearing in the equations for $\lambda_-$ and
651: $\bar{\psi}_{-i}$ are positive definite: these components can have
652: no zero modes. All zero modes live in the components $\lambda_+$
653: and $\bar{\psi}_{+i}$.
654:
655: \para
656: To see how this correlates with the chirality of the worldsheet
657: fermions, we now allow these zero modes to vary along the string
658: so that $\lambda_+=\lambda_+(x^0,x^3)$ and
659: $\bar{\psi}_{+i}=\bar{\psi}_{+i}(x^0,x^3)$. Plugging this ansatz
660: back into the Dirac equation, including now the derivatives ${\cal
661: D}_\pm$ in \eqn{ds}, we find the equations of motion
662: ${\partial}_-\lambda_+=\partial_-\bar{\psi}_{+i}=0$. We call these
663: fermions {\it right movers}.
664:
665: \para
666: Repeating this analysis for Dirac equations \eqn{etap}, we find
667: that $\eta_-$ and $\bar{\tilde{\psi}}_{-i}$ both carry zero modes
668: in the background of the vortex. They are {\it left movers} on the
669: string worldsheet\footnote{In the background of an anti-vortex,
670: with ${\cal D}_{\bar{z}}Q_i=0$, the chirality of the fermi zero
671: modes is reversed, so that $(\lambda, \bar{\psi})$ donate left
672: movers, while $(\eta,\bar{\tilde{\psi}})$ donate right movers.}.
673:
674: \subsubsection{Fermi Zero Modes}
675:
676: From the previous analysis, we learn that the right moving fermi
677: zero modes solve
678: %
679: \be \sqrt{2}{\cal D}_z\lambda_+ &=& -e^2\,Q_i\bar{\psi}_{+i} \nn\\
680: \sqrt{2}{\cal D}_{\bar{z}} \bar{\psi}_{+i}&=&-
681: Q_i^\dagger\lambda_+\ee
682: %
683: while the equations for the left moving fermi zero modes solve
684: %
685: \be \sqrt{2}i{\cal D}_{\bar{z}} \eta_- &=& -e^2
686: \bar{\tilde{\psi}}_{-i} Q_i^\dagger \nn\\
687: \sqrt{2}i{\cal D}_z \bar{\tilde{\psi}}_{-i} &=& \eta_-Q_i.\ee
688: %
689: Each of these pairs of equations is the same as the equations for
690: bosonic zero modes \eqn{bogzero} that are derived by linearizing
691: the vortex equations and imposing a gauge fixing constraint. The
692: relationship between the bosonic and fermionic zero modes is
693: given by
694: %
695: \be \lambda_+\leftrightarrow \delta A_{\bar{z}}\ \ &{\rm and}& \
696: \
697: \sqrt{2}\bar{\psi}_{+i}\leftrightarrow -i\delta Q_i^\dagger \nn\\
698: \eta_-\leftrightarrow \delta A_z \ \ &{\rm and}& \ \
699: \sqrt{2}\bar{\tilde{\psi}}_{-i}\leftrightarrow -\delta
700: Q_i.\label{relzero}\ee
701: %
702: This mapping between the zero mode profiles is a consequence of
703: the preserved supersymmetry in the background of the vortex. Using
704: this, it is trivial to derive the explicit zero modes in the case
705: of a single $k=1$ vortex.
706:
707: \subsubsection*{Goldstino Modes}
708:
709: The bosonic translational modes were given in \eqn{transzero}.
710: Their fermionic counterparts are
711: %
712: \be \lambda_+=F_{z\bar{z}}\bar{\chi}_+ \ \ &{\rm and}& \ \
713: \bar{\psi}_{+i}=-\frac{i}{\sqrt{2}} {\cal D}_{z}Q_i^\dagger
714: \bar{\chi}_+
715: \nn\\
716: \eta_-=F_{\bar{z}z}\chi_- \ \ &{\rm and}& \ \
717: \bar{\tilde{\psi}}_{-i}=-\frac{1}{\sqrt{2}} {\cal D}_{\bar{z}}Q_i
718: \chi_- .\label{goldy}\ee
719: %
720: Both of these are Goldstino modes, arising from acting on the
721: bosonic vortex profile \eqn{solutions} with the two broken
722: supersymmetries parameterized by $\chi_\pm$. The above formulae
723: hold for arbitrary $k$; if we restrict to the explicit $k=1$
724: solution, we may write these in terms of the vortex profile
725: function $q(\rho)$,
726: %
727: \be (\lambda_+)^a_{\ b}= \frac{ie^2v^2}{2r}
728: (q^2-1)\phi^a\bar{\phi}_i\bar{\chi}_+ \ \ &{\rm and}& \ \
729: \bar{\psi}^A_{+i}=-\frac{i\sqrt{2}v}{r}(\partial_zq)\,{\phi^a\bar{\phi}_b}
730: \bar{\chi}_+
731: \nn\\
732: \eta_-= -\frac{ie^2v^2}{2r} (q^2-1)\phi^a\bar{\phi}_b\chi_-
733: \ \ &{\rm and}& \ \ \bar{\tilde{\psi}}^a_{-i}=- \frac{\sqrt{2}v}{r}
734: (\partial_{\bar{z}}q)\,{\phi^a\bar{\phi}_i}
735: \chi_- .\label{goldy2}\ee
736: %
737:
738:
739: \subsubsection*{Super-Orientation Modes}
740:
741: The superpartners of the orientational modes are equally easy to
742: write down. Given the bosonic zero modes \eqn{orientzero}, we have
743: %
744: \be (\lambda_+)^a_{\ b}=2i (\partial_{\bar{z}}q)\,
745: \phi^a\bar{\xi}_{+b}
746: \ \ &{\rm and}& \ \
747: \bar{\psi}^a_{+i}=-\frac{iv}{\sqrt{2}}(q^2-1)\phi_i\bar{\xi}^a_{+}
748: \nn\\
749: (\eta_-)^a_{\ b}= -2i(\partial_z q)\, {\xi}^a_-\bar{\phi}_b \ \
750: &{\rm and}& \ \ \bar{\tilde{\psi}}^a_{-i}=
751: -\frac{v}{\sqrt{2}}(q^2-1) \xi_-^a \bar{\phi}_i .\label{somodes}\ee
752: %
753: It is clear from these expressions, that the redundancy \eqn{iden}
754: which acts among the $\phi_i$ orientational coordinates, must also
755: act on the superpartners $\xi_{\pm i}$, so that
756: %
757: \be \phi_i\rightarrow e^{i\alpha}\phi_i\ \ \ {\rm and} \ \ \
758: \xi_i\rightarrow e^{i\alpha}\xi_i .\label{iden2}\ee
759: %
760: Moreover, the fact that there do not exist orientational
761: coordinates in the $N=1$ Abelian theory means that we must impose
762: a constraint on the $\xi_{\pm i}$, namely
763: %
764: \be \sum_{i=1}^{N_c}\bar{\phi}^i\xi_{\pm i}=0 .\label{fermcons}\ee
765: %
766:
767:
768: \subsection{Supersymmetric Dynamics}
769:
770:
771: The low-energy dynamics of the vortex string arises by promoting
772: the collective coordinates $z$, $\chi_\pm$, $\phi_i$ and $\xi_{\pm
773: i}$ to dynamical fields on the string worldsheet, depending on
774: $y^0\equiv x^0$ and $y^1\equiv x^3$. The fact that the vortices
775: are BPS, preserving 1/2 of the ${\cal N}=2$ four-dimensional
776: supersymmetry, ensures that the resulting worldsheet dynamics is
777: invariant under ${\cal N}=(2,2)$ supersymmetry. Indeed, the
778: various bosonic and fermionic collective coordinates are easily
779: packaged into ${\cal N}=(2,2)$ superfields. The translational mode
780: $z$ and the two Goldstino modes $\chi_\pm$ sit in an ${\cal
781: N}=(2,2)$ chiral multiplet $Z$. Our notation is
782: standard\footnote{The one deviation from standard notation is to
783: label the complex auxiliary fields in each chiral multiplet as
784: $G$. This distinguishes them from the auxiliary $F$ fields in
785: four-dimensions.} and follows, for example, \cite{phases}
786: %
787: \be
788: Z=z+\theta^+\chi_++\theta^-\chi_-+\theta^+\theta^-G_Z+\ldots\ .\ee
789: %
790: Similarly, the orientation modes $\phi_i$ and their superpartners
791: $\xi_{\pm i}$ also sit in $(2,2)$ chiral multiplets,
792: %
793: \be
794: \Phi_i=\phi_i+\theta^+\xi_{+i}+\theta^-\xi_{-i}+\theta^+\theta^-G_i+\ldots\ .
795: \ee
796: %
797: The two constraints $\bar{\phi}^i\phi_i=r$, and
798: $\bar{\phi}^i\xi_{\pm i}=0$, together with the identification
799: \eqn{iden2}, are imposed on the worldsheet theory by introducing
800: an auxiliary ${\cal N}=(2,2)$ vector multiplet which, in
801: Wess-Zumino gauge, has components
802: %
803: \be
804: U&=&-\theta^-\bar{\theta}^-(u_0-u_1)+\theta^+\bar{\theta}^+(u_0+u_1)
805: - \theta^-\bar{\theta}^+\sigma -
806: \theta^+\bar{\theta}^-\bar{\sigma}
807: \\ && + \sqrt{2}{i}\theta^i\theta^+(\bar{}\theta}^-\bar{\zeta}_-
808: +\bar{\theta}^+\bar{\zeta}_+)+\sqrt{2}{i\bar{\theta}^+\bar{\theta}^-
809: (\theta^-\zeta_-+\theta^+\zeta_+)+2\theta^-\theta^+
810: \bar{\theta}^+\bar{\theta}^- D .\nn\ee
811: %
812: The two dimensional field strength
813: $u_{01}=\partial_0u_1-\partial_1 u_0$ is naturally housed in a
814: twisted chiral multiplet, defined by $\Sigma =
815: \bar{D}_+D_-U/\sqrt{2}$, with component expansion
816: %
817: \be \Sigma = \sigma -
818: i\sqrt{2}\theta^+\bar{\zeta}_+-i\sqrt{2}\bar{\theta}^-\zeta_-+\sqrt{2}
819: \theta^+\bar{\theta}^-(D-iu_{01})+\ldots\ .\label{twistedc}\ee
820: %
821: The fields $\sigma$, $\zeta_\pm$ and $D$ are all auxiliary. Their
822: role will become clear shortly.
823:
824: \para
825: With the exception of a single integration constant $t$, the
826: dynamics of a $k=1$ vortex string is fixed entirely by the
827: symmetries of the theory. In particular, the $SU(N_c)_{\rm diag}$
828: symmetry of \eqn{breaking} descends to an $SU(N_c)$ global
829: symmetry on the worldsheet, under which the $\Phi_i$ transform in
830: the fundamental ${\bf N_c}$ representation. The resulting
831: dynamics is given by
832: %
833: \be {\cal L}_{\rm vortex} = \int d^4\theta\ T\bar{Z}Z+
834: \sum_{i=1}^{N_c}\bar{\Phi}_ie^{2U}\Phi_i + \frac{it}{2\sqrt{2}}\int d\theta^+
835: d\bar{\theta}^-\ \Sigma\ee
836: %
837: where $T=2\pi v^2$ is the tension of the vortex, while
838: %
839: \be t = ir+\frac{\theta}{2\pi}\ee
840: %
841: is the integration constant that needs to be fixed, and plays the
842: role of a complexified worldsheet FI parameter. After integrating
843: out the auxiliary fields $G_Z$ and $G_i$, the purely bosonic part
844: of the worldsheet Lagrangian reads,
845: %
846: \be {\cal L}_{\rm bose}=T\,|\partial_m z|^2 +
847: \sum_{i=1}^{N_c}\left(|{\cal D}_m\phi^i|^2 -
848: 2|\sigma|^2|\phi_i|^2\right)
849: +D(\sum_{i=1}^{N_c}|\phi_i|^2-r)+\frac{\theta}{2\pi}u_{01}.\ \ \ \
850: \ \ee
851: %
852: Here the $\phi_i$ fields carry charge $+1$ under the gauge
853: symmetry, with ${\cal D}\phi_i=\partial\phi_i - iu\phi_i$.
854: Dividing out by this symmetry imposes the identification
855: \eqn{iden}: $\phi_i\rightarrow e^{i\alpha}\phi_i$. Meanwhile, the
856: $D$-field in this Lagrangian plays the role of a Lagrange
857: multiplier, imposing the condition \eqn{phiphi}:
858: $\sum_i|\phi_i|^2=r$. The value of $r$ is fixed by the requirement
859: that the kinetic terms for $\phi_i$ are canonical. One finds the
860: result,
861: %
862: \be r=\frac{2\pi}{e^2}.\ee
863: %
864: This result was first shown using a brane construction in
865: \cite{vib}, and later re-derived by explicitly computing the
866: overlap of zero modes\footnote{This follows by taking the time
867: dependent ansatz for the orientational modes: ${\cal
868: D}_tQ_i=\delta Q_i$ and $F_{0z}=\delta A_z$, with
869: $\delta\phi_i=\dot{\phi}_i$. Inserting this into the four
870: dimensional kinetic terms gives,
871: %
872: \be \int dx^1dx^2 \left(\frac{1}{2e^2}F_{0i}^2+|{\cal
873: D}_tQ_i|^2\right) = \int dx^1dx^2 \left(\frac{1}{e^2}|{\cal
874: D}_iq|^2 +v^2(q-1)^2(q+1)^2\right)\frac{|{\cal D}_t\phi_i|^2}{r} =
875: \frac{2\pi}{e^2} \frac{|{\cal D}_t\phi_i|^2}{r}\nn\ee where the
876: integral is recognized as the same one that appears in computing
877: the vortex tension.} in \cite{others}. Similarly, it can be shown
878: that the four-dimensional $\theta$-angle descends to a worldsheet
879: $\theta$-angle \cite{vstring,others}. The end result is that
880: worldsheet complexified FI parameter $t$ is identified with the
881: bulk complexified gauge coupling $\tau$ \eqn{tau}:
882: %
883: \be t = \tau\ .\ee
884: %
885: We now turn to the fermionic part of the worldsheet Lagrangian,
886: given by
887: %
888: \be L_{\rm fermi} &=& 2iT\left(\bar{\chi}_-\partial_+\chi_-
889: +\bar{\chi}_+\partial_-\chi_+\right) +
890: 2i\sum_{i=1}^{N_c}\left(\bar{\xi}_{-i}{\cal D}_+\xi_{-i} +
891: \bar{\xi}_{+i}{\cal D}_-\xi_{+i}\right)\\ &&
892: -\sqrt{2}\sum_{i=1}^{N_c}\left(
893: \bar{\sigma}\bar{\xi}_{+i}\xi_{-i}+\sigma \bar{\xi}_{-i}\xi_{+i}
894: +\bar{\phi}_i(\xi_{-i}\zeta_+-\xi_{+i}\zeta_-)+\phi_i(\bar{\zeta}_-\bar{\xi}_{+i}
895: -\bar{\zeta}_+\bar{\xi}_{-i})\right)\nn\ee
896: %
897: The fermions $\xi_{\pm i}$ both have charge $+1$ under the $U(1)$
898: gauge symmetry, which is now seen to implement the full
899: identification \eqn{iden2}: $\phi_i\rightarrow e^{i\alpha}\phi_i$
900: and $\xi_\pm\rightarrow e^{i\alpha}\xi_{\pm}$. The vector
901: multiplet fermions $\zeta_\pm$ have no kinetic term and act as
902: Grassmannian Lagrange multipliers, imposing the constraint
903: \eqn{fermcons}: $\sum_i\bar{\phi}_i\xi_{\pm i}=0$. Finally, the
904: role of $\sigma$ is to mediate a four-fermi interaction for the
905: super-orientation modes. Upon integrating out $\sigma$, we have
906: %
907: \be {\cal L}_{\rm 4-fermi} =
908: -2|\sigma|^2|\phi_i|^2-\sqrt{2}\bar{\sigma}\bar{\xi}_{+i}\xi_{-i}-\sqrt{2}
909: \sigma\bar{\xi}_{-i} \xi_{+i} =
910: -\frac{|\bar{\xi}_{-i}\xi_{+i}|^2}{r}.\label{224fermi}\ee
911: %
912: Four-fermi terms of this kind are typical for soliton dynamics in
913: supersymmetric theories. We pause here to review how they arise.
914: In deriving the Dirac equations \eqn{lambdap} and \eqn{etap} we
915: set $A=\tilde{Q}_i=0$. This is valid in the background of the
916: bosonic vortex. However, it is no longer true in the presence of
917: fermions since fermi bilinears act as a source for these fields.
918: For example, the Yukawa couplings involving $A^\dagger$ contribute
919: to the equation of motion,
920: %
921: \be {\cal D}^2A +
922: {i\sqrt{2}}[\lambda,\eta]-\sqrt{2}e^2\bar{\tilde{\psi}}_i\tilde{\psi}_i
923: +e^2\{Q_iQ_i^\dagger + \tilde{Q}_i^\dagger\tilde{Q}_i, A\} =0
924: \label{yukyuk}\ee
925: % \nn\\ {\cal D}^2\tilde{Q}_i &=& i\sqrt{2}\tilde{\psi}_i\lambda
926: %- \sqrt{2}\eta\psi_i\ee
927: %
928: The solution to this equation then feeds back into the Dirac
929: equations \eqn{diracs} and must be solved iteratively, order by
930: order in the number of Grassmannian collective coordinates. This
931: is a finite, but somewhat complicated procedure (see \cite{stefan}
932: for a simple quantum mechanical model where it may be carried
933: through to completion). Thankfully, the end result \eqn{224fermi}
934: is dictated by supersymmetry.
935:
936:
937: \subsubsection{Symmetries}
938: \label{symmetries}
939:
940: The four-dimensional ${\cal N}=2$ theory has two $U(1)$
941: R-symmetries\footnote{In the absence of the FI parameter $v^2$,
942: the theory has an $SU(2)_R$ symmetry, under which $(\lambda,\eta)$
943: and $(Q,\tilde{Q}^\dagger)$ both transform as doublets. The FI
944: parameter breaks $SU(2)_R\rightarrow U(1)_V$.} that we will call
945: $U(1)_R$ and $U(1)_V$. The charges of the various fields under
946: $U(1)_R\times U(1)_V$ are listed in the table.
947: %
948: \begin{center}\begin{tabular}{c|ccccccc} & $A$ & $\lambda$ & $\eta$ &
949: $Q$ & $\tilde{Q}$ & $\psi$
950: & $\tilde{\psi}$ \\ \hline $U(1)_R$ & 2 & 1 & 1 & 0 & 0 & -1 & -1 \\
951: $U(1)_V$ & 0 & 1 & -1 & 1 & 1 & 0 & 0\ .
952: \end{tabular}\end{center}
953: %
954: Both of these symmetries descend to the vortex worldsheet, where
955: they appear as the two R-symmetries of the ${\cal N}=(2,2)$
956: superalgebra. The action on the fermionic collective coordinates
957: of the vortex can can be read directly from \eqn{relzero}. We
958: have
959: %
960: \begin{center}\begin{tabular}{c|ccccccccc} & $\sigma$ & $\zeta_+$
961: & $\zeta_-$ & $\phi$ & $\xi_+$ & $\xi_-$ & $z$ & $\chi_+$ &
962: $\chi_-$ \\ \hline $U(1)_R$ & 2 & -1 & 1 & 0 & -1 & 1 & 0 & -1 & 1 \\
963: $U(1)_V$ & 0 & 1 & 1 & 0 & -1 & -1 & 0 & -1 & -1
964: \\ $U(1)_Z$ & 0 & 1 & 1 & 0 & -1 & -1 & 2 & 1 & 1\ .
965: \end{tabular}\end{center}
966: %
967: The $U(1)_R$ symmetry is axial; it suffers an anomaly in the
968: quantum theory of the vortex (as, indeed, does the $U(1)_R$ in
969: four-dimensions). In contrast $U(1)_V$ is a vector R-symmetry on
970: the worldsheet.
971:
972: \para
973: The vortex theory also includes a further global $U(1)_Z$
974: symmetry, which arises from rotating the vortex string in the
975: $z=x^1+ix^2$ plane. The charges of the worldsheet fields under
976: $U(1)_Z$ are listed in the table and follow from \eqn{goldy} and
977: \eqn{somodes}. There exists a suitable linear combination of
978: $U(1)_Z$ and $U(1)_V$ which simply rotates the phase of the chiral
979: multiplet $Z$, leaving all other fields invariant.
980:
981: \para
982: There are other, translational, symmetries of the worldsheet
983: theory that reflect the fact that $z$ and $\chi_\pm$ are all
984: Goldstone modes, arising from broken translation and supersymmetry
985: invariance respectively. In both cases, this ensures they have
986: only derivative couplings. In particular, it is the existence of
987: these symmetries that prevents the Goldstino modes $\chi_\pm$ from
988: appearing in the four-fermi term \eqn{224fermi}.
989:
990:
991:
992: \subsection{${\cal N}=2$ Preserving Deformations}
993:
994: So far we have described vortices in only the simplest ${\cal
995: N}=2$ theory with $N_f=N_c$. There are a number of ways to deform
996: and augment our theory that preserve ${\cal N}=2$ supersymmetry.
997: Here we list them and describe their effect on the worldsheet. We
998: postpone until Section 4 a discussion of deformations that break
999: the four dimensional supersymmetry to ${\cal N}=1$.
1000:
1001: \subsubsection{Adding Masses}
1002:
1003: The simplest deformation of our theory that preserves ${\cal N}=2$
1004: supersymmetry is to add a complex mass parameter $m_i$ for each
1005: hypermultiplet. The superpotential \eqn{4dn2} now becomes
1006: %
1007: \be {\cal W}_{{\cal
1008: N}=2}=\sqrt{2}\sum_{i=1}^{N_f}\tilde{Q}_i(A-m_i)Q_i .\ee
1009: %
1010: The vacuum \eqn{vac1} survives only if we turn on the adjoint
1011: scalar field $A$ to cancel the F-term contributions,
1012: %
1013: \be Q^a_{\ i}=v\delta^a_{\ i}\ \ \ ,\ \ \ \tilde{Q}_i=0\ \ \ \ , \
1014: \ \ A={\rm diag}(m_1,\ldots,m_{N_c}) .\label{vac2}\ee
1015: %
1016: The vortex moduli space does not fare well under this deformation.
1017: It can be simply shown that the masses $m_i$ lift the internal
1018: ${\bf CP}^{N_c-1}$ vortex moduli space, leaving behind $N_c$
1019: distinct, isolated vortex solutions, each of which carries
1020: magnetic flux in a different diagonal $U(1)$ subgroup of the
1021: $U(N_c)$ gauge group, supported by a different $Q_i$ winding at
1022: infinity.
1023:
1024: \para
1025: It was shown in \cite{memono,sy,vstring} that the 4d masses $m_i$
1026: induce ``twisted masses" \cite{hh} for the fields on the vortex
1027: worldsheet. In the language of ${\cal N}=(2,2)$ superfields, this
1028: deformation replaces the standard kinetic terms for $\Phi_i$
1029: by
1030: %
1031: \be \sum_{i=1}^{N_c}\bar{\Phi}_i e^{2U}\Phi_i \longrightarrow
1032: \sum_{i=1}^{N_c}\ \bar{\Phi}_i
1033: \exp\left(2U-2\theta^-\bar{\theta}^+m_i-2\theta^+\bar{\theta}^-m_i^\dagger\right)
1034: \Phi_i.\ee
1035: %
1036: In terms of components, the vortex theory (neglecting for now the
1037: $Z$ multiplet whose dynamics remains unchanged) becomes,
1038: %
1039: \be {\cal L}_{\rm vortex} &=& \sum_{i=1}^{N_c}\left(|{\cal D}_m\phi_i|^2
1040: +|F_i|^2- 2|\sigma-m_i|^2|\phi_i|^2\right)
1041: +D(\sum_{i=1}^{N_c}|\phi_i|^2-r)+\frac{\theta}{2\pi}u_{01} \nn\\
1042: && +\sum_{i=1}^{N_c} 2i\left(\bar{\xi}_{-i}{\cal D}_+\xi_{-i} +
1043: \bar{\xi}_{+i}{\cal D}_-\xi_{+i}\right) -\sqrt{2}
1044: (\bar{\sigma}-\bar{m}_i)\bar{\xi}_{+i}\xi_{-i}
1045: -\sqrt{2}(\sigma-m_i) \bar{\xi}_{-i}\xi_{+i} \nn\\ &&
1046: -i\sqrt{2}\sum_{i=1}^{N_c}\Big(
1047: \bar{\phi}_i(\xi_{-i}\zeta_+-\xi_{+i}\zeta_-)+\phi_i(\bar{\zeta}_-\bar{\xi}_{+i}
1048: -\bar{\zeta}_+\bar{\xi}_{-i})\Big).\ee
1049: %
1050: Note that the masses $m_i$ break the $U(1)_R$ symmetry, both in
1051: four-dimensions and on the vortex worldsheet. The twisted masses
1052: on the worldsheet have the desired effect of lifting the ${\bf
1053: CP}^{N_c-1}$ moduli space of the vortex theory, leaving behind
1054: $N_c$ isolated vacua given by
1055: %
1056: \be |\phi_i|^2 = r\delta_{ij}\ \ \ ,\ \ \ \sigma = m_j\ \ \ \
1057: j=1,\ldots, N_c .\ee
1058: %
1059: These different vacua of the worldsheet theory are identified with
1060: the different vortex solutions in four-dimensions. Kinks
1061: interpolating between these vacua on the worldsheet correspond to
1062: magnetic monopoles in four-dimensions, confined to lie on the
1063: vortex string by the Meissner effect \cite{memono}.
1064:
1065:
1066: \subsubsection{Adding Flavors}
1067: \label{add}
1068:
1069: We now consider the theory with $N_f>N_c$ fundamental
1070: hypermultiplets. The D-term and F-term vacuum conditions in
1071: four-dimensions read
1072: %
1073: \be \sum_{i=1}^{N_f}Q_iQ_i^\dagger - \tilde{Q}_i^\dagger
1074: \tilde{Q}_i = v^2\ \ \ {\rm and}\ \ \ \sum_{i=1}^{N_f}
1075: Q_i\tilde{Q}_i= 0 .\label{higgsbranch}\ee
1076: %
1077: When $m_i=0$, there are no further conditions, and there is a
1078: $2N_c(N_f-N_c)$ dimensional Higgs branch of the theory. For the
1079: purposes of this section, we place ourselves in the particular
1080: vacuum $\tilde{Q}_i=0$ and
1081: %
1082: \be Q^a_{\ i}=v\delta^a_{\ i} \ \ \ {a,i=1,\ldots N_c}
1083: \label{0vac}\ee
1084: %
1085: with $Q_i=0$ for $i=N_c+1,\ldots, N_f$. This is to be supplemented
1086: by $A=0$. If we now turn on masses for the hypermultiplets, the
1087: vacuum \eqn{0vac} survives, with $A={\rm
1088: diag}(m_1,\ldots,m_{N_c})$.
1089:
1090:
1091: \para
1092: Vortices in theories with $N_f>N_c$ have a rather different
1093: character than those in the $N_f=N_c$ theory. The most noticeable
1094: difference is that they gain extra bosonic collective coordinates,
1095: among them a scale size. These additional collective coordinates
1096: are non-normalizable when $m_i=0$ \cite{ward,ls} but become
1097: normalizable when finite masses $m_i$ are turned on for
1098: $i=N_c+1,\ldots, N_f$ \cite{semi}. Vortices of this kind are
1099: sometimes referred to as semi-local vortices: a review of these
1100: objects in Abelian theories can be found in \cite{semi2}, while a
1101: detailed discussion in non-Abelian theories was given in
1102: \cite{semi}.
1103:
1104:
1105: \para
1106: An effective dynamics for the vortex worldsheet in theories with
1107: $N_f>N_c$ was proposed in \cite{vib}, based on a D-brane
1108: construction. It is once again an ${\cal N}=(2,2)$ supersymmetric
1109: $U(1)$ gauge theory, now with $N_c$ chiral multiplets $\Psi_i$ of
1110: charge $+1$ and a further $(N_f-N_c)$ chiral multiplets
1111: $\tilde{\Psi}_j$ of charge $-1$. The D-term for this theory reads
1112: %
1113: \be D= \sum_{i=1}^{N_c} |\phi_i|^2 -
1114: \sum_{j=1}^{N_f-N_f}|\tilde{\phi}_j|^2-r = 0 \label{tilphi}\ee
1115: %
1116: which, together with the gauge action $\phi_i\rightarrow
1117: e^{i\alpha}\phi_i$ and $\tilde{\phi}_i\rightarrow
1118: e^{-i\alpha}\tilde{\phi}_i$, defines the Higgs branch of the
1119: vortex theory. This Higgs branch is conjectured to coincide the
1120: vortex moduli space. As in the previous section, assigning complex
1121: masses $m_i$ to the four dimensional hypermultiplets induces
1122: twisted masses $m_i$, $i=1,\ldots,N_c$ for the $\Phi_i$ fields,
1123: and twisted masses $\tilde{m}_j=m_{j+N_c}$, $j=1,\ldots, N_f-N_c$
1124: for $\tilde{\Phi}_j$.
1125:
1126: \para
1127: The presence of the negatively charged fields $\tilde{\phi}_j$
1128: means that the moduli space \eqn{tilphi} is now non-compact,
1129: corresponding to the scaling mode of the vortex. Note however that
1130: the natural metric on the Higgs branch does not coincide with the
1131: natural metric on the vortex moduli space. In particular, the
1132: non-normalizability of the scaling modes as $m_i\rightarrow 0$ is
1133: not reproduced in this model. Nonetheless, it has been shown that
1134: the vortex theory \eqn{tilphi} does indeed correctly capture the
1135: quantum dynamics of the vortex string \cite{vstring,scvs}.
1136:
1137:
1138:
1139:
1140: \subsubsection*{Higgs Expectation Values}
1141:
1142: When $N_f>N_c$ and $m_i=0$, the vacuum conditions
1143: \eqn{higgsbranch} in the four-dimensional theory have a moduli
1144: space of solutions. We may ask what happens to the vortex string
1145: as we change the expectation values of $Q_i$ and $\tilde{Q}_i$
1146: such that \eqn{higgsbranch} remains satisfied. The answer to this
1147: question was given in \cite{scvs}: turning on expectation values
1148: for $\tilde{Q}_i$ induces a superpotential on the vortex string
1149: worldsheet. For completeness, we briefly describe this deformation
1150: here.
1151:
1152: \para
1153: First some notation: define the gauge invariant meson operator
1154: %
1155: \be M_i^j\equiv \tilde{Q}^j Q_i\label{mij} .\ee
1156: %
1157: It is not hard to show that in four-dimensional vacua for which
1158: $M\neq 0$, the space of BPS vortex solutions is greatly reduced.
1159: The key point is that a vacuum expectation value for $\tilde{Q}$
1160: does not allow a BPS vortex to live in the associated part of the
1161: gauge group. This follows from the mathematical fact that there is
1162: no holomorphic line bundle of negative degree. In a more physical
1163: language, a direct analysis of the vortex equations reveals that
1164: BPS vortices do not exist in $U(1)$ theories if both negatively
1165: and positively charged fields gain an expectation value
1166: \cite{penin,davis}. The upshot of this is that the vortex moduli
1167: space is partly lifted in four dimensional vacua for which $M\neq
1168: 0$. It was shown in \cite{scvs} that this effect is captured on
1169: the vortex worldsheet by the introduction of a superpotential of
1170: the form,
1171: %
1172: \be {\cal W}_{(2,2)}\sim \sum_{i=1}^{N_f} \sum_{j=1}^{N_f-N_c}
1173: M^{j+N_c}_i \,\tilde{\Phi}_j\Phi^i .\ee
1174: %
1175:
1176:
1177:
1178:
1179: \subsection{Multiple Vortices}
1180:
1181: %
1182: \EPSFIGURE{vib.eps,height=160pt}{}
1183: %
1184: \noindent So far we have focussed on the theory for a single
1185: vortex string for which, at least in the case $N_f=N_c$,
1186: symmetries are sufficient to dictate the dynamics. In \cite{vib},
1187: a D-brane construction was used to derive a worldsheet theory
1188: which describes the interactions of $k>1$ parallel vortex strings.
1189: The D-brane construction starts with the usual Hanany-Witten
1190: set-up for ${\cal N}=2$ four-dimensional gauge theories
1191: \cite{hanwit,w}, consisting of D4-branes attached to parallel
1192: NS5-branes. Separating the NS5-branes in the direction out of the
1193: page induces the FI parameter $v^2$. The vortex strings arise as
1194: stretched D2-branes as shown in figure 2. The worldvolume theory
1195: of $k$ vortex strings is given by an ${\cal N}=(2,2)$ $U(k)$
1196: non-Abelian gauge theory with matter content,
1197: %
1198: \para
1199:
1200: \be \mbox{$U(k)$ Vector Multiplet $U$} &+& \mbox{Adjoint Chiral
1201: Multiplet $Z$}\nn\\ &+& \mbox{$N_c$ Fundamental Chiral Multiplets
1202: $\Phi_i$} \nn\\ &+& \mbox{$N_f-N_c$ Anti-Fundamental Chiral
1203: Multiplets $\tilde{\Phi}_j$} . \nn\ee
1204: %
1205: The complexified worldsheet FI parameter is again equated to the
1206: 4d complexified gauge coupling, $t=\tau$, or
1207: %
1208: \be ir+\frac{\theta_{2d}}{2\pi}=\frac{2\pi
1209: i}{e^2}+\frac{\theta_{4d}}{2\pi}\label{fi} . \ee
1210: %
1211: The D-term condition for the worldsheet theory is now $u(k)$
1212: valued and is given by,
1213: %
1214: \be \sum_{i=1}^{N_c}\phi_i\phi_i^\dagger
1215: -\sum_{j=1}^{N_f-N_c}\tilde{\phi}_j^\dagger\tilde{\phi}_j+T[z,z^\dagger]=r\,1_k .
1216: \label{vortexd}\ee
1217: %
1218: This provides $k^2$ constraints on the $2k(N_f+k)$ degrees of
1219: freedom in $\phi_i$, $\tilde{\phi}_j$ and $z$. After dividing by
1220: $U(k)$ gauge transformations, we are left with a $2kN_f$
1221: dimensional manifold which defines the target space for the vortex
1222: string sigma-model. This was conjectured in \cite{vib} to coincide
1223: with $2kN_f$ dimensional vortex moduli space. This quotient
1224: construction has subsequently been derived from a direct analysis
1225: of the non-Abelian vortex equations \cite{mmatrix,mmreview}.
1226:
1227: \para
1228: %The fermionic terms in the $U(k)$ worldsheet theory are given by,
1229: %
1230: %\be L_{\rm fermi}&=&2iT\ \Tr_k\left(\bar{\chi}_-{\cal
1231: %D}_+\chi_-+\bar{\chi}_+{\cal D}_-\chi_+\right) +
1232: %2i\sum_{i=1}^{N_c}\left(\bar{\xi}_{-i}{\cal
1233: %D}_+\xi_{-i}+\bar{\xi}_{+i}{\cal D}_-\xi_{+i}\right) \nn\\
1234: %&&
1235: %-\sqrt{2}\,\Tr_k\left([\bar{\chi}_-,[\sigma,\chi_+]]-[\bar{\chi}_+,[\bar{\zeta}_-,
1236: %z]]-[\bar{\xi}_-,[\bar{\zeta}_+,z]]\right) + {\rm h.c.} \label{22phew}\\
1237: %&& -\sqrt{2}\sum_{i=1}^{N_c}\left( \bar{\xi}_{-i}\sigma\xi_{+i}
1238: %-\bar{\xi}_{+i}\bar{\zeta}_-\phi_i +
1239: %\bar{\xi}_{-i}\bar{\zeta}_+\phi_i\right) + {\rm h.c.} \nn\ee
1240: %%
1241: The $4kN_c$ fermionic zero modes of $k$ parallel vortex strings
1242: live in the $U(k)$ adjoint valued $\chi_\pm$ and the fundamental
1243: $\xi_{\pm i}$, subject to the $2k^2$ complex constraints arising
1244: from the auxiliary fermions $\zeta_\pm$,
1245: %
1246: \be \sum_i\phi_i\bar{\xi}_{\pm i}+[z,\bar{\chi}_\pm]=0 .\ee
1247: %
1248: In the case of a single $k=1$ vortex, these reduce to the
1249: constraints \eqn{fermcons}.
1250:
1251: \para
1252: Note that the vacuum moduli space \eqn{vortexd} inherits a metric
1253: from the canonical kinetic terms for $\phi_i$ and $z$. This metric
1254: is known not to agree with the standard Manton metric
1255: \cite{manton,samols} on the vortex moduli space (except in the
1256: special case $k=1$ and $N_f=N_c$ that we described in detail
1257: earlier). This is because the limit in which the $d=1+1$ gauge
1258: theory on the D2-branes decouples from other stringy modes is
1259: different from the limit in which the D2-branes are described as
1260: vortices in the $d=3+1$ dimensional theory on the D4-branes; the
1261: two descriptions hold in different regimes of validity as we vary
1262: the parameters of the brane set-up. Nonetheless, if one is
1263: interested in computing objects protected by supersymmetry
1264: --- such as the classical, or quantum, masses of BPS states in the
1265: vortex theory ---- it should be valid to work with the gauge
1266: linear sigma model. In practice, this claim has been confirmed
1267: only for the $k=1$ theory with $N_f>N_c$. It has also been
1268: confirmed that the intricate topology of the $k=2$ vortex string
1269: moduli space in the $N_f=N_c=2$ is correctly captured by the
1270: gauged linear sigma model \cite{asy,kt,lots}.
1271:
1272:
1273: \section{${\cal N}=(0,2)$ Supersymmetry}
1274:
1275: In the previous section we have worked with both ${\cal N}=1$
1276: superfields in four dimensions, $A$, $Q_i$ and $\tilde{Q}_j$, as
1277: well as ${\cal N}=(2,2)$ superfields in two dimensions, $\Sigma$,
1278: $\Phi_i$ and $\tilde{\Phi}_j$. From now on we will deal with
1279: ${\cal N}=(0,2)$ supersymmetry in two dimensions. Since this may
1280: be less familiar to some readers we devote this section to a
1281: review of the structure of ${\cal N}=(0,2)$ superfields \cite{hw,ds} and
1282: their relationship to ${\cal N}=(2,2)$ theories. The presentation
1283: follows \cite{phases} and \cite{abs}.
1284:
1285: \subsection{Superfields}
1286:
1287: ${\cal N}=(0,2)$ supersymmetry is generated by two right-moving,
1288: and no left-moving, supersymmetries. The two chiral supercharges
1289: are $Q_+$ and $\bar{Q}_+$. The $(0,2)$ superspace is parameterized
1290: by the bosonic coordinates $y^\pm=(y^0\pm y^1)$ and their
1291: fermionic partners $\theta^+$ and $\bar{\theta}^+$. The action of
1292: the supersymmetry generators in superspace is given as
1293: %
1294: \be
1295: Q_+&=&\frac{\partial}{\partial\theta^+}+i\bar{\theta}^+(\partial_0+\partial_1)
1296: %\ \ \ \ ,\ \ \ \
1297: \nn\\
1298: \bar{Q}_+&=&-\frac{\partial}{\partial\bar{\theta}^+}-i\theta^+(\partial_0+\partial_1) .
1299: \ee
1300: %
1301: These commute with the superderivatives,
1302: %
1303: \be D_+&=&\frac{\partial}{\partial\theta^+}-i\bar{\theta}^+
1304: (\partial_0+\partial_1)
1305: %\ \ \ \ ,\ \ \ \
1306: \nn\\
1307: \bar{D}_+&=&-\frac{\partial}{\partial\bar{\theta}^+}+i\theta^+
1308: (\partial_0+\partial_1) \ee
1309: %
1310: which satisfy $\{D_+,D_+\}=\{\bar{D}_+\bar{D}_+\}=0$ and
1311: $\{D_+,\bar{D}_+\}=2i\partial_+$. We now describe the different
1312: superfields of interest.
1313:
1314: \subsubsection*{Gauge Multiplets}
1315:
1316: We start with the real, adjoint valued, gauge multiplet $U$,
1317: which has the component expansion
1318: %
1319: \be
1320: U=(u_0-u_1)-2i\theta^+\bar{\zeta}_--2i\bar{\theta}^+\zeta_-+2\theta^+\bar{\theta}^+D .
1321: \label{02vec}\ee %%
1322: Already we see the chiral nature of the supersymmetry, since only
1323: the combination $u_-=u_0-u_1$ of the two-dimensional gauge field
1324: appears in the superfield, together with a left moving fermion
1325: $\zeta_-$. The scalar field $D$ will be seen to be auxiliary. The
1326: covariant superderivatives are given by
1327: %
1328: \be {\cal D}_+&=&\frac{\partial}{\partial
1329: \theta^+}-i\bar{\theta}^+({\cal D}_0+{\cal D}_1)
1330: %\ \ \ \ {\rm and}\
1331: \ \ \ \nn\\ \bar{{\cal D}}_+&=&-\frac{\partial}{\partial
1332: \bar{\theta}^+}+i{\theta}^+({\cal D}_0+{\cal D}_1)\nn\ee
1333: %
1334: where
1335: %
1336: \be {\cal D}_0+{\cal D}_1=\partial_0+\partial_1-i(u_0+u_1) \nn\ee
1337: %
1338: includes the $u_+$ component of the gauge field, but no fermions.
1339: Meanwhile, the gauginos are included in the remaining covariant
1340: superderivative,
1341: %
1342: \be {\cal D}_0-{\cal D}_1=\partial_0-\partial_1-iU .\ee
1343: %
1344: The field strength lives naturally in a fermi multiplet (which we
1345: shall define shortly) given by the usual commutator of
1346: derivatives:
1347: %
1348: \be \Upsilon=[\bar{\cal D}_+,{\cal D}_0-{\cal
1349: D}_1]=-2\Big(\zeta_--i\theta^+(D-iv_{01})-i\theta^+\bar{\theta}^+({\cal
1350: D}_0+{\cal D}_1)\zeta_-\Big) .\ee
1351: %
1352: Here the field strength is
1353: $u_{01}=\partial_0u_1-\partial_1u_0-i[u_0,u_1]$. The kinetic terms
1354: for the gauge multiplet are then given by integration over all of
1355: superspace $d^2\theta =d\theta^+\,d\bar{\theta}^+$,
1356: %
1357: \be S_{\rm gauge}&=&\frac{1}{8g^2}\,\Tr\,\int d^2y \,d^2\theta\ \Upsilon^\dagger \Upsilon \nn\\
1358: &=&\frac{1}{g^2}\,\Tr\,\int d^2y\ \Big(\ft12\
1359: u_{01}^2+i\bar{\zeta}_-({\cal D}_0+{\cal D}_1)\zeta_-+D^2\Big)
1360: \ee
1361: %
1362: but, in fact, will not be required in the following.
1363:
1364: \subsubsection*{Chiral Multiplets}
1365:
1366: The chiral multiplets of $(0,2)$ theories are bosonic superfields
1367: $\Phi$, living in any representation $R$ of the gauge group. They
1368: satisfy
1369: %
1370: \be \bar{\cal D}_+\Phi=0 .\ee
1371: %
1372: Chiral multiplets contain right-moving fermions $\xi_+$, paired
1373: with a complex boson $\phi$. Their component expansion gives
1374: %
1375: \be
1376: \Phi=\phi+\sqrt{2}\theta^+\xi_+-i\theta^+\bar{\theta}^+(D_0+D_1)\phi
1377: \label{02chi}\ee
1378: %
1379: where $(D_0+D_1)$ is now the usual bosonic covariant derivative.
1380: The kinetic terms for the chiral multiplet are given by the
1381: action,
1382: %
1383: \be S_{\rm chiral}&=&-\frac{i}{2}\int d^2y\,d^2\theta\
1384: \bar{\Phi}({\cal D}_0-{\cal D}_1)\Phi \label{chirallag}\\ &=& \int
1385: d^2y\ \left(-|D_\alpha\phi|^2+i\bar{\xi}_+(D_0-D_1)\xi_+ -
1386: i\sqrt{2}\bar{\phi}\zeta_-\xi_+ +
1387: i\sqrt{2}\bar{\xi}_+\bar{\zeta}_-\phi + \bar{\phi}D\phi\right) .
1388: \nn\ee
1389: %
1390: The scalar field $\phi$ couples to the auxiliary field $D$, to
1391: give rise to the usual D-term (Note that for Abelian theories, if
1392: $\Phi$ has charge $p$ then one should replace $\zeta_-\rightarrow
1393: p\zeta_-$ and $D\rightarrow pD$ in the above action.).
1394:
1395: \subsubsection*{Fermi Multiplets}
1396:
1397: One novel feature of $(0,2)$ theories that is not shared by the
1398: non-chiral $(2,2)$ theories is the existence of a fermionic
1399: multiplet $\Gamma$, containing only left moving fermions $\chi_-$
1400: and no propagating bosons. Like the chiral multiplets, they can
1401: live in any representation $R$ of the gauge group. The fermi
1402: multiplet satisfies
1403: %
1404: \be \bar{\cal D}_+\Gamma=\sqrt{2}E\label{e}\ee
1405: %
1406: where $\bar{\cal D}_+E=0$, which can be solved by taking $E$ to be
1407: a holomorphic function of chiral superfields $E=E(\Phi_i)$. The
1408: fermi multiplet has component expansion
1409: %
1410: \be \Gamma = \chi_--\sqrt{2}\theta^+ G
1411: -i\theta^+\bar{\theta}^+(D_0+D_1)\chi_--\sqrt{2}\bar{\theta}^+E .
1412: \label{02fer}\ee
1413: %
1414: Note that the superfield $\Upsilon$ containing the field strength
1415: is of this type, with $\bar{\cal D}_+\Upsilon = 0$. In general,
1416: $E$ itself will also have a $\theta$ expansion,
1417: %
1418: \be E(\Phi_i)=E(\phi_i)+\sqrt{2}\theta^+\frac{\partial
1419: E}{\partial\phi_i}\xi_{+i}
1420: -i\theta^+\bar{\theta}^+(D_0+D_1)E({\phi_i}) \ee
1421: %
1422: The kinetic terms for the fermi multiplet are
1423: %
1424: \be S_{\rm fermi}&=&-\frac{1}{2}\int d^2y\,d^2\theta\
1425: \bar{\Gamma}\Gamma \label{fermilag}\\ &=&
1426: \left(i\bar{\chi}_-(D_0+D_1)\chi_-+|G|^2-|E(\phi_i)|^2
1427: -\bar{\chi}_-\frac{\partial E}{\partial\phi^i}\xi_{+i}
1428: +\bar{\xi}_{+i}\frac{\partial\bar{E}}{\partial\bar{\phi}_i}\chi_-\right)
1429: \nn\ee
1430: %
1431: We see that the complex scalar $G$ is an auxiliary field, lacking
1432: a kinetic term. Also note that the function $E(\phi)$ appears as a
1433: potential term in the Lagrangian.
1434:
1435: \subsection{Superpotentials}
1436:
1437: In ${\cal N}=(0,2)$ theories the auxiliary field $G$ lives in a fermi
1438: multiplet $\Gamma$, rather than a chiral multiplet. A
1439: superpotential $J(\Phi_i)$ is a holomorphic function of chiral
1440: superfields and a suitable action may be constructed by
1441: integrating terms of the form $\Gamma J$ over half of superspace.
1442: Most generally we can introduce a superpotential $J^a$ for each
1443: fermi multiplet $\Gamma^a$,
1444: %
1445: \be S_{J} &=& -\frac{1}{\sqrt{2}}\sum_a\int d^2y\,d\theta^+ \ \
1446: \Gamma_a\left.
1447: J^a(\Phi_i)\right|_{\bar{\theta}^+=0}+\ {\rm h.c.} \nn\\
1448: &=& \sum_{a}\int d^2y\ \
1449: G_aJ^a(\phi_i)+\sum_i\chi_{-a}\frac{\partial
1450: J^a}{\partial\phi_i}\xi_{+i}+\ {\rm h.c.}\ .\label{sj}\ee
1451: %
1452: This integration over half of superspace yields an ${\cal
1453: N}=(0,2)$ supersymmetric invariant action if and only if
1454: $\bar{D}_+(\Gamma_aJ^a)=0$, which requires
1455: %
1456: \be \sum_a E_aJ^a=0 \label{ej} .\ee
1457: %
1458: Of course, the combination $\Gamma_aJ^a$ is also required to be
1459: gauge invariant. An important example of the superpotential is the
1460: Fayet-Iliopoulos and theta term which are packaged in the complex
1461: combination $t=ir+\theta/2\pi$. The interaction can be written as
1462: %
1463: \be S_{D\theta}&=&\frac{t}{4}\,\Tr\int d^2y d\theta^+\
1464: \left.\Upsilon\right|_{\bar{\theta}^+=0} +{\rm \ h.c.}\nn\\ &=&
1465: \Tr\int d^2y\ (-rD+\frac{\theta}{2\pi}u_{01}) .\label{fipot}\ee
1466: %
1467:
1468: \subsection{${\cal N}=(2,2)$ Decomposition}
1469:
1470: It will prove useful for orientation to recall how the more
1471: familiar ${\cal N}=(2,2)$ superfields decompose into their ${\cal
1472: N}=(0,2)$ counterparts. The conventions below are taken from
1473: \cite{phases}.
1474:
1475: \para
1476: One can enlarge ${\cal N} =(0,2)$ superspace to ${\cal N}=(2,2)$ superspace through
1477: the addition of two further fermionic components $\theta^-$ and
1478: $\bar{\theta}^-$. The corresponding superderivatives are
1479: %
1480: \be
1481: D_-=\frac{\partial}{\partial\theta^-}-i\bar{\theta}^-(\partial_0-\partial_1)\
1482: \ \ \ ,\ \ \ \
1483: \bar{D}_-=-\frac{\partial}{\partial\bar{\theta}^-}+i\theta^-(\partial_0-\partial_1).\
1484: \ee
1485: %
1486: The ${\cal N}=(2,2)$ vector multiplet $V_{(2,2)}$ decomposes into
1487: an ${\cal N}=(0,2)$ vector multiplet $V$ described in \eqn{02vec},
1488: together with an ${\cal N}=(0,2)$ chiral multiplet $\Sigma$. This
1489: chiral multiplet inherits the right moving fermion $\zeta_+$ and
1490: the complex scalar field $\sigma$ contained in $V_{(2,2)}$. It is
1491: most simply described by reduction from the ${\cal N}=(2,2)$
1492: twisted chiral multiplet containing the field strength
1493: $\Sigma_{(2,2)}=(1/\sqrt{2})\{\bar{\cal D}_+,{\cal D}_-\}$, in
1494: terms of which the ${\cal N}=(0,2)$ chiral multiplet is given by
1495: %
1496: \be \Sigma=\left.\Sigma_{(2,2)}\right|_{\theta^-=\bar{\theta}_-=0}.
1497: \ee
1498: %
1499: An ${\cal N}=(2,2)$ chiral multiplet $\Phi_{(2,2)}$ satisfies
1500: $\bar{D}_+\Phi_{(2,2)}=\bar{D}_-\Phi_{(2,2)}=0$. This chiral
1501: multiplet decomposes into an ${\cal N}=(0,2)$ chiral multiplet
1502: $\Phi$ and a fermi multiplet $\Gamma$, defined by
1503: %
1504: \be \Phi &=& \left.\Phi_{(2,2)}\right|_{\theta^-=\bar{\theta}^-=0}
1505: \nn\\
1506: \Gamma&=&\frac{1}{\sqrt{2}}\,{\cal
1507: D}_-\left.\Phi_{(2,2)}\right|_{\theta^-=\bar{\theta}^-=0}.\ee
1508: %
1509: If $\Phi_{(2,2)}$ transforms under a representation $R$ of the
1510: gauge group, then both $\Phi$ and $\Gamma$ also transform under
1511: $R$. A quick computation yields $\bar{\cal D}_+\Gamma=2i \Sigma
1512: \Phi$, meaning that, in the notation of \eqn{e}, ${\cal N}=(2,2)$
1513: supersymmetry imposes,
1514: %
1515: \be E=i\sqrt{2}\Sigma \Phi .\ee
1516: %
1517: The final ${\cal N}=(2,2)$ multiplet of interest is a twisted
1518: chiral multiplet $\Sigma_{(2,2)}$, satisfying
1519: $\bar{D}_+\Sigma_{(2,2)}=D_-\Sigma_{(2,2)}=0$. Like the ${\cal
1520: N}=(2,2)$ chiral multiplet, this too decomposes into an ${\cal
1521: N}=(0,2)$ chiral multiplet $\Sigma$ and a fermi multiplet $F$.
1522: They are given by,
1523: %
1524: \be \Sigma&=&\left.\Sigma_{(2,2)}\right|_{\theta^-=\bar{\theta}^-=0}\nn\\
1525: F&=&
1526: -\frac{1}{\sqrt{2}}\bar{D}_-\left.\Sigma_{(2,2)}\right|_{\theta^-=\bar{\theta}^-=0} .
1527: \ee
1528: %
1529: Note, however, that from the expansion \eqn{twistedc}, the
1530: $\theta^+$ component of the ${\cal N}=(0,2)$ chiral multiplet
1531: $\Sigma$ contains the barred fermion, rather than the unbarred
1532: fermion,
1533: %
1534: \be \Sigma = \sigma - i\sqrt{2}\theta^+\bar{\zeta}_+
1535: -i2\theta^+\bar{\theta}^+\partial_+\sigma .\label{02tc}\ee
1536: %
1537: This subtlety will prove important in what follows. Since twisted
1538: chiral multiplets $\Sigma_{(2,2)}$ are always uncharged under the
1539: gauge group, the corresponding fermi multiplet satisfies
1540: $\bar{\cal D}_+F=0$.
1541:
1542: \subsubsection{The Vortex Theory in ${\cal N}=(0,2)$ Language}
1543:
1544: Let us finish this section by describing the ${\cal N}=(2,2)$
1545: vortex theory of Section 2 in the language of ${\cal N}=(0,2)$
1546: superfields. This will serve to fix notation for what is to come.
1547: We decompose the fields as
1548: %
1549: \be \mbox{${\cal N}=(2,2)$ $U(k)$ Vector Multiplet}
1550: &\longrightarrow& \mbox{$U(k)$ Vector Multiplet, $U$} \nn\\
1551: &&+\ \mbox{Adjoint Chiral Multiplet
1552: $\Sigma$}\nn\\
1553: \mbox{${\cal N}=(2,2)$ Adjoint Chiral Multiplet} &\longrightarrow&
1554: \mbox{Adjoint Chiral Multiplet, $Z$} \nn\\ &&+\
1555: \mbox{Adjoint Fermi Multiplet $\Xi$} \nn\\
1556: \mbox{${\cal N}=(2,2)$ Fund. Chiral Multiplets} &\longrightarrow&
1557: \mbox{Fund. Chiral Multiplets, $\Phi_i$} \nn\\
1558: &&+\ \mbox{Fund. Fermi Multiplets $\Gamma_i$}\nn\\
1559: \mbox{${\cal N}=(2,2)$ Anti-Fund. Chiral Multiplet}
1560: &\longrightarrow& \mbox{Anti-Fund. Chiral Multiplets, $\tilde{\Phi}_j$} \nn\\
1561: &&+\ \mbox{Anti-Fund. Fermi Multiplet $\tilde{\Gamma}_j$}\nn\ee
1562: %
1563: where all the objects on the right are ${\cal N}=(0,2)$
1564: superfields. As before, $i=1,\ldots, N_c$ for $\Phi_i$ and
1565: $j=1,\ldots N_f-N_c$ for $\tilde{\Phi}_j$. Appendix A contains a
1566: list of the different component fields which appear in each of
1567: these multiplets.
1568:
1569: \para
1570: The ${\cal N}=(2,2)$ supersymmetry imposes the relations,
1571: %
1572: \be \bar{\cal D}_+\Xi=2i[\Sigma,Z] \ \ \ ,\ \ \ \bar{\cal
1573: D}_+\Gamma_i=2i(\Sigma-m_i)\Phi_i\ \ \ ,\ \ \ \bar{\cal
1574: D}_+\tilde{\Gamma}^j=-2i(\Sigma-\tilde{m}_j)\tilde{\Phi}_j\ \ \ \
1575: \ \label{22rels}\ee
1576: %
1577: (There is no sum over $i$ and $j$ on the right-hand side of these
1578: equations). As we have seen, the right-hand side of each of these
1579: equations appears as a potential ``$|E|^2$" arising in equation
1580: \eqn{fermilag}. A further contribution to the worldsheet scalar
1581: potential arises from the D-term, which provides the constraint
1582: \eqn{vortexd}.
1583:
1584:
1585:
1586: \section{The ${\cal N}=(0,2)$ Dynamics of Vortex Strings}
1587:
1588: It is now time to present new results for the dynamics of vortex
1589: strings in theories with ${\cal N}=1$ supersymmetry. Most of this
1590: section is devoted to the discussion of a simple deformation of
1591: the ${\cal N}=2$ theory by the addition of a superpotential. In
1592: Section \ref{more} we discuss a second class of deformations.
1593:
1594:
1595:
1596: \subsection{Adding a Superpotential}
1597:
1598: We start by considering a ``Dijkgraaf-Vafa"-like deformation
1599: \cite{dv}, breaking ${\cal N}=2$ to ${\cal N}=1$ through the
1600: addition of a superpotential for the adjoint superfield $A$. The
1601: superpotential now reads
1602: %
1603: \be {\cal W}=\sqrt{2}
1604: \sum_{i=1}^{N_f}\tilde{Q}_i(A-m_i)Q_i+\hat{\cal
1605: W}(A)\label{def}\ee
1606: %
1607: which gives rise to the scalar potential
1608: %
1609: %
1610: \be V_{4d}&=& \frac{e^2}{2}\Tr(\,\sum_{i=1}^{N_f}\,Q_iQ_i^\dagger
1611: - \tilde{Q}_i\tilde{Q}_i^\dagger - v^2{\bf 1}_{N_c})^2 +e^2\Tr|\,
1612: \sum_{i=1}^{N_f}\tilde{Q}_iQ_i-\partial\hat{W}/\partial A|^2\label{v4dv}\\
1613: &&+\sum_{i=1}^{N_f}\left(Q_i^\dagger
1614: \{A-m_i,\bar{A}-\bar{m}_i\}Q_i +
1615: \tilde{Q}_i\{A-m_i,\bar{A}-m_i\}\tilde{Q}_i^\dagger\right) +
1616: \frac{1}{2e^2}\Tr|[A,A^\dagger]|^2 .\nn\ee
1617: %
1618: Let's look at how this superpotential affects the vacuum
1619: structure. If ${\hat{\cal W}}$ is linear in $A$ then there is
1620: merely a constant piece in the F-term above and the Lagrangian
1621: still preserves ${\cal N}=2$ supersymmetry. We can perform an
1622: $SU(2)_R$ rotation of the scalar fields
1623: $(Q_i,\tilde{Q}_i^\dagger)$ to bring the Lagrangian back to the
1624: form \eqn{v4dn2}. We will assume that $\hW$ does not contain a
1625: linear piece. In this case, for a generic superpotential $\hW(A)$,
1626: $\tilde{Q}_i$ must turn on in the vacuum. Without loss of
1627: generality, we choose the vacuum to be of the form,
1628: %
1629: \be Q^a_{\ i} = p_i\,\delta^a_{\ i} \ \ \ , \ \ \ \tilde{Q}^a_{\
1630: i}=\tilde{p}_i\,\delta^a_{\ i}\ \ \ ,\ \ \ A={\rm
1631: diag}(m_1,\ldots,m_{N_c})\label{wbac}\ee
1632: %
1633: with
1634: %
1635: \be |p_i|^2-|\tilde{p}_i|^2=v^2 \ \ \ {\rm and} \ \ \
1636: \tilde{p}_ip_i = \left.\frac{\partial\hW}{\partial a}\right|_{m_i}
1637: \ \ \ \ \ \mbox{for each $i=1,\ldots,N_c$} \label{whatsleft}\ee
1638: %
1639:
1640: \subsection{What Becomes of the Vortex?}
1641:
1642: Our goal is to understand how this deformation affects the
1643: dynamics of the vortex string\footnote{Vortices in a similar
1644: system were studied in \cite{bolog}, but in the limit with
1645: $v^2=0$, so that the vortex is built around a linear piece of
1646: $\hW$. This gives rise to somewhat different physics from that
1647: considered here.}. Let us firstly consider the case with distinct
1648: masses $m_i$. Before adding the superpotential $\hW$ there were
1649: $N_c$ different BPS vortices, each living in a different
1650: $U(1)\subset U(N_c)$ and each with a different $Q_i$,
1651: $i=1,\ldots,N_c$ carrying the asymptotic winding. What changes in
1652: the presence of $\hW$?
1653:
1654: \para
1655: The crucial point to note is that something rather special happens
1656: when the superpotential is tuned so that a critical point
1657: coincides with one of the masses, say $m_k$ for some $k=1,\ldots,
1658: N_c$
1659: %
1660: \be \left.\frac{\partial\hW(a)}{\partial
1661: a}\right|_{a=m_k}=0 .\label{survive}\ee
1662: %
1663: If this is case, the vacuum equation \eqn{wbac} sets
1664: $\tilde{Q}_k=0$. There is then no obstacle in constructing the
1665: $k^{\rm th}$ vortex in which $Q_k$ winds; indeed the ${\cal N}=2$
1666: vortex solution remains a solution in the deformed theory.
1667:
1668: \para
1669: Vortices of this type in ${\cal N}=1$ theories are often called
1670: D-term vortices (the name arises because the symmetry breaking is
1671: induced by a FI parameter, or D-term). It was shown in \cite{ddt}
1672: that such objects are 1/2 BPS, preserving two of the four
1673: supercharges of the four-dimensional ${\cal N}=1$ theory. In two
1674: dimensions, there are two distinct superalgebras with two
1675: supercharges: the non-chiral $(1,1)$ algebra, and the chiral
1676: $(0,2)$ algebra. Given that the previous section was devoted to a
1677: review of ${\cal N}=(0,2)$ theories, the reader may guess this
1678: will be relevant for the vortex string. Let's now see that this is
1679: indeed the case \cite{ddt}. The ${\cal N}=1$ supersymmetry
1680: transformations for the vector multiplet fields are,
1681: %
1682: \be \delta A_\mu &=& -i\bar{\epsilon}\sigma_\mu\lambda +
1683: i\bar{\lambda}\sigma_\mu \epsilon \nn\\ \delta
1684: D&=&\bar{\epsilon}\bar{\sigma}^\mu {\cal D}_\mu \lambda + {\cal
1685: D}_\mu \lambda \bar{\sigma}^\mu\epsilon \nn\\ \delta \lambda &=&
1686: \ft12 \sigma^{\mu\nu}\epsilon F_{\mu\nu} + i\epsilon D .\ee
1687: %
1688: For each chiral multiplet $Q_i$, they take the form
1689: %
1690: \be \delta Q_i &=& \sqrt{2}\epsilon \psi_i\nn\\ \delta F_i&=&
1691: i\sqrt{2}\bar{\epsilon}\bar{\Dslash}\psi_i -
1692: 2i\bar{\epsilon}\lambda Q_i\nn\\ \delta \psi_i &=&
1693: \sqrt{2}\epsilon F_i + i\sqrt{2}(\!\Dslash Q_i) \bar{\epsilon} .\ee
1694: %
1695: Similar transformations also hold for the chiral multiplets
1696: $\tilde{Q}_i$ with the appropriate substitutions. Finally, the
1697: supersymmetry transformations for the adjoint chiral multiplet $A$
1698: take the form,
1699: %
1700: \be \delta A &=& \sqrt{2}\epsilon \eta\nn\\ \delta F &=&
1701: i\sqrt{2}\bar{\epsilon}\bar{\Dslash}\eta -
1702: 2i\bar{\epsilon}[\lambda, A]\nn\\ \delta \eta &=& \sqrt{2}\epsilon
1703: F + i\sqrt{2}(\!\Dslash A) \bar{\epsilon} .\ee
1704: %
1705: The key point here is that the vortex equations \eqn{vort},
1706: together with the requirement that $F_i=F=0$, provide solutions to
1707: $\delta\lambda = \delta \psi_i = \delta\eta = 0$. The latter
1708: condition $F=0$ is trivially satisfied when \eqn{survive} holds,
1709: for then $\tilde{Q}_i=0$, while $A$ remains constant. To see which
1710: supersymmetries are preserved in this case, it will suffice to
1711: examine the $\delta \psi_i$ transformation. Using \eqn{ds}, in the
1712: background of a stationary vortex so that ${\cal D}_+={\cal
1713: D}_-=0$, we have
1714: %
1715: \be \delta\psi_{-i} = -2\sqrt{2}i ({\cal D}_z Q_i)
1716: \bar{\epsilon}_- =0 \ \ \ {\rm and}\ \ \ \delta\psi_{+i} =
1717: 2\sqrt{2}i ({\cal D}_{\bar{z}}Q_i)\bar{\epsilon}_+ .\ee
1718: %
1719: In the background of a vortex, with the scalar field satisfying
1720: ${\cal D}_zQ_i=0$, we learn that $\bar{\epsilon}_-$ is the
1721: preserved supersymmetry; it descends to provide the supersymmetry
1722: variation parameter on the worldsheet. Meanwhile,
1723: $\bar{\epsilon}_+$ is the broken supersymmetry which generates a
1724: single Goldstino mode on the worldsheet. In our notation
1725: \eqn{goldy}, we have $\bar{\epsilon}_+=\chi_+/4$. (The $\chi_-$
1726: collective coordinate in \eqn{goldy} arises from the second
1727: supersymmetry transformation of the ${\cal N}=2$ theory. Its fate
1728: in our ${\cal N}=1$ theory will be discussed shortly). The spinors
1729: $\epsilon_\pm$ have definite, and opposite, chirality on the
1730: worldsheet. This is the statement that the worldsheet theory
1731: preserves chiral ${\cal N}=(0,2)$ supersymmetry, rather than
1732: ${\cal N}=(1,1)$.
1733:
1734:
1735:
1736: \para
1737: We have seen that, in the special case that a critical point of
1738: $\hW$ coincides with a mass \eqn{survive}, there exists at least
1739: one BPS vortex preserving ${\cal N}=(0,2)$ supersymmetry. But what
1740: happens if this is not the case? If \eqn{survive} is not
1741: satisfied, then there can be no BPS vortex solutions. To see this,
1742: note that \eqn{whatsleft} tells us that $\tilde{Q}_k$ gains
1743: an expectation value in the 4d vacuum. This means it cannot now
1744: remain constant but, must wind asymptotically to ensure that its
1745: kinetic term remains finite. A putative BPS vortex must now
1746: satisfy,
1747: %
1748: \be {\cal D}_zQ_i={\cal D}_z \tilde{Q}_i=0 .\ee
1749: %
1750: Yet $Q_i$ and $\tilde{Q}_i$ have opposite charges. A standard
1751: theorem in mathematics
1752: --- that a line bundle of negative degree has no non-zero
1753: holomorphic section
1754: --- states that there can only be simultaneous solutions to these
1755: equations when either $\tilde{Q}_i=0$ or $Q_i=0$. (See, for
1756: example, equation (3.43) of \cite{phases}). One can reach the same
1757: conclusion by noting that $A$ is now also sourced in the vortex
1758: background and $\delta\eta\neq 0$\footnote{The lack of BPS
1759: vortices in this case is entirely analogous to the statement that
1760: $F$-term vortices are not BPS in ${\cal N}=1$ theories \cite{ddt}.
1761: In our set-up, the value of $\partial \hW/\partial a$ evaluated at
1762: $a=m_k$ plays the role of the constant in the F-term in
1763: \cite{ddt}.}. Of course, simple topological arguments imply that
1764: vortex strings still exist. However, they must satisfy the full
1765: second order equations of motion, rather than the first order
1766: Bogomolnyi equations, and their tension is strictly greater than
1767: the BPS bound $T=2\pi v^2$.
1768:
1769:
1770: \subsection{Vortex Dynamics}
1771:
1772: In section 2, we described the ${\cal N}=(2,2)$ $U(k)$ theory on
1773: the vortex worldsheet that captures the dynamics of $k$ parallel
1774: vortex strings in ${\cal N}=2$ four dimensional gauge theories.
1775: We would like to understand how the worldsheet theory reacts to
1776: the superpotential $\hW(A)$, breaking the four dimensional
1777: supersymmetry from ${\cal N}=2$ to ${\cal N}=1$. We have seen
1778: above that the vortices in the theory with superpotential $\hW(A)$
1779: are classically BPS, preserving ${\cal N}=(0,2)$ supersymmetry,
1780: when equation \eqn{survive} holds; otherwise there are no BPS
1781: vortices. We would like to see this from the worldsheet.
1782:
1783: \para
1784: In fact, there is a unique deformation on the vortex worldsheet
1785: that preserves ${\cal N}=(0,2)$ supersymmetry and reproduces the
1786: expected vacuum structure described above. Recall from Section 3.2
1787: that superpotentials in ${\cal N}=(0,2)$ theories are constructed
1788: from fermi multiplets. The only such multiplet with a suitable
1789: transformation under the $U(k)$ gauge symmetry is $\Xi$,
1790: containing $\chi_-$ and the complex auxiliary field $G_Z$. The
1791: worldsheet deformation is given by the ${\cal N}=(0,2)$
1792: superpotential,
1793: %
1794: \be S_{{\cal W}} &\equiv& -\frac{1}{\sqrt{2}} \Tr_k\left.\int
1795: d\theta^+\ \Xi\,J(\Sigma)\right|_{\bar{\theta}^+=0} - {\rm h.c.}
1796: \nn\\ &=& -\frac{1}{\sqrt{2}}\Tr_k\,\left.\int d\theta^+\
1797: \Xi\,\frac{\partial\hW(\Sigma)}{\partial\Sigma}\right|_{\bar{\theta}^+=0}-
1798: {\rm h.c.}\label{booty}\ee
1799: %
1800: (up to some overall, unfixed, constant of proportionality). Note
1801: that a superpotential of this form is a viable holomorphic term
1802: since $\bar{\cal D}_+\Xi=2i[\Sigma, Z]\equiv i\sqrt{2}E_\Xi$ and
1803: %
1804: \be \Tr\ E_\Xi J=\sqrt{2}\Tr\ \left([\Sigma,
1805: Z]\,\frac{\partial\hW(\Sigma)}{\partial\Sigma}\right)=0\ee
1806: %
1807: which satisfies the requirement \eqn{ej}. In principle there could
1808: also be $\sigma$-dependent deformations of the kinetic terms for
1809: $\Lambda_i$ and $\Xi$. As is common in supersymmetric field
1810: theories, we will have less control over these ``D-term"
1811: deformations, but will see that the superpotential \eqn{booty}
1812: captures much of the important physics.
1813:
1814: \para
1815: The deformation \eqn{booty} has implications for both the bosonic
1816: and fermionic zero modes of the vortex strings. We defer a
1817: discussion of the fermions to the next subsection; we start here
1818: by studying the bosonic zero modes. The extra bosonic term on the
1819: vortex worldsheet arising from \eqn{booty} is a
1820: potential\footnote{A note on dimensions: In 4d, $[\hW(A)]=3$,
1821: which ensures that the scalar potential has the correct
1822: dimensions: $[|\partial \hW/\partial A|^2]=4$. In 2d the auxiliary
1823: field has dimension $[\sigma]=1$, so that $[\partial
1824: \hW/\partial\sigma]=2$. The presence of the vortex tension, with
1825: $[T]=2$, means that the worldsheet scalar potential \eqn{kimmoy}
1826: has the correct scaling for the two dimensional worldsheet.}, %
1827: %
1828: \be V_{2d}= \Tr_k\left(T|G_Z|^2 +
1829: G_Z\frac{\partial\hW(\sigma)}{\partial\sigma} + {\rm h.c.}\right)
1830: =\frac{1}{T}\Tr_k\left|\frac{\partial\hW(\sigma)}{\partial\sigma}\right|^2 .
1831: \label{kimmoy}\ee
1832: %
1833: We will now show that this gives the expected vacuum structure by
1834: studying the $k=1$ vortex theory in some detail; the extension to
1835: $k>1$ then follows.
1836:
1837:
1838: \subsubsection{An Example: $k=1$ with $N_f=N_c$}
1839:
1840: To illustrate the role of the superpotential \eqn{booty}, let's
1841: look at the familiar $k=1$ theory of a single vortex in the case
1842: with $N_f=N_c$ flavors. As we discussed in detail in Section 2,
1843: when $\hW=0$ the internal moduli space is ${\bf CP}^{N_c-1}$ with
1844: $\phi_i$ providing homogeneous coordinates. Once we turn on the
1845: superpotential $\hW$, the bosonic part of the worldsheet theory is
1846: given by
1847: %
1848: \be {\cal L}_{\rm bose}= T|\partial_mz|^2 + \sum_{i=1}^{N_c}\left( |{\cal
1849: D}_m\phi_i|^2 - 2|\sigma-m_i|^2|\phi_i|^2\right) +
1850: D(\sum_{i=1}^{N_c}|\phi_i|^2-r) - \frac{1}{T}\left|\frac{\partial
1851: \hW}{\partial \sigma}\right|^2 + \frac{\theta}{2\pi}u_{01} .\nn\ee
1852: %
1853: %
1854: In the presence of distinct, non-zero masses $m_i$, this
1855: worldsheet theory has a supersymmetric ground state (i.e. with
1856: vanishing vacuum energy) at
1857: %
1858: \be |\phi_j|^2=r\delta_{ij}\ \ \ ,\ \ \ \sigma=m_i\ee
1859: %
1860: only if $\hW(\sigma)$ has a critical point at $\sigma=m_i$
1861: %
1862: \be \left.\frac{\partial \hW(\sigma)}{\partial
1863: \sigma}\right|_{\sigma=m_i}=0 .\label{again}\ee
1864: %
1865: This coincides with the expectations of the previous section: BPS
1866: vortices only exist when \eqn{again} holds.
1867:
1868: \para
1869: When the masses do not coincide with the critical points, and
1870: there are no BPS vortices, the potential $|\partial\hW/\partial
1871: \sigma |^2/T$ determines the vacuum energy of the vortex string.
1872: One could try to compare this to the excess tension of the non-BPS
1873: vortex string, above the bound $T=2\pi v^2$, but this
1874: unfortunately suffers from the previously mentioned ambiguity in
1875: classical wavefunction renormalization for $\chi_-$ which also
1876: affects the coefficient in front of $|G_Z|^2$.
1877:
1878:
1879: \para
1880: If the hypermultiplet masses vanish, $m_i=0$, then the story is a
1881: little different. We may now set $\sigma=0$ in the vacuum (recall
1882: that we assumed $\hW$ does not contain a linear piece, so
1883: $\sigma=0$ is guaranteed to be a critical point). The full ${\bf
1884: CP}^{N_c-1}$ bosonic moduli space is now restored. This is in
1885: agreement with expectations from four dimensions, where we may
1886: happily construct any vortex string, built around the vacuum with
1887: $A=\tilde{Q}_i\equiv0$. We conclude that the superpotential
1888: $\hW(A)$ does not affect the bosonic zero modes in this case, a
1889: point made previously in \cite{syss}. However, the superpotential
1890: does still affect the fermi zero modes. We now turn to a study of
1891: these.
1892:
1893: \subsection{Fermions}
1894:
1895: We will study the fermions in the case with vanishing
1896: hypermultiplet masses $m_i=0$. Of all the Dirac equations in
1897: \eqn{diracs}, only that for $\eta$ is modified by the
1898: superpotential. It now reads
1899: %
1900: \be -\frac{i}{e^2}\Dbarslash\eta-\frac{i\sqrt{2}}{e^2}
1901: [A,\bar{\lambda}]-\sqrt{2}\tilde{Q}^\dagger_i\bar{\psi}_i-
1902: \sqrt{2}\bar{\tilde{\psi}}_i{Q}_i^{\dagger} -
1903: \frac{\partial^2\hW(A)}{\partial A^2}\bar{\eta}=0 .\ee
1904: %
1905: In the background of the vortex, we may again set
1906: $A=\tilde{Q}_i=0$. This means that all right-moving fermi zero
1907: modes
1908: --- those donated by $\lambda_+$ and $\bar{\psi}_{+i}$ ---
1909: remain the same as in the ${\cal N}=(2,2)$ case, given by
1910: solutions to
1911: %
1912: \be \sqrt{2}{\cal D}_z\lambda_+ &=& -e^2\,Q_i\bar{\psi}_{+i}\ , \nn\\
1913: \sqrt{2}{\cal D}_{\bar{z}} \bar{\psi}_{+i}&=&-
1914: Q_i^\dagger\lambda_+\label{same} .\ee
1915: %
1916: If the lowest order term in the superpotential $\hW$ is cubic or
1917: higher, then the left-moving fermi zero modes are similarly
1918: unaffected. However, if the superpotential $\hW(A)$ includes a
1919: quadratic mass term
1920: %
1921: \be \hW(A)=\mu_2 A^2 + \ldots \ee
1922: %
1923: then the equations for the left moving fermi zero modes become
1924: %
1925: \be \sqrt{2}i{\cal D}_{\bar{z}} \eta_- &=& -e^2
1926: \bar{\tilde{\psi}}_{-i} Q_i^\dagger - \sqrt{2}\mu_2\bar{\eta}_-\nn\\
1927: \sqrt{2}i{\cal D}_z \bar{\tilde{\psi}}_{-i} &=&
1928: \eta_-Q_i .\label{newbo}\ee
1929: %
1930: These equations are no longer related to the bosonic zero mode
1931: equations \eqn{bogzero}: this is to be expected since, in breaking
1932: to ${\cal N}=1$ supersymmetry, we have lost the half of
1933: supersymmetry which ensured the correspondence between bosonic
1934: zero modes and left-moving fermionic zero modes. Nevertheless, as
1935: stressed in \cite{syss}, the Dirac equations \eqn{newbo} must
1936: still admit the same number of solutions as the equations with
1937: $\mu_2=0$. This follows from the fact that the zero modes are
1938: chiral on the worldsheet, and cannot gain a mass through a
1939: deformation. For a single $k=1$ vortex in the $U(2)$ gauge theory,
1940: \eqn{newbo} was analyzed in \cite{syss}, both perturbatively in
1941: $\mu_2\rho$, as well as in the large $\mu_2$ limit.
1942:
1943: \para
1944: To summarize, we learn that the deformation leaves the fermi zero
1945: modes untouched unless $\mu_2\neq 0$, in which case it deforms the
1946: profile of the left-moving fermi zero modes only. However, the
1947: number of zero modes on the worldsheet remains the same. Let us
1948: now compare this with the predictions from the proposed worldsheet
1949: deformation \eqn{booty}.
1950:
1951: \subsubsection*{Implications for Worldsheet Fermions}
1952:
1953: In the presence of the superpotential $\hW$, the fermionic terms
1954: in the $U(k)$ worldsheet theory read\footnote{As we mentioned
1955: previously, the deformation from ${\cal N}=2$ to ${\cal N}=1$ may
1956: also induce a finite wavefunction renormalization of the
1957: left-moving fermion kinetic terms. We will not consider this
1958: here.}
1959: %
1960: \be {\cal L}_{\rm fermi}&=&2iT\ \Tr_k\left(\bar{\chi}_-{\cal
1961: D}_+\chi_-+\bar{\chi}_+{\cal D}_-\chi_+\right) +
1962: 2i\sum_{i=1}^{N_c}\left(\bar{\xi}_{-i}{\cal
1963: D}_+\xi_{-i}+\bar{\xi}_{+i}{\cal D}_-\xi_{+i}\right) \nn\\
1964: &&
1965: -\sqrt{2}\,\Tr_k\left([\bar{\chi}_-,[\sigma,\chi_+]]-[\bar{\chi}_+,[\bar{\zeta}_-,
1966: z]]-[\bar{\xi}_-,[\bar{\zeta}_+,z]]\right) + {\rm h.c.} \label{phew}\\
1967: && -\sqrt{2}\sum_{i=1}^{N_c}\left( \bar{\xi}_{-i}\sigma\xi_{+i}
1968: -\bar{\xi}_{+i}\bar{\zeta}_-\phi_i +
1969: \bar{\xi}_{-i}\bar{\zeta}_+\phi_i\right) +
1970: \Tr_k\left(\chi_-\frac{\partial^2 \hW(\sigma)}{\partial
1971: \sigma^2}\bar{\zeta}_+\right) + {\rm h.c.}\ . \nn\ee
1972: %
1973: The ${\cal N}=(0,2)$ superpotential is responsible for only the
1974: final term. Integrating out the auxiliary fermions $\zeta_\pm$
1975: again gives constraints on the dynamical fermions,
1976: %
1977: \be \sum_i\phi_i\bar{\xi}_{+i}+[z,\bar{\chi}_+]=0\ \ \ {\rm and}\
1978: \ \ \ \sum_i\phi_i\bar{\xi}_{-i}+[z,\bar{\chi}_-]=
1979: \frac{\partial^2\hW(\sigma)}{\partial \sigma^2}\chi_- .\ee
1980: %
1981: We see that the right-moving fermions are unaffected by the
1982: superpotential, in agreement with the Dirac equations \eqn{same}.
1983: Similarly, if $\hW$ has no quadratic term, so $\mu_2=0$, then the
1984: left-moving constraints are also left unchanged if we set
1985: $\sigma=0$ (we shall see the role played by a non-zero $\sigma$
1986: shortly). However, when $\mu_2\neq 0$, setting $\sigma=0$ still
1987: leaves deformed constraints on the left-moving fermions. For
1988: example, in the case of a single $k=1$ vortex, the constraints
1989: read
1990: %
1991: \be
1992: \sum_{i=1}^{N_c}{\phi}_i\bar{\xi}_{-i}=\mu_2{\chi}_- .\label{02cons}\ee
1993: %
1994: It's worth making a comment on this point. In the ${\cal N}=(0,2)$
1995: theory, we have defined the left-moving worldsheet fermions such
1996: that their kinetic terms are diagonal:
1997: $\bar{\chi}_-\partial_+\chi_-+\bar{\xi}_{-i}{\cal D}_+\xi_{-i}$.
1998: The constraint \eqn{02cons} holds in this basis. It is always
1999: possible to redefine the fermions so that the constraint
2000: \eqn{02cons} reverts to the original ${\cal N}=(2,2)$ constraint
2001: \eqn{fermcons},
2002: %
2003: \be
2004: \bar{\xi}'_{-i}=\bar{\xi}_{-i}-\frac{\mu_2}{r}\bar{\phi}_i\chi_- \
2005: \ \ \Rightarrow \ \ \ \sum_{i=1}^{N_c}\phi_i\,\bar{\xi}'_{-i}=0 .\ee
2006: %
2007: This will then lead to a non-diagonal form for the fermion kinetic
2008: terms.
2009:
2010: \para
2011: It was argued in \cite{syss} that, even in the presence of the
2012: four-dimensional superpotential $\hW(A)=\mu_2A^2$, the worldsheet
2013: theory of the vortex string still retains ${\cal N}=(2,2)$
2014: supersymmetry. This argument was based on the survival of the
2015: left-moving fermi zero modes, and the lack of a suitable ${\cal
2016: N}=(0,2)$ deformation of the ${\bf CP}^{N_c-1}$ sigma-model. We
2017: disagree with this conclusion. The vortex worldsheet theory is not
2018: described by a ${\bf CP}^{N_c-1}$ sigma-model, but rather by a
2019: ${\bf C}\times {\bf CP}^{N_c-1}$ sigma-model and, as we have seen,
2020: there is a suitable deformation of the latter in which $\chi_-$,
2021: the left-moving fermion in ${\bf C}$, mixes with $\xi_{-i}$.
2022: Moreover, this mixing is necessary to correctly capture the
2023: bosonic properties of the vortex with arbitrary superpotential and
2024: masses. As we explained above, to see this mixing between $\chi_-$
2025: and $\xi_{-i}$ from an explicit analysis of the fermions would
2026: require us to solve the fermi zero mode equations \eqn{newbo}, and
2027: take their overlap to determine both the kinetic terms and the
2028: constraint condition for the Grassmann collective coordinates of
2029: the vortex.
2030:
2031: \subsection{Symmetries and Other Aspects}
2032:
2033: We now discuss various further aspects of the worldsheet theory,
2034: starting with an analysis of the symmetries. We will show that the
2035: worldsheet superpotential has the correct properties under
2036: R-symmetry transformations to be induced by the superpotential
2037: $\hW(A)$. The addition of the superpotential $\hW(A)$ breaks both
2038: the $U(1)_R$ and the $U(1)_V$ symmetries in four dimensions. If
2039: the superpotential takes the form,
2040: %
2041: \be \hW(A) = \sum_{n=2} \mu_n A^n .\label{mun}\ee
2042: %
2043: Treating the parameters $\mu_n$ as spurion fields, the symmetry is
2044: restored if $\mu_n$ carries charge $(2-2n, 2)$ under $U(1)_R\times
2045: U(1)_V$. Let us check that these charges descend to the worldsheet
2046: theory. The deformation \eqn{booty} once again destroys both
2047: $U(1)_R$ and $U(1)_V$ on the worldsheet, this time through the
2048: presence of the worldsheet fermi interactions. The final term in
2049: \eqn{phew} is\footnote{The presence of $\bar{\zeta}_+$ in this
2050: expression, rather than $\zeta_+$, is crucial in this analysis. It
2051: follows from the component expansion \eqn{02tc} and ultimately
2052: from the fact $\Sigma$ arises from the decomposition of a $(2,2)$
2053: twisted chiral multiplet as opposed to a $(2,2)$ chiral
2054: multiplet.},
2055: %
2056: \be \sum_n n(n-1)\mu_n \ \Tr_k\,\left(\chi_-\sigma^{n-2}
2057: \,\bar{\zeta_+}\right) .\label{yippee}\ee
2058: %
2059: Examining the table in Section \ref{symmetries}, we see that the
2060: $U(1)_R\times U(1)_V$ worldsheet symmetry is again restored if
2061: $\mu_n$ is assigned charges $(2-2n,2)$, in agreement with the
2062: analysis in four dimensions.
2063:
2064:
2065: \para
2066: Note that the $U(1)_Z$ symmetry on the worldsheet, which arises
2067: from rotational invariance in the $z=x^1+ix^2$ plane, is left
2068: unbroken by the deformation \eqn{yippee} as, indeed, it must be.
2069:
2070: \subsubsection*{Discrete Symmetries}
2071:
2072: One can also check that the deformation on the worldsheet is consistent
2073: with the discrete symmetries of the bulk theory\footnote{We thank M. Shifman and
2074: A. Yung for stressing the importance of this.}. We start by considering the action
2075: of parity, defined by
2076: %
2077: \be P: x^i \rightarrow -x^i\ \ \ \ i=1,2,3\ee
2078: %
2079: The original ${\cal N}=2$ theory can be written in terms of Dirac
2080: spinors. For example, the adjoint Dirac spinor is
2081: $\Psi = (\lambda,\bar{\eta})^T$. Parity maps $P: \Psi \rightarrow \gamma^0\Psi$, or
2082: %
2083: \be P: \lambda \leftrightarrow \bar{\eta} \ \ \ {\rm and}\ \ \ \
2084: P: \psi_i \leftrightarrow \bar{\tilde{\psi}}_i\label{fp}\ee
2085: %
2086: while for the complex adjoint scalar $P: A\rightarrow A^\star$. (The
2087: imaginary part is really a pseudoscalar). Because the $z=x^1+ix^2
2088: \rightarrow -z$ part of the parity transformation can be undone by
2089: the rotation $U(1)_Z$ on the worldsheet, we may restrict attention to the simpler
2090: parity transformation $P: x^3\rightarrow -x^3$, with $x^1$ and $x^2$ untouched.
2091: This is the parity action under which the vortex string remains
2092: invariant. It must therefore descend to the worldsheet. Indeed, as we
2093: reviewed in Section 2, $(\lambda,\psi)$ donate right-moving zero modes
2094: $\chi_+$ and $\xi_{+i}$,
2095: while $(\bar{\eta}, \bar{{\tilde{\psi}}})$ donate left-moving zero
2096: modes $\chi_-$ and $\xi_{- i}$. So the action of parity \eqn{fp} in the
2097: bulk also exchanges left and right-movers on the worldsheet.
2098:
2099: \para
2100: So much for the ${\cal N}=2$ theory. What happens in the presence
2101: of the ${\cal N}=1$ deformation? This pure parity symmetry \eqn{fp} is broken in
2102: the 4d theory because the interactions of $\lambda$ and $\eta$ are
2103: different. This is also seen in our ${\cal N}=(0,2)$ worldsheet
2104: theory where the interactions of left and right movers differ.
2105:
2106: \para
2107: The 4d ${\cal N}=2$ theory is also invariant under $CP$.
2108: Under charge conjugation, $C: B\rightarrow -B$ and the vortex is
2109: mapped onto the anti-vortex. So this cannot be a symmetry of the
2110: worldsheet. However, under the particular parity transformation
2111: %
2112: \be P': x^2 \rightarrow -x^2 \label{p1}\ee
2113: %
2114: with $x^1$ and $x^3$ invariant, we also have $B_3\rightarrow
2115: -B_3$. Moreover, the complex coordinate $z$ transverse to the
2116: vortex string is mapped to $P': z\rightarrow z^\star$. This ensures
2117: that the bosonic vortex solution is invariant under $CP'$.
2118: For example, we have
2119: %
2120: \be {\cal D}_z Q_i\ \stackrel{C}{\longrightarrow}\ {\cal
2121: D}_{z}Q^\dagger\ \stackrel{P'}{\longrightarrow}\ {\cal D}_{\bar
2122: z}Q^\dagger \ee
2123: %
2124: so the Bogomolnyi equation ${\cal D}_z Q=0$ remains invariant
2125: under $CP'$. When acting on the fermions, $CP'$ sends Weyl spinors to their
2126: complex conjugates,
2127: %
2128: \be CP': \psi_i\rightarrow -i\sigma_2\bar{\psi}_i\ \ \ , \ \ \ \
2129: CP': \lambda\rightarrow -i\sigma_2\bar{\lambda}\ \ \ ,\ \ \ {\rm etc.} \ee
2130: %
2131: This symmetry also descends to the worldsheet, where it acts as
2132: complex conjugation, as can be checked explicitly from the zero
2133: mode expressions of Section 2. We have,
2134: %
2135: \be CP': \phi_i \rightarrow \bar{\phi}_i \ \ \ {\rm and}\ \ \
2136: CP': \xi_{\pm i} \rightarrow \bar{\xi}_{\pm i}\ \ {\rm etc} \ee
2137: %
2138: Note that, just as $CP'$ in the 4d theory didn't exchange
2139: $\lambda$ and $\bar{\eta}$, so this symmetry on the worldsheet
2140: doesn't send left-movers to right-movers. This can be
2141: traced to the fact that the action $CP'$ under which the string
2142: is invariant doesn't affect $x^3$.
2143:
2144: \para
2145: Unlike the pure parity transformation, the $CP'$ symmetry survives
2146: the deformation to ${\cal N}=1$ supersymmetry. More precisely, the
2147: symmetry survives if the parameters in the superpotential ${\cal W}=\mu_n A^n$
2148: are real. Alternatively we can think of these parameters as transforming
2149: under $CP': \mu_n\rightarrow \mu_n^\star$. The same behavior is seen
2150: in the worldsheet theory. Invariance of the final term in \eqn{phew} requires
2151: that $CP':\mu_n\rightarrow \mu_n^\star$, in agreement with the 4d analysis.
2152:
2153:
2154: \subsubsection*{The Four-Fermi Term}
2155:
2156: So far we have neglected the role of $\sigma$ on the string
2157: worldsheet. In the ${\cal N}=(2,2)$ case, we saw that $\sigma$
2158: correctly takes into account the effect of the Yukawa couplings in
2159: four-dimensions, resulting in a four-fermi term \eqn{224fermi} on
2160: the worldsheet. It will play the same role here. The equation of
2161: motion \eqn{yukyuk} for the adjoint field $A$ is now changed by
2162: the superpotential $\hW(A)$. Even if the superpotential has
2163: $\mu_2=0$, so the profiles of both left and right-moving fermionic
2164: zero modes are the same as in the ${\cal N}=2$ theory, the
2165: solutions to the full equations of motion, including Yukawa
2166: sources for $A$, will necessarily differ. We would expect this to
2167: feed back into the worldsheet dynamics. As in the ${\cal N}=(2,2)$
2168: case, it is difficult to determine this explicitly, but thankfully
2169: the lifting of the zero modes is once again dictated by the
2170: symmetries of the problem.
2171:
2172: \para
2173: Let's start by examining the simplest case, with $\hW(A)=\mu_2
2174: A^2$, so that the fermionic constraint equation is given by
2175: \eqn{02cons}. Integrating out $\sigma$ on the worldsheet once
2176: again gives rise to a four-fermi term
2177: %
2178: \be {\cal L}_{\rm 4-fermi}=
2179: -\frac{|\bar{\xi}_{-i}\xi_{+i}|^2}{(r+2|\mu_2|^2/T)}\label{024fermi}\ee
2180: %
2181: which, up to an overall rescaling, looks the same as the ${\cal
2182: N}=(2,2)$ four-fermi term \eqn{224fermi}. However this is
2183: deceptive, for the constraints \eqn{02cons} ensure that
2184: \eqn{024fermi} now includes a component of $\chi_-$. Previously,
2185: as we discussed in Section \ref{symmetries}, $\chi_-$ was
2186: prohibited from appearing in the four-fermi term since it was a
2187: Goldstino mode in the ${\cal N}=(2,2)$ theory. It loses this
2188: protection in the ${\cal N}=(0,2)$ theory.
2189:
2190: \para
2191: If the superpotential contains quadratic and higher order terms,
2192: then integrating out $\sigma$ results not only in a four-fermi
2193: term on the worldsheet, but also in a slew of higher order fermion
2194: lifting terms. These terms are an interesting prediction of the
2195: deformation \eqn{booty}.
2196:
2197:
2198:
2199:
2200:
2201:
2202: \subsubsection*{A Comment on Anomalies}
2203:
2204:
2205: In Section \eqn{add}, we saw that additional fundamental ${\cal
2206: N}=2$ hypermultiplets in four dimensions contributed extra zero
2207: modes to the vortex string which were captured in the gauged
2208: linear sigma model by adding $(N_f-N_c)$ chiral multiplets in the
2209: anti-fundamental representation of the $U(k)$ worldsheet gauge
2210: group.
2211:
2212: \para
2213: There exists a trivial generalization in the ${\cal N}=1$ theories
2214: in which we add only four-dimensional chiral multiplets, instead
2215: of full hypermultiplets. For example, the addition of a single
2216: four dimensional chiral multiplet $Q$, transforming in the ${\bf
2217: N}_c$ of $U(N_c)$, will contribute both bosonic and fermionic zero
2218: modes to the vortex string. These live in an ${\cal N}=(0,2)$
2219: chiral multiplet $\tilde{\Phi}$ of the worldsheet theory,
2220: transforming in the $\bar{\bf k}$ of $U(k)$. In contrast, the
2221: addition of $\tilde{Q}$, transforming in the $\bar{\bf N}_c$ of
2222: $U(N_c)$, will contribute only fermi zero modes, living in an
2223: ${\cal N}=(0,2)$ fermi multiplet $\tilde{\Gamma}$ which transforms
2224: in the $\bar{\bf k}$ of $U(k)$.
2225:
2226: \para
2227: While the above observation is trivial, there is an interesting
2228: corollary in the quantum theory. The four-dimensional theory with
2229: unequal numbers of fundamental and anti-fundamental chiral
2230: multiplets is inconsistent at the quantum level, suffering a gauge
2231: anomaly. This inconsistency descends to the vortex worldsheet,
2232: which also suffers a $U(k)$ gauge anomaly unless the number of
2233: chiral multiplets $\tilde{\Phi}$ is equal to the number of fermi
2234: multiplets $\tilde{\Gamma}$. It would be interesting to study
2235: vortices in chiral, anomaly free four-dimensional gauge theories,
2236: to see if there is a corresponding delicate anomaly cancellation
2237: on the vortex worldsheet.
2238:
2239: \subsubsection*{The SQCD Limit}
2240:
2241: To reach the ${\cal N}=1$ SQCD limit of the four-dimensional
2242: theory, we send $\mu_2\rightarrow \infty$ to decouple the adjoint
2243: chiral multiplet $A$. On the worldsheet, this has the effect of
2244: decoupling the $U(k)$ adjoint chiral multiplet $\Sigma$. At the
2245: same time, the constraint on the left-moving fermions \eqn{02cons}
2246: becomes simply $\chi_-=0$, which effectively removes the fermi
2247: multiplet $\Xi$. The right-moving fermions on the worldsheet are
2248: still constrained to obey $\bar{\phi}\xi_{+i}=0$ (in the case
2249: $N_f=N_c$) while the left-moving fermions $\xi_{-i}$ are
2250: unconstrained. Nonetheless, the theory appears to be free of
2251: worldsheet gauge anomalies.
2252:
2253: \para
2254: In this limit, the four-dimensional theory develops an enhanced, chiral flavor
2255: symmetry $S[U(N_f)\times U(N_f)]$, rotating left and right movers independently.
2256: (The ``S'' here is to remind us that the overall $U(1)_B$ is part of the gauge
2257: group).
2258: In the presence of the FI parameter, this is broken spontaneously and the
2259: surviving symmetry in the vacuum is,
2260: %
2261: \be S[U(N_c)\times U(N_f-N_c)] \times U(N_f) \times U(1)_R\ee
2262: %
2263: Here the $U(1)_R$ is the anomaly-free R-symmetry. The same
2264: symmetry enhancement is also seen on the vortex worldsheet theory
2265: proposed above. There is once again a particular choice for the
2266: anomaly free R-current.
2267:
2268: \para
2269: There is an issue with the normalizability of the fermi zero modes
2270: in this limit. As $\mu_2\rightarrow 0$, the Dirac equation for the
2271: left-moving fermi zero modes become ${\cal
2272: D}_z\bar{\tilde{\psi}}_{-i}=0$ which has only non-normalizable
2273: solutions. This could be mirrored on the worldsheet by infinite
2274: kinetic terms for $\Gamma_i$, of the type that we neglected in the
2275: discussion above. Alternatively, one could add a suitable
2276: deformation to the 4d theory, such as the meson field considered
2277: in \cite{gsy}, which once again renders these zero modes finite.
2278:
2279:
2280:
2281:
2282: \subsection{A D-Brane Construction}
2283:
2284: %
2285: \EPSFIGURE{vib2.eps,height=160pt}{}
2286: %
2287: One can construct a D-brane configuration whose low-energy
2288: dynamics is governed by the four-dimensional theory of interest,
2289: namely ${\cal N}=2$ super QCD, broken to ${\cal N}=1$ by the
2290: addition of a superpotential $\hW(A)$ for the adjoint chiral
2291: multiplet. One starts with the usual Hanany-Witten set-up for four
2292: dimensional ${\cal N}=2$ gauge theories \cite{hanwit,w}. This
2293: consists of two parallel NS5-branes lying in the $012345$
2294: directions and separated a distance $\Delta X^6\sim l_s/e^2$ in
2295: the $X^6$ direction. The ${\cal N}=2$ $U(N_c)$ gauge theory lives
2296: on $N_c$ D4-branes, with worldvolume $01236$, which are suspended
2297: between these two NS5-branes, while $N_f$ D6-branes with
2298: worldvolume $0123789$ provide the hypermultiplets. To describe the
2299: deformation \eqn{def} to ${\cal N}=1$ supersymmetry, we introduce
2300: the complex coordinates
2301: %
2302: \be v= X^4+iX^5\ \ \ ,\ \ \ w=X^8+iX^9 .\ee
2303: %
2304: A superpotential $\hW(A)$ is induced on the D4-brane worldvolume
2305: if we bend the right-hand NS5-brane so that it no longer lies at
2306: the point $w=0$, but rather on the complex curve \cite{deboeroz}
2307: %
2308: \be w= \hW(v) .\nn\ee
2309: %
2310: Note that in the limit $\mu_n\rightarrow \infty$, with $\mu_n$
2311: defined in \eqn{mun}, the curved NS5-brane becomes multiple flat
2312: NS5-branes, lying a constant values of $v=X^4+iX^5$, given by the
2313: roots of $\hW$. This is the description of the superpotential
2314: first presented in \cite{first,second}.
2315:
2316: \para
2317: We may now pass through the series of moves described in
2318: \cite{vib}, turning on a FI parameter by separating the two
2319: NS5-branes in the $X^7$ direction, and identifying the vortices as
2320: stretched D2-branes. The final result is shown in figure 3 in the
2321: case of $N_f=N_c$. The figure shows a slice through $w=0$. The
2322: dots depict the roots of $\hW(v)$, where the curved NS5-brane
2323: intersects the $w=0$ plane; the ghostly dotted line shows where
2324: the NS5-brane has left this plane and is living at some other
2325: value of $w$. Figure 3 corresponds to a quartic superpotential,
2326: with three critical points. One can check that the theory on the
2327: D2-brane preserves ${\cal N}=(0,2)$ supersymmetry. It is clear
2328: from the brane set-up that the D2-brane has a supersymmetric
2329: ground state only when it may safely stretch from the curved
2330: NS5-brane to a D4-brane, remaining at constant $v=m_i$ and without
2331: leaving the safety of $w=0$. This requires
2332: %
2333: \be \left.\frac{\partial \hW(v)}{\partial v}\right|_{m_i}=0 .
2334: \label{itsgood}\ee
2335: %
2336: This is the brane perspective on the statement that BPS vortices
2337: only exist when \eqn{itsgood} is satisfied. It provides further
2338: evidence that a worldsheet superpotential of the form \eqn{booty}
2339: is required.
2340:
2341:
2342:
2343:
2344: \subsection{A Different Superpotential}
2345: \label{more}
2346:
2347: To end this section, we consider a different deformation of the
2348: ${\cal N}=2$ theory which breaks the four dimensional
2349: supersymmetry to ${\cal N}=1$. We add a superpotential of the
2350: form,
2351: %
2352: \be {\cal W}_{{\cal N}=2} = \sqrt{2}\sum_{i=1}^{N_f}\,\tilde{Q}_i
2353: {\cal V}_i(A) Q_i .\label{another}\ee
2354: %
2355: Here ${\cal V}_i(A)$ is an arbitrary holomorphic function of $A$.
2356: The four-dimensional quantum dynamics of theories of this type was
2357: previously studied in \cite{kap,seiji,rab,deboeroz}. We are here
2358: interested in the effect on the vortex worldsheet. In fact, we
2359: have already met one example of such a deformation that preserves
2360: ${\cal N}=2$ supersymmetry, because the complex mass term is of
2361: this form with ${\cal V}_i(A)=A-m_i$. In that case, we saw that
2362: the effect was not to induce a superpotential on the worldsheet,
2363: but instead to change the relationship between $(0,2)$ fermi and
2364: chiral fields,
2365: %
2366: \be \bar{\cal D}_+\Gamma_i=2i\Sigma\Phi_i\ \longrightarrow \
2367: \bar{\cal D}_+\Gamma_i=2i(\Sigma-m_i)\Phi_i . \ee
2368: %
2369: Given this, it is natural to conjecture that the general
2370: deformation \eqn{another} is captured by the worldsheet theory
2371: with the relationship,
2372: %
2373: \be \bar{\cal D}_+\Gamma_i= 2i{\cal
2374: V}_i(\Sigma)\Phi_i .\label{vsup}\ee
2375: %
2376: We will now provide evidence that this is indeed the case. We will
2377: show that the deformation \eqn{vsup} is in agreement with all
2378: symmetries of the theory, and reproduces the known behavior of the
2379: vortex. The details of the calculations are similar to those
2380: presented earlier, so we shall be brief.
2381:
2382: \para
2383: Let us firstly study what becomes of the vortex. We take the
2384: vacuum of the four-dimensional theory to be
2385: %
2386: \be Q^a_{\ i}=v\delta^a_{\ i}\ \ \ ,\ \ \ \tilde{Q}_i=0\ \ \ \
2387: A={\rm diag}(\nu_1,\ldots,\nu_{N_c})\label{4dvac}\ee
2388: %
2389: where $\nu_i$ is one of the roots of ${\cal V}_i$. If the $\nu_i$
2390: are all distinct, the situation is the same as the one we
2391: encountered in Section 2.5.1 with distinct masses $m_i$: there are
2392: $N_c$ different vortices, each supported by the winding of a
2393: different $Q_i$. In contrast, if all $\nu_i$ coincide, the full
2394: ${\bf CP}^{N_c-1}$ internal moduli space of the vortex is
2395: restored.
2396:
2397: \para
2398: Let us see how this is reproduced on the vortex worldsheet by the
2399: deformation \eqn{vsup}. For definiteness, we take a single $k=1$
2400: vortex string in the $N_f=N_c$ theory. The bosonic part of the
2401: worldsheet theory is given by,
2402: %
2403: \be {\cal L}_{\rm bose}=T\,|\partial_m z|^2 +
2404: \sum_{a=1}^{N_c}\left(|{\cal D}_m\phi^a|^2 - 2|{\cal
2405: V}_i(\sigma)|^2|\phi_i|^2\right)
2406: +D(\sum_{i=1}^{N_c}|\phi_i|^2-r)+\frac{\theta}{2\pi}u_{01} .\nn\ee
2407: %
2408: If the roots of $\nu_i$ of ${\cal V}_i(\sigma)$ are distinct, this
2409: theory has isolated vacua, given by
2410: %
2411: \be |\phi_j|^2=r\delta_{ij}\ \ \ ,\ \ \ \sigma
2412: =\nu_i .\label{tricky}\ee
2413: %
2414: However, there is an ambiguity here since ${\cal V}_i$ has
2415: multiple roots $\nu_i$. Suppose, for definiteness, that ${\cal
2416: V}_i(\sigma)$ is a polynomial of degree $P_i$. Then it appears
2417: that, for each $i=1,\ldots, N_c$, there are $P_i$ different vacua
2418: of the worldsheet theory. How are we to interpret these? In past
2419: examples \cite{memono,sy,vstring}, different vacua of the
2420: worldsheet corresponded to different physical vortices --- see
2421: Section 2.5.1. But we certainly don't want the same interpretation
2422: here because the four-dimensional theory doesn't have $P_i$
2423: distinct vortices, each with $Q_i$ winding asymptotically.
2424: Thankfully, the interpretation of the multiple worldsheet vacua in
2425: the present case is somewhat different. For fixed
2426: $i=1,\ldots,N_c$, the $P_i$ different vacua differ only in the
2427: value of the auxiliary field $\sigma$. The field $\sigma$ is to be
2428: integrated out, set equal to its classical, algebraic equation of
2429: motion. But there are $P_i$ different solutions to this algebraic
2430: equation. The theory is only complete if we specify which of these
2431: solutions we are to take. This means that the vacuum $\sigma =
2432: \nu_i$ chosen in \eqn{tricky} is not a dynamical variable, but
2433: rather a parameter of the worldsheet theory. We are therefore free
2434: to fix it as we please, and the only natural candidate is to
2435: equate it with the four-dimensional vacuum value $\nu_i$ in
2436: \eqn{4dvac}\footnote{The equation of motion for $\sigma$ includes
2437: a term bilinear in the fermions, seen explicitly in \eqn{ria}.
2438: The root of the equation of motion is taken to be the
2439: four-dimensional vacuum value $\nu_i$ when the fermions vanish,
2440: and is continuously connected to $\nu_i$ when the fermions turn
2441: on.}. The end result is a situation where the same worldsheet
2442: Lagrangian describes the vortex string in different
2443: four-dimensional vacua; the specific four-dimensional vacuum of
2444: interest appears as a boundary condition on the auxiliary $\sigma$
2445: field.
2446:
2447: \para
2448: As a check of the conjecture \eqn{vsup}, we can confirm that the
2449: $U(1)_R\times U(1)_V$ charges are consistent. If we write the
2450: superpotential as
2451: %
2452: \be {\cal V}_i(A) = \sum_{n=0}\ h^{(i)}_n A^n\ee
2453: %
2454: then we are required to assign spurion charge $(2-2n,0)$ to
2455: $h_n^{(i)}$. Let's check that this is in agreement with the
2456: worldsheet. The deformation \eqn{vsup} gives rise to the terms
2457: %
2458: \be L_{\rm vortex} = \ldots + \sqrt{2}\sum_n \ (n h_n^{(i)}
2459: \bar{\xi}_{-i}\sigma^{n-1}\phi_i\bar{\zeta_+}\ +\
2460: h_n^{(i)}\bar{\xi}_{-i}\sigma^n\xi_{+i})+\ldots \label{ria}\ee
2461: %
2462: from which we learn that $h_n^{(i)}$ must again be assigned charge
2463: $(2-2n,0)$ under the worldsheet $U(1)_R\times U(1)_V$.
2464:
2465:
2466:
2467: \newpage
2468: \section*{Appendix: The Alphabet}
2469:
2470: This appendix is included to help the reader keep track of the
2471: burgeoning conventions. The four dimensional fields are all
2472: components of ${\cal N}=1$ superfields,
2473: %
2474: \be
2475: \begin{array}{ll}
2476: A_\mu: & \mbox{4d gauge field in the vector multiplet $V$} \\
2477: A: & \mbox{Adjoint valued 4d scalar in the chiral multiplet $A$} \\
2478: Q_i: & \mbox{Fundamental 4d scalar in the chiral multiplet $Q_i$} \\
2479: \tilde{Q}_j: & \mbox{Fundamental 4d scalar in the chiral
2480: multiplet $\tilde{Q}_j$}
2481: \\ \lambda: & \mbox{Adjoint valued 4d
2482: fermion in the vector multiplet $V$} \\
2483: \eta: & \mbox{Adjoint valued 4d fermion in the chiral multiplet
2484: $A$} \\ \psi_i: & \mbox{Fundamental 4d
2485: fermion living in the chiral multiplet $Q_i$} \\
2486: \tilde{\psi}_j: & \mbox{Anti-fundamental 4d fermion in the chiral
2487: multiplet $\tilde{Q}_j$} .\end{array}\nn\ee
2488: %
2489: The worldsheet fields are all components of ${\cal N}=(0,2)$
2490: superfields as described in Section 3:
2491: %
2492: \be
2493: \begin{array}{ll}
2494: z: & \mbox{Worldsheet scalar arising from broken translational
2495: invariance,} \\ & \mbox{ in the chiral multiplet $Z$} \\
2496: \phi_i: & \mbox{Worldsheet scalar corresponding to orientation
2497: modes
2498: of the string,} \\ & \mbox{in the chiral multiplet $\Phi_i$} \\
2499: \sigma: & \mbox{Worldsheet auxiliary scalar in the chiral multiplet $\Sigma$} \\
2500: u_m: & \mbox{Worldsheet gauge field in the vector multiplet $U$} \\
2501: {\chi}_+ & \mbox{Worldsheet Goldstino fermion in the chiral
2502: multiplet $Z$}
2503: \\ \chi_- & \mbox{Worldsheet fermion in the fermion
2504: multiplet $\Xi$} \\
2505: \xi_{+i}: & \mbox{Worldsheet fermions living in the fermion
2506: multiplet $\Phi_i$.} \\ \xi_{-i}: & \mbox{Worldsheet fermions in
2507: the fermion multiplet $\Gamma_i$} \\ \bar{\zeta}_+: &
2508: \mbox{Worldsheet
2509: auxiliary fermion living in the chiral multiplet $\Sigma$.} \\
2510: \zeta_-: & \mbox{Worldsheet auxiliary fermion in the vector
2511: multiplet $U$} .
2512:
2513: \nn\end{array}\ee
2514:
2515: \section*{Acknowledgement}
2516: We would like to thank Philip Argyres, Adam Ritz, and especially
2517: Misha Shifman
2518: and Alyosha Yung for helpful discussions. M.E. is supported in
2519: part by DOE grant FG02-84ER-40153. D.T. is supported by the Royal
2520: Society.
2521:
2522: \begin{thebibliography}{99}
2523:
2524: \small
2525: \parskip=0pt plus 2pt
2526:
2527: \bibitem{vib} A.~Hanany and D.~Tong,
2528: ``{\it Vortices, instantons and branes},''
2529: JHEP {\bf 0307}, 037 (2003)
2530: [arXiv:hep-th/0306150].
2531: %%CITATION = HEP-TH 0306150;%%
2532:
2533: \bibitem{auzzi}
2534: R.~Auzzi, S.~Bolognesi, J.~Evslin, K.~Konishi and A.~Yung,
2535: ``{\it Nonabelian superconductors: Vortices and confinement in N = 2 SQCD},''
2536: Nucl.\ Phys.\ B {\bf 673}, 187 (2003)
2537: [arXiv:hep-th/0307287].
2538: %%CITATION = HEP-TH 0307287;%%
2539:
2540:
2541: \bibitem{sy} M.~Shifman and A.~Yung,
2542: ``{\it Non-Abelian string junctions as confined monopoles},''
2543: Phys.\ Rev.\ D {\bf 70}, 045004 (2004)
2544: [arXiv:hep-th/0403149].
2545: %%CITATION = HEP-TH 0403149;%%
2546:
2547:
2548: \bibitem{vstring} A.~Hanany and D.~Tong,
2549: ``{\it Vortex strings and four-dimensional gauge dynamics},''
2550: JHEP {\bf 0404}, 066 (2004)
2551: [arXiv:hep-th/0403158].
2552: %%CITATION = HEP-TH 0403158;%%
2553:
2554: \bibitem{nick}
2555: N.~Dorey,
2556: ``{\it The BPS spectra of two-dimensional supersymmetric gauge theories with
2557: twisted mass terms},''
2558: JHEP {\bf 9811}, 005 (1998)
2559: [arXiv:hep-th/9806056].
2560: %%CITATION = HEP-TH 9806056;%%
2561:
2562: \bibitem{dht}
2563: N.~Dorey, T.~J.~Hollowood and D.~Tong,
2564: ``{\it The BPS spectra of gauge theories in two and four dimensions},''
2565: JHEP {\bf 9905}, 006 (1999)
2566: [arXiv:hep-th/9902134].
2567: %%CITATION = HEP-TH 9902134;%%
2568:
2569: \bibitem{scvs} D.~Tong,
2570: ``{\it Superconformal vortex strings},''
2571: JHEP {\bf 0612}, 051 (2006)
2572: [arXiv:hep-th/0610214].
2573: %%CITATION = HEP-TH 0610214;%%
2574:
2575: \bibitem{tasi} D.~Tong,
2576: ``{\it TASI lectures on solitons},''
2577: arXiv:hep-th/0509216.
2578: %%CITATION = HEP-TH 0509216;%%
2579:
2580:
2581: \bibitem{other}
2582: V.~Markov, A.~Marshakov and A.~Yung,
2583: ``{\it Non-Abelian vortices in N = 1* gauge theory},''
2584: Nucl.\ Phys.\ B {\bf 709}, 267 (2005)
2585: [arXiv:hep-th/0408235]. \\
2586: %%CITATION = HEP-TH 0408235;%%
2587:
2588: \bibitem{others}
2589: A.~Gorsky, M.~Shifman and A.~Yung,
2590: ``{\it Non-Abelian Meissner effect in Yang-Mills theories at weak
2591: coupling},'' Phys.\ Rev.\ D {\bf 71}, 045010 (2005)
2592: [arXiv:hep-th/0412082].
2593: %%CITATION = HEP-TH 0412082;%%
2594:
2595: \bibitem{bolog} S.~Bolognesi,
2596: ``{\it The holomorphic tension of vortices},''
2597: JHEP {\bf 0501}, 044 (2005)
2598: [arXiv:hep-th/0411075]; ``{\it The holomorphic tension of nonabelian vortices and the quark = dual-quark
2599: condensate},''
2600: Nucl.\ Phys.\ B {\bf 719}, 67 (2005)
2601: [arXiv:hep-th/0412241].
2602: %%CITATION = HEP-TH 0412241;%%
2603: %%CITATION = HEP-TH 0411075;%%
2604:
2605: \bibitem{syss} M.~Shifman and A.~Yung,
2606: ``{\it Non-abelian flux tubes in SQCD: Supersizing world-sheet supersymmetry},''
2607: arXiv:hep-th/0501211.
2608: %%CITATION = HEP-TH 0501211;%%
2609:
2610: \bibitem{gsy} A.~Gorsky, M.~Shifman and A.~Yung,
2611: ``{\it N = 1 supersymmetric quantum chromodynamics: How confined non-Abelian
2612: monopoles emerge from quark condensation},''
2613: arXiv:hep-th/0701040.
2614: %%CITATION = HEP-TH 0701040;%%
2615:
2616:
2617:
2618: \bibitem{no} H.~B.~Nielsen and P.~Olesen,
2619: ``{\it Vortex-Line Models For Dual Strings},''
2620: Nucl.\ Phys.\ B {\bf 61}, 45 (1973).
2621: %%CITATION = NUPHA,B61,45;%%
2622:
2623: \bibitem{jackr} R.~Jackiw and P.~Rossi,
2624: ``{\it Zero Modes Of The Vortex-Fermion System},''
2625: Nucl.\ Phys.\ B {\bf 190}, 681 (1981).
2626: %%CITATION = NUPHA,B190,681;%%
2627:
2628: \bibitem{wb} J.~Wess and J.~Bagger,
2629: ``{\it Supersymmetry and supergravity},'' Princeton Univ. Pr. (1992)
2630:
2631: \bibitem{stefan} S.~Vandoren and P.~van Nieuwenhuizen,
2632: ``{\it New instantons in the double-well potential},''
2633: Phys.\ Lett.\ B {\bf 499}, 280 (2001)
2634: [arXiv:hep-th/0010130].
2635: %%CITATION = HEP-TH 0010130;%%
2636:
2637: \bibitem{memono} D.~Tong,
2638: ``{\it Monopoles in the Higgs phase},''
2639: Phys.\ Rev.\ D {\bf 69}, 065003 (2004)
2640: [arXiv:hep-th/0307302].
2641: %%CITATION = HEP-TH 0307302;%%
2642:
2643: \bibitem{hh} A.~Hanany and K.~Hori,
2644: ``{\it Branes and N=2 theories in two dimensions},''
2645: Nucl.\ Phys.\ B {\bf 513}, 119 (1998)
2646: [arXiv:hep-th/9707192].
2647: %%CITATION = HEP-TH 9707192;%%
2648:
2649:
2650:
2651:
2652: \bibitem{ward}
2653: R.~S.~Ward,
2654: ``{\it Slowly Moving Lumps In The $CP^1$ Model In (2+1)-Dimensions},''
2655: Phys.\ Lett.\ B {\bf 158}, 424 (1985).
2656: %%CITATION = PHLTA,B158,424;%%
2657:
2658: \bibitem{ls} R.~A.~Leese and T.~M.~Samols,
2659: ``{\it Interaction of semilocal vortices},''
2660: Nucl.\ Phys.\ B {\bf 396}, 639 (1993).
2661: %%CITATION = NUPHA,B396,639;%%
2662:
2663: \bibitem{semi}
2664: M.~Shifman and A.~Yung,
2665: ``{\it Non-Abelian semilocal strings in N=2 supersymmetric QCD},''
2666: Phys.\ Rev.\ D {\bf 73}, 125012 (2006)
2667: [arXiv:hep-th/0603134].
2668: %%CITATION = HEP-TH 0603134;%%
2669:
2670: \bibitem{semi2} A.~Achucarro and T.~Vachaspati,
2671: ``{\it Semilocal and electroweak strings},''
2672: Phys.\ Rept.\ {\bf 327}, 347 (2000)
2673: [Phys.\ Rept.\ {\bf 327}, 427 (2000)]
2674: [arXiv:hep-ph/9904229].
2675: %%CITATION = HEP-PH 9904229;%%
2676:
2677: \bibitem{penin} A.~A.~Penin, V.~A.~Rubakov, P.~G.~Tinyakov and S.~V.~Troitsky,
2678: ``{\it What becomes of vortices in theories with flat directions},''
2679: Phys.\ Lett.\ B {\bf 389}, 13 (1996)
2680: [arXiv:hep-ph/9609257].
2681: %%CITATION = HEP-PH 9609257;%%
2682:
2683: \bibitem{davis} A.~Achucarro, A.~C.~Davis, M.~Pickles and J.~Urrestilla,
2684: ``{\it Vortices in theories with flat directions},''
2685: Phys.\ Rev.\ D {\bf 66}, 105013 (2002)
2686: [arXiv:hep-th/0109097].
2687: %%CITATION = HEP-TH 0109097;%%
2688:
2689: \bibitem{hanwit} A.~Hanany and E.~Witten,
2690: ``{\it Type IIB superstrings, BPS monopoles, and three-dimensional gauge
2691: dynamics},''
2692: Nucl.\ Phys.\ B {\bf 492}, 152 (1997)
2693: [arXiv:hep-th/9611230].
2694: %%CITATION = HEP-TH 9611230;%%
2695:
2696:
2697: \bibitem{w} E.~Witten,
2698: ``{\it Solutions of four-dimensional field theories via M-theory},''
2699: Nucl.\ Phys.\ B {\bf 500}, 3 (1997)
2700: [arXiv:hep-th/9703166].
2701: %%CITATION = HEP-TH 9703166;%%
2702:
2703: \bibitem{mmatrix} M.~Eto, Y.~Isozumi, M.~Nitta, K.~Ohashi and N.~Sakai,
2704: ``{\it Moduli space of non-Abelian vortices},''
2705: Phys.\ Rev.\ Lett.\ {\bf 96}, 161601 (2006)
2706: [arXiv:hep-th/0511088].
2707: %%CITATION = HEP-TH 0511088;%%
2708:
2709: \bibitem{mmreview} M.~Eto, Y.~Isozumi, M.~Nitta, K.~Ohashi and N.~Sakai,
2710: ``{\it Solitons in the Higgs phase: The moduli matrix approach},''
2711: J.\ Phys.\ A {\bf 39}, R315 (2006)
2712: [arXiv:hep-th/0602170].
2713: %%CITATION = HEP-TH 0602170;%%
2714:
2715:
2716: \bibitem{manton} N.~S.~Manton, ``{\em A Remark On The Scattering Of BPS Monopoles}'',
2717: Phys.\ Lett.\ B {\bf 110}, 54 (1982).
2718: %%CITATION = PHLTA,B110,54;%%
2719:
2720: \bibitem{samols} T.~M.~Samols,
2721: ``{\em Vortex Scattering}'' Commun.\ Math.\ Phys.\ {\bf 145}, 149
2722: (1992).
2723: %%CITATION = CMPHA,145,149;%%
2724:
2725:
2726:
2727: \bibitem{asy} R.~Auzzi, M.~Shifman and A.~Yung,
2728: ``{\it Composite non-Abelian flux tubes in N = 2 SQCD},''
2729: Phys.\ Rev.\ D {\bf 73}, 105012 (2006)
2730: [arXiv:hep-th/0511150].
2731: %%CITATION = HEP-TH 0511150;%%
2732:
2733: \bibitem{kt} K.~Hashimoto and D.~Tong,
2734: ``{\it Reconnection of non-abelian cosmic strings},''
2735: JCAP {\bf 0509}, 004 (2005)
2736: [arXiv:hep-th/0506022].
2737: %%CITATION = HEP-TH 0506022;%%
2738:
2739: \bibitem{lots} M.~Eto, K.~Hashimoto, G.~Marmorini, M.~Nitta, K.~Ohashi and W.~Vinci,
2740: ``{\it Universal reconnection of non-Abelian cosmic strings},''
2741: arXiv:hep-th/0609214.
2742: %%CITATION = HEP-TH 0609214;%%
2743:
2744:
2745: \bibitem{hw}
2746: C.~M.~Hull and E.~Witten, ``{\it Supersymmetric Sigma Models And
2747: The Heterotic String}'' Phys.\ Lett.\ B {\bf 160}, 398 (1985).
2748: %%CITATION = PHLTA,B160,398;%%
2749:
2750: \bibitem{ds}
2751: M.~Dine and N.~Seiberg, ``{\it $(2,0)$ Superspace},''
2752: Phys.\ Lett.\ B {\bf 180}, 364 (1986).
2753: %%CITATION = PHLTA,B180,364;%%
2754:
2755: \bibitem{phases}
2756: E.~Witten, ``{\it Phases of N = 2 theories in two dimensions},''
2757: Nucl.\ Phys.\ B {\bf 403} (1993) 159 [arXiv:hep-th/9301042].
2758: %%CITATION = HEP-TH 9301042;%%
2759:
2760: \bibitem{abs}
2761: A.~Adams, A.~Basu and S.~Sethi, ``{\it $(0,2)$ duality},''
2762: Adv.\ Theor.\ Math.\ Phys.\ {\bf 7}, 865 (2004)
2763: [arXiv:hep-th/0309226].
2764: %%CITATION = HEP-TH 0309226;%%
2765:
2766: \bibitem{dv} R.~Dijkgraaf and C.~Vafa,
2767: ``{\it A perturbative window into non-perturbative physics},''
2768: arXiv:hep-th/0208048.
2769: %%CITATION = HEP-TH 0208048;%%
2770:
2771:
2772:
2773:
2774: \bibitem{ddt} S.~C.~Davis, A.~C.~Davis and M.~Trodden,
2775: ``{\it N=1 supersymmetric cosmic strings},''
2776: Phys.\ Lett.\ B {\bf 405}, 257 (1997)
2777: [arXiv:hep-ph/9702360].
2778: %%CITATION = HEP-PH 9702360;%%
2779:
2780: \bibitem{deboeroz} J.~de Boer and Y.~Oz,
2781: ``{\it Monopole condensation and confining phase of N = 1 gauge theories via
2782: M-theory fivebrane},''
2783: Nucl.\ Phys.\ B {\bf 511}, 155 (1998)
2784: [arXiv:hep-th/9708044].
2785: %%CITATION = HEP-TH 9708044;%%
2786:
2787: \bibitem{first} S.~Elitzur, A.~Giveon and D.~Kutasov,
2788: ``{\it Branes and N = 1 duality in string theory},''
2789: Phys.\ Lett.\ B {\bf 400}, 269 (1997)
2790: [arXiv:hep-th/9702014].
2791: %%CITATION = HEP-TH 9702014;%%
2792:
2793: \bibitem{second} S.~Elitzur, A.~Giveon, D.~Kutasov, E.~Rabinovici and A.~Schwimmer,
2794: ``{\it Brane dynamics and N = 1 supersymmetric gauge theory},''
2795: Nucl.\ Phys.\ B {\bf 505}, 202 (1997)
2796: [arXiv:hep-th/9704104].
2797: %%CITATION = HEP-TH 9704104;%%
2798:
2799:
2800:
2801: \bibitem{kap} A.~Kapustin,
2802: ``{\it The Coulomb branch of N = 1 supersymmetric gauge theory with adjoint and
2803: fundamental matter},''
2804: Phys.\ Rev.\D {\bf 66}, 010001 (2002)
2805: arXiv:hep-th/9611049.
2806: %%CITATION = HEP-TH 9611049;%%
2807:
2808: \bibitem{seiji}
2809: T.~Kitao, S.~Terashima and S.~K.~Yang,
2810: ``{\it N = 2 curves and a Coulomb phase in N = 1 SUSY gauge theories with adjoint
2811: and fundamental matters},''
2812: Phys.\ Lett.\ B {\bf 399}, 75 (1997)
2813: [arXiv:hep-th/9701009].
2814: %%CITATION = PHLTA,B399,75;%%
2815:
2816:
2817: \bibitem{rab} A.~Giveon, O.~Pelc and E.~Rabinovici,
2818: ``{\it The Coulomb phase in N = 1 gauge theories with an LG-type
2819: superpotential},''
2820: Nucl.\ Phys.\ B {\bf 499}, 100 (1997)
2821: [arXiv:hep-th/9701045].
2822: %%CITATION = HEP-TH 9701045;%%
2823:
2824:
2825: %\bibitem{seiberg} N.~Seiberg,
2826: % ``{\it Electric-magnetic duality in supersymmetric nonAbelian gauge theories},''
2827: % Nucl.\ Phys.\ B {\bf 435}, 129 (1995)
2828: % [arXiv:hep-th/9411149].
2829: % %%CITATION = HEP-TH 9411149;%%
2830:
2831:
2832:
2833:
2834: \end{thebibliography}
2835:
2836:
2837: \end{document}
2838: