1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %%% %%%
3: %%% Sigma-Model Solitons in the Noncommutative Plane: %%%
4: %%% Construction and Stability Analysis %%%
5: %%% %%%
6: %%% by %%%
7: %%% %%%
8: %%% Andrei V. Domrin, Olaf Lechtenfeld, Stefan Petersen %%%
9: %%% %%%
10: %%% --- long version for hep-th --- %%%
11: %%% %%%
12: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: %%% %%%
14: %%% Latex source file: latex twice %%%
15: %%% %%%
16: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
17: %%% %%%
18: %%% hep-th/0412001 %%%
19: %%% Hannover preprint ITP-UH-25/04 %%%
20: %%% 1+33 pages, 5 eps figures, macros included %%%
21: %%% released 01 December 2004 %%%
22: %%% revised 18 March 2005 %%%
23: %%% %%%
24: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
25:
26: \documentclass[11pt]{article}
27: \usepackage{amssymb,amsmath,amsthm,amscd,bbm} % AMS packages
28: %\usepackage{latexsym}
29: \usepackage{graphicx}
30: \usepackage{here}
31: \usepackage{psfrag}
32:
33: \oddsidemargin -1mm
34: \evensidemargin -1mm
35: \topmargin -15mm
36: \textheight 230mm
37: \textwidth 165mm
38: \setlength{\parskip}{\medskipamount}
39: \setlength{\unitlength}{1mm}
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41:
42: \catcode`@=11
43: \renewcommand{\section}{\@startsection{section}{1}{0pt}{\medskipamount}
44: {\medskipamount}{\large\bf}}
45: \renewcommand{\subsection}{\@startsection{subsection}{1}{0pt}{\smallskipamount}
46: {\smallskipamount}{\normalsize\bf}}
47: \numberwithin{equation}{section}
48: \catcode`@=12
49:
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51:
52: \newcommand{\mbf}[1]{{\boldsymbol{#1}}}
53: \newcommand{\unity}{\mathbbm{1}}
54: \newcommand{\R}{{\mathbbm{R}}}
55: \newcommand{\C}{{\mathbbm{C}}}
56: \newcommand{\Z}{{\mathbbm{Z}}}
57: \newcommand{\NN}{{\mathbbm{N}}}
58: \newcommand{\Hcal}{{\mathcal H}}
59: \newcommand{\Dcal}{{\mathcal D}}
60: \newcommand{\Ecal}{{\mathcal E}}
61: \newcommand{\adag}{a^{\dagger}}
62:
63: \def\a{\alpha}
64: \def\ab{{\bar\alpha}}
65: \def\de{\delta}
66: \def\th{\theta}
67: \def\>{\rangle}
68: \def\<{\langle}
69: \def\={\ =\ }
70: \def\+{\dagger}
71: \def\pa{\partial}
72: \def\diff{\mathrm{d}}
73: \def\e{\mathrm{e}}
74: \def\ic{\mathrm{i}}
75: \def\tr{\mathrm{tr}}
76: \def\Tr{\mathrm{Tr}}
77: \def\sfrac#1#2{{\textstyle\frac{#1}{#2}}}
78:
79: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
80:
81: \begin{document}
82: \begin{titlepage}
83: \setcounter{page}{0}
84: \begin{flushright}
85: hep-th/0412001\\
86: ITP--UH-25/04\\
87: \end{flushright}
88: \vskip 2.0cm
89:
90: \begin{center}
91: {\LARGE\bf
92: Sigma-Model Solitons in the Noncommutative Plane: \\[8pt]
93: Construction and Stability Analysis
94: }
95: \\
96: \vspace{14mm}
97: {\Large
98: Andrei V. Domrin}
99: \\[5mm]
100: { \em
101: Department of Mathematics and Mechanics, Moscow State University\\
102: Leninskie gory, 119992, GSP-2, Moscow, Russia}
103: \\[5mm]
104: email: \texttt{domrin@mi.ras.ru}
105: \\[12mm]
106: {\Large
107: Olaf Lechtenfeld} \ \ and \ \
108: {\Large
109: Stefan Petersen}
110: \\[5mm]
111: { \em
112: Institut f\"ur Theoretische Physik, Universit\"at Hannover \\
113: Appelstra\ss{}e 2, 30167 Hannover, Germany }
114: \\[5mm]
115: email: \texttt{lechtenf, petersen @itp.uni-hannover.de}
116: \\[12mm]
117: %\small \today
118: \end{center}
119: \vspace{15mm}
120:
121: \begin{abstract}
122: \noindent
123: Noncommutative multi-solitons are investigated in Euclidean two-dimensional
124: U($n$) and Grassmannian sigma models, using the auxiliary Fock-space formalism.
125: Their construction and moduli spaces are reviewed in some detail,
126: unifying abelian and nonabelian configurations.
127: The analysis of linear perturbations around these backgrounds reveals
128: an unstable mode for the U($n$) models but shows stability
129: for the Grassmannian case. For multi-solitons which are diagonal
130: in the Fock-space basis we explicitly evaluate the spectrum of the Hessian
131: and identify all zero modes. It is very suggestive but remains to be proven
132: that our results qualitatively extend to the entire multi-soliton moduli space.
133: \end{abstract}
134:
135: \vfill
136: \end{titlepage}
137:
138: \tableofcontents
139:
140: \pagebreak
141:
142: \section{Introduction }
143: \noindent
144: Multi-solitons in noncommutative Euclidean two-dimensional sigma models are
145: of interest as static D0-branes inside D2-branes with a constant B-field
146: background~\cite{Lechtenfeld:2001uq} but also per se as nonperturbative
147: classical field configurations. Adding a temporal dimension, they represent
148: static solutions not only of a (WZW-extended) sigma model but also of the
149: Yang-Mills-Higgs BPS equations on noncommutative $\R^{2+1}$. In fact,
150: the full BPS sector of the Yang-Mills-Higgs system is, in a particular gauge,
151: given by the (time-dependent) solutions to the sigma-model equations of motion.
152:
153: In the commutative case, the classical solutions of Euclidean two-dimensional
154: sigma models have been investigated intensively by physicists as well as
155: by mathematicians (for a review see, e.g.~\cite{Zakrzewski}).
156: Prominent target spaces are U($n$) group manifolds or their Grassmannian
157: cosets Gr$(n,r)=\frac{\textrm{U}(n)}{\textrm{U}(r){\times}\textrm{U}(n{-}r)}$
158: for $1\le r<n$ (which are geodesic submanifolds of U($n$)).\footnote{
159: Most familiar are the Gr$(n,1)=\C P^{n-1}$ models.}
160: Any classical solution~$\Phi$ for the U($n$) sigma model can be constructed
161: iteratively, with at most $n{-}1$ so-called unitons as building blocks
162: \cite{Uhlenbeck, Wood}.
163: The subset of hermitian solutions, $\Phi^\+=\Phi$, coincides with the space
164: of Grassmannian solutions, $\Phi\in\text{Gr}(n,r)$ for some rank $r<n$,
165: of which again a subset is distinguished by a BPS property.\footnote{
166: not to be confused with the BPS condition for the Yang-Mills-Higgs system}
167: These BPS configurations are precisely the one-uniton solutions and can be
168: interpreted as (static) multi-solitons.
169:
170: Let us now perturb such multi-solitons within the configuration space
171: of the two-dimensional (i.e.~static) sigma-model, either within their
172: Grassmannian or, more widely, within the whole group manifold.
173: A linear stability analysis then admits a two-fold interpretation.
174: First, it is relevant for the semiclassical evaluation of
175: the Euclidean path integral, revealing potential {\it quantum\/} instabilities
176: of the two-dimensional model. Second, it yields the (infinitesimal) time
177: evolution of fluctuations around the static multi-soliton in the time-extended
178: three-dimensional theory, indicating {\it classical\/} instabilities if they
179: are present. More concretely, any static perturbation of a classical
180: configuration can be taken as (part of the) Cauchy data for a classical time
181: evolution, and any negative eigenvalue of the quadratic fluctuation operator
182: will give rise to an exponential runaway behavior, at least within the linear
183: response regime. Furthermore, fluctuation zero modes are expected to belong
184: to moduli perturbations of the classical configuration under consideration.
185: The current knowledge on the effect of quantum fluctuations is summarized in
186: \cite{Zakrzewski}.
187:
188: The Moyal deformation of Euclidean two-dimensional sigma models has been
189: described in~\cite{Lechtenfeld:2001aw},
190: including the construction of their BPS solutions.
191: The operator formulation of the noncommutative U($n$) theory turns it into a
192: (zero-dimensional) matrix model with U($\C^n{\otimes}\Hcal$) as its target
193: space, where $\Hcal$ denotes a Heisenberg algebra representation module.
194: Each BPS configuration again belongs to some Grassmannian, whose rank~$r$
195: can be finite or infinite. The latter case represents a smooth
196: deformation of the known commutative multi-solitons, while the former situation
197: realizes a noncommutative novelty, namely abelian multi-solitons
198: (they even occur for the U(1) model).
199: Although these solutions are available in explicit form, very little is known
200: about their stability.\footnote{ A few recent works \cite{Solovyov:2000xy,
201: Aganagic:2000mh,Gross:2000ss,Hadasz:2001cn,Fujii:2001wp,Durhuus:2001nj}
202: address perturbations of noncommutative solitons but not for sigma models.}
203: Our work sheds some light on this issue,
204: in particular for the interesting abelian case.
205:
206: The paper is organized as follows. In the next section we review the
207: noncommutative two-dimensional U($n$) sigma-model and present nested
208: classes of finite-energy solutions to its equation of motion. We focus
209: on BPS configurations (multi-solitons) and their moduli spaces, unifying
210: the previously different descriptions of abelian and nonabelian multi-solitons.
211: Section~3 is devoted to a study of the fluctuations around generic
212: multi-solitons and detects a universal unstable mode.
213: Section~4 analyzes the spectrum of the Hessian specifically for
214: U(1)~backgrounds and settles the issue for diagonal BPS solutions,
215: analytically as well as numerically.
216: Section~5 addresses the fluctuation problem in the nonabelian case
217: for the example of the noncommutative U(2)~model.
218: We close with a summary and a list of open questions.
219:
220:
221: \section{Noncommutative 2d sigma model and its solutions}
222: \noindent
223: Before addressing fluctuations, it is necessary to present the
224: noncommutative two-dimensional sigma model and its multi-soliton
225: solutions which we will set out to perturb later on.
226: These solutions have been discussed as static solutions of
227: a $2{+}1$ dimensional noncommutative sigma model
228: \cite{Lechtenfeld:2001aw,Lechtenfeld:2001gf,Wolf:2002jw,Ihl:2002kz}
229: which results from Moyal deforming the WZW-modified integrable sigma model
230: \cite{Ward:1988ie,Ward:1990vc,Ioannidou:1998jh}.
231: Here we review the results pertaining to the static situation.
232:
233: \subsection{Noncommutativity}
234: \noindent
235: A Moyal deformation of Euclidean $\R^2$ with coordinates $(x,y)$
236: is achieved by replacing the ordinary pointwise product of smooth functions
237: on it with the noncommutative but associative Moyal star product.
238: The latter is characterized by a constant positive real parameter~$\th$ which
239: prominently appears in the star commutation relation between the coordinates,
240: \begin{equation} \label{nccoord}
241: x\star y - y\star x \ \equiv\ [\,x\,,\,y\,]_\star \= \ic\,\th \ .
242: \end{equation}
243: For a concise treatment of the Moyal star product
244: see~\cite{Harvey:2001yn,Douglas:2001ba,Szabo:2001kg}.
245: It is convenient to work with the complex combinations
246: \begin{equation}
247: z \= x+\ic y \quad\textrm{and}\quad \bar{z} \= x-\ic y
248: \qquad\Longrightarrow\qquad [\,z\,,\,\bar{z}\,]_\star \= 2\th
249: \end{equation}
250: and to scale them to
251: \begin{equation}
252: a \= \tfrac{z}{\sqrt{2\th}} \quad\textrm{and}\quad
253: a^\dagger \= \tfrac{\bar{z}}{\sqrt{2\th}}
254: \qquad\Longrightarrow\qquad [\,a\,,\,a^\dagger\,]_\star \= 1 \ .
255: \end{equation}
256: A different realization of this Heisenberg algebra promotes the coordinates
257: (and thus all their functions) to noncommuting operators acting on an
258: auxiliary Fock space~$\Hcal$ but keeps the ordinary operator product.
259: The Fock space is a Hilbert space with orthonormal basis states
260: \begin{equation} \label{oscbasis}
261: \begin{aligned}
262: {}& \qquad\qquad\qquad\quad |m\>\=\sfrac{1}{\sqrt{m!}}\,(\adag)^m\,|0\>
263: \qquad\text{for}\quad m\in\NN_0 \quad\text{and}\quad a|0\>=0 \ ,\\[4pt]
264: {}& a\,|m\> \= \sqrt{m}\,|m{-}1\> \ ,\qquad
265: \adag\,|m\> \= \sqrt{m{+}1}\,|m{+}1\> \ ,\qquad
266: N\,|m\> \ := \adag a\,|m\> \= m\,|m\> \ ,
267: \end{aligned}
268: \end{equation}
269: therewith characterizing $a$ and $a^\dagger$ as standard annihilation
270: and creation operators. The star-product and operator formulations are
271: tightly connected through the Moyal-Weyl map:
272: Coordinate derivatives correspond to commutators with coordinate operators,
273: \begin{equation}
274: \sqrt{2\th}\,\pa_z \ \leftrightarrow\ -\text{ad}(\adag)\quad,\qquad
275: \sqrt{2\th}\,\pa_{\bar z} \ \leftrightarrow\ \text{ad}(a) \quad,
276: \end{equation}
277: and the integral over the noncommutative plane reads
278: \begin{equation}
279: \int\!\diff^2x\;f_{\star}(x) \= 2\pi\,\th\,\Tr_{\Hcal}\,f_{\mathrm{op}} \ ,
280: \end{equation}
281: where the function $f_{\star}$ corresponds to the operator~$f_{\mathrm{op}}$
282: via the Moyal-Weyl map and the trace is over the Fock space~$\Hcal$.
283: We shall work with the operator formalism but refrain from introducing
284: special notation indicating operators, so all objects are operator-valued
285: if not said otherwise.
286:
287: \subsection{Two-dimensional sigma model}
288: \noindent
289: The fields $\Phi$ of the noncommutative two-dimensional U($n$) sigma model
290: are unitary $n{\times}n$ matrices with operator-valued entries,
291: i.e.~$\Phi\in\text{U}(\C^n\otimes\Hcal)=\text{U}(\Hcal^{\oplus n})$,
292: and thus subject to the constraint
293: \begin{equation}
294: \Phi\,\Phi^\+ \= \unity_n\otimes\unity_{\Hcal} \= \Phi^\+\,\Phi \ .
295: \end{equation}
296: The Euclidean action of the model
297: coincides with the energy functional of its $2{+}1$ dimensional extension
298: evaluated on static configurations~\cite{Lechtenfeld:2001aw},\footnote{
299: $|A|^2 = \Tr(A^\+A)$ is the squared Hilbert-Schmidt norm of U($n$)-valued
300: operators on $\Hcal$. We restrict ourselves to finite-energy configurations,
301: i.e.~we demand that $[a,\Phi]$ exists and is Hilbert-Schmidt. Furthermore we
302: only consider solutions for which $\Delta\Phi$ is traceclass and $\Phi$
303: is a bounded operator in order for the below expressions to be well defined.}
304: \begin{equation} \label{E}
305: \begin{aligned}
306: E[\Phi] &\= 2\pi\,\th\,\Tr \bigl(
307: \pa_z\Phi^\+\,\pa_{\bar z}\Phi + \pa_{\bar z}\Phi^\+\,\pa_z\Phi \bigr) \\
308: &\= \pi\,\Tr \bigl( [a,\Phi]^\+\,[a,\Phi] + [a,\Phi^\+] [a,\Phi^\+]^\+ \bigr)\\
309: &\= 2\,\pi\,\Tr \bigl( [a,\Phi]^\+\,[a,\Phi] \bigr)
310: \= 2\,\pi\,\bigl| [a,\Phi] \bigr|^2 \\
311: &\= \pi\,\Tr \bigl( \Delta\Phi^\+\,\Phi + \Phi^\+\Delta\Phi \bigr) \ ,
312: \end{aligned}
313: \end{equation}
314: where the trace is taken over the Fock space $\Hcal$ as well as
315: over the U($n$) group space.
316: Here, we have introduced the hermitian Laplace operator $-\Delta$
317: which is defined via
318: \begin{equation}
319: \Delta{\cal O} \ :=\ \bigl[a,[\adag,{\cal O}]\bigr]
320: \= \bigl[\adag,[a,{\cal O}]\bigr]\ .
321: \end{equation}
322: Its kernel is spanned by functions only of $a$ or only of $\adag$.
323: Varying $E[\Phi]$ under the above constraint one finds the equation of
324: motion,\footnote{assuming that the appearing total derivative terms vanish}
325: \begin{equation} \label{fulleom}
326: 0 \= \bigl[a\,,\Phi^\+[\adag,\Phi]\bigr] + \bigl[\adag,\Phi^\+[a\,,\Phi]\bigr]
327: \= \Phi^\+ \, \Delta\Phi - \Delta\Phi^\+ \, \Phi \ .
328: \end{equation}
329:
330: Since the annihilation and creation operators above should be read as
331: $\unity_n\otimes a$ and $\unity_n\otimes\adag$, respectively,
332: this model enjoys a global U(1)$\times$SU$(n)\times$SU($n$) invariance under
333: \begin{equation} \label{globalsym}
334: \Phi \quad\longrightarrow\quad
335: (V\otimes\unity_{\Hcal})\,\Phi\,(W\otimes\unity_{\Hcal})
336: \end{equation}
337: with $V V^\+ = \unity_n = W W^\+$, which generates a moduli space
338: to any nontrivial solution of~(\ref{fulleom}).
339: Another obvious symmetry is induced by the ISO(2) Euclidean group
340: transformations of the noncommutative plane,
341: as generated by the adjoint action of $a$, $\adag$ and $N$.
342: Specifically, a global translation of the noncommutative plane,
343: \begin{equation}
344: (z\,,\,\bar z)\quad\longrightarrow\quad
345: (z+\zeta\,,\,\bar z+\bar\zeta)\=\sqrt{2\th}\,(a+\a\,,\,\adag+\ab)\ ,
346: \end{equation}
347: induces on (operator-valued) scalar functions~$f$ the unitary operation
348: \begin{equation} \label{globaltrans}
349: f\quad\longrightarrow\quad \e^{-\zeta\pa_z-\bar\zeta\pa_{\bar z}}\,f
350: \= \e^{\a\,\text{ad}(\adag)-\ab\,\text{ad}(a)}\,f
351: \= \e^{\a\adag-\ab a}\,f\,\e^{-\a\adag+\ab a}\ =:\ D(\a)\,f\,D(\a)^\+\ .
352: \end{equation}
353: Its action on states leads to coherent states, e.g.
354: \begin{equation} \label{coherent}
355: D(\a)\,|0\> \= \e^{\a\adag-\ab a}\,|0\>
356: \= \e^{-\frac12\ab\a}\,\e^{\a\adag}\,|0\> \ =:\ |\a\> \ .
357: \end{equation}
358: A global rotation of the noncommutative plane,
359: \begin{equation}
360: (z\,,\,\bar z)\quad\longrightarrow\quad
361: (\e^{\ic\,\vartheta} z\,,\,\e^{-\ic\,\vartheta} \bar z)
362: \= \sqrt{2\th}\,(\e^{\ic\,\vartheta} a\,,\,\e^{-\ic\,\vartheta} \adag)\ ,
363: \end{equation}
364: induces the unitary transformation
365: \begin{equation} \label{globalrot}
366: f\quad\longrightarrow\quad \e^{\ic\vartheta\,\text{ad}(\adag a)}\,f
367: \= \e^{\ic\vartheta\,\adag a}\,f\,\e^{-\ic\vartheta\,\adag a}
368: \ =:\ R(\vartheta)\,f\,R(\vartheta)^\+
369: \end{equation}
370: for $\vartheta\in\R/2\pi\Z$,
371: because $[N,f]=\bar{z}\pa_{\bar z}f-\pa_z f z$.
372: Applying $R(\vartheta)$ to a coherent state we obtain
373: \begin{equation}
374: R(\vartheta) |\a\> \= |\e^{\ic\,\vartheta} \a\>\ .
375: \end{equation}
376: Since the adjoint actions of both~$D(\a)$ and~$R(\vartheta)$ commute with
377: that of~$\Delta$, the energy functional is invariant under them,
378: and all translates and rotations of a solution to~(\ref{fulleom})
379: also qualify as solutions, with equal energy.
380: However, other unitary transformations will in general change the value of~$E$.
381:
382: There is a wealth of finite-energy configurations~$\Phi$ which fulfil
383: the equation of motion~(\ref{fulleom}).
384: A subclass of those is distinguished by possessing a smooth commutative limit,
385: in which $\Phi$ merges with a commutative solution. The latter have been
386: classified by \cite{Uhlenbeck, Wood}. We call these ``nonabelian'' because they
387: cannot appear in the U(1) case, where the commutative model is a free field
388: theory and does not admit finite-energy solutions.
389: Yet, there can (and do) exist non-trivial abelian finite-energy solutions
390: at finite~$\th$, whose $\th\to0$ limit is necessarily singular and
391: produces a discontinuous configuration. We term such solutions ``abelian''
392: even in case they are imbedded in a nonabelian group.
393:
394: Of particular interest are
395: configurations diagonal in the oscillator basis~(\ref{oscbasis}) as well as
396: in~$\C^n$, namely $\Phi=
397: \bigoplus_{i=1}^n\text{diag}\bigl(\{\e^{\ic\a_i^\ell}\}_{\ell=0}^\infty\bigr)$
398: with $\a_i^\ell\in\R$.
399: A short computation reveals that such a configuration
400: obeys the equation of motion~(\ref{fulleom}) if and only if the phases
401: satisfy $\e^{\ic\a_i^\ell}=\pm\e^{\ic\a_i}$ where the sign depends on~$\ell$.
402: Its energy is finite if the number of positive signs or the number of
403: negative signs in each block labelled by~$i$ is finite.
404:
405: \subsection{Grassmannian configurations}
406: \noindent
407: It is a daunting task to classify all solutions to the full equation of
408: motion~(\ref{fulleom}), and we shall focus on the subset of hermitian ones.
409: Any (not necessarily classical) hermitian configuration obeys
410: \begin{equation}
411: \Phi^\+ \= \Phi \qquad\Longrightarrow\qquad
412: \Phi^2 \= \unity_n\otimes\unity_{\Hcal} \ =:\ \unity \ ,
413: \end{equation}
414: and is conveniently parametrized by a hermitian projector~$P{=}P^\+$ via
415: \begin{equation}
416: \Phi \ =:\ \unity - 2P \= P^\perp - P\= \e^{\ic\pi P}
417: \qquad\text{with}\quad P^2 = P\ ,
418: \end{equation}
419: where $P^\perp=\unity{-}P$ denotes the complementary projector.
420: This allows one to define a topological charge like in the commutative case as
421: \cite{Matsuo:2000pj}
422: \begin{equation} \label{Qdef}
423: \begin{aligned}
424: Q[\Phi] &\= \sfrac{1}{4}\,\th\,\Tr \bigl(
425: \Phi\,\pa_z\Phi\,\pa_{\bar z}\Phi - \Phi\,\pa_{\bar z}\Phi\,\pa_z\Phi \bigr)\\
426: &\= \Tr \bigl( P\,[\adag,P]\,[a\,,P]-P\,[a\,,P]\,[\adag,P] \bigr) \\
427: &\= \bigl| P\,[a\,,P] \bigr|^2 - \bigl| [a\,,P]\,P \bigr|^2 \\
428: &\= \Tr \bigl( P - P\,a\,P\,\adag P + P\,\adag P\,a\,P \bigr) \\
429: &\= \Tr \bigl(P\,a\,(\unity{-}P)\,\adag P-P\,\adag(\unity{-}P)\,a\,P\bigr)\ ,
430: \end{aligned}
431: \end{equation}
432: which may be compared with the energy
433: \begin{equation} \label{Edef}
434: \begin{aligned}
435: \sfrac{1}{8\pi}\,E[\Phi] &\= \sfrac12\,\th\,\Tr \bigl(
436: \pa_z \Phi\,\pa_{\bar z}\Phi \bigr) \\
437: &\= \Tr \bigl( [\adag,P]\;[P\,,a] \bigr) \\
438: &\= \bigl| P\,[a\,,P] \bigr|^2 + \bigl| [a\,,P]\,P \bigr|^2 \\
439: &\=\Tr\bigl( P\,(a\,\adag{+}\adag a)\,P
440: - P\,a\,P\,\adag P - P\,\adag P\,a\,P \bigr) \\
441: &\= \Tr \bigl(P\,a\,(\unity{-}P)\,\adag P+P\,\adag(\unity{-}P)\,a\,P\bigr)\ .
442: \end{aligned}
443: \end{equation}
444:
445: Each hermitian projector~$P$ gives rise to the complementary projector
446: $\unity{-}P$, with the properties\footnote{
447: By a slight abuse of notation, we denote the energy and topological charge
448: as functionals of~$P$ again with the symbols $E$ and~$Q$, respectively.}
449: \begin{equation} \label{coP}
450: E[\unity{-}P] \= E[P] \qquad\text{and}\qquad
451: Q[\unity{-}P] \= -Q[P] \ .
452: \end{equation}
453: Comparing (\ref{Qdef}) and (\ref{Edef}) we get the relations
454: \begin{equation} \label{EandQ}
455: \sfrac{1}{8\pi}\,E[P] \= Q[P] + 2 \bigl| [a\,,P]\,P \bigr|^2
456: \= -Q[P] + 2 \bigl| P\,[a\,,P] \bigr|^2 \ ,
457: \end{equation}
458: which yield the BPS bound
459: \begin{equation} \label{bound}
460: E[P]\ \ge\ 8\pi\,\bigl| Q[P] \bigr|
461: \end{equation}
462: for any hermitian configuration.
463:
464: The set of all projectors unitarily equivalent to $P$ is called
465: the Grassmannian, and it is given by the coset
466: \begin{equation} \label{defGr}
467: \text{Gr}(P) \=
468: \frac{\text{U}(\Hcal^{\oplus n})}{\text{U(im$P$)}\times\text{U(ker$P$)}}\ .
469: \end{equation}
470: Thus, each given $P$ (and thus hermitian $\Phi$) belongs to a certain
471: Grassmannian, and the space of all hermitian~$\Phi$ decomposes into a
472: disjoint union of Grassmannians. The restriction to hermitian $\Phi$
473: reduces the unitary to a Grassmannian sigma model, whose configuration space
474: is parametrized by projectors $P$.
475: In the commutative case the topological charge $Q$ is an element of
476: $\pi_2(\text{Gr}(P))=\Z$, where $S^2$ is the compactified plane. Hence,
477: it is an invariant of the Grassmannian if one excludes from (\ref{defGr})
478: any singular unitary transformations with nontrivial winding at infinity.
479: It is less obvious how to properly extend this consideration to the
480: infinite-dimensional cases encountered here~\cite{Harvey:2001pd}.
481: We therefore take a pragmatic viewpoint and demand $Q[P]$ to be constant
482: throughout the Grassmannian. This may downsize the above coset $\text{Gr}(P)$
483: by restricting the set of admissible unitaries~$U$.
484:
485: Let us make this more explicit by computing $Q[U\!PU^\+]$.
486: To this end, we define
487: \begin{equation} \label{defomega}
488: \omega\ :=\ U^\+\,[a\,,\,U] \qquad\text{and}\qquad
489: \omega^\+\= U^\+\,[\adag,\,U] \ ,\qquad\text{with}\qquad
490: [a+\omega\,,\adag+\omega^\+]\=\unity \ ,
491: \end{equation}
492: as elements of the Lie algebra of U$(\Hcal^{\oplus n})$.\footnote{
493: In fact, the finite-energy condition (see footnote before (\ref{E}))
494: enforces $[a\,,U\!PU^\+]$ to be Hilbert-Schmidt which implies that
495: $[\omega\,,P]$ is Hilbert-Schmidt as well. Unfortunately, this does not
496: suffice to guarantee the constancy of $Q$ in $\text{Gr}(P)$.
497: On the other hand, $\omega$ itself need not even be bounded, as in the
498: example of (\ref{globalrot}) where $\omega=(\e^{\ic\vartheta}{-}1)a$.}
499: A short calculation yields
500: \begin{align} \label{QU}
501: Q[U\!PU^\+] &\= \Tr \bigl(
502: P\,[\adag{+}\omega^\+,P]\,[a{+}\omega\,,P] -
503: P\,[a{+}\omega\,,P]\,[\adag{+}\omega^\+,P] \,\bigr) \\[4pt]
504: &\= Q[P]\;+\;\Tr\,P\,\bigl( \,
505: [\adag\!,P]\omega + \omega[\adag\!,P] - [a,P]\omega^\+\! - \omega^\+[a,P]
506: - \omega^\+(\unity{-}P)\omega + \omega(\unity{-}P)\omega^\+ \bigr) P \ ,
507: \nonumber
508: \end{align}
509: which constrains $\omega$ in terms of~$P$. In case of $\Tr P<\infty$,
510: this indeed reduces to
511: \begin{equation}
512: Q[U\!PU^\+] \= Q[P]\ +\ \Tr\,P\,\bigl( \,
513: [\omega\,,\adag] + [a\,,\omega^\+] + [\omega\,,\omega^\+] \bigr) \= Q[P] \ .
514: \end{equation}
515:
516: Alternatively, we may calculate the infinitesimal variation of $Q[P]$
517: under $P\to P{+}\de P$. Remembering that $P$ and $\de P$ are bounded and that
518: their commutators with $a$ or $\adag$ are Hilbert-Schmidt, we get
519: \begin{equation} \label{Qvar}
520: \de Q[P] \= \Tr \bigl(
521: \bigl[\adag,[P,\de P][a,P]\bigr]-\bigl[a,[P,\de P][\adag,P]\bigr] \bigr)
522: \ +\ 3\,\Tr \,\de P \bigl( [\adag,P][a,P] - [a,P][\adag,P] \bigr)
523: \end{equation}
524: with the first trace being a ``boundary term''.
525: Variations inside the Grassmannian are given by
526: \begin{equation}
527: \de P \= [ \Lambda_{\text{o}}\,,P ]
528: \qquad\text{with}\qquad \Lambda_{\text{o}}^\+ = -\Lambda_{\text{o}}
529: \qquad\text{and}\qquad
530: P\Lambda_{\text{o}}P \=0\= (\unity{-}P)\Lambda_{\text{o}}(\unity{-}P)
531: \end{equation}
532: (see Section~3.2 below), and thus
533: \begin{equation}
534: \de Q[P] \= \Tr \bigl(
535: \bigl[a , \Lambda_{\text{o}} [\adag,P]\bigr] -
536: \bigl[\adag , \Lambda_{\text{o}} [a,P]\bigr] \bigr)
537: \ +\ 3\,\Tr \, \bigl[ P\,,\,
538: \Lambda_{\text{o}} [a,P][\adag,P] - \Lambda_{\text{o}} [\adag,P][a,P] \bigr]\ .
539: \end{equation}
540: Hence, if $\Lambda_{\text{o}}$ is bounded, the second term vanishes
541: and $\de Q[P]$ reduces to the boundary term.\footnote{
542: Yet again, this condition is too strong a demand as the examples of
543: (\ref{globaltrans}) and (\ref{globalrot}) indicate.}
544:
545: Formally, another invariant is the rank of the projector,
546: \begin{equation}
547: r \ :=\ \text{rank}(P) \= \text{dim(im}P) \ ,
548: \end{equation}
549: which may differ from $Q$ when being infinite.
550: Let us consider the class of projectors which can be decomposed as
551: \begin{equation} \label{specialP}
552: P \= U\,\bigl( \widehat{P}\,+\,P' \bigr) \, U^\+ \qquad\text{with}\qquad
553: \widehat{P} = \bar{P}\otimes\unity_\Hcal \qquad\text{and}\qquad
554: \widehat{P}\,P' =0= P'\,\widehat{P} \ ,
555: \end{equation}
556: where $U$ is an admissible unitary transformation, $\bar{P}$ denotes a
557: constant projector, and the rank~$r'$ of $P'$ is finite.
558: In this case the difference of $r$ and $Q$ is determined by the invariant
559: U($n$) trace~\cite{Otsu}
560: \begin{equation}
561: R[P] \= \tr\,\bar P\ \in \{0,1,\dots,n\}\ ,
562: \end{equation}
563: and such projectors are characterized by the pair $(R,Q)$.
564: Formally, the total rank then becomes $r=R{\cdot}\infty+Q$.
565: Since $a$ and $\adag$ commute with $\widehat{P}$,
566: the topological charge depends only on the finite-rank piece,
567: \begin{equation} \label{specialQE}
568: Q[P] \= Q[\widehat{P}{+}P'] \= Q[P'] \= \Tr\,P' \= r' \ .
569: \end{equation}
570: For $\bar P{=}0$ one obtains the abelian solutions because they fit
571: into~U($\Hcal$).
572: For $R{>}0$ we have nonabelian solutions because more than one copy of
573: $\Hcal$ is needed to accomodate them.
574:
575: Any projector in $\C^n\otimes\Hcal=\Hcal^{\oplus n}$ can be parametrized as
576: \begin{equation} \label{Tdef}
577: P \= |T\>\,\<T|T\>^{-1}\<T|
578: \end{equation}
579: where
580: \begin{equation} \label{Tarray}
581: |T\> \=
582: \Bigl(\, |T^1\> \quad |T^2\> \quad \dots \quad |T^r\> \Bigr) \=
583: \begin{pmatrix}
584: |T_1^1\> & |T_1^2\> & \dots & |T_1^r\> \\[4pt]
585: |T_2^1\> & |T_2^2\> & \dots & |T_2^r\> \\[4pt]
586: \vdots & \vdots & \ddots & \vdots \\[4pt]
587: |T_n^1\> & |T_n^2\> & \dots & |T_n^r\> \end{pmatrix} \=
588: \begin{pmatrix}
589: |T_1\> \\[4pt] |T_2\> \\[4pt] \vdots \\[4pt] |T_n\> \end{pmatrix}
590: \end{equation}
591: denotes an $n{\times}r$ array of kets in $\Hcal$,
592: with $r$ possibly being infinite. Thus,
593: \begin{equation}
594: \<T|T\> \= \Bigl( \<T^\ell|T^m\> \Bigr) \=
595: \Bigl( \textstyle{\sum_{i=1}^n} \<T_i^\ell|T_i^m\> \Bigr)
596: \end{equation}
597: stands for an invertible $r{\times}r$ matrix, and $r$ is the rank of~$P$.
598: The column vectors $|T^\ell\>$ span the image im$P$ of the projector in
599: $\Hcal^{\oplus n}$.
600: There is some ambiguity in the definition~(\ref{Tdef}) of~$|T\>$ since
601: \begin{equation}
602: |T\>\ \to\ |T\>\,\Gamma
603: \qquad\text{for}\quad \Gamma\in\text{GL}(r)
604: \end{equation}
605: amounts to a change of basis in im$P$ and does not change the projector~$P$.
606: This freedom may be used to normalize $\<T|T\>=\unity_r$, which is still
607: compatible with $\Gamma\in\text{U}(r)$.
608: On the other hand, a unitary transformation
609: \begin{equation}
610: |T\>\ \to\ U\,|T\> \qquad\text{for}\quad U\in\text{U}(\Hcal^{\oplus n})
611: \end{equation}
612: yields a unitarily equivalent projector $UP\,U^\+$.
613: Altogether, we have the bijection
614: \begin{equation} \label{Ttransform}
615: \tilde{P} \= U\,P\,U^\+ \qquad\Longleftrightarrow\qquad
616: |\tilde{T}\> \= U\,S\,|T\> \= U\,|T\>\,\Gamma
617: \qquad\text{with}\quad S \in \text{GL(im$P$)} \ .
618: \end{equation}
619: Here, the trivial action of $U\in\text{U(im$P$)}{\times}\text{U(ker$P$)}$
620: on $P$ can be subsumed in the $S$ action on~$|T\>$.
621:
622: Any diagonal projector can be cast into the form
623: \begin{equation} \label{Pdiag}
624: P_{\text{d}} \= \text{diag}\,\Bigl(
625: \underbrace{\unity_\Hcal,\dots,\unity_\Hcal}_{R\text{\ times}}\,,\,P_Q\,,
626: \underbrace{{\bf0}_\Hcal,\dots,{\bf0}_\Hcal}_{n{-}R{-}1\text{\ times}} \Bigr)
627: \qquad\text{with}\qquad P_Q \= \sum_{m=0}^{Q-1} |m\>\<m| \ .
628: \end{equation}
629: Formally, $P_{\text{d}}$ has rank~$r$ in $\Hcal^{\oplus n}$.
630: A corresponding $r{\times}n$ array of kets via (\ref{Tdef}) would be
631: \begin{equation} \label{specialket}
632: |T_{\text{d}}\> \=
633: \begin{pmatrix}
634: |\Hcal\> & \emptyset & \dots & \emptyset & 0_Q \\
635: \emptyset & |\Hcal\> & & \emptyset & 0_Q \\
636: \vdots & & \ddots & & \vdots \\
637: \emptyset & \emptyset & & |\Hcal\> & 0_Q \\
638: \emptyset & \emptyset & \dots & \emptyset & |T_Q\> \\
639: \vdots & \vdots & & \vdots & \\
640: \emptyset & \emptyset & \dots & \emptyset & 0_Q
641: \end{pmatrix}
642: \qquad\text{with}\quad\begin{cases} {}\\[-12pt]
643: \quad|\Hcal\> &\!\!=\ \bigl( |0\>\ |1\>\ |2\>\ |3\>\ \dots \bigr) \\[4pt] {}\
644: \quad\emptyset &\!\!=\ \bigl(\ 0\quad0\quad0\quad0\ \ \dots \bigr) \\[4pt]
645: \quad|T_Q\> &\!\!=\ \bigl( |0\>\ |1\>\ \dots\ |Q{-}1\> \bigr) \\[4pt] {}\
646: \quad0_Q &\!\!=\
647: \bigl(\ \underbrace{0\quad0\quad\dots\quad0}_{Q\text{\ times}}\ \bigr)
648: \end{cases} \quad.
649: \end{equation}
650: In the abelian case, $n{=}1$ and $R{=}0$,
651: this reduces to $|T_{\text{d}}\>=|T_Q\>$,
652: and any projector is unitarily equivalent to~$P_Q$.
653: Due to (\ref{coP}),
654: analogous results are valid in the complementary case $P\to\unity{-}P$.
655:
656: \subsection{BPS solutions}
657: \noindent
658: Let us focus on classical solutions within a Grassmannian.
659: For $\Phi=\unity{-}2P$, the equation of motion~(\ref{fulleom}) reduces to
660: \begin{equation} \label{Peom}
661: 0 \= [ \Delta P\,,\,P ] \=
662: [ \adag,\,(\unity{-}P)a P ] + [ a\,,\,P \adag(\unity{-}P) ] \=
663: [ a\,,\,(\unity{-}P)\adag P ] + [ \adag,\,P a(\unity{-}P) ] \ .
664: \end{equation}
665: Still, we do not know how to characterize its full solution space.
666: However, (\ref{Peom}) is identically satisfied by projectors subject to
667: \cite{Gopakumar:2001yw,Hadasz:2001cn}
668: \begin{align}
669: \text{either the BPS equation}&&
670: 0 &\= [a\,,P]\,P \= (\unity{-}P)\,a\,P &\label{BPS}\\
671: \text{or the anti-BPS equation}&&
672: 0 &\= [\adag,P]\,P \= (\unity{-}P)\,\adag P &\label{aBPS}
673: \end{align}
674: which are only ``first-order''.
675: Solutions to (\ref{BPS}) are called solitons while those to (\ref{aBPS}) are
676: named anti-solitons. Hermitian conjugation shows that the latter are obtained
677: {}from the former by exchanging $P\leftrightarrow\unity{-}P$, and so we can
678: ignore the anti-BPS solutions for most of the paper.
679: Equation (\ref{BPS}) means that
680: \begin{equation}
681: a \quad\text{maps}\quad \text{im}P\ \hookrightarrow\ \text{im}P
682: \end{equation}
683: and, hence, characterizes subspaces of $\Hcal^{\oplus n}$ which are
684: stable under the action of~$a$.
685: The term ``BPS equation'' derives from the observation that (\ref{BPS})
686: inserted in (\ref{EandQ}) implies the saturation of the BPS
687: bound~(\ref{bound}), which simplifies to
688: \begin{equation}
689: E[P] \= 8\pi\,Q[P] \=
690: 8\pi\,\Tr \bigl( P\,a\,(\unity{-}P)\,\adag P \bigr) \=
691: 8\pi\,\Tr \bigl( P\,a\,\adag P - a\,P\,\adag \bigr) \ .
692: \end{equation}
693: For BPS solutions of the particular form (\ref{specialP}) we indeed
694: recover that $E[P]=8\pi\,\Tr P'$.
695: Clearly, these BPS solutions constitute the absolute minima of the energy
696: functional within each Grassmannian.
697:
698: When the parametrization (\ref{Tdef}) is used, the BPS condition (\ref{BPS})
699: simplifies to
700: \begin{equation} \label{BPS2}
701: a\,|T_i^\ell\> \= |T_i^{\ell'}\>\,\gamma_{\ell'}^{\ \ell}
702: \qquad\text{for some $r{\times}r$ matrix}\quad
703: \gamma\=\bigl(\gamma_{\ell'}^{\ \ell}\bigr)
704: \end{equation}
705: which represents the action of~$a$ in the basis chosen for im$P$.
706: For instance, the ket $|T_Q\>$ in (\ref{specialket}) indeed obeys
707: \begin{equation}
708: a\,|T_Q\> \= |T_Q\>\,\gamma_Q \qquad\text{with}\qquad \gamma_Q \=
709: \left( \begin{smallmatrix}
710: 0 & \sqrt{1} & 0 & \dots & 0 \\
711: 0 & 0 & \sqrt{2} & & 0 \\
712: \vdots & \vdots & \ddots & \ddots & \\
713: 0 & 0 & \dots & 0 & \sqrt{Q{-}1} \\
714: 0 & 0 & \dots & 0 & 0
715: \end{smallmatrix} \right)
716: \quad,
717: \end{equation}
718: and the diagonal projector $P_{\text{d}}$ in (\ref{Pdiag}) is BPS,
719: but it is by far not the only one.
720:
721: Any basis change in im$P$ induces a similarity transformation
722: $\gamma\mapsto\Gamma\gamma\Gamma^{-1}$, which leaves $P$ unaltered and
723: thus has no effect on the value of the energy. Therefore, it suffices
724: to consider $\gamma$ to be of Jordan normal form. More generally,
725: a unitary transformation~(\ref{Ttransform}) is compatible with the
726: BPS condition (\ref{BPS2}) only if
727: \begin{equation}
728: U^\+ a\,U\,|T\> \= |T\>\,\gamma_U
729: \qquad\text{for some $r{\times}r$ matrix}\ \gamma_U \ ,
730: \end{equation}
731: which implies that
732: \begin{equation} \label{BPScomp}
733: \omega\,|T\> \= |T\>\,\gamma_\omega
734: \qquad\text{with}\qquad \gamma_\omega = \gamma_U-\gamma \ .
735: \end{equation}
736: The trivially compatible transformations are those in
737: $\text{U(im$P$)}{\times}\text{U(ker$P$)}$, which can be subsumed in
738: $S\in\text{GL(im$P$)}$ and lead to $\gamma_U{=}\Gamma^{-1}\gamma\Gamma$.
739: Another obvious choice are rigid symmetries of the energy functional,
740: as given in (\ref{globaltrans}), (\ref{globalrot}), and (\ref{globalsym})
741: for $W{=}V^\+$.
742: The challenging task then is to identify the nontrivial BPS-compatible unitary
743: transformations, since these relate different $a$-stable subspaces of fixed
744: dimension in $\Hcal^{\oplus n}$ and thus generate the multi-soliton moduli
745: space. For $|T_Q\>$ in (\ref{specialket}) we shall accomplish this
746: infinitesimally in Section~3.3.
747:
748: \subsubsection{Abelian solitons}
749: \noindent
750: Let us take a closer look at the BPS solutions of the noncommutative
751: U(1) sigma model. All finite-energy configurations are based on $R{=}0$
752: and have rank $r=Q<\infty$, thus $E=8\pi r$.
753: The task is to solve the ``eigenvalue equation''
754: \begin{equation} \label{BPSabelian}
755: a\,|T\> \= |T\>\,\gamma \qquad\text{for}\qquad \gamma \= \bigoplus_{s=1}^q
756: \left( \begin{smallmatrix}
757: \a_s & 1 & 0 & \dots & 0 \\
758: 0 & \a_s & 1 & & 0 \\
759: \vdots & & \ddots & \ddots & \\
760: 0 & 0 & & \a_s & 1 \\
761: 0 & 0 & \dots & 0 & \a_s
762: \end{smallmatrix} \right) \qquad\text{with}\quad \a_s\in\C \ ,
763: \end{equation}
764: where the Jordan cells have sizes~$r_s$ for $s=1,\dots,q$,
765: with $\sum_{s=1}^q r_s=r$.
766: For a given rank~$r$, the above matrices $\gamma$ parametrize the
767: $r$-soliton moduli space.
768: The general solution (unique up to cell-wise normalization and basis
769: changes $|T^{(s)}\>\to|T^{(s)}\>\Gamma^{(s)}$) reads
770: \begin{equation}
771: |T\> \= \Bigl( |T^{(1)}\>\ \dots\ |T^{(q)}\> \Bigr)
772: \qquad\text{with}\qquad |T^{(s)}\> \= \Bigl(
773: |\a_s\>\ \ a^\+|\a_s\>\ \dots\ \ \sfrac{1}{(r_s-1)!}(a^\+)^{r_s-1}|\a_s\>
774: \Bigr)
775: \end{equation}
776: and is based on the coherent states~(\ref{coherent}).
777: In the star-product picture, the corresponding $\Phi$ represents $r$ lumps
778: centered at positions $\a_s$ with degeneracies $r_s$ in the $xy$~plane.
779: Lifting a degeneracy by ``point-splitting'', the related Jordan cell dissolves
780: into different eigenvalues. Hence, the generic situation has
781: $r_s{=}1\ \forall s$, and so
782: \begin{equation} \label{Tcoherentr}
783: \gamma \= \text{diag} (\a_1,\a_2,\dots,\a_r) \qquad\Longleftrightarrow\qquad
784: |T\> \= \Bigl( |\a_1\>\ |\a_2\>\ \dots\ |\a_r\> \Bigr) \ ,
785: \end{equation}
786: which yields the projector
787: \begin{equation} \label{Pcoherentr}
788: P \= \sum_{k,\ell=1}^r
789: |\a_k\> \,\Bigl( \<\a_.|\a_.\> \Bigr)^{-1}_{k\ell} \<\a_\ell| \ .
790: \end{equation}
791:
792: Each solution can be translated via a unitary transformation mediated
793: by $D(\beta)$, which shifts $\a_\ell\mapsto\a_\ell{+}\beta\ \forall\ell$,
794: and rotated by $R(\vartheta)$, which moves
795: $\a_\ell\mapsto\e^{\ic\vartheta}\a_\ell\ \forall\ell$.
796: The individual values of~$\a_\ell$ (the soliton locations) may also be moved
797: around by appropriately chosen unitary transformations,
798: so that any $r$-soliton configuration can be reached from the diagonal one,
799: which describes $r$ solitons on top of each other at the coordinate origin
800: \cite{Hadasz:2001cn}:
801: \begin{equation} \label{diagonalBPS}
802: |T\> \= U\,|T_r\>\,\Gamma \qquad\Longleftrightarrow\qquad
803: P \= U\,P_r\,U^\+ \qquad\text{with}\quad
804: P_r\= \sum_{m=0}^{r-1} |m\>\<m| \ .
805: \end{equation}
806: We illustrate the latter point with the example of $r=2$. Generically,
807: \begin{equation}
808: \bigl( |\a_1\>\ |\a_2\> \bigr) \= U\,S\,\bigl( |0\>\ |1\> \bigr)
809: \= U\,\bigl( |0\>\ |1\> \bigr)\,\Gamma \qquad\text{with}\quad
810: U\,S\, \= |\a_1\>\<0| + |\a_2\>\<1| + \dots \ ,
811: \end{equation}
812: where the omitted terms annihilate $|0\>$ and $|1\>$. Factorizing $US$ yields
813: \begin{equation} \label{U2}
814: \begin{aligned}
815: U &\= \frac{1}{\sqrt{1{-}|\sigma|^2}}\,\Bigl(|\a_1\>\ |\a_2\>\Bigr)
816: \begin{pmatrix} \e^{-\ic\gamma} & 0 \\ 0 & \e^{-\ic\gamma'} \end{pmatrix}
817: \begin{pmatrix} -\sin\beta' & \phantom{-}\cos\beta' \\
818: \phantom{-}\sin\beta & -\cos\beta \end{pmatrix}
819: \begin{pmatrix} \<0| \\ \<1| \end{pmatrix} \ +\ \dots \\[4pt]
820: \text{and}\qquad \Gamma &\=
821: \begin{pmatrix} \cos\beta & \cos\beta' \\ \sin\beta & \sin\beta' \end{pmatrix}
822: \begin{pmatrix} \e^{\ic\gamma} & 0 \\ 0 & \e^{\ic\gamma'} \end{pmatrix}
823: \ ,\\[4pt] \text{with}\qquad \sigma &\=
824: \<\a_1|\a_2\> \= \e^{\bar{\a}_1\a_2-\frac12|\a_1|^2-\frac12|\a_2|^2}
825: \= \e^{-\ic(\gamma-\gamma')} \cos(\beta{-}\beta') \ ,
826: \end{aligned}
827: \end{equation}
828: so that $|\sigma|^2=\e^{-|\a_1-\a_2|^2}$, and $\beta{+}\beta'$ and
829: $\gamma{+}\gamma'$ remain undetermined. The rank-2 projector becomes
830: \begin{equation}
831: P \= \frac{1}{1{-}|\sigma|^2}\,\Bigl( |\a_1\>\<\a_1| + |\a_2\>\<\a_2|
832: - \sigma |\a_1\>\<\a_2| - \bar\sigma |\a_2\>\<\a_1| \Bigr) \=
833: U\, \Bigl( |0\>\<0| + |1\>\<1| \Bigr)\, U^\+ \ .
834: \end{equation}
835:
836: \subsubsection{Nonabelian solitons}
837: \noindent
838: We proceed to infinite-rank projectors. For simplicity, let us discuss
839: the case of U(2) solitons -- the results will easily generalize to U($n$).
840: Clearly, we can imbed two finite-rank BPS solutions (with $R{=}0$) into
841: U($\Hcal{\oplus}\Hcal$) by letting each act on a different copy of~$\Hcal$.
842: Such configurations are noncommutative deformations of the trivial projector
843: $\bar{P}=(\begin{smallmatrix} 0&0\\0&0 \end{smallmatrix})$
844: and thus represent a combination of abelian solitons. Therefore, we turn to
845: projectors with $R=1$ and so, formally, $r=\infty+Q$. Given (\ref{specialP})
846: such solutions may be considered as noncommutative deformations of the
847: U(2) projector $\bar{P}=(\begin{smallmatrix} 1&0\\0&0 \end{smallmatrix})$.
848: For this reason, one expects the generic solution $|T\>$ to (\ref{BPS2})
849: to be combined from a full set of states for the first copy of $\Hcal$
850: and a finite set of coherent states $|\a_\ell\>$ in the second copy of $\Hcal$,
851: \begin{equation}
852: |T\> \= \begin{pmatrix}
853: |\Hcal\> & 0 & 0 & \dots & 0 \\[4pt]
854: \emptyset & |\a_1\> & |\a_2\> & \dots & |\a_Q\> \end{pmatrix}
855: \qquad\Longrightarrow\qquad
856: P \= \unity_\Hcal\ \oplus\ \sum_{k,\ell=1}^Q
857: |\a_k\> \,\Bigl( \<\a_.|\a_.\> \Bigr)^{-1}_{k\ell} \<\a_\ell| \ ,
858: \end{equation}
859: with $E=8\pi Q$.
860: This projector is of the special form (\ref{specialP}), with
861: $\bar{P}=(\begin{smallmatrix}1&0\\0&0\end{smallmatrix})$ and $U=\unity$.
862: By a unitary transformation on the second copy of~$\Hcal$
863: such a configuration can be mapped to the diagonal form
864: \begin{equation} \label{simplespecialP}
865: |T_{\text{d}}\> \= \begin{pmatrix}
866: |\Hcal\> & 0 & 0 & \dots & 0 \\[4pt]
867: \emptyset & |0\> & |1\> & \dots & |Q{-}1\> \end{pmatrix}
868: \qquad\Longrightarrow\qquad
869: P_{\text{d}} \= \unity_\Hcal\ \oplus\ P_Q\ .
870: \end{equation}
871: It is convenient to reorder the basis of im$P_{\text{d}}$ such that
872: \begin{equation} \label{Thatdef}
873: |T_{\text{d}}\> \= \begin{pmatrix}
874: 0 & 0 & \dots & 0 & |\Hcal\> \\[4pt]
875: |0\> & |1\> & \dots & |Q{-}1\> & \emptyset \end{pmatrix}
876: \= \begin{pmatrix} S_Q \\[4pt] P_Q \end{pmatrix} |\Hcal\> \
877: =:\ \hat{T}_{\text{d}}\,|\Hcal\>\ ,
878: \end{equation}
879: where
880: \begin{equation}
881: P_Q \= \sum_{m=0}^{Q-1} |m\>\<m| \qquad\text{and}\qquad
882: S_Q \= (a\,\sfrac{1}{\sqrt{N}})^Q \= \sum_{m=Q}^\infty |m{-}Q\>\<m|
883: \end{equation}
884: denotes the $Q$th power of the shift operator.
885: The form of (\ref{Thatdef}) suggests to pass from states
886: $|T\>=\begin{pmatrix} |T_1\> \\ |T_2\> \end{pmatrix}$ with $R{=}1$
887: to operators $\hat{T}=\begin{pmatrix} \hat{T}_1 \\ \hat{T}_2 \end{pmatrix}$
888: on $\Hcal$:
889: \begin{equation}
890: |T\> \= \hat{T}\,|\Hcal\> \qquad\Longrightarrow\qquad
891: P \= \hat{T}\,(\hat{T}^\+ \hat{T})^{-1} \hat{T}^\+ \ .
892: \end{equation}
893: In fact, it is always possible to introduce $\hat{T}$ as
894: \begin{equation}
895: \hat{T}_i \= \sum_{\ell=1}^r |T_i^\ell\>\<\ell{-}1|
896: \qquad\Longleftrightarrow\qquad
897: |T_i^\ell\> \= \hat{T}_i\,|\ell{-}1\>
898: \qquad \textrm{for} \quad i=1,2 \quad\textrm{and}\quad \ell=1,\dots,r\ .
899: \end{equation}
900: We may even put $\hat{T}^\+\hat{T}=\unity_\Hcal$ by using the
901: freedom $\hat{T}\to\hat{T}\,\hat\Gamma$ with an operator
902: $\hat\Gamma=(\hat{T}^\+\hat{T})^{-1/2}$. Our example
903: of $|T_{\text{d}}\>$ in~(\ref{Thatdef}) is already normalized since
904: \begin{equation}
905: S_1\,S_1^\+\=\unity_\Hcal \qquad\text{but}\qquad
906: S_1^\+ S_1 \=\unity_\Hcal-|0\>\<0|
907: \qquad\Longrightarrow\qquad S^\+_Q S_Q + P_Q \= \unity_\Hcal \ .
908: \end{equation}
909:
910: It is instructive to turn on a BPS-compatible unitary transformation
911: in~(\ref{specialP}). In our example~(\ref{Thatdef}), we apply~\cite{Lee:2004dt}
912: \begin{equation} \label{Umu}
913: U(\mu) \= \begin{pmatrix}
914: S_Q \sqrt{\frac{N_Q}{N_Q{+}\mu\bar\mu}} \, S^\+_Q &
915: S_Q \frac{\bar\mu}{\sqrt{N_Q{+}\mu\bar\mu}} \\[12pt]
916: \frac{\mu}{\sqrt{N_Q{+}\mu\bar\mu}} \, S^\+_Q &
917: \frac{\mu\,P_Q\,-\,\sqrt{N_Q}}{\sqrt{N_Q{+}\mu\bar\mu}}
918: \end{pmatrix}
919: \qquad\text{with}\quad
920: N_Q \= {\adag}^Q a^Q \= N(N{-}1)\cdots(N{-}Q{+}1)
921: \end{equation}
922: to $\ \hat{T}_{\text{d}}=(S_Q,P_Q)^t\ $ of (\ref{Thatdef}). With the help of
923: \begin{equation}
924: S_Q P_Q \= 0 \= N_Q P_Q \qquad\text{and}\qquad
925: S_Q \sqrt{N_Q} \= a^Q
926: \end{equation}
927: we arrive at
928: \begin{equation} \label{Ttilde}
929: \hat{T}(\mu) \= U(\mu)\,\hat{T}_{\text{d}} \=
930: \begin{pmatrix} a^Q \\ \mu \end{pmatrix} \,
931: \frac{1}{\sqrt{N_Q{+}\mu\bar\mu}} \=
932: \begin{pmatrix} a^Q \\ \mu \end{pmatrix} \,\hat\Gamma
933: \ =:\ \check{T}(\mu)\,\hat\Gamma \ .
934: \end{equation}
935: This transformation can be regarded as a regularization
936: of $\hat{T}_{\text{d}}$ since
937: \begin{equation}
938: \lim_{\mu\to0} \hat{T}(\mu) \=
939: \bigl(\begin{smallmatrix} 1 & 0 \\ 0 & \e^{\ic\delta}\!\end{smallmatrix}\bigr)
940: \,\hat{T}_{\text{d}} \quad\text{with}\quad
941: \delta=\lim_{\mu\to0}\arg\mu \qquad\text{and thus}\qquad
942: \lim_{\mu\to0} P(\mu) \= P_{\text{d}} \ ,
943: \end{equation}
944: but note the singular normalization in the limit! For completeness we also
945: display the transformed projector,
946: \begin{equation}
947: P(\mu) \= U(\mu) \begin{pmatrix}
948: \unity_\Hcal & {\bf0}_\Hcal \\[8pt] {\bf0}_\Hcal & P_Q
949: \end{pmatrix} U^\+(\mu) \= \begin{pmatrix}
950: a^Q \frac{1}{N_Q{+}\mu\bar\mu} {\adag}^Q &
951: a^Q \frac{\bar\mu}{N_Q{+}\mu\bar\mu} \\[12pt]
952: \frac{\mu}{N_Q{+}\mu\bar\mu} {\adag}^Q &
953: \frac{\mu\bar\mu}{N_Q{+}\mu\bar\mu} \end{pmatrix} \ .
954: \end{equation}
955:
956: How do we see that such projectors are BPS?
957: Writing $|T\>=\hat{T}|\Hcal\>$, the BPS condition
958: \begin{equation} \label{TBPS}
959: a\,|T_i^\ell\> \= |T_i^{\ell'}\>\,\gamma_{\ell'}^{\ \ell}
960: \qquad\text{implies}\qquad
961: a\,\hat{T}_i \= \hat{T}_i \,\hat\gamma \qquad\textrm{with}\quad i=1,2
962: \end{equation}
963: for some operator $\hat\gamma$ in $\Hcal$.
964: We do not know the general solution for arbitrary~$\hat\gamma$.
965: However, an important class of solutions arises for the choice $\hat\gamma=a$
966: where the BPS condition reduces to the ``holomorphicity condition''\footnote{
967: This may even be the general case: If there exists an invertible operator
968: $\hat\Gamma$ solving $a\hat\Gamma=\hat\Gamma\hat\gamma$, then the general
969: solution to~(\ref{TBPS}) reads $\hat{T}_i=\check{T}_i\hat\Gamma$ with
970: $[a,\check{T}_i]=0$, and $\hat\Gamma$ can be scaled to unity.}
971: \begin{equation}
972: [\,a\,,\,\hat{T}_i\,] \= 0 \qquad\textrm{for}\quad i=1,2\ .
973: \end{equation}
974: These equations are satisfied by {\em any\/} set of functions
975: $\{\hat{T}_1,\hat{T}_2\}$ of $a$ alone, i.e.~not depending on~$\adag$.
976: Indeed, for the example of $\hat{T}_{\text{d}}$ in (\ref{Thatdef}),
977: we concretely have
978: \begin{equation}
979: \hat\gamma\= a\,P_Q \ +\ (\unity_\Hcal{-}P_Q)\,a\,\sqrt{\sfrac{N-Q}{N}} \ ,
980: \end{equation}
981: while $\check{T}(\mu)$ in (\ref{Ttilde}) is obviously holomorphic and thus BPS.
982: Because $\hat{T}(\mu)$ emerges from $\hat{T}_{\text{d}}$ via a BPS-compatible
983: unitary transformation it shares the topological charge~$Q$ and the
984: energy $E=8\pi Q$ with the latter.
985:
986: The generalization to arbitrary values of $n$ and $R<n$ is straightforward:
987: $\hat{T}$ becomes an $n{\times}R$ array of operators $(\hat{T}_i^L)$,
988: on which left multiplication by $a$ amounts to right multiplication by
989: an $R{\times}R$ array of operators~$\hat\gamma$. For the special choice
990: $\hat\gamma_{L'}^{\ L}=\delta_{L'}^{\ L}a$, any collection of holomorphic
991: functions of $a$ serves as a solution for~$\hat{T}_i^L$.
992: Quite generally, one can show that for polynomial functions $\hat{T}_i^L(a)$
993: the resulting projector has a finite topological charge~$Q$ given by the
994: degree of the highest polynomial and is thus of finite energy
995: \cite{Lee:2000ey,Lechtenfeld:2001aw,Foda:2002nt}.
996: In this formulation it becomes evident that nonabelian solutions have a
997: smooth commutative limit, where $\sqrt{2\theta}a\to{z}$ and $\theta\to0$.
998: Indeed, they are seen as deformations of the well known solitons in the
999: $\text{Gr}(n,R)=\frac{\text{U}(n)}{\text{U}(R){\times}\text{U}(n{-}R)}$
1000: Grassmannian sigma model.\footnote{
1001: For the above-discussed example one has $\text{Gr}(2,1)=\C P^1$.}
1002: Hence, the moduli space of the nonabelian solitons coincides with that
1003: of their commutative cousins. By rescaling $\hat{T}\to\hat{T}\Gamma$
1004: with a $\C$-valued $R{\times}R$ matrix~$\Gamma$ we can eliminate $R$
1005: complex parameters from $nR$ independent polynomials.
1006: For a charge-$Q$ solution, there remain $nRQ+(n{-}1)R$ complex moduli,
1007: of which $(n{-}1)R$ parametrize the vacuum and $nRQ$ describe the position
1008: and shape of the multi-soliton~\cite{Lee:2000ey}.
1009: For the case of $\C P^1$ this yields a complex $2Q$ dimensional soliton
1010: moduli space represented by
1011: $\bigl(\begin{smallmatrix} \check{T}_1 \\ \check{T}_2 \end{smallmatrix}\bigr)
1012: =\bigl(\begin{smallmatrix} \ a^Q+\ldots+\nu \\
1013: \lambda a^Q+\ldots+\mu \end{smallmatrix}\bigr)$.
1014: Our sample calculation above suggests that taking $\check{T}_2\to\mu$ and
1015: then performing the limit $\mu\to 0$ one recovers the complex $Q$ dimensional
1016: moduli space of the abelian solitons given by
1017: $|T\>=\bigl(|\a_1\>\,\dots\,|\a_{Q-1}\>\bigr)$ as a boundary.
1018:
1019: \subsection{Some non-BPS solutions}
1020: \noindent
1021: For the record we also present a particular class of non-BPS (and non-anti-BPS)
1022: Grassmannian solutions to the equation of motion $\ [\Delta P,P]=0$.
1023: {}From the action of $\Delta$ on a basis operator,
1024: \begin{equation}
1025: \Delta\,|m\>\<n| \= (m{+}n{+}1)\,|m\>\<n| -
1026: \sqrt{mn}\,|m{-}1\>\<n{-}1| - \sqrt{(m{+}1)(n{+}1)}\,|m{+}1\>\<n{+}1| \ ,
1027: \end{equation}
1028: we infer that $\Delta$ maps the $k$-th off-diagonal into itself,
1029: so in particular it retains the diagonal:
1030: \begin{equation} \label{Deltadiag}
1031: \Delta\,|m\>\<m| \= (2m{+}1)\,|m\>\<m| -
1032: m\,|m{-}1\>\<m{-}1| - (m{+}1)\,|m{+}1\>\<m{+}1| \ .
1033: \end{equation}
1034: It follows that {\em every\/} diagonal projector is a solution.
1035: Let us first consider the case of U(1).
1036: Given the natural ordering of the basis $\{|m\>\}$ of~$\Hcal$,
1037: any diagonal projector~$P'$ of finite rank~$r{=}Q$ can be written as
1038: \begin{equation} \label{nonBPSP}
1039: P' \= \sum_{s=1}^q \sum_{k=0}^{r_s-1} |m_s{+}k\>\<m_s{+}k|
1040: \= \sum_{s=1}^q S^\+_{m_s}\,P_{r_s}\;S_{m_s}
1041: \qquad\text{with}\quad m_{s+1}>m_s{+}r_s \quad\forall s
1042: \end{equation}
1043: and $\sum_{s=1}^q r_s =r$. Via
1044: \begin{equation} \label{DeltanonBPSP}
1045: \begin{aligned}
1046: \Delta P' &\= \sum_{s=1}^q \Bigl\{
1047: m_s\,|m_s\>\<m_s|\ -\ m_s\,|m_s{-}1\>\<m_s{-}1| \\ &\qquad -\
1048: (m_s{+}r_s)\,|m_s{+}r_s\>\<m_s{+}r_s|\ +\
1049: (m_s{+}r_s)\,|m_s{+}r_s{-}1\>\<m_s{+}r_s{-}1| \Bigr\}
1050: \end{aligned}
1051: \end{equation}
1052: its energy is easily calculated as
1053: \begin{equation} \label{nonBPSenergy}
1054: \sfrac{1}{8\pi}\,E[P'] \= Q\ +\ 2\sum_{s=1}^q m_s \ ,
1055: \end{equation}
1056: which is obviously minimized for the BPS case $q=1$ and $m_1=0$.
1057: The energy is additive as long as the two projectors to be combined
1058: are not getting ``too close''.
1059: This picture generalizes to the nonabelian case by formally allowing
1060: $r_s$ and $m_s$ to become infinite.
1061: In particular, the energy does not change when one imbeds $P'$
1062: (or several copies of it) into U($\Hcal^{\oplus n}$)
1063: and adds to it a constant projector as in (\ref{specialP}).
1064: Hence, with the proper redefinition of the $m_s$, (\ref{nonBPSenergy}) holds
1065: for nonabelian diagonal solutions as well.
1066:
1067: The inversion $P\to\unity{-}P$ generates additional solutions,
1068: which for the structure in (\ref{specialP}) and BPS-compatible unitaries~$U$
1069: are represented as
1070: \begin{equation}
1071: P \= U\,\bigl( \widehat{P}-P' \bigr) \, U^\+ \qquad\text{with}\qquad
1072: \widehat{P} = (\unity_n{-}\bar{P})\otimes\unity_\Hcal \qquad\text{and}\qquad
1073: (\unity{-}\widehat{P})P' =0= P'(\unity{-}\widehat{P}) \ ,
1074: \end{equation}
1075: When $\unity{-}\widehat{P}{+}P'$ is BPS, then $P$ becomes anti-BPS
1076: with topological charge $Q[P]=-\Tr P'$,
1077: producing an anti-soliton with energy $E[P]=8\pi\Tr P'$.
1078: It is possible to combine solitons and anti-solitons to a non-BPS solution via
1079: \begin{equation}
1080: P \= P^{\phantom{\+}}_{\text{sol}} + P_{\overline{\text{sol}}}
1081: \qquad\text{provided}\qquad
1082: P^{\phantom{\+}}_{\text{sol}} P_{\overline{\text{sol}}} \=0\=
1083: P_{\overline{\text{sol}}} P^{\phantom{\+}}_{\text{sol}}\ ,
1084: \end{equation}
1085: so that their topological charges and energies simply add to
1086: \begin{equation}
1087: Q[P] \= Q[P^{\phantom{\+}}_{\text{sol}}] + Q[P_{\overline{\text{sol}}}]
1088: \qquad\text{and}\qquad
1089: E[P] \= E[P^{\phantom{\+}}_{\text{sol}}] + E[P_{\overline{\text{sol}}}] \ .
1090: \end{equation}
1091: For the diagonal case, this is included in the solutions discussed above.
1092: Examples for U(1) and U(2) (which appeared in~\cite{Furuta:2002nv}) are
1093: (for $m{>}r$)
1094: \begin{equation}
1095: \begin{aligned}
1096: P &\= \unity_\Hcal - P_m + P_r \= \unity_\Hcal - \sum_{k=r}^{m-1} |k\>\<k|
1097: \qquad&\text{with}&\quad
1098: Q= r{-}m \ &\text{and}&\quad \sfrac{1}{8\pi}E= r{+}m \ , \\
1099: P &\= P_{r_1} \oplus (\unity_\Hcal{-}P_{r_2})
1100: \qquad&\text{with}&\quad
1101: Q= r_1{-}r_2 \ &\text{and}&\quad \sfrac{1}{8\pi}E= r_1{+}r_2 \ ,
1102: \end{aligned}
1103: \end{equation}
1104: respectively.
1105: Besides the global translations and rotations, other unitary transformations
1106: are conceivably compatible with the equation of motion $\ [\Delta P,P]=0$,
1107: generating moduli spaces of non-BPS Grassmannian solutions. Outside the
1108: Grassmannian manifolds, many more classical configurations are to be found.
1109:
1110:
1111: \section{Fluctuation analysis}
1112: \noindent
1113: In order to investigate the stability of the classical configurations
1114: constructed in the previous section, we must study the energy functional~$E$
1115: in the neighborhood of the solution under consideration. Since the latter
1116: is a minimum or a saddle point of~$E$, all the linear stability information
1117: is provided by the Hessian, i.e.~the second variation of~$E$ evaluated at
1118: the solution. The Hessian is viewed as a linear map on the solution's tangent
1119: space of fluctuations, and its spectrum encodes the invariant information:
1120: Zero modes belong to field directions of marginal stability
1121: (and extend to moduli if they remain zero to higher orders) while negative
1122: eigenvalues signal instabilities. A perturbation in such a field direction
1123: provides (part of) the initial conditions of a runaway solution in a time
1124: extension of the model. We cannot be very specific about the stability of
1125: a general classical configuration. Therefore, we shall restrict our attention
1126: to the stability of BPS solutions (as introduced above), at which the
1127: Hessian simplifies sufficiently to obtain concrete results.
1128: Since any BPS configuration is part of a moduli space which is imbedded
1129: in some Grassmannian which itself lies inside the full configuration space
1130: of the noncommutative U($n$) sigma model, the total fluctuation space contains
1131: the subspace of Grassmannian fluctuations which in turn includes the subspace
1132: of BPS perturbations, the latter being zero modes associated with moduli.
1133: In order to simplify the problem of diagonalizing the Hessian we shall search
1134: for decompositions of the fluctuation space into subspaces which are invariant
1135: under the action of the Hessian. As we consider configurations of finite
1136: energy only, admissible fluctuations~$\phi$ must render $\delta^2 E$ finite
1137: and keep the ``background'' $\Phi$ unitary. Furthermore, they need to be
1138: subject to the same conditions as $\Phi$ itself: $\phi$~is bounded, $[a,\phi]$
1139: and $[\adag,\phi]$ are Hilbert-Schmidt, and $\Delta\phi$ is traceclass.
1140: Finally, in keeping with our restricted notion of Grassmannian, we do not
1141: admit hermitian fluctuations\footnote{
1142: Perturbations inside the Grassmannian must be hermitian
1143: (see Section~3.2 below).}
1144: which alter the topological charge.
1145:
1146: \subsection{The Hessian}
1147: \noindent
1148: The Taylor expansion of the energy functional
1149: around some finite-energy configuration~$\Phi$ reads
1150: \begin{equation}
1151: \begin{aligned}
1152: E[\Phi{+}\phi] &\= E[\Phi]\,+
1153: \int\!\!\text{d}^2z\;\frac{\delta E}{\delta\Phi(z)}[\Phi]\;\phi(z)\,+\,
1154: \frac12 \int\!\!\text{d}^2z\!\int\!\!\text{d}^2z'\;\phi(z)\,
1155: \frac{\delta^2 E}{\delta\Phi(z)\,\delta\Phi(z')}[\Phi]\;\phi(z')\,+ \dots
1156: \\[4pt]
1157: & \ =:\ E[\Phi]\,+\, E^{(1)}[\Phi,\phi]\,+\, E^{(2)}[\Phi,\phi]\,+\,\dots\ ,
1158: \end{aligned}
1159: \end{equation}
1160: where the U($n$) traces are included in $\int\!\text{d}^2z$.
1161: The perturbation~$\phi$ is to be constrained as to keep the background
1162: $\Phi$ unitary. Since we compute to second order in~$\phi$,
1163: it does not suffice to take $\phi\in T_\Phi\text{U}(\Hcal^{\oplus n})$.
1164: Rather, we must include the leading correction stemming from the exponential
1165: map onto $\text{U}(\Hcal^{\oplus n})$, which generates the finite
1166: perturbation $\phi=\Phi'{-}\Phi$. The latter is subject to the constraint
1167: \begin{equation} \label{phidagger}
1168: (\Phi^\+{+}\phi^\+)(\Phi{+}\phi) \= \unity \qquad\Longrightarrow\qquad \phi^\+
1169: \= -\Phi^\+\phi\,\Phi^\+ + \Phi^\+\phi\,\Phi^\+\phi\,\Phi^\+ + O(\phi^3)\ ,
1170: \end{equation}
1171: which we shall use to eliminate $\phi^\+$ from the variations.
1172: It is important to realize that in this way the term linear in $\phi^\+$
1173: generates a contribution to $E^{(2)}[\Phi,\phi]$.
1174: Performing the expansion for the concrete expression (\ref{E}) and
1175: using (\ref{phidagger}) we arrive at\footnote{
1176: The same result is obtained by considering $E[\exp\{t\phi\}\Phi]$ up
1177: to $O(t^2)$.}
1178: \goodbreak
1179: \begin{align}
1180: E^{(1)}[\Phi,\phi] &\= \pi\,\Tr\bigl\{
1181: [\adag,\,\Phi^\+\phi\,\Phi^\+]\,[a\,,\,\Phi] +
1182: [\adag,\,\Phi]\,[a\,,\,\Phi^\+\phi\,\Phi^\+] -
1183: [\adag,\,\Phi^\+]\,[a\,,\,\phi] -
1184: [\adag,\,\phi]\,[a\,,\,\Phi^\+]
1185: \bigr\} \nonumber\\[4pt]
1186: &\= 2\pi\,\Tr\bigl\{
1187: (\Delta\Phi^\+\,\Phi - \Phi^\+\,\Delta\Phi)\,\Phi^\+\,\phi\bigr\} \= 0\ ,
1188: \\[6pt]
1189: E^{(2)}[\Phi,\phi] &\= \pi\,\Tr\bigl\{
1190: [\adag,\,\Phi^\+\phi\,\Phi^\+]\,[a\,,\,\phi] +
1191: [\adag,\,\phi]\,[a\,,\,\Phi^\+\phi\,\Phi^\+] \nonumber\\ &\qquad\ \ -\
1192: [\adag,\,\Phi^\+\phi\,\Phi^\+\phi\,\Phi^\+]\,[a\,,\,\Phi] -
1193: [\adag,\,\Phi]\,[a\,,\,\Phi^\+\phi\,\Phi^\+\phi\,\Phi^\+]
1194: \bigr\} \nonumber\\[4pt]
1195: &\= 2\pi\,\Tr\bigl\{
1196: \Phi^\+\phi\,\Phi^\+\phi\,\Phi^\+\,\Delta\Phi -
1197: \Phi^\+\phi\,\Phi^\+\,\Delta\phi \bigr\} \nonumber\\[4pt]
1198: &\= 2\pi\,\Tr\bigl\{
1199: \phi^\+\,\Delta\phi - \phi^\+\,(\Phi\Delta\Phi^\+)\;\phi \} + O(\phi^3) \ =:\
1200: 2\pi\,\Tr\bigl\{ \phi^\+\,H\,\phi \bigr\} + O(\phi^3) \ ,
1201: \label{Qform}
1202: \end{align}
1203: defining the Hessian $H=\Delta-(\Phi\Delta\Phi^\+)$ as a self-adjoint operator.
1204: Hence, our task is essentially reduced to working out the spectrum of the
1205: Hessian. Since $\Delta$ is clearly a positive semidefinite operator,
1206: an instability can only occur in directions for which $\<\Phi\Delta\Phi^\+\>$
1207: is sufficiently large.
1208:
1209: For later reference, we present the action of $H$ in the oscillator basis,
1210: \begin{equation} \label{Hosc}
1211: \begin{aligned}
1212: {}& H \ \sum_{m,\ell} \phi_{m,\ell}\,|m\>\<\ell| \=
1213: \sum_{m,\ell} (H\phi)_{m,\ell}\,|m\>\<\ell|
1214: \qquad\qquad\text{with} \\[4pt]
1215: {}& (H\phi)_{m,\ell} \= (m{+}\ell{+}1)\,\phi_{m,\ell} -
1216: \sqrt{(m{+}1)(\ell{+}1)}\,\phi_{m+1,\ell+1} -
1217: \sqrt{m\ell}\,\phi_{m-1,\ell-1} \\[4pt]
1218: {}& \qquad\qquad -\ \sum_{j,k} \Phi_{m,j}\,\bigl\{
1219: (j{+}k{+}1)\,\Phi_{j,k} -
1220: \sqrt{(j{+}1)(k{+}1)}\,\Phi_{j+1,k+1} -
1221: \sqrt{jk}\,\Phi_{j-1,k-1}
1222: \bigr\}\,\phi_{k,\ell} \ ,
1223: \end{aligned}
1224: \end{equation}
1225: where $\Phi_{m,\ell}$ as well as $\phi_{m,\ell}$ are still $n{\times}n$
1226: matrix-valued. At diagonal abelian backgrounds, as given in (\ref{nonBPSP}),
1227: the matrices reduce to numbers and the latter expression simplifies to
1228: \begin{equation} \label{Hdiag}
1229: \begin{aligned}
1230: (H\phi)_{m,\ell} &\= (m{+}\ell{+}1-2b_m)\,\phi_{m,\ell} -
1231: \sqrt{(m{+}1)(\ell{+}1)}\,\phi_{m+1,\ell+1} -
1232: \sqrt{m\ell}\,\phi_{m-1,\ell-1}
1233: \qquad\text{with} \\[4pt]
1234: b_m &\= \sum_{s=1}^q \bigl\{
1235: \delta_{m,m_s}m_s + \delta_{m,m_s-1}m_s +
1236: \delta_{m,m_s+r_s}(m_s{+}r_s) +
1237: \delta_{m,m_s+r_s-1}(m_s{+}r_s) \bigr\} \ ,
1238: \end{aligned}
1239: \end{equation}
1240: which for fixed $\ell$ differs from $(\Delta\phi)_{m,\ell}$ in at most
1241: $4q$ entries. For diagonal U(1) BPS backgrounds $\Phi=\Phi_r=\unity-2P_r$
1242: this reduces further to
1243: \begin{equation} \label{bdiag}
1244: b_m \= r\,\bigl( \delta_{m,r-1} + \delta_{m,r} \bigr) \ .
1245: \end{equation}
1246: It is important to note that these expressions do not yet yield a
1247: matrix representation of $H$ because the constraint~(\ref{phidagger})
1248: on the allowable perturbations must still be taken into account.
1249: We can do this by replacing $\phi$ with $\phi-\Phi\phi^\+\Phi$
1250: everywhere; then $\phi$ becomes unconstrained.
1251:
1252: \subsection{Decomposition into even and odd fluctuations}
1253: \noindent
1254: We specialize to Grassmannian backgrounds, $\Phi=\unity-2P=\Phi^\+$,
1255: characterized by a hermitian projector~$P$ and obeying $\Phi^2=\unity$.
1256: Any such projector induces an orthogonal decomposition
1257: \begin{equation}
1258: \C^n\otimes\Hcal\= P\,(\C^n\otimes\Hcal)\oplus(\unity{-}P)(\C^n\otimes\Hcal)
1259: \ =:\ \text{im}P \oplus \text{ker}P \ ,
1260: \end{equation}
1261: and a fluctuation~$\phi$ decomposes accordingly as
1262: \begin{equation}
1263: \phi \=
1264: \underbrace{P\,\phi\,P\ +\ (\unity{-}P)\,\phi\,(\unity{-}P)}_{\phi_{\text{e}}}
1265: \ +\
1266: \underbrace{P\,\phi\,(\unity{-}P)\ +\ (\unity{-}P)\,\phi\,P}_{\phi_{\text{o}}}
1267: \ ,
1268: \end{equation}
1269: where the subscripts refer to ``even'' and ``odd'', respectively.
1270: Since $\Phi$ acts as $-\unity$ on $\text{im}P$ but as $+\unity$ on
1271: $\text{ker}P$, we infer that
1272: \begin{equation} \label{hermfluct}
1273: \Phi\,\phi_{\text{e}} \= \phi_{\text{e}} \Phi \qquad\text{and}\qquad
1274: \Phi\,\phi_{\text{o}} \=-\phi_{\text{o}} \Phi \qquad\Longrightarrow\qquad
1275: \phi^\+_{\text{e}} \=-\phi_{\text{e}} \qquad\text{and}\qquad
1276: \phi^\+_{\text{o}} \= \phi_{\text{o}}
1277: \end{equation}
1278: to leading order from (\ref{phidagger}),
1279: i.e.~even fluctuations are anti-hermitian while odd ones are hermitian.
1280: This implies that odd fluctuations keep $\Phi$ inside its Grassmannian,
1281: but even ones perturb away from it.
1282: It also follows that in
1283: \begin{equation} \label{cross1}
1284: \Tr \bigl( \phi^\+\,\Delta\phi \bigr) \=
1285: \Tr \bigl( \phi^\+_{\text{e}}\,\Delta\phi_{\text{e}} \bigr) +
1286: \Tr \bigl( \phi^\+_{\text{o}}\,\Delta\phi_{\text{o}} \bigr) +
1287: \Tr \bigl( \phi^\+_{\text{e}}\,\Delta\phi_{\text{o}} \bigr) +
1288: \Tr \bigl( \phi^\+_{\text{o}}\,\Delta\phi_{\text{e}} \bigr)
1289: \end{equation}
1290: the last two terms cancel each other. Furthermore,
1291: the equation of motion (\ref{Peom}), $[\Delta P,P]=0$, implies that
1292: \begin{equation} \label{cross2}
1293: (\Phi\Delta\Phi)_{\text{o}} \= 0 \qquad\Longrightarrow\qquad
1294: \Tr \bigl( \phi^\+_{\text{e}}\;\Phi\Delta\Phi\;\phi_{\text{o}} \bigr) \= 0 \=
1295: \Tr \bigl( \phi^\+_{\text{o}}\;\Phi\Delta\Phi\;\phi_{\text{e}} \bigr) \ ,
1296: \end{equation}
1297: because only an even number of odd terms in a product survives under the trace.
1298: Combining (\ref{cross1}) and (\ref{cross2}) we conclude that
1299: \begin{equation} \label{eosplit}
1300: E^{(2)}[\Phi,\phi_{\text{e}}{+}\phi_{\text{o}}] \=
1301: E^{(2)}[\Phi,\phi_{\text{e}}] + E^{(2)}[\Phi,\phi_{\text{o}}] \ ,
1302: \end{equation}
1303: which allows us to treat these two types of fluctuations separately.
1304:
1305: The above decomposition has another perspective.
1306: Recall that any background configuration $\Phi$ being unitary
1307: can be diagonalized by some unitary transformation,
1308: \begin{equation} \label{diagPhi}
1309: \Phi \= U\,\Phi_{\text{d}}\,U^\+
1310: \= U\,\text{diag}\bigl(\{\e^{\ic\lambda_i}\}\bigr)\,U^\+\ .
1311: \end{equation}
1312: When $\Phi$ is hermitian, i.e.~inside some Grassmannian, the diagonal phase
1313: factors can be just $+1$ or $-1$, and $U$ is determined only up to a factor
1314: $V\in\text{U(im}P)\times\text{U(ker}P)$
1315: which keeps the two eigenspaces $\text{im}P$ and $\text{ker}P$ invariant.
1316: Adding a perturbation~$\phi$ lifts the high degeneracy of $\Phi$,
1317: so that the diagonalization of $\Phi{+}\phi$ requires an infinitesimal
1318: ``rotation'' $K$ of $\text{im}P$ and $\text{ker}P$ inside~$\Hcal$ as well as
1319: a ``large'' rediagonalization $V$ inside the two eigenspaces.
1320: Modulo higher order terms we may write
1321: \begin{equation}
1322: \begin{aligned}
1323: \Phi + \phi &\= U\,(1{+}K)\,V\,
1324: (\Phi_{\text{d}}+\phi_{\text{d}})\,V^\+(1{-}K)\,U^\+ \= \Phi\ +\ U \bigl\{
1325: V\,\phi_{\text{d}}\,V^\+\,+\,[K\,,\,\Phi_{\text{d}}] \bigr\} U^\+ \\[4pt]
1326: & \text{with}\qquad [V,\Phi_{\text{d}}]=0 \qquad\text{and}\qquad
1327: K = -K^\+ \quad\text{infinitesimal}\ ,
1328: \end{aligned}
1329: \end{equation}
1330: where $\phi_{\text{d}}$ is a purely diagonal and anti-hermitian fluctuation.
1331: Since $V$ depends on $\phi$ it should not be absorbed into~$U$.
1332: It rather generates all non-diagonal fluctuations inside
1333: $\text{im}P$ and $\text{ker}P$, allowing us to rewrite
1334: \begin{equation}
1335: V\,\phi_{\text{d}}\,V^\+ \= \phi'_{\text{d}}\ +\
1336: [ \Lambda_{\text{e}}\,,\,\phi_{\text{d}} ]\ ,
1337: \qquad\text{with}\qquad \Lambda_{\text{e}}^\+ = -\Lambda_{\text{e}}
1338: \end{equation}
1339: being a generator of $\text{U(im}P)\times\text{U(ker}P)$ and
1340: a modified diagonal perturbation~$\phi'_{\text{d}}$.
1341: Redenoting also $K=\epsilon\Lambda_{\text{o}}$ with a real and
1342: infinitesimal $\epsilon$ and a generator $\Lambda_{\text{o}}$ of the
1343: Grassmannian, the general fluctuation is parametrized as
1344: \begin{equation}
1345: \phi \= U\,\bigl\{ \phi'_{\text{d}}\ +\
1346: [ \Lambda_{\text{e}}\,,\,\phi_{\text{d}} ]\ +\
1347: [ \Lambda_{\text{o}}\,,\,\epsilon\Phi_{\text{d}} ] \bigr\}\, U^\+
1348: \end{equation}
1349: and decomposed (after diagonalizing the background via $U$) into
1350: a ``radial'' part $\phi'_{\text{d}}$ and an ``angular'' part
1351: $\phi_{\text{a}}=[\Lambda,\text{any}]$ with $\Lambda$ generating
1352: U($\Hcal^{\oplus n}$)~\cite{Gopakumar:2000zd}.
1353: For a Grassmannian background all terms have definite hermiticity properties,
1354: and we can identify
1355: \begin{equation}
1356: \phi_{\text{e}} \= U\,\bigl\{ \phi'_{\text{d}}\ +\
1357: [ \Lambda_{\text{e}}\,,\,\phi_{\text{d}} ] \bigr\}\, U^\+
1358: \qquad\text{and}\qquad
1359: \phi_{\text{o}} \=
1360: U\,\bigl\{ [ \Lambda_{\text{o}}\,,\,\epsilon\Phi_{\text{d}} ] \bigr\}\, U^\+
1361: \= \epsilon\,[ U \Lambda_{\text{o}} U^\+\,,\Phi ] \ .
1362: \end{equation}
1363: We have seen in (\ref{eosplit}) above that the even and odd fluctuations can
1364: be disentangled in~$E^{(2)}$. It is not clear, however, whether the diagonal
1365: perturbations can in turn be separated from the even angular ones in the
1366: fluctuation analysis (but see below for diagonal backgrounds where $U=\unity$).
1367:
1368: \subsection{Odd or Grassmannian perturbations}
1369: \noindent
1370: As far as stability of BPS configurations is concerned, the odd
1371: perturbations are easily dealt with by a general argument.
1372: Since a shift by $\phi_{\text{o}}$ keeps $\Phi$ inside its Grassmannian,
1373: wherein $\Phi$ already minimizes the energy, such a perturbation cannot lower
1374: the energy any further and we can be sure that negative modes are absent here.
1375: Therefore, solitons in the noncommutative {\em Grassmannian\/} sigma model
1376: are stable, up to possible zero modes.
1377: An obvious zero mode is generated by the translational and rotational symmetry.
1378: A glance at (\ref{globaltrans}) and (\ref{globalrot}) shows that
1379: the corresponding infinitesimal operators $\Lambda$ are given by
1380: \begin{equation}
1381: \Lambda_{\text{trans}} \= \a\,\adag - \ab\,a
1382: \qquad\text{and}\qquad
1383: \Lambda_{\text{rot}} \= \ic\vartheta\,\adag a\ ,
1384: \end{equation}
1385: which indeed leads to the annihilation of $[\Lambda_{\text{trans}},\Phi]$
1386: and $[\Lambda_{\text{rot}},\Phi]$ by~$H$ as we shall see.
1387:
1388: {}From the discussion at the end of Section~2.4
1389: we know that for $r{>}1$ there are additional zero modes inside the
1390: Grassmannian, because the multi-soliton moduli spaces are higher-dimensional.
1391: Parametrizing these BPS-compatible perturbation as follows,
1392: \begin{equation} \label{BPSpert}
1393: \Phi + \phi_{\text{o}} \= U^{\phantom{\+}}_{\text{B}}\,\Phi\, U^\+_{\text{B}}
1394: \= \Phi + \epsilon\,[\Lambda^{\phantom{\+}}_{\text{B}},\Phi] + O(\epsilon^2)
1395: \qquad\text{for}\quad
1396: U^{\phantom{\+}}_{\text{B}} = \e^{\epsilon\Lambda^{\phantom{\+}}_{\text{B}}}
1397: \quad\text{with}\quad
1398: \Lambda^\+_{\text{B}} = -\Lambda^{\phantom{\+}}_{\text{B}} \ ,
1399: \end{equation}
1400: the corresponding Lie-algebra element (\ref{defomega}) becomes
1401: \begin{equation}
1402: \omega \= U^\+_{\text{B}}\,[a\,,U^{\phantom{\+}}_{\text{B}}]
1403: \= (\e^{-\epsilon\,\text{ad}\Lambda^{\phantom{\+}}_{\text{B}}}-1)\,a
1404: \= \epsilon\,[a\,,\Lambda^{\phantom{\+}}_{\text{B}}] + O(\epsilon^2)\ ,
1405: \end{equation}
1406: and the condition~(\ref{BPScomp}) of BPS compatibility
1407: to leading order in $\epsilon$ reads
1408: \begin{equation} \label{BPSfluct}
1409: [a\,,\Lambda^{\phantom{\+}}_{\text{B}}]\,|T\> \= |T\>\,\gamma_\omega
1410: \qquad\text{for some $r{\times}r$ matrix $\gamma_\omega$}\ .
1411: \end{equation}
1412:
1413: Let us try to find $\Lambda^{\phantom{\+}}_{\text{B}}$ in the abelian case
1414: by perturbing around the diagonal BPS configuration
1415: \begin{equation}
1416: |T_r\> \= \bigl( |0\>\ |1\>\ \dots\ |r{-}1\> \bigr)
1417: \qquad\Longleftrightarrow\qquad
1418: \Phi_r \= \unity_\Hcal\ -\ 2\sum_{m=0}^{r-1} |m\>\<m| \ .
1419: \end{equation}
1420: Expanding the generator
1421: \begin{equation}
1422: \Lambda^{\phantom{\+}}_{\text{B}}\=\sum_{m,\ell}\Lambda_{m,\ell}\,|m\>\<\ell|
1423: \qquad\text{with}\quad \bar\Lambda_{m,\ell} = -\Lambda_{\ell,m}
1424: \end{equation}
1425: it is easy to see that the even part of $\Lambda^{\phantom{\+}}_{\text{B}}$
1426: automatically fulfils~(\ref{BPSfluct}), and so restrictions to
1427: $\Lambda_{m,\ell}$ arise only for the odd components.
1428: In fact, only the terms with $m{\ge}r$ and $\ell{<}r$ in the above sum
1429: may violate~(\ref{BPSfluct}), which leads to the conditions
1430: \begin{equation} \label{Leqs}
1431: \sqrt{m{+}1}\,\Lambda_{m+1,\ell} - \sqrt{\ell}\,\Lambda_{m,\ell-1} \= 0
1432: \qquad\text{for all}\quad m\ge r \quad\text{and}\quad \ell<r \ .
1433: \end{equation}
1434: Since $\ell{=}0$ yields $\Lambda_{m+1,0}{=}0$ as a boundary condition,
1435: this hierarchy of equations puts most components to zero, except for
1436: $m=r,\dots,r{+}\ell$ at any fixed $\ell{<}r$. These remaining $r(r{+}1)/2$
1437: components are subject to $r(r{-}1)/2$ equations from~(\ref{Leqs}),
1438: whose solution
1439: \begin{equation} \label{Lsol}
1440: \Lambda_{r+j,\ell+j} \= \sqrt{ \sfrac{
1441: (\ell{+}j)(\ell{+}j{-}1)\cdots(\ell{+}1)}{(r{+}j)(r{+}j{-}1)\cdots(r{+}1)}}\,
1442: \Lambda_{r,\ell}
1443: \qquad\text{for}\quad j=1,\dots,r{-}1{-}\ell
1444: \quad\text{at}\quad \ell\le r{-}2
1445: \end{equation}
1446: fixes $r(r{-}1)/2$ components in terms of the $r{-}1$ components
1447: appearing on the right hand side, which therefore are free complex parameters.
1448: The $r$th free parameter $\Lambda_{r,r-1}$ does not enter and is associated
1449: with the rigid translation mode.\footnote{
1450: The rigid rotation mode is absent because $\Phi_r$ is spherically symmetric.}
1451: Finally, the ensueing BPS perturbation~(\ref{BPSpert}) is found to be
1452: \begin{equation} \label{BPSmode}
1453: \phi_{\text{o}} \= \epsilon\,[\Lambda^{\phantom{\+}}_{\text{B}},\Phi_r] \=
1454: \epsilon \sum_{m=r}^{\infty} \sum_{\ell=0}^{r-1} \Bigl(
1455: \Lambda_{m,\ell}\,|m\>\<\ell|\ +\ \bar\Lambda_{m,\ell}\,|\ell\>\<m| \Bigr) \ ,
1456: \end{equation}
1457: with $\Lambda_{m,\ell}$ taken from (\ref{Lsol}).
1458: To higher orders in~$\epsilon$, the BPS-compatibility condition is not
1459: automatically satisfied by our solution~(\ref{Lsol}) but this can be repaired
1460: by adding suitable even components to~$\Lambda^{\phantom{\+}}_{\text{B}}$.
1461: Our result ties in nicely with the observation of $r$ complex moduli~$\a_k$ in
1462: (\ref{Pcoherentr}) whose shifts produce precisely $r$ complex zero modes.
1463: For the simplest non-trivial case of $r{=}2$, one can also extract these modes
1464: from differentiating (\ref{U2}) with respect to $\a_1$ or~$\a_2$.
1465:
1466: \subsection{Even or non-Grassmannian perturbations}
1467: \noindent
1468: The stability analysis of BPS configurations inside the full noncommutative
1469: U($n$) sigma model requires the investigation of the even fluctuations
1470: $\phi_{\text{e}}$ as well. Here, we have only partial results
1471: to offer.\footnote{
1472: Even for the commutative sigma model this is an open problem
1473: \cite{Zakrzewski}.}
1474: Yet, there is the following general argument which produces an unstable
1475: even fluctuation mode~$\phi_{\text{neg}}$ (but not an eigenmode)
1476: for {\em any\/} noncommutative multi-soliton $\Phi=\unity-2P$ with $Q>0$.
1477: For this, consider some other multi-soliton $\tilde\Phi=\unity-2\tilde P$
1478: which is contained in $\Phi$ in the sense that
1479: \begin{equation}
1480: \text{im}(\tilde P) \ \subset\ \text{im}(P)
1481: \qquad\Longleftrightarrow\qquad
1482: \tilde P\,P \= P\,\tilde P \= \tilde P \ .
1483: \end{equation}
1484: We then simply say that $\tilde P\subset P$.
1485: It follows that their difference $\Pi$ is the orthogonal complement
1486: of $\tilde P$ in im($P$),
1487: \begin{equation} \label{diffP}
1488: \Pi \= P - \tilde P \ \subset\ P \qquad\Longrightarrow\qquad
1489: \Pi^2 = \Pi \qquad\text{and}\qquad \Pi\,\tilde P =0= \tilde P\,\Pi \ .
1490: \end{equation}
1491: In particular, we may choose $\tilde P=0$.
1492: For any such pair ($P,\tilde P$) there exists a continuous path
1493: \begin{equation} \label{path}
1494: \begin{aligned}
1495: \Phi(s) \= \e^{\ic s\Pi} (\unity{-}2P) \=
1496: \unity-2P +(1{-}\e^{\ic s})\Pi\=\unity-2\tilde P -(1{+}\e^{\ic s})\Pi \\[4pt]
1497: \text{connecting} \qquad \Phi(0)=\Phi=\unity{-}2P \qquad
1498: \text{with} \qquad \Phi(\pi)=\tilde\Phi=\unity{-}2\tilde P \ .
1499: \end{aligned}
1500: \end{equation}
1501: Note that $\Phi(s)$ interpolates between two different Grassmannians,
1502: touching them only at $s{=}0$ and $s{=}\pi$.
1503: Since we assumed that $\Phi$ and $\tilde\Phi$ are BPS we know that
1504: \begin{equation}
1505: E[\Phi] \= 8\pi\,Q \qquad\text{and}\qquad E[\tilde\Phi] \= 8\pi\,\tilde Q
1506: \end{equation}
1507: with $Q$ and $\tilde Q$ being the topological charges of $P$ and $\tilde P$,
1508: respectively.\footnote{
1509: Note that $\tilde Q$ need not be smaller than $Q$ when $r(P)$ is infinite.}
1510: Inserting (\ref{path}) into the expression (\ref{E}) for the energy
1511: and abbreviating $1{-}\e^{\ic s}=\rho$ we compute
1512: \begin{align}
1513: E(s) \ :=\ E[\Phi(s)]
1514: &\= 8\pi\,Q
1515: -2\pi(\rho{+}\bar\rho)\,\Tr\bigl([a,P][\Pi,\adag]+[\adag,P][\Pi,a]\bigr)
1516: +2\pi\,\rho\bar\rho\,\Tr\bigl([a,\Pi][\Pi,\adag]\bigr) \nonumber \\[4pt]
1517: &\= 8\pi\,Q -4\pi(1{-}\cos s)\,
1518: \Tr\bigl([a,P][\Pi,\adag]+[\adag,P][\Pi,a]-[a,\Pi][\Pi,\adag]\bigr)\ ,
1519: \label{Escalc}
1520: \end{align}
1521: where the two traces could be combined due to the relation
1522: \begin{equation}
1523: \rho + \bar\rho \= \rho\bar\rho \= 2(1{-}\cos s) \ .
1524: \end{equation}
1525: Luckily, we do not need to evaluate the traces above.
1526: Knowing that $E(\pi)=8\pi\tilde Q$ we infer that the last trace in
1527: (\ref{Escalc}) must be equal to $Q{-}\tilde Q$
1528: and hence
1529: \begin{equation} \label{Epath}
1530: \begin{aligned}
1531: E(s) &\= 8\pi\,Q\,-\,4\pi(1{-}\cos s)(Q-\tilde Q)
1532: \=8\pi\bigl(Q\,\cos^2\sfrac{s}{2}\,+\,\tilde Q\,\sin^2\sfrac{s}{2}\bigr)\\[4pt]
1533: &\= 8\pi\,Q\,-\,2\pi(Q-\tilde Q)\,s^2\,+\,O(s^4)\ .
1534: \end{aligned}
1535: \end{equation}
1536: Evidently,
1537: for $Q>0$ we can always lower the energy of a given BPS configuration
1538: by applying an even perturbation $\phi_{\text{neg}}=-\ic\epsilon\Pi$ towards a
1539: BPS solution with smaller charge $\tilde Q<Q$.
1540: The higher the charge of $P$ the more such modes are present.
1541: Yet, they are not independent of one another but rather span a cone
1542: extending from the background, as we shall see in examples below. In fact,
1543: there is no reason to expect any of these unstable fluctuations to represent
1544: an eigenmode of the Hessian, and in general they do not. Nevertheless, their
1545: occurrence again demonstrates that there must be (at least) one negative
1546: eigenvalue of~$H$, and for diagonal U(1) BPS backgrounds we shall prove in
1547: Section~4.3 that there is exactly one.
1548: The only stable BPS solutions are therefore the ``vacua'' defined by
1549: $P'{=}0$ in (\ref{specialP}) and based on a constant U($n$)
1550: projector~$\bar{P}$. This leaves no stable solutions in the abelian case
1551: besides $\Phi=\unity_\Hcal$.
1552:
1553: In addition to the unstable mode, there exist also a number of non-Grassmannian
1554: zero modes around each BPS configuration, which generate nearby non-BPS
1555: solutions to the equation of motion. This will become explicit in the examples
1556: discussed below.
1557:
1558:
1559: \section{Perturbations of U(1) backgrounds}
1560: \subsection{Invariant subspaces}
1561: \noindent
1562: We specialize further to diagonal U(1) backgrounds $\Phi$ as given by
1563: (\ref{nonBPSP}) (not necessarily BPS).
1564: Using (\ref{Hosc}) it is easy to see that $H$ maps any off-diagonal
1565: into itself. Let us parametrize the $k$-th upper diagonal $\Dcal_k$ as
1566: \begin{equation} \label{Dkdef}
1567: \phi_{(+k)} \= \sum_{m=0}^\infty \mu_{(k)m}\,|m\>\<m{+}k|
1568: \qquad\text{with}\quad \mu_{(k)m} \in\C
1569: \qquad\text{for}\quad k=0,1,2,\dots\ .
1570: \end{equation}
1571: The hermiticity properties~(\ref{hermfluct}) of the perturbations demand that
1572: we combine $\Dcal_k$ and $\Dcal_k^\+$ into a subspace $\Ecal_k$ of the $k$-th
1573: upper plus lower diagonals by defining
1574: \begin{equation}
1575: \Ecal_k\ :=\ \bigl\{ \phi_{(k)}\,\bigm|\,
1576: \phi_{(k)} \= \phi_{(+k)} - \Phi\,\phi_{(+k)}^\+ \Phi \bigr\}\ ,
1577: \qquad\text{which implies}\qquad
1578: \phi_{(k)}^\+ \= - \Phi\,\phi_{(k)}\,\Phi \ .
1579: \end{equation}
1580: The direct sum of all $\Ecal_k$ is the full admissible tangent space to
1581: U($\Hcal$), and so we may decompose
1582: \begin{equation} \label{finesplit}
1583: \phi \= \sum_{k=0}^\infty \phi_{(k)} \=
1584: \sum_{k=0}^\infty \sum_{m=0}^\infty
1585: \Bigl\{ \mu_{(k)m}\,|m\>\<m{+}k|\ \mp\ \bar\mu_{(k)m}\,|m{+}k\>\<m| \Bigr\}\ ,
1586: \end{equation}
1587: with the sign depending on whether the component is even or odd.
1588: Since $H$ maps $\Ecal_k$ into itself, the bilinear form
1589: defined in (\ref{Qform}) is block-diagonal on the set of $\Ecal_k$,
1590: \begin{equation}
1591: 2\pi\,\Tr \bigl\{ \phi_{(k)}^\+\,H\,\phi_{(\ell)} \bigr\} \
1592: \sim\ \delta_{k\ell} \ ,
1593: \end{equation}
1594: which implies the factorization
1595: \begin{equation} \label{offdiag}
1596: E^{(2)}\,[\Phi,\textstyle{\sum_k}\phi_{(k)}] \=
1597: \sum_k E^{(2)}\,[\Phi,\phi_{(k)}] \ .
1598: \end{equation}
1599: In other words,
1600: $\Ecal_k$ forms an $H$-invariant subspace for each value of~$k$.
1601: In particular, $\Ecal_0$ is the space of admissible skew-hermitian diagonal
1602: matrices, and the (purely imaginary) diagonal fluctuations
1603: $\phi_{(0)}\equiv\phi_{\text{d}}\in\Ecal_0$ can be considered on their own.
1604:
1605: \subsection{Results for diagonal U(1) backgrounds}
1606: \noindent
1607: Apparently, for diagonal U(1) backgrounds it suffices to study the second
1608: variation form $E^{(2)}[\Phi,.]$ on each subspace $\Ecal_k$ ($k\ge0$)
1609: separately.
1610: Let us give more explicit formulae for the restriction of $E^{(2)}[\Phi,.]$
1611: to these subspaces. To this end, we denote the non-zero entries of $\Phi$
1612: by $\de_j:=\Phi_{jj}$ ($j=0,1,2,\dots$).
1613: We have $\de_j=\pm 1$ for each $j$, and the set
1614: \begin{equation}
1615: J\ :=\ \{j\ge0\;|\;\de_{j+1}\neq\de_j\}
1616: \end{equation}
1617: is finite. Then, a straightforward calculation shows that the restriction
1618: of $E^{(2)}[\Phi,.]$ to
1619: \begin{equation} \label{phi0}
1620: \begin{aligned}
1621: &\Ecal_0 \= \bigl\{ \phi_{(0)}\,\bigm|\,
1622: \phi_{(0)} \= \ic \sum_{m=0}^\infty \phi_m\,|m\>\<m|
1623: \quad\text{with}\quad \phi_m \in\R \bigr\}
1624: \qquad\text{is given by} \\
1625: &\sfrac{1}{2\pi} E^{(2)}[\Phi,\phi_{(0)}] \=
1626: \sum_{m=0}^\infty (m{+}1)\,(\phi_{m+1}-\phi_m)^2
1627: -2 \sum_{j\in J} (j{+}1)\,(\phi_{j+1}^2 + \phi_j^2) \ .
1628: \end{aligned}
1629: \end{equation}
1630:
1631: The formula for $E^{(2)}[\Phi,.]$ on $\Ecal_{k>0}$ is more complicated.
1632: Considering a fixed $k$-th upper diagonal $\Dcal_k$ as parametrized in
1633: (\ref{Dkdef}), we first evaluate (suppressing the subscript $(k)$)
1634: \begin{equation}
1635: \Tr \{ \phi^\+_{(+k)}\,\Delta\,\phi_{(+k)} \} \= \sum_{m=0}^\infty
1636: |\sqrt{m{+}1}\,\mu_{m+1}-\sqrt{m{+}k{+}1}\,\mu_m|^2 \=
1637: \sfrac12 k\,|\mu_0|^2 + \sfrac12 \sum_{m=0}^\infty R_m(\mu) \ ,
1638: \end{equation}
1639: with
1640: \begin{align} \label{Rmdef}
1641: R_m(\mu) &\ :=\
1642: |\sqrt{m{+}1}\,\mu_{m+1}-\sqrt{m{+}k{+}1}\,\mu_{m}|^2+
1643: |\sqrt{m{+}k{+}1}\,\mu_{m+1}-\sqrt{m{+}1}\,\mu_{m}|^2 \\[6pt] \nonumber
1644: &\= \bigl(\sqrt{m{+}k{+}1}-\sqrt{m{+}1}\bigr)^2
1645: \bigl(|\mu_{m+1}|^2+|\mu_m|^2\bigr)+
1646: 2\sqrt{m{+}k{+}1}\sqrt{m{+}1}\,|\mu_{m+1}-\mu_m|^2\ .
1647: \end{align}
1648: Armed with this expressions, we compute the second variation on~$\Ecal_k$.
1649: To state the outcome, it is useful to introduce the index sets
1650: \begin{equation}
1651: J-k\ :=\ \{j{-}k\in\NN_0\,|\, j\in J\} \qquad\text{and}\qquad
1652: A\ :=\ (J\cup(J-k))\setminus (J\cap(J-k)) \ ,
1653: \end{equation}
1654: i.e.~$A$ is the symmetric difference of $J$ and $J{-}k$.
1655: For $\phi_{(k)}\in\Ecal_k$ a direct calculation results in
1656: \begin{align} \label{QPhik}
1657: \sfrac{1}{2\pi} E^{(2)}[\Phi,\phi_{(k)}] &\=
1658: k\,|\mu_0|^2\ +\sum_{j\in\NN_0\setminus A}R_j(\mu)\ +\
1659: \sum_{j\in A}(2j{+}2{+}k)\bigl(|\mu_j|^2+|\mu_{j+1}|^2\bigr) \\[6pt] \nonumber
1660: &\quad -2\,\sum_{j\in J}\,(j{+}1)\bigl(|\mu_j|^2+|\mu_{j+1}|^2\bigr)\ -\
1661: 2\!\sum_{j\in J-k}\!(j{+}k{+}1)|\mu_j|^2\ -\
1662: 2\!\!\!\sum_{j\in J-k+1}\!\!\!(j{+}k)|\mu_j|^2 \ .
1663: \end{align}
1664:
1665: This expression simplifies when $k$ is greater than the largest element
1666: of~$J{+}1$. Then the sets $J{-}k$ and $J{-}k{+}1$ are empty,
1667: so $A=J$, and (\ref{QPhik}) takes the form
1668: \begin{equation} \label{QPhik2}
1669: \sfrac{1}{2\pi} E^{(2)}[\Phi,\phi_{(k)}] \=
1670: k\,|\mu_0|^2\ +\sum_{j\in\NN_0\setminus J}R_j(\mu)
1671: \ +\ k\,\sum_{j\in J}\bigl(|\mu_j|^2+|\mu_{j+1}|^2\bigr) \ .
1672: \end{equation}
1673: One sees that for each $j{\in}J$ the corresponding coefficient $\mu_j$
1674: decouples from all successive coefficients~$\mu_{m>j}$.
1675: Since the elements of~$J$ signify in the string of $\mu_m$ the boundaries
1676: between even and odd fluctuations, this observation confirms the
1677: decomposition~(\ref{eosplit}) of~$E^{(2)}$ into an even and odd part also after
1678: the restriction to~$\Ecal_k$.\footnote{
1679: For $|J|>1$ the even and/or odd part of $E^{(2)}$ is split further. In total,
1680: $E^{(2)}$ decomposes into at least $|J|{+}1$ blocks.}
1681: Furthermore, we note that $R_j(\mu)>0$ unless $\mu_j=\mu_{j+1}=0$,
1682: which makes it obvious from the expression~(\ref{QPhik2}) that the
1683: quadratic form~$E^{(2)}$ is strictly positive on each $\Ecal_k$ with
1684: $k>\max_{j\in J}(j{+}1)$.
1685: For the remaining~$\Ecal_k$ we have to work a little harder.
1686:
1687: \subsection{Results for diagonal U(1) BPS backgrounds}
1688: \noindent
1689: The idea is to pursue the {\it reduction of~$E^{(2)}$ to a sum of squares\/}
1690: whenever possible. For diagonal BPS solutions~(\ref{diagonalBPS})
1691: \begin{equation}
1692: \Phi_r \= \unity_\Hcal\ -\ 2\sum_{m=0}^{r-1} |m\>\<m|
1693: \= \sum_{m=0}^\infty \de_m\,|m\>\<m|
1694: \qquad\text{with}\quad \de_m \=
1695: \begin{cases} -1 & \text{for $m<r$} \\ +1 & \text{for $m\ge r$} \end{cases}\ ,
1696: \end{equation}
1697: this strategy turns out to be successful at all $k\ge1$
1698: but breaks down at $k=0$.
1699: We note that now $J=\{r{-}1\}$, which implies a distinction of cases:
1700: ``very off-diagonal'' perturbations have $k>r$,
1701: ``slightly off-diagonal'' perturbations occur for $1\le k\le r$,
1702: and diagonal perturbations mean $k=0$, to be discussed last.
1703: It will also be instructive to visualize the Hessian on the space $\Ecal_k$
1704: in matrix form,
1705: \begin{equation}
1706: E^{(2)}[\Phi_r,\phi_{(k)}]\=
1707: 2\pi\sum_{m,\ell=0}^\infty \bar\mu_m\,H^{(k)}_{m\ell}\,\mu_\ell\ ,
1708: \end{equation}
1709: defining an infinite-dimensional matrix $H^{(k)}=\bigl(H^{(k)}_{m\ell}\bigr)$.
1710:
1711: \subsubsection{Very off-diagonal perturbations}
1712: \noindent
1713: Since $J$ consists just of one element, at $k>r$ the matrix~$H^{(k)}$
1714: splits into two parts only, of which the odd one has finite size~$r$.
1715: More explicitly,
1716: \begin{equation}
1717: H^{(k)} \= H^{(k)}_{\text{Gr}(P)}\,\oplus\,H^{(k)}_{\text{ker}P} \ ,
1718: \end{equation}
1719: with
1720: \begin{align}
1721: \sfrac{1}{2} H^{(k)}_{\text{Gr}(P)} = \begin{pmatrix}
1722: k+1 & -\sqrt{1(k{+}1)} & & & & \\[8pt]
1723: -\sqrt{1(k{+}1)} & k+3 & -\sqrt{2(k{+}2)} & & & \\[8pt]
1724: & -\sqrt{2(k{+}2)} & k+5 & \qquad\ddots\!\!\!\!\!\!\!\! & & \\[8pt]
1725: & & \!\!\!\!\!\!\!\!\ddots & \;\ddots & \ddots & \\[8pt]
1726: & & & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\ddots
1727: & 2r+k-3 & -\sqrt{(r{-}1)(r{+}k{-}1)} \\[8pt]
1728: & & & & \!\!\!\!\!\!\!\!\!\!\!\!-\sqrt{(r{-}1)(r{+}k{-}1)}
1729: & 2r+k-1{\mbf{\,-\,r}}
1730: \end{pmatrix}
1731: \end{align}
1732: \begin{align} \label{QkerP}
1733: \sfrac{1}{2} H^{(k)}_{\text{ker}P} \= \begin{pmatrix}
1734: 2r+k+1{\mbf{\,-\,r}} & -\sqrt{(r{+}1)(r{+}k{+}1)} & & & \\[8pt]
1735: -\sqrt{(r{+}1)(r{+}k{+}1)} & 2r+k+3 & -\sqrt{(r{+}2)(r{+}k{+}2)} & & \\[8pt]
1736: & -\sqrt{(r{+}2)(r{+}k{+}2)} & 2r+k+5 & \qquad\ddots\!\!\!\!\!\!\!\! & \\[8pt]
1737: & & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\ddots & \ddots & & \\[8pt]
1738: \end{pmatrix} \ ,
1739: \end{align}
1740: where the boldface contributions disturb the systematics and originate
1741: from the $\Phi\Delta\Phi^\+$ term in the Hessian.
1742: The previous paragraph asserts strict positivity of~$H$ for $k>r$.
1743: Due to the finiteness of $H^{(k)}_{\text{Gr}(P)}$, the Hessian
1744: on $\Ecal_{k>r}$ features precisely $r$ positive eigenvalues\footnote{
1745: meaning that the corresponding modes are normalizable eigenvectors,
1746: i.e.~Hilbert-Schmidt}
1747: in its spectrum.
1748: We shall argue below that $H^{(k)}_{\text{ker}P}$ contributes a purely
1749: continuous spectrum~$\R_+$ for any~$k$. Thus, the above eigenvalues are
1750: not isolated but imbedded in the continuum.
1751:
1752: It is instructive to look at the edge of the continuum. The non-normalizable
1753: zero mode of $H^{(k)}_{\text{Gr}(P)}$ is explicitly given by
1754: \begin{equation} \label{edge}
1755: \mu_{r+m}^{\text{zero}} \= \mu_r\sqrt{\frac{
1756: (r{+}k{+}1)(r{+}k{+}2)\cdots(r{+}k{+}m)}{(r{+}1)(r{+}2)\cdots(r{+}m)}}
1757: \= \mu_r\sqrt{\frac{
1758: (r{+}m{+}1)(r{+}m{+}2)\cdots(r{+}m{+}k)}{(r{+}1)(r{+}2)\cdots(r{+}k)}} \ ,
1759: \end{equation}
1760: which grows like $m^{k/2}$ when $m=0,1,2,\dots$ gets large.
1761: Being unbounded for $k{>}0$ this infinite vector does not yield an admissible
1762: perturbation of~$\Phi_r$ however: $\delta^2 E$ is infinite.
1763:
1764: \subsubsection{Slightly off-diagonal perturbations}
1765: \noindent
1766: For each value of $k$ in the range $1\le k\le r$ we already established
1767: in Section~3.3 the existence of an odd complex normalizable zero mode
1768: connected with a moduli parameter. Nevertheless, as we will show now,
1769: $E^{(2)}[\Phi_r,\phi_{(k)}]$ remains positive for $1\le k\le r$ albeit
1770: not strictly so. In order to simplify (\ref{QPhik}) we make use of
1771: the following property for $R_j(\mu)$ as defined in~(\ref{Rmdef}):
1772: \begin{equation} \label{Rid}
1773: \begin{aligned}
1774: \sum_{j=m}^{l-1}R_j(\mu)
1775: &\= k\,|\mu_l|^2\ -\ k\,|\mu_m|^2\ +\ 2\sum_{j=m}^{l-1}
1776: |\sqrt{j{+}1}\,\mu_{j+1}-\sqrt{j{+}k{+}1}\,\mu_{j}|^2 \\
1777: &\= k\,|\mu_m|^2\ -\ k\,|\mu_l|^2\ +\ 2\sum_{j=m}^{l-1}
1778: |\sqrt{j{+}k{+}1}\,\mu_{j+1}-\sqrt{j{+}1}\,\mu_{j}|^2\ .
1779: \end{aligned}
1780: \end{equation}
1781: Here, by definition, all integers are non-negative and all sums are taken
1782: to vanish if $m\ge l$. The two equations above are easily proved by induction
1783: over $l$, starting from the trivial case $l=m$.
1784: We now employ the algebraic identities~(\ref{Rid}) to rewrite~(\ref{QPhik})
1785: for $1\le k\le r$ and obtain
1786: \begin{align} \label{QPhik3}
1787: \sfrac{1}{2\pi} E^{(2)}[\Phi_r,\phi_{(k)}] &\=
1788: k\,|\mu_r|^2\ +\ \sum_{j=r}^{\infty}R_j(\mu) \\
1789: &\quad +\ 2\sum_{j=0}^{r-k-2}
1790: |\sqrt{j{+}1}\,\mu_{j+1}-\sqrt{j{+}k{+}1}\,\mu_{j}|^2
1791: \ +\ 2\!\sum_{j=r-k}^{r-2}
1792: |\sqrt{j{+}k{+}1}\,\mu_{j+1}-\sqrt{j{+}1}\,\mu_{j}|^2\ , \nonumber
1793: \end{align}
1794: which is indeed positive semi-definite.
1795: We observe that the even part of~$E^{(2)}$ now consists of two disjoint pieces,
1796: containing $\{\mu_0,\dots,\mu_{r-k-1}\}$ and $\{\mu_{j\ge r}\}$,
1797: which are separated by the odd perturbations $\{\mu_{r-k},\dots,\mu_{r-1}\}$.
1798: If the right-hand side of~(\ref{QPhik3}) is equal
1799: to zero, then all components $\mu_{j\ge r}$ must vanish because each $R_j$
1800: is a strictly positive quadratic form of $\mu_j$ and $\mu_{j+1}$. In contrast,
1801: the remaining coefficients $\mu_{j<r}$ need not vanish at the zero set
1802: of $E^{(2)}[\Phi_r,\phi_{(k)}]$. In order to find the zero modes of
1803: the Hessian, it is convenient to visualize it again in matrix form, namely
1804: \begin{equation}
1805: H^{(k)} \= H^{(k)}_{\text{im}P}\,
1806: \oplus\,H^{(k)}_{\text{Gr}(P)}\,\oplus\,H^{(k)}_{\text{ker}P} \ ,
1807: \end{equation}
1808: where the blocks have sizes $r{-}k$, $k$, and $\infty$, respectively.
1809: {}From (\ref{QPhik3}) we extract
1810: \goodbreak
1811: \begin{align}
1812: \sfrac{1}{2} H^{(k)}_{\text{im}P} \= \begin{pmatrix}
1813: k+1 & -\sqrt{1(k{+}1)} & & & & \\[8pt]
1814: -\sqrt{1(k{+}1)} & k+3 & -\sqrt{2(k{+}2)} & & & \\[8pt]
1815: & -\sqrt{2(k{+}2)} & k+5 & \qquad\ddots\!\!\!\!\!\!\!\! & & \\[8pt]
1816: & & \!\!\!\!\!\!\!\!\ddots & \;\ddots & \ddots & \\[8pt]
1817: & & & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\ddots
1818: & 2r-k-3 & -\sqrt{(r{-}k{-}1)(r{-}1)} \\[8pt]
1819: & & & & \!\!\!\!\!\!\!\!\!\!\!\!-\sqrt{(r{-}k{-}1)(r{-}1)}
1820: & 2r-k-1{\mbf{\,-\,r}}
1821: \end{pmatrix}
1822: \end{align}
1823: %\vskip-5mm
1824: \begin{align}
1825: \sfrac{1}{2} H^{(k)}_{\text{Gr}(P)} \= \begin{pmatrix}
1826: 2r-k+1{\mbf{\,-\,r}} & -\sqrt{(r{-}k{+}1)(r{+}1)} & & & \\[8pt]
1827: -\sqrt{(r{-}k{+}1)(r{+}1)} & 2r-k+3 & \qquad\ddots\!\!\!\!\!\!\!\! & & \\[8pt]
1828: & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\ddots & \!\!\!\!\!\!\!\!\;\ddots
1829: & \qquad \ddots & & \\[8pt]
1830: & & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!
1831: \!\!\!\!\!\!\!\!\ddots & \!\!\!\!\!\!\!\!\!\!\!\! 2r+k-3
1832: & \!\!\!\!-\sqrt{(r{-}1)(r{+}k{-}1)} \\[8pt]
1833: & &
1834: & \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!-\sqrt{(r{-}1)(r{+}k{-}1)}
1835: & \!\!\!\ 2r+k-1{\mbf{\,-\,r}}
1836: \end{pmatrix}
1837: \end{align}
1838: and $H^{(k)}_{\text{ker}P}$ being identical to the matrix in~(\ref{QkerP}).
1839: The extreme cases are $\ H^{(1)}_{\text{Gr}(P)}=(0)\ $ and
1840: $\ H^{(r-1)}_{\text{im}P}=(0)$, while $H^{(r)}_{\text{im}P}$ is empty.
1841:
1842: We already know that $H^{(k)}_{\text{ker}P}$ is strictly positive definite.
1843: As claimed earlier, its spectrum is $\R_+$ and purely continuous,
1844: with a non-normalizable zero mode given by~(\ref{edge}).
1845: Being finite-dimensional, the other two blocks jointly yield again just $r$
1846: non-negative and non-isolated proper eigenvalues for fixed~$k{>}0$.
1847: Interestingly, two of these eigenvalues are now zero and give us one even and
1848: one odd complex normalizable zero mode. With $\beta_k,\gamma_k\in\C$ the latter
1849: take the following form:\footnote{
1850: The even zero mode is of course absent for $k=r$.}
1851: \goodbreak
1852: \begin{align} \label{evenzero}
1853: H^{(k)}_{\text{im}P}:\qquad
1854: (\mu_\ell)_{\text{zero}} &\= \gamma_k\,\begin{pmatrix}
1855: \qquad\ \;\sqrt{1}\sqrt{2}\sqrt{3} \cdots
1856: \sqrt{r{-}k{-}2}\sqrt{r{-}k{-}1} \\[4pt]
1857: \quad\ \sqrt{k{+}1}\sqrt{2}\sqrt{3} \cdots
1858: \sqrt{r{-}k{-}2}\sqrt{r{-}k{-}1} \\[4pt]
1859: \sqrt{k{+}1}\sqrt{k{+}2}\sqrt{3} \cdots
1860: \sqrt{r{-}k{-}2}\sqrt{r{-}k{-}1} \\[-2pt]
1861: \vdots \\[2pt]
1862: \sqrt{k{+}1}\sqrt{k{+}2}\sqrt{k{+}3} \cdots
1863: \sqrt{r{-}2}\sqrt{r{-}k{-}1} \\[4pt]
1864: \sqrt{k{+}1}\sqrt{k{+}2}\sqrt{k{+}3} \cdots
1865: \sqrt{r{-}2}\sqrt{r{-}1} \quad\
1866: \end{pmatrix} \ , \\[10pt] \label{oddzero}
1867: H^{(k)}_{\text{Gr}(P)}:\qquad
1868: (\mu_\ell)_{\text{zero}} &\= \beta_k\,\begin{pmatrix}
1869: \qquad\; \sqrt{r{+}1}\sqrt{r{+}2}\sqrt{r{+}3} \cdots
1870: \sqrt{r{+}k{-}2}\sqrt{r{+}k{-}1}\\[4pt]
1871: \quad\ \sqrt{r{-}k{+}1}\sqrt{r{+}2}\sqrt{r{+}3} \cdots
1872: \sqrt{r{+}k{-}2}\sqrt{r{+}k{-}1}\\[4pt]
1873: \sqrt{r{-}k{+}1}\sqrt{r{-}k{+}2}\sqrt{r{+}3} \cdots
1874: \sqrt{r{+}k{-}2}\sqrt{r{+}k{-}1}\\[-2pt]
1875: \vdots \\[2pt]
1876: \sqrt{r{-}k{+}1}\sqrt{r{-}k{+}2}\sqrt{r{-}k{+}3} \cdots
1877: \sqrt{r{-}2}\sqrt{r{+}k{-}1}\\[4pt]
1878: \sqrt{r{-}k{+}1}\sqrt{r{-}k{+}2}\sqrt{r{-}k{+}3} \cdots
1879: \sqrt{r{-}2}\sqrt{r{-}1} \quad\
1880: \end{pmatrix} \ .
1881: \end{align}
1882: Altogether, there are $2r{-}2$ real zero modes
1883: in $H_{\text{im}P}$ and $2r$ real zero modes in $H_{\text{Gr}(P)}$.
1884: Identifying $\mu_{(k)j}=\Lambda_{j,k+j}=-\bar\Lambda_{k+j,j}$,
1885: the latter precisely agree with the BPS moduli found in~(\ref{Lsol}).
1886: The former zero modes correspond to moduli outside the Grassmannian, thus
1887: generating nearby non-diagonal non-BPS solutions to the equation of motion.
1888:
1889: \subsubsection{Diagonal perturbations}
1890: \noindent
1891: We come to the diagonal (or radial) perturbations $\phi_{(0)}\in\Ecal_0$
1892: (see~(\ref{phi0})). The transformation of
1893: \begin{equation}
1894: \sfrac{1}{2\pi} E^{(2)}[\Phi_r,\phi_{(0)}] \=
1895: \sum_{m=0}^\infty (m{+}1)\,(\phi_{m+1}-\phi_m)^2
1896: -2r\,( \phi_r^2 + \phi_{r-1}^2 )
1897: \end{equation}
1898: to a sum of squares is much easier than that of~(\ref{QPhik}),
1899: but the result shows one minus in the signature:
1900: \begin{equation} \label{QPhi0}
1901: \sfrac{1}{2\pi} E^{(2)}[\Phi_r,\phi_{(0)}] \= -r\,(\phi_{r-1}+\phi_{r})^2\
1902: +\!\sum_{j\ge0,\,j\neq r-1}\!\!(j{+}1)\,(\phi_{j+1}-\phi_j)^2\ .
1903: \end{equation}
1904: However, we can use this expression to conclude that the second variation
1905: form $E^{(2)}$ cannot have more than one negative mode on $\Ecal_0$.
1906:
1907: {\bf Lemma}.
1908: {\it Let $V\subset\Ecal_0$ be a real vector subspace such that
1909: $E^{(2)}[\Phi,\phi]<0$ for all nonzero $\phi\in V$. Then $V$ is at
1910: most one-dimensional. In other words, one cannot find linearly
1911: independent vectors $\phi,\psi\in\Ecal_0$ such that $E^{(2)}[\Phi,.]$ takes
1912: negative values on each non-zero linear combination of $\phi$ and~$\psi$.
1913:
1914: Proof}. A necessary and sufficient condition for a pair of linearly
1915: independent vectors $\phi,\psi\in\Ecal_0$ to span a negative subspace
1916: for $E^{(2)}(.):=E^{(2)}[\Phi,.]$ is that
1917: \begin{equation}
1918: E^{(2)}(\phi)<0\ ,\quad E^{(2)}(\psi)<0 \quad\text{and}\quad
1919: B(\phi,\psi)^2-E^{(2)}(\phi)E^{(2)}(\psi)<0 \ ,
1920: \end{equation}
1921: where
1922: $B(\phi,\psi):=\frac12\{E^{(2)}(\phi{+}\psi)-E^{(2)}(\phi)-E^{(2)}(\psi)\}$ is
1923: the corresponding bilinear form. If $\phi,\psi\in\Ecal_0$
1924: satisfy these inequalities, then the same holds for their truncations
1925: (that is, vectors obtained by replacing all the coordinates
1926: $\phi_j$, $\psi_j$ with $j\ge n$ by zero) if $n$ is large enough.
1927: Thus, we may assume the existence of $n$ such
1928: that $\phi_j=\psi_j=0$ for $j\ge n$. But the restriction of $E^{(2)}$
1929: to the space $\R^n$ of all vectors
1930: $\phi=(\phi_0,\phi_1,\dots,\phi_{n-1},0,0,\dots)$
1931: has signature $(1,n{-}1)$ according to~(\ref{QPhi0}) and, therefore,
1932: this space contains no two-dimensional negative subspaces.\qed
1933:
1934: For an alternative visualization, we write the second-order fluctuation form as
1935: \begin{equation}
1936: E^{(2)}[\Phi_r,\phi_{(0)}]\=
1937: 2\pi\sum_{m,\ell=0}^\infty \phi_m\,H^{(0)}_{m\ell}\,\phi_\ell\ ,
1938: \end{equation}
1939: where the matrix entries $H^{(0)}_{m\ell}$ can be extracted from (\ref{Hdiag})
1940: or from (\ref{QPhi0}) as
1941: \begin{equation}
1942: \sfrac{1}{2\pi}\,H^{(0)}_{m\ell}
1943: \= (2m{+}1-2r\delta_{m,r-1}-2r\delta_{m,r})\,\delta_{m,\ell}
1944: - m\,\delta_{m,\ell+1} - (m{+}1)\,\delta_{m+1,\ell} \ .
1945: \end{equation}
1946: In matrix form this reads as ($m,\ell\in\NN_0$)
1947: \begin{equation} \label{Qzero}
1948: \begin{aligned}
1949: (H^{(0)}_{m\ell}) &\= \begin{pmatrix}
1950: \phantom{-}1 & -1 \\[2pt]
1951: -1 & \phantom{-}3 & -2 \\[2pt]
1952: & -2 & \phantom{-}5 & -3 \\[-2pt]
1953: & & -3 & \phantom{-}7 & \ddots \\[-2pt]
1954: & & & \ddots & \ddots
1955: \end{pmatrix}
1956: \ -\ 2\,r\,(\delta_{m,r-1}\delta_{\ell,r-1}\,+\,\delta_{m,r}\delta_{\ell,r})
1957: \\[8pt]
1958: &\= \begin{pmatrix}
1959: \ddots & \ddots & & & & \\[-2pt]
1960: \ddots & 2r{-}3 & -r{+}1 & & & \\[2pt]
1961: & -r{+}1 & \mbf{-1} & -r & & \\[2pt]
1962: & & -r & \mbf{+1} & -r{-}1 & \\[-2pt]
1963: & & & -r{-}1 & 2r{+}3 & \ddots \\[-2pt]
1964: & & & & \ddots & \ddots
1965: \end{pmatrix} \ .
1966: \end{aligned}
1967: \end{equation}
1968: In distinction to the earlier cases, this infinite Jacobian (i.e.~tri-diagonal)
1969: matrix does not split. We shall compute its spectrum in a moment.
1970: Prior to this, some observations can already be made.
1971:
1972: First of all, the unique zero mode of $H^{(0)}$ is found from (\ref{edge})
1973: for $k{=}0$ and given by ($\gamma_0\in\R$)
1974: \begin{equation}
1975: \phi_{\text{zero}}\=\ic\gamma_0\,\Phi_r
1976: \qquad\Longleftrightarrow\qquad (\phi_m)_{\text{zero}} \= \gamma_0\,
1977: (\underbrace{-1,\ldots,-1}_{r\:\text{times}},+1,+1,\ldots) \ ,
1978: \end{equation}
1979: which clearly generates the global phase rotation symmetry
1980: $\Phi\to\e^{\ic\gamma_0}\Phi$ already noted in~(\ref{globalsym}).
1981: In contrast to the zero modes in~$\Ecal_{k>0}$ depicted in (\ref{edge}),
1982: (\ref{evenzero}) and~(\ref{oddzero}), $\phi_{\text{zero}}$ is not
1983: Hilbert-Schmidt but bounded. Although it does not belong to a proper zero
1984: eigenvalue of the Hessian, $\phi_{\text{zero}}$, being proportional
1985: to~$\Phi_r$, still meets all our requirements for an admissible perturbation.
1986:
1987: In order to identify unstable modes we turn on all diagonal perturbations
1988: in a particular manner suggested by~(\ref{path}),
1989: \begin{equation}
1990: \Phi(\{\alpha_m\}) \= \unity\ -\
1991: \sum_{m=0}^{r-1} (\e^{\ic\alpha_m}{+}1)\,|m\>\<m| \ +\
1992: \sum_{m=r}^\infty (\e^{\ic\alpha_m}{-}1)\,|m\>\<m| \ .
1993: \end{equation}
1994: The energy of this configuration is easily calculated to be
1995: \begin{equation}
1996: \sfrac{1}{8\pi}\,E(\{\alpha_m\}) \=
1997: \sin^2\sfrac{\alpha_0-\alpha_1}{2} +\dots
1998: +(r{-}1)\sin^2\sfrac{\alpha_{r-2}-\alpha_{r-1}}{2}
1999: +r\cos^2\sfrac{\alpha_{r-1}-\alpha_r}{2}
2000: +(r{+}1)\sin^2\sfrac{\alpha_r-\alpha_{r+1}}{2} +\dots
2001: \end{equation}
2002: with a single $\cos^2$ appearing only in the indicated place.
2003: Expanding up to second order in the real parameters~$\alpha_m$ and
2004: putting $\alpha_m=0$ for $m\ge r$, we find that
2005: \begin{equation}
2006: H^{(0)}<0 \qquad\Longleftrightarrow\qquad
2007: (\alpha_0{-}\alpha_1)^2 + 2(\alpha_1{-}\alpha_2)^2 + \dots +
2008: (r{-}1)(\alpha_{r-2}{-}\alpha_{r-1})^2 - r\,\alpha_{r-1}^2 \ \le\ 0 \ ,
2009: \end{equation}
2010: which determines a convex cone in the $r$-dimensional restricted fluctuation
2011: space. The most strongly negative mode is given by ($\alpha\in\R$)
2012: \begin{equation}
2013: \phi_{\text{neg}}\=-\ic\alpha P_r
2014: \qquad\Longleftrightarrow\qquad (\phi_m)_{\text{neg}} \= -\alpha\,
2015: (\underbrace{1,\ldots,1}_{r\:\text{times}},0,0,\ldots) \ .
2016: \end{equation}
2017: Comparison with (\ref{Epath}) reveals that following this mode one arrives at
2018: $\Phi=\unity$. More generally,
2019: \begin{equation}
2020: (\phi_m) \= -\alpha\,
2021: (\underbrace{0,\ldots,0}_{\tilde r\:\text{times}},
2022: \underbrace{1,\ldots,1}_{(r-\tilde r)\:\text{times}},0,0,\ldots)
2023: \end{equation}
2024: perturbs in the direction of $\Phi=\unity-2P_{\tilde r}$.
2025: None of these modes are eigenvectors of the matrix~$H^{(0)}$.
2026:
2027: It is instructive to confirm these assertions numerically.
2028: To this end, we truncate the matrix $(H^{(0)}_{m\ell})$ at some cut-off value
2029: $m_{\text{max}}=\ell_{\text{max}}$ and diagonalize the resulting
2030: finite-dimensional matrix. Truncating at values $m_{\text{max}}<r$ only
2031: allows for positive eigenvalues, whereas for values $m_{\text{max}}\geq r$
2032: we obtain exactly {\em one\/} negative eigenvalue.
2033: Its numerical value depends roughly linearly on $r$ and converges very quickly
2034: with increasing cut-off as can be seen from the following graphs.
2035: \vskip-5mm
2036: \begin{figure}[H]
2037: \psfrag{$k_max$}{$\scriptstyle{m_{\text{max}}}$}
2038: \psfrag{$lambda$}{$\scriptstyle{\lambda}$}
2039: \begin{center}
2040: \includegraphics[origin=ct,width=80mm]{ev1}
2041: \hspace{\fill}
2042: \includegraphics[origin=ct,width=80mm]{ev2}
2043: \end{center}
2044: \caption{Possible eigenvalues $\lambda$ versus the cut-off parameter
2045: $m_{\text{max}}$ for $r=5$ and $r=20$}
2046: \end{figure}
2047:
2048: \subsubsection{Spectrum of the Hessian}
2049: \noindent
2050: The spectrum of the restriction $H^{(k)}=\Delta_k-(\Phi_r\Delta\Phi_r)_k$
2051: of the Hessian $H$ to $\Ecal_k$ can actually be calculated by
2052: passing to a basis of Laguerre polynomials. The following considerations
2053: are based on spectral theory and require all vectors to be in $l_2$,
2054: i.e.~all operators to be Hilbert-Schmidt.
2055:
2056: Let us first consider~$\Ecal_0$.
2057: We use the following property of the noncommutative
2058: Laplacian $-\Delta_0$: sending each basis vector $|m\>$ of $\Hcal$ to the
2059: $m$-th Laguerre polynomial $L_m(x)$, we get a unitary map
2060: $U:\Hcal\to L^2(\R_+,\e^{-x}\diff x)$ such that $U\Delta_0U^{-1}$ is just
2061: the operator of multiplication by~$x$:
2062: \begin{equation} \label{xL}
2063: x\,L_m(x) \= -m\,L_{m-1}(x) + (2m{+}1)L_m(x) - (m{+}1)L_{m+1}(x) \ ,
2064: \end{equation}
2065: which is to be compared with (\ref{Qzero}).
2066: Hence, in the basis of the Laguerre polynomials, the eigenvalue equation
2067: $H^{(0)}|\phi\>=\lambda|\phi\>$ is rewritten in terms of $f:=U|\phi\>$ as
2068: \begin{equation} \label{ev}
2069: x\,f(x)-2r\bigl(\<f,L_{r-1}\>L_{r-1}(x)+\<f,L_r\>L_r(x)\bigr)\=\lambda\,f(x)\ .
2070: \end{equation}
2071: Clearly, a function $f\in L^2(\R_+,\e^{-x}\diff x)$ satisfies (\ref{ev})
2072: if and only if it is given by
2073: \begin{equation} \label{fsol}
2074: f(x)\=\frac{c_0L_{r-1}(x)+c_1L_r(x)}{x-\lambda}
2075: \end{equation}
2076: and the constant coefficients $c_0,c_1\in\R$ satisfy the linear system
2077: \begin{equation} \label{linsys}
2078: \begin{cases}
2079: \left(I_{r-1,r-1}(\lambda)-\frac1{2r}\right)c_0+I_{r-1,r}(\lambda)\,c_1\=0
2080: \\[6pt]
2081: I_{r,r-1}(\lambda)\,c_0+\left(I_{r,r}(\lambda)-\frac{1}{2r}\right)c_1\=0
2082: \end{cases} \ ,
2083: \end{equation}
2084: which is obtained simply by inserting (\ref{fsol}) into (\ref{ev}).
2085: Here we use the notation
2086: \begin{equation} \label{ints}
2087: I_{k,l}(\lambda)\ :=\
2088: \int_0^{\infty} \frac{\e^{-x}\,\diff x}{x-\lambda}\; L_k(x)\,L_l(x)
2089: \qquad\text{for}\quad k,l\ge0\ .
2090: \end{equation}
2091:
2092: The integrals (\ref{ints}) are simple (though not elementary)
2093: special functions of $\lambda$. Indeed,
2094: $I_{00}(\lambda)$ is a version of the integral logarithm:
2095: $I_{00}(\lambda)=-\e^{-\lambda}\operatorname{li}(\e^{\lambda})$
2096: for all $\lambda<0$. On the other hand, using the recursion relations
2097: \begin{equation}
2098: (k{+}1)\,L_{k+1}(x)\=(2k{+}1{-}x)\,L_k(x)\ -\ k\,L_{k-1}(x)\ ,
2099: \end{equation}
2100: one can show by induction over $k$ and $l$ that all functions
2101: $I_{k,l}(\lambda)$ are expressed in terms of $I_{00}(\lambda)$:
2102: \begin{equation}
2103: I_{kl}(\lambda)\=A_{kl}(\lambda)\,I_{00}(\lambda)\ +\ B_{kl}(\lambda)\ ,
2104: \end{equation}
2105: where $A_{kl}$ and $B_{kl}$ are polynomials in $\lambda$ of degree at
2106: most~$k{+}l$. Hence, the determinant
2107: \begin{equation}
2108: F_r(\lambda)\ :=\ \left\|\begin{matrix}
2109: I_{r-1,r-1}(\lambda)-\frac{1}{2r} & I_{r-1,r}(\lambda) \\[6pt]
2110: I_{r,r-1}(\lambda) & I_{r,r}(\lambda)-\frac{1}{2r} \end{matrix}\right\|
2111: \end{equation}
2112: of the linear system (\ref{linsys}) is a known special function
2113: of $\lambda$, whose zeros $\lambda_r$ on the negative semiaxis are precisely
2114: the negative eigenvalues of the Hessian operator $H^{(0)}[\Phi_r]$ for the
2115: diagonal BPS background of rank~$r$.
2116:
2117: One can now prove the existence of negative eigenvalues.
2118: By verifying that the real numbers
2119: $F_r(-\infty):=\lim_{\lambda\to-\infty}F_r(\lambda)$ and
2120: $F_r(0):=\lim_{\lambda\to0-} F_r(\lambda)$ have different signs,
2121: we can conclude that $F_r(\lambda)$ possesses at least one zero on the
2122: negative semiaxis.
2123: For example, take $r=1$. In this case we find
2124: \begin{equation}
2125: \left.\begin{matrix}
2126: I_{01}(\lambda)\= I_{10}(\lambda) & \!=\ (1{-}\lambda)I_{00}(\lambda)-1
2127: \hfill \\[6pt] \hfill
2128: I_{11}(\lambda) & \!=\ (1{-}\lambda)^2 I_{00}(\lambda)+\lambda-1
2129: \end{matrix}
2130: \right\}\quad\Longrightarrow\qquad
2131: 2\,F_1(\lambda)\=-\lambda-\sfrac12-\lambda^2 I_{00}(\lambda) \ .
2132: \end{equation}
2133: Passing to the limit under the integral sign, we see that
2134: $F_1(-\infty)=\frac14$ and $F_1(0)=-\frac14$. This proves the existence
2135: of a negative eigenvalue of $H^{(0)}[\Phi_1]$.
2136:
2137: Can $H^{(0)}[\Phi_r]$ also have non-negative eigenvalues? To disprove this
2138: possibility, we consider the eigenvalue equation~(\ref{ev}) and show that
2139: it admits only the trivial solution $f(x)\equiv 0$ if $\lambda\ge 0$.
2140: Indeed, for non-negative~$\lambda$ the solution~(\ref{fsol}) is not
2141: square-integrable unless
2142: \begin{equation}
2143: c_0L_{r-1}(\lambda)+c_1L_{r}(\lambda)\=0 \qquad\Longrightarrow\qquad
2144: c_0\=c\,L_{r}(\lambda) \qquad\text{and}\qquad c_1\=-c\,L_{r-1}(\lambda)
2145: \end{equation}
2146: for some constant~$c$.
2147: Therefore, the solution~(\ref{fsol}) is completely determined as
2148: \begin{equation} \label{fsol2}
2149: f(x)\=c\,\frac{L_r(\lambda)L_{r-1}(x)-L_{r-1}(\lambda)L_r(x)}{x-\lambda}
2150: \=c\,\sum_{k=0}^{r-1} L_k(\lambda)\,L_k(x)
2151: \end{equation}
2152: via the Christoffel-Darboux formula.
2153: It follows that $f$ is orthogonal to $L_{r}$. Inserting $\<f,L_r\>=0$
2154: into~(\ref{ev}), we learn that the polynomial $(x{-}\lambda)f(x)$ is
2155: proportional to $L_{r-1}(x)$. But then the first equality in~(\ref{fsol2})
2156: demands that $L_{r-1}(\lambda)=0$, constraining~$\lambda$. Feeding this
2157: into the second expression for~$f$ in~(\ref{fsol2}) we see that the sum
2158: runs to $r{-}2$ only, implying that $f$ is orthogonal to $L_{r-1}$ as well.
2159: This finally simplifies the eigenvalue equation~(\ref{ev}) to the one for
2160: $\Delta_0$, namely
2161: \begin{equation}
2162: x\,f(x) \= \lambda\,f(x) \ ,\qquad\text{whence}\qquad
2163: f(x)\ \equiv\ 0 \ ,
2164: \end{equation}
2165: as required.
2166: Hence, the non-negative part of the spectrum of~$H^{(0)}$ is purely continuous.
2167:
2168: A similar analysis can be applied to the Hessian on~$\Ecal_{k>0}$.
2169: In this case, the appropriate polynomials are the normalized (generalized)
2170: Laguerre polynomials~$L^k_m$, which form an orthonormal basis for
2171: $L^2(\R_+,x^k\e^{-x}\diff{x})$. One rediscovers the $r$ proper eigenvalues
2172: as zeroes of the characteristic polynomial for
2173: $H^{(k)}_{\text{im}P}\oplus H^{(k)}_{\text{Gr}(P)}$,
2174: accompanied by a continuous spectrum covering the positive semiaxis for
2175: $H^{(k)}_{\text{ker}P}$, as claimed earlier.
2176:
2177: If we do not care for resolving non-isolated eigenvalues, we may also infer
2178: the spectrum of the Hessian~$H$ from the classical Weyl theorem
2179: (see~\cite{Kato}, Theorem IV.5.35 and \cite{RS}, \S~XIII.4, Example~3).
2180:
2181: {\bf Assertion}.
2182: {\it Let $\Phi=\Phi_r$ be the diagonal\/} BPS {\it solution of rank~$r>0$. Then
2183: \begin{equation} \label{assert}
2184: \operatorname{spectrum}\,(H[\Phi_r])\=\{\lambda_r\}\cup\,[0,+\infty)\ ,
2185: \end{equation}
2186: where $\lambda_r$ is a negative eigenvalue of multiplicity~$1$ and
2187: $[0,+\infty)$ is the essential spectrum.\footnote{
2188: We recall that the essential spectrum of a self-adjoint operator is the whole
2189: spectrum minus all isolated eigenvalues of finite multiplicity.
2190: The multiplicity of an eigenvalue is the dimension of the corresponding
2191: eigenspace.}
2192: Furthermore, we have $-2r<\lambda_r<0$.
2193:
2194: Proof}. Weyl's theorem asserts that the essential spectrum is preserved
2195: under compact perturbations. The essential spectrum of $\Delta$ (on all
2196: subspaces~$\Ecal_k$) is known to be equal to $[0,+\infty)$. (This follows,
2197: for example, from the explicit diagonalization of $\Delta_k$ in terms of
2198: the Laguerre polynomials, as indicated above.)
2199: Since the perturbation $(\Phi_r\Delta\Phi_r)$ is finite-dimensional,
2200: Weyl's theorem applies to show that the essential spectrum of~$H$ is also
2201: equal to $[0,+\infty)$. Although we have explicitly shown that $H^{(k)}$ is
2202: positive semi-definite for~$k{>}0$, the essential spectrum~$\R_+$ cannot
2203: exhaust the whole spectrum because the operator~$H^{(0)}$ is not positive
2204: semi-definite on $\Ecal_0$. (Indeed, (\ref{QPhi0}) shows that the quadratic
2205: form~$E^{(2)}$ is negative on many finite vectors.) Hence, $H^{(0)}$ must
2206: have negative eigenvalues of finite multiplicity.
2207: The above lemma implies that there can be only one such eigenvalue, and
2208: its multiplicity must be equal to~1. This proves (\ref{assert}) and the
2209: inequality $\lambda_r<0$. The estimate $-2r<\lambda_r$ follows because the
2210: norm of the operator $(\Phi_r\Delta\Phi_r)$ is equal to $2r$, so
2211: $H[\Phi_r]+2r\unity_\Hcal\ge\Delta$ is non-negative definite and hence
2212: cannot have negative eigenvalues.\qed
2213:
2214: \subsection{Results for non-diagonal U(1) BPS backgrounds}
2215: \noindent
2216: Even though any non-diagonal BPS background~$\Phi$ can be reached from a
2217: diagonal one by a unitary transformation~$U$ which does not change
2218: the value of the energy, the fluctuation problem will get modified
2219: under such a transformation because
2220: \begin{equation}
2221: U\,a\,U^\+ \= f(a,\adag) \qquad\text{and}\qquad
2222: U\,\adag U^\+ \= f^\+(a,\adag)
2223: \end{equation}
2224: will in general not have an action as simple as $a$ and $\adag$ in the
2225: original oscillator basis.
2226: The exceptions are the symmetry transformations, of which
2227: only the rigid translations $D(\a)$ and rotations $R(\vartheta)$
2228: keep $\Phi$ within the Grassmannian.
2229: To be specific, we recall that
2230: \begin{equation}
2231: \begin{aligned}
2232: D(\a)^\+\,a\,D(\a) \= a + \a
2233: &\qquad\Longrightarrow\qquad&
2234: E\,[D(\a)\,\Phi\,D(\a)^\+] \= E\,[\Phi] \ ,\\[4pt]
2235: R(\vartheta)^\+\,a\,R(\vartheta) \= \e^{\ic\vartheta} a
2236: &\qquad\Longrightarrow\qquad&
2237: E\,[R(\vartheta)\,\Phi\,R(\vartheta)^\+] \= E\,[\Phi] \ .
2238: \end{aligned}
2239: \end{equation}
2240: The invariance of the spectrum of~$H$ under translations (or rotations)
2241: can also be seen directly:
2242: \begin{equation}
2243: \begin{aligned}
2244: H\,\phi_n &\= \epsilon_n\,\phi_n \qquad\text{and}\qquad
2245: \Delta\,(D\,f\,D^\+) \= D\,(\Delta f)\,D^\+ \quad\Longrightarrow \\[4pt]
2246: H'\phi_n' &\ \equiv\
2247: \bigl[ \Delta-(D\Phi D^\+)\Delta(D\Phi D^\+) \bigr] D\phi_n D^\+ \=
2248: D\,\bigl([\Delta-\Phi\Delta\Phi]\phi_n \bigr)D^\+ \=\epsilon_n\,\phi_n' \ .
2249: \end{aligned}
2250: \end{equation}
2251: More general unitary transformations will not simply commute with the action
2252: of $\Delta$ and thus change the spectrum of~$H$.
2253:
2254: In the rank-one BPS situation, any solution is a translation of~$\Phi_1$,
2255: and hence our previous discussion of fluctuations around diagonal U(1) BPS
2256: backgrounds completely covers that case. This is no longer true for higher-rank
2257: BPS backgrounds. For a complete stability analysis of abelian $r$-solitons
2258: it is therefore necessary to investigate separately the spectrum of~$H$
2259: around each soliton configuration (\ref{Pcoherentr}) based on $r$~coherent
2260: states, whose center of mass may be chosen to be the origin. Since the
2261: background holomorphically depends on $r$ parameters~$\a_1,\dots,\a_r$ and
2262: passes to $\Phi_r$ when all parameters go to zero, one may hope to show
2263: that the spectrum of~$H$ does not change qualitatively when $\Phi$ varies
2264: inside the rank-$r$ moduli space.
2265:
2266:
2267: \section{Perturbations of U(2) backgrounds}
2268: \noindent
2269: In this section we investigate the behavior of a simple U(2) BPS
2270: solution under perturbations and conclude that the fluctuation analysis
2271: reduces to the U(1) case discussed in the previous section.
2272:
2273: \subsection{Results for diagonal U(2) BPS backgrounds}
2274: \noindent
2275: We consider a nonabelian BPS projector of the diagonal type considered
2276: in \eqref{simplespecialP}:
2277: \begin{equation}
2278: P \= \unity_\Hcal\,\oplus\,P_Q \qquad\text{where}\qquad
2279: P_Q \= \sum_{k=0}^{Q-1} |k\>\<k|\ .
2280: \end{equation}
2281: Clearly this describes a BPS solution
2282: \begin{equation}
2283: \Phi \= \begin{pmatrix} \Phi^{(1,1)} & \Phi^{(1,2)} \\[4pt]
2284: \Phi^{(2,1)} & \Phi^{(2,2)} \end{pmatrix}
2285: \= \begin{pmatrix} - \unity_\Hcal & 0\\[4pt]
2286: 0 & \unity_\Hcal-2P_Q \end{pmatrix}
2287: \end{equation}
2288: of energy $E=8 \pi Q$.
2289: Inserting this into the expression (\ref{Qform}) for the quadratic
2290: energy correction~$E^{(2)}$ one obtains
2291: %\goodbreak
2292: \begin{equation}
2293: \begin{aligned}
2294: E^{(2)}[\Phi,\phi] &\= 2\pi\,\Tr\bigl\{
2295: \phi^\+\,\Delta\phi\ -\ \phi^\+\,(\Phi\Delta\Phi^\+)\;\phi \} \\[4pt]
2296: &\= 2\pi\,\Tr\bigl\{
2297: \phi^\+\,\bigl(\begin{smallmatrix}1&0\\0&1\end{smallmatrix}\bigr)\,
2298: \Delta\phi\ -\
2299: \phi^\+\,\bigl(\begin{smallmatrix}0&0\\0&1\end{smallmatrix}\bigr)\,
2300: 2Q \bigl( |Q{-}1\>\<Q{-}1| + |Q\>\<Q| \bigr)\,\phi \bigr\} \ .
2301: \end{aligned}
2302: \end{equation}
2303: Since the action of the Hessian is obviously diagonal, we find that
2304: \begin{equation}
2305: \phi \= \begin{pmatrix} \phi^{(1,1)} & \phi^{(1,2)} \\
2306: \phi^{(2,1)} & \phi^{(2,2)} \end{pmatrix}
2307: \qquad\Longrightarrow\qquad
2308: E^{(2)}[\Phi,\phi] \=
2309: \sum_{i,j=1}^2 E^{(2)}\bigl[\Phi^{(i,i)},\phi^{(i,j)}\bigr] \ ,
2310: \end{equation}
2311: which reduces the fluctuation analysis to a collection of abelian cases.
2312: For the case at hand, the Hessian in the $(1,.)$ sectors is given by
2313: the Laplacian and thus non-negative, while in each $(2,.)$ sector
2314: it is identical to the Hessian for the abelian rank-$Q$ diagonal BPS case.
2315: Hence, the relevant results of the previous section carry over completely.
2316:
2317: This mechanism extends to any diagonal U(2) background (not necessarily BPS),
2318: mapping the U(2) fluctuation spectrum to a collection of fluctuation spectra
2319: around corresponding diagonal U(1) configurations.
2320: The fluctuation problem around non-diagonal backgrounds, in contrast, is not
2321: easily reduced to an abelian one, except when the background is related to
2322: a diagonal one by a rigid symmetry, as defined in (\ref{globalsym}),
2323: (\ref{globaltrans}) and (\ref{globalrot}). Only a small part of the moduli,
2324: however, is generated by rigid symmetries, as our example of $U(\mu)$ in
2325: (\ref{Umu}) demonstrates even for $Q{=}1$.
2326: Finally, it is straightforward to extend these considerations to the
2327: general U($n$) case.
2328:
2329:
2330: \section{Conclusions}
2331: \noindent
2332: After developing a unified description of abelian and nonabelian
2333: multi-solitons in noncommutative Euclidean two-dimensional sigma models with
2334: U($n$) or Grassmannian target space, we have analyzed the issue of their
2335: stability. Thanks to the BPS bound, multi-solitons in Grassmannian sigma models
2336: are always stable. As in the commutative case, their imbedding into a unitary
2337: sigma model renders them unstable however, as there always exists one negative
2338: eigenvalue of the Hessian which triggers a decay to the vacuum configuration.
2339:
2340: Our results are concrete and complete for abelian and nonabelian $Q$-soliton
2341: configurations which are diagonal in the oscillator basis (or related to
2342: such by global symmetry).
2343: For this case, we proved that the spectrum of the Hessian consists of the
2344: essential spectrum $[0,\infty)$ and an eigenvalue $\lambda_Q$ of multiplicity
2345: one with $-2Q<\lambda_Q<0$.
2346: This assertion was confirmed numerically, and the value of $\lambda_Q$
2347: was given as a zero of a particular function composed of monomials in~$\lambda$
2348: and the special function $\e^{-\lambda}\operatorname{li}(\e^{\lambda})$.
2349:
2350: Furthermore, the complete set of zero modes of the Hessian was identified.
2351: Each abelian diagonal $Q$-soliton background is characterized by a diagonal
2352: projector~$P$ of rank~$Q$, whose image and kernel trigger a decomposition of
2353: the space of fluctuations into three invariant subspaces, namely
2354: $u(\text{im}P)$, $u(\text{ker}P)$ and $\diff\,\text{Gr}(P)$. In addition,
2355: every side diagonal together with its transpose is seperately invariant
2356: under the action of the Hessian. This leads to a particular distribution
2357: of the admissible zero modes of the Hessian, displayed here for the example
2358: of~$Q{=}4$:
2359: \vskip-5mm
2360: \begin{figure}[H]
2361: \psfrag{phi=}{$\phi\ =\ $}
2362: \psfrag{im P}{$\scriptstyle{\text{im}P}$}
2363: \psfrag{ker P}{$\scriptstyle{\text{ker}P}$}
2364: \psfrag{d Gr(P)}{$\scriptstyle{\text{d\,Gr($P$)}}$}
2365: \psfrag{u(im P)}{$\scriptstyle{u(\text{im}P)}$}
2366: \psfrag{u(ker P)}{$\scriptstyle{u(\text{ker}P)}$}
2367: \begin{center}
2368: \includegraphics[origin=ct,width=80mm]{mat}
2369: \end{center}
2370: %\caption{}
2371: \end{figure}
2372: \vskip-5mm \noindent
2373: where the double line denotes the single negative eigenvector,
2374: each solid diagonal segment represents a real normalizable zero mode,
2375: the dashed line depicts an admissible non-normalizable zero mode,
2376: and empty areas do not contain admissible zero modes.
2377: In addition, each side diagonal features a non-admissible zero mode
2378: at the edge of the continuous part $[0,\infty)$ of the spectrum.
2379: We plot the complete spectrum of the Hessian at~$Q{=}4$
2380: (cut off at size $m_{\text{max}}{=}30$) for each invariant subspace
2381: $\Ecal_k^{\text{Gr}(P)}$ (boxes), $\Ecal_k^{\text{im}P}$ (stars),
2382: $\Ecal_k^{\text{ker}P}$ (crosses) and $\Ecal_0$ (circles), up to $k{=}6$:
2383: \vskip-5mm
2384: \begin{figure}[H]
2385: \psfrag{$k$}{$\scriptstyle{k}$}
2386: \psfrag{$lambda$}{$\scriptstyle{\lambda}$}
2387: \begin{center}
2388: \includegraphics[origin=ct,width=80mm]{ev3}
2389: \hspace{\fill}
2390: \includegraphics[origin=ct,width=80mm]{ev4}
2391: \end{center}
2392: \caption{Eigenvalues $\lambda$ of the cut-off Hessian (size~30) for $Q{=}4$
2393: in subspaces $\Ecal_k$}
2394: \end{figure}
2395: Since most soliton moduli are not associated with global symmetries,
2396: our explicit results for diagonal backgrounds do not obviously extend to
2397: generic (non-diagonal) backgrounds. We have not been able to compute the
2398: fluctuation spectrum across the entire soliton moduli space.
2399: The only exception is the abelian single-soliton solution which always is
2400: a translation of the diagonal configuration and therefore fully covered
2401: by our analysis. Already for the case of two U(1) solitons, the unitary
2402: transformation which changes their distance in the noncommutative plane
2403: is only partially known.
2404:
2405: This leads us to a number of unsolved problems. The most pressing one seems to
2406: be the extension of our fluctuation analysis to non-diagonal backgrounds.
2407: Next, following the even zero modes one can now find new non-BPS solutions.
2408: Also, it is worthwhile to investigate the commutative limit of the Hessian
2409: and its spectrum. Another interesting aspect is the existence of infinite-rank
2410: abelian projectors, i.e.~via an infinite array of coherent states, associated
2411: with BPS solutions of infinite energy. Furthermore, some technical questions
2412: concerning the admissible set of fields and their fluctuations have remained.
2413: Finally, it would be rewarding to extend the geometrical understanding of
2414: sigma models to the noncommutative realm.
2415:
2416: \bigskip
2417:
2418: \noindent
2419: {\large{\bf Acknowledgements}} \\
2420: The authors are indebted to A.D.~Popov for numerous fruitful discussions
2421: and thankful to R.~Wimmer and M.~Wolf for comments on the manuscript.
2422: This work was partially supported by the Deutsche Forschungsgemeinschaft (DFG)
2423: and the joint RFBR--DFG grant 436~RUS~113/669/0-2. In addition,
2424: A.V.D.~acknowledges support by the Russian Foundation for Basic Research,
2425: grants no.~04-01-00236 and 05-01-00981, as well as the Program in support of
2426: leading scientific schools no.~NSh-2040.2003.1. He is also very grateful to the
2427: Institut f\"ur Theoretische Physik at Universit\"at Hannover for hospitality.
2428:
2429:
2430: %\pagebreak
2431: \bigskip
2432:
2433: \begin{thebibliography}{99}
2434: %\addtolength{\itemsep}{-3pt}
2435:
2436: \bibitem{Lechtenfeld:2001uq}
2437: O.~Lechtenfeld, A.D.~Popov and B.~Spendig,\\
2438: {\em Noncommutative solitons in open N=2 string theory,}
2439: {\sl JHEP }{\bf 0106} (2001) 011 [hep-th/0103196].
2440: %%CITATION = HEP-TH 0103196;%%
2441:
2442: \bibitem{Zakrzewski}
2443: W.J.~Zakrzewski,
2444: ``Low dimensional sigma models,''
2445: Adam Hilger, 1989.
2446:
2447: \bibitem{Uhlenbeck}
2448: K.~Uhlenbeck,
2449: {\em Harmonic maps into Lie groups:
2450: classical solutions of the chiral model,}\\
2451: {\sl J.\ Diff.\ Geom.\ }{\bf 30} (1989) 1--50.
2452:
2453: \bibitem{Wood}
2454: J.C. Wood,
2455: {\em Explicit construction and parameterization of harmonic two spheres
2456: in the unitary group,}
2457: {\sl Proc.\ London Math.\ Soc.\ }{\bf 58} (1989) 608.
2458:
2459: \bibitem{Lechtenfeld:2001aw}
2460: O.~Lechtenfeld and A.D.~Popov,
2461: {\em Noncommutative multi-solitons in 2+1 dimensions,}\\
2462: {\sl JHEP }{\bf 0111} (2001) 040 [hep-th/0106213].
2463: %%CITATION = HEP-TH 0106213;%%
2464:
2465: \bibitem{Solovyov:2000xy}
2466: A.~Solovyov,
2467: {\em On noncommutative solitons,}\\
2468: {\sl Mod.\ Phys.\ Lett.\ A}{\bf 15} (2000) 2205 [hep-th/0008199].
2469: %%CITATION = HEP-TH 0008199;%%
2470:
2471: \bibitem{Aganagic:2000mh}
2472: M.~Aganagic, R.~Gopakumar, S.~Minwalla and A.~Strominger,\\
2473: {\em Unstable solitons in noncommutative gauge theory,}
2474: {\sl JHEP }{\bf 0104} (2001) 001 [hep-th/0009142].
2475: %%CITATION = HEP-TH 0009142;%%
2476:
2477: \bibitem{Gross:2000ss}
2478: D.J.~Gross and N.A.~Nekrasov,
2479: {\em Solitons in noncommutative gauge theory,}\\
2480: {\sl JHEP }{\bf 0103} (2001) 044 [hep-th/0010090].
2481: %%CITATION = HEP-TH 0010090;%%
2482:
2483: \bibitem{Hadasz:2001cn}
2484: L.~Hadasz, U.~Lindstr\o{}m, M.~Ro\v{c}ek and R.~von Unge,
2485: {\em Noncommutative multisolitons:\
2486: Moduli spaces, quantization, finite theta effects and stability,}
2487: {\sl JHEP }{\bf 0106} (2001) 040 [hep-th/0104017].
2488: %%CITATION = HEP-TH 0104017;%%
2489:
2490: \bibitem{Fujii:2001wp}
2491: A.~Fujii, Y.~Imaizumi and N.~Ohta,
2492: {\em Supersymmetry, spectrum and fate of D0-Dp systems with B-field,}
2493: {\sl Nucl.\ Phys.\ B} {\bf 615} (2001) 61 [hep-th/0105079].
2494: %%CITATION = HEP-TH 0105079;%%
2495:
2496: \bibitem{Durhuus:2001nj}
2497: B.~Durhuus, T.~Jonsson and R.~Nest,
2498: {\em The existence and stability of noncommutative scalar solitons,}
2499: {\sl Commun.\ Math.\ Phys.\ }{\bf 233} (2003) 49 [hep-th/0107121].
2500: %%CITATION = HEP-TH 0107121;%%
2501:
2502: \bibitem{Lechtenfeld:2001gf}
2503: O.~Lechtenfeld and A.D.~Popov,
2504: {\em Scattering of noncommutative solitons in 2+1 dimensions,}\\
2505: {\sl Phys.\ Lett.\ B} {\bf 523} (2001) 178 [hep-th/0108118].
2506: %%CITATION = HEP-TH 0108118;%%
2507:
2508: \bibitem{Wolf:2002jw}
2509: M.~Wolf,
2510: {\em Soliton antisoliton scattering configurations in a noncommutative
2511: sigma model in 2+1 dimensions,}
2512: {\sl JHEP }{\bf 0206} (2002) 055 [hep-th/0204185].
2513: %%CITATION = HEP-TH 0204185;%%
2514:
2515: \bibitem{Ihl:2002kz}
2516: M.~Ihl and S.~Uhlmann,
2517: {\em Noncommutative extended waves and soliton-like configurations
2518: in N=2 string theory,}
2519: {\sl Int.\ J.\ Mod.\ Phys.\ A }{\bf 18} (2003) 4889 [hep-th/0211263].
2520: %%CITATION = HEP-TH 0211263;%%
2521:
2522: \bibitem{Ward:1988ie}
2523: R.S.~Ward,
2524: {\em Soliton solutions in an integrable chiral model in 2+1 dimensions,}\\
2525: {\sl J.\ Math.\ Phys.\ }{\bf 29} (1988) 386--389.
2526: %%CITATION = JMAPA,29,386;%%
2527:
2528: \bibitem{Ward:1990vc}
2529: R.S.~Ward,
2530: {\em Classical solutions of the chiral model, unitons,
2531: and holomorphic vector bundles,}\\
2532: {\sl Commun.\ Math.\ Phys.\ }{\bf 128} (1990) 319.
2533: %%CITATION = CMPHA,128,319;%%
2534:
2535: \bibitem{Ioannidou:1998jh}
2536: T.A.~Ioannidou and W.J.~Zakrzewski,
2537: {\em Solutions of the modified chiral model in (2+1) dimensions,}
2538: {\sl J.\ Math.\ Phys.\ }{\bf 39} (1998) 2693 [hep-th/9802122].
2539: %%CITATION = HEP-TH 9802122;%%
2540:
2541: \bibitem{Harvey:2001yn}
2542: J.A.~Harvey,
2543: {\em Komaba lectures on noncommutative solitons and D-branes,}
2544: hep-th/0102076.
2545: %%CITATION = HEP-TH 0102076;%%
2546:
2547: \bibitem{Douglas:2001ba}
2548: M.R.~Douglas and N.A.~Nekrasov,
2549: {\em Noncommutative field theory,}\\
2550: {\sl Rev.\ Mod.\ Phys.\ }{\bf 73} (2001) 977 [hep-th/0106048].
2551: %%CITATION = HEP-TH 0106048;%%
2552:
2553: \bibitem{Szabo:2001kg}
2554: R.J.~Szabo,
2555: {\em Quantum field theory on noncommutative spaces,}\\
2556: {\sl Phys.\ Rept.\ }{\bf 378} (2003) 207 [hep-th/0109162].
2557: %%CITATION = HEP-TH 0109162;%%
2558:
2559: \bibitem{Matsuo:2000pj}
2560: Y.~Matsuo,
2561: {\em Topological charges of noncommutative soliton,}\\
2562: {\sl Phys.\ Lett.\ B }{\bf 499} (2001) 223 [hep-th/0009002].
2563: %%CITATION = HEP-TH 0009002;%%
2564:
2565: \bibitem{Harvey:2001pd}
2566: J.A.~Harvey,
2567: {\em Topology of the gauge group in noncommutative gauge theory,}
2568: hep-th/0105242.
2569: %%CITATION = HEP-TH 0105242;%%
2570:
2571: \bibitem{Otsu}
2572: H.~Otsu, T.~Sato, H.~Ikemori and S.~Kitakado,
2573: {\em New BPS solitons in 2+1 dimensional noncommutative CP(1) model,}
2574: {\sl JHEP }{\bf 0307} (2003) 054 [hep-th/0303090];\\
2575: %%CITATION = HEP-TH 0303090;%%
2576: {\em Lost equivalence of nonlinear sigma and CP(1) models
2577: on noncommutative space,}\\
2578: {\sl JHEP }{\bf 0406} (2004) 006 [hep-th/0404140].
2579: %%CITATION = HEP-TH 0404140;%%
2580:
2581: \bibitem{Gopakumar:2001yw}
2582: R.~Gopakumar, M.~Headrick and M.~Spradlin,
2583: {\em On noncommutative multi-solitons,}\\
2584: {\sl Commun.\ Math.\ Phys.\ }{\bf 233} (2003) 355 [hep-th/0103256].
2585: %%CITATION = HEP-TH 0103256;%%
2586:
2587: \bibitem{Lee:2004dt}
2588: K.M.~Lee,
2589: {\em Chern-Simons solitons, chiral model, and (affine) Toda model on
2590: noncommutative space,}
2591: {\sl JHEP }{\bf 0408} (2004) 054 [hep-th/0405244].
2592: %%CITATION = HEP-TH 0405244;%%
2593:
2594: \bibitem{Lee:2000ey}
2595: B.H.~Lee, K.M.~Lee and H.S.~Yang,
2596: {\em The CP(n) model on noncommutative plane,}\\
2597: {\sl Phys.\ Lett.\ B }{\bf 498} (2001) 277 [hep-th/0007140].
2598: %%CITATION = HEP-TH 0007140;%%
2599:
2600: \bibitem{Foda:2002nt}
2601: O.~Foda, I.~Jack and D.R.T.~Jones,
2602: {\em General classical solutions in the noncommutative $CP^{N-1}$ model,}
2603: {\sl Phys.\ Lett.\ B }{\bf 547} (2002) 79 [hep-th/0209111].
2604: %%CITATION = HEP-TH 0209111;%%
2605:
2606: \bibitem{Furuta:2002nv}
2607: K.~Furuta, T.~Inami, H.~Nakajima and M.~Yamamoto,
2608: {\em Non-BPS solutions of the noncommutative CP(1) model in (2+1)-dimensions,}
2609: {\sl JHEP }{\bf 0208} (2002) 009 [hep-th/0207166].
2610: %%CITATION = HEP-TH 0207166;%%
2611:
2612: \bibitem{Gopakumar:2000zd}
2613: R.~Gopakumar, S.~Minwalla and A.~Strominger,
2614: {\em Noncommutative solitons,}\\
2615: {\sl JHEP }{\bf 0005} (2000) 020 [hep-th/0003160].
2616: %%CITATION = HEP-TH 0003160;%%
2617:
2618: \bibitem{Kato}
2619: T.~Kato,
2620: ``Perturbation theory for linear operators,''
2621: Springer, 1966.
2622:
2623: \bibitem{RS}
2624: M.~Reed and B.~Simon,
2625: ``Methods of modern mathematical physics, volume~IV: Analysis of operators,''
2626: Academic Press, 1978.
2627:
2628: \end{thebibliography}
2629:
2630: \end{document}
2631:
2632: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2633:
2634:
2635:
2636: