hep-th0305063/zm.tex
1: \documentclass[12pt]{article}
2: \usepackage{mathbbol}
3: 
4: \setlength{\textwidth}{155mm}
5: \setlength{\oddsidemargin}{0mm}
6: \setlength{\textheight}{210mm}
7: \setlength{\topmargin}{00mm}
8: \parskip=1ex plus0.5ex minus0.2ex
9: 
10: \def\beq{\begin{equation}}
11: \def\eeq{\end{equation}}
12: \def\bea{\begin{array}}
13: \def\eea{\end{array}}
14: \def\beqa{\begin{eqnarray}}
15: \def\eeqa{\end{eqnarray}}
16: \def\Pexp{{\rm Pexp}}
17: \def\cO{{\cal{O}}}
18: \def\cP{{\cal{P}}}
19: \def\cA{{\cal{A}}}
20: \def\cN{{\cal{N}}}
21: \def\cF{{\cal{F}}}
22: \def\cV{{\cal{V}}}
23: \def\cR{{\cal{R}}}
24: \def\cD{{\cal{D}}}
25: \def\k{{\bf k}}
26: \def\ff{{\rm{ff}}}
27: \def\zm{{\rm{zm}}}
28: \def\ddz{{\frac{d}{dz}}}
29: \def\Tr{{\rm Tr}}
30: \def\tr{{\rm tr}}
31: \def\diag{{\rm diag}}
32: \def\pl{{{\cal P}_\infty}}
33: \def\plo{{{\cal P}_\infty^0}}
34: \newcommand{\re}{\relax{\rm I\kern-.18em R}}
35: \newcommand{\qu}{{\rm I \! H}}
36: \newcommand{\refeq}[1]{\mbox{Eq.~(\ref{eq:#1})}}
37: \newcommand{\half}{{\scriptstyle{{1\over 2}}}}
38: \newcommand{\hhalf}{{\scriptstyle{{1/2}}}}
39: \newcommand{\third}{{\scriptstyle{{1\over 3}}}}
40: \newcommand{\quart}{{\scriptstyle{{1\over 4}}}}
41: \newcommand{\veps}{\varepsilon}
42: \newcommand{\vphi}{\varphi}
43: \newcommand{\zahlen}{{\mathbb{Z}}}
44: \newcommand{\ein}{{\Eins}}
45: 
46: \begin{document}
47: 
48: \hfill\hbox{INLO-PUB-06/03}
49: \vskip5mm
50: \begin{center}{\Large\bf Constituent monopoles through the eyes of fermion 
51: zero-modes}\\[1cm]
52: {\bf Falk Bruckmann, D\'aniel N\'ogr\'adi} and {\bf Pierre van Baal}\\[3mm]
53: {\em Instituut-Lorentz for Theoretical Physics, University
54: of Leiden,\\ P.O.Box 9506, NL-2300 RA Leiden, The Netherlands}
55: \end{center}
56: 
57: \section*{Abstract}
58: We use the fermion zero-modes in the background of multi-caloron solutions 
59: with non-trivial holonomy as a probe for constituent monopoles. We find in 
60: general indication for an extended structure. However, for well separated
61: constituents these become point-like. We analyse this in detail for the 
62: $SU(2)$ charge 2 case, where one is able to solve the relevant Nahm equation 
63: exactly, beyond the piecewize constant solutions studied previously. 
64: Remarkably the zero-mode density can be expressed in the high temperature 
65: limit as a function of the conserved quantities that classify the solutions 
66: of the Nahm equation. 
67: 
68: \section{Introduction}\label{sec:intro}
69: 
70: To describe regular monopoles in gauge theories a Higgs field is required. 
71: This de\-fines the abelian subgroup of the gauge field. Yet in the full 
72: non-abelian theory there is no Dirac string and a regular solution results, 
73: the well-known 't Hooft-Polyakov monopole~\cite{THPo}. In the strong 
74: interactions no such Higgs field should be present, but nevertheless it has 
75: been conjectured that a dual superconductor description, in which monopoles 
76: form the dual charges that condense, could explain confinement~\cite{DuSC}. 
77: This scenario receives some support from the studies in supersymmetric 
78: theories through Seiberg-Witten duality~\cite{SeWi}, although also the old 
79: center vortex picture is still under active study~\cite{GrRe}. Lattice studies 
80: based on abelian and center projections, and their respective notions of 
81: dominance~\cite{Suzu} are the main means through which one tries to address 
82: these issues. One relies on so-called gauge fixing singularities to identify 
83: the relevant monopole~\cite{AbPr} or center vortex degrees of freedom.
84: 
85: A more recent alternative to study the monopole content of gauge theories, 
86: without the need of addressing singular configurations, gauge fixing, nor 
87: introducing an extra Higgs field, has been through calorons, which are 
88: instantons at finite temperature. It has been found that calorons are actually 
89: made up from constituent monopoles~\cite{NCal,Lee,KvB}, which becomes 
90: most apparent when the background Polyakov loop is non-trivial (as in the 
91: confined phase), and the size of the caloron is larger than the inverse 
92: temperature (the extent of the euclidean time direction). The background 
93: Polyakov loop is defined in the periodic gauge $A_\mu(\vec x,t)=
94: A_\mu(\vec x,t+\beta)$ by its asymptotic value, or the so-called holonomy, 
95: \beq
96: \pl=\lim_{x\to\infty}\Pexp(\int_0^\beta A_0(\vec x,t)dt)=g^\dagger
97: \exp(2\pi i\diag(\mu_1,\mu_2,\ldots,\mu_n))g,
98: \eeq
99: where $g$ is the gauge rotation used to diagonalize $\pl$, whose eigenvalues 
100: $\exp(2\pi i\mu_j)$ can be ordered on the circle such that $\mu_1\leq\mu_2
101: \leq\ldots\leq\mu_n\leq\mu_{n+1}$, with $\mu_{n+j}\equiv 1+\mu_j$ and 
102: $\sum_{i=1}^n\mu_i=0$. The masses of the constituent monopoles are given 
103: by $8\pi^2\nu_j/\beta$, with $\nu_j\equiv\mu_{j+1}-\mu_j$, which add up 
104: to $8\pi^2/\beta$, consistent with the instanton action.
105: 
106: Solutions are known explicitly~\cite{Dubna} for $SU(n)$ charge 1 calorons. The 
107: $n$ constituents can have any spatial location, although all have the same 
108: location in time (they do, however, become static when well separated), 
109: and $n-1$ abelian phases complete its $4n$ parameters. Charge $k$ calorons 
110: can be viewed as composed of $kn$ monopoles, of which a class of axially 
111: symmetric configurations was constructed explicitly~\cite{BrvB}.
112: 
113: The purpose of this paper is to study in more detail these higher charge 
114: calorons, where the emphasis is on constructing the chiral fermion 
115: zero-modes. Charge 1 calorons have exactly one fermion zero-mode, which was 
116: shown for well separated constituents to be supported on one and only one 
117: constituent~\cite{MTCP,MTP}. We may change the constituent that supports the
118: zero-mode, by changing the fermion boundary conditions from (anti)periodic,
119: to being periodic up to a phase $\exp(2\pi iz)$ (from now on we will use the 
120: classical scale invariance to set $\beta=1$). For $z\in[\mu_j,\mu_{j+1}]$ the 
121: zero-mode is localized to what we will call type $j$ constituent monopoles 
122: (with mass $8\pi^2\nu_j$, and the appropriate $U^{n-1}(1)$ charge associated 
123: to their embedding in $SU(n)$). 
124: 
125: Lattice evidence has been gathered over recent years that these monopole 
126: constituents are present in dynamical configurations in the confined phase of 
127: gauge theories for $SU(2)$ using cooling~\cite{Ilg1,Ilg2}, and for $SU(3)$ 
128: using fermion zero-modes~\cite{Gattp} as a filter. It is somewhat of a  puzzle 
129: that these constituent monopoles had not been seen in earlier cooling studies 
130: (apart from when using twisted boundary conditions~\cite{MTAP}). That they 
131: remained unnoticed when using fermion zero-modes as a filter is, however, 
132: simply a consequence of the fact that these studies were restricted to the 
133: use of fixed fermion boundary conditions. Only when cycling through boundary 
134: conditions specified by periodicity up to a phase, the $SU(3)$ charge 1 
135: instanton configurations will show three separate constituent monopoles. 
136: In Fig.~\ref{fig:zmcycle} we show a typical example based on the exact 
137: solutions for $SU(3)$, closely following the observed behavior~\cite{Gattp} 
138: based on actual lattice simulations in the confined phase, which guarantees 
139: the background Polyakov loop to be non-trivial. In the high temperature 
140: phase, where the Polyakov loop is trivial, two of the constituents are 
141: massless and only one peak will be seen. These massless constituents are 
142: interesting in their own right, giving rise to so-called non-abelian 
143: clouds~\cite{Wein}, but they will not concern us here.
144: 
145: \begin{figure}[htb]
146: \vskip2.7cm
147: \special{psfile=zm12d60.ps voffset=-75 hoffset=-10 vscale=30.0 hscale=30.0}
148: \special{psfile=zm19d60.ps voffset=-75 hoffset=125 vscale=30.0 hscale=30.0}
149: \special{psfile=zm30d60.ps voffset=-75 hoffset=260 vscale=30.0 hscale=30.0}
150: \hskip0.5cm$z=12/60$\hskip3.1cm$z=19/60$\hskip3.1cm$z=30/60$
151: \vskip2.7cm
152: \special{psfile=zm43d60.ps voffset=-75 hoffset=-10 vscale=30.0 hscale=30.0}
153: \special{psfile=zm48d60.ps voffset=-75 hoffset=125 vscale=30.0 hscale=30.0}
154: \special{psfile=zm58d60.ps voffset=-75 hoffset=260 vscale=30.0 hscale=30.0}
155: \hskip0.5cm$z=43/60$\hskip3.1cm$z=48/60$\hskip3.1cm$z=58/60$
156: \caption{The logarithm of the properly normalized zero-mode density for a 
157: typical $SU(3)$ caloron of charge 1, cycling through $z$. Shown are $z=\mu_j$ 
158: (for linear plots see Fig.~\ref{fig:zmrcycle}) and three values of $z$ roughly 
159: in the middle of each interval $z\in[\mu_j,\mu_{j+1}]$. All plots are on the 
160: same scale, cutoff for values of the logarithm below -5. The zero-mode with 
161: anti-periodic boundary conditions is found at $z=30/60$. For the action density
162: of the associated gauge field, see Ref.~\cite{MTP,WWW}.}\label{fig:zmcycle}
163: \vskip2.9cm
164: \special{psfile=zm19d60l.ps voffset=-75 hoffset=-10 vscale=30.0 hscale=30.0}
165: \special{psfile=zm43d60l.ps voffset=-75 hoffset=125 vscale=30.0 hscale=30.0}
166: \special{psfile=zm58d60l.ps voffset=-75 hoffset=260 vscale=30.0 hscale=30.0}
167: \hskip0.3cm$z=19/60$\hskip3.1cm$z=43/60$\hskip3.1cm$z=58/60$
168: \caption{The properly normalized delocalized zero-mode densities for 
169: $z=\mu_j$, on a {\em linear} scale (cmp. Fig.~\ref{fig:zmcycle}). 
170: }\label{fig:zmrcycle}
171: \end{figure}
172: 
173: We do not only study the fermion zero-modes for the higher charge calorons 
174: to compare with lattice simulations, but also as a tool to understand to 
175: which extend the constituent monopoles can be unambiguously identified in 
176: the caloron solutions. In the high temperature limit the non-abelian cores 
177: of the monopoles shrink to zero-sizes, and one is left with abelian gauge 
178: fields. Without taking the high temperature limit, but excluding the 
179: non-abelian cores, the same physics describes what we called the far field 
180: region~\cite{BrvB}. It would be desirable if the abelian field in this region
181: is described by point-like Dirac monopoles (actually dyons because of 
182: self-duality), when the constituents are well separated. For charge 1 calorons 
183: and the class of axially symmetric solutions studied before~\cite{BrvB} the 
184: density of the fermion zero-modes become Dirac delta functions at the 
185: locations of the constituent monopoles in the high temperature limit, 
186: for any constituent separation. This infinite localization in the high 
187: temperature limit can be understood from the fact that for most $z$ values 
188: there is an effective mass for the fermions. Therefore, by studying these 
189: zero-modes in the general case, we hope to learn more about the localization 
190: of the monopoles. 
191: 
192: It is not directly obvious that any higher charge caloron can be described 
193: by point-like constituents in the far field limit. The tool to construct 
194: these solutions is the Nahm equation~\cite{Nahm,NCal}, which is a duality 
195: transformation that maps the problem of finding charge $k$ calorons to that 
196: of $U(k)$ gauge fields on a circle with specific singularities at $z=\mu_j$. 
197: In the cases studied so far, the dual (Nahm) gauge field can be made piecewize 
198: constant, from which one easily reads off the constituent monopole locations. 
199: In general, however, the Nahm gauge fields depend non-trivially on $z$, and it 
200: is important to understand what this implies for the localization of the 
201: constituents. Quite remarkably, we will nevertheless find that in the high 
202: temperature limit the fermion zero-mode density does not depend on $z$ and can 
203: be expressed in terms of the conserved quantities of the Nahm equation. We 
204: compute it explicitly for charge 2, revealing in general an extended
205: structure. However, the extended structure collapses to isolated points for 
206: well separated constituents. This is so as long as $z\neq\mu_j$, which is the 
207: value where the localization of the zero-modes jumps. We separately study 
208: in the high temperature limit the case where $z=\mu_j$, for which the fermion 
209: zero-modes delocalize (decaying algebraically). Support of the zero-modes is 
210: now on constituent monopoles of types $j-1$ and $j$. We analyse this in 
211: detail for the class of axially symmetric solutions of Ref.~\cite{BrvB}, in 
212: the light of some puzzles concerning so-called bipole zero-modes~\cite{Bip}.
213: 
214: This paper is organized as follows: first we describe the caloron zero-modes 
215: in Sect.~\ref{sec:zm} and set up the Green's function calculation in 
216: Sect.~\ref{sec:zmlim}, to be able to discuss the zero-mode and far field 
217: limits. Sect.~\ref{sec:bipole} deals with the case where $z=\mu_j$, for which 
218: zero-modes are delocalized. In Sect.~\ref{sec:cons} we relate the zero-mode 
219: density for $z\neq\mu_j$ to conserved quantities of the Nahm equation and 
220: study in Sect.~\ref{sec:chtwo} the general solution for $SU(2)$ charge 2 
221: caloron. We conclude with a discussion of the implications of our results 
222: and the problems that still need to be addressed. Two appendices provide the 
223: details for the zero-mode and far field limit calculations. 
224: 
225: \section{Fermion zero-modes}\label{sec:zm}
226: 
227: We wish to construct the zero-modes of the chiral Dirac equation in 
228: the background of a self-dual gauge field at finite temperature. 
229: The Dirac equation in its two-component Weyl form, with $\bar
230: \sigma_\mu=\sigma_\mu^\dagger=(\ein_2,-i\vec \tau)$ ($\tau_i$ are the
231: usual Pauli matrices), reads
232: \beq
233: \bar D\hat\Psi_z(x)\equiv\bar\sigma_\mu D_\mu\hat\Psi_z(x)\equiv
234: \bar\sigma_\mu(\partial_\mu+A_\mu(x))\hat\Psi_z(x).\label{eq:Dirac}
235: \eeq
236: Assuming the gauge field is periodic in the imaginary time direction, with 
237: period $\beta=1$, we seek the zero-modes that satisfy the boundary condition 
238: \beq
239: \hat\Psi_z(t+1,\vec x)=\exp(-2\pi iz)\hat\Psi_z(t,\vec x),
240: \eeq
241: A simple abelian gauge transformation, $\Psi_z(x)=\exp(2\pi i zt)
242: \hat\Psi_z(x)$, makes the zero-mode periodic. This gauge transformation
243: replaces the gauge field by $A_\mu(x)-2\pi i z_\mu\ein_n$, but it does 
244: not change the field strength, such that existence of the appropriate 
245: number of zero-modes is guaranteed by the index theorem. 
246: 
247: The caloron solutions are obtained by Fourier transformation, reformulating 
248: the algebraic ADHM (Atiyah-Drinfeld-Hitchin-Manin) construction~\cite{ADHM} 
249: of multi-instanton solutions in $R^4$, as the Nahm transformation~\cite{NCal}. 
250: For this the instantons in $R^4$ periodically repeat themselves in the 
251: imaginary time direction up to a gauge rotation with $\pl$. This Fourier 
252: transformation also selects out of the infinite number of fermion zero-modes 
253: in $R^4$, those that satisfy the correct periodicity.
254: 
255: In the ADHM formulation the $k$ normalized fermion zero-modes for a charge $k$ 
256: $SU(N)$ instanton are given by ($\veps\equiv i\tau_2=\sigma_2$, $i=1,\ldots,n$ 
257: is the color index, $l=1,\ldots,k$ the charge index and $I=1,2$ the spinor 
258: index)
259: \beq
260: \Psi_{iI}^l(x)=\pi^{-1}\left(\phi^{-\hhalf}(x) u^\dagger(x)f_x\veps
261: \right)_{iI}^l,\quad\phi(x)=\ein_n+u^\dagger(x)u(x),\label{eq:Psi}
262: \eeq
263: with $u^\dagger(x)$ given explicitly in terms of the ADHM parameters by 
264: \beq
265: u^\dagger(x)=\lambda(B-\ein_k x)^{-1},\quad B=\sigma_\mu B_\mu,
266: \quad x=x_\mu\sigma_\mu,
267: \eeq
268: where $\lambda=(\lambda_1,\ldots,\lambda_k)$, with $\lambda_i^\dagger$ a 
269: two-component spinor in the $\bar n$ representation of $SU(n)$ ($\lambda$ 
270: can be seen as a $n\times 2k$ complex matrix), and $B_\mu$ hermitian 
271: $k\times k$ matrices. As $\phi(x)$ is an $n\times n$ positive hermitian 
272: matrix (for $n=2$ proportional to $\ein_2$), its square root is well-defined. 
273: We also recall the gauge field is given by
274: \beq
275: A_\mu(x)=\phi^{-\hhalf}(x)(u^\dagger(x)\partial_\mu u(x))\phi^{-\hhalf}(x)
276: +\phi^{\hhalf}(x)\partial_\mu\phi^{-\hhalf}(x).
277: \label{eq:aadhm}
278: \eeq
279: which can be shown to be self-dual provided the quadratic ADHM constraint is 
280: satisfied,
281: \beq
282: \lambda^\dagger\lambda+(B-\ein_k x)^\dagger(B-\ein_k x)=\sigma_0 f^{-1}_x,
283: \label{eq:adhmconstr}
284: \eeq
285: which implicitly defines $f_x$ as a hermitian $k\times k$ Green's function,
286: thereby completing the description of \refeq{Psi}. A further 
287: simplication~\cite{KvB,Temp} will be helpful, which uses the fact 
288: that $2(B_\mu-x_\mu)=\partial_\mu f_x^{-1}$ and $u^\dagger(x)=\phi(x)
289: \lambda f_x(B-\ein_k x)^\dagger$, implying
290: \beqa
291: A_\mu(x)&=&\half\phi^\hhalf(x)\lambda\bar\eta_{\mu\nu}\partial_\nu f_x
292: \lambda^\dagger\phi^\hhalf(x)+\half[\phi^{-\hhalf}(x),\partial_\mu
293: \phi^\hhalf(x)],\nonumber\\ \phi^{-1}(x)&=&1-\lambda f_x\lambda^\dagger,
294: \quad\Psi_{iI}^l(x)=(2\pi)^{-1}(\phi^\hhalf(x)\lambda\partial_\mu 
295: f_x\bar\sigma_\mu\veps)_{iI}^l,\label{eq:defA}
296: \eeqa
297: with $\bar\eta_{\mu\nu}=\bar\eta^j_{\mu\nu}\sigma_j=\bar\sigma_{[\mu}
298: \sigma_{\nu]}$ the anti-selfdual 't Hooft tensor.
299: 
300: As mentioned above, calorons are obtained by arranging the instantons in 
301: $R^4$ to be periodic (up to a gauge rotation). The time interval $[0,1]$ 
302: will contain as many instantons as the topological charge of the caloron. 
303: One splits the charge index $l$ as $l=pk+a$, where $a$ labels the $k$ 
304: instantons in the interval $[0,1]$ and $p$ labels the infinite number of 
305: repeated time-intervals, playing the role of the Fourier index. We find, 
306: suppressing the gauge and spinor index (cmp. Ref.~\cite{MTCP,MTP}), 
307: \beq
308: \hat\Psi_z^a(x)=(2\pi)^{-1}\phi^\hhalf(x)\partial_\mu\int_0^1 dz'
309: \hat\lambda_b(z')\hat f^{ba}_x(z',z)\bar\sigma_\mu\veps.\label{eq:Psihat}
310: \eeq
311: where the Fourier transforms of $\lambda$ and $f_x$ are denoted by 
312: $\hat\lambda_a(z)$ and $\hat f_x^{ba}(z',z)$.
313: The fermion zero-modes thus constructed are in the so-called algebraic gauge, 
314: for which $\hat\Psi_z(t+1,\vec x)=\exp(-2\pi iz)\pl\hat\Psi_z(t,\vec x)$. In 
315: this gauge all components of $A_\mu$ vanish at spatial infinity and the 
316: non-trivial holonomy is encoded in the boundary condition, which for the 
317: gauge field reads $A_\mu(t+1,\vec x)=\pl A_\mu(t,\vec x){\cal P}_\infty^{-1}$. 
318: A simple time-dependent gauge transformation allows one to transform to the 
319: periodic gauge, after which $A_0$ goes to a constant at spatial infinity.
320: 
321: To encode the appropriate periodicity in the ADHM parameters we need to take 
322: $\lambda_{pk+a}={\cal P}_\infty^p\zeta_a$~\cite{BrvB}. Introducing the 
323: projections $P_m$ on the eigenvalues of $\pl$, i.e. $\pl\equiv\sum_{m=1}^n
324: \exp(2\pi i\mu_m)P_m$, we find $\hat\lambda_a(z)=\sum_{m=1}^n\delta(z-\mu_m)
325: P_m\zeta_a$, which makes the expression for the zero-modes particularly simple,
326: \beq
327: \hat\Psi_z^a(x)=(2\pi)^{-1}\phi^\hhalf(x)\sum_{m=1}^n P_m\zeta_b
328: \bar\sigma_\mu\veps\partial_\mu\hat f^{ba}_x(\mu_m,z).\label{eq:zm}
329: \eeq
330: The fact that we are dealing with higher charge calorons, is reflected in 
331: the presence of the indices $a,b=1,\ldots,k$. Making use of the well-known 
332: identity~\cite{Temp,Osb,CoG} in $R^4$, $\Psi_{iI}^l(x)^*\Psi_{iI}^m(x)=
333: -(2\pi)^{-2}\partial_\mu^2 f^{lm}_x$, Fourier transformation gives the
334: appropriate expression for the caloron zero-mode density
335: \beq
336: \hat\Psi_z^a(x)^\dagger\hat\Psi_z^b(x)=-(2\pi)^{-2}
337: \partial_\mu^2\hat f^{ab}_x(z,z).\label{eq:zdens}
338: \eeq
339: Using the fact that $\lim_{|\vec x|\rightarrow\infty}|\vec x|\hat 
340: f^{ab}_x(z,z)=\pi\delta^{ab}$, the zero-modes $\hat\Psi_z^a$ are seen 
341: to be orthonormal.
342: 
343: We close this section by remarking that for $SU(2)$ an alternative 
344: construction is possible, as part of the $Sp(n)$ series (since $Sp(1)=
345: SU(2)$). The ADHM construction for $Sp(n)$ is based on quaternions. 
346: In particular $\lambda_l$ is assumed to be a quaternion, whereas $B_\mu$ 
347: is now real-symmetric. All formulae presented above remain valid, but it 
348: should be noted that the transformation $\lambda\rightarrow\lambda 
349: T^\dagger$, $B_\mu\rightarrow T B_\mu T^\dagger$, with $T\in U(k)$, 
350: leaving the gauge field and the ADHM constraint untouched, has to 
351: be replaced by $T\in O(k)$.
352: 
353: \section{Zero-mode and far field limit}\label{sec:zmlim}
354: 
355: As we have seen above, all physical quantities can be reconstructed, once 
356: we have found the Green's function $\hat f^{ab}_x(z,z')$. Here we review the
357: necessary ingredients. We start with the fact that the Green's function is 
358: defined through an impurity scattering problem~\cite{BrvB}
359: \beq
360: \left\{-\frac{d^2}{dz^2}+V(z;\vec x)\right\}
361: \!f_x(z,z')=4\pi^2 \ein_k\delta(z\!-\!z'),\label{eq:Green}
362: \eeq
363: where $f_x(z,z')$ is related to $\hat f_x(z,z')$ by a $U(k)$ gauge 
364: transformation 
365: \beq
366: f_x(z,z')=\hat g(z)\hat f_x(z,z')\hat g^\dagger(z'),\quad
367: \hat g(z)=\exp\left(2\pi i(\xi_0-x_0\ein_k)z\right).\label{eq:dftrans}
368: \eeq
369: The ``potential" $V$, which includes ``impurity" contributions,
370: is determined by the (dual) $U(k)$ Nahm gauge field $\hat A_\mu(z)$ 
371: \beqa
372: &&V(z;\vec x)=4\pi^2\vec R^2(z;\vec x)+2\pi\sum_m\delta(z-\mu_m)S_m,\quad
373: S_m=\hat g(\mu_m)\hat S_m\hat g^\dagger(\mu_m),\nonumber\\&&R_j(z;\vec x)
374: =x_j\ein_k-(2\pi i)^{-1}\hat g(z)\hat A_j(z)\hat g^\dagger(z),\qquad\,
375: \hat S_m^{ab}=\pi\tr_2(\zeta_a^\dagger P_m\zeta_b).\label{eq:Rdef}
376: \eeqa
377: Fourier transformation of $B_\mu$ gives $(2\pi i)^{-1}\delta(z-z')(
378: \delta_{\mu0}\ein_k\ddz+\hat A_\mu(z))$ and defines the Nahm gauge field. 
379: We have further used the fact that one can choose a $U(k)$ gauge in which 
380: $\hat A_0(z)\equiv 2\pi i\xi_0$ is constant, which itself can be transformed 
381: to zero by $\hat g(z)$. Note that $\hat g(1)$ plays the role of the holonomy 
382: associated to the dual Nahm gauge field $\hat A_\mu(z)$. Crucial is that we 
383: can formulate the zero-mode and far field limits without specifying the 
384: solutions of the Nahm equations. These Nahm equations, which are equivalent 
385: to the ADHM constraint, are given by 
386: \beqa
387: &&\ddz\hat A_j(z)+[\hat A_0(z),\hat A_j(z)]+\half\veps_{jk\ell}[\hat A_k(z),
388: \hat A_\ell(z)]=2\pi i\sum_m\delta(z-\mu_m)\rho_m^{\,j},\nonumber\\
389: &&\hskip4cm\vec\rho_m^{\,ab}\equiv-\pi\tr_2\left(\zeta_a^\dagger P_m\zeta_b
390: \vec\tau\right).\label{eq:nahm}
391: \eeqa
392: This will be discussed in more detail in Sect.~\ref{sec:chtwo}.
393: 
394: To solve the second order equation for $f_x(z,z')$ it is convenient to convert 
395: it to a first order equation, involving $2k\times 2k$ matrices, 
396: \beq
397: \left(\ddz-\pmatrix{0&\ein_k\cr V(z;\vec x)&0\cr}\right)\pmatrix{\quad 
398: f_x(z,z')\cr\ddz f_x(z,z')\cr}=-4\pi^2\delta(z-z')\pmatrix{0\cr\ein_k\cr},
399: \eeq
400: which can be solved as
401: \beq
402: \pmatrix{\quad f_x(z,z')\cr\ddz f_x(z,z')\cr}=-4\pi^2 W(z,z_0)\left\{
403: (\ein_{2k}-\cF_{z_0})^{-1}-\theta(z'-z)\ein_{2k}\right\}W^{-1}(z',z_0)
404: \pmatrix{0\cr\ein_k\cr},\label{eq:fdf}
405: \eeq
406: where $z_0$ can be arbitrary and 
407: \beq
408: W(z_2,z_1)=\Pexp\left[\int_{z_1}^{z_2}\pmatrix{0&\ein_k\cr V(z;\vec x)&0\cr}
409: dz\right],\quad\cF_{z_0}=\hat g^\dagger(1)W(z_0+1,z_0).\label{eq:Wdef}
410: \eeq
411: In particular, one can show that~\cite{BrvB} 
412: \beq
413: -\half\Tr_n F_{\mu\nu}^2(x)=-\half\partial_\mu^2\partial_\nu^2\log 
414: \psi(x),\quad\psi=\det\left(ie^{-\pi ix_0}(\ein_{2k}-\cF_{z_0})
415: /\sqrt{2}\right).\label{eq:defpsi}
416: \eeq
417: To isolate the exponential contributions in \refeq{Wdef}, one introduces two 
418: solutions of the Riccati equation~\cite{BrvB},
419: \beq
420: R_m^\pm(z)^2\pm\frac{1}{2\pi}\frac{d}{dz}\,R_m^\pm(z)=\vec
421: R^2(z;\vec x).\label{eq:Riccati} 
422: \eeq
423: Since
424: $\vec R(z;\vec x)\to\vec x\ein_k$ for $|\vec x|\to\infty$, we find in this
425: limit that $R_m^\pm(z)\to|\vec x|\ein_k$. Defining
426: \beq 
427: f_m^\pm(z)=P\exp\left[\pm 2\pi\int_{\mu_m}^z \!\!\!R_m^\pm(z)dz\right],
428: \quad z\in[\mu_m,\mu_{m+1}],\label{eq:fdef}
429: \eeq
430: we see that $f_m^\pm(z)\to\exp\left(\pm 2\pi|\vec x|(z-\mu_m)\ein_k\right)$.
431: These are the exponentially rising and decaying solutions of \refeq{Green},
432: in terms of which we can rewrite for $z,z'\in(\mu_m,\mu_{m+1})$
433: \beq
434: W(z,z')=W_m(z)W_m^{-1}(z'),\quad W_m(z)\equiv\hat W_m(z)F_m(z),
435: \eeq
436: where $F_m(z)$ contains all exponential factors, 
437: \beq
438: \hat W_m(z)=\pmatrix{\ein_k&\ein_k\cr 2\pi R^+_m(z)&-2\pi R^-_m(z)\cr},
439: \quad F_m(z)=\pmatrix{f_m^+(z)&0\cr0&f_m^-(z)\cr}.\label{eq:Wmdef}
440: \eeq
441: 
442: As an illustration let us assume $\vec R(z;\vec x)^2=(\vec x\ein_k-\vec e\,
443: Y_m)^2\equiv R^2_m$, independent of $z$, for $z\in[\mu_m,\mu_{m+1}]$, as is 
444: the case for charge one~\cite{KvB} and a class of axially symmetric 
445: solutions~\cite{BrvB} (see Sect.~\ref{sec:bipole}). We then find $R^+_m(z)=
446: R^-_m(z)=R_m$ and $f_m^\pm(z)=\exp(2\pi(z-\mu)R_m)$. Diagonalizing $R_m$ 
447: defines $k$ locations (the eigenvalues of $Y_m$), which are sharply defined 
448: and give rise to point-like constituents in the high temperature limit. When 
449: $\vec R(z;\vec x)$ is not piecewize constant, these locations are a priori not 
450: sharply defined. We still expect the cores to be the regions in $\vec x$ where 
451: $\vec R(z;\vec x)$ remains small (cmp. \refeq{Rdef}). The separations between 
452: the cores of monopoles of different type is controlled by the discontinuity in 
453: $\hat A_j(z)$, \refeq{nahm}, and can in general be chosen large. This allows 
454: us to define the zero-mode limit, where $\vec x$ is assumed to be far removed 
455: from any constituent core {\em not} of type $m$. In technical terms that means 
456: that $R_{m'}^\pm(z)$ is large for all $m'\neq m$. Accordingly, $f_{m'}^-(z)$ 
457: and $f_{m'}^+(z)^{-1}$ are exponentially small for these values of $\vec x$ 
458: (cmp. \refeq{fdef}). Results that are valid up to these exponential correction 
459: are denoted by a subscript ``$\zm$" for the zero-mode limit and ``$\ff$" for 
460: the far field limit. In the latter case, $\vec x$ is assumed to be far removed 
461: from {\em all} constituents.
462: 
463: In appendix A we show that for $\mu_m\leq z'\leq z\leq\mu_{m+1}$ 
464: \beq
465: f^{\zm}_x(z,z')=\pi(e_m^+(z)-e_m^-(z)Z_{m+1}^-)\left(e_m^+-Z_m^+e_m^-
466: Z_{m+1}^-\right)^{-1}(\tilde e_m^+(z')-Z_m^+\tilde e_m^-(z'))R_m^{-1}(z').
467: \label{eq:zmlim}
468: \eeq
469: (for $z'>z$ one uses $f_x(z',z)=f_x^\dagger(z,z')$) with
470: \beqa
471: e_m^\pm(z)\equiv f_m^\mp(z)f_m^\mp(\mu_{m+1})^{-1},&&e_m^\pm\equiv 
472: e_m^\pm(\mu_m),\quad\tilde e_m^\pm(z)\equiv f_m^\mp(z)^{-1},\nonumber\\
473: Z_m^-\equiv\ein_k-2\Sigma_m^{-1}R_{m-1}(\mu_m),&& Z_m^+\equiv\ein_k-2
474: \Sigma_m^{-1}R_m(\mu_m),\label{eq:Zedef}\\R_m(z)\equiv\half(R_m^+(z)+
475: R_m^-(z)),&&\Sigma_m\equiv R_m^-(\mu_m)+R_{m-1}^+(\mu_m)+S_m.\nonumber
476: \eeqa
477: This result is valid up to {\em exponential} corrections as long as $\vec x$ 
478: is far removed from all constituents of type $m'\neq m$. Note that in this 
479: limit $Z_m^{\pm}=\ein_k$, however, only up to {\em algebraic} corrections, 
480: which is why we have kept them. In Fig.~\ref{fig:zmlocal} we give the 
481: zero-mode densities for a charge 2 axially symmetric solution in $SU(2)$, 
482: with well-separated constituents. We found no differences to 1 part in 
483: $10^6$ between the exact result and the result obtained with the zero-mode 
484: limit, \refeq{zmlim}. The choice of basis for the zero-modes involves the 
485: gauge rotation that for each interval $z\in[\mu_m,\mu_{m+1}]$ identifies 
486: the constituent locations, cmp. Sect.~\ref{sec:bipole} and Ref.~\cite{BrvB}. 
487: This ensures that each zero-mode is localized to only one constituent monopole. 
488: 
489: \begin{figure}[htb]
490: \vspace{4.4cm}
491: \special{psfile=zm2a11.ps voffset=-110 hoffset=-95 vscale=50.0 hscale=50.0}
492: \special{psfile=zm2a22.ps voffset=-155 hoffset=-95 vscale=50.0 hscale=50.0}
493: \special{psfile=zm2p11.ps voffset=-110 hoffset=175 vscale=50.0 hscale=50.0}
494: \special{psfile=zm2p22.ps voffset=-155 hoffset=175 vscale=50.0 hscale=50.0}
495: \special{psfile=ExWe.eps  voffset=-160 hoffset=40 vscale=50.0 hscale=50.0}
496: \caption{Zero-mode densities for a typical charge 2, $SU(2)$ axially symmetric 
497: solution. For comparison the action density (cmp. Fig.~2 of Ref.~\cite{BrvB})
498: is shown in the middle. All are on a logarithmic scale, cutoff below $e^{-3}$. 
499: On the left is shown the two periodic zero-modes ($z=0$) and on the right 
500: the two anti-periodic zero-modes ($z=1/2$). 
501:  }\label{fig:zmlocal}
502: \end{figure}
503: 
504: It is now almost trivial to read off from this the far field limit for 
505: $f_x(z,z)$, where $\vec x$ is assumed to be far from {\em all} constituents. 
506: As long as $z\neq\mu_m$ and $z\neq\mu_{m+1}$, $e_m^-(z)$ and $\tilde e_m^-(z)$ 
507: are exponentially small, and $f^\ff_x(z,z)=\pi e_m^+(z)(e_m^+)^{-1}\tilde 
508: e_m^+(z)R_m^{-1}(z)$. But using the definitions of $e_m^\pm(z)$ and 
509: $\tilde e_m^\pm(z)$ all exponential factors precisely cancel and in 
510: the far field limit we are left with
511: \beq
512: f^\ff_x(z,z)=\pi R_m^{-1}(z),\quad z\in(\mu_m,\mu_{m+1}),\label{eq:fzzffl}
513: \eeq
514: which will play a very important role in Sects.~\ref{sec:cons} and 
515: \ref{sec:chtwo}. We can also read off the far field limit for $f_x(z,z')$ 
516: evaluated at the impurities $\mu_m$ and $\mu_{m+1}$, noting that by 
517: definition $e_m^\pm(\mu_{m+1})=\tilde e_m^\pm(\mu_m)=\ein_k$,
518: \beqa
519: f^\ff_x(\mu_{m+1},\mu_{m+1})&=&\pi(\ein_k-Z_{m+1}^-)R_m^{-1}(\mu_{m+1})=2\pi
520: \Sigma_{m+1}^{-1},\nonumber\\f^\ff_x(\mu_m,\mu_m)&=&\pi(\ein_k-Z_m^+)
521: R_m^{-1}(\mu_m)=2\pi\Sigma_m^{-1},\label{eq:fmmffl}
522: \eeqa
523: and $f^\ff_x(\mu_{m+1},\mu_m)=0$ (as well as $f^\ff_x(\mu_m,\mu_{m+1})=0$,
524: using $f_x(z,z')=f_x^\dagger(z',z)$), verifying the results of 
525: Ref.~\cite{BrvB}. Although $f_x(z,z)$ is continuous at the impurities, 
526: comparing \refeq{fzzffl} with \refeq{fmmffl} we see, as anticipated, that 
527: in the high temperature limit $f_x(z,z)$ is discontinuous at the impurities. 
528: At finite temperature, the transition across the impurity has a ``width" 
529: inversely proportional to the temperature.
530: 
531: For $k=1$, where $e_m^\pm(z)=\exp(\pm 2\pi(\mu_{m+1}-z)r_m)$ and 
532: $\tilde e_m^\pm(z)=\exp(\pm 2\pi(z-\mu_m)r_m)$ one finds for 
533: $z',z\in(\mu_m,\mu_{m+1})$
534: \beq
535: f^\zm_x(z',z)=\frac{2\pi\sinh\left(2\pi r_m(\mu_{m+1}-z'+\gamma^-_{m+1})\right)
536: \sinh\left(2\pi r_m(z-\mu_m+\gamma^+_m)\right)}{r_m\sinh\left(2\pi r_m
537: (\nu_m+\gamma_m^++\gamma_{m+1}^-)\right)},
538: \eeq
539: which agrees with the result of Ref.~\cite{MTP}, where only the limit with
540: $\gamma_m^+\equiv-\half\log Z_m^+$ and $\gamma_{m+1}^-\equiv-\half\log
541: Z_{m+1}^-$ neglected and $z'=z$ was considered. We stress again that the 
542: presence of $\gamma$ implies a subtle algebraically decaying influence 
543: due to the constituents of type $m-1$ and $m+1$ (the influence of all other 
544: constituents is decaying exponentially, although this is only relevant for 
545: $SU(n>3)$). In terms of constituent radii $r_m=|\vec x-\vec y_m|$ and 
546: locations $\vec y_m$ one has 
547: \beq
548: Z_{m+1}^-=\frac{|\vec y_{m+1}-\vec y_m|+r_{m+1}-r_m}{|\vec y_{m+1}-
549: \vec y_m|+r_{m+1}+r_m},\quad Z_m^+=\frac{|\vec y_{m-1}-\vec y_m|+
550: r_{m-1}-r_m}{|\vec y_{m-1}-\vec y_m|+r_{m-1}+r_m}.
551: \eeq
552: For $SU(2)$, with $\vec y_{m+1}=\vec y_{m-1}$, therefore $Z_m^+=Z_{m+1}^-$, 
553: or $\gamma_m^+=\gamma_{m+1}^-$, and the influence of the other constituent 
554: is only felt by a renormalization of the mass $\nu_m$, 
555: \beq
556: SU(2):\  f^\zm_x(z,z)=\pi\frac{\cosh\left(2\pi r_m(\nu_m+2\gamma_m)\right)-
557: \cosh\left(2\pi r_m(2z-\mu_m-\mu_{m+1})\right)}{r_m\sinh\left(2\pi r_m(\nu_m+
558: 2\gamma_m)\right)}.
559: \eeq
560: Note that for $SU(2)$ $\mu_1+\mu_2=0$ and $\mu_2+\mu_3=1$. This leads to the 
561: well-know result for the monopole zero-mode density $f^\zm_x(z,z)=\pi 
562: r^{-1}_m\tanh\left(\pi r_m(\nu_m+2\gamma_m)\right)$, with $z=0$ for $m=1$ 
563: (periodic zero-mode) and $z=1/2$ for $m=2$ (anti-periodic zero-mode).
564: 
565: \section{Bipole zero-modes}\label{sec:bipole}
566: 
567: In this section we discuss the zero-modes in the high temperature limit for
568: $z=\mu_m$, which means that $\pl e^{-2\pi iz}$ has one of its eigenvalues
569: equal to 1, which leads to one of the components of the fermion to become
570: massless. Indeed, using Eqs.~(\ref{eq:zdens},\ref{eq:dftrans},\ref{eq:fmmffl}) 
571: we find that in the far field limit
572: \beq
573: z=\mu_m:\qquad\hat\Psi_z^a(x)^\dagger\hat\Psi_z^b(x)=-(2\pi)^{-1}
574: \partial_i^2\left(\hat\Sigma_m^{-1}\right)^{ab},\quad\hat\Sigma_m\equiv
575: \hat g^\dagger(\mu_m)\Sigma_m\hat g(\mu_m),\label{eq:dens}
576: \eeq
577: decays algebraically and has support on the constituents of type $m-1$ and 
578: $m$, as is easy to see for $k=1$, where $\Sigma_m=|\vec x-\vec y_{m-1}|+
579: |\vec x-\vec y_m|+|\vec y_{m-1}-\vec y_m|$. Here we will restrict ourselves 
580: to $SU(2)$, particularly interesting for the axially symmetric caloron 
581: solutions, since in the high temperature limit its gauge field has the form 
582: of the so-called bipole ansatz (see \refeq{bipans}), which always has an 
583: integrable chiral fermion zero-mode~\cite{Bip}. In the bipole ansatz all Dirac 
584: strings have to run in the same direction, but other than that, the locations 
585: of the self-dual Dirac monopoles can be arbitrary. However, for the axially 
586: symmetric caloron solutions the constituents have to alternate between 
587: opposite charges on a line~\cite{BrvB}. In this case, with the solution coming 
588: from a regular caloron, there are always as many zero-modes as the number of
589: constituents with a given charge (equal to the topological charge $k$). 
590: By considering the case of solutions with topological charge 2, we will 
591: find the expressions for the bipole zero-mode and the extra zero-mode, in 
592: terms of the constituent locations only (which should be possible, since 
593: the abelian gauge field has this property in the high temperature limit). 
594: Remarkably, we will find that rearranging the order of the constituents, 
595: so as to violate the constraint coming from the axially symmetric caloron, 
596: the second zero-mode is no longer integrable (while the gauge field and 
597: bipole zero-mode remain well defined).
598: 
599: A particular class of axially symmetric caloron solutions is obtained by 
600: taking $Y^j_m\equiv\hat g(z)\frac{\hat A_j(z)}{2\pi i}\hat g^\dagger(z)$ 
601: to be piecewize constant. This can be shown to satisfy the Nahm or ADHM 
602: equation when we take~\cite{BrvB}
603: \beq
604: \zeta_a=\rho_a\exp(2\pi i\alpha_a)\zeta,\quad\alpha_a\equiv\sum_{m=1}^n
605: \alpha_a^m P_m,\quad\Tr_n\alpha_a=0,\quad \vec Y_m=Y_m\vec e,\label{eq:zadef}
606: \eeq
607: where $\rho_a$ are positive, not to be confused with 
608: \beq
609: \vec\rho_m^{\,ab}=\rho^a\rho^b\exp(2\pi i(\alpha_b^m-\alpha_a^m))\Delta
610: \vec y_m,\quad\Delta\vec y_m=\Delta y_m\vec e\equiv-\pi\tr_2(\zeta^\dagger 
611: P_m\zeta\vec\tau).\label{eq:rhodef}
612: \eeq
613: For $SU(2)$ one has $\Delta\vec y_1=-\Delta\vec y_2$ (for $SU(n>2)$ a
614: constraint on $\zeta$ is required, to guarantee that all $\Delta\vec y_m$ 
615: are parallel). We will take $\vec e=\Delta\vec y_2/|\Delta\vec y_2|=(0,0,1)$
616: (hence $\Delta y_2>0$) and $\pl=\exp(2\pi i\mu_2\tau_3)$ (therefore $\zeta
617: =\sqrt{\Delta y_2/\pi}\ein_2$). This can always be arranged to be the case 
618: by a global gauge rotation, and a spatial rotation. We define $\vec y_m$, 
619: up to an irrelevant overall shift, through $\Delta\vec y_m\equiv\vec y_m-
620: \vec y_{m-1}$. This fixes $Y_m$ to be
621: \beq
622: Y_m^{ab}=(\xi_a+\rho_a^2 y_m)\delta_{ab}+i(1-\delta_{ab})\rho_a\rho_b
623: \sum_{j= 1}^n\Delta y_j\frac{\exp\left(2\pi i\left[\alpha_b^j-\alpha_a^j
624: -(\mu_j+s_j^m)(\xi_0^b-\xi_0^a)\right]\right)}{2\sin\left(\pi
625: \left[\xi_0^b-\xi_0^a\right]\right)}.\label{eq:Ymfull}
626: \eeq
627: The constituent locations are found from the eigenvalues of $Y_m$. Although 
628: constant, it is not true in general that the $Y_m$ can be diagonalized 
629: simultaneously, making this a non-abelian solution of the Nahm equation. 
630: Yet, as we remarked before, one easily computes the Green's function, 
631: since $R^\pm_m(z)=R_m=\sqrt{(\vec x\ein_k-\vec e\,Y_m)\cdot(\vec x\ein_k-
632: \vec e\,Y_m)}$ is constant in $z$. Using that for the axially symmetric 
633: solutions $\vec\rho_m^{\,ab}=\hat S_m^{ab}\vec y_m/|\Delta\vec y_m|$, one 
634: shows~\cite{BrvB} that in the high temperature limit the (abelian) gauge 
635: field can be written in the form of the bipole ansatz~\cite{Bip}
636: \beq
637: A_\mu(x)=-\frac{i}{2}\tau_3\bar\eta^3_{\mu\nu}\partial_\nu\log\phi(x),
638: \quad\phi(x)=\phi_\ff(x)\equiv\frac{\det(R_1+R_2+S_2)}{\det(R_1+R_2-S_2)}.
639: \label{eq:bipans}
640: \eeq
641: In the bipole ansatz, one splits $A_\mu(x)$ in an isospin up and isospin down 
642: component (with inverted abelian charges). However, all that concerns us here 
643: is the fact that for {\em any} $\phi$, $A_\mu(x)$ as given above is self-dual 
644: (and hence a solution of the Maxwell equations) provided $\log\phi$ is 
645: harmonic away from Dirac string singularities (defined by $\phi^{-1}=0$). 
646: This always gives rise to at least one normalizable zero-mode of the chiral 
647: Dirac equation
648: \beq
649: \hat\Psi_{mI}(x)=(2\pi\tilde\rho)^{-1}\phi^{-\hhalf}(x)(\tau_1\bar\sigma_\mu
650: \veps)_{mI}\partial_\mu\log\phi(x).\label{eq:bipzm}
651: \eeq
652: were $\tilde\rho$ is simply a normalization factor. Here $m$ corresponds to the 
653: isospin component that survives for $z=\mu_m$ (with the other component
654: related to $\hat f_x(\mu_1,\mu_2)$ vanishing in the far field limit,
655: see \refeq{Psihat}). 
656: 
657: We now work out the explicit form of all $k$ zero-modes for the axially 
658: symmetric caloron solution, showing how the zero-mode in \refeq{bipzm} is
659: recovered from these. Using Eqs.~(\ref{eq:zm},\ref{eq:fmmffl},\ref{eq:rhodef}), 
660: together with the fact that $\hat f_x(\mu_1,\mu_2)$ is exponentially small
661: ($\tau_1$ is used for picking out the surviving component), we find for the 
662: normalized zero-modes at $z=\mu_m$ ($|\zeta|=\sqrt{\Delta y_2/\pi}$)
663: \beq
664: \hat\Psi_{mI}^a(x)=\phi_\ff^{\hhalf}(x)|\zeta|\rho_b e^{2\pi i\alpha_b^m}
665: (\tau_1\bar\sigma_\mu\veps)_{mI}\partial_\mu\left(\hat\Sigma_m^{-1}
666: \right)^{ba}.\label{eq:axzm}
667: \eeq
668: Using the fact that (cmp. \refeq{defA})
669: \beq
670: \phi^{-1}_\ff(x)=1-2\pi|\zeta|^2\rho_a e^{2\pi i\alpha_a^m}\left(
671: \hat\Sigma_m^{-1}\right)^{ab}e^{-2\pi i\alpha_b^m}\rho_b,
672: \eeq
673: the following linear combination recovers the bipole zero-mode
674: \beq
675: \hat\Psi_{mI}(x)=\frac{\rho_a e^{-2\pi i\alpha_a^m}}{\tilde\rho}\hat
676: \Psi_{mI}^a(x)=\frac{\phi_\ff^{\hhalf}(x)}{2\pi\tilde\rho}(\tau_1\bar
677: \sigma_\mu\veps)_{mI}\partial_\mu\left(1-\phi_\ff^{-1}(x)\right),\quad
678: \tilde\rho^2\equiv\sum_a\rho_a^2|\zeta|^2.
679: \eeq
680: By defining $\hat\Psi^{(j)}_{mI}(x)=\rho^{(j)}_a\hat\Psi_{mI}^a(x)$ with 
681: $\rho^{(j)}$ an orthonormal set of complex vectors, with $\rho^{(1)}_a\equiv
682: \rho_a e^{2\pi i\alpha_a^m}|\zeta|/\tilde\rho$ we form a complete set of 
683: orthonormal zero-modes. That these are solutions of \refeq{Dirac} is guaranteed
684: by the general formalism we developed, but one may of course check this by 
685: substitution in the Dirac equation. This requires $\partial_i\det\Phi
686: \partial_i\Phi^{-1}=0$, with $\Phi\equiv\left(\ein_k-2\hat S_m
687: \hat\Sigma_m^{-1}\right)^{-1}$.
688: 
689: For axially symmetric $SU(2)$ solutions with charge 2 we choose $\rho^{(2)}_a
690: =\veps_{ab}\rho^{(1)}_b$ and find for the two orthonormal zero-modes at 
691: $z=\mu_2$
692: \beq
693: \hat\Psi^{(i)}=\frac{1}{2\pi\tilde\rho}\phi_\ff^\hhalf(x)\pmatrix{\partial_2+
694: i\partial_1\cr-i\partial_3\cr}\phi_{(i)}^{-1}(x),\quad\phi_{(i)}^{-1}(x)\equiv
695: 2\pi\tilde\rho^2\left(\rho^{(i)}\right)^\dagger\hat\Sigma_2^{-1}\rho^{(1)}.
696: \eeq
697: with $\phi_{(1)}^{-1}=1-\phi_\ff^{-1}$, as shown in Ref.~\cite{BrvB}, only 
698: depending on the constituent locations $y_m^{(a)}$ read off from the 
699: eigenvalues of $Y_m$. Many choices of $y_m$, $\xi_a$, $\xi_0^a$, $\alpha_a^2$, 
700: $\rho_a$ and $\mu_2$ actually give rise to the {\em same} constituent 
701: locations, and hence the {\em same} expressions for $A_\mu(x)$ and 
702: $\hat\Psi^{(1)}$. It is important for consistency that this will hold for 
703: $\hat\Psi^{(2)}$ as well. Apart from an irrelevant phase, this is indeed the 
704: case (checked for many random choices of the parameters). The explicit analytic 
705: formulae in terms of the 4 arbitrary constituent locations, apart from the 
706: constraint on the ordering $y_1^{(1)}<y_2^{(1)}<y_1^{(2)}<y_2^{(2)}$, read
707: \beqa
708: \phi_{(1)}^{-1}(x)&=&1-\phi_\ff^{-1}=1-\prod_{i=1}^2\frac{r_1^{(i)}+r_2^{(i)}
709: -|y_1^{(i)}-y_2^{(i)}|}{r_1^{(i)}+r_2^{(i)}+|y_1^{(i)}-y_2^{(i)}|},
710: \label{eq:phis}\\ \phi_{(2)}^{-1}(x)&=&e^{i\gamma}N\frac{(y_1^{(2)}-y_1^{(1)})
711: (r_2^{(1)}-r_2^{(2)})+(y_2^{(2)}-y_2^{(1)})(r_1^{(1)}-r_1^{(2)})}{
712: \sum_{i,j=1}^2N_i^{(j)}r_i^{(j)}}\phi_{(1)}^{-1},\nonumber
713: \eeqa
714: where $r_i^{(j)}=|\vec x-\vec e y_i^{(j)}|$ and the constants $N,N_i^{(j)}$ 
715: are given by
716: \beq
717: N\equiv\sqrt{\frac{(y_2^{(2)}-y_1^{(1)})(y_1^{(2)}-y_2^{(1)})
718:                  }{(y_2^{(1)}-y_1^{(1)})(y_2^{(2)}-y_1^{(2)})}},\ 
719: N_i^{(j)}\equiv\frac{(y_{i'}^{(j)}-y_{i'}^{(j')})(y_{i'}^{(j)}-y_{i}^{(j')})}{
720:                      (y_{2}^{(j)}-y_{1}^{(j)})},\ j'\neq j,i'\neq i.
721: \eeq
722: The phase $\gamma$ vanishes when $\alpha_a^m=\xi_0^a=0$, but is of course 
723: irrelevant for checking $\hat\Psi^{(2)}$ to be a properly normalized fermion
724: zero-mode, orthogonal to $\hat\Psi^{(1)}$. 
725: In Fig.~\ref{fig:zmdeloc} we give an example for the behavior of these 
726: zero-mode densities. We choose $y_1^{(1)}=-6.031$, $y_2^{(1)}=-2.031$, 
727: $y_1^{(2)}=2.031$ and $y_2^{(2)}=6.031$. These are the constituent locations 
728: also found in Fig.~\ref{fig:zmlocal}, based on the axially symmetric solution 
729: with $\mu_2=1/4$, $\alpha_a=\xi_0=0$, $\xi=3.5$, $\Delta y_2=1$ and $\rho_1
730: =\rho_2=2$. Shown are the results for both zero-modes (bipole zero-mode on the 
731: right) at finite temperature, $\beta=1$ (bottom), and at infinite temperature 
732: (top). Note that these two only differ in the cores of the constituents, 
733: regular at finite temperature, but singular for the self-dual Dirac monopoles 
734: one is left with in the high temperature limit. The bipole zero-mode density 
735: is shown on a scale enhanced by a factor 5. Its reduced height is due to 
736: the fact that this zero-mode decays much slower than the other one, as 
737: can be read off from the behavior of $\phi_{(i)}^{-1}(x)$ in \refeq{phis}.
738: 
739: Crucial for the normalizability of both zero-modes is that $1/\phi_{(i)}$ is 
740: constant on the Dirac strings, where $1/\phi_\ff$ vanishes 
741: \beqa
742: \phi_{(1)}^{-1}(0,0,x_3)=1,&\phi_{(2)}^{-1}(0,0,x_3)=-e^{i\gamma}N\left(
743: \frac{y_2^{(2)}-y_1^{(2)}}{y_2^{(2)}-y_1^{(1)}}\right)\quad\mbox{for}\quad 
744: y_1^{(1)}\leq x_3\leq y_2^{(1)},\nonumber\\\phi_{(1)}^{-1}(0,0,x_3)=1,&
745: \phi_{(2)}^{-1}(0,0,x_3)=~~e^{i\gamma}N\left(\frac{y_2^{(1)}-y_1^{(1)}}{y_2^{
746: (2)}-y_1^{(1)}}\right)\quad\mbox{for}\quad y_1^{(2)}\leq x_3\leq y_2^{(2)}.
747: \eeqa
748: We are now in a position to answer the question what happens when violating 
749: \begin{figure}[htb]
750: \vskip6.7cm
751: \special{psfile=zm2secal.ps voffset=-125 hoffset=-100 vscale=60.0 hscale=60.0}
752: \special{psfile=zm2secl.ps voffset=-200 hoffset=-100 vscale=60.0 hscale=60.0}
753: \special{psfile=zm2bipal5.ps voffset=-125 hoffset=115 vscale=60.0 hscale=60.0}
754: \special{psfile=zm2bipl5.ps voffset=-200 hoffset=115 vscale=60.0 hscale=60.0}
755: \caption{The two zero-mode densities at $z=\mu_2=1/4$  (same configuration
756: as Fig.~\ref{fig:zmlocal}). The bipole zero-mode (right) is at 5 times the 
757: vertical scale of the second zero-mode (left). Top for the high temperature 
758: limit and bottom for finite temperature ($\beta=1$).}\label{fig:zmdeloc}
759: \vspace{4.7cm}
760: \special{psfile=zm2secwr.ps voffset=-190 hoffset=-100 vscale=60.0 hscale=60.0}
761: \special{psfile=zm2bipwr.ps voffset=-190 hoffset=115 vscale=60.0 hscale=60.0}
762: \caption{The two zero-mode densities for \refeq{phis} with the order of 
763: $y_1^{(2)}$ and $y_2^{(1)}$ interchanged. The second zero-mode (left) is 
764: at the same scale, whereas the bipole zero-mode (right) has its 
765: vertical scale magnified by a factor 10, compared to Fig.~\ref{fig:zmdeloc}.
766: }\label{fig:singzm}
767: \end{figure}
768: the constraint on the alternating order of oppositely charged constituents,
769: by smoothly deforming from $y_2^{(1)}<y_1^{(2)}$ to $y_2^{(1)}>y_1^{(2)}$.
770: Under this deformation, both $\hat\Psi^{(1)}$ and $\hat\Psi^{(2)}$ remain 
771: zero-mode solutions of the Dirac equation, and $A_\mu$ remains a self-dual 
772: abelian gauge field. We now have a Dirac string for $y_1^{(2)}<x_3<y_2^{(1)}$ 
773: of double the usual strength ($\phi_\ff^{-1}$ behaving as $x_\perp^4$ as 
774: opposed to $x_\perp^2$), where the second zero-mode density diverges, as 
775: illustrated in Fig.~\ref{fig:singzm}. This is because $\phi_{(2)}$ will no 
776: longer be constant on the double Dirac string. The bipole zero-mode, on the 
777: other hand, remains well defined. It actually vanishes identically on the 
778: double Dirac string (cmp. Fig.~1 of Ref.~\cite{Bip}), and no longer ``sees" 
779: the two inner self-dual Dirac monopoles.
780: 
781: It would be interesting if one could formulate an index theorem for these 
782: abelian field configurations with singularities, but this will not be 
783: straightforward as our analysis shows. It is yet another subtlety in 
784: describing the monopole content of non-abelian gauge fields. Developing 
785: a better understanding of these constraints, that affect the long range 
786: properties of configurations, is our main motivation for these studies. 
787: 
788: \section{Appearance of conserved quantities}\label{sec:cons}
789: 
790: Our analysis has shown that in all cases, as long as $z$ stays away from the 
791: impurities, the zero-mode density is exponentially localized to the cores of 
792: the constituents in the far field limit. The sizes of the monopole cores 
793: shrink to zero in the high temperature limit (with masses scaling proportional 
794: with the temperature), therefore we expect 
795: \beq
796: \cV_m(\vec x)\equiv(4\pi)^{-1}\Tr\left(R_m^{-1}(z)\right),
797: \label{eq:Rcons}
798: \eeq
799: cmp. Eqs.~\ref{eq:zdens} and \ref{eq:fzzffl}, to be harmonic almost everywhere
800: except for singularities tracing the cores of type $m$ monopoles for $z\in
801: ( \mu_m,\mu_{m+1})$. Since the caloron gauge field does not depend on $z$, 
802: this interpretation requires $\cV_m$ not to depend on $z$. The trace in 
803: \refeq{Rcons} is necessary to remove any $z$ dependence due to the fact that 
804: the basis of zero-modes is only defined up to a -- possibly $z$ dependent --
805: unitary transformation. All this is obviously true when $\vec R(z;\vec x)$ 
806: is piecewize constant, as for $k=1$ and for the class of axially symmetric 
807: solutions of Ref.~\cite{BrvB}. In this case the zero-mode density in the 
808: high temperature limit reduces to the sum of $k$ delta function, located 
809: at the appropriate constituent monopole locations.
810: 
811: To show that $\cV_m$ is independent of $z$, even when $\vec R(z,\vec x)$ is 
812: {\em not} piecewize constant we solve the Riccati equation, \refeq{Riccati},
813: iteratively in $1/|\vec x|$ and obtain the multipole expansion for $\cV_m
814: (\vec x)$. We can then use the Nahm equation for $\hat A_j(z)$ to check if 
815: every moment of this expansion is independent of $z$. We restrict our 
816: attention to the region between $\mu_m$ and $\mu_{m+1}$ and perform a 
817: rotation $U(\hat x)$ 
818: \beq
819: Y_i(z)\equiv U_{ij}(\hat x)\hat g(z)\frac{\hat A_j(z)}{2\pi i}\hat 
820: g^\dagger(z),\quad
821: U(\hat x)=\pmatrix{\sin(\theta)\cos(\vphi)&\sin(\theta)\sin(\vphi)&\hphantom{-}
822:   \cos(\theta)\cr\cos(\theta)\cos(\vphi)&\cos(\theta)\sin(\vphi)&-
823:   \sin(\theta)\cr-\sin(\vphi)&\cos(\vphi)&0\cr},\label{eq:defU}
824: \eeq
825: with $(\theta,\vphi)$ defined such that $U_{1j}=\hat x_j\equiv x_j/|\vec x|$
826: (the dependence of $\vec Y(z)$ on $\hat x$ will always be implicitly assumed).
827: This leaves rotations around $\hat x$ as a remaining freedom, which we will
828: make use of later. For the axially symmetric solutions discussed in 
829: Sect.~\ref{sec:bipole}, $\vec Y(z)=U(\hat x)\vec Y_m$, cmp. \refeq{Ymfull}. 
830: 
831: The Nahm equation, which is invariant under rotations, 
832: is equivalent to (working in the gauge where $\hat A_0$ is constant, removed 
833: by the gauge transformation with $\hat g(z)$)
834: \beq
835: \ddz Y_i(z)=-\pi i\veps_{ijk}[Y_j(z),Y_k(z)].\label{eq:snahm}
836: \eeq
837: We introduced $\vec Y(z)$ in \refeq{defU} such that 
838: \beq
839: \vec R(z;\vec x)^2=\ein_k|\vec x|^2-2|\vec x|Y_1(z)+\vec Y^{\,2}(z),
840: \eeq
841: has a simple form. Writing
842: \beq
843: R_m^\pm(z)^2\equiv\ein_k|\vec x|^2-|\vec x|Q_\pm(z;|\vec x|^{-1}),
844: \label{eq:wdef}
845: \eeq
846: we can now formulate the Riccati equation, \refeq{Riccati}, in terms of 
847: $\vec Y(z)$ and $Q_\pm(z;|\vec x|^{-1})$, 
848: \beq
849: \ein_k-\frac{Q_\pm(z;|\vec x|^{-1})}{|\vec x|}=\ein_k-2\frac{Y_1(z)}{|\vec x|}+
850: \frac{\vec Y^{\,2}(z)}{|\vec x|^2}\pm\frac{1}{2\pi|\vec x|}\ddz\sqrt{\ein_k-
851: \frac{Q_\pm(z;|\vec x|^{-1})}{|\vec x|}},\label{eq:QRic}
852: \eeq
853: which can be solved by iteration, expanding in powers of $1/|\vec x|$, 
854: something that is easily automated. We used the algebraic program 
855: FORM~\cite{FORM} for its superior memory management and speed to push
856: this calculation to a high order. We find for the first few terms (also
857: easily obtained by hand),
858: \beqa
859: R_m^\pm(z)/|\vec x|&=&\ein_k-Y_1/|\vec x|+\half
860: (Y_2^2+Y_3^2\pm i[Y_3,Y_2])/|\vec x|^2+\\&&\half(Y_2Y_1Y_2+Y_3Y_1Y_3\mp i
861: Y_2Y_1Y_3\pm iY_3Y_1Y_2)/|\vec x|^3+\ldots\label{eq:Qs}\nonumber
862: \eeqa
863: where the $z$ dependence of $Y_j(z)$ is suppressed for ease of notation
864: and any derivatives with respect to $z$ are eliminated with the help of the
865: Nahm equation, \refeq{snahm}. We substitute this expansion in \refeq{Rcons},
866: with $R_m(z)=\half(R_m^+(z)+R_m^-(z))$, from which we obtain its multipole 
867: expansion. The first few terms are 
868: \beq
869: \cV_m(\vec x)=\frac{1}{4\pi|\vec x|}\Tr\left(\ein_k+Y_1/|\vec x|+\half
870: (3Y_1^2-\vec Y^{\,2})/|\vec x|^2+\half Y_1(5Y_1^2-3\vec Y^{\,2})
871: /|\vec x|^3+\ldots\right).\label{eq:Rexp}
872: \eeq
873: 
874: A number of checks can be performed on this result. First of all $\cV_m$ 
875: and $Q_\pm$ are invariant under any rotation among $Y_2$ and $Y_3$,
876: $\cV_m$ is real and $Q_\pm$ are hermitian, as should be. By construction,
877: cmp. Eqs.~(\ref{eq:zdens}), (\ref{eq:fzzffl}) and (\ref{eq:Rcons}),
878: \beq
879: \sum_{a=1}^k\hat\Psi_z^a(x)^\dagger\hat\Psi_z^a(x)=-\partial_i^2\cV_m(\vec x),
880: \quad z\in(\mu_m,\mu_{m+1})\label{eq:Trzdens}
881: \eeq
882: in the far field limit, and one verifies that this integrates to $k$, 
883: as required by the proper normalization of the zero-modes. But most 
884: importantly, we verify with the help of the Nahm equation, \refeq{snahm}, that 
885: $\ddz\cV_m=0$ to the order given (whereas in general neither $R^\pm_m(z)$, 
886: nor $\Tr(R_m(z))$ are constant). For the dipole term this is immediate, 
887: since $\ddz\Tr Y_1=-2\pi i \Tr[Y_2,Y_3]=0$. For the quadrupole term we have 
888: $\ddz\Tr(3Y_1^2-\vec Y^{\,2})=-4\pi i\Tr(3Y_1[Y_2,Y_3]-\veps_{ijk}Y_iY_jY_k)
889: =0$, using the cyclic property of the trace, etc. Note that $\Tr\hat A_i(z)$ 
890: plays a special role. It corresponds to the $U(1)$ part of the $U(k)$ Nahm 
891: gauge field, and therefore decouples in the Nahm equations. As is obvious 
892: from the definition of $\vec R(z;\vec x)$, \refeq{Rdef}, it can actually 
893: be absorbed in a shift of $\vec x$. Therefore, where this simplifies matters 
894: we may assume $\vec Y(z)$ to be traceless. 
895: 
896: Finally we check, as conjectured above, that each of the terms is harmonic. 
897: For this we have to note that $\vec Y(z)$ depends on $\vec x$ through the 
898: rotation $U(\hat x)$, see \refeq{defU}. In the expression for $\cV_m(\vec x)$ 
899: the $\vec x$ dependence is easily recovered since $\vec Y^{\,2}$ is independent 
900: of $\vec x$ and $Y_1(z)=\hat g(z)\hat A_j(z)\hat g^\dagger(z)x_j/(2\pi i
901: |\vec x|)$. It is now straightforward to verify that each term is harmonic. 
902: When $k=1$ we may use that $Y_1=\hat x\cdot\vec y_m$ is no longer a matrix, 
903: and one indeed finds $\cV_m(\vec x)$ in \refeq{Rexp} to be the multipole 
904: expansion for $(4\pi|\vec x-\vec y_m|)^{-1}$ for that case. Since this is
905: harmonic (for $\vec x\neq \vec y_m$), each term in the multipole expansion of 
906: $\cV_m(\vec x)$ has to be harmonic. For arbitrary $k$ and $\vec Y(z)$ we have 
907: checked these properties to order $|\vec x|^{-14}$ in the multipole expansion. 
908: 
909: From now on we take charge 2. In this case there are 5 independent conserved 
910: quantities that characterize the solutions of the Nahm equation on a given 
911: interval $z\in[\mu_m,\mu_{m+1}]$, apart from the 3 translational degrees of 
912: freedom contained in $\Tr\hat A_i(z)$. They are given by the entries of the 
913: traceless and symmetric matrix $M$,
914: \beq
915: M_{ij}\equiv-\half\left(\Tr(\hat A_i(z)\hat A_j(z))-\third
916: \delta_{ij}\Tr(\hat A_k(z)\hat A_k(z))\right).\label{eq:defM}
917: \eeq
918: One easily checks with the Nahm equation that this is conserved, as for 
919: the quadrupole term considered above. Although not needed here, it is 
920: known~\cite{NCal} that for any $k$ a solution to the Nahm equation implies 
921: $\det(y_j\hat A_j(z)-2\pi iy_jx_j)$ is constant, provided $\vec y\in C^3$, 
922: with $y_j^2=0$. Indeed, for $k=2$ and $\vec x=\vec 0$, using that 
923: $\det(y_j\hat A_j(z))=-\half\Tr(y_j\hat A_j(z))^2$, this is equivalent to 
924: $M$ being constant if and only if $y_j^2=0$. Since $M$ gives all the 
925: independent invariants, it should be possible to express $\cV_m(\vec x)$ 
926: in terms of $M$ and $\vec x$. Considerable simplifications occur, because 
927: we can write 
928: \beq
929: \hat A_j(z)\equiv i\cA_{ja}(z)\hat g^\dagger(z)\tau_a\hat g(z)\label{eq:Aia}
930: \eeq
931: (absorbing the trace part in a shift of $\vec x$), and use $\tau_a\tau_b=
932: \delta_{ab}\ein_2+i\veps_{abc}\tau_c$ to reduce the matrix products to scalar 
933: products, $\cA_{ia}(z)\cA_{ja}(z)=M_{ij}+\third\delta_{ij}\cA^2_{ka}(z)$.
934: We do indeed find that $\cV_m(\vec x)$ is a function of $M$ and $\vec x$ only, 
935: \beq
936: \cV_m(\vec x)=\frac{1}{2\pi|\vec x|}\left(1+\frac{3}{2|\vec x|^2}\hat M_{11}
937: (\hat x)+\cO(|\vec x|^{-4})\right).
938: \eeq
939: where for convenience we introduced 
940: \beq
941: \hat M_{ij}(\hat x)=\frac{1}{4\pi^2}\left(U(\hat x)MU^{-1}(\hat x)\right)_{ij}
942: =\half\Tr\left(Y_i(z)Y_j(z)-\third\delta_{ij}\vec Y^{\,2}(z)\right).
943: \label{eq:defMhat}
944: \eeq
945: We performed the multipole expansion for charge 2 to order $|\vec x|^{-21}$. 
946: Only the odd orders appear, because $\vec Y(z)$ is assumed to be traceless.
947: In appendix B we give the term of order 21, and show how from this all lower 
948: order multipole coefficients can be recovered.
949: 
950: \section{Exact results for charge 2}\label{sec:chtwo}
951: 
952: In this section we will construct $SU(2)$ charge 2 solutions for which $\vec 
953: R(z;\vec x)$ is {\em not} piecewize constant and analyse the localization 
954: of the fermion zero-modes in the far field limit. We already saw in 
955: Sect.~\ref{sec:cons} that on each interval $(\mu_m,\mu_{m+1})$ we have
956: information on the zero-mode density, \refeq{Trzdens}, in terms of 8
957: parameters. Three of these are associated with $\Tr\vec Y(z)$, which give the 
958: center of mass coordinates for the constituents of type $m$. Of the other 5, 
959: given by the $3\times 3$  traceless symmetric matrix $M$, 3 are associated to
960: a rotation $\cR$ that diagonalizes $M$, whereas the remaining 2 parameterize 
961: the eigenvalues of $M$. They will give a scale ($D$) and shape ($\k$) 
962: parameter, see below.
963: 
964: \subsection{Solutions to the Nahm equation}\label{sec:k0}
965: Explicit solutions in terms of Jacobi elliptic functions were first considered 
966: in the context of $SU(2)$ charge 2 monopoles~\cite{NahmM,BrDa}. 
967: These solutions can be adopted for the calorons provided the appropriate 
968: boundary conditions, read off from \refeq{nahm}, are implemented. In terms 
969: of the $3\times 3$ matrix $\cA_{ia}(z)$ defined in \refeq{Aia}, the Nahm 
970: equation becomes (away from $z=\mu_m$) 
971: \beq
972: \frac{1}{2}\ddz\cA^t(z)=\det(\cA(z))\cA^{-1}(z),
973: \eeq
974: from which we find
975: \beq
976: \frac{1}{4}\ddz\left(\cA(z)\cA^t(z)\right)=\ein_3\det\cA(z)=\ein_3\sqrt{\det
977: \left(\cA(z)\cA^t(z)\right)}.\label{eq:AA}
978: \eeq
979: The traceless part of $\cA(z)\cA^t(z)$ is therefore independent of $z$, once 
980: again verifying that $M=\cA(z)\cA^t(z)-\third\ein_3\Tr(\cA(z)\cA^t(z))$ is 
981: constant. Here we are, however, interested in the equation for the trace part 
982: \beq
983: \ddz F(z)=4\sqrt{\det\left(F(z)\ein_3+M\right)},\quad
984: F(z)\equiv\third\Tr\left(\cA(z)\cA^t(z)\right).
985: \eeq
986: When we diagonalize $M$ with a suitable rotation $\cR$, $M=\cR\diag(c_1,c_2,
987: c_3)\cR^t$, fixing $\cR$ such that $c_2\leq c_1\leq c_3$ (note that in 
988: addition $c_1+c_2+c_3=\Tr M=0$), this can be cast in the form
989: \beq
990: \ddz F(z)=4\sqrt{(F(z)+c_1)(F(z)+c_2)(F(z)+c_3)}.
991: \eeq
992: Defining $F(z)+c_i=\quart D^2 f_i^2(Dz)$ and introducing $D$ and 
993: $\k$ to parameterize the $c_i$,
994: \beq
995: D^2\equiv\quart(c_3-c_2),\quad \k^2\equiv\frac{c_3-c_1}{c_3-c_2}\leq 1,
996: \label{eq:Dkdef}
997: \eeq
998: shows that the solution can be written in terms of the Jacobi elliptic 
999: functions\footnote{The Jacobi elliptic functions are defined by $sn_\k(u)=
1000: \sin(\vphi(u))$, $cn_\k(u)=\cos(\vphi(u))$ and $dn_\k(u)=\sqrt{1-\k^2sn^2_\k
1001: (u)}$, with $u=\int_0^{\vphi(u)}d\theta(1-\k^2\sin^2\theta)^{-\hhalf}$. We 
1002: use boldface for $\k$ to avoid confusion with charge $k$. One also encounters 
1003: the notation~\cite{AbSt} $sn(u|m)$ for $sn_\k(u)$, with $m=\k^2$.} 
1004: \beq
1005: f_1(z)=\frac{\k'}{cn_\k(z)},\quad f_2(z)=\frac{\k'sn_\k(z)}{cn_\k(z)},
1006: \quad f_3(z)=\frac{dn_\k(z)}{cn_\k(z)},\quad\k'\equiv\sqrt{1-\k^2}.
1007: \eeq
1008: The overall sign of the functions $f_i(z)$ is chosen such that $df_1(z)/dz=
1009: f_2(z)f_3(z)$, and cyclic, such that in terms of these the most general 
1010: solution of the Nahm equation is given by
1011: \beq
1012: \hat A_j(z;\vec a,\cR,h,\cD,\k)\equiv 2\pi i\hat g^\dagger(z)h^\dagger
1013: \Bigl(a_j\ein_2+\cD\cR_{jb}f_{b}(Dz)\tau_b\Bigr)h\hat g(z),
1014: \quad\cD\equiv(4\pi)^{-1}D,\label{eq:fnahm}
1015: \eeq
1016: where $\cR$ is the rotation that diagonalizes $M$ and $h$ is a global gauge 
1017: rotation (leaving $\cA(z)\cA^t(z)$ invariant). In the $Sp(1)$ formalism for 
1018: constructing $SU(2)$ calorons one requires $\hat A^t_j(z)=\hat A_j(-z)$, a 
1019: property shared by $f_b(Dz)\tau_b$ for each $b$. Arranging $f_2(z)$ to be 
1020: odd and $f_{1,3}(z)$ to be even in $z$ was the reason for choosing 
1021: $c_2\leq c_1\leq c_3$. 
1022: 
1023: We see that $\k'=0$ (i.e. $\k=1$) recovers the case where $\vec R(z;\vec x)$ 
1024: is piecewize constant, for which $\cV^{\k=1}_m(\vec x)=(4\pi|\vec x-\vec y|
1025: )^{-1}+(4\pi|\vec x+\vec y|)^{-1}$, with $\vec y=(0,0,\cD)$ and $\pm\vec y$ 
1026: the two locations of the equal charge constituents (the center of mass assumed 
1027: to be zero), see Sect.~\ref{sec:cons}. The combined zero-mode density in the 
1028: far field limit is given by $-\partial_i^2\cV_m(\vec x)=\delta(\vec x-\vec y)
1029: +\delta(\vec x+\vec y)$, the sum of two delta-functions at these constituent 
1030: locations. Actually, two point-like constituents necessarily implies $\k=1$. 
1031: Comparing $3\hat M_{11}/(16\pi^3|\vec x|^3)=3\sum_{j=1}^3c_jx_j^2/(16\pi^3|
1032: \vec x|^5)$, see \refeq{defMhat}, with the quadrupole term for $\cV^{\k=1}_m
1033: (\vec x)$, $(2x_3^2-x_2^2-x_1^2)\cD^2/(4\pi|\vec x|^5)$, one finds that 
1034: $c_1=c_2$, which forces $\k=1$. It is in this context that we define 
1035: $\vec y=\pm(0,0,\cD)$ as would-be constituent locations even when $\k\neq 1$.
1036: In which way we will approach point-like constituents when $\k\to1$ will 
1037: become clear in Sect.~\ref{sec:ext}. So far we have studied the Nahm equation 
1038: away from the ``impurities". We first want to convince ourselves that there 
1039: are more general (not piecewize constant) solutions to the Nahm equation 
1040: that describe calorons with non-trivial holonomy, for which we need to solve 
1041: the boundary conditions at the ``impurities".
1042: 
1043: \subsection{Matching at the impurities}
1044: We will consider here the case of $SU(2)$ charge 2 calorons in the symplectic 
1045: representation, for which $\hat A_j(-z)=\hat A^t_j(z)$. This condition is 
1046: preserved under the gauge transformation $\hat g(z)$ defined in 
1047: \refeq{dftrans}, but requires us to further constrain $h$ appearing in 
1048: \refeq{fnahm} to be generated by $\tau_2$. It means that for $z\in[\mu_1,
1049: \mu_2]$ the Nahm gauge field is given by $\hat A_j(z)=\hat A_j(z;\vec a^{(1)},
1050: \cR^{(1)},e^{-\frac{i}{2}\theta^{(1)}\tau_2}, \cD^{(1)},\k^{(1)})$. For 
1051: $z\in[\mu_2,1+\mu_1]$ we use that periodicity of $\hat A_j(z)$ implies 
1052: $\hat A_j(1-z)=\hat A_j(-z)=\hat A_j^t(z)$, such that $\hat A_j(z)=
1053: \hat A_j(z-\half;\vec a^{(2)},\cR^{(2)},e^{-\frac{i}{2}\theta^{(2)}
1054: \tau_2},\cD^{(2)},\k^{(2)})$. This was studied before by Houghton and Kraan 
1055: for trivial holonomy ($\mu_2=0$)~\cite{HoKr}, where one of the monopole types 
1056: is massless. For general holonomy, the Nahm equation reduces at $z=\mu_2$ to 
1057: (see \refeq{nahm})
1058: \beq
1059: \hat A_j(\mu_2-\half;\vec a^{(2)},\cR^{(2)},e^{-\frac{i}{2}\theta^{(2)}\tau_2},
1060: \cD^{(2)},\k^{(2)})-\hat A_j(\mu_2;\vec a^{(1)},\cR^{(1)},e^{-\frac{i}{2}
1061: \theta^{(1)}\tau_2},\cD^{(1)},\k^{(1)})=2\pi i\rho_2^j.\label{eq:jump}
1062: \eeq
1063: At $z=\mu_1=-\mu_2$, using $\hat A_j(-z)=\hat A_j(z)^t$, the same condition 
1064: is found (one can check~\cite{BrvB} that $\rho_1=-\rho_2^t$). 
1065: 
1066: \begin{figure}[htb]
1067: \vspace{4.5cm}
1068: \special{psfile=rectangle.ps voffset=-330 hoffset=-90 vscale=100 hscale=100}
1069: \caption{Plotting $|\Delta\vec a|$ versus $D$, \refeq{bcI}, for $\k=0,$ 
1070: 0.9, 0.99, 0.999 and 0.9999 (left to right).}\label{fig:kDa-plot} 
1071: \end{figure}
1072: 
1073: A particularly simple solution is obtained for $\mu_2=1/4$ (equal mass 
1074: constituents) by taking the same parameters in both intervals, up to a 
1075: shift $\vec a$, $\cD^{(m)}=\cD$, $\k^{(m)}=\k$, $\cR^{(m)}=\ein_3$ and 
1076: $\theta^{(m)}=0$, collapsing \refeq{jump} to
1077: \beq
1078: \vec\rho_2=\Delta\vec a\ein_2-2(0,1,0)\tau_2\cD\k'\frac{sn_\k(\quart D)}{
1079: cn_\k(\quart D)},\quad\Delta\vec a\equiv\vec a^{(2)}-\vec a^{(1)}
1080: \eeq
1081: This can be solved by taking $P_2=\half(\ein_2+\tau_2)$, $\zeta_a=\rho
1082: \exp(2\pi i\alpha_a\tau_2)$, with $\alpha_1-\alpha_2=1/4$, which
1083: gives $\vec\rho_2^{\,ab}=-2\pi\rho^2P_2^{ab}(0,1,0)$, such that
1084: \beq
1085: \frac{D\k'sn_\k(\quart D)}{2\pi cn_\k(\quart D)}(0,1,0)=
1086: -\Delta\vec a=\pi\rho^2(0,1,0).\label{eq:bcI}
1087: \eeq
1088: For increasing $D$, $cn_\k(\quart D)$ will reach zero at $\quart D=K(\k)$,
1089: the half-period\footnote{$K(\k)=\int_0^{\pi/2}d\theta(1-\k^2\sin^2\theta
1090: )^{-\hhalf}$, satisfying $K(\k)=-\log(\quart\k')(1+\cO(\k'^2))$~\cite{AbSt}}. 
1091: To keep $|\Delta\vec a|$ finite, $\k'$ has to approach zero as well. 
1092: 
1093: In Fig.~\ref{fig:kDa-plot} we plot $|\Delta\vec a|$ as a function of $\k$ 
1094: and $D$, from which we confirm that, at fixed $|\Delta\vec a|$, $\k$ has 
1095: to approach 1 to have well separated constituents. In this point-like limit, 
1096: monopole constituents of type 1 are located at $(0,-\half\pi \rho^2,\pm\cD)$ 
1097: and those of opposite charge (type 2) at $(0,\half\pi\rho^2,\pm\cD)$ (choosing 
1098: the overall center of mass at the origin). This configuration has parallel 
1099: magnetic moments and placing self-dual Dirac monopoles at the mentioned 
1100: locations would give a gauge field in the bipole ansatz, \refeq{bipans}. 
1101: However, the abelian component of the gauge field coming from the above caloron
1102: can at best take the bipole form in the limit discussed. For $\k=1$ and $\cD$ 
1103: finite, $|\Delta\vec a|=0$, and one is left with two singular instantons (of 
1104: zero size) at $(0,0,\pm\cD)$. 
1105: 
1106: For this reason we now consider a class of solutions which contains 
1107: {\em regular} axially symmetric solutions with $\k=1$. Again we take 
1108: $\mu_2=1/4$, $\cD^{(m)}=\cD$, $\k^{(m)}=\k$, $\zeta_a=\rho\exp(2\pi 
1109: i\alpha_a\tau_2)$, but now with $\alpha_2=-\alpha_1\equiv\pi^{-1}\alpha$, 
1110: $P_2=\half(1+\tau_3)$, $\theta^{(2)}=-\theta^{(1)}\equiv\theta$ and 
1111: \beq
1112: \cR^{(2)}=\left(\cR^{(1)}\right)^{-1}=\pmatrix{\cos\vphi&0&\sin
1113: \vphi\cr0&1&0\cr-\sin\vphi&0&\cos\vphi}.
1114: \eeq
1115: One finds 
1116: \beq
1117: \vec\rho_2^{\,ab}=-\pi\rho^2(-\sin\alpha\tau_3,-\sin\alpha\tau_2,
1118: \tau_1+\cos\alpha\ein_1)^{ab},
1119: \eeq
1120: and \refeq{jump} takes the following form
1121: \beqa
1122: 2\cD(f_1(\quart D)\sin\theta\cos\vphi+f_3(\quart D)\cos\theta\sin\vphi)
1123: \tau_3+\Delta a_1\ein_2&=&\pi\rho^2\sin\alpha\tau_3\nonumber\\
1124: -2\cD f_2(\quart D)\tau_2+\Delta a_2\ein_2&=&\pi\rho^2\sin\alpha\tau_2\\
1125: -2\cD(f_1(\quart D)\cos\theta\sin\vphi+f_3(\quart D)\sin\theta\cos\vphi)
1126: \tau_1+\Delta a_3\ein_2&=&-\pi\rho^2(\tau_1+\cos\alpha\ein_2).\nonumber
1127: \eeqa
1128: This can be simplified to
1129: \beqa
1130: \cD\sin(\theta-\vphi)\Bigl(f_3(\quart D)-f_1(\quart D)\Bigr)=\half\pi\rho^2
1131: (1-\sin\alpha),&\Delta\vec a=-\pi\rho^2\cos\alpha(0,0,1),\label{eq:sol}\\
1132: \cD\sin(\theta+\vphi)\Bigl(f_3(\quart D)+f_1(\quart D)\Bigr)=\half\pi\rho^2
1133: (1+\sin\alpha),&\cD f_2(\quart D)=-\half\pi\rho^2\sin\alpha,\nonumber
1134: \eeqa
1135: It gives a three parameter family of solutions with would-be point-like 
1136: constituents at
1137: \beq
1138: \vec y_m^{\,(j)}=((-1)^{j}\cD\sin\vphi,0,(-1)^{m+j}\cD\cos\vphi-(-1)^m
1139: \half\rho^2\cos\alpha).\label{eq:conloc}
1140: \eeq
1141: 
1142: To have an exact point-like far field limit we need to impose $\k=1$, 
1143: implying $\sin\alpha=0$ and $\cos\theta\sin\vphi=0$. The first possibility 
1144: is that $\cos\theta=0$, for which $|\cos\vphi|\cD=\half\rho^2$. 
1145: One finds two constituents of opposite charge to coincide. Such a solution 
1146: describes a singular (zero-size) instanton on top of a smooth caloron. 
1147: Excluding this singular case we are left with the choice $\sin\vphi=0$, 
1148: for which $|\sin\theta|\cD=\half\rho^2$. We now find axially symmetric 
1149: solutions with constituent locations at
1150: \beq
1151: \vec y_m^{\,(j)}=\mp\half\pi\rho^2\left((-1)^m+(-1)^j|\sin\theta|^{-1}
1152: \right)(0,0,1), 
1153: \eeq
1154: where the overall sign comes from the fact that $\cos\alpha=\pm1$. For $\cos
1155: \theta\neq0$, all constituents are now separated from each other, giving a 
1156: regular solution. Both cases were already studied in Ref.~\cite{BrvB}, based 
1157: on assuming that $\vec\rho_2$ is one dimensional (cmp. Sect.~\ref{sec:bipole}). 
1158: It can be shown that for $SU(2)$ {\em exact} point-like constituents, 
1159: $\k^{(1)}=\k^{(2)}=1$, forces $\vec\rho_2$ to be one dimensional for 
1160: {\em any} choice of the remaining parameters. 
1161: 
1162: \begin{figure}[htb]
1163: \vspace{6.5cm}
1164: \special{psfile=kofrho.ps voffset=-310 hoffset=-180 vscale=100 hscale=100}
1165: \special{psfile=butterfly.ps voffset=-310 hoffset=40 vscale=100 hscale=100}
1166: \caption{On the left we plot $\k$ versus $\rho$ for $\alpha=0$ ($\k\equiv1$), 
1167: $\alpha=-\pi/100$ and $\alpha=-\pi/2$ (giving the lower bound for $\k$ at 
1168: fixed $\rho$). On the right are shown the locations, \refeq{conloc}, of 
1169: monopoles and antimonopoles (fat vs. thin curves) in the $1,3$-plane for 
1170: $\rho=1/4$, by varying $\alpha$ from $-\pi$ (indicated by the arrows) to 0.
1171: }\label{fig:conloc}
1172: \end{figure}
1173: 
1174: Nevertheless, when $\sin\alpha\neq0$, insisting as before that equal charge 
1175: constituents are well separated, while keeping the centers of mass of these 
1176: pairs at a fixed distance $\pi\rho^2\cos\alpha$, one forces $\k\to1$ while 
1177: increasing $\cD$, and hence approximate point-like constituents. We will 
1178: illustrate this behavior for $\theta=\pi/4$. In Fig.~\ref{fig:conloc} 
1179: we plot for a typical value of $\rho$ the constituent locations as given 
1180: by \refeq{conloc}, varying $\alpha$ between $-\pi$ and 0 (given $\rho$, 
1181: $\alpha$ and $\theta$, one can use \refeq{sol} to solve for $\vphi$, $\cD$ 
1182: and $\k$). We also plot $\k$ as a function of $\rho$ for some values of 
1183: $\vphi$, showing the rapid uniform approach to $\k=1$. The asymptotic 
1184: behavior for $\alpha=-\pi/2$ is determined by 
1185: \beq
1186: \k'=\frac{4\exp(-D/4)}{3+2\sqrt{2}}\left(1+\cO(\k'^2)\right),
1187: \quad D=4\sqrt{2}\pi^2\rho^2\left(1+\cO(\k'^2)\right).\label{eq:kpasym}
1188: \eeq
1189: 
1190: It is also interesting to inspect $\hat A_i(z)$ in the limit $\k\to1$ (or 
1191: $\cD\to\infty$), to understand to which extent we retrieve the piecewize 
1192: constant behavior of $\vec R(z;\vec x)$, on which the point-like limit is 
1193: based. For this we plot $f_i(D(\k)z)$ in Fig.~\ref{fig:plfs}, which apart 
1194: from fixed rotations and an overall factor $\cD$ would represent the 
1195: constituent locations. Since $\cD(\k)/(\pi\rho^2)$ approaches $\sqrt{2}$ for 
1196: $\k\to1$, it means we normalize the constituent locations with respect to 
1197: height of the jumps in the Nahm data. At the impurities ($z=\pm1/4$) we 
1198: therefore expect $f_i(D(\k)z)$ to go to a fixed value. The plotted cases,
1199: $1-\k=10^{-4}$, $D(\k)=15.53$ (left) and $1-\k=10^{-8}$, $D(\k)=33.95$ 
1200: (right), clearly demonstrate how in the bulk $f_3\to 1$ and $f_{1,2}\to0$, 
1201: but that they differ at $z=\pm1/4$, to accommodate the discontinuities 
1202: of the Nahm equation at the impurities. The cross-over from bulk behavior 
1203: to the impurity values scales as $D^{-1}$, and is only absent for axially 
1204: symmetric solutions.
1205: 
1206: \begin{figure}[htb]
1207: \vskip11mm\hskip73mm$f_3$\vskip16mm\hskip73mm$f_1$\vskip13mm\hskip73mm$f_2$
1208: \vskip2mm
1209: \special{psfile=plfs4-9s.ps voffset=-35 hoffset=5 vscale=70.0 hscale=70.0}
1210: \special{psfile=plfs8-9s.ps voffset=-35 hoffset=220 vscale=70.0 hscale=70.0}
1211: \vskip2mm
1212: \caption{Plots of $f_j(D(\k)z)$ for $z\in[-\quart,\quart]$ at $\k=1-10^{-4}$
1213: (left) and $\k=1-10^{-8}$ (right), illustrating the approach to the point-like
1214: limit $\k\to1$.}\label{fig:plfs}
1215: \end{figure}
1216: 
1217: \subsection{Extended structure}\label{sec:ext}
1218: When $\vec R(z;\vec x)$ is not piecewize constant, it is not clear what plays
1219: the role of the constituent locations. We therefore do expect some extended 
1220: structure for $\k\neq1$. For this reason we now come back to analysing 
1221: $\cV_m(\vec x)=(4\pi)^{-1}\Tr R_m^{-1}(z)$ in more detail. Quite remarkably, 
1222: the discussion in Sect.~\ref{sec:cons} implies that the result for $\k=0$ 
1223: can be obtained from that at $\k=1$. For this we may use the symmetry 
1224: $c_2\leftrightarrow c_3$, which leaves $\cV_m(\vec x)$ invariant, apart from 
1225: interchanging $x_2$ and $x_3$ ($\hat M$ is left unchanged when absorbing the 
1226: rotation $\cR$ that interchanges $c_2$ and $c_3$ in $U$, see \refeq{defMhat}). 
1227: With the definitions of $D=4\pi\cD$ and $\k$ in \refeq{Dkdef} one finds that 
1228: this implies $\cD^2\to-\cD^2$ and $\k\to\k'$. Therefore,
1229: \beq
1230: \k=0:\quad\cV_m(\vec x)=\frac{\sqrt{2}\sqrt{|\vec x|^2-\cD^2+\sqrt{(|\vec 
1231: x|^2-\cD^2)^2+4\cD^2x_2^2}}}{4\pi\sqrt{(|\vec x|^2-\cD^2)^2+4\cD^2x_2^2}}.
1232: \label{eq:k0p}
1233: \eeq
1234: For this we use that the $\k=1$ result, $\cV_m(\vec x)=(4\pi|\vec x-\vec y|
1235: )^{-1}+(4\pi|\vec x+\vec y|)^{-1}$, with $\vec y=\cD(0,0,1)$, can be rewritten 
1236: in the form of \refeq{k0p} by changing the sign of $\cD^2$ and interchanging 
1237: $x_2$ and $x_3$. In some sense we may say that the result for $\k=0$ describes 
1238: point-like constituents that have moved into a complex direction\footnote{This 
1239: was observed before in the monopole context~\cite{NahmM}. It is also 
1240: worthwhile to point out that the axially symmetric monopole solutions 
1241: discussed there can appear as such in the caloron context, as we read off 
1242: for $\k=0$ from \refeq{bcI}. In this case we expect to find another class of 
1243: axially symmetric caloron solutions, with $\Delta\vec a$ giving the symmetry
1244: axes, but since we are more interested here in the case of well-separated 
1245: monopole constituents, we did not analyse this in further detail.}. We note 
1246: that \refeq{k0p} is singular on the ring $x_2=0$, $|\vec x|^2=\cD^2$, i.e. on 
1247: the circle of radius $|\cD|$ in the $1,3$-plane. But there is more of a 
1248: surprise: the $x_2$ derivative is discontinuous on the disk bounded by the 
1249: ring (away from the disc the function is smooth and harmonic), $\cV_m(\vec x)
1250: =(2\pi)^{-1}|\cD x_2|/(\cD^2-r^2)^{3/2}(1+\cO(x_2^2))$, with $r^2\equiv x_1^2
1251: +x_3^2$. This implies indeed an extended structure, with singularities on the 
1252: entire disk. Before showing how to deal with this singularity structure we 
1253: consider general values of $\k$.
1254: 
1255: As we have seen in Sect.~\ref{sec:cons}, $\Tr R_m^{-1}(z)$ is conserved 
1256: as a consequence of the Nahm equation. Using for $R_m^\pm(z)$ the Riccati 
1257: equation, \refeq{Riccati}, the fact that $\Tr R_m^{-1}(z)=2\Tr(R_m^+(z)
1258: +R_m^-(z))^{-1}$ is independent of $z$ imposes a severe constraint. 
1259: We make use of this by expanding $R_m^\pm(z)$ as a Taylor series in $z$. 
1260: The Taylor coefficients can be expressed in terms of the initial conditions 
1261: $R_m^\pm(0)$, through the explicit solution of $\hat A_j(z)$, \refeq{fnahm} 
1262: (we make use of the invariance under translations and rotations to put 
1263: $\vec a=\vec 0$, $h=\ein_2$ and $\cR=\ein_3$, as in Sect.~\ref{sec:cons}).  
1264: We thus obtain the Taylor series for $\Tr R_m^{-1}(z)$, of which all 
1265: coefficients should vanish except for the 0th order. This gives a set 
1266: of {\em algebraic} equations for the initial conditions, $R_m^\pm(0)$, 
1267: encoded in $X_\mu$ and $\tilde X_\mu$ through
1268: \beq
1269: R_m(0)=\half(R_m^+(0)+R_m^-(0))\equiv\cD(X_0\ein_2+X_j\tau_j),\quad
1270: \half(R_m^+(0)-R_m^-(0))\equiv\cD(\tilde X_0\ein_2+\tilde X_j\tau_j),
1271: \eeq
1272: Note that $\cD$ enters as an overall scale factor, cmp. \refeq{fnahm}, 
1273: such that
1274: \beq
1275: \cV_m(\vec x)=\frac{1}{4\pi}\Tr R_m^{-1}(0)=\frac{1}{2\pi\cD}\tilde
1276: \cV(\cD\vec x),\quad\tilde\cV=\half\Tr\left(\frac{1}{X_0+X_j\tau_j}\right)=
1277: \frac{X_0}{X_0^2-X_j^2}.
1278: \eeq
1279: Dependence on the coordinates and on $\k$ is mostly left implicit. The system 
1280: of equations is of course hugely overdetermined, and some amount of good 
1281: fortune was required in that the first 11 orders in the Taylor expansion were 
1282: sufficient to solve for $X_\mu$ and $\tilde X_\mu$. Quite remarkably these 
1283: imply that $\tilde X_0=\tilde X_1=\tilde X_3=X_2=0$, which considerably 
1284: simplifies the task of solving for the remaining 4 variables. We found that 
1285: \beq
1286: \delta\equiv X_0^2-X_1^2-X_3^2-\tilde X_2^2
1287: \eeq
1288: satisfies the cubic equation
1289: \beqa
1290: &&\hskip-9mm(2-\k^2)\k^4+4\k^2(x_1^2-x_3^2)-\k^4(3x_1^2-x_2^2-x_3^2)+
1291: \left((2-\k^2)(3x_1^2-x_2^2+x_3^2)-4x_1^2\right)|\vec x|^2\nonumber\\ 
1292: &&\hskip-5mm -|\vec x|^6-\left(\k^4+2\k^2(x_1^2-x_2^2-x_3^2)+4x_2^2+
1293: |\vec x|^4\right)\delta+\left(\k^2-2+|\vec x|^2\right)\delta^2+\delta^3=0,
1294: \eeqa
1295: whereas $X_1/X_0$, $X_3/X_0$ and $\tilde X_2$ can be solved for in terms of 
1296: $\delta$, 
1297: \beqa
1298: X_1/X_0&=&\frac{|\vec x|^4-\delta^2+2(\k'^2(\delta-2x_1^2)+x_1^2+x_2^2-x_3^2)+
1299:          (2-\k^2)\k^2}{4x_1\k'\k^2}\nonumber\\
1300: X_3/X_0&=&\frac{|\vec x|^4-\delta^2+2(\delta-2x_3^2-\k'^2(x_1^2-x_2^2-x_3^2))-
1301:          (2-\k^2)\k^2}{4x_3\k^2}\nonumber\\
1302: \tilde X_2&=&\frac{|\vec x|^4-\delta^2+2\k^2(x_1^2-x_2^2-x_3^2)+4x_2^2+
1303:           \k^4}{4x_2\k'}.
1304: \eeqa
1305: Therefore, $\tilde\cV^2$ is a rational function of $\delta$, $\vec x$ 
1306: and $k$,
1307: \beq
1308: \tilde\cV^2=\frac{1}{(\delta+\tilde X_2^2)(1-X_1^2X_0^{-2}-X_3^2X_0^{-2})},
1309: \eeq
1310: and the proper root of the cubic equation for $\delta$ to use is fixed by 
1311: $\tilde\cV\to1/|\vec x|$. Indeed, the asymptotic expansion for $\delta$,
1312: \beq
1313: \delta=|\vec x|^2\left(1+\frac{2x_2^2 +\k^2(x_1^2-x_2^2-x_3^2)}{|\vec x|^4}
1314: +\ldots\right),
1315: \eeq
1316: reproduces the multipole expansion of $\tilde\cV$. This was verified to the 21 
1317: orders given in \refeq{m20}. On general grounds it can be argued, since 
1318: $\delta$ satisfies a cubic equation, that $\tilde\cV^2$ has to satisfy a 
1319: cubic equation as well. Its coefficients (polynomials in $\k$ and $\vec x$) 
1320: are somewhat lengthy, and therefore not reproduced here, in part because
1321: we will shortly present an exact integral equation for $\cV_m(\vec x)$, valid
1322: for any $\k$. The exact results for $\k=0$ and $\k=1$ is most easily recovered
1323: from the cubic equation for $\tilde \cV^2$, but the integral representation
1324: will be valid for these two cases as well.
1325: 
1326: Like for $\k=0$, we find that $\cV_m(\vec x)$ is harmonic everywhere except 
1327: on a disk, now bounded by an ellipse with major and minor axes $\cD$, resp. 
1328: $\k'\cD$. On this disk the function vanishes, satisfying in the direction 
1329: perpendicular to the disk the expansion $\cV_m(\vec x)=(2\pi)^{-1}\k'|\cD 
1330: x_2| /(\cD^2\k'^2-x_1^2-\k'^2x_3^2)^{3/2}(1+\cO(x_2^2))$ (cmp. the 
1331: discussion above, for $\k=0$). By introducing the 
1332: ``polar" coordinates $(x_1,x_3)=(\k'r\cos\vphi,r\sin\vphi)$, in terms of which 
1333: the ellipse at the boundary of the disk is characterized by $r=\cD$, this can 
1334: be written as $\cV_m(\vec x)=(2\pi\k')^{-1}|\cD x_2|/(\cD^2-r^2)^{3/2}(1+
1335: \cO(x_2^2))$.  Care is required to deal correctly with the behavior at the 
1336: edge of the disk when computing the Laplacian. We find
1337: \beq
1338: -\partial_i^2\cV_m(\vec x)=-\delta(x_2)\frac{\cD}{\pi r\k'}\frac{\partial}{
1339: \partial r}\frac{\theta(\cD-r)}{\sqrt{\cD^2-r^2}},\label{eq:chdist}
1340: \eeq
1341: with $\theta(\cD-r)$ the step function, giving the following integral 
1342: representation 
1343: \beq
1344: \cV_m(\vec x)=\frac{1}{2\pi|\vec x|}+\frac{\cD}{4\pi^2}\int_0^{2\pi}
1345: \!\!\!\!\!d\vphi\int_0^\cD\!\!\!\frac{dr}{\sqrt{\cD^2-r^2}}\partial_r
1346: \frac{1}{\sqrt{(x_1-\k'r\cos\vphi)^2+(x_3-r\sin\vphi)^2+x_2^2}}.\label{eq:res}
1347: \eeq
1348: Note that $\k'$, appearing in the denominator of \refeq{chdist}, cancels due to
1349: the change of variables to ``polar" coordinates. By numerical evaluation, we 
1350: checked this formula against the exact results. It gives us confidence that 
1351: we interpreted the singularity structure correctly. Taking an arbitrary test 
1352: function $f(\vec x)$ we find
1353: \beq
1354: \cN(f)\equiv-\int f(\vec x)\partial_i^2\cV_m(\vec x)d^3x=2f(\vec 0)+
1355: \frac{\cD}{\pi}\int_0^{2\pi}d\vphi\int_0^\cD dr~\frac{\partial_r f
1356: (\k'r\cos\vphi,0,r\sin\vphi)}{\sqrt{\cD^2-r^2}}.\label{eq:distest}
1357: \eeq
1358: The integral over $r$ is well defined for any $\k'$ and can be used to check 
1359: the correct normalization for the integrated zero-mode density, $\cN(1)=2$. 
1360: \refeq{distest} is also particularly convenient for studying the limit 
1361: $\k'\to0$. Using the fact that $\cN(f)$  is even in $\k'$, we may write 
1362: $\cN(f)=\cN_0(f)+\k'^2 \cN_2(f)+\cO(\k'^4)$, with in particular
1363: \beq
1364: \cN_0(f)=2f(\vec 0)+\frac{\cD}{\pi}\int_{-\cD}^\cD dy\int_{_{-\sqrt{\cD^2-y^2}}
1365: }^{^{\sqrt{\cD^2-y^2}}}dx~\frac{y\partial_yf(0,0,y)}{(x^2+y^2)
1366: \sqrt{\cD^2-x^2-y^2}},
1367: \eeq
1368: reintroducing cartesian coordinates 
1369: $x=r\cos\vphi$ and $y=r\sin\vphi$. The integral over $x$ is easily 
1370: performed and we find for $\cN_0(f)$
1371: \beq
1372: \cN_0(f)=2f(\vec 0)+\int_{-\cD}^\cD dy~{\rm sign}(y)\partial_yf(0,0,y)=
1373: f(0,0,\cD)+f(0,0,-\cD),
1374: \eeq
1375: whereas $\cN_2(f)$ gets contributions from $f$ on the line between $(0,0,\cD)$ 
1376: and $(0,0,-\cD)$. Therefore, we conclude that in the limit $\k \to1$ two 
1377: point-like constituents are found, and that this limit is approached in a 
1378: smooth way (despite the behavior observed in Fig.~\ref{fig:plfs}). 
1379: \begin{figure}[htb]
1380: \vskip5.1cm
1381: \special{psfile=testf.ps voffset=-45 hoffset=100 vscale=80.0 hscale=80.0}
1382: \caption{$\cN(f)$, \refeq{distest}, as a function of $\k^2$ for $f(\vec x)=
1383: \exp[-10(x_1^2+x_2^2+(x_3-1)^2)]$ and $\cD=1$.}\label{fig:testf}
1384: \end{figure}
1385: As an illustration we plot in Fig.~\ref{fig:testf} $\cN(f)$ as a function of 
1386: $\k^2$ for a Gaussian centered at one of the would-be constituent locations, 
1387: where it takes the value 1. We see that indeed $\cN(f)$ reaches 1 (linear 
1388: in $\k'^2$) for $\k\to1$. We also recall that \refeq{kpasym} implies the 
1389: point-like limit is reached exponentially in the constituent monopole 
1390: separation ($2\cD$).
1391: 
1392: For any $\k\neq1$ the core has an extended structure in the high temperature 
1393: limit. One might expect it to be extended along a line, since $\hat A_j(z)$ 
1394: depends on a single parameter. With $\vec R(z;\vec x)$ not constant, the 
1395: Riccati equation apparently ``smears out" the core, but suprisingly only to a
1396: disk. Note that inside this disk the singular zero-mode density is negative. 
1397: To guarantee the proper normalization, this is compensated by the singular 
1398: behavior at the ellipse which forms the edge of the disk. It should be noted, 
1399: however, that the singularity structure obtained in the high temperature limit 
1400: will only tell us to which region the core will be restricted. We will need 
1401: to resolve the non-abelian field components inside the core to understand how 
1402: this limiting behavior comes about. 
1403: 
1404: \section{Summary and Discussion}\label{sec:disc}
1405: 
1406: We studied the zero-modes of higher charge calorons, in particular in the 
1407: far field limit. In this limit one considers the regions outside the cores 
1408: of the constituents monopoles, where the gauge field is abelian. The
1409: constituent monopoles are most prominent when they all have a non-zero mass,
1410: implying a non-trivial value of the holonomy. Since the holonomy can be seen 
1411: as a background Polyakov loop, these calorons are therefore more relevant 
1412: for the confined phase, where the average Polyakov loop is non-trivial. 
1413: Lattice evidence for the relevance of these configurations is steadily 
1414: increasing~\cite{Ilg2,Gattp}, and much analytic understanding of the charge 1 
1415: calorons has been gained~\cite{KvB}. These reveal $n$ constituents for 
1416: $SU(n)$, which when well separated become static BPS monopoles. To localize 
1417: these monopoles it is useful to consider the high temperature limit, for 
1418: which the masses go to infinity and the non-abelian cores shrink to zero 
1419: size. We note that the high temperature limit is equivalent to the limit  
1420: of infinite Higgs vacuum expectation value (recall that $A_0$ plays the role 
1421: of the Higgs field in the adjoint representation). Since also chiral fermions 
1422: in a caloron background generically have a mass, likewise going to infinity 
1423: in the high temperature limit, one finds these zero-modes to be localized 
1424: entirely to the non-abelian cores. Indeed for charge 1, the zero-mode density 
1425: becomes a delta function at the location of one of the constituent monopoles. 
1426: Which one of them, is uniquely determined by the holonomy, and the phase up 
1427: to which the zero-mode is chosen to be periodic in the imaginary time 
1428: direction. We generalize away from anti-periodic boundary conditions since 
1429: this freedom in choosing the phase allows us to control on which constituent 
1430: the zero-mode localizes. It thus allows us to probe the underlying gauge 
1431: field configuration, see Fig.~\ref{fig:zmcycle}, as has been convincingly 
1432: demonstrated in Monte Carlo studies as well~\cite{Gattp}.
1433: 
1434: For higher charge $SU(n)$ calorons, $k>1$, there will be $kn$ constituent
1435: monopoles, or more precisely $k$ constituent monopoles of given type. Within 
1436: each type the magnetic charge associated to the embedding in $U^{n-1}(1)$, 
1437: and the mass is fixed. In a previous paper~\cite{BrvB} we constructed 
1438: axially symmetric solutions, which shared many of the properties of the 
1439: charge 1 solutions. In particular the high temperature limit gave 
1440: point-like constituents and we have verified here that the zero-mode 
1441: density indeed localizes to these constituents in the expected way, see 
1442: Fig.~\ref{fig:zmlocal}. This, however, is not expected to hold for higher 
1443: charge calorons in general. The reason is that one constructs these solutions 
1444: with the help of the Nahm equation, described by a dual $U(k)$ gauge field 
1445: $\hat A_j(z)$ on the circle, with singularities at the eigenvalues $\mu_m$ 
1446: of the logarithm of the holonomy. In case of the axially symmetric solutions 
1447: of Ref.~\cite{BrvB} this Nahm gauge field is actually piecewize constant 
1448: (apart from some trivial $U(k)$ gauge rotation), and its value on the interval 
1449: $z\in[\mu_m,\mu_{m+1}]$ directly determines the locations of the type $m$ 
1450: monopole constituents. However, in general the Nahm gauge field will not be 
1451: piecewize constant. In that case it remained to be seen if the constituent 
1452: locations are smeared out over all values of $\hat A(z)$, or worse.
1453: 
1454: Indeed in this paper we have found for charge 2 calorons in $SU(2)$ that in 
1455: general the location of the constituents is smeared out over a disk. To be
1456: more precise, we have shown that the zero-mode density vanishes everywhere,
1457: except on a disk that is bounded by an ellipse. We wish to stress that
1458: this result is found by solving the Nahm equation on an interval, without
1459: imposing the boundary conditions. In other words, a disk is described by 
1460: the parameters of the solution for each interval, that is for each monopole 
1461: type, namely the center of mass, the orientation, a shape and a scale 
1462: parameter (eight in total). These should also be the building blocks for 
1463: higher gauge groups $SU(n)$ since their dual descriptions differ only in 
1464: the number of intervals. Moreover, only the boundary conditions distinguish 
1465: between calorons and monopole solutions, and we believe our result is of 
1466: value in the context of the latter as well.
1467: 
1468: In particular we want to draw attention to the fact that (the trace of) the 
1469: chiral fermion zero-mode density in the high temperature limit is the Laplacian
1470: of a function $-\cV_m(\vec x)$ that is determined in terms of the Nahm gauge 
1471: field and turns out to be a constant of the motion with respect to the Nahm 
1472: equation. We have verified this property by computing the multipole expansion 
1473: of $\cV_m(\vec x)$ to a high order, but we expect this can be proven from the 
1474: integrability of the Nahm equations, an interesting problem to be pursued in 
1475: the future. The fact that $\cV_m(\vec x)$ is conserved is a powerful result 
1476: indeed, since it allowed us to calculate this function exactly, on which we 
1477: base the findings mentioned above. 
1478: 
1479: This makes us conjecture that the cores of the constituent monopoles are in 
1480: general extended, collapsing to the disk in the high temperature limit. In 
1481: itself this is not surprising, since one knows from the study of monopoles, 
1482: when two are closer together than the size of their core, they show an extended 
1483: structure, for charge 2 indeed in the shape of a doughnut~\cite{MonRing}. It 
1484: should also be noted that we found, when approaching the disk from above with 
1485: a test function smaller than the ring, the resulting zero-mode density can be 
1486: negative. There are two reasons why we should not be too worried about this. 
1487: If the core would collapse to points, the proper normalization of the 
1488: zero-modes requires the contribution from the singularity to be quantized, 
1489: giving a delta function of unit strength. When the core is extended, we have 
1490: no such constraint. In addition, our equations are derived by ignoring the 
1491: exponentially small terms that occur in the core. Therefore our result is 
1492: only valid outside the core, where we indeed find the zero-mode density to 
1493: vanish. This means the cores have to lie within the disk, whereas the only 
1494: thing we can say about their contributions is that they have to integrate to 
1495: 2 (since $-\partial_i^2\cV_m(\vec x)$ adds the two zero-mode densities). 
1496: Although this may seem to resemble a singularity like the Dirac string, it 
1497: is more subtle than that, because only like-charge monopoles are involved 
1498: here. Nevertheless, it does show that the non-abelian fields inside the 
1499: core have an intricate behavior. 
1500: 
1501: It will therefore be interesting in the future to try and get access to the 
1502: gauge field and to see if the field strength shows further localization within 
1503: the disks we have found on the basis of the zero-modes. Another useful object 
1504: is the Polyakov loop because it traces the constituent monopoles~\cite{MTAP}
1505: and is able to reveal extended structures in the context of Abelian 
1506: projection~\cite{AbPr,Ford}. We believe to have gained sufficient information 
1507: to get access to the solutions at finite temperature, with a fully resolved 
1508: core to answer these questions, if necessary by numerical means.
1509: 
1510: However, from the physical point of view, the most important result of this
1511: paper is our demonstration that for well separated constituents, when the 
1512: scale parameter $D$ is large, the shape of the caloron zero-mode density leads 
1513: to point-like constituents, i.e. the shape parameter $\k$ approaches 1. It 
1514: does so exponentially in $D$ and we have shown no structure is left on the 
1515: disk bounded by the ellipse, collapsing to a line in this limit. Any trace 
1516: left over on this line is proportional to $1-\k^2$, and therefore vanishes 
1517: exponentially in $D$. This comes about through the boundary conditions 
1518: $\hat A(z)$ has to satisfy at the ``impurities" $\mu_m$, relating the size 
1519: and shape parameters, $D$ and $\k$. We leave it to a future publication to 
1520: more fully describe the moduli space of solutions, solving for these 
1521: constraints. Also here some remarkable simplifications seem to occur, 
1522: related to the integrability of the Nahm equations. 
1523: 
1524: The results of this paper therefore provide one further step in establishing
1525: that non-abelian gauge field configurations can be described on a large 
1526: distance scale in terms of abelian monopoles. Of course it is only a small
1527: step, because we use exact self-dual solutions to establish these results.
1528: This has been in part because we discovered earlier~\cite{BrvB}, somewhat to 
1529: our surprise, that it is far from trivial to write down superpositions of 
1530: these monopole fields without having visible Dirac strings all over the 
1531: place, that would carry too much energy for comfort. In part this is due to 
1532: the crudeness involved in the superposition, instead requiring a fine-tuning 
1533: of the non-abelian tails with the exponential components in the abelian gauge 
1534: field (to properly absorb the return flux). It has been this problem to deal 
1535: with Dirac strings that for so long has hampered attempts to describe an 
1536: interacting theory of magnetic monopoles~\cite{Zwan}. 
1537: 
1538: In the light of this it would of course be welcome if more lattice studies
1539: are performed to get a handle on the dynamics of these constituent monopoles. 
1540: It should be pointed out that instantons larger than $\beta$ will no longer 
1541: reveal themselves as lumps of size $\rho$. Rather there is a transition region 
1542: beyond which $\rho$ should be interpreted to set the inter-constituent 
1543: distance (typically of order $\pi\rho^2/\beta$), whereas the size of the lumps 
1544: is in this region set by the mass of the constituent monopoles. This may lead 
1545: to a natural infrared cutoff in the size distribution of instantons, provided 
1546: by the temperature (rather than the spatial volume). Because of this it may be 
1547: worthwhile to reinvestigate the issue of the instanton size distribution, also 
1548: with the criticism presented in Ref.~\cite{Horv} in mind.
1549: 
1550: \section*{Appendix A}
1551: 
1552: We will derive in this appendix the zero-mode limit. Our starting point is the 
1553: explicit expression for the Green's function, \refeq{fdf}. It is useful to 
1554: note that this Green's function satisfies $f_x(z,z')=f_x^\dagger(z',z)$. To 
1555: show that \refeq{fdf} indeed respects this relation, one can use 
1556: (see \refeq{Wdef})
1557: \beqa
1558: &&\pmatrix{0&-\ein_k\cr\ein_k&0\cr}W^\dagger(z,z')\pmatrix{0&\ein_k\cr-\ein_k
1559: &0\cr}=W^{-1}(z,z'),\label{eq:Wdagger}\\&&\pmatrix{0&-\ein_k\cr\ein_k&0\cr}
1560: (\ein_{2k}-\cF_{z_0}^\dagger)^{-1}\pmatrix{0&\ein_k\cr-\ein_k&0\cr}=(
1561: \ein_{2k}-\cF_{z_0}^{-1})^{-1}=\ein_{2k}-(\ein_{2k}-\cF_{z_0})^{-1}.\nonumber
1562: \eeqa
1563: We will take $z_0=\mu_m+0$, which leads to the following decomposition of 
1564: $\cF_{\mu_m}$ in terms of contributions coming from the impurities ($T_m$) 
1565: and from the propagation between the impurities ($H_m$),
1566: \beqa
1567: &&\cF_{\mu_m}=T_mH_{m-1}\cdots T_2H_1T_1\hat g^\dagger(1)H_n T_nH_{n-1}
1568: \cdots T_{m+1}H_m,\nonumber\\&& T_m=\exp\pmatrix{0&0\cr 2\pi S_m&0\cr},\quad 
1569: H_m=\hat W_m(\mu_{m+1})F_m(\mu_{m+1})\hat W_m^{-1}(\mu_m).\label{eq:THs}
1570: \eeqa
1571: Note that by definition $F_m(\mu_m)=\ein_k$. It is convenient to absorb 
1572: the algebraic contributions coming from $\hat W_m$ in the ``impurity" 
1573: contributions $T_m$.
1574: \beqa
1575: \Theta_m&=&\pmatrix{\theta^m_{++}&\theta^m_{+-}\cr\theta^m_{-+}&\theta^m_{--}
1576: \cr}\equiv\hat W_m^{-1}(\mu_m)T_m\hat W_{m-1}(\mu_m)\label{eq:Theta}\\
1577: &=&\half R_m^{-1}(\mu_m)\pmatrix{
1578: R_m^-(\mu_m)+R_{m-1}^+(\mu_m)+S_m&R_m^-(\mu_m)-R_{m-1}^-(\mu_m)+S_m\cr
1579: R_m^+(\mu_m)-R_{m-1}^+(\mu_m)-S_m&R_m^+(\mu_m)+R_{m-1}^-(\mu_m)-S_m\cr}.
1580: \nonumber
1581: \eeqa
1582: The following identities are noteworthy
1583: \beq
1584: \theta^m_{+\pm}+\theta_{-\pm}^m=\ein_k,\quad\theta^m_{\pm+}-\theta_{\pm-}^m=
1585: \pm R_m^{-1}(\mu_m)R_{m-1}(\mu_m),\quad 2R_m(\mu_m)\theta^m_{++}=\Sigma_m,
1586: \label{eq:thetas}
1587: \eeq
1588: as well as the fact that $\Theta^{-1}_m=\hat W_{m-1}^{-1}(\mu_m)T_m^{-1}
1589: \hat W_m(\mu_m)$ can be computed explicitly
1590: \beq
1591: \Theta^{-1}_m=\pmatrix{\theta_m^{++}&\theta_m^{+-}\cr\theta_m^{-+}&
1592: \theta_m^{--}\cr}=R_{m-1}^{-1}(\mu_m)R_m(\mu_m)\pmatrix{\theta^m_{--}&
1593: -\theta^m_{+-}\cr-\theta^m_{-+}&\theta^m_{++}\cr}.\label{eq:Thetainv}
1594: \eeq
1595: Finally, introducing 
1596: \beq
1597: \hat\cF_{\mu_m}=W^{-1}_m(\mu_m)\cF_{\mu_m}W_m(\mu_m)=\Theta_mF_{m-1}
1598: \cdots\Theta_2F_1\hat g^\dagger(1)\Theta_nF_{n-1}\cdots\Theta_{m+1}F_m,
1599: \label{eq:cFhat}
1600: \eeq
1601: we find for $\mu_m\leq z'\leq z\leq\mu_{m+1}$ (cmp. \refeq{fdf}) 
1602: \beq
1603: f_x(z,z')=-4\pi^2\pmatrix{\ein_k\cr0\cr}^tW_m(z)(\ein_{2k}-\hat
1604: \cF_{\mu_m})^{-1}W_m^{-1}(z')\pmatrix{0\cr\ein_k\cr}.\label{eq:farf}
1605: \eeq
1606: It will be useful to write $\hat\cF_{\mu_m}=F_m^{-1}\Theta_{m+1}^{-1}LK
1607: \Theta_{m+1}F_m$, with $L\equiv\Theta_{m+1}F_m\Theta_m$, because 
1608: $K\equiv F_{m-1}\Theta_{m-1}\cdots\Theta_1\hat g^\dagger(1)F_n\Theta_n
1609: \cdots\Theta_{m+2}F_{m+1}$ contains the exponential factors in terms of 
1610: which we can take the zero-mode limit. 
1611: 
1612: With $(\ein_{2k}-\hat\cF_{\mu_m})^{-1}=F_m^{-1}\Theta_{m+1}^{-1}(1-LK)^{-1}
1613: \Theta_{m+1}F_m$, we reduce the problem to approximating $(\ein_{2k}-LK)^{-1}$.
1614: For this it is convenient to write $LK\equiv\hat L\hat K+\tilde L\tilde K$, 
1615: with
1616: \beq
1617: \hat K\equiv\pmatrix{K_{++}&K_{+-}\cr 0&\ein_k\cr},\quad
1618: \tilde K\equiv\pmatrix{0&0\cr K_{-+}&K_{--}\cr},\quad
1619: \hat L\equiv\pmatrix{L_{++}&0\cr L_{-+}&0\cr},\quad
1620: \tilde L\equiv\pmatrix{0&L_{+-}\cr0&L_{--}\cr}.
1621: \eeq
1622: after which we find
1623: \beq
1624: (\ein_{2k}-LK)^{-1}=\hat K^{-1}\left(\hat K^{-1}-\hat L
1625: -\tilde L\tilde K\hat K^{-1}\right)^{-1}.
1626: \eeq
1627: As we will show next, the advantage of all this is that terms containing
1628: $K_{\pm\pm}$ are of the form $K_{++}^{-1}$, $K_{++}^{-1}K_{+-}$, $K_{-+}
1629: K_{++}^{-1}$ or $(K^{--})^{-1}\equiv K_{--}-K_{-+}K_{++}^{-1}K_{+-}$ and
1630: that these are all exponentially decaying. For the first three this is 
1631: easily seen using that $K$ is of the form $F\Theta F\Theta\cdots F\Theta F$, 
1632: whereas for the last term we recall a well-known formula for the inverse of 
1633: a $2\times 2$ matrix with as entries ($k\times k$) matrices
1634: \beq
1635: K^{-1}=\pmatrix{K_{++}&K_{+-}\cr K_{-+}&K_{--}\cr}^{-1}=
1636: \pmatrix{(K_{++}-K_{+-}K^{-1}_{--}K_{-+})^{-1}&
1637:          (K_{-+}-K_{--}K^{-1}_{+-}K_{++})^{-1}\cr
1638:          (K_{+-}-K_{++}K^{-1}_{-+}K_{--})^{-1}&
1639:          (K_{--}-K_{-+}K^{-1}_{++}K_{+-})^{-1}}.\label{eq:Kinv}
1640: \eeq
1641: From this we find that $(K^{--})^{-1}=(K^{-1})_{--}$ (hence the upper 
1642: indices). With $K^{-1}$ having the form $F^{-1}\Theta^{-1}F^{-1}\Theta^{-1}
1643: \cdots F^{-1}\Theta^{-1} F^{-1}$, which interchanges the role of $f^+$ and 
1644: $f^-$, we conclude that $K^{--}$ behaves as $K_{++}$ and that therefore 
1645: $(K^{--})^{-1}$ is exponentially decaying as well. Using 
1646: \beq
1647: \hat K^{-1}=\pmatrix{K^{-1}_{++}&-K^{-1}_{++}K_{+-}\cr 0&\ein_k\cr},\quad
1648: \tilde K\hat K^{-1}=\pmatrix{0&0\cr K_{-+}K^{-1}_{++}&(K^{--})^{-1}\cr}
1649: \eeq
1650: and neglecting the exponentially decaying terms, we find the simple result
1651: \beq
1652: (\ein_{2k}-LK)^{-1}\to\pmatrix{0&0\cr0&\ein_k}\pmatrix{-L_{++}&0\cr
1653: -L_{-+}&\ein_k}^{-1}=\pmatrix{0&0\cr -L_{-+}L_{++}^{-1}&\ein_k},\label{eq:ffl}
1654: \eeq
1655: or including subleading terms
1656: \beq
1657: (\ein_{2k}-LK)^{-1}=\pmatrix{\cO(K_{++}^{-1})&\cO(K_{++}^{-1})\cr -L_{-+}
1658: L_{++}^{-1}+\cO(X)&\ein_k+\cO(X)},\label{eq:subffl}
1659: \eeq
1660: where $\cO(X)\equiv\cO(K_{++}^{-1}K_{+-})+\cO(K_{-+}K_{++}^{-1})+
1661: \cO((K^{--})^{-1})$.
1662: 
1663: Using \refeq{ffl} and $\Theta_{m+1}F_m=L\Theta_m^{-1}$ we find
1664: \beq
1665: f^{\zm}_x(z,z')=\pi\pmatrix{\ein_k\cr\ein_k\cr}^t F_m(z)F_m^{-1}
1666: \Theta_{m+1}^{-1}\pmatrix{0&0\cr -L_{-+}L_{++}^{-1}&\ein_k\cr}L\Theta_m^{-1}
1667: F_m(z')\pmatrix{-\ein_k\cr\hphantom{-}\ein_k\cr}R^{-1}_m(z'),
1668: \eeq
1669: This can be simplified further using $L^{--}\equiv(L^{-1})_{--}=(L_{--}-
1670: L_{-+}L^{-1}_{++}L_{+-})^{-1}$ (cmp. \refeq{Kinv}), such that 
1671: \beq
1672: f^{\zm}_x(z,z')=\pi\pmatrix{f_m^+(z)\cr f_m^-(z)\cr}^tF_m^{-1}
1673: \Theta_{m+1}^{-1}\pmatrix{0&0\cr0&(L^{--})^{-1}\cr}\Theta_m^{-1}
1674: \pmatrix{-f_m^+(z')^{-1}\cr\hphantom{-}f_m^-(z')^{-1}\cr}R^{-1}_m(z').
1675: \eeq
1676: With Eqs.~(\ref{eq:thetas},\ref{eq:Thetainv}), noting that $Z_m^-=-
1677: \theta_m^{+-}(\theta_m^{--})^{-1}$ and $Z_m^+=(\theta_m^{--})^{-1}
1678: \theta_m^{-+}$ (see \refeq{Zedef}), we find after some algebra the relatively 
1679: simple result given in \refeq{zmlim}. We note that it is not directly obvious 
1680: that $f^\zm_x(z,z')=f_x^\zm(z',z)^\dagger$. Nevertheless, this is guaranteed 
1681: to be true from the fact that the exact Green's function respects this 
1682: property. All we wish to mention here, is that \refeq{Wdagger} implies rather 
1683: non-trivial relations involving $f_m^\pm(z)^\dagger$ and $R_m^\pm(z)^\dagger$, 
1684: which could be used to explicitly verify that $f^\zm_x(z,z')=
1685: f_x^\zm(z',z)^\dagger$. 
1686: 
1687: \section*{Appendix B}
1688: 
1689: Using the invariance under a one-parameter set of rotations around $\hat x$ 
1690: for $M$, \refeq{defM}, or equivalently around $(1,0,0)$ for $\hat M(\hat x)$, 
1691: \refeq{defMhat}, we can for charge 2 express the multipole expansion of 
1692: $\cV_m(\vec x)$ in the following 4 independent parameters,
1693: \beqa
1694: &&\!\!\!p\equiv\frac{3}{2}\hat M_{11}(\hat x),\quad w^2\equiv\hat M_{12}^2(\hat
1695: x)+\hat M_{13}^2(\hat x),\quad q^2\equiv\frac{1}{8}\left(\hat M_{11}(\hat x)+
1696: 2\hat M_{22}(\hat x)\right)^2\!\!+\frac{1}{2}\hat M_{23}^2(\hat x),\nonumber\\
1697: &&\!\!\!\!s^3\equiv4\hat M_{12}(\hat x)\hat M_{13}(\hat x)\hat M_{23}(\hat x)+
1698: \left(\hat M_{12}^2(\hat x)-\hat M_{13}^2(\hat x)\right)\left(\hat M_{11}
1699: (\hat x)+2\hat M_{22}(\hat x)\right),
1700: \eeqa
1701: where 
1702: $|\vec x|^2p$, $|\vec x|^4w^2$, $|\vec x|^4q^2$ and $|\vec x|^6s^3$ can be
1703: written as monomials in $\vec x$. This choice has the particular advantage 
1704: that for charge 2 the following remarkably simple form can be used 
1705: \beq
1706: \cV_m(\vec x)=\frac{1}{2\pi}\sum_{n=0}\frac{a_n}{|\vec x|^{2n+1}},\quad
1707: \partial_pa_n=na_{n-1},\label{eq:mpol}
1708: \eeq
1709: checked to order $|\vec x|^{-21}$, but likely to be true to all orders.
1710: Therefore, the result to this order can be read off from the $l=2n=20$ 
1711: multipole coefficient 
1712: \beqa
1713: a_{10}&=&\!\!p^{10}\!+45p^8(q^2\!-2w^2)+180p^7s^3\!+315p^6(q^4\!-8q^2w^2\!+
1714:   4w^4)+630p^5s^3(3q^2\!-4w^2)\nonumber\\ 
1715: &+&\!\!\frac{525}{2}p^4(2q^6-36q^4w^2+60q^2w^4-16w^6+3s^6)+
1716:   3150p^3s^3(q^4-4q^2w^2+2w^4)\nonumber\\ 
1717: &+&\!\!\frac{315}{8}p^2(5q^8-160q^6w^2+560q^4w^4-448q^2w^6+40q^2s^6+
1718:   80w^8-48w^2s^6)\nonumber\\
1719: &+&\!\!\frac{105}{2}ps^3(15q^6-120q^4w^2+168q^2w^4-48w^6+2s^6)+ 
1720:   \frac{63}{8}\Bigl(q^{10}-50q^8w^2\nonumber\\
1721: &&\quad+300q^6w^4+5q^4(5s^6-96w^6)+80q^2(3w^8-w^2s^6)+40w^4s^6-32w^{10}\Bigr).
1722: \label{eq:m20}
1723: \eeqa
1724: 
1725: \section*{Acknowledgements}
1726: 
1727: We thank Conor Houghton for sharing his insights and unpublished notes, 
1728: as well as Andreas Wipf, David Adams and in particular Chris Ford for 
1729: discussions. We also thank Michael M\"uller-Preussker, Michael Ilgenfritz, 
1730: Boris Martemyanov, Stanislav Shcheredin and Christof Gattringer for 
1731: stimulating discussions concerning calorons with non-trivial holonomy 
1732: on the lattice. The research of FB is supported by FOM. 
1733: 
1734: \begin{thebibliography}{99}
1735: \bibitem{THPo}G. 't Hooft, Nucl. Phys. B79 (1974) 276; 
1736: %%CITATION = NUPHA,B79,276;%%
1737: A.M. Polyakov, JETP Lett. 20 (1974) 194. %%CITATION = JTPLA,20,194;%%
1738: \bibitem{DuSC}S. Mandelstam, Phys. Rep. 23 (1976) 245;
1739: %%CITATION = PRPLC,23,245;%%
1740: G. 't Hooft, in: {\em High Energy Physics}, ed. A. Zichichi (Editrice 
1741: Compositori, Bolognia, 1976); Nucl. Phys. B138 (1978) 1. 
1742: %%CITATION = NUPHA,B138,1;%%
1743: \bibitem{SeWi}N. Seiberg and E. Witten, Nucl. Phys. B426 (1994) 19; erratum 
1744: B430, (1994) 485 [hep-th/9407087]. %%CITATION = HEP-TH 9407087;%%
1745: \bibitem{GrRe}For a review see J. Greensite, {\em The confinement problem in 
1746: lattice gauge theory}, hep-lat/0301023 (to appear in Prog. Part. Nucl. Phys.)
1747: %%CITATION = HEP-LAT 0301023;%%
1748: \bibitem{Suzu}T. Suzuki and I. Yotsuyanagi, Phys.Rev. D42 (1990) 4257;
1749: %%CITATION = PHRVA,D42,4257;%%
1750: For a review see the contributions of M. Chernodub and M. Polikarpov
1751: [hep-th/9710205], A. Di Giacomo [hep-th/9710080] and T. Suzuki, in: 
1752: {\em Confinement, Duality and Non-perturbative Aspects of QCD}, 
1753: ed. P. van Baal, NATO ASI Series B: Vol. 368 (Plenum Press, 1998).
1754: %%CITATION = HEP-TH 9710205;%%
1755: %%CITATION = HEP-TH 9710080;%%
1756: \bibitem{AbPr}G. 't Hooft, Nucl. Phys. B190 [FS3] (1981) 455;
1757: %%CITATION = NUPHA,B190,455;%%
1758: Physica Scripta 25 (1982) 133.
1759: %%CITATION = PHSTB,25,133;%%
1760: \bibitem{NCal}W. Nahm, {\em Self-dual monopoles and calorons}, in:
1761: Lect. Notes in Physics. 201, eds. G. Denardo, e.a. (1984) p. 189.
1762: \bibitem{Lee}K. Lee and P. Yi, Phys. Rev. D56 (1997) 3711 [hep-th/9702107]; 
1763: %%CITATION = HEP-TH 9702107;%%
1764: K. Lee, Phys. Lett. B426 (1998) 323 [hep-th/9802012]; 
1765: %%CITATION = HEP-TH 9802012;%%
1766: K. Lee and C. Lu, Phys. Rev. D58 (1998) 025011 [hep-th/9802108].
1767: %%CITATION = HEP-TH 9802108;%%
1768: \bibitem{KvB}T.C. Kraan and P. van Baal, Phys. Lett. B428 (1998) 268
1769: [hep-th/9802049]; %%CITATION = HEP-TH 9802049;%%
1770: Nucl. Phys. B533 (1998) 627 [hep-th/9805168]; %%CITATION = HEP-TH 9805168;%%
1771: Phys. Lett. B435 (1998) 389 [hep-th/9806034]. %%CITATION = HEP-TH 9806034;%%
1772: \bibitem{Dubna}P. van Baal, in: {\em Lattice fermions and structure of the 
1773: vacuum}, eds. V. Mitrjushkin and G. Schierholz (Kluwer, Dordrecht, 2000), 
1774: p. 269 [hep-th/9912035]. %%CITATION = HEP-TH 9912035;%%
1775: \bibitem{BrvB}F. Bruckmann and P. van Baal, Nucl. Phys. B645 (2002) 105 
1776: [hep-th/0209010]. %%CITATION = HEP-TH 0209010;%%
1777: \bibitem{MTCP}M. Garc\'{\i}a P\'erez, A. Gonz\'alez-Arroyo, C. Pena and P. van
1778: Baal, Phys. Rev. D60 (1999) 031901 [hep-th/9905016].
1779: %%CITATION = HEP-TH 9905016;%%
1780: \bibitem{MTP}M.N. Chernodub, T.C. Kraan and P. van Baal, Nucl. Phys. 
1781: B(Proc.Suppl.) 83-84 (2000) 556 [hep-lat/9907001].
1782: %%CITATION = HEP-LAT 9907001;%%
1783: \bibitem{Ilg1}E.-M. Ilgenfritz, M. M\"uller-Preussker, and A.I. Veselov, in: 
1784: {\em Lattice fermions and structure of the vacuum}, eds. V. Mitrjushkin and 
1785: G. Schierholz (Kluwer, Dordrecht, 2000), 345 [hep-lat/0003025];
1786: %%CITATION = HEP-LAT 0003025;%%
1787: E.-M. Ilgenfritz, B.V. Martemyanov, M. M\"uller-Preussker and A.I. Veselov, 
1788: Nucl. Phys. B(Proc.Suppl.)94 (2001) 407 [hep-lat/0011051]; 
1789: %%CITATION = HEP-LAT 0011051;%%
1790: Nucl. Phys. B(Proc. Suppl.)106 (2002) 589 [hep-lat/0110212].
1791: %%CITATION = HEP-LAT 0110212;%%
1792: \bibitem{Ilg2} E.-M. Ilgenfritz, B.V. Martemyanov, M. M\"uller-Preussker, 
1793: S. Shcheredin and A.I. Veselov, Phys. Rev. D66 (2002) 074503 [hep-lat/0206004].
1794: %%CITATION = HEP-LAT 0206004;%%
1795: \bibitem{Gattp}C. Gattringer and S. Schaefer, Nucl. Phys. B654 (2003) 30 
1796: [hep-lat/0212029];\\ %%CITATION = HEP-LAT 0212029;%%
1797: see also C. Gattringer, Phys. Rev. D67 (2003) 034507 [hep-lat/0210001].
1798: %%CITATION = HEP-LAT 0210001;%%
1799: \bibitem{MTAP}M. Garc\'{\i}a P\'erez, A. Gonz\'alez-Arroyo, A. Montero and P.
1800: van Baal, Jour. of High Energy Phys. 06 (1999) 001 [hep-lat/9903022].
1801: %%CITATION = HEP-LAT 9903022;%%
1802: \bibitem{Wein}K. Lee, E.J. Weinberg and P. Yi, Phys. Lett. B376 (1996) 97
1803: [hep-th/9601097]; %%CITATION = HEP-TH 9601097;%%
1804: Phys. Rev. D54 (1996) 6351 [hep-th/9605229]; %%CITATION = HEP-TH 9605229;%%
1805: E.J. Weinberg, {\em Massive and Massless Monopoles and Duality}, hep-th/9908095;
1806: %%CITATION = HEP-TH 9908095;%%
1807: C.J. Houghton and E.J. Weinberg, Phys. Rev. D66 (2002) 125002 [hep-th/0207141].
1808: %%CITATION = HEP-TH 0207141;%%
1809: \bibitem{WWW}www.lorentz.leidenuniv.nl/vanbaal/Caloron.html
1810: \bibitem{Nahm}W. Nahm, Phys. Lett. 90B (1980) 413.
1811: %%CITATION = PHLTA,B90,413;%%
1812: \bibitem{Bip}P. van Baal, in: {\em Confinement, Topology, and other 
1813: Non-Perturbative Aspects of QCD}, eds. J. Greensite and S. Olejnik, NATO 
1814: Science Series, Vol. 83 (Kluwer, Dordrecht, 2002), p. 1 [hep-th/0202182]. 
1815: %%CITATION = HEP-TH 0202182;%%
1816: \bibitem{ADHM}M.F. Atiyah, N.J. Hitchin, V.G. Drinfeld, Yu. I. Manin,  
1817: Phys. Lett. 65 A (1978) 185; %%CITATION = PHLTA,A65,185;%%
1818: M.F. Atiyah, {\em Geometry of Yang-Mills fields}, Fermi lectures, 
1819: (Scuola Normale Superiore, Pisa, 1979).
1820: \bibitem{Temp}E.F. Corrigan, D.B. Fairlie, S. Templeton and P. Goddard,
1821: Nucl. Phys. B140 (1978) 31. %%CITATION = NUPHA,B140,31;%%
1822: \bibitem{Osb}H. Osborn, Nucl. Phys.  B159 (1979) 497.
1823: %%CITATION = NUPHA,B159,497;%%
1824: \bibitem{CoG}E.F. Corrigan and P. Goddard, Annals Phys. (N.Y.) 154 (1984) 253.
1825: %%CITATION = APNYA,154,253;%%
1826: \bibitem{FORM}J.A.M.Vermaseren, {\em New features of FORM}, math-ph/0010025.
1827: %%CITATION = MATH-PH 0010025;%%
1828: \bibitem{NahmM}W. Nahm, Multi-monpoles in the ADHM construction (preprint 
1829: IC/81/238), in: {\em Gauge theories and lepton hadron interactions}, eds. 
1830: Z. Horvath, e.a., (CRIP, Budapest, 1982).
1831: \bibitem{BrDa}S.A. Brown, H. Panagopoulos and M.K. Prasad, 
1832: Phys. Rev. D26 (1982) 854;\\ %%CITATION = PHRVA,D26,854;%%
1833: A.S. Dancer, Comm. Math. Phys. 158 (1993) 545. %%CITATION = CMPHA,158,545;%%
1834: \bibitem{AbSt}M. Abramowitz and I. Stegun (eds.), {\em Handbook of Mathematical
1835: Functions}, (Dover Publ., New York, 1972).
1836: \bibitem{HoKr}C.J. Houghton and T.C. Kraan, {\em Notes on two-caloron Nahm
1837: data}, July 1998 and June 1999, unpublished.
1838: \bibitem{MonRing}P. Forg\'acs, Z.~Horv\'ath and L. Palla, 
1839: Nucl. Phys. B192 (1981) 141; %%CITATION = NUPHA,B192,141;%%
1840: M.F.~Atiyah, and N.J.~Hitchin, {\em The Geometry and Dynamics of
1841: Magnetic Monopoles}, (Princeton Univ. Press, 1988).
1842: \bibitem{Ford}C. Ford, U.G. Mitreuter, T. Tok, A. Wipf and J.M. Pawlowski,
1843: Annals Phys. 269 (1998) 26 [hep-th/9802191]. %%CITATION = HEP-TH 9802191;%%
1844: \bibitem{Zwan}D. Zwanziger, Phys. Rev. 176 (1968) 1489;
1845: %%CITATION = PHRVA,176,1489;%%
1846: T.S. Tu, T.T. Wu and C.N. Yang, Sci. Sin. 21 (1978) 317,
1847: %%CITATION = SSINA,21,317;%%
1848: reprinted in C.N. Yang, {\em Selected Papers 1945-1980, with Commentary},
1849: (W.H. Freeman and Co., New York, 1983). 
1850: \bibitem{Horv}I. Horvath, N. Isgur, J. McCune and H.B. Thacker,
1851: Phys. Rev. D65 (2002) 014502 [hep-lat/0102003]. %%CITATION = HEP-LAT 0102003;%%
1852: \end{thebibliography}
1853: \end{document}
1854: