cond-mat0601347/sc3.tex
1: \documentclass[twocolumn,showpacs,showkeys]{revtex4}
2: \usepackage{graphicx,subfigure}
3: \input{psfig.sty}
4: \newcommand {\gs}{\
5:   \raisebox{-0.2ex}{$\stackrel{\scriptstyle>}{\scriptstyle\sim}$}\ }
6: \newcommand{\ls}{\
7:   \raisebox{-0.2ex}{$\stackrel{\scriptstyle<}{\scriptstyle\sim}$}\ }
8: \parindent=0.5cm
9: \parskip=0.2 cm
10: \newcommand{\la}[1]{\label{#1}}
11: \newcommand{\bastar}{\begin{eqnarray*}}
12: \newcommand{\eastar}{\end{eqnarray*}}
13: \newskip\humongous \humongous=0pt plus 1000pt minus 1000pt
14: \def\caja{\mathsurround=0pt}
15: \def\eqalign#1{\,\vcenter{\openup1\jot \caja
16:         \ialign{\strut \hfil$\displaystyle{##}$&$
17:         \displaystyle{{}##}$\hfil\crcr#1\crcr}}\,}
18: \newif\ifdtup
19: \def\panorama{\global\dtuptrue \openup1\jot \caja
20:         \everycr{\noalign{\ifdtup \global\dtupfalse
21:         \vskip-\lineskiplimit \vskip\normallineskiplimit
22:         \else \penalty\interdisplaylinepenalty \fi}}}
23: \def\eqalignno#1{\panorama \tabskip=\humongous
24:         \halign to\displaywidth{\hfil$\displaystyle{##}$
25:         \tabskip=0pt&$\displaystyle{{}##}$\hfil
26:         \tabskip=\humongous&\llap{$##$}\tabskip=0pt
27:         \crcr#1\crcr}}
28: \relax
29: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
30: \newcommand{\be}{\begin{equation}}
31: \newcommand{\ee}{\end{equation}}
32: \newcommand{\bea}{\begin{eqnarray}}
33: \newcommand{\eea}{\end{eqnarray}}
34: \newcommand{\X}{{\vec X}}
35: \newcommand{\pro}{\partial}
36: \newcommand{\n}{\hat n}
37: \newcommand{\oneg}{\displaystyle\frac{1}{g}}
38: \newcommand{\alfa}{\hat{\alpha}}
39: \newcommand{\D}{{\hat D}}
40: \newcommand{\hfi}{{\hat \phi}}
41: \newcommand{\B}{{\vec B}}
42: \newcommand{\C}{\stackrel{\ast}{C}}
43: \newcommand{\ksi}{{\hat \xi}}
44: \newcommand{\A}{{\vec A}}
45: \newcommand{\valpha}{{\vec \alpha}}
46: \newcommand{\vbeta}{{\vec \beta}}
47: \newcommand{\dfrac}{\displaystyle\frac}
48: \newcommand{\ba}{\begin{array}}
49: \newcommand{\ea}{\end{array}}
50: \newcommand{\dpst}{\displaystyle}
51: \newcommand{\vareps}{\varepsilon}
52: \newcommand{\uone}{\mbox{$U(1)$\,\,}}
53: \newcommand{\suu}{\mbox{$SU(2)$\,\,}}
54: \newcommand{\nn}{\nonumber}
55: \newcommand{\hn}{\hat n}
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57: \begin{document}
58: \title{Topological Objects in Two-gap Superconductor:I}
59: \bigskip
60: \author{Y. M. Cho}
61: \email{ymcho@yongmin.snu.ac.kr}
62: \author{Pengming Zhang}
63: \date{\today}
64: \email{zhpm@phya.snu.ac.kr}
65: \affiliation{Center for Theoretical Physics and School of Physics \\
66: College of Natural Sciences, Seoul National University,
67: Seoul 151-742, Korea  \\}
68: \begin{abstract}
69: We discuss topological objects, in particular the non-Abrikosov
70: vortex and the magnetic knot made of the
71: twisted non-Abrikosov vortex, in two-gap superconductor.
72: We show that there are two types of non-Abrikosov vortex in
73: Ginzburg-Landau theory of two-gap superconductor, the D-type which has
74: no concentration of the condensate at the core
75: and the N-type which has a non-trivial
76: profile of the condensate at the core, under a wide
77: class of realistic interaction potential.
78: Furthermore, we show that we can construct a stable magnetic knot
79: by twisting the non-Abrikosov vortex and connecting two periodic ends
80: together, whose knot topology $\pi_3(S^2)$ is described by the Chern-Simon
81: index of the electromagnetic potential.
82: We discuss how these topological objects can be constructed in
83: $\rm MgB_2$ or in liquid metallic hydrogen.
84: \end{abstract}
85: \pacs{74.20.-z, 74.20.De, 74.60.Ge, 74.60.Jg, 74.90.+n}
86: \keywords{non-Abrikosov magnetic vortex, fractional magnetic flux,
87: magnetic knot in two-gap superconductor}
88: \maketitle
89: 
90: \section{Introduction}
91: 
92: Topological objects, in particular finite energy topological
93: objects (monopoles, vortices, skyrmions, and knots),
94: have played increasingly important role
95: in physics \cite{dirac,abri,skyr,fadd1,prl01}. In condensed matter the
96: best known topological objects are the Abrikosov
97: vortex in one-gap superconductors and similar ones in Bose-Einstein
98: condensates and superfluids, which have been the subject
99: of intensive studies.
100: A recent advent of two-component Bose-Einstein condensates
101: and two-gap superconductors \cite{bec,sc},
102: however, has opened up an exciting new possibility
103: for us to construct far more interesting topological
104: objects in laboratories.
105: It has already been shown that non-Abrikosov vortices whose
106: topology is fixed by $\pi_2(S^2)$ and finite energy topological
107: knots whose topology is fixed by $\pi_3(S^2)$ exist in these
108: condensed matters \cite{ijpap,ruo,pra05,prb05,baba1,cm2}.
109: The reason for this is
110: that these condensed matters are made of
111: two components which can be viewed as an $SU(2)$ multiplet.
112: In general this type of topological objects is possible
113: when one has a multi-component condensates,
114: which allows the non-Abelian topology.
115: 
116: The purpose of this paper is to discuss new
117: topological objects in Ginzburg-Landau theory of
118: two-gap superconductor in detail.
119: {\it With a most general $U(1)\times U(1)$ symmetric potential
120: which can describe a wide class of
121: two-gap superconductors we first show that there are
122: two types of non-Abrikosov vortex,
123: D-type and N-type, in two-gap superconductor.
124: The D-type has no concentration of the condensate
125: at the core, but the N-type has a non-trivial
126: profile of the condensate at the core.  The reason why the
127: two-gap superconductor has two types of vortex is
128: that the vortex in two-gap superconductor allows
129: two different boundary conditions. In terms of topology
130: the non-Abrikosov vortex is described by two
131: types of topology, non-Abelian $\pi_2(S^2)$ topology
132: or Abelian $\pi_1(S^1)$ topology.
133: And within the same topology both D-type and N-type vortices
134: exist. In particular, there are infinitely many D-type
135: vortices classified by the natural number $k$ which have the same topology.
136: Moreover, the magnetic flux of these non-Abrikosov vortices
137: can be integral or fractional, and the integral flux vortex has
138: the $\pi_2(S^2)$ topology and the fractional flux vortex has
139: the $\pi_1(S^1)$ topology.}
140: We show that the N-type vortex
141: has a $2\pi n/g$-flux or a fractional flux (a fraction of
142: $2\pi n/g$), but the D-type vortex has $2\pi k/g$ more
143: flux than the N-type vortex.
144: These characteristic features of the non-Abrikosov vortex
145: are clearly absent in the Abrikosov vortex
146: which carries $2\pi n/g$-flux whose topology is fixed by
147: $\pi_1(S^1)$.
148: 
149: Next, we show that the non-Abrikosov vortex
150: can be twisted to form a helical vortex which is periodic
151: in $z$-axis. {\it More importantly, we show that
152: we can construct a stable magnetic knot in two-gap superconductors
153: by smoothly bending the helical vortex and
154: connecting the periodic ends together. The vortex ring
155: acquires the knot topology $\pi_3(S^2)$ which is fixed by
156: the Chern-Simon index of the electromagnetic potential.}
157: Because of the helical structure of the
158: magnetic flux the knot has two magnetic flux linked together,
159: one around the knot tube and one along the knot,
160: whose linking number is given by the knot quantum number.
161: And the flux trapped inside the vortex ring provides
162: a stabilizing repulsive force which prevents the collapse
163: of the knot, because it can not be squeezed out.
164: This means that the knot has dynamical (as well as topological)
165: stability.
166: 
167: It is well-known that multi-gap superconductor may have
168: interband Josephson interaction \cite{maz}.
169: We consider a most general quartic Josephson interaction
170: in two-gap superconductor, and show that the presence of
171: the Josephson interaction does not affect the existence of the
172: above topological objects, but can alter the shape of the
173: solutions drastically. {\it We show that in the presence of
174: the Josephson interaction we have a magnetic vortex
175: which can be viewed as a bound state of two fluxes,
176: which becomes a braided magnetic vortex when twisted.}
177: 
178: The paper is organized as follows. In Section II we discuss
179: a most general quartic potential in Ginzburg-Landau theory of
180: two-gap superconductor in mean field approximation, and
181: study the vacuum structure. In Section III we show that
182: the Ginzburg-Landau theory of two-gap superconductor
183: can be understood as a theory of $CP^1$ field coupled to
184: a scalar field and the electromagnetic field, and argue that
185: the topology of the theory can be described by the $CP^1$ field
186: and the electromagnetic field. In Section IV we
187: construct the non-Abrikosov vortex in two-gap superconductor,
188: and show that there are two types of boundary condition
189: which allow two types of magnetic vortex,
190: the D-type which has no concentration of condensate at the core
191: and the N-type which has a non-trivial concentration of condensate
192: at the core. Moreover we show that there are two types of topology,
193: non-Abelian $\pi_2(S^2)$ and Abelian $\pi_1(S^1)$,
194: which describes these vortices.
195: We show that the magnetic flux of these vortices can be
196: integral or fractional depending on
197: the parameters of the potential, but the D-type
198: vortex has $2\pi k/g$ more flux than the N-type vortex.
199: In Section V we show that we can construct a
200: helical magnetic vortex in two-gap superconductor,
201: by twisting the non-Abrikosov vortex and making it
202: periodic in $z$-axis.
203: In Section VI we construct the magnetic knot bending
204: the helical vortex and connecting the periodic ends
205: together, and show that the knot topology $\pi_3(S^2)$
206: is described by the Chern-Simon index of the
207: electromagnetic potential.
208: In Section VII we consider the Josephson interaction,
209: and show that the inclusion of the Josephson interaction
210: does not affect the existence of the topological objects in
211: two-gap superconductor but alter the shape of the solutions
212: drastically. In Section VIII we
213: discuss the non-Abelian superconductivity which can describe
214: a two-gap superconductor made of two condensates which carry
215: opposite charge, and argue that the non-Abelian superconductivity
216: can be realized in liquid metallic hydrogen (LMH).
217: Finally in Section IX we discuss the physical implications of
218: our results, and discuss how one can identify these topological
219: objects in $\rm MgB_2$ and LMH.
220: 
221: \section{Effective potential of Two-gap Superconductor}
222: 
223: In mean field approximation
224: the free energy of the two-gap
225: superconductor could be expressed by \cite{baba1,maz,baba2}
226: \bea
227: &{\cal H} = \dfrac{\hbar^2}{2m_1} |(\mbox{\boldmath $\nabla$}
228: + ig \mbox{\boldmath $A$}) \tilde \phi_1|^2
229: +\dfrac{\hbar^2}{2m_2} |(\mbox{\boldmath $\nabla$}
230: + ig \mbox{\boldmath $A$}) \tilde \phi_2|^2 \nn\\
231: &+ \tilde V(\tilde \phi_1,\tilde \phi_2)
232: + \dfrac{1}{2} (\mbox{\boldmath $\nabla$}
233: \times \mbox{\boldmath $A$})^2,
234: \label{scfe1}
235: \eea
236: where $\tilde V$ is the effective potential.
237: We choose the potential to be the most general quartic
238: potential which has the $U(1)\times U(1)$ symmetry,
239: \bea
240: &\tilde V =\dfrac{\tilde{\lambda}_{11}}2|\tilde{\phi}_1|^4
241: +\tilde{\lambda}_{12}|\tilde{\phi}_1|^2|\tilde{\phi}_2|^2
242: +\dfrac{\tilde{\lambda}_{22}}2| \tilde{\phi}_2|^4 \nn\\
243: &-\tilde{\mu}_1|\tilde{\phi}_1|^2
244: -\tilde{\mu}_2|\tilde{\phi}_2|^2,
245: \label{scpot1}
246: \eea
247: where $\tilde{\lambda}_{ij}$ are the quartic coupling
248: constants and $\tilde \mu_i$ are the chemical potentials of
249: $\tilde \phi_i~(i=1,2)$. One might like to include the
250: Josephson interaction to the potential which breaks the
251: $U(1)\times U(1)$ symmetry down to $U(1)$.
252: The Josephson interaction will be discussed separately
253: in the following. But as we will see, the inclusion of the
254: Josephson interaction does not alter the qualitative features of
255: the topological objects we discuss in this paper.
256: 
257: With the normalization
258: of $\tilde \phi_1$ and $\tilde \phi_2$ to $\phi_1$ and $\phi_2$,
259: \bea
260: \phi_1=\dfrac \hbar {\sqrt{2m_1}}\tilde{\phi}_1,\;\;\;\;
261: \phi_2=\frac \hbar {\sqrt{2m_2}}\tilde{\phi}_2.
262: \eea
263: one can simplify the above Hamiltonian (\ref{scfe1}) to
264: \bea
265: &{\cal H} = |(\mbox{\boldmath $\nabla$}
266: + ig \mbox{\boldmath $A$}) \phi|^2 + V(\phi_1,\phi_2) \nn\\
267: &+ \dfrac{1}{2} (\mbox{\boldmath $\nabla$}
268: \times \mbox{\boldmath $A$})^2,
269: \label{scfe2}
270: \eea
271: where $V$ is the normalized potential,
272: \bea
273: &V =\dfrac{\lambda_{11}}{2}|\phi_1|^4+\lambda_{12}|\phi_1|^2
274: |\phi_2|^2+\dfrac{\lambda_{22}}{2}|\phi_2|^4 \nn\\
275: &-\mu_1|\phi_1|^2-\mu_2|\phi_2|^2 \nn\\
276: &= \dfrac{\lambda_{11}}{2} \big( |\phi_1|^2
277: -\hat \phi_1^2 \big)^2 +\dfrac{\lambda_{22}}{2}\big( |\phi_2|^2
278: - \hat \phi_2^2 \big)^2 \nn\\
279: &+\lambda_{12} \big(|\phi_1|^2 -\hat \phi_1^2 \big)
280: \big(|\phi_2|^2 -\hat \phi_2^2 \big) + V_0, \nn\\
281: &\hat \phi_1^2 = \dfrac{\mu_1\lambda_{22}-\mu_2\lambda_{12}}\Delta,
282: ~~~~~\hat \phi_2^2 = \dfrac{\mu_2\lambda_{11}
283: -\mu_1\lambda_{12}}\Delta \nn\\
284: &V_0 = -\dfrac{\lambda_{11} \mu_2^2
285: +\lambda_{22} \mu_1^2 -2 \lambda_{12} \mu_1 \mu_2}{2\Delta}, \nn\\
286: &\Delta = \lambda_{11} \lambda_{22} - \lambda_{12}^2.
287: \label{scpot2}
288: \eea
289: Notice that the potential can also be written as
290: \bea
291: &V = \dfrac{1}{2\lambda_{11}} \Big[(\lambda_{11} |\phi_1|^2
292: +\lambda_{12} |\phi_2|^2 -\mu_1)^2 \nn\\
293: &+ \Delta \big( |\phi_2|^2 -\hat \phi_2^2 \big)^2 \Big] +  V_0 \nn\\
294: &= \dfrac{1}{2\lambda_{22}} \Big[(\lambda_{12} |\phi_1|^2
295: +\lambda_{22} |\phi_2|^2 -\mu_2)^2 \nn\\
296: &+ \Delta \big( |\phi_1|^2 -\hat \phi_1^2 \big)^2 \Big] +  V_0.
297: \label{scpot3}
298: \eea
299: From now on we will assume that all coupling constants
300: except $\lambda_{12}$ are positive.
301: 
302: To find the vacuum of the potential let
303: \bea
304: &\dfrac{\partial V}{\partial |\phi _1|^2}=\lambda _{11}|\phi _1|^2
305: +\lambda_{12}|\phi _2|^2-\mu _1=0, \nn\\
306: &\dfrac{\partial V}{\partial |\phi _2|^2}=\lambda _{12}|\phi _1|^2
307: +\lambda_{22}|\phi _2|^2-\mu _2=0,
308: \label{vcon}
309: \eea
310: and find the extremum
311: \bea
312: &|\phi _1|^2=\hat \phi_1^2,
313: ~~~~~|\phi_2|^2= \hat \phi_2^2.
314: \label{vac}
315: \eea
316: To check whether this extremum is the maximum or minimum,
317: consider the Hessian
318: \bea
319: &\det H=\det \dfrac{\partial^2V}
320: {\partial |\phi _i|^2\partial |\phi _j|^2}
321: =\lambda _{11}\lambda _{22}-\lambda _{12}^2 \nn\\
322: &=\Delta.
323: \label{hess}
324: \eea
325: There are three possibilities; positive,
326: zero, or negative $\Delta$. We consider each case separately.
327: 
328: A. $\Delta >0$: In this case we have
329: $\lambda_{12}^2<\lambda_{11}\lambda_{22}$,
330: and the extremum (\ref{vac}) becomes the local minimum.
331: But since $|\phi _i|$ have to be positive we have
332: the following vacuum
333: \bea
334: &\Bigg( \matrix{<|\phi_1|> \cr <|\phi_2|> } \Bigg)
335: =\Bigg( \matrix{\hat \phi_1
336: \cr \hat \phi_2 }\Bigg),
337: \label{vaca1}
338: \eea
339: for $\lambda_{12} \leq 0$ or for
340: \bea
341: &0< \lambda_{12},~~~\dfrac{\lambda_{12}}{\lambda_{22}}
342: <\dfrac{\mu _1}{\mu _2}
343: <\dfrac{\lambda _{11}}{\lambda _{12}}.
344: \label{vcona1}
345: \eea
346: Notice that both $<|\phi_1|>$ and $<|\phi_2|>$ are
347: non-vanishing. But for
348: \bea
349: &0< \lambda_{12},~~~\dfrac{\lambda_{12}}{\lambda_{22}}
350: <\dfrac{\lambda_{11}}{\lambda_{12}} <\dfrac{\mu_1}{\mu_2},
351: \label{vcona21}
352: \eea
353: we have the following vacuum from (\ref{scpot3}),
354: \bea
355: &\Bigg( \matrix{<|\phi_1|> \cr <|\phi _2|> } \Bigg)
356: =\Bigg( \matrix{\sqrt{\mu_1/\lambda_{11}} \cr 0 }\Bigg),
357: \label{vaca2}
358: \eea
359: Finally, when
360: \bea
361: &0< \lambda_{12},~~~\dfrac{\mu_1}{\mu_2}
362: < \dfrac{\lambda_{12}}{\lambda_{22}}
363: <\dfrac{\lambda_{11}}{\lambda_{12}},
364: \label{vcona22}
365: \eea
366: we can always transform this case to the case
367: (\ref{vcona21}) by re-labeling $\phi_1$ and
368: $\phi_2$ as $\phi_2$ and $\phi_1$, so that in this case
369: we can assume that the vacuum is still given by
370: (\ref{vaca2}) without loss of generality.
371: 
372: B. $\Delta =0$: In this case we have
373: $\lambda_{12}^2=\lambda_{11} \lambda_{22}$,
374: and the potential (\ref{scpot2}) is reduced to
375: \bea
376: &V =\dfrac{1}{2\lambda_{11}} \Big[ \big( \lambda_{11}|\phi_1|^2
377: +\lambda_{12} |\phi_2|^2 -\mu_1 \big)^2 \nn\\
378: &-2(\mu_2 \lambda_{11} -\mu_1 \lambda_{12}) |\phi_2|^2 \Big] \nn\\
379: &=\dfrac{1}{2\lambda_{22}} \Big[ \big(\lambda_{12}|\phi_1|^2
380: +\lambda_{22} |\phi_2|^2 -\mu_2 \big)^2 \nn\\
381: &-2(\mu_1 \lambda_{22} -\mu_2 \lambda_{12}) |\phi_1|^2 \Big].
382: \label{scpot4}
383: \eea
384: So for $\lambda_{12} <0$, the potential becomes unbounded from below,
385: so that it has no minimum. For
386: \bea
387: 0< \lambda_{12},
388: ~~~~~\dfrac{\lambda_{11}}{\lambda_{12}}
389: =\dfrac{\lambda_{12}}{\lambda_{22}}<\dfrac{\mu_1}{\mu_2},
390: \label{vconb1}
391: \eea
392: we have the following vacuum
393: \bea
394: \Bigg( \matrix{ <|\phi_1|> \cr <|\phi_2|> } \Bigg)
395: =\Bigg( \matrix{ \sqrt{\mu_1/\lambda_{11}} \cr
396: 0  } \Bigg).
397: \label{vacb1}
398: \eea
399: Next, consider the case
400: \bea
401: 0< \lambda_{12},
402: ~~~~~\dfrac{\mu_1}{\mu_2}< \dfrac{\lambda_{11}}{\lambda_{12}}
403: =\dfrac{\lambda_{12}}{\lambda_{22}}.
404: \label{vconb2}
405: \eea
406: But this can be transformed to (\ref{vconb1}) with
407: the re-labeling. So we can assume that the vacuum is given by
408: (\ref{vacb1}) without loss of generality.
409: Finally, when
410: \bea
411: \dfrac{\mu_1}{\mu_2} =\dfrac{\lambda_{11}}{\lambda_{12}}
412: =\dfrac{\lambda_{12}}{\lambda_{22}},
413: \label{vconb3}
414: \eea
415: we have the degenerate vacuum
416: \bea
417: \mu_1 <|\phi_1|>^2 + \mu_2 <|\phi_2|>^2
418: = \dfrac{\mu_1 \mu_2}{\lambda_{12}}.
419: \label{vacb3}
420: \eea
421: This case includes the special (and familiar) $SU(2)$ symmetric
422: case
423: \bea
424: &\lambda_{11} =\lambda_{12} = \lambda_{22} = \lambda,
425: ~~~u_1 =\mu_2 =\mu, \nn\\
426: &<|\phi_1|>^2 + <|\phi_2|>^2 = \dfrac{\mu}{\lambda}.
427: \label{vacsu2}
428: \eea
429: In this case the Hamiltonian (\ref{scfe2}) has the full
430: $SU(2)$ symmetry.
431: 
432: C. $\Delta <0$: In this case we have
433: $\lambda_{12}^2>\lambda_{11}\lambda_{22}$,
434: and the extremum (\ref{vac}) becomes the local maximum.
435: So the minimum state must satisfy
436: \bea
437: |\phi_1|^2|\phi_2|^2=0.
438: \eea
439: Now, by inspection one can show that when
440: \bea
441: 0< \lambda_{12}, ~~~\dfrac{\lambda_{11}}{\lambda_{12}}
442: <\sqrt{\dfrac{\lambda_{11}}{\lambda_{22}}}
443: <\dfrac{\mu_1}{\mu_2},
444: \label{vconc1}
445: \eea
446: the vacuum must be
447: \bea
448: &\Bigg( \matrix{<|\phi_1|> \cr <|\phi_2|>} \Bigg)
449: = \Bigg( \matrix{\sqrt{\mu_1/\lambda_{11}} \cr 0 } \Bigg).
450: \label{vacc1}
451: \eea
452: Next, consider the case
453: \bea
454: \dfrac{\mu_1}{\mu_2}<\sqrt{\dfrac{\lambda_{11}}{\lambda_{22}}}
455: <\dfrac{\lambda_{12}}{\lambda_{22}}.
456: \eea
457: This case can be reduced
458: to the above case (by re-labeling $\phi_1$ and $\phi_2$),
459: so that when $\lambda_{12}$ is positive
460: one can assume that the vacuum is given by (\ref{vacc1})
461: without loss of generality.
462: Finally when $\lambda_{12}$ is negative the potential
463: has no minimum, because it is unbounded from below.
464: This must be clear from (\ref{scpot4}).
465: 
466: In summary, we have three types of vacuum state: \\
467: A. Type I: Integer flux vacuum
468: \bea
469: \Bigg( \matrix{ <|\phi_1|> \cr <|\phi_2|> } \Bigg)
470: =\Bigg( \matrix{ \sqrt{\mu_1/\lambda_{11}} \cr
471: 0  } \Bigg).
472: \label{vacb}
473: \eea
474: This is possible when we have one of the following
475: three cases,
476: \bea
477: &&(a)~~~~0< \lambda_{12},~~~\dfrac{\lambda_{12}}{\lambda_{22}}
478: <\dfrac{\lambda_{11}}{\lambda_{12}} \leq \dfrac{\mu_1}{\mu_2}, \nn\\
479: &&(b)~~~~0< \lambda_{12},~~~\dfrac{\lambda_{11}}{\lambda_{12}}
480: <\sqrt{\dfrac{\lambda_{11}}{\lambda_{22}}}
481: <\dfrac{\mu_1}{\mu_2}, \nn\\
482: &&(c)~~~~0< \lambda_{12},~~~\dfrac{\lambda_{11}}{\lambda_{12}}
483: =\dfrac{\lambda_{12}}{\lambda_{22}}<\dfrac{\mu_1}{\mu_2}.
484: \eea
485: We call this integer flux vacuum because, as we will see,
486: for this type of vacuum the magnetic vortex has
487: an integer flux. \\
488: B. Type II: Fractional flux vacuum
489: \bea
490: \left( \begin{array}{l}
491: <|\phi_1|> \\ <|\phi_2|> \end{array} \right)
492: =\left( \begin{array}{l} \hat{\phi}_1 \\ \hat{\phi}_2
493: \end{array} \right).
494: \label{vaca}
495: \eea
496: This is possible when we have one of the following
497: three cases,
498: \bea
499: &&(a)~~~~\lambda_{12} <0,~~~\dfrac{|\lambda_{12}|}{\lambda_{22}}
500: <\dfrac{\lambda_{11}}{|\lambda_{12}|}, \nn\\
501: &&(b)~~~~0< \lambda_{12},~~~\dfrac{\lambda_{12}}{\lambda_{22}}
502: <\dfrac{\mu_1}{\mu_2}<\dfrac{\lambda_{11}}{\lambda_{12}}, \nn\\
503: &&(c)~~~~\lambda_{12} =0.
504: \eea
505: We call this fractional flux vacuum because, as we will see,
506: for this type of vacuum the magnetic vortex has
507: a fractional flux. \\
508: C. Type III: Degenerate vacuum
509: \bea
510: \mu_1 <|\phi_1|>^2 + \mu_2 <|\phi_2|>^2 = \dfrac{\mu_1 \mu_2}{\lambda_{12}}.
511: \label{vacc}
512: \eea
513: This is what we have when
514: \bea
515: \dfrac{\mu_1}{\mu_2}=\dfrac{\lambda_{12}}{\lambda_{11}}
516: =\dfrac{\lambda_{22}}{\lambda_{12}}.
517: \eea
518: Notice that the potential (\ref{scpot2}) has no vacuum
519: when
520: \bea
521: \lambda_{12}< 0,~~~\dfrac{\lambda_{11}}{|\lambda_{12}|}
522: \leq \dfrac{|\lambda_{12}|}{\lambda_{22}}.
523: \eea
524: All other cases can be reduced to one of the above cases
525: by re-labelling $\phi_1$ and $\phi_2$.
526: As we will see the vacuum structure will play an important
527: role in the following.
528: 
529: Notice that with
530: \bea
531: &\lambda =\dfrac{\lambda_{11}+\lambda _{22}+2\lambda _{12}}4, \nn\\
532: &\alpha =\dfrac{\lambda_{11}-\lambda_{22}}2,
533: ~~~\beta =\dfrac{\lambda_{11}+\lambda_{22}-2\lambda_{12}}4, \nn\\
534: &\mu =\dfrac{\mu_1+\mu_2}2,~~~\gamma =\dfrac{\mu _1-\mu _2}2,
535: \eea
536: we have
537: \bea
538: &\lambda_{11}=\lambda+\beta+\alpha,
539: ~~~~~\lambda_{22}=\lambda+\beta-\alpha,  \nn\\
540: &\lambda_{12}=\lambda-\beta, \nn\\
541: &\mu_1=\mu+\gamma,~~~~\mu_2=\mu-\gamma,
542: \eea
543: so that the potential (\ref{scpot2}) can be written as
544: \bea
545: &V= \dfrac {\lambda} 2(|\phi_1|^2+|\phi_2|^2-\dfrac{\mu}\lambda)^2
546: +\dfrac {\alpha}2 (|\phi_1|^4-|\phi_2|^4) \nn\\
547: &+\dfrac {\beta}2 (|\phi_1|^2-|\phi_2|^2)^2
548: -\gamma (|\phi_1|^2-|\phi_2|^2)-\dfrac{\mu^2}{2\lambda}.
549: \label{scpot5}
550: \eea
551: In terms of the new parameters we have
552: \bea
553: &\hat \phi_1^2= \dfrac {2 (\beta\mu+\gamma\lambda)
554: -\alpha(\mu+\gamma)}{\Delta}, \nn\\
555: &\hat \phi_2^2= \dfrac {2 (\beta\mu-\gamma\lambda)
556: +\alpha(\mu-\gamma)}{\Delta},
557: \eea
558: where $\Delta=4 \beta\lambda-\alpha^2$.
559: 
560: \section{Dynamics of two-gap condensates}
561: 
562: With this preliminary one may study the topological objects
563: of two-gap superconductor minimizing the free energy.
564: On the other hand, to study a static solution,
565: one might as well start from the following relativistic
566: Ginzburg-Landau Lagrangian
567: \bea
568: &{\cal L} = - |D_\mu \phi|^2 + V(\phi)
569: - \dfrac{1}{4} F_{\mu \nu}^2, \nn\\
570: &D_\mu \phi = (\partial_\mu + ig A_\mu) \phi,
571: \label{sclag}
572: \eea
573: which reproduces the the free energy (\ref{scfe2}) in
574: the static limit.
575: The Lagrangian has the equation of motion
576: \bea
577: &D^2\phi_1 = \dfrac{\partial V}{\partial |\phi_1|^2} \phi_1, \nn\\
578: &D^2\phi_2 = \dfrac{\partial V}{\partial |\phi_2|^2} \phi_2, \nn\\
579: &\partial_\mu F_{\mu \nu} = j_\nu =  i g \Big[(D_\nu
580: \phi)^{\dagger}\phi - \phi ^{\dagger}(D_\nu \phi) \Big].
581: \label{sceq1}
582: \eea
583: To understand the meaning of this we let
584: \bea
585: &\phi =\dfrac{1}{\sqrt 2} \rho \xi,
586: ~~~~~|\phi^{\dagger}\phi|=\dfrac{\rho}{2},
587: ~~~~~{\xi}^{\dagger}\xi = 1, \nn\\
588: &\hat n = \xi^{\dagger} \vec \sigma \xi,
589: \label{ndef}
590: \eea
591: and find the following identities
592: \bea
593: & (\pro_\mu \hn)^2 = 4 \big(|\pro_\mu \xi|^2
594: -|\xi^\dag \pro_\mu \xi|^2 \big), \nn\\
595: & -\dfrac{1}{g} \hn \cdot (\pro_\mu \hn \times \pro_\nu \hn)
596: = \dfrac{2i}{g} (\pro_\mu \xi^\dag \pro_\nu \xi
597: - \pro_\nu \xi^\dag \pro_\mu \xi ) \nn\\
598: &= \pro_\mu C_\nu - \pro_\nu C_\mu, \nn\\
599: &\Big[\pro_\mu  +\dfrac{1}{2} (ig C_\mu - \vec \sigma
600: \cdot \pro_\mu \n) \Big] \xi =0, \nn\\
601: &C_\mu =\dfrac{2i}g\xi ^{\dagger }\partial _\mu \xi.
602: \label{id}
603: \eea
604: From these we can reduce (\ref{sceq1}) to \cite{ijpap,prb05}
605: \begin{eqnarray}
606: &\partial ^2\rho -\Big( \dfrac 14(\partial _\mu \hn)^2+g^2(A_\mu
607: -\dfrac 12C_\mu )^2\Big) \rho   \nonumber \\
608: &=\Big[\dfrac \lambda 2 (\rho ^2- \bar \rho^2)
609: +\big(\dfrac \alpha 2\rho^2-\gamma \big) n_3
610: +\dfrac \beta 2\rho^2 n_3^2 \Big] \rho, \nonumber \\
611: &\hat{n}\times \partial ^2\hat{n}+2\dfrac{\partial _\mu \rho }%
612: \rho \hat{n}\times \partial _\mu \hat{n}-\dfrac 2{g\rho^2}
613: \partial_\mu F_{\mu \nu }\partial_\nu \hat{n}  \nonumber \\
614: &=\Big(2\gamma -(\dfrac{\alpha}2 + \beta n_3) \rho^2 \Big)
615: \hat{k}\times \hat{n},  \nonumber \\
616: &\partial_\mu F_{\mu \nu }=j_\nu =g^2\rho ^2\Big(A_\nu
617: -\dfrac12 C_\nu \Big),  \nonumber \\
618: &\bar \rho^2=\dfrac{2\mu}\lambda.
619: \label{sceq2}
620: \end{eqnarray}
621: This is the equation for two-gap superconductor,
622: which allows a large class of interesting topological
623: objects, straight magnetic vortex, helical magnetic vortex,
624: and magnetic knot, all with $4\pi/g$-flux, $2\pi/g$-flux,
625: or fractional flux.
626: 
627: The equation (\ref{sceq1}) is an equation of the complex doublet
628: $\phi$ which has four degrees. But notice that the equation
629: (\ref{sceq2}) is, except for $C_\mu$, expressed
630: completely in terms of the $CP^1$ field $\hn$ and
631: the scalar field $\rho$. Moreover,
632: (\ref{id}) tells that $C_\mu$ can also be written in terms of
633: $\hn$. In fact $\hn$ uniquely defines a righthanded
634: orthonormal frame ($\hat n_1,\hat n_2,\hat n$), with
635: $\hat n_1 \times \hat n_2=\hat n$, up to the $U(1)$ rotation
636: which leaves $\hat n$ invariant. Then $C_\mu$ is given (up to a $U(1)$
637: gauge transformation) by the Mermin-Ho relation \cite{ho,cho79,cho80,cho81}
638: \bea
639: &C_\mu = -\dfrac{1}{g} \hat n_1 \cdot \pro_\mu \hat n_2, \nn\\
640: &\pro_\mu C_\nu - \pro_\nu C_\mu
641: =-\dfrac{1}{g} \hn \cdot (\pro_\mu \hn \times \pro_\nu \hn).
642: \eea
643: This tells that we can transform the equation (\ref{sceq1})
644: of the complex doublet condensate $\phi$ to
645: the equation (\ref{sceq2}) of the $CP^1$ field $\hn$
646: and the scalar field $\rho$. In fact, with
647: \bea
648: &B_\mu= A_\mu-\dfrac12 C_\mu, \nn\\
649: &G_{\mu\nu}= \pro_\mu B_\nu - \pro_\nu B_\mu,
650: \label{bm}
651: \eea
652: we can express (\ref{sceq2}) completely in terms of $\hn$,
653: $\rho$, and $B_\mu$. This is not accidental.
654: Indeed with (\ref{ndef}), (\ref{id}), and (\ref{bm}),
655: we can express the Hamiltonian (\ref{scfe2}) as
656: \bea
657: &{\cal H}=\dfrac 12 (\partial_\mu \rho)^2 + \dfrac 12 g^2\rho^2 B^2_\mu
658: +\dfrac 18\rho^2(\partial_\mu \hat n)^2 + V(\rho,n_3) \nn\\
659: &+\dfrac 14 \big[G_{\mu\nu}-\dfrac {1}{2g} \hat n
660: \cdot (\partial_\mu \hat n \times \partial_\nu \hat n)  \big]^2,
661: \label{scfe3}
662: \eea
663: where
664: \bea
665: &V(\rho,n_3)=\dfrac \lambda 8 \big(1 +\dfrac{\alpha}\lambda n_3
666: +\dfrac{\beta}\lambda  n_3^2 \big) \rho^4 \nn\\
667: &-\dfrac\mu 2 \big(1+\dfrac{\gamma}\mu n_3 \big) \rho^2,
668: \eea
669: and $n_3 = \xi_1^* \xi_1 - \xi_2^* \xi_2$.
670: This means that the Ginzburg-Landau
671: theory of two-gap superconductor can be understood
672: as a theory of $CP^1$ field $\hn$ (coupled to $\rho$
673: and $B_\mu$) \cite{ijpap,prb05,baba1,cm2}.
674: This is because the $U(1)$ gauge invariance of (\ref{sclag})
675: reduces the physical degrees
676: of the complex doublet $\phi$ to $\rho$ and $\hn$, and
677: the massive photon $B_\mu$.
678: As we will see, this has a very important physical implication,
679: because this tells that the topology of two-gap superconductor
680: can be described by the topology of $\hn$ and $A_\mu$.
681: 
682: The equation (\ref{sceq2}) allows two conserved currents,
683: the electromagnetic current $j_\mu$ and the
684: neutral current $k_\mu$ \cite{cm2},
685: \bea
686: &j_\mu = g^2 \rho^2(A_\mu -\dfrac 12C_\mu) \nn\\
687: &k_\mu = g^2 \rho^2 \Big[ A_\mu \big(\xi_1^{*}\xi_1
688: -\xi_2^{*}\xi_2 \big)
689: +\dfrac{i}{g} \big(\partial_\mu \xi_1^{*}\xi_1 \nn\\
690: &+\xi_2^{*}\partial_\mu \xi_2 \big) \Big],
691: \label{2sc1}
692: \eea
693: which are nothing but the Noether
694: currents of the $U(1)\times U(1)$ symmetry of the Hamiltonian
695: (\ref{scfe2}). Indeed they are
696: the sum and difference of two electromagnetic currents of
697: $\phi_1$ and $\phi_2$
698: \bea
699: j_\mu = j^{(1)}_\mu + j^{(2)}_\mu,
700: ~~~~~k_\mu = j^{(1)}_\mu - j^{(2)}_\mu.
701: \label{2sc2}
702: \eea
703: Clearly the conservation of $j_\mu$ follows from
704: the last equation of (\ref{sceq2}). But the conservation of
705: $k_\mu$ comes from the second equation of (\ref{sceq2}),
706: which (together with the last equation) tells the
707: existence of a partially conserved $SU(2)$ current
708: $\vec j_\mu$ \cite{cm2},
709: \bea
710: \vec j_\mu = g\rho^2 \Big(\frac 12\vec{n} \times \partial_\mu
711: \vec{n} -g(A_\mu -\frac 12C_\mu )\vec{n} \Big).
712: \eea
713: This $SU(2)$ current is exactly conserved when
714: $\alpha=\beta=\gamma=0$. But notice that
715: $k_\mu=\hat k \cdot \vec j_\mu$.
716: This assures that we have the conservation of $k_\mu$
717: even when $\alpha \beta \gamma \neq 0$. It is interesting to notice
718: that $j_\mu$ and $k_\mu$ are precisely the $\hn$ and $\hat k$
719: components of $\vec j_\mu$.
720: 
721: \section{Non-Abrikosov Vortex}
722: 
723: The two-gap superconductor allows different types of
724: interesting magnetic vortex \cite{ijpap,prb05,cm2,baba2}.
725: In terms of the structure of the vortex
726: they are classified to two types, the D-type vortex
727: which has no concentration of the condensates at the core
728: and the N-type which has a non-vanishing concentration of
729: the condensates at the core.
730: The reason for this is that, unlike
731: the Abrikosov vortex in ordinary superconductor,
732: the vortex in two-gap superconductor allows two different
733: boundary conditions at the core.
734: 
735: To discuss the straight vortex let
736: $(\varrho,\varphi,z)$ be the cylindrical coordinates and
737: choose the ansatz
738: \begin{eqnarray}
739: &\rho =\rho (\varrho ),~~~~~\xi =\left(
740: \begin{array}{c}
741: \cos \dfrac{f(\varrho )}2\exp (-in\varphi ) \\
742: \sin \dfrac{f(\varrho )}2
743: \end{array} \right) , \nn\\
744: &A_\mu =\dfrac ngA(\varrho)\partial_\mu \varphi.
745: \label{svans}
746: \end{eqnarray}
747: With this we have
748: \bea
749: &\hat{n}=\xi ^{\dag }\vec{\sigma}\xi =\left(
750: \begin{array}{c}
751: \sin f(\varrho )\cos n\varphi  \\
752: \sin f(\varrho )\sin n\varphi  \\
753: \cos f(\varrho )
754: \end{array} \right),  \nonumber \\
755: &C_\mu =n \dfrac{\cos {f(\varrho )}+1}g \partial_\mu \varphi, \nn\\
756: & j_\mu =ng\rho^2\Big(A-\dfrac{\cos {f}+1}{2} \Big)
757:  \partial_\mu \varphi,  \nn\\
758: & k_\mu = ng\rho^2\Big(A \cos f-\dfrac{\cos {f}+1}{2} \Big)
759: \partial_\mu \varphi,
760: \label{sc}
761: \eea
762: and the Hamiltonian (\ref{scfe3}) becomes
763: \begin{eqnarray}
764: &{\cal H} =\dfrac 12\dot{\rho}^2+\dfrac 18\rho
765: ^2 \Big(\dot{f}^2+\dfrac{n^2}{\varrho^2}\sin ^2f \Big)  \nn\\
766: &+ \dfrac{n^2\rho
767: ^2}{2\varrho^2}\Big( A - \dfrac {\cos f+1}{2} \Big) ^2
768: +\dfrac{n^2}{2g^2\varrho^2}\dot{A}^2 \nn \\
769: &+\dfrac \lambda 8  \Big[(\rho ^2-\bar \rho^2)^2
770: +\dfrac{\alpha}\lambda (\rho^2-\dfrac{4\gamma }\alpha) \rho^2 \cos f \nn\\
771: &+\dfrac{\beta}\lambda \rho^4\cos ^2f \Big]
772: -\dfrac{\mu^2}{2\lambda}.
773: \label{scfe4}
774: \end{eqnarray}
775: With this (\ref{sceq2}) becomes
776: \begin{eqnarray}
777: &\ddot{\rho}+\dfrac{1}{\varrho}\dot{\rho}-\Big[\dfrac 14
778: \Big(\dot{f}^2+\dfrac{n^2}{\varrho^2}\sin^2f\Big) \nn\\
779: &+\dfrac{n^2}{\varrho^2}
780: \Big(A-\dfrac{\cos {f}+1}{2} \Big)^2\Big]\rho   \nonumber \\
781: &=\dfrac \lambda 2  \Big[(\rho^2- \bar \rho^2)
782: +\dfrac{\alpha}\lambda (\rho^2-\dfrac{2\gamma }\alpha)\cos f
783: +\dfrac{\beta}\lambda \rho^2\cos ^2f \Big]\rho,  \nonumber \\
784: &\ddot{f}+\Big(\dfrac{1}{\varrho}+2\dfrac{\dot{\rho}}\rho %
785: \Big)\dot{f}-2\dfrac{n^2}{\varrho^2}\Big(A-\dfrac 12\Big)%
786: \sin f  \nonumber \\
787: &=\Big(2\gamma -(\dfrac \alpha 2
788: +\beta \cos f)\rho^2 \Big) \sin f,  \nonumber \\
789: &\ddot{A}-\dfrac{1}{\varrho}\dot{A}-g^2\rho ^2\Big(A-\dfrac{\cos {f}+1}{2} \Big)=0.
790: \label{sveq}
791: \end{eqnarray}
792: Notice that this can also be derived by minimizing the
793: Hamiltonian (\ref{scfe4}).
794: 
795: To solve the equation, we have to fix the boundary
796: conditions. To determine the possible boundary condition
797: at the core we expand $\rho (\varrho)$, $f(\varrho)$,
798: and $A(\varrho)$ near the origin as
799: \begin{eqnarray}
800: &\rho (\varrho) \simeq \rho _0+\rho _1\varrho
801: +\rho _2\varrho^2+\rho _3\varrho^3+..., \nn\\
802: &f(\varrho) \simeq f_0+f_1\varrho+f_2\varrho^2+f_3\varrho^3+..., \nn\\
803: &A(\varrho) \simeq a_0+a_1\varrho+a_2\varrho^2+a_3\varrho^3+...,
804: \end{eqnarray}
805: and find that the smoothness at the core requires
806: \begin{widetext}
807: \bea
808: &\dfrac{\rho_0}4 \Big( (2a_0- \cos f_0 -1)^2+\sin^2 f_0 \Big)
809: \dfrac{1}{\varrho^2} + \Big[ \rho_0 \Big( \dfrac12 (2a_0-1)
810: f_1 \sin f_0 + (2 a_0 -\cos f_0-1) a_1 \Big) \nn\\
811: &+\dfrac{\rho_1}4 \Big((2a_0- \cos f_0 -1)^2+\sin^2 f_0
812: -\dfrac{4}{n^2} \Big) \Big] \dfrac1\varrho...
813: +\Big[ ...+ \dfrac{\rho_k}4 \Big((2a_0- \cos f_0 -1)^2+\sin^2 f_0
814: -\dfrac{4 k^2}{n^2} \Big) \Big] \varrho^{k-2}+...=0, \nn\\
815: &\left \{ \matrix {\dfrac{(2a_0-1)\sin f_0}{\varrho^2}
816: +\Big(\big((2a_0-1)\cos f_0-\dfrac{1}{n^2} \big) f_1 +2a_1\sin f_0\Big)
817: \dfrac 1\varrho +...=0,~~~\rho_0\neq 0  \cr
818: \dfrac{(2a_0-1)\sin f_0}{\varrho^2}
819: +\Big(\big((2a_0-1)\cos f_0-\dfrac{3}{n^2} \big) f_1 +2a_1\sin f_0\Big)
820: \dfrac 1\varrho +...=0,~~~\rho_0=0,~\rho_1\neq 0 \cr
821: .  \cr .  \cr .  \cr
822: \dfrac{(2a_0-1)\sin f_0}{\varrho^2}
823: +\Big(\big((2a_0-1)\cos f_0-\dfrac{2k+1}{n^2} \big) f_1 +2a_1\sin f_0\Big)
824: \dfrac 1\varrho +...=0,~~~\rho_0=\rho_1=...\rho_{k-1}=0,
825: ~\rho_k\neq 0 } \right. \nn\\
826: &\dfrac{a_1}\varrho+\dfrac{g^2\rho_0^2}{2}
827: \big(2a_0-\cos f_0 -1 \big)
828: -\Big(3a_3-g^2\rho_0 \rho_1(2a_0-\cos f_0-1) \nn\\
829: &-g^2\rho_0^2(a_1+\dfrac 12 f_1\sin f_0) \Big) \varrho+...=0.
830: \label{sccbc}
831: \eea
832: \end{widetext}
833: Now, consider the case $\rho_0\neq 0$ first.
834: In this case the last equation
835: requires $a_1=0$ and $2a_0-1=\cos f_0$,
836: and the first equation requires
837: $\sin f_0=0$ and $\rho_1=0$. With $\sin f_0=0$
838: we must either have $f_0=\pi$ and $a_0=0$ or $f_0=0$ and $a_0=1$.
839: And with $f_0=\pi$, we have $f_1=0$
840: for $n \neq \pm 1$ from the second equation.
841: One might choose $f_0=0$ instead of
842: $f_0=\pi$, but this does not lead us to a new solution.
843: Next, consider the case $\rho_0=0,~\rho_1 \neq 0$.
844: In this case the first two sets of equations tells
845: that we may have $f_0=\pi$, $a_0=\pm 1/n$, and $f_1=0$
846: for $n\neq \mp 1$ or $n\neq \pm 3$.
847: Similarly, for $\rho_0=\rho_1=...=\rho_{k-1}=0,~\rho_k \neq 0$,
848: (\ref{sccbc}) tells that we may have $f_0=\pi$, $a_0=\pm k/n$,
849: and $f_1=0$ for $n\neq \mp 1$ or $n\neq \pm (2k+1)$.
850: This tells that we can choose the following boundary condition at
851: the core for the vortex described by the ansatz (\ref{svans})
852: in two-gap superconductor \cite{ijpap,prb05,cm2}: \\
853: A. Dirichlet boundary condition
854: \bea
855: &\rho(0)=0,~~~\dot{\rho}(0) \neq 0, \nn\\
856: &A(0)=-\dfrac{1}{n},~~~f(0)=\pi,~~~\dot f=0 ~for ~n \neq 1.
857: \label{dbc}
858: \eea
859: In particular, for $n=1$ we must have $A(0)=-1$.
860: In general, the Dirichlet boundary
861: condition can be written as
862: \bea
863: &\rho(0)=\dfrac{d \rho}{d \varrho}(0)
864: =...=\dfrac{d^{k-1} \rho}{d \varrho^{k-1}}(0)=0,
865: ~~~\dfrac{d^k \rho}{d \varrho^k}(0) \neq 0, \nn\\
866: &A(0)=-\dfrac{k}{n},~~~f(0)=\pi,~~~\dot f=0 ~for ~n \neq 1,
867: \label{dbcg}
868: \eea
869: so that for $n=1$ we must have $A(0)=-k$. \\
870: B. Neumann boundary condition
871: \bea
872: &\rho(0)\neq 0,~~~\dot{\rho}(0)=0, \nn\\
873: &A(0)=0,~~~f(0)=\pi,~~~\dot f=0 ~for ~n \neq 1.
874: \label{nbc}
875: \eea
876: So here we have $A(0)=0$ for all $n$.
877: This shows that the magnetic vortex in two-gap superconductor
878: allows two types of boundary condition which are different
879: from what we have in ordinary superconductor.
880: This is a new feature of two-gap superconductor
881: which will have a deep impact in the following.
882: 
883: To determine the boundary condition at the infinity
884: notice that at the infinity all fields must assume the
885: vacuum values. In particular, the electromagnetic current
886: must vanish at the infinity.
887: This means that we must have
888: \begin{eqnarray}
889: &\rho(\infty)= \sqrt{2(<|\phi_1|>^2+<|\phi_2|>^2)},\nn\\
890: &\cos f(\infty)=<n_3>=\dfrac{<|\phi_1|>^2-<|\phi_2|>^2}
891: {<|\phi_1|>^2+<|\phi_2|>^2}, \nn\\
892: & A(\infty)=\dfrac {\cos f(\infty)+1}{2} \nn\\
893: &=\dfrac {<|\phi_1|>^2} {<|\phi_1|>^2+<|\phi_2|>^2}.
894: \label{bcinf}
895: \end{eqnarray}
896: Notice that for the integer flux vacuum we have $A(\infty)=1$,
897: but for the fractional flux vacuum $A(\infty)$ becomes
898: fractional.
899: 
900: At this point one might worry about the apparent singularity
901: at the core in the gauge potential when $A(0) \neq 0$.
902: But this singularity is a coordinate singularity
903: which can easily be removed by a gauge transformation.
904: Indeed one can always choose a gauge where $A(0)$ becomes zero
905: to remove the coordinate singularity. But notice that
906: the gauge transformation also changes $A(\infty)$
907: by the same amount, leaving $A(\infty)-A(0)$ invariant.
908: 
909: \begin{figure}[t]
910: \includegraphics[scale=0.4]{sv4pi.eps}
911: \caption{The D-type straight vortex with $n=1$ and $k=1$
912: which has $4\pi/g$ flux.
913: Three solutions are shown: $\alpha=\beta=\gamma=0$ (solid
914: lines), $\alpha=\beta=0$, $\gamma =0.005$ (dashed lines),
915: and $\alpha=\gamma=0$, $\beta =-0.005$ (dotted lines).
916: Here the unit of the scale is $1/{\bar \rho}$ and
917: we have put $\lambda/g^2=2$.}
918: \label{sv4pi}
919: \end{figure}
920: 
921: The existence of two types of boundary conditions
922: in two-gap superconductor has an important impact.
923: To understand this notice that the magnetic flux of vortex
924: is given by
925: \begin{eqnarray}
926: \Phi = \oint A_\mu dx^\mu = \big( A(\infty)-A(0) \big)
927: \dfrac {2 \pi n}{g}.
928: \label{mflux}
929: \end{eqnarray}
930: Now, it is clear that the magnetic flux becomes fractional
931: when $A(\infty)$ is fractional, which happens when
932: $<n_3> \neq 1$ (or equivalently $<|\phi_2|>\neq 0$).
933: As importantly, when $A(\infty)=1$ the magnetic flux becomes
934: $2\pi(n+k)/g$ with $A(0)=-k/n$. This was impossible
935: in ordinary superconductor.
936: Now we classify the magnetic vortex in terms of the flux.
937: 
938: \subsection{$4\pi/g$-flux vortex}
939: 
940: Let us choose the Dirichlet boundary condition
941: at the core and the integer flux vacuum
942: at the infinity \cite{ijpap,cm2}. With $k=1$ we require
943: \begin{eqnarray}
944: &\rho (0)=0,~~\dot \rho (0) \neq 0,~~~\rho (\infty )
945: =\sqrt{\dfrac{2(\mu +\gamma )}{(\lambda +\alpha+\beta )}}, \nn\\
946: &f(0)=\pi,~~~~~f(\infty ) =0, \nn\\
947: &A(0)=-\dfrac{1}{n},~~~~~A(\infty )=1.
948: \label{dbc4pi}
949: \end{eqnarray}
950: With this we can integrate (\ref{sveq}) to find the vortex solutions.
951: The solutions with $n=1$ with different parameters
952: are shown in Fig.~\ref{sv4pi}.
953: We call this a D-type vortex, because this comes from
954: the Dirichlet boundary condition at the core.
955: Both $\phi_1$ and $\phi_2$ start from zero at the core.
956: However, notice that $\phi_1$ approaches the finite vacuum value
957: but $\phi_2$ approaches zero at the infinity. So $\phi_2$
958: has a maximum concentration at a finite distance
959: from the core. This is a generic feature of a D-type vortex.
960: 
961: \begin{figure}[t]
962: \includegraphics[scale=0.4]{sv6pi.eps}
963: \caption{The D-type straight vortex with $n=1$ and $k=2$
964: which has $6\pi/g$ flux.
965: Three solutions are shown: $\alpha=\beta=\gamma=0$ (solid lines),
966: $\alpha=\beta=0$, $\gamma =0.005$ (dashed lines), and
967: $\alpha=\gamma=0$, $\beta =-0.005$ (dotted lines). Here the unit
968: of the scale is $1/{\bar \rho}$ and we have put $\lambda/g^2=2$.}
969: \label{sv6pi}
970: \end{figure}
971: 
972: With (\ref{dbc4pi}) the magnetic flux is given by
973: \begin{eqnarray}
974: &\Phi =\dfrac{}{} \int F_{\varrho \varphi}d^2x
975: =\dfrac{}{} \int \partial_\varrho A_\varphi d^2x \nn\\
976: &=\Big(1+ \dfrac{1}{n} \Big) \dfrac{2\pi n}g=\dfrac{2\pi}{g}(n+1),
977: \end{eqnarray}
978: so that when $n=1$ the vortex has $4\pi/g$ flux.
979: Moreover, the solution has a non-Abelian topology. To see this notice that
980: $\hn$ defines a mapping $\pi_2(S^2)=n$ from the compactifed
981: $xy$-plane $S^2$ to the $CP^1$ space $S^2$.
982: Clearly this is non-Abelian.
983: 
984: In general we may require
985: \bea
986: &\rho(0)=\dfrac{d \rho}{d \varrho}(0)
987: =...=\dfrac{d^{k-1} \rho}{d \varrho^{k-1}}(0)=0, \nn\\
988: &\dfrac{d^k \rho}{d \varrho^k}(0) \neq 0,~~~\rho(\infty )
989: =\sqrt{\dfrac{2(\mu +\gamma )}{(\lambda +\alpha+\beta )}}, \nn\\
990: &f(0)=\pi,~~~~~f(\infty ) =0, \nn\\
991: &A(0)=-\dfrac{k}n,~~~~~A(\infty )=1,
992: \eea
993: and obtain a different D-type vortex whose magnetic
994: flux is given by
995: \bea
996: &\Phi =\Big(1+ \dfrac{k}{n} \Big) \dfrac{2\pi n}g=\dfrac{2\pi}{g}(n+k).
997: \eea
998: The $6\pi/g$-flux vortex with $n=1$
999: and $k=2$ is shown in Fig. \ref{sv6pi}. This tells that there exist
1000: infinitely many D-type vortices which have the same
1001: topology $\pi_2(S^2)=n$. Again this
1002: is completely unexpected.
1003: 
1004: \subsection{$2\pi/g$-flux vortex}
1005: 
1006: Now we choose the Neumann boundary condition
1007: at the core and the integer flux vacuum at the infinity \cite{ijpap,cm2},
1008: \begin{eqnarray}
1009: &\rho(0) \neq 0,~~~\dot \rho(0)=0,~~~\rho (\infty)
1010: =\sqrt{\dfrac{2(\mu +\gamma )}{(\lambda +\alpha +\beta )}}, \nn\\
1011: &f(0)=\pi,~~~~~f(\infty ) =0, \nn\\
1012: &A(0)=0,~~~~~A(\infty )=1,
1013: \end{eqnarray}
1014: and find the vortex solutions. The solutions with $n=1$
1015: but with different parameters are shown in Fig.~\ref{sv2pi}.
1016: We call this a N-type vortex, because this comes from
1017: the Neumann boundary condition at the core.
1018: In this case $\phi_1$ behavior is the same as before.
1019: But notice that $\phi_2$ has a maximum concentration at the core,
1020: and approaches zero at the infinity.
1021: This is a generic feature of a N-type vortex.
1022: 
1023: \begin{figure}[t]
1024: \includegraphics[scale=0.4]{sv2pi.eps}
1025: \caption{The N-type straight vortex with $n=1$ and $2\pi/g$ flux.
1026: Three solutions are shown: $\alpha=\beta=\gamma=0$ (solid
1027: lines), $\alpha=\beta=0$, $\gamma =0.005$ (dashed lines),
1028: and $\alpha=\gamma=0$, $\beta =-0.005$ (dotted lines).
1029: Here the unit of the scale is $1/{\bar \rho}$ and
1030: we have put $\lambda/g^2=2$.}
1031: \label{sv2pi}
1032: \end{figure}
1033: 
1034: The magnetic flux of the vortex is given by
1035: \begin{eqnarray}
1036: &\Phi =\dfrac{}{} \int F_{\varrho \varphi}d^2x
1037: = \dfrac{}{} \int \partial_\varrho A_\varphi d^2x \nn\\
1038: &=\dfrac{2\pi n}g,
1039: \end{eqnarray}
1040: so that it has the same flux as the Abrikosov vortex.
1041: But notice that the topology of the $CP^1$ field $\hn$ is
1042: still non-Abelian as before, $\pi_2(S^2)=n$.
1043: The reason why there exist two types of vortices which have
1044: different magnetic fluxes but have the same topology
1045: is that the magnetic flux is determined by
1046: the boundary condition $A(\infty)-A(0)$, not by
1047: the topology. The topology assures only the quantization of the flux,
1048: and does not determine what is the unit flux quantum.
1049: 
1050: \subsection{Fractional flux vortex}
1051: 
1052: This is possible when we have the fractional flux vacuum
1053: at infinity
1054: \begin{eqnarray}
1055: &\rho (\infty )=2\sqrt{\dfrac{2 \beta \mu -\alpha \gamma}
1056: {4\beta \lambda-\alpha^2}},
1057: ~~~\cos f(\infty )=\dfrac{2\gamma \lambda -\alpha \mu}
1058: {2\beta \mu-\alpha \gamma}, \nn\\
1059: &A(\infty )=\dfrac 12 \dfrac{2(\gamma \lambda + \beta \mu)
1060: -\alpha(\mu+\gamma)}{2\beta \mu-\alpha \gamma}.
1061: \end{eqnarray}
1062: At the core we can impose either the Dirichlet condition
1063: (\ref{dbc}) or the Neumann condition (\ref{nbc}).
1064: 
1065: \begin{figure}[t]
1066: \includegraphics[scale=0.4]{sf4pi.eps}
1067: \caption{The D-type straight vortices with $n=1$  and $k=1$
1068: which have a fractional flux with
1069: $\alpha=\gamma=0,~\beta=1.0$ (solid lines),
1070: $\alpha=0,~\beta=\lambda,~\gamma=-0.2$ (dashed lines),
1071: and $\alpha=-0.25,~\beta=\lambda,~\gamma=0$ (dotted lines).
1072: Here the unit of the scale is $1/{\bar \rho}$ and
1073: we have put $\lambda/g^2=2$.}
1074: \label{sf4pi}
1075: \end{figure}
1076: 
1077: We consider two special cases: \\
1078: 1. $\lambda _{11}=\lambda _{22}$ ($\alpha =0$).
1079: In this case we have
1080: \begin{eqnarray}
1081: \rho(\infty )=\sqrt {\dfrac{2\mu}{\lambda}},
1082: ~~~~\cos f(\infty)=\dfrac{\gamma \lambda}{\beta \mu}.
1083: \end{eqnarray}
1084: So with the Dirichlet boundary condition at the core
1085: the magnetic flux is given by
1086: \begin{equation}
1087: \Phi=\Big(\dfrac{\gamma \lambda}
1088: {2\beta \mu} +\dfrac12+\dfrac{k}{n} \Big) \dfrac{2\pi n}{g},
1089: \end{equation}
1090: but with the Neumann boundary condition at the core
1091: the magnetic flux is given by
1092: \begin{equation}
1093: \Phi=\Big(\dfrac{\gamma \lambda}{2\beta \mu} +\dfrac12 \Big)
1094: \dfrac{2\pi n}{g}.
1095: \end{equation}
1096: Clearly they are fractional. \\
1097: 2. $\lambda_{12}=0$ ($\lambda=\beta$). In this case two
1098: condensates $\phi_1$ and $\phi_2$ have no direct coupling,
1099: and we have
1100: \begin{equation}
1101: <|\phi _1|> =\sqrt{\dfrac{\mu_1}{\lambda_{11}}},
1102: ~~~~~<|\phi _2|> =\sqrt{\dfrac{\mu _2}{\lambda_{22}}}.
1103: \end{equation}
1104: With this we have
1105: \begin{eqnarray}
1106: &\rho(\infty)=2 \sqrt {\dfrac{2\lambda \mu-\alpha\gamma}
1107: {4\lambda ^2-\alpha^2}},
1108: ~~~\cos f(\infty )=\dfrac{2\gamma \lambda -\alpha \mu}
1109: {2\lambda \mu-\alpha \gamma}, \nn\\
1110: &A(\infty)=\dfrac 12 \dfrac{(2\lambda-\alpha)(\mu+\gamma)}
1111: {2\mu \lambda -\alpha \gamma}.
1112: \end{eqnarray}
1113: So with the Dirichlet boundary condition at the core
1114: the magnetic flux is given by
1115: \begin{equation}
1116: \Phi=\Big(\dfrac 12 \dfrac{(2\lambda-\alpha)(\mu+\gamma)}
1117: {2\lambda \mu-\alpha \gamma}
1118: +\dfrac{k}{n} \Big) \dfrac{2\pi n}{g},
1119: \end{equation}
1120: but with the Neumann boundary condition at the core
1121: the magnetic flux is given by
1122: \begin{equation}
1123: \Phi=\dfrac 12 \dfrac{(2\lambda-\alpha)(\mu+\gamma)}
1124: {2\lambda \mu-\alpha \gamma} \dfrac{2\pi n}{g}.
1125: \end{equation}
1126: Again they are fractional, in spite of the fact that
1127: $\phi_1$ and $\phi_2$ have no direct coupling.
1128: This is because they are coupled through the electromagnetic
1129: potential, which tells that the two-gap superconductor is
1130: not a naive superposition of two one-gap superconductor.
1131: 
1132: \begin{figure}[t]
1133: \includegraphics[scale=0.4]{sf2pi.eps}
1134: \caption{The N-type straight vortices with $n=1$ which have
1135: a fractional flux with $\alpha=\gamma=0,~\beta=1.0$ (solid lines),
1136: $\alpha=0,~\beta=\lambda,~\gamma=-0.2$ (dashed lines),
1137: and $\alpha=-0.25,~\beta=\lambda,~\gamma=0$ (dotted lines).
1138: Here the unit of the scale is $1/{\bar \rho}$ and
1139: we have put $\lambda/g^2=2$.}
1140: \label{sf2pi}
1141: \end{figure}
1142: 
1143: According to the different boundary condition
1144: at the core there are two types of fractional flux vortices,
1145: D-type and N-type. These fractional flux vortices with $n=1$
1146: are plotted in Fig.~\ref{sf4pi} and Fig.~\ref{sf2pi}.
1147: The fractional vortex is also topological,
1148: but the topology of the fractional vortex is different from
1149: that of integer flux vortex.
1150: Notice that for the fractional flux vortices
1151: the $\pi_2(S^2)$ topology of $\hn$ becomes trivial, $\pi_2(S^2)=0$.
1152: This is because $\hn$ does not cover the target space
1153: $S^2$ fully. But in this case we still have
1154: a $U(1)$ topology $\pi_1(S^1)$, the topology of the $U(1)$
1155: symmetry which leaves $\hn$ invariant. And this Abelian
1156: topology describes the topology of
1157: the fractional flux vortex. So the topology of
1158: the fractional flux vortices is the same as that of
1159: the Abrikosov vortex.
1160: 
1161: An important feature of the fractional flux vortex is that
1162: the energy per unit length of the vortex is logarithmically
1163: divergent, which can be shown from the Hamiltonian (\ref{scfe4}).
1164: This is because the fractional flux vortex has a non-vanishing
1165: neutral current $k_\mu$ at the infinity.
1166: This, however, does not make the fractional flux vortex unphysical.
1167: In laboratory setting one can observe such vortex because
1168: one has a natural cutoff
1169: parameter $\Lambda$ fixed by the size of the superconductor,
1170: which can effectively make the
1171: energy of the fractional flux vortex finite.
1172: Indeed in $\rm He^3$ superfluid one often encounters
1173: the vorticity vortex whose energy is logarithmically
1174: divergent \cite{pra05,vol}.
1175: 
1176: The existence of fractional flux vortices
1177: in two-gap superconductor has been pointed out
1178: before in London limit \cite{baba2}. Our analysis in this paper
1179: shows that the London limit does not fully describe the vortex
1180: in two-gap superconductor. This is because the magnetic flux
1181: is determined by the boundary condition at the origin
1182: (as well as the boundary condition at the infinity).
1183: Clearly the existence of two types of vortices
1184: which have different core structure and
1185: different magnetic flux can not be understood in London limit.
1186: 
1187: In this section we have shown that the two-gap superconductor
1188: can have totally different magnetic vortices which can not be
1189: found in ordinary superconductor. There are two types of vortices,
1190: D-type and N-type, and both have two different topologies, $\pi_2(S^2)$
1191: and $\pi_1(S^1)$, which describe the vortices.
1192: The integral flux vortex is described by the $\pi_2(S^2)$
1193: topology, but the fractional vortex is described by the
1194: $\pi_1(S^1)$ topology. As importantly, there are infinitely
1195: many different vortices within the same topological sector.
1196: Moreover, the magnetic flux
1197: of the D-type vortex is larger than
1198: that of the N-type vortex by a factor $2\pi k/g$,
1199: so that in the same topological sector the D-type vortex
1200: has more energy than the N-type vortex.
1201: 
1202: Obviously all these vortices are non-Abrikosov. This does
1203: not mean that two-gap superconductor can not admit an
1204: Abrikosov vortex. With $f=\pi$ (or $f=0$) and
1205: $\alpha=\beta=\gamma=0$, (\ref{sveq}) describes an
1206: Abrikosov vortex. This is because with $\phi_1=0$
1207: (or with $\phi_2=0$) the two-gap superconductor reduces to an
1208: ordinary superconductor.
1209: 
1210: \section{Helical vortex}
1211: 
1212: In this section we show that the above non-Abrikosov
1213: vortices can be twisted to form a twisted magnetic vortex.
1214: With the twisting we obtain the helical vortex
1215: which is periodic in $z$-coordinate. To show this
1216: we choose the following ansatz \cite{ijpap,cm2},
1217: \begin{eqnarray}
1218: &\rho =\rho (\varrho ),~~~~~\xi =\left(
1219: \begin{array}{c}
1220: \cos \dfrac{f(\varrho )}2\exp (-in\varphi )\nonumber \\
1221: \sin \dfrac{f(\varrho )}2\exp (imkz)
1222: \end{array}
1223: \right) ,  \nonumber \\
1224: &A_\mu =\dfrac 1g\Big(nA_1(\varrho )\partial _\mu \varphi
1225: +mkA_2(\varrho )\partial _\mu z\Big).
1226: \label{hvans}
1227: \end{eqnarray}
1228: Obviously the ansatz is periodic in $z$-coordinate,
1229: with the period $2\pi/k$.
1230: 
1231: With the ansatz we have
1232: \bea
1233: &\hat{n}=\xi ^{\dag }\vec{\sigma}\xi =\left(
1234: \begin{array}{c}
1235: \sin f(\varrho )\cos (n\varphi +mkz) \\
1236: \sin f(\varrho )\sin (n\varphi +mkz) \\
1237: \cos f(\varrho ) \end{array} \right),  \nonumber \\
1238: &C_\mu =n\dfrac{\cos {f(\varrho )}+1}g \partial _\mu \varphi
1239: +mk \dfrac{\cos {f(\varrho )}-1}g \partial _\mu z, \nn\\
1240: &j_\mu =g\rho ^2\Big(n\big(A_1-\dfrac{\cos {f}+1}2\big)
1241: \partial _\mu \varphi   \nonumber \\
1242: &+mk\big(A_2-\dfrac{\cos {f}-1}2\big) \partial _\mu z\Big), \nn\\
1243: & k_\mu = g\rho ^2\Big(n\big(A_1 \cos f-\dfrac{\cos {f}+1}2\big)
1244: \partial _\mu \varphi   \nonumber \\
1245: &+mk\big(A_2 \cos f+\dfrac{\cos {f}-1}2\big) \partial _\mu z\Big),
1246: \label{schv}
1247: \eea
1248: and the following Hamiltonian
1249: \bea
1250: &{\cal H} =\dfrac 12\dot{\rho}^2+\dfrac 18\rho^2
1251: \Big(\dot{f}^2+ \big( \dfrac{n^2}{\varrho^2}
1252: +m^2 k^2 \big ) \sin ^2f \Big)  \nn\\
1253: &+ \dfrac{\rho^2}2 \Big[\dfrac{n^2}{\varrho^2}
1254: \Big( A_1 - \dfrac{\cos f+1}{2}\Big)^2
1255: + m^2k^2\Big(A_2- \dfrac{\cos f -1}{2} \Big)^2 \Big] \nn\\
1256: &+\dfrac {1}{2g^2}\Big( \dfrac{n^2}{\varrho^2}\dot{A_1}^2+m^2k^2
1257: \dot{A_2}^2 \Big)
1258: +\dfrac \lambda 8  \Big[(\rho ^2-\bar \rho^2)^2  \nn\\
1259: &+\dfrac{\alpha}\lambda (\rho^2-\dfrac{4\gamma }\alpha)\rho^2 \cos f
1260: +\dfrac{\beta}\lambda \rho^4\cos ^2f \Big]-\dfrac{\mu^2}{2\lambda}.
1261: \label{hvham}
1262: \eea
1263: With this (\ref{sceq2}) becomes
1264: \begin{eqnarray}
1265: &\ddot{\rho}+\dfrac 1\varrho \dot{\rho}-\Big[\dfrac 14
1266: \big(\dot{f}^2+(\dfrac{n^2}{\varrho^2}+m^2k^2)\sin^2f \big)  \nonumber \\
1267: &+\dfrac{n^2}{\varrho^2}(A_1-\dfrac{\cos {f}+1}2 )^2
1268: +m^2k^2(A_2- \dfrac{\cos {f}-1}2)^2\Big]\rho   \nonumber \\
1269: &=\dfrac \lambda 2\Big[\big(\rho ^2-\rho _0^2\big) +
1270: \dfrac{\alpha}{\lambda} \big(\rho^2-\dfrac{2\gamma}\alpha \big)
1271: \cos f +\dfrac{\beta}{\lambda} \rho^2\cos^2 f \Big] \rho,  \nonumber \\
1272: &\ddot{f}+\big(\dfrac 1\varrho +2\dfrac{\dot{\rho}}\rho %
1273: \big)\dot{f}-2\Big[\dfrac{n^2}{\varrho^2}\big(A_1-
1274: \dfrac 12\big)  \nonumber \\
1275: &+m^2k^2\big(A_2+\dfrac 12\big) \Big] \sin f \nonumber \\
1276: &=\Big(2\gamma - \big(\dfrac \alpha 2
1277: +\beta \cos f \big) \rho^2 \Big) \sin f,  \nonumber \\
1278: &\ddot{A_1}-\dfrac 1\varrho \dot{A}_1-g^2\rho ^2\Big(A_1
1279: - \dfrac{\cos {f}+1}2 \Big)=0,  \nonumber \\
1280: &\ddot{A_2}+\dfrac 1\varrho \dot{A}_2-g^2\rho ^2\Big(A_2
1281: -\dfrac{\cos {f}-1}2 \Big)=0.
1282: \label{hveq}
1283: \end{eqnarray}
1284: This is an obvious generalization of (\ref{sveq}).
1285: 
1286: \begin{figure}[t]
1287: \includegraphics[scale=0.4]{hv4pi.eps}
1288: \caption{The D-type helical vortex with $4\pi/g$-flux along the vortex
1289: with $m=n=1$ (with $k=1$). Three solutions are shown:
1290: $\alpha=\beta=\gamma=0$ (solid lines),
1291: $\alpha=\beta=0$ and $\gamma =0.005$ (dashed lines),
1292: $\alpha=\gamma=0$ and $\beta =-0.005$ (dotted lines).
1293: Here the unit of the scale is $1/{\bar \rho}$ and
1294: we have put $k=0.12 \bar \rho$ and $\lambda/g^2=2$.}
1295: \label{hv4pi}
1296: \end{figure}
1297: 
1298: To obtain the helical vortex we first consider the integer flux
1299: boundary condition at the infinity
1300: \begin{eqnarray}
1301: &\rho (\infty )=\sqrt {\dfrac{2(\mu +\gamma )}
1302: {(\lambda +\beta +\alpha)}},~~~~~f(\infty )=0, \nn\\
1303: &A_1(\infty )=1,~~~~~A_2(\infty )=0.
1304: \end{eqnarray}
1305: Just like the straight vortex, there are two types of
1306: boundary conditions at the core. Here we consider only the case $n=1$
1307: for simplicity: \\
1308: A. Dirichlet boundary condition
1309: \bea
1310: &\rho(0)=0,~~~~~f(0)=\pi, \nn\\
1311: &A_1(0)=-1,~~~~~\dot A_2(0)=0.
1312: \label{dbchv}
1313: \eea
1314: B. Neumann boundary condition
1315: \bea
1316: &\dot{\rho}(0)=0,~~~~~f(0)=\pi, \nn\\
1317: &A_1(0)=0,~~~~~\dot A_2(0)=0.
1318: \label{nbchv}
1319: \eea
1320: With the Dirichlet boundary condition we have the
1321: D-type helical vortex which has $4\pi/g$-flux along the
1322: vortex shown in Fig.~\ref{hv4pi}, but with the Neumann boundary
1323: condition we have the N-type helical vortex which has
1324: $2\pi/g$-flux along the vortex shown in Fig.~\ref{hv2pi}.
1325: 
1326: \begin{figure}[t]
1327: \includegraphics[scale=0.4]{hv2pi.eps}
1328: \caption{The N-type helical vortex with $2\pi/g$-flux along the vortex
1329: with $m=n=1$. Three solutions are shown:
1330: $\alpha=\beta=\gamma=0$ (solid lines),
1331: $\alpha=\beta=0$ and $\gamma =0.005$ (dashed lines),
1332: $\alpha=\gamma=0$ and $\beta =-0.005$ (dotted lines).
1333: Here the unit of the scale is $1/{\bar \rho}$ and
1334: we have put $k=0.12 \bar \rho$ and $\lambda/g^2=2$.}
1335: \label{hv2pi}
1336: \end{figure}
1337: 
1338: For the fractional flux helical vortex we impose the
1339: fractional flux boundary condition at the infinity
1340: \begin{eqnarray*}
1341: &\rho (\infty )=2\sqrt{\dfrac{2\beta \mu-\alpha \gamma}
1342: {4\beta \lambda-\alpha ^2}},
1343: ~~~\cos f(\infty )=\dfrac{2\gamma \lambda -\alpha \mu}
1344: {2\beta \mu -\alpha \gamma}, \nn\\
1345: &A_1(\infty )=\dfrac 12 \dfrac{2(\gamma \lambda+\beta \mu)
1346: -\alpha(\mu+\gamma)}{2\beta \mu -\alpha \gamma}, \nn\\
1347: &A_2(\infty )=\dfrac 12 \dfrac{2(\gamma \lambda-\beta \mu)
1348: -\alpha(\mu-\gamma)}{2\beta \mu -\alpha \gamma}.
1349: \end{eqnarray*}
1350: Now, with the Dirichlet boundary condition at the core,
1351: we obtain the D-type helical vortex which has a fractional flux along
1352: the vortex shown in Fig.~\ref{hf4pi}.
1353: But with the Neumann boundary
1354: condition (\ref{nbchv}) at the core,
1355: we obtain the N-type helical vortex which has a fractional flux along
1356: the vortex shown in Fig.~\ref{hf2pi}.
1357: Notice that, just as the fractional flux straight vortex,
1358: the energy per one period of the fractional flux helical vortex
1359: is logarithmically divergent.
1360: 
1361: A new feature of the helical vortex is that the magnetic
1362: flux becomes helical.
1363: Indeed the ansatz (\ref{hvans}) tells that the magnetic
1364: flux can be decomposed to the one along
1365: the vortex and the other around the vortex \cite{ijpap,cm2}
1366: \bea
1367: &F_{\hat \varrho \hat \varphi}=\dfrac ng\dfrac{\dot{A}_1}\varrho,
1368: ~~~F_{\hat{z}\hat{\varrho}}=-\dfrac{mk}g\dot{A}_2,
1369: \eea
1370: so that we have two magnetic fluxes linked together,
1371: \bea
1372: &\Phi_{\hat z}=\dfrac{}{}\int F_{\hat{\varrho}\hat{\varphi}}\varrho
1373: d\varrho d\varphi =\Big(A_1(\infty )-A_1(0)\Big)\dfrac{2\pi n}g, \nn\\
1374: &\Phi_{\hat \varphi}=\dfrac{}{}\int F_{\hat{z}\hat{\varrho}}dzd\varrho
1375: =-\Big(A_2(\infty )-A_2(0)\Big)\dfrac{2\pi m}g.
1376: \eea
1377: Obviously $\Phi_{\hat \varphi}$ is due to the helical structure
1378: of the vortex, which becomes fractional in general.
1379: 
1380: \begin{figure}[t]
1381: \includegraphics[scale=0.4]{hf4pi.eps}
1382: \caption{The D-type helical vortices which have a fractional flux
1383: with $m=n=1$ (with $k=1$). Two solutions with $\alpha=\gamma=0,~\beta= 0.5$
1384: (solid lines), and $\alpha=\gamma=0,~\beta = 1.0$ (dashed lines)
1385: are shown. Here the unit of the scale is $1/{\bar \rho}$ and
1386: we have put $\lambda/g^2=2,~k=0.1 \bar \rho$.}
1387: \label{hf4pi}
1388: \end{figure}
1389: 
1390: Another important feature of the helical vortex is that
1391: the electromagnetic current $j_\mu$ which is responsible for
1392: the Meissner effect also becomes helical.
1393: In particular, it has a non-trivial electromagnetic current
1394: $j_{\hat z}$ along the vortex which generates the magnetic flux
1395: $\Phi_{\hat \varphi}$, in addition to the usual electromagnetic
1396: current $j_{\hat \varphi}$ around the vortex which is responsible
1397: for $\Phi_{\hat z}$. But notice that the total electromagnetic
1398: current $i_{\hat z}$ along the vortex becomes zero.
1399: This, together with (\ref{sc}), tells that $\phi_1$ and $\phi_2$
1400: generate non-vanishing electromagnetic currents $i_{\hat z}^{(1)}$
1401: and $i_{\hat z}^{(2)}$ which flow oppositely and cancel each other.
1402: In this sense we may call the helical vortex superconducting,
1403: even though it has no net electromagnetic current
1404: $i_{\hat z}$ along the vortex \cite{ijpap,cm2}.
1405: 
1406: \section{magnetic knot in two-gap superconductor}
1407: 
1408: Clearly the helical vortex is unstable unless the periodicity
1409: condition is enforced by hand. Nevertheless it has an important
1410: implication, because the helical vortex predicts the existence of a
1411: topological knot in two-gap superconductor. This is because
1412: we can make it a twisted magnetic vortex ring
1413: smoothly bending and connecting two
1414: periodic ends together. The resulting twisted magnetic vortex ring
1415: becomes a knot whose topology is described by the Chern-Simon index
1416: of the electromagnetic potential \cite{ijpap,cm2}.
1417: 
1418: There have been two objections against the existence of
1419: a stable magnetic vortex ring in Abelian superconductor.
1420: First, it is supposed to be unstable
1421: due to the tension created by the ring \cite{huang}.
1422: Indeed if one constructs a vortex ring from an Abrikosov vortex,
1423: it becomes unstable because of the tension.
1424: But we can easily overcome this difficulty by twisting
1425: the magnetic vortex first and connecting the periodic ends together.
1426: In this case the non-trivial twist of the magnetic field
1427: forbids the untwisting of the vortex ring
1428: by any smooth deformation of field configuration, and
1429: the vortex ring becomes a stable knot.
1430: The other objection is that the Abelian gauge theory
1431: is supposed to have no non-trivial knot topology
1432: which allows a stable vortex ring.
1433: This again is a common misconception. As we have seen,
1434: the theory has a well-defined knot
1435: topology $\pi_3(S^2)$ described by the Chern-Simon index
1436: of the electromagnetic potential.
1437: This tells that there is
1438: no reason whatsoever why the Abelian superconductor can not have
1439: a topological knot.
1440: 
1441: \begin{figure}[t]
1442: \includegraphics[scale=0.4]{hf2pi.eps}
1443: \caption{The N-type helical vortices which have a fractional flux
1444: with $m=n=1$. Two solutions with $\alpha=\gamma=0,~\beta= 0.5$
1445: (solid lines), and $\alpha=\gamma=0,~\beta = 1.0$ (dashed lines)
1446: are shown. Here the unit of the scale is $1/{\bar \rho}$ and
1447: we have put $\lambda/g^2=2,~k=0.1 \bar \rho$.}
1448: \label{hf2pi}
1449: \end{figure}
1450: 
1451: To demonstrate the existence of a topological knot in the two-gap
1452: superconductor, we introduce the toroidal coordinates
1453: ($\eta,\gamma,\varphi $) defined by
1454: \begin{eqnarray}
1455: &x=\dfrac{a}{D}\sinh{\eta}\cos{\varphi}, ~~~y=\dfrac{%
1456: a}{D}\sinh{\eta}\sin{\varphi},  \nonumber \\
1457: &z=\dfrac{a}{D}\sin{\gamma},  \nonumber \\
1458: &D=\cosh{\eta}-\cos{\gamma},  \nonumber \\
1459: &ds^2=\dfrac{a^2}{D^2} \Big(d\eta^2+d\gamma^2+\sinh^2\eta
1460: d\varphi^2 \Big),  \nonumber \\
1461: &d^3x=\dfrac{a^3}{D^3} \sinh{\eta} d\eta d\gamma d\varphi,
1462: \label{tc}
1463: \end{eqnarray}
1464: where $a$ is the radius of the knot defined by $\eta=\infty$.
1465: Notice that in toroidal coordinates, $\eta=\gamma=0$ represents
1466: spatial infinity of $R^3$, and $\eta=\infty$ describes the torus
1467: center.
1468: 
1469: Now we choose the following ansatz,
1470: \begin{eqnarray}
1471: &\phi=\dfrac{1}{\sqrt 2} \rho (\eta, \gamma) \Bigg(\matrix{\cos
1472: \dfrac{f(\eta,\gamma )}{2} \exp (-im\varphi ) \cr \sin
1473: \dfrac{f(\eta ,\gamma)}{2} \exp (in\omega (\eta,\gamma))} \Bigg),
1474: \nonumber \\
1475: &A_\mu =\dfrac ng A_0(\eta, \gamma) \partial_\mu \eta
1476: +\dfrac ng A_1(\eta, \gamma) \partial_\mu \gamma \nn\\
1477: &+\dfrac mg A_2(\eta, \gamma)\partial_\mu \varphi.
1478: \label{sckans}
1479: \end{eqnarray}
1480: With this we have
1481: \begin{eqnarray}
1482: &\hn=\Bigg(\matrix {
1483: \sin f(\eta,\gamma )\cos (n\omega +m\varphi) \cr
1484: \sin f(\eta,\gamma )\sin (n\omega +m\varphi) \cr
1485: \cos f(\eta,\gamma ) } \Bigg), \nn\\
1486: &C_\mu = \dfrac{m}{2g} (\cos f +1)\partial_\mu \varphi +
1487: \dfrac{n}{2g} (\cos f -1) \partial_\mu \omega, \nn\\
1488: &F_{\eta \gamma }=\dfrac ng \big (\partial_\eta A_1
1489: -\partial_\gamma A_0 \big), \nn\\
1490: &F_{\gamma\varphi }=\dfrac mg \partial_\gamma A_2,
1491: ~~~~~F_{\varphi \eta }=-\dfrac mg \partial_\eta A_2.
1492: \label{kvf1}
1493: \end{eqnarray}
1494: Notice that, in the orthonormal frame $(\hat \eta, \hat \gamma,
1495: \hat \varphi)$, we have
1496: \begin{eqnarray}
1497: &A_{\hat{\eta}}=\dfrac Da A_\eta,
1498: ~~A_{\hat{\gamma}}=\dfrac Da A_\gamma,
1499: ~~A_{\hat{\varphi}}=\dfrac {D}{a \sinh \eta} A_\varphi, \nn\\
1500: &F_{\hat{\eta}\hat{\gamma}}=\dfrac {nD^2}{g a^2}
1501: \big (\partial_\eta A_1-\partial_\gamma A_0 \big), \nn\\
1502: &F_{\hat{\gamma}\hat{\varphi}}=\dfrac {mD^2}{g a^2 \sinh \eta}
1503: \partial_\gamma A_2, \nn\\
1504: &F_{\hat{\varphi}\hat{\eta}}=-\dfrac {mD^2}{ga^2 \sinh \eta}
1505: \partial_\eta A_2. \nn
1506: \label{kvf2}
1507: \end{eqnarray}
1508: Next, we adopt the $SU(2)$ symmetric potential
1509: with $\lambda_{11}=\lambda_{22}=\lambda_{12}=\lambda$ and
1510: $\mu_1=\mu_2=\mu$ for simplicity.
1511: In this case we have the following knot equation
1512: \begin{widetext}
1513: \begin{eqnarray}
1514: &\Bigg[ \partial _\eta ^2+\partial _\gamma ^2
1515: +\Big( \dfrac{\cosh \eta }{\sinh \eta }-\dfrac{\sinh \eta }D\Big)
1516: \partial _\eta -\dfrac{\sin \gamma }D\partial _\gamma \Bigg] \rho
1517: -\dfrac 14\Bigg[  \big(\partial _\eta f \big)^2
1518: + \big(\partial_\gamma f \big)^2 \nn\\
1519: &+\sin ^2f\Big( n^2 \big(\partial_\eta \omega
1520:  \big)^2+n^2 \big(\partial_\gamma \omega
1521: \big)^2+\dfrac{m^2}{\sinh ^2\eta }\Big) \Bigg] \rho
1522: -\Bigg[ n^2\Big(A_0- \dfrac{\cos f +1}2 \partial_\eta \omega \Big)^2
1523: +n^2 \Big(A_1- \dfrac{\cos f +1}2 \partial_\gamma \omega \Big)^2 \nn\\
1524: &+\dfrac{m^2}{\sinh ^2\eta } \Big(A_2
1525: - \dfrac{\cos f -1}2 \Big)^2\Bigg] \rho
1526: =\dfrac \lambda 2\dfrac{a^2}{D^2}
1527: \Big(\rho ^2-\bar \rho^2 \Big)\rho, \nn\\
1528: &\Bigg[ \partial _\eta ^2+\partial_\gamma ^2
1529: +\Big( \dfrac{\cosh \eta }{\sinh \eta }-\dfrac{\sinh \eta }D\Big)
1530: \partial _\eta -\dfrac{\sin \gamma }D
1531: \partial _\gamma \Bigg] f+\dfrac 2\rho \Big(\partial _\eta
1532: f\partial _\eta \rho
1533: +\partial _\gamma f\partial _\gamma \rho  \Big) \nn\\
1534: &=2\sin f\Bigg[ n^2 \Big(A_0-\dfrac 12\partial _\eta \omega
1535:  \Big)\partial _\eta \omega +n^2 \Big(A_1
1536: -\dfrac 12\partial _\gamma \omega \Big)\partial _\gamma \omega
1537: +\dfrac{m^2}{\sinh ^2\eta } \Big(A_2+\dfrac 12 \Big)\Bigg], \nn\\
1538: &\Bigg[ \partial _\eta ^2+\partial _\gamma ^2+\Big( \dfrac{\cosh \eta }{%
1539: \sinh \eta }-\dfrac{\sinh \eta }D\Big) \partial _\eta -\dfrac{\sin \gamma }%
1540: D\partial _\gamma \Bigg] \omega +\dfrac 2\rho \Big(\partial _\eta
1541: \rho \partial_\eta \omega +\partial _\gamma \rho
1542: \partial_\gamma \omega \Big)
1543: -\dfrac{\sin f}{1+\cos f} \Big(\partial _\eta f\partial _\eta
1544: \omega +\partial_\gamma f\partial_\gamma \omega \Big) \nn\\
1545: &+\dfrac2{\sin f} \Big(\partial_\eta fA_0+\partial_\gamma fA_1 \Big)=0, \nn\\
1546: & \Big(\partial _\gamma +\dfrac{\sin \gamma }D \Big)\Big(\partial
1547: _\gamma A_0-\partial _\eta A_1 \Big)=\dfrac{a^2}{D^2}g^2\rho ^2
1548: \Big(A_0- \dfrac{\cos f +1}2 \partial_\eta \omega \Big), \nn\\
1549: & \Big(\partial _\eta +\dfrac{\cosh \eta }{\sinh \eta }
1550: +\dfrac{\sinh \eta }D \Big) \Big(\partial _\eta A_1-\partial _\gamma A_0
1551: \Big)=\dfrac{a^2}{D^2}g^2\rho ^2 \Big(A_1
1552: - \dfrac{\cos f +1}2 \partial_\gamma \omega \Big), \nn\\
1553: &\Bigg[ \partial _\eta ^2+\partial _\gamma ^2-\Big( \dfrac{\cosh
1554: \eta }{\sinh \eta }-\dfrac{\sinh \eta }D\Big)
1555: \partial _\eta +\dfrac{\sin \gamma }D
1556: \partial _\gamma \Bigg] A_2=\dfrac{a^2}{D^2}g^2\rho ^2
1557: \Big(A_2- \dfrac{\cos f -1}2  \Big).
1558: \label{sckeq}
1559: \end{eqnarray}
1560: Moreover, from (\ref{scfe2}) and (\ref{sckans}) we have the following
1561: Hamiltonian for the knot
1562: \begin{eqnarray}
1563: &{\cal H}=\dfrac{D^2}{2a^2}\Bigg[ (\partial _\eta \rho
1564: )^2+(\partial _\gamma \rho )^2+\dfrac 14\rho ^2\big( (\partial
1565: _\eta f)^2+(\partial _\gamma f)^2\big) +\dfrac 14\rho ^2\sin
1566: ^2f\Big( n^2(\partial _\eta \omega )^2+n^2(\partial
1567: _\gamma \omega )^2+\dfrac{m^2}{\sinh ^2\eta }\Big) \Bigg] \nn\\
1568: &+\dfrac{D^2}{2a^2}\Bigg[ n^2(A_0- \dfrac{\cos f +1}2
1569: \partial_\eta \omega )^2+n^2(A_1-\dfrac{\cos f +1}2
1570: \partial_\gamma \omega )^2+\dfrac{m^2}{\sinh
1571: ^2\eta }(A_2 -\dfrac{\cos f -1}2 )^2\Bigg] \rho ^2 \nn\\
1572: &+\dfrac{D^4}{2g^2a^4}\Bigg[ n^2(\partial _\eta A_1-\partial _\gamma A_0)^2+%
1573: \dfrac{m^2}{\sinh ^2\eta }\Big( (\partial _\eta A_2)^2+(\partial
1574: _\gamma A_2)^2\Big) \Bigg]  +\dfrac \lambda 8(\rho ^2-\bar \rho
1575: ^2)^2.
1576: \label{sckham}
1577: \end{eqnarray}
1578: \end{widetext}
1579: Minimizing the energy we reproduce the knot equation
1580: (\ref{sckeq}).
1581: 
1582: In toroidal coordinates, $\eta=\gamma=0$ represents
1583: spatial infinity of $R^3$, and $\eta=\infty$ describes
1584: the torus center. So we can impose
1585: the following Neumann boundary condition
1586: \begin{eqnarray}
1587: &\rho(0,0)=\bar \rho, ~~~~~\dot \rho(\infty,\gamma)=0,  \nonumber \\
1588: &f(0,\gamma)=0, ~~~~~f(\infty,\gamma)=\pi,  \nonumber \\
1589: &\omega(\eta,0)=0, ~~~~~\omega(\eta,2 \pi)=2 \pi,  \nonumber \\
1590: &A_0(0,\gamma)=0, ~~~~~A_0(\infty,\gamma)=0,  \nonumber \\
1591: &A_1(0,\gamma)=1, ~~~~~A_1(\infty,\gamma)=0,  \nonumber \\
1592: &A_2(0,\gamma)=0, ~~~~~A_2(\infty,\gamma)=-1,
1593: \label{sckbc}
1594: \end{eqnarray}
1595: to obtain the desired knot.
1596: A numerical integration of (\ref{sckeq}) with the boundary
1597: conditions (\ref{sckbc}) is difficult to perform.
1598: But we can obtain the actual knot profile of $\rho,~f$, and
1599: $\omega$ by minimizing the energy. From (\ref{sckham}) the knot
1600: energy is given by
1601: \begin{eqnarray}
1602: &E=\dfrac{}{} \int {\cal H} \dfrac{a^3}{D^3} \sinh \eta d\eta
1603: d\gamma d\varphi.
1604: \label{scke1}
1605: \end{eqnarray}
1606: We find that, for $m=n=1$,
1607: the radius of knot which minimizes the energy (\ref{scke1})
1608: is given by
1609: \begin{eqnarray}
1610: a \simeq \dfrac{1.2}{\sqrt{2\mu}}.
1611: \label{sckrad}
1612: \end{eqnarray}
1613: From this we obtain the three-dimensional energy profile
1614: of the lightest axially
1615: symmetric knot in two-gap superconductor, which is shown in
1616: Fig.~\ref{2sck3d}.
1617: 
1618: With this we can estimate the energy of the axially symmetric
1619: knot,
1620: \begin{eqnarray}
1621: E \simeq 51 \dfrac{\bar \rho}{\sqrt \lambda}.
1622: \label{scke2}
1623: \end{eqnarray}
1624: We can also calculate the magnetic flux of the knot. Since the flux is
1625: helical, we have two fluxes, the flux $\Phi_{\hat \gamma}$ passing
1626: through the knot disk of radius $a$ in the $xy$-plane and the flux
1627: $\Phi_{\hat \varphi}$ which surrounds it. From the knot solution
1628: we find
1629: \begin{eqnarray}
1630: &\Phi _{\hat \eta} =\dfrac{}{}\int F_{\gamma \varphi }d\gamma d\varphi =0.
1631: \nonumber\\
1632: &\Phi _{\hat \gamma} =\dfrac{}{}\int F_{\varphi \eta }d\varphi d\eta
1633: =\dfrac{2m\pi }g, \nn\\
1634: &\Phi _{\hat \varphi} =\dfrac{}{}\int F_{\eta \gamma }d\eta d\gamma
1635: =-\dfrac{2n\pi }g
1636: \label{kflux}
1637: \end{eqnarray}
1638: The flux is quantized in the unit of $2\pi/g$, but this is
1639: due to the $SU(2)$ symmetric potential and the Neumann
1640: boundary condition (\ref{sckbc}). In general they can be fractional.
1641: But independent of this the two fluxes are linked,
1642: whose linking number becomes $mn$. This is important.
1643: 
1644: \begin{figure}[t]
1645: \includegraphics[scale=0.4]{ttsc1.eps}
1646: \caption{The energy profile of a N-type knot with $m=n=1$. Here we have
1647: put $\lambda/g^2=2$.} \label{2sck3d}
1648: \end{figure}
1649: 
1650: This confirms that the knot can be viewed as a twisted
1651: magnetic vortex ring, where the linking of two magnetic fluxes
1652: provides the knot topology. There is a natural
1653: candidate which can describe this topology of twisted
1654: magnetic vortex ring, the Chern-Simon index of
1655: the electromagnetic potential which describes
1656: the $\pi_3(S^2)$ topology.
1657: We can calculate the knot quantum number
1658: from the Chern-Simon index, and find
1659: \begin{eqnarray}
1660: Q_{CS} =-\dfrac{mn}{8\pi ^2}\int \varepsilon^{ijk} A_i
1661: F_{jk}dx^3 = mn.
1662: \label{sckqncs}
1663: \end{eqnarray}
1664: This confirms that the Chern-Simon index is indeed given by the
1665: linking number of two magnetic fluxes $\Phi_{\hat \gamma}$
1666: and $\Phi_{\hat \varphi}$.
1667: 
1668: It has been asserted that the knot topology is described by
1669: the $\pi_3(S^2)$ topology of the $CP^1$ field $\hn$ \cite{baba1}.
1670: We emphasize that this is only partially true,
1671: which becomes correct only when the knot carries an integer
1672: magnetic flux. Indeed, in this case
1673: $\hn$ acquires a non-trivial knot topology
1674: $\pi_3(S^2)$, which is given by \cite{ijpap}
1675: \begin{eqnarray}
1676: &Q =-\dfrac{mn}{8\pi ^2}\int (\partial_\eta f\partial_\gamma \omega
1677: -\partial_\gamma f \partial_\eta \omega)
1678: \sin f d\eta d\gamma d\varphi \nonumber \\
1679: &= - \dfrac{mn}{4\pi} \int \sin f df d\omega  = mn,
1680: \label{sckqn}
1681: \end{eqnarray}
1682: where the last equality comes from the boundary condition (\ref{sckbc}).
1683: Notice, however, that the $\pi_3(S^2)$ topology of $\hn$
1684: becomes trivial when the knot has a fractional flux. This is
1685: because the fractional flux knot comes from a fractional flux vortex,
1686: which has a trivial $\pi_2(S^2)$ topology. This means that
1687: the $\pi_3(S^2)$ topology of $\hn$ can not describe the
1688: topology of a fractional flux knot. In contrast,
1689: the Chern-Simon index of the electromagnetic potential
1690: can still provide the non-trivial $\pi_3(S^2)$,
1691: even when the knot has a fractional flux.
1692: This is because the Chern-Simon index describes
1693: the knot topology of the twisted magnetic flux.
1694: 
1695: \section{Josephson Interaction}
1696: 
1697: It has been well-known that two-gap superconductor may allow
1698: the prototype Josephson interaction \cite{maz}
1699: \bea
1700: \eta (\phi_1^{*}\phi_2 +\phi_1\phi_2^{*}),
1701: \eea
1702: where $\eta$ is a coupling constant. But we can consider
1703: a more general quartic Josephson interaction,
1704: \bea
1705: &V_J = \bar \eta \Big[ \big(\phi_1^{*}\phi_2 \exp(i\theta_1)
1706: +\phi_2^{*}\phi_1 \exp(-i\theta_1) \big) \Big]\nn\\
1707: &+\bar \eta_1 \Big[(\phi_1^{*}\phi_2)^2 \exp(i\theta_2)
1708: +(\phi_2^{*}\phi_1)^2 \exp(-i\theta_2) \Big] \nn\\
1709: &+ \bar \eta_2 (|\phi_1|^2 + |\phi_2|^2)
1710: \Big[\big(\phi_1^{*}\phi_2 \exp(i\theta_3)
1711: +\phi_2^{*}\phi_1 exp(-i\theta_3) \big) \Big] \nn\\
1712: &+ \bar \eta_3 (|\phi_1|^2 - |\phi_2|^2)
1713: \Big[\big(\phi_1^{*}\phi_2 \exp(i\theta_4) \nn\\
1714: &+\phi_2^{*}\phi_1 \exp(-i\theta_4) \big) \Big].
1715: \label{jpot1}
1716: \eea
1717: Clearly the Josephson interaction breaks the
1718: $U(1)\times U(1)$ symmetry of the potential (\ref{scpot2})
1719: down to $U(1)$.
1720: In general it is not easy to accommodate this type of
1721: generalized Josephson interaction. But for simpler
1722: Josephson interactions we can
1723: accommodate them within the framework of the potential
1724: (\ref{scpot2}). To understand how,
1725: notice that the potential (\ref{scpot2}) already has
1726: a Josephson interaction in the sense that it allows
1727: an interband transition between $\phi_1$ and $\phi_2$
1728: when $\lambda_{12}$ is not zero. This implies that
1729: the above Josephson interaction could be included
1730: in the $\lambda_{12}$ interaction.
1731: 
1732: \begin{figure*}[t]
1733: \centering\subfigure[~~$|\phi_1|^2$]
1734: {\includegraphics[scale=0.3]{jphi1.eps}}\hfil
1735: \subfigure[~~$|\phi_2|^2$]
1736: {\includegraphics[scale=0.3]{jphi2.eps}}
1737: \caption{The density profile of $|\phi_1|^2$ and $|\phi_2|^2$ of
1738: the N-type magnetic vortex in the presence of Josephson interaction.
1739: Here we have put $\bar \rho=1$, $\gamma=0.05$, $\eta = 0.25$, and
1740: $\lambda/g^2=2$.}
1741: \label{jphi}
1742: \end{figure*}
1743: 
1744: To exploit this point we generalize
1745: the $U(1)\times U(1)$ symmetric potential (\ref{scpot2}) to
1746: \bea
1747: &V \rightarrow \bar V =V + V_1, \nn\\
1748: &V_1 = \eta \Big[ \big(\phi_1^{*}\phi_2 \exp(i\theta)
1749: +\phi_2^{*}\phi_1 \exp(-i\theta) \big) \Big] \nn\\
1750: &+\eta_1 \Big[ \big(\phi_1^{*}\phi_2)^2 \exp(2i\theta)
1751: +(\phi_2^{*}\phi_1)^2 \exp(-2i\theta) \big) \Big] \nn\\
1752: &+ \eta_2 (|\phi_1|^2 + |\phi_2|^2)
1753: \Big[\big(\phi_1^{*}\phi_2 \exp(i\theta)
1754: +\phi_2^{*}\phi_1 \exp(-i\theta) \big) \Big] \nn\\
1755: &+\eta_3 (|\phi_1|^2 - |\phi_2|^2)
1756: \Big[\big(\phi_1^{*}\phi_2 \exp(i\theta) \nn\\
1757: &+\phi_2^{*}\phi_1 \exp(-i\theta) \big) \Big],
1758: \label{jpot2}
1759: \eea
1760: and introduce a new doublet $\psi$ with
1761: an $SU(2)$ transformation of $\phi$,
1762: \bea
1763: &\psi = {\cal M} \phi, \nn\\
1764: &{\cal M} = \Bigg( \matrix {\cos \dfrac a2 \exp(ib)
1765: ~~~-\sin \dfrac a2 \exp(ic) \cr
1766: \sin \dfrac a2 \exp(-ic)~~\cos \dfrac a2 \exp(-ib) } \Bigg), \nn\\
1767: &\tan a= \dfrac{\eta}\gamma.
1768: \label{jt}
1769: \eea
1770: Now, we can show that when
1771: \bea
1772: &\theta=-b+c, \nn\\
1773: &\eta_1= \dfrac{\eta^2}{2\gamma^2-\eta^2} \beta,
1774: ~~~~\eta_2 = -\dfrac{\eta}{2\gamma} \alpha, \nn\\
1775: &\eta_3= \dfrac{2\gamma \eta}{2\gamma^2-\eta^2} \beta,
1776: \label{jcon}
1777: \eea
1778: the potential (\ref{jpot2}) can be written as
1779: \bea
1780: &\bar V=\dfrac{\lambda_{11}'}{2}|\psi_1|^4+\lambda_{12}'|\psi_1|^2
1781: |\psi_2|^2+\dfrac{\lambda_{22}'}{2}|\psi_2|^4 \nn\\
1782: &-\mu_1'|\psi_1|^2-\mu_2'|\psi_2|^2 \nn\\
1783: &=\dfrac {\lambda'} 2(|\psi _1|^2+|\psi _2|^2 -\dfrac{\mu'}{\lambda'})^2
1784: +\dfrac {\alpha'} 2(|\psi _1|^4-|\psi _2|^4) \nn\\
1785: &+\dfrac {\beta'}2 (|\psi_1|^2-|\psi _2|^2)^2
1786: -\gamma' (|\psi _1|^2-|\psi _2|^2) \nn\\
1787: &- \dfrac{\mu'}{2\lambda'},
1788: \label{jpot3}
1789: \eea
1790: where
1791: \bea
1792: &\lambda' = \lambda - \eta_1, \nonumber \\
1793: &\alpha'= \dfrac{\sqrt{\gamma^2+\eta^2}}\gamma \alpha,
1794: ~~~\beta' = \dfrac{2(\gamma^2+\eta^2)}{2\gamma^2-\eta^2} \beta,  \nn\\
1795: &\mu' = \mu,~~~~~\gamma' = \sqrt{\gamma^2+\eta^2}.
1796: \eea
1797: So in terms of $\psi_1$ and $\psi_2$ the potential
1798: (\ref{jpot2}) becomes the potential (\ref{scpot5})
1799: which has no Josephson interaction.
1800: This means that, with (\ref{jcon}), we can formally
1801: absorb the Josephson interaction to $\lambda_{12}$ interaction.
1802: From this we concludes that the presence of the Josephson
1803: interaction does not affect the existence of the topological
1804: objects in two-gap superconductor.
1805: 
1806: This does not mean that the Josephson interaction does not affect
1807: the topological solutions. On the contrary, it does change
1808: the shape of the solutions drastically. This is because
1809: under the transformation (\ref{jt}) the profile of $\phi_1$
1810: and $\phi_2$ change drastically.
1811: To demonstrate this we let
1812: $\alpha=\beta=0$ for simplicity, and adopt the potential
1813: which has the following Josephson interaction
1814: \bea
1815: &V= \dfrac{\lambda}2 (|\phi_1|^2 + |\phi_2|^2-\dfrac{\mu}\lambda)^2
1816: - \gamma (|\phi_1|^2 - |\phi_2|^2) \nn\\
1817: &+\eta (\phi_1^{*}\phi_2 \exp(i\theta)
1818: +\phi_1\phi_2^{*} \exp(-i\theta)).
1819: \label{jpot}
1820: \eea
1821: Notice that, in terms of $\psi$, the potential is written as
1822: \bea
1823: &V= \dfrac{\lambda}2 (|\psi_1|^2 + |\psi_2|^2-\dfrac{\mu}\lambda)^2 \nn\\
1824: &- \sqrt{\gamma^2+\eta^2}(|\phi_1|^2 - |\phi_2|^2),
1825: \eea
1826: so that it clearly has two types of straight vortex solution
1827: of the following form
1828: \bea
1829: &\psi =\dfrac {\rho}{\sqrt{2}}
1830: \Bigg( \matrix {\cos \dfrac f2 \exp(-in\varphi) \cr
1831: \sin \dfrac f2 } \Bigg),
1832: ~~~\rho= \rho(\varrho),  \nn\\
1833: &A_\mu=\dfrac ng A(\varrho) \pro_\mu \varphi.
1834: \label{jvans1}
1835: \eea
1836: Now, in terms of $\phi$, the solution acquires the form
1837: \bea
1838: &\phi =\dfrac {\rho}{\sqrt{2}} \Big( \matrix {\xi_1 \cr
1839: \xi_2} \Big), \nn\\
1840: &\xi_1=\cos \dfrac f2 \cos \dfrac a2 \exp(-in\varphi-ib) \nn\\
1841: &+ \sin \dfrac f2 \sin \dfrac a2 \exp(i\theta+ib), \nn\\
1842: &\xi_2=-\cos \dfrac f2 \sin \dfrac a2 \exp(-in\varphi-i\theta-ib) \nn\\
1843: &+\sin \dfrac f2 \cos \dfrac a2 \exp(ib).
1844: \label{jvans2}
1845: \eea
1846: So, in the N-type vortex both $\phi_1$ and $\phi_2$ have a non-vanishing
1847: concentration of at the core in the presence of the Josephson
1848: interaction. The density profile of $\phi_1$ and $\phi_2$ of the
1849: N-type $2\pi/g$-flux vortex with $b=0$ and $\theta=0$
1850: is plotted in Fig.~\ref{jphi}.
1851: Obviously the solution is not axially symmetric. More importantly
1852: the vortex appears as a ``bound state" of two vortices
1853: made of $\phi_1$ and $\phi_2$. This confirms that the
1854: Josephson interaction does not prevent the the existence of
1855: two types of magnetic vortex,
1856: but changes the profile of the solutions drastically.
1857: We notice that a similar vortex has been discussed in
1858: two-component Bose-Einstein condensate \cite{ueda}.
1859: 
1860: Furthermore we can construct a helical vortex by
1861: twisting the above vortex and making it periodic in $z$-coordinate.
1862: In this case the helical vortex becomes a ``braided" magnetic vortex
1863: made of two vortices of $\phi_1$ and $\phi_2$.
1864: To see this we let
1865: \begin{eqnarray}
1866: &\phi= {\cal M}^{-1} \psi, \nn\\
1867: &\psi =\dfrac {\rho}{\sqrt{2}} \left(
1868: \begin{array}{l}
1869: {\cos \dfrac f2 \exp{(-in\varphi) }} \\
1870: {\sin \dfrac f2 \exp{(imkz)}}
1871: \end{array}  \right), \nn\\
1872: &A_\mu=\dfrac ng A_1(\varrho) \pro_\mu \varphi
1873: +\dfrac{mk}g A_2(\varrho) \pro_\mu z,
1874: \label{hjvans}
1875: \end{eqnarray}
1876: and again obtain two types of vortex. In this case the solution has
1877: the following particle densities for $\phi_1$ and
1878: $\phi_2$,
1879: \bea
1880: &|\phi _1|^2 =\dfrac{\rho ^2}4\Big( 1
1881: +\dfrac{\gamma}{\sqrt{\gamma^2+\eta^2}} \cos f \nn\\
1882: &+\dfrac{\eta}{\sqrt{\gamma^2+\eta^2}} \sin f
1883: \cos (n\varphi+mkz+\theta+2b)  \Big), \nn \\
1884: &|\phi _2|^2 =\dfrac{\rho ^2}4\Big( 1
1885: -\dfrac{\gamma}{\sqrt{\gamma^2+\eta^2}} \cos f \nn\\
1886: &-\dfrac{\eta}{\sqrt{\gamma^2+\eta^2}} \sin f
1887: \cos (n \varphi+mkz +\theta+2b)  \Big).
1888: \label{hjvp}
1889: \eea
1890: Clearly this shows that in the presence of the Josephson
1891: interaction the helical magnetic vortex
1892: becomes a braided vortex in which $\phi_1$-flux and $\phi_2$-flux
1893: are braided together.
1894: 
1895: Now, it goes without saying that we can make a braided knot,
1896: a twisted vortex ring, with the braided magnetic vortex.
1897: This tells that the Josephson interaction makes
1898: the topological objects in two-gap superconductor
1899: more interesting.
1900: 
1901: \section{Non-Abelian Superconductor}
1902: 
1903: So far we have discussed an Abelian gauge theory of
1904: two-gap superconductor. But our analysis implies that
1905: the doublet ($\phi_1,\phi_2$) can be treated as an $SU(2)$
1906: doublet. Indeed, when $\alpha=\beta=\gamma=0$,
1907: the Lagrangian (\ref{sclag}) has
1908: an exact $SU(2)$ symmetry. Even when there is no $SU(2)$
1909: symmetry, one may still regard that the theory has an approximate
1910: $SU(2)$ symmetry which is broken by the $\alpha$, $\beta$, and
1911: $\gamma$ terms.
1912: In this sense one may conclude that the Abelian two-gap superconductor
1913: has a (broken) $SU(2)$ symmetry.
1914: On the other hand the Ginzburg-Landau Lagrangian (\ref{sclag})
1915: is still based on the Abelian electromagnetic interaction.
1916: This leads us to wonder whether one can have
1917: a genuine non-Abelian superconductor
1918: in which the superconductivity is described by
1919: a non-Abelian dynamics.
1920: 
1921: To discuss this issue, notice that in the above two-gap
1922: superconductor the two condensates
1923: $\phi_1$ and $\phi_2$ carry the same charge,
1924: because the doublet is coupled to the Abelian
1925: electromagnetic field. Now we show that
1926: when the two condensates are made of opposite charges
1927: (made of one electron-electron pair condensate and one hole-hole
1928: pair condensate) the two-gap superconductor can be described by
1929: a genuine non-Abelian $SU(2)$ gauge
1930: theory. Moreover, we show that this type of non-Abelian superconductor
1931: also allows a non-Abrikosov magnetic vortex and topological
1932: knot identical to
1933: what we have discussed in this paper.
1934: 
1935: To construct a theory of non-Abelian superconductivity which
1936: is based on a genuine non-Abelian gauge theory, we need to understand
1937: the mathematical structure of the non-Abelian gauge theory.
1938: In non-Abelian gauge theory one can always decompose the
1939: gauge potential into the restricted potential $\hat A_\mu$
1940: and the valence potential $\X_\mu$.
1941: Consider the $SU(2)$ gauge theory and let $\hn$ be
1942: a gauge covariant unit triplet
1943: which selects the charge direction of $SU(2)$. In this case
1944: we have the following decomposition \cite{cho80,cho81},
1945: \bea
1946: & \vec{A}_\mu =A_\mu \n -
1947: \oneg \n\times\pro_\mu\n+\X_\mu\nonumber
1948:          = \hat A_\mu + \X_\mu, \nn\\
1949: &  (A_\mu = \n\cdot \vec A_\mu,~ \n^2 =1,~
1950: \hat{n}\cdot\vec{X}_\mu=0),
1951: \label{cdecom}
1952: \eea
1953: where $ A_\mu$ is the
1954: ``electric'' potential. Notice that the restricted potential
1955: is precisely the potential which leaves $\n$
1956: invariant under the parallel transport,
1957: \bea
1958: \D_\mu \n = \pro_\mu \n
1959: + g {\hat A}_\mu \times \n = 0.
1960: \eea
1961: Under the infinitesimal
1962: gauge transformation \bea \delta \n = - \vec \alpha \times \n
1963: \,,\,\,\,\, \delta \A_\mu = \oneg  D_\mu \vec \alpha, \eea one has
1964: \bea &&\delta A_\mu = \oneg \n \cdot \pro_\mu \valpha,\,\,\,\
1965: \delta \hat A_\mu = \oneg \D_\mu \valpha  ,  \nn \\
1966: &&\hspace{1.2cm}\delta \X_\mu = - \valpha \times \X_\mu  .
1967: \eea
1968: This tells three things. First, $\hat A_\mu$ by itself describes an $SU(2)$
1969: connection which enjoys the full $SU(2)$ gauge degrees of
1970: freedom. Secondly, the valence potential $\vec X_\mu$ forms a
1971: gauge covariant vector field under the gauge transformation.
1972: Most importantly, this tells that the decomposition is
1973: gauge-independent. Once the gauge covariant topological field
1974: $\hat n$ is given, the decomposition follows automatically
1975: independent of the choice of a gauge \cite{cho80,cho81}.
1976: 
1977: The importance of the decomposition (\ref{cdecom}) for our
1978: purpose is that one can construct a non-Abelian gauge theory,
1979: a restricted gauge theory which has a full non-Abelian gauge
1980: degrees of freedom, with the restricted potential
1981: $\hat A_\mu$ alone \cite{cho80,cho81}. This is because
1982: the valence potential $\vec X_\mu$ can be treated as
1983: a gauge covariant source, so that one can exclude it from
1984: the theory without compromising the gauge invariance.
1985: Indeed it is this restricted gauge theory which describes
1986: the non-Abelian gauge theory of
1987: superconductivity \cite{prb05}.
1988: 
1989: Remarkably the restricted potential $\hat{A}_\mu$ retains
1990: all the essential topological characteristics of the original
1991: non-Abelian potential.
1992: In fact, $\hat{n}$ defines the $\pi_2(S^2)$ topology
1993: which describes the non-Abelian monopoles
1994: and the $\pi_3(S^3)$ topology which characterizes
1995: the topologically distinct vacua \cite{cho79,cho80,cho81}.
1996: Furthermore it has a dual structure,
1997: \begin{eqnarray}
1998: & \hat{F}_{\mu\nu} = (F_{\mu\nu}+ H_{\mu\nu})\hat{n}\mbox{,}\nonumber \\
1999: & F_{\mu\nu} = \partial_\mu A_{\nu}-\partial_{\nu}A_\mu \mbox{,}\nonumber \\
2000: & H_{\mu\nu} = -\dfrac{1}{g} \hat{n}\cdot(\partial_\mu
2001: \hat{n}\times\partial_\nu\hat{n}) = \partial_\mu
2002: C_\nu-\partial_\nu C_\mu,
2003: \end{eqnarray}
2004: where $C_\mu$ is the ``magnetic'' potential.
2005: Notice that this is exactly the potential $C_\mu$
2006: that we have introduced in (\ref{id}).
2007: This is an indication that the Ginzburg-Landau
2008: theory of two-gap superconductor is closely related to
2009: the restricted $SU(2)$ gauge theory.
2010: 
2011: With these preliminaries we now demonstrate a
2012: non-Abelian superconductivity and non-Abelian Meissner effect.
2013: Consider a $SU(2)$ gauge theory
2014: described by the Lagrangian in which a doublet $\Phi$ couples
2015: to the restricted $SU(2)$ gauge potential,
2016: \bea
2017: &{\cal L} = -|\hat D_\mu \Phi|^2 - V(\Phi,\Phi^{\dagger})
2018: -\dfrac{1}{4} {\hat F}_{\mu\nu}^2, \nn\\
2019: &\hat D_\mu \Phi = ( \partial_\mu + \dfrac{g}{2i} \vec \sigma
2020: \cdot \hat A_\mu ) \Phi.
2021: \label{nasclag1}
2022: \eea
2023: The equation of
2024: motion of the Lagrangian is given by
2025: \bea
2026: &{\hat D}^2\Phi
2027: =\dfrac{d V}{d \Phi^{\dagger}} \Phi, \nn\\
2028: &\hat D_\mu \hat F_{\mu \nu}
2029: = g \Big[(\hat D_\nu \Phi)^{\dagger}\dfrac{\vec\sigma}{2i} \Phi
2030: - \Phi ^{\dagger} \dfrac{ \vec\sigma}{2i}(\hat D_\nu \Phi) \Big].
2031: \label{nasceq}
2032: \eea
2033: Let $\xi$ and $\eta$
2034: be two doublets which form an orthonormal basis,
2035: \bea
2036: &\xi{\dagger}\xi =1,~~~~~\eta^{\dagger} \eta = 1,
2037: ~~~~~\xi^{\dagger}\eta = \eta^{\dagger} \xi = 0, \nn\\
2038: &\xi^{\dagger} \vec \sigma\xi = \hat n,
2039: ~~~~~\eta^{\dagger} \vec \sigma\eta= -\hat n, \nn\\
2040: &(\hn \cdot \vec \sigma) ~\xi = \xi,
2041: ~~~~~(\hn \cdot \vec \sigma) ~\eta = -\eta,
2042: \label{odbasis}
2043: \eea
2044: and let
2045: \bea
2046: \Phi = \phi_+ \xi + \phi_- \eta, ~~~~~(\phi_+= \xi^{\dagger}\Phi,
2047: ~~~\phi_-= \eta^{\dagger} \Phi).
2048: \label{phidecom}
2049: \eea
2050: With this we have the identity
2051: \bea
2052: &\Big[\partial_\mu - \dfrac{g}{2i} \big(C_\mu \hn
2053: + \dfrac{1}{g} \hn \times \partial_\mu \hn \big)
2054: \cdot \vec \sigma \Big] \xi = 0, \nn\\
2055: &\Big[\partial_\mu + \dfrac{g}{2i} \big(C_\mu \hn
2056: - \dfrac{1}{g} \hn \times \partial_\mu \hn \big)
2057: \cdot \vec \sigma \Big] \eta = 0,
2058: \label{cid0}
2059: \eea
2060: and find
2061: \bea
2062: \hat D_\mu \Phi = (D_\mu \phi_+) \xi + (D_\mu \phi_-)  \eta,
2063: \label{phidef}
2064: \eea
2065: where
2066: \bea
2067: &D_\mu \phi_+ = (\partial_\mu + \dfrac{g}{2i} {\cal A}_\mu) \phi_+,
2068: ~~~D_\mu \phi_- = (\partial_\mu - \dfrac{g}{2i} {\cal A}_\mu) \phi_-, \nn\\
2069: &{\cal A}_\mu = A_\mu + C_\mu, \nn\\
2070: &C_\mu = \dfrac{2i}{g} \xi^{\dagger}\partial_\mu \xi
2071: = - \dfrac{2i}{g} \eta^{\dagger}\partial_\mu \eta. \nn
2072: \eea
2073: From this we can express (\ref{nasclag1}) as
2074: \bea
2075: &{\cal L} = - |D_\mu \phi_+|^2 - |D_\mu \phi_-|^2
2076: - V(\phi_+,\phi_-)  \nn\\
2077: &- \dfrac{\lambda}{2} (\phi_+^{\dagger}\phi_+ + \phi_-^{\dagger}\phi_-)^2
2078: -\dfrac{1}{4} {\cal F}_{\mu\nu}^2,
2079: \label{nasclag2}
2080: \eea
2081: where
2082: \bea
2083: {\cal F}_{\mu\nu} = \partial_\mu {\cal A}_\nu
2084: - \partial_\nu {\cal A}_\mu. \nn
2085: \eea
2086: This tells that the restricted $SU(2)$ gauge theory
2087: (\ref{nasclag1}) is reduced to
2088: an Abelian gauge theory coupled to oppositely charged
2089: scalar fields $\phi_+$ and $\phi_-$. We emphasize that
2090: this Abelianization is achieved without any gauge fixing.
2091: 
2092: The Abelianization assures that the non-Abelian Ginzburg-Landau
2093: theory is not different from the Abelian Ginzburg-Landau theory.
2094: Indeed with
2095: \begin{equation}
2096: \chi = \left(\begin{array}{rr}
2097: \phi_+\\
2098: \phi_-^*
2099: \end{array}\right),
2100: \end{equation}
2101: we can express the Lagrangian (\ref{nasclag2}) as
2102: \bea
2103: &{\cal L} = - |D_\mu \chi|^2 + \mu^2 \chi^{\dagger}\chi
2104: - \dfrac{\lambda}{2} (\chi^{\dagger} \chi)^2
2105: - \dfrac{1}{4} {\cal F}_{\mu \nu}^2, \nn\\
2106: &D_\mu \chi = (\partial_\mu + ig {\cal A}_\mu) \chi.
2107: \label{nalag}
2108: \eea
2109: This is formally identical to the Lagrangian (\ref{sclag}) of
2110: the Abelian two-gap superconductor discussed in Section III.
2111: The only difference is
2112: that here $\phi$ and $A_\mu$ are replaced by $\chi$ and ${\cal A}_\mu$.
2113: This establishes that, with the proper redefinition of field
2114: variables (\ref{phidecom}) and (\ref{phidef}), our non-Abelian
2115: restricted gauge theory (\ref{nasclag1}) can in fact be made
2116: identical to the Abelian gauge theory of two-gap superconductor. This
2117: proves the existence of non-Abelian superconductors
2118: made of the doublet consisting of oppositely charged
2119: condensates \cite{prb05}. As importantly our analysis tells that
2120: the two-gap Abelian superconductor has a hidden
2121: non-Abelian gauge symmetry, because it can be transformed to
2122: the non-Abelian restricted gauge theory. This implies that
2123: the underlying dynamics of the two-gap superconductor
2124: is indeed the non-Abelian gauge symmetry.
2125: In the non-Abelian formalism it is explicit.
2126: But in the Abelian formalism it is hidden,
2127: where the full non-Abelian gauge symmetry only
2128: becomes transparent when one embeds the doublet
2129: $(\phi_1,\phi_2)$ properly into the non-Abelian symmetry.
2130: 
2131: Once the equivalence of two Lagrangians (\ref{sclag})
2132: and (\ref{nasclag1}) is established, it must be evident
2133: that the non-Abelian gauge theory of two-gap superconductor also
2134: admits a non-Abrikosov vortex and magnetic knot.
2135: This confirms the existence of a non-Abelian
2136: Meissner effect and non-Abelian superconductivity.
2137: All the above results of Abelian superconductor become
2138: equally valid here.
2139: 
2140: \section{Discussion}
2141: 
2142: In this paper we have shown that the two-gap
2143: superconductor can admit non-Abrikosov vortex
2144: and topological knot. There are two types of non-Abrikosov vortex,
2145: D-type and N-type. The D-type has no concentration of
2146: the condensate at the core, but the N-type has a non-trivial profile of
2147: the condensate at the core. In terms of topology
2148: there are two, the non-Abelian topology $\pi_2(S^2)$
2149: defined by $\hn$ and the Abelian topology
2150: $\pi_1(S^1)$ defined by the invariant subgroup of $\hn$.
2151: And both D-type and N-type vortices exist within the same
2152: topological sector. In particular, we have infinitely many
2153: D-type vortices which have the same topology.
2154: The magnetic flux of the vortices can be integral or fractional.
2155: The N-type vortex can have a $2\pi n/g$-flux or a fraction of this flux,
2156: but the D-type vortex
2157: has $2\pi k/g$ more flux than the N-type. And we have shown that
2158: these non-Abrikosov vortices can be twisted
2159: to form a helical vortex which is periodic in $z$-coordinate.
2160: 
2161: Perhaps a most interesting topological object in two-gap
2162: superconductor is the magnetic knot, a twisted magnetic vortex
2163: ring made of helical vortex. Our analysis suggests that
2164: we have two types of knot, the D-type and the N-type.
2165: They are made of two magnetic fluxes linked together,
2166: one flux along the knot axis
2167: and one flux along
2168: the knot tube. And the linking number of two fluxes
2169: provides the knot topology $\pi_3(S^2)$, which is described by
2170: the Chern-Simon index of the electromagnetic potential.
2171: The knot is stable dynamically as well as topologically.
2172: The topological stability follows from the fact that
2173: two flux rings linked together can not be separated by
2174: any continuous deformation of the field configuration.
2175: The dynamical stability follows from the fact that
2176: the flux trapped inside of the knot ring can not be squeezed
2177: out, which means that it provides
2178: a repulsive force against the collapse of the knot.
2179: Another way to understand this dynamical stability
2180: is to notice that the supercurrent
2181: along the knot generates a net
2182: angular momentum around the knot axis.
2183: And this provides the centrifugal
2184: repulsive force preventing the knot to collapse.
2185: This makes the knot dynamically stable.
2186: 
2187: The Josephson interaction makes these topological
2188: objects more interesting. The straight vortex becomes
2189: a bound state of two magnetic vortices made of two condensates
2190: $\phi_1$ and $\phi_2$, and the helical vortex becomes
2191: a braided magnetic vortex of two condensates.
2192: Moreover the knot acquires the form of a braided magnetic
2193: vortex ring. And we have two of them.
2194: 
2195: It must be emphasized, however, that the actual magnetic flux
2196: of vortex and knot is determined by the two-gap superconductor
2197: at hand because it is fixed by the parameters of
2198: the potential which characterizes
2199: the superconductor. Independent of this all two-gap superconductors
2200: have two types of vortex and knot.
2201: On the other hand one must keep the followings in mind.
2202: First, compared with the N-type vortex the D-type vortex
2203: has more energy in general because the D-type carries
2204: more flux. This opens the possibility that,
2205: within the same topological sector,
2206: the D-type vortices could decay to the N-type vortices.
2207: Secondly, the energy (per unit length) of
2208: the fractional flux vortex and knot
2209: is logarithmically divergent, so that they can exist only
2210: when there is a cutoff parameter which makes the energy finite.
2211: This tells that the N-type $2\pi/g$ vortex forms the true
2212: finite energy ground state vortex of two-gap
2213: superconductor.
2214: 
2215: Another important lesson from our analysis is that
2216: the non-Abelian dynamics
2217: could play a crucial role in condensed matter physics.
2218: Indeed we have shown that we can actually
2219: construct an $SU(2)$ gauge theory of superconductivity
2220: which is mathematically equivalent to
2221: the Abelian gauge theory of two-gap superconductor.
2222: This means that, implicitly or explicitly, the underlying dynamics of
2223: multi-gap superconductor can ultimately be related to
2224: a non-Abelian gauge theory.
2225: 
2226: In this paper we have studied the topological objects in
2227: two-gap superconductor. But from our discussion it must
2228: be clear that similar topological objects should also exist
2229: in multi-gap superconductor in general. This is because
2230: the multi-gap superconductor is described by a multi-component
2231: condensate, which naturally accommodates the non-trivial
2232: non-Abelian topology.
2233: 
2234: Clearly the above theory of two-gap superconductor
2235: is closely related to the Gross-Pitaevskii
2236: theory of two-component Bose-Einstein
2237: condensate (BEC), which tells that similar topological objects
2238: can also exist in two-component BEC \cite{ijpap,ruo,pra05}.
2239: This is because in the absence of the
2240: electromagnetic interaction the above Ginzburg-Landau
2241: Lagrangian reduces to the Gross-Pitaevskii Lagrangian
2242: of two-component BEC. But there is an important difference.
2243: In two-component BEC only the N-type vortex and knot
2244: exist, because it allows only the N-type boundary
2245: condition \cite{ijpap,pra05}.
2246: In this sense it is really remarkable that two-gap
2247: superconductor allows two types of topological objects.
2248: 
2249: We close with the following remarks: \\
2250: 1. Recently similar non-Abelian vortices and knots
2251: have been asserted to exist almost everywhere, in atomic physics
2252: in two-component BEC \cite{ijpap,ruo,pra05},
2253: in condensed matter physics in multi-gap
2254: superconductors \cite{ijpap,baba1},
2255: in nuclear physics in Skyrme theory \cite{prl01,plb04},
2256: in high energy physics in QCD \cite{plb05}.
2257: The major difference here is that our vortex and knot are made of a
2258: real magnetic flux. We emphasize that at the center of these
2259: topological objects lies the baby skyrmion and the Faddeev-Niemi
2260: knot in Skyrme theory. In fact, one can show that
2261: our magnetic vortex and knot (as well as those in
2262: two-component BEC) are a straightforward
2263: generalization of the baby skyrmion and the Faddeev-Niemi
2264: knot \cite{prl01,ijpap,plb04}. This is because both the Ginzburg-Landau
2265: theory of two-gap superconductor and the Skyrme theory
2266: are described by a $CP^1$ field $\hn$ which obeys the same
2267: non-linear dynamics. \\
2268: 2. From our analysis there should be no doubt that the non-Abrikosov
2269: vortex and the magnetic knot must exist in two-gap
2270: superconductor. If so, the challenge now is
2271: to verify the existence of these topological
2272: objects experimentally. Constructing the knot
2273: might not be a simple task at present moment. But the construction
2274: of the non-Abrikosov vortex could be rather straightforward (at
2275: least in principle) \cite{exp}.
2276: To identify the non-Abrikosov vortex, there are two points
2277: one has to keep in mind. First, the magnetic flux of the
2278: non-Abrikosov vortex need not be $2\pi/g$, and can be fractional
2279: in general. More importantly, there are two types of vortex,
2280: the D-type which has no concentration of the condensate
2281: at the core and the N-type which has
2282: a non-trivial concentration of the condensate at the core.
2283: These are the crucial points which distinguish the non-Abrikosov
2284: vortex from the Abrikosov vortex.
2285: With this in mind, one should be able to construct and identify
2286: the non-Abrikosov vortex in two-gap superconductor
2287: without much difficulty. \\
2288: 3. The non-Abelian gauge theory of superconductivity
2289: is not just an academic curiosity. There is an excellent
2290: example of non-Abelian two-gap
2291: superconductor, the liquid metallic
2292: hydrogen (LMH) \cite{ash}. Under high pressure the LMH becomes
2293: a superconducting state in low temperature,
2294: due to the electron Cooper pairs. But
2295: in a lower temperature the proton Cooper pairs can coexist with
2296: the electron pairs. And obviously it has no
2297: Josephson interaction and probably a weak or no $\lambda_{12}$
2298: interaction. So the LMH becomes an excellent
2299: candidate of non-Abelian two-gap superconductor.
2300: This implies that the LMH can have all
2301: the topological objects we have discussed in this paper.
2302: In particular it must have two types of
2303: vortex and knot.
2304: 
2305: In this paper we have discussed
2306: the topological objects which we obtain with the ansatz (\ref{svans}),
2307: (\ref{hvans}), (\ref{sckans}), or (\ref{jvans2}). But we emphasize that
2308: there are other topological objects which can be obtained
2309: with different ansatz. These objects and the physical
2310: implications of these topological objects in two-gap
2311: superconductor will be discussed
2312: in an accompanying paper \cite{sc5}.
2313: 
2314: {\bf ACKNOWLEDGEMENT}
2315: 
2316: ~~~The work is supported in part by the ABRL Program of
2317: Korea Science and Engineering Foundation (Grant R02-2003-000-10043-0).
2318: 
2319: \begin{thebibliography}{99}
2320: \bibitem{dirac}P. A. M. Dirac, Proc. Roy. Soc. {\bf A113}, 60 (1931);
2321: Phys. Rev. {\bf 74}, 817 (1948).
2322: \bibitem{abri} A. Abrikosov, Sov. Phys. JETP {\bf 5}, 1174 (1957);
2323: H. Nielsen and P. Olesen, Nucl. Phys. {\bf 61}, 45 (1973).
2324: \bibitem{skyr}T. H. R. Skyrme, Proc. Roy. Soc. (London) {\bf 260}, 127
2325: (1961); {\bf 262}, 237 (1961); Nucl. Phys. {\bf 31}, 556 (1962).
2326: \bibitem{fadd1} L. Faddeev and A. Niemi, Nature {\bf 387}, 58 (1997);
2327: J. Gladikowski and M. Hellmund, Phys. Rev. {\bf D56}, 5194 (1997);
2328: R. Battye and P. Sutcliffe, Phys. Rev. Lett. {\bf 81}, 4798 (1998).
2329: \bibitem{prl01} Y. M. Cho, Phys. Rev. Lett. {\bf 87}, 252001 (2001).
2330: \bibitem{bec} C. Myatt {\it at al.}, Phys. Rev. Lett. {\bf 78}, 586 (1997);
2331: D. Stamper-Kurn, {\it at al.}, Phys. Rev. Lett. {\bf 80}, 2027 (1998).
2332: \bibitem{sc}J. Nagamatsu et al., Nature {\bf 410}, 63 (2001);
2333: S. L. Bud'ko et al., Phys. Rev. Lett. {\bf 86}, 1877 (2001);
2334: C. U. Jung et al., Appl. Phys. Lett. {\bf 78}, 4157 (2001).
2335: \bibitem{ijpap} Y. M. Cho, cond-mat/0112325;
2336: Int. J. Pure Appl. Phys. {\bf 1}, 246 (2005); Y. M. Cho and N. S. Yong,
2337: cond-mat/0308182, submitted to Int. J. Pure Appl. Phys.
2338: \bibitem{prb05} Y. M. Cho, cond-mat/0112498;
2339: Phys. Rev. {\bf B72}, 212516 (2005).
2340: \bibitem{baba1} E. Babaev, L. Faddeev, and A. Niemi,
2341: Phys. Rev. {\bf B65}, 100512 (2002).
2342: \bibitem{ruo} H. Stoof {\it at al.}, Phys. Rev. Lett. {\bf
2343: 87}, 120407 (2001); C. Savage and
2344: J. Ruostekoski, Phys. Rev. Lett. {\bf 91}, 010403 (2003).
2345: \bibitem{pra05} Y. M. Cho, Hyojoong Khim, and Pengming Zhang,
2346: Phys. Rev. {\bf A72}, 063603 (2005).
2347: \bibitem{cm2} Y. M. Cho, cond-mat/0601347, Phys. Rev. {\bf B73}, in press.
2348: \bibitem{maz} A. Leggett, Prog. Theo. Phys. {\bf 36}, 901 (1966);
2349: Y. Tanaka, Phys. Rev. Lett. {\bf 88}, 017002 (2002);
2350: A. Gurevich, Phys. Rev. {\bf B67}, 184515 (2003);
2351: M. Zhitomirsky and V. Dao, Phys. Rev. {\bf B69}, 054508 (2004).
2352: \bibitem{baba2} E. Babaev, Phys. Rev. Lett. {\bf 89}, 067001 (2002).
2353: \bibitem{ho} N. D. Mermin and T. L. Ho, Phys. Rev. Lett. {\bf 36},
2354: 594 (1976).
2355: \bibitem{cho79} Y. M. Cho, Phys. Lett. {\bf B81}, 25 (1979);
2356: hep-th/0409246.
2357: \bibitem{cho80}Y. M. Cho, Phys. Rev. {\bf D21}, 1080 (1980);
2358: Y. M. Cho, Phys. Rev. {\bf D62}, 074009 (2000).
2359: \bibitem{cho81}Y. M. Cho, Phys. Rev. Lett. {\bf 46}, 302 (1981);
2360: Phys. Rev. {\bf D23}, 2415 (1981); W. S. Bae, Y. M. Cho, and S. W. Kimm,
2361: Phys. Rev. {\bf D65}, 025005 (2002).
2362: \bibitem{vol} See, for example, G. Volovik, {\it The
2363: Universe in a Helium Droplet}, Clarendon Press (Oxford), 2003.
2364: \bibitem{plb04} Y. M. Cho, Phys. Lett. {\bf B603}, 88 (2004);
2365: W. S. Bae, Y. M. Cho, and B. S. Park, hep-th/0404181.
2366: \bibitem{huang} K. Huang and R. Tipton, Phys. Rev. {\bf D23}, 3050
2367: (1981).
2368: \bibitem{ueda} K. Kasamatsu, M. Tsubota and M. Ueda,
2369: Phys. Rev. Lett. {\bf 93}, 250406 (2004);
2370: Phys. Rev. {\bf A71}, 043611 (2005).
2371: \bibitem{plb05} Y. M. Cho, Phys. Lett. {\bf B616}, 101 (2005).
2372: \bibitem{exp} M. Eskildsen {\it et al.} Phys. Rev. Lett. {\bf 89},
2373: 187003 (2002); A. Koshelev and A. Golubov, Phys. Rev. Lett. {\bf 90},
2374: 177002 (2003).
2375: \bibitem{ash} N. Ashcroft, Phys. Rev. Lett. {\bf 92}, 187002 (2004);
2376: E. Babaev, A. Sudbo, and N. Ashcroft, Nature {\bf 431}, 666 (2004);
2377: Phys. Rev. Lett. {\bf 95}, 105301 (2005).
2378: \bibitem{sc5} Y. K. Bang, Y.M. Cho, and Pengming Zhang,
2379: to be published.
2380: \end{thebibliography}
2381: 
2382: \end{document}
2383: