1: \documentclass[twocolumn,showpacs,showkeys]{revtex4}
2: \usepackage{graphicx}
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{Knots in Condensed Matters}
59: \author{Y. M. Cho}
60: \email{ymcho@yongmin.snu.ac.kr}
61: \affiliation{Center for Theoretical Physics and School of Physics \\
62: College of Natural Sciences,
63: Seoul National University,
64: Seoul 151-742, Korea \\
65: }
66: \begin{abstract}
67: ~~~~~We propose two types of topologically stable knot solitons
68: in condensed matters, one in two-component
69: Bose-Einstein condensates and one in two-gap superconductors.
70: We identify the knot in Bose-Einstein condensates as
71: a twisted vorticity flux ring and the knot in two-gap superconductors
72: as a twisted magnetic flux ring.
73: In both cases we show that there is a remarkable interplay between
74: topology and dynamics which transforms the topologcal stability
75: to the dynamical stability, and vise versa.
76: We discuss how these knots can be constructed in the spin-1/2
77: condensate of $^{87}{\rm Rb}$ atoms and in two-gap superconductor
78: of $MgB_2$.
79: \end{abstract}
80: \pacs{03.75.Fi, 05.30.Jp, 67.40.Vs, 74.72,-h}
81: \keywords{Topological knot in BEC, Knot in two-gap superconductor}
82: \maketitle
83:
84:
85: \section{Introduction}
86:
87: Topological objects (monopoles, vortices, skyrmions, etc.)
88: have played an increasingly important role
89: in physics \cite{dirac,abri}. In particular finite energy topological
90: objects have been widely studied in theoretical physics
91: \cite{skyrme,thooft}. A recent addition to this family of finite
92: energy solitons has been the knots \cite{fadd1,prl01}.
93: The interest on these topological objects,
94: however, has been confined mainly to theoretical physics,
95: because most of them (exept the vortices) exist in
96: ``hypothetical'' worlds which are very difficult to
97: create in laboratories. The only
98: topological objects which one can realistically expect to exist
99: in the ``standard'' models are the electroweak
100: monopoles and dyons in Weinberg-Salam model
101: which have a non-trivial $W_\mu$
102: and $Z_\mu$ dressing \cite{plb97,yang}. Unfortunately these objects
103: could carry an infinite energy, which makes it impossible to
104: create them experimentally.
105:
106: Fortunately the recent experimental advances in condensed matter
107: physics, in particular the construction of multi-component Bose-Einstein
108: condensates (BEC) made of $^{87}{\rm Rb}$ \cite{bec,ruo}
109: and two-gap superconductors made of $MgB_2$ \cite{exp1,exp2},
110: have widely opened new opportunities for
111: us to create new topological objects experimentally which so far
112: have been only of theoretical interest.
113: The purpose of this paper is to argue that these new
114: multi-component condensates could allow us to have real
115: knots, topologically stable finite energy 3-dimensional solitons.
116: This is because, due to the multi-component structure, the new condensates
117: have a non-Abelian symmetry which provides the needed topology
118: for the stable knots.
119:
120: To understand how the realistic knots can appear in these
121: condensed matters, it is necessary to understand
122: the prototype Faddeev-Niemi knot in Skyrme theory.
123: The Skyrme theory is well known to have a magnetic vortex
124: known as the baby skyrmion \cite{piet}, and the
125: Faddeev-Niemi knot can be identified as a twisted vortex ring
126: made of a helical baby skyrmion \cite{prl01,plb04}. In the following
127: we show how one can construct similar knots from helical vortices
128: in two-component Bose-Einstein condensates and two-gap superconductors.
129:
130: The paper is organized as follows. In Section II we discuss
131: how the helical vortex can give rise to the prototype knot
132: in Skyrme theory for later purpose.
133: In Section III we discuss gauge theory of two-component BEC
134: which has the vorticity interaction, and show that the theory
135: is closely related to Skyrme theory. With this we show that
136: the theory allows a topological knot similar to the one in
137: Skyrme theory. We also identify that this knot is a vorticity knot
138: very similar to the one in Gross-Pitaevskii theory of two-component BEC.
139: In Section IV we discuss the Landau-Ginzburg theory of two-gap
140: superconductor and show that the theory is closely related to
141: the gauge theory of two-component BEC. With this we argue
142: that the theory can also admit a knot, a twisted magnetic vortex
143: ring, similar to the vorticity knot in two-component BEC.
144: In doing so we also establish the non-Abelian flux quantization
145: in two-gap superconductor. We demonstrate the existence of magnetic
146: vortex whose flux is quantized in the unit $4\pi/g$, not $2\pi/g$,
147: in two-gap superconductor.
148: Finally in Section V we discuss physical implications of
149: our results.
150:
151: \section{Knot in Skyrme theory}
152:
153: The Skyrme theory has a rich topological structure.
154: It has skyrmion and baby skyrmion \cite{skyrme,piet}.
155: But recently it has been shown that the theory allows
156: a knotted soliton identical to
157: the Faddeev-Niemi knot in Skyrme-Faddeev
158: non-linear sigma model, which can be identified as a twisted
159: vortex ring made of helical baby skyrmion \cite{prl01,plb04}.
160: To see this, we review the knot in Skyrme theory first.
161:
162: Let $\omega$ and $\hn$ be the scalar field and
163: the non-linear sigma field in Skyrme theory. With
164: \bea
165: &U = \exp (\omega \dfrac{\vec \sigma}{2i} \cdot \hat n)
166: = \cos \dfrac{\omega}{2} - i (\vec \sigma \cdot \hat n)
167: \sin \dfrac{\omega}{2}, \nn\\
168: &L_\mu = U\partial_\mu U^{\dagger}, ~~~~~({\hat n}^2 = 1)
169: \eea
170: the Skyrme Lagrangian is expressed as
171: \bea
172: &{\cal L} = \dfrac{\mu^2}{4} {\rm tr} ~L_\mu^2 + \dfrac{\alpha}{32}
173: {\rm tr} \left( \left[ L_\mu, L_\nu \right] \right)^2 \nn \\
174: &= - \dfrac{\mu^2}{4} \Big[ \dfrac{1}{2} (\partial_\mu \omega)^2
175: +(1-\cos\omega)(\partial_\mu \hat n)^2 \Big] \nn \\
176: & -\dfrac{\alpha}{16} \Big[ \dfrac{1-\cos\omega}{2} (\partial_\mu
177: \omega \partial_\nu \hat n - \partial_\nu \omega \partial_\mu \hat n)^2 \nn\\
178: & + (1-\cos\omega)^2 (\partial_\mu \hat n \times \partial_\nu \hat
179: n)^2 \Big].
180: \label{slag}
181: \eea
182: The equation of motion is given by
183: \bea
184: &\dfrac{\mu^2}{4} \left[ \partial^2 \omega -\sin\omega
185: (\partial_\mu \hat n)^2 \right] \nn\\
186: &+\dfrac{\alpha}{32} \sin\omega (\partial_\mu
187: \omega \partial_\nu \hat n -\partial_\nu \omega \partial_\mu \hat n)^2 \nn\\
188: &+\dfrac{\alpha}{8} (1-\cos\omega)
189: \partial_\mu \big[ (\partial_\mu
190: \omega \partial_\nu \hat n -\partial_\nu \omega \partial_\mu \hat
191: n) \cdot \partial_\nu \hat n \big] \nn\\
192: & - \dfrac{\alpha}{8} (1-\cos\omega) \sin\omega (\partial_\mu \hat n \times
193: \partial_\nu \hat n)^2 =0, \nn \\
194: &\partial_\mu \Big\{ \dfrac{\mu^2}{4} (1-\cos\omega) \hat n \times
195: \partial_\mu \hat n \nn\\
196: & + \dfrac{\alpha}{16} (1-\cos\omega)
197: \big[ (\partial_\nu \omega)^2
198: \hat n \times \partial_\mu \hat n
199: -(\partial_\mu \omega \partial_\nu \omega) \hat n \times
200: \partial_\nu \hat n \big] \nn\\
201: &+\dfrac{\alpha}{8} (1-\cos\omega)^2
202: (\hat n \cdot \partial_\mu \hat n \times
203: \partial_\nu \hat n) \partial_\nu \hat n \Big\}=0.
204: \label{seq}
205: \eea
206: With the spherically symmetric ansatz
207: \bea
208: \omega = \omega (r),~~~~~\hat n = \hat r,
209: \eea
210: (3) is reduced to
211: \bea
212: &\dfrac{d^2 \omega}{dr^2} +\dfrac{2}{r} \dfrac{d\omega}{dr} -\dfrac{2
213: \sin\omega}{r^2}
214: -\dfrac{\alpha}{\mu^2} \Big[ \dfrac{\sin^2 (\omega /2)}{r^2}
215: \dfrac{d^2 \omega}{dr^2} \nn\\
216: &+\dfrac{\sin\omega}{2 r^2} (\dfrac{d\omega}{dr})^2
217: -\dfrac{2 \sin\omega \sin^2 (\omega /2)}{r^4} \Big] =0.
218: \eea
219: Imposing the boundary condition $\omega(0)= 2\pi$ and $\omega(\infty)= 0$,
220: one has the well-known skyrmion \cite{skyrme}.
221:
222: A remarkable point of Skyrme theory is that $\omega=\pi$,
223: independent of $\hn$, becomes a classical solution \cite{prl01}.
224: So restricting $\omega$ to $\pi$, one can reduce the Lagrangian
225: (\ref{slag}) to the Skyrme-Faddeev Lagrangian
226: \bea
227: {\cal L} \rightarrow -\dfrac{\mu^2}{2} (\partial_\mu
228: \hat n)^2-\dfrac{\alpha}{4}(\partial_\mu \hat n \times
229: \partial_\nu \hat n)^2,
230: \label{sflag}
231: \eea
232: in which case the equation of motion (\ref{seq}) is reduced to
233: \bea
234: &\hn \times
235: \partial^2 \hn + \dfrac{\alpha}{\mu^2} ( \partial_\mu N_{\mu\nu} )
236: \partial_\nu \hn = 0, \nn\\
237: &N_{\mu\nu} = \hn \cdot (\partial_\mu \hn \times \partial_\nu \hn)
238: =\pro_\mu C_\nu - \pro_\nu C_\mu.
239: \label{sfeq}
240: \eea
241: Notice that since $N_{\mu\nu}$ forms a closed two-form,
242: it always admits a $U(1)$ potential $C_\mu$.
243:
244: The equation (\ref{sfeq}) allows non-Abelian monopole,
245: baby skyrmion, helical baby skyrmion, and Faddeev-Niemi
246: knot as its solutions \cite{prl01,plb04}. But for our purpose it is
247: important to understand the helical baby skyrmion, because this
248: plays a crucial role for us to construct the knot.
249: So we review the helical baby skyrmion.
250:
251: To construct the desired helical baby skyrmion
252: let $(\varrho,\varphi,z)$ the cylindrical coodinates,
253: and choose the ansatz
254: \bea
255: &\n= \Bigg(\matrix{\sin{f(\varrho)}\cos{(mkz+n\varphi)} \cr
256: \sin{f(\varrho)}\sin{(mkz+n\varphi)} \cr \cos{f(\varrho)}}\Bigg).
257: \label{hvans}
258: \eea
259: With this the equation (\ref{sfeq}) is reduced to
260: \bea
261: &\Big(1+(m^2 k^2+\dfrac{n^2}{\varrho^2})
262: \dfrac{\sin^2{f}}{g^2 \rho^2}\Big) \ddot{f}
263: + \Big( \dfrac{1}{\varrho}+ 2\dfrac{\dot{\rho}}{\rho} \nn\\
264: &+(m^2 k^2+\dfrac{n^2}{\varrho^2})
265: \dfrac{\sin{f}\cos{f}}{g^2 \rho^2} \dot{f}
266: + \dfrac{1}{\varrho} (m^2 k^2-\dfrac{n^2}{\varrho^2})
267: \dfrac{\sin^2{f}}{g^2 \rho^2} \Big) \dot{f} \nn\\
268: &- (m^2 k^2+\dfrac{n^2}{\varrho^2}) \sin{f}\cos{f}=0.
269: \label{hveq}
270: \eea
271: So with the boundary condition
272: \bea
273: f(0)=\pi,~~f(\infty)=0,
274: \label{bc}
275: \eea
276: we obtain the non-Abelian vortex solutions shown in Fig.\ref{hbs}.
277: Notice that, when $m=0$, the solution describes
278: the well-known baby skyrmion \cite{piet}. But when $m$ is not zero,
279: it describes a helical baby skyrmion, a twisted magnetic vortex
280: which is periodic in $z$-coodinate \cite{plb04}.
281: In this case, the vortex has a quantized magnetic flux
282: not only along the $z$-axis but also around the $z$-axis.
283:
284: \begin{figure}[t]
285: \includegraphics[scale=0.5]{hbs.eps}
286: \caption{The baby skyrmion (dashed line) with $m=0,n=1$
287: and the helical baby skyrmion (solid line) with $m=n=1$ in Skyrme
288: theory. Here $\varrho$ is in the unit ${\sqrt \alpha}/\mu$
289: and $k=0.8~\mu/{\sqrt \alpha}$.}
290: \label{hbs}
291: \end{figure}
292:
293: The helical baby skyrmion will become
294: unstable unless the periodicity condition is enforced by
295: hand. But it plays a very important role
296: because one can make it a vortex ring by smoothly connecting
297: two periodic ends. In this case the vortex ring
298: acquires the topology of a knot, and thus becomes a knot
299: itself \cite{prl01,plb04}. In fact it becomes a knot made of two
300: magnetic fluxes linked together, whose knot topology is described by
301: the Chern-Simon index of the potential $C_\mu$,
302: \bea
303: &Q = \dfrac{1}{32\pi^2} \int \epsilon_{ijk} C_i N_{jk} d^3x = mn,
304: \label{kqn}
305: \eea
306: which describes the Hopf mapping $\pi_3(S^2)$ defined by $\hn$.
307:
308: Clearly the knot has a topological stability, because two
309: flux rings linked together can not be disjointed by a smooth
310: deformation of the field configuration.
311: Moreover the topological stability is
312: backed up by a dynamical stability. This is because
313: the knot can be viewed as two magnetic fluxes linked together,
314: and the magnetic flux trapped in the knot disk
315: can not be squeezed out. This provides
316: a stablizing repulsive force which prevent the collapse
317: of the knot \cite{plb04}.
318:
319: From our discussion it becomes clear that the existence of
320: the helical baby skyrmion is crucial for the existence of a topologically
321: stable knot. In the following we show that an identical mechanism
322: works in two-component BEC and two-gap superconductor
323: which guarantees the existence of a stable knot.
324:
325: \section{Knot in Two-component Bose-Einstein Condensates}
326:
327: The recent advent of multi-component BEC (in particular
328: the spin-1/2 condensate of $^{87}{\rm Rb}$ atoms)
329: has widely opened a new opportunity for
330: us to study novel topological objects which can not be
331: realized in ordinary (one-component) BEC \cite{bec,ruo}.
332: This is because the multi-component BEC naturally allows
333: a non-Abelian structure which accommodates a non-trivial
334: topological objects, in particular a topolgical knot which is
335: very similar to the knot in Skyrme theory \cite{bec1,bec5}.
336:
337: There are two competing theories of two-component BEC,
338: the popular Gross-Pitaevskii theory \cite{ruo} and the gauge theory
339: of two-component BEC proposed recently \cite{bec1}.
340: Both theories predict topological knots.
341: Many authors have already claimed
342: the existence of a knot in Gross-Pitaevskii theory \cite{ruo}.
343: Moreover it has been shown that this knot can be identified
344: as a vorticity knot which is made of
345: two vorticity fluxes linked together, whose topology $\pi_3(S^2)$
346: is fixed by the Chern-Simon index of the velocity potential
347: of the condensate \cite{bec5}.
348:
349: So in this section we discuss the gauge theory of two-component BEC,
350: and show that this theory also has a vorticity knot very similar
351: to the knot in Gross-Pitaevskii theory.
352: Consider a ``charged'' two-component BEC
353: described by a complex doublet $\phi$ interacting
354: ``electromagnetically'', which can be described by the following
355: non-relativistic gauged Gross-Pitaevskii type Lagrangian (we
356: will discuss the relativistic generalization later)
357: \bea
358: &{\cal L} = i \dfrac {\hbar}{2} \Big[\phi^\dag (\tilde D_t \phi)
359: -(\tilde D_t \phi)^\dag \phi \Big]
360: - \dfrac {\hbar^2}{2m} |\tilde D_i \phi|^2 \nn\\
361: & + \mu^2 \phi^\dag \phi - \dfrac {\lambda}{2}
362: (\phi^\dag \phi)^2 - \dfrac {1}{4} \tilde F_{\mu \nu} ^2,
363: \label{beclag}
364: \eea
365: where $\tilde D_\mu \phi = ( \pro_\mu - i g \tilde A_\mu ) \phi$,
366: and $\mu^2$ and $\lambda$ are the chemical potential
367: and the quartic coupling constant.
368: Normalizing $\phi$ to $(\sqrt{2m}/\hbar)\phi$ and putting
369: \bea
370: \phi = \dfrac {1}{\sqrt 2} \rho \xi , ~~~~~(\xi^\dag \xi = 1)
371: \eea
372: we have the following Hamiltonian in the static limit,
373: \bea
374: &{\cal H} = \dfrac {1}{2} (\pro_i \rho)^2
375: + \dfrac {\rho^2}{2} |\tilde D_i \xi |^2 \nn \\
376: &- \dfrac {\mu^2}{2} \rho^2 + \dfrac{\lambda}{8} \rho^4
377: + \dfrac {1}{4} \tilde F_{ij} ^2.
378: \label{becham}
379: \eea
380: Notice that here we have also
381: normalized $\mu^2$ and $\lambda$ to
382: $(\hbar^2/2m) \mu^2$ and $(\hbar^2/2m)^2 \lambda$.
383: From now on we will use the normalized Hamiltonian (\ref{becham}).
384:
385: At this point it is important to realize that the ``electromagnetic''
386: interaction here should be self-induced, since we are dealing
387: with neutral condensates. So we identify
388: the ``electromagnetic'' potential by the velocity field of
389: the doublet $\xi$,
390: \bea
391: \tilde A_\mu = -\dfrac {i}{g} \xi^\dag \pro_\mu \xi.
392: \label{am}
393: \eea
394: A justification for this is that the $U(1)$ gauge invariance almost
395: dictates us to identify the velocity field as the gauge potential.
396: Indeed in the absence of the Maxwell term, (\ref{am}) becomes
397: an equation of motion. With this identification
398: the field strength becomes the vorticity field
399: \bea
400: &\tilde F_{\mu\nu} = \pro_\mu \tilde A_\nu - \pro_\nu \tilde A_\mu \nn\\
401: &= -\dfrac {i}{g} (\pro_\mu \xi^\dag \pro_\nu \xi
402: - \pro_\nu \xi^\dag \pro_\mu \xi).
403: \label{fmn}
404: \eea
405: One might wonder why we have included
406: the vorticity interaction in the Lagrangian (\ref{beclag}),
407: when this is absent in the Gross-Pitaevskii Lagrangian.
408: The reason is because it costs energy to create
409: the vorticity in two-component BEC. In ordinary BEC one does
410: not have to worry about the vorticity because the vorticity is identically
411: zero (since the velocity is given by the gradient of the phase
412: of the one-component complex condensate). But a non-Abelian
413: (multi-component) BEC has a non-vanishing vorticity,
414: in which case it is natural to include
415: the vorticity interaction in the Lagrangian \cite{bec1,bec5}.
416:
417: With (\ref{am}) the Hamiltonian (\ref{becham}) becomes
418: \bea
419: &{\cal H} = \dfrac {1}{2} (\pro_i \rho)^2 + \dfrac {\rho^2}{2}\Big(|\pro_i
420: \xi |^2 - |\xi^\dag \pro_i \xi|^2 \Big)
421: - \dfrac {\lambda}{8} (\rho^2 - \rho^2_0)^2 \nn\\
422: &- \dfrac {1}{4 g^2} (\pro_i \xi^\dag \pro_j \xi
423: - \pro_j \xi^\dag \pro_i \xi)^2, \nn\\
424: &\rho^2_0= \dfrac{2\mu^2}{\lambda}.
425: \label{becham1}
426: \eea
427: Notice that now the coupling constant $g$ represents
428: the coupling strength of the vorticity.
429: Minimizing the Hamiltonian we have
430: the following equation of motion
431: \bea
432: & \pro^2 \rho - \Big(|\pro_i \xi |^2 - |\xi^\dag \pro_i \xi|^2 \Big)
433: \rho = \dfrac {\lambda}{2} (\rho^2 - \rho_0^2) \rho,\nn \\
434: &\Big\{(\pro^2 - \xi^\dag \pro^2 \xi) + 2 (\dfrac {\pro_i
435: \rho}{\rho} - \xi^\dag \pro_i\xi)(\pro_i - \xi^\dag \pro_i \xi) \nn\\
436: &- \dfrac {1}{g^2 \rho^2} \Big[\pro_i (\pro_i
437: \xi^\dag \pro_j \xi - \pro_j \xi^\dag \pro_i \xi) \Big]
438: (\pro_j - \xi^\dag \pro_j \xi) \Big\} \xi \nn\\
439: &= 0.
440: \label{beckeq1}
441: \eea
442: To understand the meaning of this notice that with
443: \bea
444: \n = \xi^\dag \vec \sigma \xi,
445: \label{hn}
446: \eea
447: we have
448: \bea
449: & |\pro_\mu \xi|^2 - |\xi^\dag \pro_\mu \xi|^2
450: = \dfrac{1}{4} (\pro_\mu \hn)^2, \nn\\
451: &i (\pro_\mu \xi^\dag \pro_\nu \xi
452: - \pro_\nu \xi^\dag \pro_\mu \xi )
453: = \dfrac{1}{2} \hn \cdot (\pro_\mu \hn \times \pro_\nu \hn) \nn\\
454: &= \dfrac{1}{2} N_{\mu\nu},
455: \label{id}
456: \eea
457: where $N_{\mu\nu}$ is mathematically identical to what we have
458: in Skyrme theory in (\ref{sfeq}).
459: This tells that the potential $C_\mu$ for the two-form $N_{\mu\nu}$
460: in (\ref{sfeq}) is given by (up to a gauge transformation)
461: \bea
462: C_\mu = 2g \tilde A_\mu = -2i \xi^\dag \pro_\mu \xi.
463: \label{cm}
464: \eea
465: More importantly, with (\ref{id}) the Hamiltonian (\ref{becham1})
466: can be expressed as
467: \bea
468: &{\cal H} = \dfrac{1}{2} (\pro_i \rho)^2
469: + \dfrac{\rho^2}{8} (\pro_i \hn)^2
470: + \dfrac{\lambda}{8}(\rho^2-\rho^2_0)^2 \nn\\
471: &+ \dfrac{1}{16 g^2} (\pro_i \hn \times \pro_j \hn)^2.
472: \eea
473: So the theory becomes very similar to Skyrme-Faddeev
474: theory, which strongly indicates that
475: the gauge theory of two-component BEC can allow a knot.
476: As importantly this strongly implies that the Skyrme-Faddeev theory
477: could play an important role in condensed matter
478: physics.
479:
480: To simplify the equation (\ref{beckeq1}) notice that
481: from (\ref{am}) and (\ref{hn}) we have
482: \bea
483: \pro_\mu \xi = (ig \tilde A_\mu + \dfrac{1}{2} \vec \sigma
484: \cdot \pro_\mu \n) \xi.
485: \eea
486: With the identity the second equation of (\ref{beckeq1}) is reduced to
487: \bea
488: &(A + \B \cdot \vec \sigma) \xi = 0, \nn\\
489: &A = (\pro_i \n)^2, \nn \\
490: &\B = \pro^2 \n + 2 \dfrac{\pro_i \rho}{\rho} \pro_i \n
491: - \dfrac{i}{2 g^2 \rho^2} \pro_i N_{ij} \pro_j \n,
492: \eea
493: which can be written as
494: \bea
495: &A+ \vec B \cdot \hn=0, \nn\\
496: & \hn \times \vec B - i \hn \times (\hn \times \vec B) =0,
497: \eea
498: or
499: \bea
500: \n \times (\pro^2 \n + 2 \dfrac{\pro_i \rho}{\rho} \pro_i \n)
501: + \dfrac{1}{2 g^2 \rho^2} \pro_i N_{ij} \pro_j \n = 0.
502: \eea
503: So we can put (\ref{beckeq1}) into the form
504: \bea
505: &\pro^2 \rho - \dfrac{1}{4} (\pro_i \n)^2 \rho
506: = \dfrac{\lambda}{2}(\rho^2 - \rho_0^2)
507: \rho, \nn \\
508: &\n \times \pro^2 \n + 2 \dfrac{\pro_i \rho}{\rho} \n \times \pro_i
509: \n + \dfrac{1}{g^2\rho^2} \pro_i N_{ij} \pro_j \n = 0.
510: \label{beckeq2}
511: \eea
512: This is the equation of two-component BEC that we are
513: looking for. The similarity between this and the equation
514: (\ref{sfeq}) of Skyrme-Faddeev theory is unmistakable.
515:
516: Notice that the target space of $\xi$ and $\n$ is $S^3$ and
517: $S^2$, but here we have transformed the equation for $\xi$
518: in (\ref{beckeq1}) completely into the equation
519: for $\n$ in (\ref{beckeq2}).
520: This is made possible because, with the Abelian gauge invariance of (6),
521: the physical target space of $\xi$ becomes the gauge orbit space
522: $S^2 = S^3/S^1$ which forms a $CP^1$ space.
523: This means that we can view the theory as a self interacting
524: $CP^1$ theory (coupled to a scalar field $\rho$), and replace
525: $\xi$ completely in terms of $\n$.
526:
527: To show that the theory has a knot solution we construct
528: a helical vortex solution first. To do this
529: we choose the ansatz
530: \bea
531: &\rho= \rho(\varrho),
532: ~~~~~\xi = \Bigg( \matrix{\cos \dfrac{f(\varrho)}{2}
533: \exp (-in\varphi) \cr
534: \sin \dfrac{f(\varrho)}{2} \exp (imkz)} \Bigg), \nn\\
535: &\n= \xi^\dag \vec \sigma \xi
536: = \Bigg(\matrix{\sin{f(\varrho)}\cos{(mkz+n\varphi)} \cr
537: \sin{f(\varrho)}\sin{(mkz+n\varphi)} \cr \cos{f(\varrho)}}\Bigg). \nn\\
538: \label{bhvans}
539: \eea
540: With this (\ref{beckeq2}) is reduced to
541: \bea
542: &\ddot{\rho}+\dfrac{1}{\varrho}\dot\rho
543: - \dfrac{1}{4}\Big(\dot{f}^2+(m^2 k^2+\dfrac{n^2}{\varrho^2})
544: \sin^2{f}\Big)\rho \nn\\
545: &= \dfrac{\lambda}{2}(\rho^2-\rho_0^2)\rho, \nn\\
546: &\Big(1+(m^2 k^2+\dfrac{n^2}{\varrho^2})
547: \dfrac{\sin^2{f}}{g^2 \rho^2}\Big) \ddot{f}
548: + \Big( \dfrac{1}{\varrho}+ 2\dfrac{\dot{\rho}}{\rho} \nn\\
549: &+(m^2 k^2+\dfrac{n^2}{\varrho^2})
550: \dfrac{\sin{f}\cos{f}}{g^2 \rho^2} \dot{f}
551: + \dfrac{1}{\varrho} (m^2 k^2-\dfrac{n^2}{\varrho^2})
552: \dfrac{\sin^2{f}}{g^2 \rho^2} \Big) \dot{f} \nn\\
553: &- (m^2 k^2+\dfrac{n^2}{\varrho^2}) \sin{f}\cos{f}=0.
554: \label{bveq}
555: \eea
556: So with the boundary condition
557: \bea
558: &\dot \rho(0)=0,~~~~~\rho(\infty)=\rho_0, \nn\\
559: &f(0)=\pi,~~~~~f(\infty)=0,
560: \label{becbc}
561: \eea
562: we obtain the non-Abelian vortex solutions shown in Fig.\ref{becheli}.
563:
564: \begin{figure}[t]
565: \includegraphics[scale=0.5]{becknot.eps}
566: \caption{The non-Abelian vortex (dashed line) with $m=0,n=1$ and
567: the helical vortex (solid line) with $m=n=1$ in the gauge theory
568: of two-component BEC. Here we have put $\lambda/g^2=1$,
569: $k=0.64~{\sqrt \lambda}\rho_0$, and $\varrho$ is in the unit of
570: $1/{\sqrt \lambda}\rho_0$.}
571: \label{becheli}
572: \end{figure}
573:
574: There are three points that have to be
575: emphasized here. First, when $m=0$, the solution describes
576: the untwisted non-Abelian vortex \cite{bec1}. But when $m$ is not zero,
577: it describes a helical vortex which is periodic in $z$-coodinate.
578: In this case, the vortex has a non-vanishing velocity current
579: (not only around the vortex but also) along the $z$-axis.
580: Secondly, the doublet $\phi$ starts from
581: the second component at the core,
582: but the first component takes over completely at the infinity.
583: This is due to the boundary condition $f(0)=\pi$ and $f(\infty)=0$,
584: which assures that our solution describes a genuine
585: non-Abelian vortex. Thirdly, when $f=0$ (or $f=\pi$)
586: the doublet effectively becomes
587: a singlet, and (\ref{bveq}) describes the well-known
588: vortex in single-component BEC. Only when $f$ has a non-trivial
589: profile, we have a non-Abelian vortex.
590:
591: In Skyrme theory the helical vortex is interpreted as
592: a twisted magnetic vortex whose flux is quantized \cite{prl01,plb04}
593: Now we show that the above vortex is a twisted vorticity flux
594: which is also quantized.
595: To see this notice that the non-Abelian structure of
596: the vortex is represented by the doublet $\xi$.
597: Moreover, the velocity field
598: of the doublet is given by
599: \bea
600: &\tilde A_\mu =-\dfrac{n}{2g}(\cos{f}+1) \pro_\mu \varphi \nn\\
601: &-\dfrac{mk}{2g}(\cos{f}-1) \pro_\mu z,
602: \label{gpvel}
603: \eea
604: which generates the vorticity
605: \bea
606: &H_{\mu\nu}= -\dfrac{i}{g}(\pro_\mu \xi^{\dagger} \pro_\nu \xi
607: -\pro_\nu \xi^{\dagger} \pro_\mu \xi) \nn\\
608: &=\dfrac{\dot{f}}{2g} \sin{f}\Big(n(\pro_\mu \varrho \pro_\nu \varphi
609: -\pro_\nu \varrho \pro_\mu \varphi) \nn\\
610: &+mk(\pro_\mu \varrho \pro_\nu z
611: - \pro_\nu \varrho \pro_\mu z) \Big).
612: \label{gpvor}
613: \eea
614: This has two vorticity fluxes, $\phi_{\hat z}$
615: along the $z$-axis
616: \bea
617: &\phi_{\hat z}=\dfrac{}{} \int H_{{\hat \varrho}{\hat \varphi}}
618: \varrho d \varrho d \varphi
619: = -\dfrac{2\pi n}{g},
620: \label{gpfluxz}
621: \eea
622: and $\phi_{\hat \varphi}$ around the the $z$-axis (in one
623: period section from $z=0$ to $z=2\pi/k$)
624: \bea
625: &\phi_{\hat \varphi}=\dfrac{}{} \int_{z=0}^{z=2\pi/k}
626: H_{{\hat z}{\hat \varrho}} d \varrho dz
627: = \dfrac{2\pi m}{g}.
628: \label{gpfluxphi}
629: \eea
630: This shows that the helical vortex is made of two quantized vorticity
631: fluxes, the $\phi_{\hat z}$-flux which is concentrated at the core
632: and the $\phi_{\hat \varphi}$-flux which surrounds it.
633: This confirms that the helical vortex is a twisted vorticity
634: flux which is very similar to the helical vortex in Gross-Pitaevskii
635: theory \cite{bec5}.
636:
637: Now, with the helical vortex, one can easily make a twisted vortex ring
638: smoothly connecting the periodic ends together.
639: And just as in Skyrme theory the twisted vortex ring
640: becomes a topological knot. But
641: here it is $\pi_3(S^3)$ of the doublet $\xi$ which provides the
642: non-trivial quantum number,
643: \bea
644: q = - \dfrac {1}{4\pi^2} \int \epsilon_{ijk} \xi^{\dagger}
645: \partial_i \xi ( \partial_j \xi^{\dagger}
646: \partial_k \xi ) d^3 x.
647: \label{beckqn}
648: \eea
649: Of course this is identical to the expression (\ref{kqn}), due to the
650: Hopf fibering of $S^3$ to $S^2 \times S^1$ \cite{bec5}. This
651: tells that we can express the knot quantum number either
652: by $\pi_3(S^3)$ or by $\pi_3(S^2)$.
653:
654: Obviously this knot has a topological stability, because
655: the knot topology (\ref{kqn}) can not be changed by a smooth
656: deformation of the field configuration. Moreover it has
657: a dynamical stability. To understand this notice that
658: the knot has a twisted velocity field so that it has a
659: non-vanishing velocity around the $z$-axis.
660: This means that it carries a non-vanishing angular momentum along
661: the $z$-axis. And this angular momentum provides
662: the dynamical stability, because it creates a centrifugal force that
663: prevents the collapse of the knot. Notice that this dynamical
664: stability originates from the knot topology, because
665: the angular momentum comes from the twisted velocity field.
666: In this sense the topological stability and the dynamical stability
667: have one and the same origin. It is this remarkable
668: interplay between topology and dynamics which assures
669: the stability of the knot. The nontrivial twisted topology of
670: the knot expresses itself in the form of the angular momentum,
671: which in turn provides the dynamical stability of the knot.
672: This presence of the angular momnetum is
673: what differentiates our
674: knot from the untwisted Abrikosov-type vortex ring
675: which has no dynamical stability.
676:
677: There have been assertions that two-component BEC admits
678: a knot \cite{ruo}. But notice that this knot is based on
679: Gross-Pitaevskii theory of two-component BEC, which has no
680: vorticity interaction. In contrast our knot is based on
681: the gauge theory in which the vorticity
682: interaction plays a crucial role. Nevertheless physically two knots
683: are very similar \cite{bec5}. Both can be identified as
684: a vorticity knot. This implies that both theories should be
685: taken seriously as a theory of two-component BEC.
686:
687: \section{Knot in Two-gap Superconductors}
688:
689: In the above gauge theory of spin-1/2
690: condensates the gauge interaction was a
691: self-induced interaction. But when the doublet is charged,
692: the gauge interaction can be treated as independent.
693: In this case the theory can describe a two-gap superconductor.
694: But even in this case the knot topology and thus the knot itself
695: should survive. This implies that two-gap
696: superconductor should also have
697: a topological knot.
698:
699: The knot in two-gap superconductor could be either relativistic
700: or non-relativistic, and appear in both Abelian and non-Abelian
701: setting \cite{cm2}. In this paper we will discuss the relativistic
702: knot in the Abelian setting (a non-relativistic
703: Gross-Pitaevskii type theory gives an identical result).
704: Consider a charged doublet scalar field $\phi$ coupled to
705: the real electromagnetic field,
706: \bea
707: &{\cal L} = - |D_\mu \phi|^2 + \mu^2 \phi^{\dagger}\phi
708: - \dfrac{\lambda}{2} (\phi^{\dagger} \phi)^2
709: - \dfrac{1}{4} F_{\mu \nu}^2 , \nn\\
710: &D_\mu \phi = (\partial_\mu - ig A_\mu) \phi.
711: \label{sclag}
712: \eea
713: The Lagrangian has the equation of motion
714: \bea
715: &D^2\phi =\lambda(\phi^{\dagger} \phi
716: -\dfrac{\mu^2}{\lambda})\phi, \nn\\
717: &\partial_\mu F_{\mu \nu} = j_\nu = i g \Big[(D_\nu
718: \phi)^{\dagger}\phi - \phi ^{\dagger}(D_\nu \phi) \Big].
719: \label{sceq1}
720: \eea
721: Now, with
722: \bea
723: \phi =\dfrac{1}{\sqrt 2} \rho \xi,~~~~~{\xi}^{\dagger}\xi = 1,
724: ~~~~~\hat n = \xi^{\dagger} \vec \sigma \xi,
725: \eea
726: we can reduce (\ref{sceq1}) to \cite{cm2}
727: \bea
728: &\partial ^2 \rho - \Big(
729: \dfrac{1}{4} (\partial _\mu \hat n)^2+ g^2
730: (A_\mu + \tilde A_\mu)^2 \Big) \rho
731: = \dfrac{\lambda}{2} (\rho^2-\rho_0^2)\rho, \nn\\
732: &\hat n \times \partial ^2 \hat n + 2 \dfrac{\partial_\mu \rho}{
733: \rho} \hat n \times \partial_\mu \hat n
734: + \dfrac{2}{g\rho^2} \partial_\mu F_{\mu\nu} \partial_\nu \hat n =0, \nn\\
735: &\partial_\mu F_{\mu\nu} = j_\mu
736: =g^2 \rho^2 (A_\mu + \tilde A_\mu), \nn\\
737: &\tilde A_\mu=-\dfrac{i}{g} \xi^{\dagger}\partial_\mu \xi,
738: ~~~~~\rho_0=\dfrac{2\mu^2}{\lambda}.
739: \label{sceq2}
740: \eea
741: This is the equation for two-gap superconductor.
742: Notice that with $A_\mu=-\tilde A_\mu$ the first two equations
743: reduce to (\ref{beckeq2}). This tells that the gauge theory of
744: two-component BEC and the above theory of two-gap superconductor
745: are closely related.
746:
747: To obtain the desired knot we first construct a superconducting
748: helical magnetic vortex. Let
749: \bea
750: &\rho=\rho(\varrho),
751: ~~~~~\xi = \Bigg( \matrix{\cos \dfrac{f(\varrho)}{2} \exp (-in\varphi) \cr
752: \sin \dfrac{f(\varrho)}{2} \exp (imkz)} \Bigg), \nn\\
753: &A_\mu= \dfrac{1}{g} \big(n A_1(\varrho) \partial_\mu\varphi
754: + mk A_2(\varrho) \partial_\mu z \big), \nn\\
755: &\n= \xi^\dag \vec \sigma \xi
756: = \Bigg(\matrix{\sin{f(\varrho)}\cos{(n\varphi+mkz)} \cr
757: \sin{f(\varrho)}\sin{(n\varphi+mkz)} \cr \cos{f(\varrho)}}\Bigg), \nn\\
758: &\tilde A_\mu=-\dfrac{n}{2g} \big(\cos{f(\varrho)}+1\big)
759: \partial_\mu \varphi \nn\\
760: &- \dfrac{mk}{2g} \big(\cos{f(\varrho)}-1\big) \partial_\mu z.
761: \label{scans}
762: \eea
763: With this we have
764: \bea
765: &j_\mu = g\rho^2 \Big(n \big(A_1-\dfrac{\cos{f}+1}{2}\big)
766: \partial_\mu \varphi \nn\\
767: &+ mk\big(A_2-\dfrac{\cos{f}-1}{2}\big)
768: \partial_\mu z \Big),
769: \label{sc}
770: \eea
771: and (\ref{sceq2}) becomes
772: \bea
773: &\ddot{\rho}+\dfrac{1}{\varrho}\dot\rho
774: - \Big[\dfrac{1}{4} \Big(\dot{f}^2
775: +\big(\dfrac{n^2}{\varrho^2} + m^2 k^2 \big) \sin^2{f}\Big) \nn\\
776: &+\dfrac{n^2}{\varrho^2} \Big(A_1-\dfrac{\cos{f}+1}{2}\Big)^2
777: + m^2 k^2 \Big(A_2-\dfrac{\cos{f}-1}{2}\Big)^2 \Big]\rho \nn\\
778: &= \dfrac{\lambda}{2}(\rho^2-\rho_0^2)\rho, \nn\\
779: &\ddot{f} + \big(\dfrac{1}{\varrho}
780: +2\dfrac{\dot{\rho}}{\rho} \big)\dot{f}
781: -2 \Big(\dfrac{n^2}{\varrho^2}\big(A_1-\dfrac{1}{2}\big) \nn\\
782: &+ m^2 k^2 \big(A_2+\dfrac{1}{2}\big) \Big)\sin{f} = 0, \nn\\
783: &\ddot{A}_1-\dfrac{1}{\varrho}\dot{A}_1 -g^2 \rho^2
784: \Big(A_1-\dfrac{\cos{f}+1}{2}\Big) = 0, \nn\\
785: &\ddot{A}_2+\dfrac{1}{\varrho}\dot{A}_2 -g^2 \rho^2
786: \Big(A_2-\dfrac{\cos{f}-1}{2}\Big) = 0.
787: \label{sceq3}
788: \eea
789: Now, we impose the following boundary condition for the non-Abelian
790: vortices \cite{cm2},
791: \bea
792: &\rho (0) = 0,~~~\rho(\infty) = \rho_0,
793: ~~~f (0) = \pi,~~~f (\infty) = 0, \nn\\
794: & A_1 (0) = -1,~~~A_1 (\infty) =1.
795: \label{scbc}
796: \eea
797: This need some explanation, because the boundary
798: value $A_1(0)$ is chosen to be $-1$, not $0$. This is to
799: assure the smoothness of $\rho(\varrho)$ and $f(\varrho)$
800: at the origin. Only with this boundary value they
801: become analytic at the origin.
802: At this point one might object the boundary condition, because it creates
803: an apparent singularity in the gauge potential at the origin.
804: But notice that this singularity is an unphysical
805: (coordinate) singularity which can easily be removed by a gauge transformation.
806: In fact the singularity disappears with the gauge transformation
807: \bea
808: \phi \rightarrow \phi \exp(in\varphi),
809: ~~~~A_\mu \rightarrow A_\mu + \dfrac{n}{g} \partial_\mu \varphi,
810: \eea
811: which changes the boundary condition $A_1(0)=-1$ and
812: $A_1(\infty)=1$ to $A_1(0)=0$ and $A_1(\infty) =2$.
813: Mathematically this boundary condition
814: has a deep origin, which has to do with the fact
815: that the Abelian $U(1)$ runs from $0$ to $2\pi$, but the $S^1$
816: fiber of $SU(2)$ runs from $0$ to $4\pi$ \cite{cm2}.
817: As for $A_2(\varrho)$, we choose $A_2(\infty)=0$
818: to make the supercurrent vanishing at infinity
819: and require the vortex superconducting. As we will see,
820: this requires a logarithmic divergence for $A_2(0)$.
821: The boundary condition will
822: have an important consequence in the following.
823:
824: \begin{figure}
825: \includegraphics[scale=0.7]{nascknot.eps}
826: \caption{The non-Abelian vortex (dashed line) with
827: $m=0,n=1$ and the helical vortex (solid line) with $m=n=1$ in
828: two-gap superconductor. Here we have put $g=1,~\lambda=2$,
829: $k=\rho_0/10$, and $\varrho$ is in the unit of $1/\rho_0$.
830: Notice that $A_2$ has a logarithmic singularity at the origin.}
831: \label{heli-sc}
832: \end{figure}
833:
834: With the boundary condition we can integrate (\ref{sceq3})
835: and obtain the non-Abelian vortex solution
836: of the two-gap superconductor, which is shown in Fig.\ref{heli-sc}.
837: The solution is very similar to the one we have in
838: two-component BEC. When $m=0$, the solution (with $A_2=0$)
839: describes an untwisted non-Abelian vortex \cite{cm2}.
840: But when $m$ is not zero, it describes a helical magnetic vortex
841: which is periodic in $z$-coordinate.
842: Moreover, the vortex starts from
843: the second component at the core,
844: but the first component takes over completely at the infinity.
845: This is due to the boundary condition $f(0)=\pi$ and $f(\infty)=0$,
846: which assures that our solution describes a genuine
847: non-Abelian vortex. This is true even when $m=0$. Only when
848: $f=0$ (or $f=\pi$) the doublet effectively becomes
849: a singlet, and (\ref{sceq3}) describes the Abelian
850: Abrikosov vortex of one-gap superconductor.
851:
852: There are important differences between the non-Abelian vortex
853: and the Abrikosov vortex. First the non-Abelian vortex
854: has a non-Abelian magnetic flux quantization \cite{cm2}.
855: Indeed the quantized magnetic flux $\hat \phi_z$
856: of the non-Abelian vortex along the $z$-axis is given by
857: \bea
858: &H_z= \dfrac{n}{g} \dfrac{\dot A_1}{\varrho}, \nn\\
859: &\hat \phi_z = \dfrac{}{} \int H_z d^2x
860: = \dfrac{2\pi n}{g} \big[A_1(\infty) - A_1(0)\big] \nn\\
861: &= \dfrac{4\pi n}{g}.
862: \label{zflux}
863: \eea
864: Notice that the unit of the non-Abelian flux is $4\pi/g$,
865: not $2\pi/g$. This is a direct
866: consequence of the boundary condition (\ref{scbc}).
867: This non-Abelian quantization of magnetic flux comes
868: from the non-Abelian topology $\pi_2(S^2)$ of the doublet $\xi$,
869: or equivalently the triplet $\hn$, whose topological
870: quantum number is given by \cite{cm2}
871: \bea
872: &q = \dfrac {g}{4\pi } \int H_z d^2 x
873: = - \dfrac {1}{4\pi} \int \epsilon_{ij} \partial_i \xi^{\dagger}
874: \partial_j \xi d^2 x \nn\\
875: &= \dfrac {1}{8\pi} \int \epsilon_{ij} \hn \cdot (\partial_i \hn
876: \times \partial_j \hn) d^2 x = n.
877: \label{scqn}
878: \eea
879: This distinguishes our non-Abelian vortex from
880: the Abelian vortex whose topology is fixed by $\pi_1(S^1)$.
881:
882: Another important feature of the non-Abelian vortex
883: is that it carries a non-vanishing supercurrent
884: along the $z$-axis,
885: \bea
886: &i_z = mkg \dfrac{}{} \int \rho^2 \big(A_2
887: -\dfrac{\cos f +1}{2}\big)\varrho d\varrho d\varphi \nn\\
888: &= \dfrac{2\pi mk}{g} \int \big(\ddot A_2
889: +\dfrac{1}{\varrho} \dot A_2 \big) \varrho d\varrho \nn\\
890: &= \dfrac{2\pi mk}{g} (\varrho \dot A_2) \Big|_{\varrho=0}^{\varrho=\infty}
891: =-\dfrac{2\pi mk}{g} (\varrho \dot A_2) \Big|_{\varrho=0}.
892: \label{scz}
893: \eea
894: This is due to the logarithmic divergence of $A_2$ at the origin.
895:
896: Notice that the superconducting helical vortex
897: has only a heuristic value, because one needs
898: an infinite energy to create it (since the magnetic flux
899: around the vortex becomes divergent because of the singularity
900: of $A_2$ at the origin).
901: With the helical vortex, however, one can make
902: a vortex ring by smoothly bending and connecting
903: two periodic ends. In the vortex ring the infinite magnetic
904: flux of $A_2$ can be made finite making the finite supercurrent
905: (\ref{scz}) of the vortex ring produce a finite flux,
906: and we can fix the flux to have the value $4\pi m/g$
907: by adjusting the current with
908: $k$. With this the vortex ring now
909: becomes a topologically stable knot.
910:
911: To see this notice that the doublet $\xi$, after forming a knot,
912: acquires a non-trivial topology $\pi_3(S^2)$ which provides
913: the knot quantum number,
914: \bea
915: &Q = - \dfrac {1}{4\pi^2} \int \epsilon_{ijk} \xi^{\dagger}
916: \partial_i \xi ( \partial_j \xi^{\dagger}
917: \partial_k \xi ) d^3 x \nn\\
918: &= \dfrac{g^2}{32\pi^2} \int \epsilon_{ijk} C_i
919: (\partial_j C_k - \partial_k C_j) d^3x
920: =mn.
921: \label{bkqn}
922: \eea
923: This is nothing but the Chern-Simon index of the potential
924: $C_\mu$, which is mathematically identical to
925: the quantum number of the knots we discussed before.
926: This tells that our
927: knot is also made of two quantized magnetic flux rings
928: linked together whose knot quantum number is fixed by
929: the linking number $mn$. Obviously two flux rings linked together
930: can not be separated by any continuous deformation of
931: the field configuration. This provides the topological stability
932: of the knot.
933:
934: Again this topological stability is backed up
935: by a dynamical stability. To see this notice that the supercurrent
936: of the knot has two components, the one around
937: the knot tube which confines the magnetic flux along the knot,
938: but more importantly the other along the knot
939: which creates a magnetic flux
940: passing through the knot disk. This component of supercurrent
941: along the knot now generates a net
942: angular momentum which provides
943: the centrifugal repulsive force preventing the knot to collapse.
944: This makes the knot dynamically stable.
945:
946: To compare our knot with the Abrikosov vortex ring (made of the
947: Abrikosov vortex in conventional superconductor),
948: notice that the Abrikosov knot is empty (i.e.,
949: does not carry a net supercurrent).
950: As importantly it is unstable, and collapses immediately.
951: In contrast our knot has a helical supercurrent, and is stable.
952: Furthermore these two features are deeply related.
953: The helical supercurrent plays a crucial role to stablize the vortex
954: ring by providing the net angular momentum, which prevents the
955: collapse of the vortex ring. And this helical supercurrent
956: originates from the knot topology. This remarkable
957: interplay between topology and dynamics is what provides
958: the stability of the knot. The nontrivial topology expresses
959: itself in the form of the helical supercurrent, which in turn
960: provides the dynamical stability of the knot. We emphasize that
961: this supercurrent is what distinguishes our knot from
962: the Abrikosov vortex ring, which has neither topological
963: nor dynamical stability.
964:
965: \section{Discussion}
966:
967: \begin{figure}[t]
968: \includegraphics[scale=0.7]{2schv.eps}
969: \caption{The regular helical magnetic vortex with $m=n=1$ in
970: two-gap superconductor. Here we have put $g=1,~\lambda=2$,
971: $k=0.12 \rho_0$, $A_2(0)=0.5$, and $\varrho$ is in the unit of $1/\rho_0$.
972: Notice that the solution is completely regular.}
973: \label{2schv}
974: \end{figure}
975:
976: In this paper we have presented a compelling argument
977: for the existence of topological
978: knots in two-component BEC and two-gap superconductor.
979: Similar knots have popped out almost everywhere,
980: in particular in high energy physics in QCD \cite{plb05} and
981: Weinberg and Salam model \cite{ewknot}.
982: But we emphasize that at the center of these
983: topological objects lies the baby skyrmion and the Faddeev-Niemi
984: knot \cite{prl01,plb04}. In fact, our helical vortices
985: and knots in this paper are a straightforward
986: generalization of the baby skyrmion and the Faddeev-Niemi
987: knot.
988:
989: It has been assumed that the topological
990: objects in Skyrme theory can only be realized at high energy,
991: at the hadronic scale. But our analysis shows that
992: similar objects could exist in a completely different environment,
993: at a much lower scale, in low energy condensed matters.
994: If so, the challenge now is
995: to verify the existence of the topological
996: knot experimentally in condensed matters. Constructing the knot
997: might not be so easy at present moment.
998: Nevertheless, with some experimental ingenuity,
999: one should be able to construct the knots
1000: in condensed matters.
1001:
1002: Note Added: One might doubt the existence of a superconducting
1003: knot because the superconducting helical vortex we discussed
1004: in Section IV was singular (and thus unphysical). In this note
1005: we report a regular superconducting helical vortex which has a finite
1006: magnetic flux around the axis and thus a finite energy.
1007: The regular solution is obtained linking $A_2(0)$ with $k$.
1008: For example for $k=0.12$ we obtain the regular solution shown
1009: in Fig.\ref{2schv}, with $A_2(0)=0.5$. This type of regular
1010: helical vortex has vanishing supercurrent $i_z$,
1011: but could still be called superconducting
1012: because it has a non-trivial supercurrent density $j_z$
1013: which generates a net magnetic flux $H_\varphi$ around the vortex.
1014: The regular helical vortex strongly support the existence of
1015: a regular knot. The details will be published elsewhere.
1016:
1017: {\bf ACKNOWLEDGEMENT}
1018:
1019: ~~~The work is supported in part by the ABRL Program of
1020: Korea Science and Engineering Foundation (R14-2003-012-01002-0)
1021: and by the BK21 Project of the Ministry of Education.
1022:
1023: \begin{thebibliography}{99}
1024: \bibitem{dirac}P. A. M. Dirac, Proc. Roy. Soc. {\bf A113}, 60 (1931);
1025: Phys. Rev. {\bf 74}, 817 (1948).
1026: \bibitem{abri} A. Abrikosov, Sov. Phys. JETP {\bf 5}, 1174 (1957).
1027: \bibitem{skyrme}T. H. R. Skyrme, Proc. Roy. Soc. (London) {\bf 260}, 127
1028: (1961); {\bf 262}, 237 (1961); Nucl. Phys. {\bf 31}, 556 (1962).
1029: See also, for example, I. Zahed and G. Brown,
1030: Phys. Rep. {\bf 142}, 1 (1986), and the references therein.
1031: \bibitem{thooft}G. 'tHooft, Nucl. Phys. {\bf 79}, 276 (1974);
1032: A. M. Polyakov, JETP Lett. {\bf 20}, 194 (1974).
1033: \bibitem{fadd1} L. Faddeev and A. Niemi, Nature {\bf 387}, 58 (1997);
1034: J. Gladikowski and M. Hellmund, Phys. Rev. {\bf D56}, 5194 (1997);
1035: R. Battye and P. Sutcliffe, Phys. Rev. Lett. {\bf 81}, 4798
1036: (1998).
1037: \bibitem{prl01} Y. M. Cho, Phys. Rev. Lett. {\bf 87}, 252001 (2001).
1038: \bibitem{plb97} Y. M. Cho and D. Maison, Phys. Lett. {\bf B391},
1039: 360 (1997); W. S. Bae and Y. M. Cho, JKPS {\bf 46}, 791 (2005).
1040: \bibitem{yang} Y. Yang, {\it Solitons in Field Theory and Nonlinear
1041: Analysis} (Springer) 2001.
1042: \bibitem{bec} C. Myatt {\it at al.}, Phys. Rev. Lett. {\bf 78}, 586 (1997);
1043: D. Stamper-Kurn, {\it at al.}, Phys. Rev. Lett. {\bf 80}, 2027 (1998);
1044: J. Stenger {\it at al.}, Nature {\bf 396}, 345 (1998).
1045: \bibitem{ruo} J. Ruostekoski and J. Anglin, Phys. Rev. Lett.
1046: {\bf 86}, 3934 (2001); U. Al Khawaja and H. Stoof, Nature {\bf 411},
1047: 818 (2001); Phys. Rev. {\bf A64}, 043612 (2001);
1048: H. Stoof {\it at al.}, Phys. Rev. Lett. {\bf 87}, 120407 (2001);
1049: R. Battye, N. Cooper, and P. Sutcliffe, Phys. Rev. Lett. {\bf 88},
1050: 080401 (2002); C. Savage and J. Ruostekoski, Phys. Rev. Lett. {\bf 91},
1051: 010403 (2003).
1052: \bibitem{exp1}J. Nagamatsu et al., Nature {\bf 410}, 63 (2001);
1053: S. L. Bud'ko et al., Phys. Rev. Lett. {\bf 86}, 1877 (2001);
1054: C. U. Jung et al., Appl. Phys. Lett. {\bf 78}, 4157(2001).
1055: \bibitem{exp2}H. D. Yang et al., Phys. Rev. Lett. {\bf 87}, 167003 (2001);
1056: J. J. Tu et al., Phys. Rev. Lett. {\bf 87}, 277001 (2001).
1057: \bibitem{piet} B. Piette, B. Schroers, and W. Zakrzewski,
1058: Nucl. Phys. {\bf 439}, 205 (1995).
1059: \bibitem{plb04} Y. M. Cho, Phys. Lett. {\bf B603}, 88 (2004);
1060: hep-th/0404181.
1061: \bibitem{bec1} Y. M. Cho, cond-mat/0112325.
1062: \bibitem{bec5} Y. M. Cho, cond-mat/0409636.
1063: \bibitem{cm2} Y. M. Cho, cond-mat/0112498; cond-mat/0308182.
1064: \bibitem{plb05}Y. M. Cho, Phys. Lett. {\bf B616}, 101 (2005).
1065: \bibitem{ewknot} Y. M. Cho, hep-th/0110076.
1066: \end{thebibliography}
1067:
1068: \end{document}
1069: