cond-mat0505120/prof.tex
1: \documentclass[twocolumn,pra,showpacs,preprintnumbers]{revtex4}
2: \usepackage{graphicx}
3: \begin{document}
4: \title{Stationary Josephson effect in a weak-link between nonunitary
5: triplet superconductors}
6: \author{G. Rashedi$^{1,3}$ and Yu. A.
7: Kolesnichenko$^{2}$}
8: \address{$^1$ Institute for Advanced Studies in Basic Sciences,
9:  Zanjan, 45195-1159, Iran\\ $^2$ B.Verkin Institute for Low
10: Temperature Physics
11:  Engineering of National Academy of Sciences of Ukraine, 47,
12:   Lenin ave , 61103, Kharkov, Ukraine\\
13: $^3$ Department of Physics, Faculty of Science, University of
14: Shahrekord, Shahrekord, P.O.Box 115, Iran}
15: 
16: \date{\today}
17: 
18: \begin{abstract}
19: A stationary Josephson effect in a weak-link between misorientated
20: nonunitary triplet superconductors is investigated theoretically.
21: The non-self-consistent quasiclassical Eilenberger equation for
22: this system has been solved analytically. As an application of
23: this analytical calculation, the current-phase diagrams are
24: plotted for the junction between two nonunitary bipolar $f-$wave
25: superconducting banks. A spontaneous current parallel to the
26: interface between superconductors has been observed. Also, the
27: effect of misorientation between crystals on the Josephson and
28: spontaneous currents is studied. Such experimental investigations
29: of the current-phase diagrams can be used to test the pairing
30: symmetry in the above-mentioned superconductors.
31: \end{abstract}
32: \pacs{74.20.Rp,74.50.+r,74.70.Tx,85.25.Cp,85.25.Dq}
33: \maketitle
34: \section{Introduction}
35: In recent years, the triplet superconductivity has become one of
36: the modern subjects for researchers in the field of
37: superconductivity \cite {Ueda,Maeno,Mackenzie}. Particularly, the
38: nonunitary spin triplet state in which Cooper pairs may carry a
39: finite averaged intrinsic spin momentum has attracted much
40: attention in the last decade \cite{Tou,Machida}. A triplet state
41: in the momentum space $\mathbf{k}$ can be described by the order
42: parameter
43: ${\hat{\Delta}}(\mathbf{k})=i(\mathbf{d}(\mathbf{k})\cdot {\hat{
44: \sigma}})\hat{\sigma}_{y}$ in a 2$\times $2 matrix form in which
45: $\hat{\sigma}_{j}$ are 2$\times $2 Pauli matrices $\left(
46: {\hat{\sigma}=}\left(
47: \hat{\sigma}_{x},\hat{\sigma}_{y},\hat{\sigma}_{z}\right) \right)
48: $. The three dimensional complex vector $\mathbf{d}(\mathbf{k)}$
49: (gap vector) describes the triplet pairing state. In the
50: nonunitary state, the product
51: ${\hat{\Delta}}(\mathbf{k}){\hat{\Delta}}(\mathbf{k})^{\dagger
52: }=\mathbf{d}(\mathbf{k})\cdot \mathbf{d}^{\ast
53: }(\mathbf{k})+i(\mathbf{d}(\mathbf{k})\times \mathbf{d}^{\ast
54: }(\mathbf{k}))\cdot {\hat{\sigma}}$ is not a multiple of the unit
55: matrix. Thus in a non-unitary state the time reversal symmetry is
56: necessarily broken spontaneously and a spontaneous moment
57: $\mathbf{m}(\mathbf{k})=i\mathbf{d}(\mathbf{k})\times
58: \mathbf{d}^{\ast }(\mathbf{k})$ appears at each point $\mathbf{k}$
59: of the momentum space. In this case the macroscopically averaged
60: moment $<\mathbf{m}(\mathbf{k})>$ integrated on the Fermi surface
61: does not vanish. The value $\mathbf{m}(\mathbf{k})$ is related to
62: the net spin average by $Tr[{\hat{\Delta}}(\mathbf{k})^{\dag
63: }{\hat{\sigma}_{j}}{\hat{\Delta}}(\mathbf{k})]$. It is clear that
64: the total spin average over the Fermi surface can be nonzero. As
65: an application, the nonunitary bipolar state of $f-$wave pairing
66: symmetry has been considered for the $B-$phase of
67: superconductivity in the $UPt_{3}$ compound which has been created
68: at low temperatures $T$ and small values of the magnetic field $H$
69: \cite{Machida,Ohmi}.
70: 
71: In the present paper, the ballistic Josephson weak link via an
72: interface between two superconducting bulks with different
73: orientations of the crystallographic axes is investigated. This
74: type of weak link structure can be used for the demonstration of
75: the pairing symmetry in the superconducting phase
76: \cite{Stefanakis}. Consequently, we generalize the formalism of
77: paper \cite{Mahmoodi} for the weak link between triplet
78: superconducting bulks with a nonunitary order parameter. In the
79: paper \cite {Mahmoodi}, the Josephson effect in the point contact
80: between unitary $f-$wave triplet superconductors has been studied.
81: Also, the effect of misorientation on the charge transport has
82: been investigated and a spontaneous current tangential to the
83: interface between the $f-$wave superconductors, has been observed.
84: 
85: In this paper the nonunitary bipolar $f-$wave model of the order
86: parameter is considered. It is shown that the current-phase
87: diagrams are totally different from the current-phase diagrams of
88: the junction between the unitary triplet ( axial and planar)
89: $f-$wave superconductors \cite{Mahmoodi} . Roughly speaking, these
90: different characters can be used to distinguish between nonunitary
91: bipolar $f-$wave superconductivity and the other types of
92: superconductivity. In the weak-link structure between the
93: nonunitary $f-$wave superconductors, the spontaneous current
94: parallel to the interface has been observed as a fingerprint for
95: unconventional superconductivity and spontaneous time reversal
96: symmetry breaking. The effect of misorientation on the spontaneous
97: and Josephson currents is investigated. It is possible to find the
98: value of the phase difference in which the Josephson current is
99: zero but the spontaneous current, which is produced by the
100: interface and is tangential to the interface, is present. In some
101: configurations and at the zero phase difference, the Josephson
102: current is not generally zero but has a finite value. This finite
103: value corresponds to a spontaneous phase difference which is
104: related to the misorientation between the gap vectors$\mathbf{d}$.
105: 
106: The arrangement of the rest of this paper is as follows. In
107: Sec.\ref{section2} we describe the configuration that we have
108: investigated. For a non-self-consistent model of the order
109: parameter, the quasiclassical Eilenberger equations
110: \cite{Eilenberger} are solved and suitable Green functions have
111: been obtained analytically. In Sec.\ref{section3} the formulas
112: obtained for the Green functions have been used for the
113: calculation of the current densities at the interface. An analysis
114: of numerical results will be presented in Sec.\ref{section4}
115: together with some conclusions in Sec.\ref{section5}.
116: \section{Formalism and Basic Equations}
117: \label{section2} We consider a model of a flat interface $y=0$
118: between two misorientated nonunitary $f-$wave superconducting
119: half-spaces (Fig.1) as a ballistic Josephson junction. In the
120: quasiclassical ballistic approach, in order to calculate the
121: current, we use ``transport-like'' quasiclassical Eilenberger
122: equations \cite{Eilenberger} for the energy integrated Green
123: functions $\breve{g}\left(
124: \mathbf{\hat{v}}_{F},\mathbf{r},\varepsilon _{m}\right) $
125: \begin{equation}
126: \mathbf{v}_{F}\cdot \nabla \breve{g}+\left[ \varepsilon
127: _{m}\breve{\sigma} _{3}+i\breve{\Delta},\breve{g}\right] =0,
128: \label{Eilenberger}
129: \end{equation}
130: and the normalization condition $\breve{g}\breve{g}=\breve{1}$,
131: where $\varepsilon _{m}=\pi T(2m+1)$ are discrete Matsubara
132: energies $m=0,1,2,...,$ $T$ is the temperature, $\mathbf{v}_{F}$
133: is the Fermi velocity and
134: $\breve{\sigma}_{3}=\hat{\sigma}_{3}\otimes \hat{I}$ in which
135: $\hat{\sigma} _{j}\left( j=1,2,3\right) $ are Pauli matrices.
136: \begin{figure}[tbp]
137: \includegraphics[width=0.9\columnwidth]{fig1.eps}
138: \caption{Scheme of a flat interface between two superconducting
139: bulks which are misorientated as much as $\protect\alpha $.}
140: \label{fig1}
141: \end{figure}
142: The Matsubara propagator $\breve{g}$ can be written in the form:
143: \begin{equation}
144: \breve{g}=\left(
145: \begin{array}{cc}
146: g_{1}+\mathbf{g}_{1}\cdot \mathbf{\hat{\sigma}} & \left(
147: g_{2}+\mathbf{g} _{2}\cdot \hat{\sigma}\right) i\hat{\sigma}_{2} \\
148: i\hat{\sigma}_{2}\left( g_{3}+\mathbf{g}_{3}\cdot
149: \hat{\sigma}\right)  & g_{4}-\hat{\sigma}_{2}\mathbf{g}_{4}\cdot
150: \hat{\sigma}\hat{\sigma}_{2}
151: \end{array}
152: \right),\label{Green's function}
153: \end{equation}
154: where the matrix structure of the off-diagonal self energy
155: $\breve{\Delta}$ in the Nambu space is
156: \begin{equation}
157: \breve{\Delta}=\left(
158: \begin{array}{cc}
159: 0 & \mathbf{d}\cdot \hat{\sigma}i\hat{\sigma}_{2} \\
160: i\hat{\sigma}_{2}\mathbf{{d^{\ast }}\cdot \hat{\sigma}} & 0
161: \end{array}
162: \right).\label{order parameter}
163: \end{equation}
164: The nonunitary states, for which $\mathbf{d\times d} ^{\ast }\neq
165: 0,$ are investigated. Fundamentally, the gap vector (order
166: parameter) $\mathbf{d}$ has to be determined numerically from the
167: self-consistency equation \cite{Ueda}, while in the present paper,
168: we use a non-self-consistent model for the gap vector which is
169: much more suitable for analytical calculations \cite{Kulik}.
170: Solutions to Eq. (\ref{Eilenberger}) must satisfy the conditions
171: for the Green functions and the gap vector $\mathbf{d}$ in the
172: bulks of the superconductors far from the interface as follow:
173: \begin{equation}
174: \breve{g}=\frac{1}{\Omega _{n}}\left(
175: \begin{array}{cc}
176: \varepsilon _{m}(1-\mathbf{A}_{n}\cdot \mathbf{\hat{\sigma}}) &
177: \left[i\mathbf{d}_{n}-\mathbf{d}_{n}\times \mathbf{A}_{n}\right]
178: \cdot
179: \hat{\sigma}i\hat{\sigma}_{2} \\
180: i\hat{\sigma}_{2}\left[ i\mathbf{d}_{n}^{\ast
181: }+\mathbf{d}_{n}^{\ast }\times \mathbf{A}_{n}\right] \cdot
182: \hat{\sigma} & -\varepsilon \hat{\sigma}_{2}(1+\mathbf{A}_{n}\cdot
183: \hat{\sigma})\hat{\sigma}_{2}
184: \end{array}
185: \right)   \label{Bulk solution}
186: \end{equation}
187: where
188: \begin{equation}
189: \hspace{-0.4cm}\mathbf{A}_{n}=\frac{i\mathbf{d}_{n}\times
190: \mathbf{d}_{n}^{\ast }}{\varepsilon _{m}^{2}+\mathbf{d}_{n}\cdot
191: \mathbf{d}_{n}^{\ast }+\sqrt{(\varepsilon
192: _{m}^{2}+\mathbf{d}_{n}\cdot \mathbf{d}_{n}^{\ast
193: })^{2}+(\mathbf{d}_{n}\times \mathbf{d}_{n}^{\ast })^{2}}}
194: \end{equation}
195: and
196: \begin{equation}
197: \hspace{-0.55cm}\Omega _{n}=\sqrt{\frac{2[(\varepsilon
198: _{m}^{2}+\mathbf{d} _{n}\cdot \mathbf{d}_{n}^{\ast
199: })^{2}+(\mathbf{d}_{n}\times \mathbf{d} _{n}^{\ast
200: })^{2}]}{\varepsilon _{m}^{2}+\mathbf{d}_{n}\cdot \mathbf{d}
201: _{n}^{\ast }+\sqrt{(\varepsilon _{m}^{2}+\mathbf{d}_{n}\cdot
202: \mathbf{d} _{n}^{\ast })^{2}+(\mathbf{d}_{n}\times
203: \mathbf{d}_{n}^{\ast })^{2}}}}
204: \end{equation}
205: \begin{equation}
206: \mathbf{d}\left( \pm \infty \right) =\mathbf{d}_{2,1}\left(
207: T,\mathbf{\hat{v}}_{F}\right) \exp \left( \mp \frac{i\phi
208: }{2}\right) \label{Bulk order parameter}
209: \end{equation}
210: where $\phi $ is the external phase difference between the order
211: parameters of the bulks and $n=1,2$ label the left and right half
212: spaces respectively. It is clear that poles of the Green function
213: in the energy space are in
214: \begin{equation} \Omega _{n}=0.
215: \end{equation}
216: Consequently,
217: \begin{equation}
218: (-E^{2}+\mathbf{d}_{n}\cdot \mathbf{d}_{n}^{\ast
219: })^{2}+(\mathbf{d}_{n}\times \mathbf{d}_{n}^{\ast })^{2}=0
220: \end{equation}
221: and
222: \begin{equation}
223: E=\pm \sqrt{\mathbf{d}_{n}\cdot \mathbf{d}_{n}^{\ast }\pm
224: i\mathbf{d} _{n}\mathbf{\times d}_{n}^{\ast }}
225: \end{equation}
226: in which $E$ is the energy value of the poles. The Eq.
227: (\ref{Eilenberger}) has to be supplemented by the continuity
228: conditions at the interface between superconductors. For all
229: quasiparticle trajectories, the Green functions satisfy the
230: boundary conditions both in the right and left bulks as well as at
231: the interface. The system of equations (\ref{Eilenberger}) and the
232: self-consistency equation for the gap vector $\mathbf{d}$
233: \cite{Ueda} can be solved only numerically.
234: \begin{figure}[tbp]
235: \includegraphics[width=0.9\columnwidth]{fig2.eps}
236: \caption{Component of the current normal to the interface
237: (Josephson current) versus the phase difference $\protect\phi $
238: for the junction between nonunitary bipolar $f-$wave bulks ,
239: $T/T_{c}=0.15$, geometry (i) and different misorientations.
240: Currents are given in units of $j_{0}=\frac{\protect\pi
241: }{2}eN(0)v_{F}\Delta _{0}(0).$} \label{fig2}
242: \end{figure}
243: For unconventional superconductors such solution requires the
244: information about the interaction between the electrons in the
245: Cooper pairs and the nature of unconventional superconductivity in
246: novel compounds which in most cases are unknown. Also, it has been
247: shown that the absolute value of a self-consistent order parameter
248: is suppressed near the interface and at the distances of the order
249: of the coherence length, while its dependence on the direction in
250: the momentum space almost remains unaltered \cite{Barash}. This
251: suppression of the order parameter changes the amplitude value of
252: the current, but does not influence the current-phase dependence
253: drastically. For example, it has been verified in
254: Refs.\cite{Coury} for the junction between unconventional $d$-wave
255: superconductors, in Ref.\cite{Barash} for the case of unitary
256: ``$f$-wave'' superconductors and in Ref.\cite{Viljas} for pinholes
257: in $^{3}He,$ that there is good qualitative agreement between
258: self-consistent and non-self-consistent results for not very large
259: angles of misorientation. It has also been observed that the
260: results of the non-self-consistent model in \cite{Yip} are similar
261: to experiment \cite{Backhaus}. Consequently, despite the fact that
262: this solution cannot be applied directly to a quantitative
263: analysis of a real experiment, only a qualitative comparison of
264: calculated and experimental current-phase relations is possible.
265: In our calculations, a simple model of the constant order
266: parameter up to the interface is considered and the pair breaking
267: and scattering on the interface are ignored. We believe that under
268: these strong assumptions our results describe the real situation
269: qualitatively. In the framework of such a model, the analytical
270: expressions for the current can be obtained for a certain form of
271: the order parameter.
272: \section{Analytical results}\label{section3}
273: The solution of Eq. (\ref{Eilenberger}) allows us to calculate the
274: current densities. The expression for the current is:
275: \begin{equation}
276: \mathbf{j}\left( \mathbf{r}\right) =2i\pi eTN\left( 0\right)
277: \sum_{m}\left\langle \mathbf{v}_{F}g_{1}\left(
278: \mathbf{\hat{v}}_{F},\mathbf{r},\varepsilon _{m}\right)
279: \right\rangle
280:  \label{charge-current}
281: \end{equation}
282: where $\left\langle ...\right\rangle $ stands for averaging over
283: the directions of an electron momentum on the Fermi surface
284: ${\mathbf{\hat{v}}} _{F},$ and $N\left( 0\right) $ is the electron
285: density of states at the Fermi level of energy.
286: \begin{figure}[tbp]
287: \includegraphics[width=0.9\columnwidth]{fig3.eps}
288: \caption{Component of the current normal to the interface
289: (Josephson current) versus the phase difference $\protect\phi $
290: for the junction between nonunitary bipolar $f-$wave bulks ,
291: $T/T_{c}=0.15$, geometry (ii) and different misorientations.}
292: \label{fig3}
293: \end{figure}
294: We assume that the order parameter is constant in space and in
295: each half-space it equals its value (\ref{Bulk order parameter})
296: far from the interface in the left or right bulks. For such a
297: model, the current-phase dependence of a Josephson junction can be
298: calculated analytically. It enables us to analyze the main
299: features of current-phase dependence for any model of the
300: nonunitary order parameter. The Eilenberger equations (\ref
301: {Eilenberger}) for Green functions $\breve{g}$, which are
302: supplemented by the condition of continuity of solutions across
303: the interface, $y=0$, and the boundary conditions at the bulks,
304: are solved for a non-self-consistent model of the order parameter
305: analytically. In the ballistic case the system of equations for
306: functions $g_{i}$ and $\mathbf{g}_{i}$ can be decomposed into
307: independent blocks of equations. The set of equations which
308: enables us to find the Green function $g_{1}$ is:
309: \begin{eqnarray}
310: v_{F}\hat{\mathbf{k}}\nabla g_{1} &=&i\left( \mathbf{d}\cdot
311: \mathbf{g}_{3}-\mathbf{d}^{\ast }\cdot \mathbf{g}_{2}\right) ;
312: \label{a} \\
313: v_{F}\hat{\mathbf{k}}\nabla \mathbf{g}_{-} &=&-2\left(
314: \mathbf{d\times
315: g} _{3}+\mathbf{d}^{\ast }\mathbf{\times g}_{2}\right) ;  \label{b} \\
316: v_{F}\hat{\mathbf{k}}\nabla \mathbf{g}_{2} &=&-2\varepsilon
317: _{m}\mathbf{g}_{2}+2ig_{1}\mathbf{d}+\mathbf{d}\times
318: \mathbf{g}_{-}; \label{d} \\ v_{F}\hat{\mathbf{k}}\nabla
319: \mathbf{g}_{3} &=&2\varepsilon _{m}\mathbf{g}
320: _{3}-2ig_{1}\mathbf{d}^{\ast }+\mathbf{d}^{\ast }\times
321: \mathbf{g}_{-}; \label{c}
322: \end{eqnarray}
323: where $\mathbf{g}_{-}=\mathbf{g}_{1}-\mathbf{g}_{4}.$ The Eqs.
324: (\ref{a})-(\ref{d}) can be solved by integrating over the
325: ballistic trajectories of electrons in the\ right and left
326: half-spaces. The general solution satisfying the boundary
327: conditions (\ref{Bulk solution}) at infinity is \begin{equation}
328: g_{1}^{\left( n\right) }=\frac{\varepsilon _{m}}{\Omega
329: _{n}}+a_{n}e^{-2s\Omega _{n}t}; \label{e}
330: \end{equation}
331: \begin{equation}
332: \mathbf{g}_{-}^{\left( n\right) }=-2\frac{\varepsilon _{m}}{\Omega
333: _{n}}\mathbf{A}_{n}+\mathbf{C}_{n}e^{-2s\Omega _{n}t};  \label{f}
334: \end{equation}
335: \begin{equation}
336: \hspace{-0.3cm}\mathbf{g}_{2}^{\left( n\right)
337: }=\frac{i\mathbf{d}_{n}-\mathbf{d}_{n}\times
338: \mathbf{A}_{n}}{\Omega _{n}}-\frac{2ia_{n}\mathbf{d}
339: _{n}+\mathbf{d}_{n}\times \mathbf{C}_{n}}{2s\eta \Omega
340: _{n}-2\varepsilon _{m}}e^{-2s\Omega _{n}t};  \label{g}
341: \end{equation}
342: \begin{equation}
343: \hspace{-0.3cm}\mathbf{g}_{3}^{\left( n\right)
344: }=\frac{i\mathbf{d}_{n}^{\ast }+\mathbf{d}_{n}^{\ast }\times
345: \mathbf{A}_{n}}{\Omega _{n}}+\frac{2ia_{n}\mathbf{d}_{n}^{\ast
346: }-\mathbf{d}_{n}^{\ast }\times \mathbf{C}_{n}}{2s\eta \Omega
347: _{n}+2\varepsilon _{m}}e^{-2s\Omega _{n}t};  \label{h}
348: \end{equation}
349: where $t$ is the time of flight along the trajectory, $sgn\left(
350: t\right) =sgn\left( y\right) =s$ and $\eta =sgn\left( v_{y}\right)
351: .$ By matching the solutions (\ref{e}-\ref{h}) at the interface
352: $\left( y=0,t=0\right) $, we find constants $a_{n}$ and
353: $\mathbf{C}_{n}.$ Indices $n=1,2$ label the left and right
354: half-spaces respectively. The function $g_{1}\left( 0\right)
355: =g_{1}^{\left( 1\right) }\left( -0\right) =g_{1}^{\left( 2\right)
356: }\left( +0\right)$ which is a diagonal term of the Green matrix
357: and determines the current density at the interface, $y=0$, is as
358: follows:
359: \begin{equation}
360: g_{1}\left( 0\right) =\frac{\eta (\mathbf{d}_{2}\cdot
361: \mathbf{d}_{2}(\eta \Omega _{1}+\varepsilon
362: )^{2}-\mathbf{d}_{1}\cdot \mathbf{d}_{1}(\eta \Omega
363: _{2}-\varepsilon )^{2}+B)}{[\mathbf{d}_{2}(\eta \Omega
364: _{1}+\varepsilon )+\mathbf{d}_{1}(\eta \Omega _{2}-\varepsilon
365: )]^{2}} \label{charge-term}
366: \end{equation}
367: where $B=i\mathbf{d}_{1}\times \mathbf{d}_{2}\cdot
368: (\mathbf{A}_{1}\mathbf{+A}_{2})(\eta \Omega _{2}-\varepsilon
369: )(\eta \Omega _{1}+\varepsilon ).$ We consider a rotation
370: $\breve{R}$ only in the right superconductor (see Fig.1), i.e.,
371: $\mathbf{d}_{2}(\hat{\mathbf{k}})=\breve{R}\mathbf{d}_{1}(
372: \breve{R}^{-1}\hat{\mathbf{k}});$ $\hat{\mathbf{k}}$ is the unit
373: vector in the momentum space.
374: \begin{figure}[tbp]
375: \includegraphics[width=0.9\columnwidth]{fig4.eps}
376: \caption{The $x-$component of the current tangential to the
377: interface versus the phase difference $\protect\phi $ for the
378: junction between nonunitary bipolar $f-$wave superconducting
379: bulks, $T/T_{c}=0.15$, geometry (i) and the different
380: misorientations.} \label{fig4}
381: \end{figure}
382: The crystallographic $c$-axis in the left half-space is selected
383: parallel to the partition between the superconductors (along the
384: $z$-axis in Fig.1). To illustrate the results obtained by
385: computing the formula (\ref{charge-term}), we plot the
386: current-phase diagrams for two different geometries. These
387: geometries correspond to the different orientations of the
388: crystals in the right and left sides of the interface
389: (Fig.1):\newline (i) The basal $ab$-plane in the right side has
390: been rotated around the $c$-axis by $\alpha $;
391: $\hat{\mathbf{c}}_{1}\Vert \hat{\mathbf{c}}_{2}$. \newline (ii)
392: The $c$-axis in the right side has been rotated around the
393: $b$-axis by $\alpha $; $\hat{\mathbf{b}}_{1}\Vert
394: \hat{\mathbf{b}}_{2}$.\newline Further calculations require a
395: certain model of the gap vector (order parameter) $\mathbf{d}$.
396: \section{Analysis of numerical results}
397: \label{section4} In the present paper, the nonunitary $f-$wave gap
398: vector in the $B-$phase (low temperature $T$ and low field $H$) of
399: superconductivity in $UPt_{3}$ compound has been considered. This
400: nonunitary bipolar state which explains the weak spin-orbit
401: coupling in $UPt_{3}$ is \cite{Machida}: \begin{equation}
402: \mathbf{d}(T,\mathbf{v}_{F})=\Delta
403: _{0}(T)k_{z}(\hat{\mathbf{x}}\left( k_{x}^{2}-k_{y}^{2}\right)
404: +\hat{\mathbf{y}}2ik_{x}k_{y}). \label{Model} \end{equation} The
405: coordinate axes
406: $\hat{\mathbf{x}},\hat{\mathbf{y}},\hat{\mathbf{z}}$ are selected
407: along the crystallographic axes
408: $\hat{\mathbf{a}},\hat{\mathbf{b}},\hat{\mathbf{c}}$ in the left
409: side of Fig.\ref{fig1}. The function $\Delta _{0}=$$\Delta
410: _{0}\left( T\right) $ describes the dependence of the gap vector
411: on the temperature $T$ (our numerical calculations are done at the
412: low value of temperature $T/T_{c}=0.1$). Using this model of the
413: order parameter (\ref{Model}) and solution to the Eilenberger
414: equations (\ref {charge-term}), we have calculated the current
415: density at the interface numerically. These numerical results are
416: listed below:\newline 1) The nonunitary property of Green's matrix
417: diagonal term consists of two parts. The explicit part which is in
418: the $B$ mathematical expression in Eq.(\ref{charge-term}) and the
419: implicit part in the $\Omega _{1,2}$ and $\mathbf{d}_{1,2}$ terms.
420: These $\Omega _{1,2}$ and $\mathbf{d}_{1,2}$ terms are different
421: from their unitary counterparts. In the mathematical expression
422: for $\Omega _{1,2}$ the nonunitary mathematical terms
423: $\mathbf{A}_{1,2}$ are presented. The explicit part will be
424: present only in the presence of misorientation between gap
425: vectors,\newline $B=i\mathbf{d}_{1}\times \mathbf{d}_{2}\cdot
426: (\mathbf{A}_{1}\mathbf{+A}_{2})(\eta \Omega _{2}-\varepsilon
427: )(\eta \Omega _{1}+\varepsilon )$, but the implicit part will be
428: present always. So, in the absence of misorientation
429: $(\mathbf{d}_{1}\mathbf{\Vert d}_{2})$, although the implicit part
430: of nonunitary exists the explicit part is absent.
431: \begin{figure}[tbp]
432: \includegraphics[width=0.9\columnwidth]{fig5.eps}
433: \caption{Current tangential to the interface versus the phase
434: difference $\protect\phi $ for the junction between nonunitary
435: bipolar $f-$wave superconducting bulks, $T/T_{c}=0.15$, geometry
436: (ii) and the different misorientations ($x$component) .}
437: \label{fig5}
438: \end{figure}
439: This means that in the absence of misorientation, current-phase
440: diagrams for planar unitary and nonunitary bipolar systems are the
441: same but the maximum values are  slightly different.\newline 2) A
442: component of current parallel to the interface $j_{z}$ for
443: geometry (i) is zero similar to the unitary case \cite{Mahmoodi}
444: while the other parallel component $j_{x}$ has a finite value (see
445: Fig.\ref{fig4}). This latter case is a difference between unitary
446: and nonunitary cases. Because in the junction between unitary
447: $f-$wave superconducting bulks all parallel components of the
448: current ($j_{x}$ and $j_{z}$) for geometry (i) are absent
449: \cite{Mahmoodi}.\newline 3) In Figs.\ref{fig2},\ref{fig3}, the
450: Josephson current $j_{y}$ is plotted for a certain nonunitary
451: model of $f-$wave and different geometries. Figs.\ref
452: {fig2},\ref{fig3} are plotted for the geometry (i) and geometry
453: (ii) respectively. They are completely unusual and totally
454: different from their unitary counterparts which have been obtained
455: in \cite{Mahmoodi}.\newline 4) In Fig.\ref{fig2} for geometry (i),
456: it is observed that by increasing the misorientation, some small
457: oscillations appear in the current-phase diagrams as a result of
458: the non-unitary property of the order parameter, . Also, the
459: Josphson current at the zero external phase difference $\phi =0$
460: is not zero but has a finite value. The Josephson current will be
461: zero at the some finite values of the phase difference.\newline 5)
462: In Fig.\ref{fig3} for geometry (ii), it is observed that by
463: increasing the misorientation the new zeros in current-phase
464: diagrams appear and the maximum value of the current will be
465: changed non-monotonically.
466: \begin{figure}[tbp]
467: \includegraphics[width=0.9\columnwidth]{fig6.eps}
468: \caption{Current tangential to the interface versus the phase
469: difference $\protect\phi $ for the junction between nonunitary
470: bipolar $f-$wave superconducting bulks, $T/T_{c}=0.15$, geometry
471: (ii) and the different misorientations ($z-$component).}
472: \label{fig6}
473: \end{figure}
474: In spite of the Fig\ref{fig2} for geometry (i), the Josephson
475: currents at the phase differences $\phi =0$, $\phi =\pi ,$ and
476: $\phi =2\pi $ are zero exactly.\newline 6) The current-phase
477: diagram for geometry (i) and $x-$component (Fig.\ref {fig4}) is
478: totally unusual. By increasing the misorientation, the maximum
479: value of the current increases. The components of current parallel
480: to the interface for geometry (ii) are plotted in Fig.\ref{fig5}
481: and Fig\ref{fig6}. All the terms at the phase differences $\phi
482: =0$, $\phi =\pi,$ and $\phi =2\pi $ are zero. The maximum value of
483: the current-pase diagrams is not a monotonic function of the
484: misorientation.\newline
485: \section{Conclusions}
486: \label{section5} Thus, we have theoretically studied the
487: supercurrents in the ballistic Josephson junction in the model of
488: an ideal transparent interface between two misoriented $UPt_{3}$
489: crystals with nonunitary bipolar $f-$wave superconducting bulks
490: which are subject to a phase difference $\phi $. Our analysis has
491: shown that misorientation between the gap vectors creates a
492: current parallel to the interface and different misorientations
493: between gap vectors influence the spontaneous parallel and normal
494: Josephson currents. These have been shown for the currents in the
495: point contact between two bulks of unitary axial and planar
496: $f$-wave superconductor in \cite{Mahmoodi} separately. Also, it is
497: shown that the misorientation of the superconductors leads to a
498: spontaneous phase difference that corresponds to the zero
499: Josephson current and to the minimum of the weak link energy in
500: the presence of the finite spontaneous current. This phase
501: difference depends on the misorientation angle. The tangential
502: spontaneous current is not generally equal to zero in the absence
503: of the Josephson current. The difference between unitary planar
504: and nonunitary bipolar states can be used to distinguish between
505: them. This experiment can be used to test the pairing symmetry and
506: recognize the different phases of $UPt_{3}$.
507: \begin{thebibliography}{99}
508: \bibitem{Ueda}  M. Sigrist and K. Ueda, \emph{Rev. Mod. Phys.} \textbf{63}, 239
509: (1991).
510: \bibitem{Maeno}  Y. Maeno, H. Hashimoto, K. Yoshida, S. Nashizaki, T.
511: Fujita, J.G. Bednorz, and F. Lichenberg, \emph{Nature},
512: \textbf{372}, 532 (1994).
513: \bibitem{Mackenzie}  A.P. Mackenzie and Y. Maeno, \emph{Rev. Mod. Phys.}
514: \textbf{75} , 657 (2003).
515: \bibitem{Tou}  H. Tou, Y. Kitaoka, K. Ishida, K. Asayama, N. Kimura, Y.
516: Onuki, E. Yamamoto, Y. Haga, and K. Maezawa, \emph{Phys. Rev.
517: Lett.} \textbf{80}, 3129 (1998).
518: \bibitem{Machida}  K. Machida, T. Nishira, and T. Ohmi, \emph{J. Phys. Soc.
519: Jpn.}, \textbf{68, }3364 (1999).
520: \bibitem{Ohmi}  T. Ohmi and K. Machida,\emph{ Phys. Rev. Lett.} \textbf{71},
521: 625 (1993).
522: \bibitem{Stefanakis}  N. Stefanakis, \emph{Phys. Rev. B} \textbf{65}, 064533
523: (2002).
524: \bibitem{Mahmoodi}  R. Mahmoodi, S.N. Shevchenko and Yu.A.
525: Kolesnichenko, \emph{Fiz. Nizk. Temp.} \textbf{28}, 262(2002)
526: [\emph{Sov. J. Low Temp. Phys.} \textbf{28} , 184 (2002)].
527: \bibitem{Kulik}  I.O. Kulik and A.N. Omelyanchouk, \emph{Fiz. Nizk. Temp.}
528: \textbf{4}, 296 (1978) [\emph{Sov. J. Low Temp. Phys.},
529: \textbf{4}, 142 (1978)].
530: \bibitem{Coury}  M.H.S. Amin, M.Coury, S.N. Rashkeev, A.N. Omelyanchouk,
531: and A.M. Zagoskin, \emph{Physica B,} \textbf{318}, 162 (2002).
532: \bibitem{Eilenberger}  G. Eilenberger,\emph{ Z. Phys.,} \textbf{214}, 195
533: (1968).
534: \bibitem{Barash}  Yu. S. Barash, A.M. Bobkov, and M. Fogelstr\"{o}m,
535: \emph{Phys. Rev. B} \textbf{64}, 214503 (2001).
536: \bibitem{Viljas}  J.K. Viljas, \emph{cond-mat}/0004246.
537: \bibitem{Yip}  S.-K. Yip, \emph{Phys. Rev. Lett.} \textbf{83}, 3864 (1999).
538: \bibitem{Backhaus}  S. Backhaus, S. Pereverzev, R.W. Simmonds, A.
539: Loshak, J.C. Davis, and R.E. Packard, \emph{Nature} \textbf{392},
540: 687 (1998).
541: \end{thebibliography}
542: \end{document}
543: