0810.0717/a4.tex
1: \documentclass[prb,aps,superscriptaddress,amssymb,floatfix,showpacs,amsfonts,eqsecnum]{revtex4} 
2: \usepackage{graphicx,psfrag,amsmath,amssymb,float,subfigure}
3: %\usepackage{epstopdf}
4: \usepackage{color}
5: \input{epsf}
6: \usepackage{graphicx,natbib}
7: 
8: \newcommand{\bc}{\begin{center}}
9: \newcommand{\ec}{\end{center}}
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\bea}{\begin{eqnarray}}
13: \newcommand{\eea}{\end{eqnarray}}
14: \newcommand{\nn}{\nonumber \\}
15: 
16: \newcommand{\e}{\text{e}}
17: \newcommand{\im}{\text{i}}
18: \def\l{\left}
19: \def\r{\right}
20: \def\12{\frac{1}{2}}
21: 
22: \makeatletter
23: \usepackage{psfrag}
24: 
25: \begin{document}
26: 
27: \title{Vortex interactions in a thin platelet superconductor}
28: 
29: 
30: \author{Cedric Yen-Yu Lin} 
31: \author{Ian Affleck}
32: 
33: 
34: \affiliation{Department of Physics and Astronomy, University of British Columbia,
35: Vancouver, BC, Canada, V6T 1Z1}
36: 
37: \date{\today}
38: 
39: \begin{abstract}
40: The thermal fluctations of vortices in a superconductor  
41: can be usefully mapped onto the quantum fluctations of a collection of bosons at $T=0$ moving in
42:  2 dimensions. 
43: When the superconductor is a thin platelet with the magnetic field parallel to its surface, the interacting 
44: quantum bosons are effectively moving in 1 dimension, allowing for powerful Luttinger liquid 
45: methods to be applied. Here we consider how this 1 dimensional limit is approached, 
46: studying the interaction of vortices with the platelet surfaces 
47: and each other.
48: Using realistic parameters and vortex interactions for an underdoped YBCO platelet
49: we determine the scattering length, $a$, characterizing the low energy interaction of a vortex pair as 
50: a function of the platelet thickness.  
51: $a$ determines the Luttinger parameter, $g$, for the quantum system at low densities, $n_0$: $g\to 1-2an_0$.
52: 
53: \end{abstract}
54: 
55: \pacs{72.15.Qm, 73.21.La, 73.23.Hk}
56: 
57: \maketitle
58: \section{Introduction}
59: Thermal fluctuations of vortices, taking into account pinning by impurities and vortex-vortex interactions, is a challenging 
60: and technologically important problem in statistical physics. An elegant approach to this subject is to 
61: map each fluctating vortex line into the world line of a quantum particle in a Feynman path integral, with 
62: the magnetic field direction becoming the imaginary time direction and the bosons moving 
63: in the other two spatial directions.\cite{Nelson,Fisher} Such a mapping is especially powerful 
64: for studying columnar defects which become static point defects in the quantum model. When the 
65: field direction is tilted relative to the (parallel) pins a novel non-Hermitean 2-dimensional many-body quantum problem arises.\cite{Nelson2} 
66: If the superconductor is a thin platelet, of thickness of order the penetration depth, with the field lying in the plane, 
67: then the quantum bosons are essentially restricted to one dimension.  This allows theoretical techniques 
68: including Tomonoga Luttinger liquid (TLL) theory and Density Matrix Renormalization Group (DMRG) to be brought to 
69: bear,\cite{Hofstetter,Affleck} 
70: rendering tractible a formidable problem.  It may be feasible to realize this classical analogue of a Luttinger liquid experimentally 
71: using high-T$_c$ superconductors. A promising candidate would be a very clean highly underdoped YBCO single 
72: crystal in which the penetration depth $\lambda_c$ can be as large as 50 microns.  
73: A platelet should be cleaved with thickness in the $a$-direction of order .1  to 1 mm. and then a magnetic 
74: field should be applied in the $b$-direction.
75: 
76: It was shown in [\onlinecite{Hofstetter,Affleck}] that critical phenomena connected with rotating 
77: the field direction away from the pin direction is controlled by the dimensionless Luttinger parameter, $g$. 
78:  In the case $g>1$ columnar 
79: defects are irrelevant and have little effect on the long-distance properties of the vortices. 
80: On the other hand, for $g<1$ they are relevant and an arbitrarily weak pinning potential 
81: drastically alters the system. Thus it is of considerable interest to determine $g$ 
82: and how it depends on the parameters of the system, including the density, $n_0$. In the dilute limit 
83: $g$ approaches unity and the interacting boson system becomes equivalent to non-interacting fermions. 
84: The leading density dependent correction is\cite{Affleck}
85: \be g=1-2an_0.\label{Lutt}\ee
86: Here $a$ is the 1-dimensional scattering length. This is defined in terms of the 
87: 1-dimensional even-channel phase shift, $\delta (k)$.  The even wave functions 
88: have the asymptotic long-distance behaviour:
89: \be \psi_e(x)\to \sin [k|x|-\delta (k)],\ee
90: where $x$ is the separation of the 2 bosons. 
91: At $k\to 0$ the phase shift is linear in $k$:
92: \be \delta (k)\to ak,\label{da}\ee
93: implying:
94: \be \psi_e(x)\to \sin k(|x|-a).\ee
95: Thus the relevance or irrelevance of pinning, in the dilute limit 
96: is determined by the sign of $a$. 
97: It is important to realize that this crucial sign is {\it not} 
98: fixed by the requirement that the boson-boson interaction be repulsive. 
99: For example, an infinite hard core repulsion of range $a_0$ leads 
100: to a scattering length $a=a_0>0$. On the other hand, a 
101: repulsive $\delta$-function interaction, $v\delta (x)$, 
102: leads to a {\it negative} scattering length, $a=-1/(\mu v)$ 
103: (where $\mu$ is the reduced mass). 
104: In a confined geometry the 1 dimensional scattering length 
105: depends not only on the direct inter-particle interaction but also 
106: on the effects of the boundaries. This problem was solved by Olshanii\cite{Olshanii} for
107: the case of ultra-cold atoms in a harmonic cylindrical trap, where 
108: it was shown that the sign of $a$ can be positive or negative 
109: depending on the ratio of the (positive) 3 dimensional scattering length 
110: to the trap radius. 
111: 
112: In this paper we study the properties of two interacting vortices in a thin 
113: platelet, or equivalently of 2 interacting bosons restricted to a narrow strip. 
114: We begin with the usual modified Bessel-function interaction between vortices 
115: given by anisotropic London theory. Standard boundary conditions at 
116: the edges of the platelet imply the existence of an infinite set of 
117: image vortices for each physical vortex. The interaction of 
118: an isolated vortex with its images, and with the external magnetic 
119: field determines its wave-function, $f (y)$, in the quantum mechanical analogue, determining 
120: the probability of the vortex being at a distance $y$ from the 
121: centre of the platelet. It is 
122: peaked near the centre, $y=0$. We then consider the scattering of two physical vortices, 
123: taking into account the interactions with all image vortices.  In general 
124: the two vortices could move off centre (away from $y_1=y_2=0$) as they scatter. 
125: Thus the calculation of the 
126: effective 1D scattering length requires, in principle, solving for a 2-dimensional 
127: 2-body wave-function.
128: 
129: Fortunately, there is a very large dimensionless number that appears 
130: quite generally in the thermodynamics of vortices, and which simplifies 
131: our calculations considerably. Consider, for simplicity, a macroscopic 
132: isotropic London superconductor of penetration depth $\lambda$ and 
133: coherence length $\xi$. We approximate the Gibbs free energy 
134: for $N$ vortices as:
135: \be G\approx \int d\tau \left[\tilde \epsilon_1\sum_{i=1}^N\left({d\vec r_i\over d\tau }\right)^2
136: +{\phi_0^2\over 8\pi^2\lambda^2}\sum_{i<j}K_0(|\vec r_i(\tau )-\vec r_j(\tau )|/\lambda )\right]. 
137: \label{Gg}\ee
138: Here $\phi_0=hc/(2e)\approx 2\times 10^{-7}\hbox{G.-cm.}^2$ is the flux quantum and 
139: \be \tilde \epsilon_1\approx {\phi_0^2\over 16\pi^2\lambda^2}\ln (\lambda /\xi ),\ee
140: the tilt modulus, is simply the energy per unit length of the vortex.\cite{DeGennes} 
141:  $\tau$ is the spatial 
142: co-ordinate  along the field direction and $\vec r_i(\tau )$ describes the shape 
143: of the $i^{th}$ vortex. We have approximated the vortex-vortex interaction 
144: as only depending on the difference of the $\vec r_i$'s at the same value of $\tau$ and 
145: used the standard London model result for the interaction energy per unit length 
146: between straight parallel vortices,\cite{DeGennes} given by the modified Bessel function, $K_0$.  
147: (This needs to be cut off at short distances of order $r_{ij}\approx \xi$.)
148: We approximate the partition function by an integral over vortex paths, $\vec r_i(\tau )$, 
149: weighted by the Boltzmann factor, $\exp [-G/(k_BT)]$. 
150: By identifying $G/(k_BT)$ with $S/\hbar$ where $S$ is the classical action for 
151: $N$ interacting bosons, the classical partition function describing thermal 
152: fluctuations of vortices becomes equivalent to the Feynman path integral 
153: for interacting bosons. In this way various 
154: thermal properties of the vortex system can be conveniently obtained 
155: from the quantum system.\cite{Nelson,Fisher} 
156:  (We set $\hbar$ and $k_B=1$.) The corresponding Hamiltonian is:
157: \be H=-{1\over 2m}\sum_{i=1}^N\nabla_i^2+{\phi_0^2\over 8\pi^2\lambda^2T}
158: \sum_{i<j}K_0(|\vec r_i(\tau )-\vec r_j(\tau )|/\lambda ) \ee
159: with
160: \be m=\tilde \epsilon_1/T.\ee
161: (Strictly speaking, even after inserting appropriate factors of $\hbar$ and $k_B$, 
162: both terms in this Hamiltonian have dimensions of inverse length rather than 
163: energy. This can 
164: be traced back to the fact that in Eq. (\ref{Gg}) $\tau$ is a spatial coordinate 
165: in the classical model but is treated as an imaginary time in the analogue 
166: quantum one. This creates no problems for our analysis since physical 
167: quantities calculated using this quantum approach involve appropriate 
168: ratios of parameters with the correction dimensions, as we shall see.)
169: It is convenient to change to dimensionless length variables, letting:
170: \be \vec u_i\equiv \vec r_i/\lambda .\ee
171: We then may write the Hamiltonian in dimensionless form:
172: \be 2m\lambda^2H=-\sum_{i=1}^N\nabla_{i}^2+V_0\sum_{i<j}K_0(|\vec u_i-\vec u_j|)\ee
173: where the dimensionless parameter which measures the interaction strength is:
174: \be V_0={\tilde \epsilon_1\phi_0^2\over 4\pi^2T^2}\approx \left({\phi_0^2\over 8\pi^2T\lambda}\right)^2
175: \ln (\lambda /\xi).\ee
176: Noting that
177: \be \phi_0^2/(8\pi^2k_B)=3.9223\times 10^4\ \mu\hbox{m}\ -\ \hbox{K}\ee
178: we see that $V_0\gg 1$ for essentially 
179: any superconductor  at any $T<T_c$. This means 
180: that the  analogue quantum mechanical bosons have very strong short-range interactions when measured 
181: in dimensionless units. Variants of this large number will appear when 
182: we consider the potential energy function that holds the 
183: vortices in the middle of the slab and the interaction between vortices inside the slab. 
184: This implies that the vortices 
185: stay near the centre of the slab up to rather large slab widths 
186: justifying a 1 dimensional approximaton.  It also 
187: allows an unusual but powerful semi-classical approximation to be applied to the 1 dimensional problem
188:  yielding an explicit formula for the scattering length as a function 
189: the platelet thickness $d$ and other parameters ($\lambda_a$, $\lambda_c$, 
190: $\xi_a$, $\xi_c$ and $T$). Our conclusion is that 
191:  $a$ is positive and large for narrow platelets, increasing 
192: with $d$ and having a value $a\approx 19 \lambda_a$ for 
193: $d=10\lambda_c\approx .5 mm$.  This implies that columnar pins are relevant 
194: and also that the system rapidly leaves the dilute regime at 
195: low densities of order $1/(20\lambda_a)$, corresponding to fields, $H$ 
196: only slightly above $H_{c1}$.
197: 
198: In the next section we briefly review London theory and discuss properties of a single vortex in a thin 
199: platelet.  In Sec. III we consider 2 interacting vortices, determining 
200: the scattering length.  Sec. IV contains conclusions. 
201: 
202: \section{A single vortex in a thin platelet superconductor}
203: 
204: \begin{figure}
205: \centerline{\includegraphics[width=7.5cm,clip]{platelet.eps}}
206: \caption{A thin platelet superconductor.}
207: \label{fig:slab}
208: \end{figure}
209: 
210: We make the London approximation,\cite{DeGennes} valid when the penetration depth is 
211: much longer than the coherence length, $\lambda \gg \xi$. We label 
212: the direction perpendicular to the platelet the $y$-direction, 
213: and label the direction of the magnetic field the $z$ direction. 
214: In a YBCO crystal the most promising geometry may be choosing $y$ and $z$ 
215: to be the $a$ and $b$ directions (or vice versa). See Fig. (\ref{fig:slab}). 
216:  Thus the thin direction of the platelet is the $a$ direction, not 
217: the usual growth direction, which is $c$. Such a sample could presumably 
218: be obtained by cleaving a macroscopic sample. The magnetic field of 
219: a single vortex, centered at $\vec r=(x,y)=0$ thus obeys:
220: \be h - \left[\lambda_a^2\frac{\partial^2 h}{\partial x^2} + \lambda_c^2 \frac{\partial^2 h}{\partial y^2} \right] = \phi_0\delta^2(r).
221: \label{london}\ee
222:  The Dirac $\delta$-function at 
223: the vortex core 
224: should actually be smeared over a distance of order $\xi$, the coherence length. 
225: Note that the decay of the magnetic field in the $x=c$ direction is governed 
226: by supercurrents running in the in the $y=a$ direction and hence involves $\lambda_a$ 
227: whereas the decay in the $y=a$ direction is governed by supercurrents 
228: running in the $x=c$ direction and hence involves $\lambda_c$.  In extremely 
229: underdoped YBCO crystals typical parameter values are 
230: \bea \lambda_c&=&50 \mu \text{m}\nn
231: \lambda_a&=& .5\mu \text{m}\nn
232: \xi_a &=& 5 \text{nm} \nn
233: \xi_c&=&.05 \text{nm}\nn
234: T_c&=& 17 K.
235: \label{par}\eea
236: (The value of $\xi_c$ may be a bit small compared to existing measurements but 
237: it is convenient to assume the result which follows from anisotropic 
238: Ginsburg-Landau theory: $\lambda_c/\lambda_a=\xi_a/\xi_c$. In any 
239: event our results only depend logarithmically on the $\xi$'s.)
240: Thus the vortex is extremely elliptical: much more extended in the 
241: $y=a$ direction. To acheive the two-dimensional limit, we need the 
242: sample thickness to be of order the vortex size.  (Actually, as we 
243: shall see a thickness of up to ten times the vortex size or more is alright.) 
244: Thus we can take advantage of the larger $\lambda_c$ by cleaving 
245: our crystal in the $a$-direction. We will refer to the parameters 
246: in Eq. (\ref{par}) at temperature $T\approx T_c$, with $a=y$ 
247: the thin direction as the standard parameters. However, our 
248: results should also apply at lower temperatures.  Note that 
249: the vortex wandering that we are concerned with here
250:  occurs primarily in the $x=c$ direction. The vortices presumably 
251: feel a periodic potential with wave-length given by 
252: the lattice constant in the $c$-direction. We will ignore this 
253: here.  In the dilute limit that we are considering it is 
254: not expected to have an important effect. 
255: 
256: The solution of Eq. (\ref{london}) 
257: at distances $r\gg \xi$, ignoring for now the boundaries, can be found by using a Fourier transform and is 
258: \bea h &=& \frac{\phi_0}{4\pi^2} \int_{-\infty}^\infty \int_{-\infty}^\infty \frac{e^{i(xk_x+yk_y)}}{\lambda_a^2k_x^2 + \lambda_c^2k_y^2 + 1}dk_xdk_y \\
259: &=& \frac{\phi_0}{2\pi\lambda_a\lambda_c}K_0\left(\sqrt{\left(\frac{x}{\lambda_a}\right)^2 + \left(\frac{y}{\lambda_c}\right)^2}\right)
260: \label{h}\eea
261: where $K_0$ is a modified Bessel function.
262: 
263: The energy per unit length of a vortex is 
264: \be \epsilon = -\frac{1}{8\pi}\int h \left[\lambda_a^2\frac{\partial h}{\partial x}\hat{x} + 
265: \lambda_c^2\frac{\partial h}{\partial y}\hat{y}\right] \cdot d\sigma, \ee
266: where the integral is taken over an ellipse of radii  $\xi_a$, $\xi_c$ around the vortex core. Thus 
267: \be \epsilon \approx \frac{\phi_0^2}{16\pi^2\lambda_a\lambda_c} \ln\left(\frac{\lambda_c}{\xi_a}\right)
268: , \label{line}\ee
269: to logarithmic accuracy. 
270: For the parameters in Eq. (\ref{par})  
271: we have $\epsilon = 10^{-8} \text{erg/cm}.$
272: 
273: We remark that the interaction energy per unit length between 2 straight parallel vortices separated by a vector $\vec r$
274: is simply:
275: \be U_{12}=\frac{\phi_0h_{12}}{4\pi},\ee
276: where $h_{12}(\vec r)$ is the magnetic field at the location of one vortex produced by the other, Eq. (\ref{h}).
277: 
278: We now consider a single straight vortex in an infinite slab, of thickness $d$, extending from $-d/2$ to $d/2$.
279: The presence of boundaries of the superconductor at $y=\pm d/2$, imposes boundary conditions, 
280: \be 0=j_y= {\partial h\over \partial x}.\ee
281: A vortex at position $y$ in a slab of thickness $d$ creates image vortices with magnetic field in the opposite 
282: direction at positions $(2n+1)d - y$ for integral $n$, and creates image vortices with field in the 
283: same direction at positions $2nd + y$ for all nonzero integral $n$. Thus the magnetic field at location $(x,y_2)$ 
284: for a vortex at $(0,y_1)$ is:
285: \bea h(x;y_2,y_1)&=& \frac{\phi_0}{2\pi\lambda_a\lambda_c}\left[\sum_{n=-\infty}^\infty K_0\left(\sqrt{\left(\frac{x}{\lambda_a}\right)^2
286: +\left(\frac{2nd + y_1-y_2}{\lambda_c}\right)^2}\right)\right.\nn  & & \left.- \sum_{n=-\infty}^\infty 
287: K_0\left(\sqrt{\left(\frac{x}{\lambda_a}\right)^2+\left(\frac{(2n+1)d - y_1-y_2}{\lambda_c}\right)^2}\right)\right]. \label{hI} \eea
288: In addition to the field of the vortex and its images an additional field occurs inside the superconductor when 
289: a field $H$ is applied outside of it:
290: \be h_1(y) = H\frac{\cosh(y/\lambda_c)}{\cosh(d/2\lambda_c)}. \label{h_1}\ee
291: The magnetic field at the position of the vortex, $(0,y)$ due to all its image vortices is:
292: \be h_2(y) = \frac{\phi_0}{2\pi\lambda_a\lambda_c}\left[- \sum_{n=\pm 1, \pm 3...} K_0\left(\frac{nd - 2y}{\lambda_c}\right) +
293:  \sum_{n=\pm 2, \pm 4...} K_0\left(\frac{nd}{\lambda_c}\right)\right].\label{h2}\ee
294: The Gibbs free energy depends on the position $y$ of the vortex as:
295: \be V_1(y) = \frac{\phi_0}{4\pi}[h_1(y) + \frac{1}{2}h_2(y)]+\hbox{constant}. \label{V_1}\ee
296: 
297: This is plotted in Fig. (\ref{Vy}) and (\ref{V100} at $H=H_{c1}$.  It has a minimum at $y=0$, large barriers at intermediate $y$ 
298: and then appears to diverge to $-\infty$ at $y\to \pm d/2$:
299: \be V_1(y)\to -{\phi_0^2\over 16\pi^2\lambda_a\lambda_c}\ln [\lambda_c/(d\pm 2y)].\ee
300:  This divergence is due to 
301: the interaction of the vortex with its image at $y\mp d$. This divergence should actually 
302: be cut off at $y\mp d/2$ of order $\xi$ due to corrections to London theory. 
303: We take this into account by replacing $K_0[(nd-2y)/\lambda_c]$ by $K_0[(nd-2y)/\lambda_c]-
304: K_0[(nd-2y)/\xi ]$ in Eq. (\ref{h2}).
305: For $H>H_{c1}$,  we might expect the true minimum energy 
306: to be at $y=0$. This gives the usual formula for $H_{c1}$ for an anistropic superconductor:
307: \be H_{c1}\approx {4\pi \over \phi_0}\epsilon .\label{Hc1}\ee
308: where $\epsilon$ is the energy per unit length of a bulk vortex, Eq. (\ref{line}). 
309: 
310: \begin{figure}[htp]
311:   \begin{center}
312:     \subfigure[Gibbs potential, $d = 10\lambda_c$] {\label{Vy} \includegraphics[width = 3.7cm, clip]{potentialy10.eps}}
313:     \subfigure[Gibbs potential, $d = 100\lambda_c$] {\label{V100} \includegraphics[width = 3.7cm, clip]{potentialy100.eps}} \\
314:     \subfigure[Wave function, $d = 10\lambda_c$] {\label{psi10} \includegraphics[width = 3.7cm, clip]{wavey10_2.eps}}
315:     \subfigure[Wave function, $d = 100\lambda_c$] {\label{psi100} \includegraphics[width = 3.7cm, clip]{wavey100.eps}}
316:   \end{center}
317:   \caption{(a) and (b):The Gibbs potential,  Eqs. (\ref{h_1}) - (\ref{V_1}),
318:  for a single vortex with our standard parameters, $H=H_{c1}$ and $d = 10,\ 100\lambda_c$; 
319:  $y$ is in units of $\lambda_c$.
320: (c) and (d): The corresponding wave functions $f(y)$.}
321:   \label{fig:edge}
322: \end{figure}
323: 
324: We now turn to a study of thermal fluctuations of a single vortex inside the slab for 
325: such an anisotropic superconductor. With the magnetic field along the $b$-axis, the 
326: tilt modulus is very different for vortex tilting in the $a=y$ or $c=x$ direction. The 
327: elastic energy is written:
328: \be G_0=\int d\tau \left[{\tilde \epsilon_a\over 2}\left({dy\over d\tau }\right)^2+
329: {\tilde \epsilon_c\over 2}\left({dx\over d\tau }\right)^2\right] .\ee
330: Due to the assumed symmetry under rotations in the $a$-$b$ plane, the tilt 
331: modulus for tilting in the $a$ direction is just given by the energy per unit 
332: length:
333: \be \tilde \epsilon_a=\epsilon \ee
334: where $\epsilon$ is given in Eq. (\ref{line}). On the other hand, 
335: the energy per unit length of a vortex aligned parallel to the $c=x$ axis is much larger, 
336: resulting in the tilt modulus\cite{Ivley}
337: \be \tilde \epsilon_c={\phi_0^2\lambda_c\over 8\pi^2\lambda_a^3}\ln (\lambda_c/\xi)\approx 
338: {2\lambda_c^2\over \lambda_a^2}\tilde \epsilon_a.\label{epc}\ee
339: To this must be added the $y$-dependent free energy:
340: \be G_1=\int d\tau V_1[y(\tau )].\ee
341: Here we have again considered only ``instantaneous'' interactions between the vortex and 
342: its images, at a fixed value of $\tau$. 
343: We do not expect this approximation to qualitatively change the long distance physics in the dilute limit.  
344:   If we consider 
345: a very long vortex, in a sample of macroscopic length in the $z=\tau=b$ direction, then 
346: the probability of the displacement of the vortex from the centre of the slab having some value $y$ 
347: is simply given by $|f(y)|^2$ where $f$ is the ground state wave-function of the one-dimensional 
348: Hamiltonian:
349: \be H_1=-{1\over 2m_a}\left({d\over dy}\right)^2+{V_1(y)\over T}\ee
350: with
351: \be m_a\equiv \tilde {\epsilon_a\over T}.\label{ma}\ee
352: It is now convenient to define the dimensionless length variable:
353: \be \tilde y \equiv y/\lambda_c,\ee
354: in terms of which:
355: \be 2m_a\lambda_c^2H=-\left({d\over d\tilde y}\right)^2+V_{0y}\left[-\sum_nK_0[(2n+1)d/\lambda_c-2\tilde y]
356: +\ln (\lambda_c/\xi_a){\cosh (\tilde y/2)\over \cosh (d/2\lambda_c)}\right] \ee
357: where
358: \be V_{0y}\equiv {2\tilde \epsilon_a\phi_0^2\lambda_c\over 16\pi^2\lambda_aT^2}=
359: 2\left({\phi_0^2\over 16\pi^2\lambda_aT}\right)^2\ln \left({\lambda_c\over \xi_a}\right). \label{V0y}
360: \ee
361: [We have set $H=H_{c1}$ given in Eq. (\ref{Hc1}).]
362: Since we will choose $d$ of order $\lambda_c$, we see that the dimensionless number $V_{0y}$ 
363: characterizes the height of the barriers holding the vortex at the center of the platelet. 
364: Using the numbers in Eq. (\ref{par})  and choosing $T=T_c$ we find:
365: \be V_{0y}\approx 10^8.\ee
366: We solve this Schroedinger equation numerically for the groundstate, for $d=10\lambda_c$, 
367:  Fig. [\ref{psi10}], finding that the particle makes only 
368: very small quantum fluctuations away from $y=0$ due to the huge barriers. 
369: For the above parameters we find 
370: \be \sqrt{<y^2>}=.01419\lambda_c=.001419 d.\ee
371: We should estimate the true critical field, $H_{c1}(d)$ 
372: using the ground state energy  of the quantum Hamiltonian. If the 
373: field is too low the particle can tunnel through the barrier corresponding 
374: to a vortex terminating at some value of $\tau$ with the magnetic flux 
375: leaving the superconductor. However, due to the exceedingly high barrier, 
376: the tunnelling probability is miniscule and the system will not 
377: be very sensitive to the precise value of $H$. 
378: We should also remark that, to solve the Schroedinger equation precisely we need 
379: to specifiy some boundary conditions on the wave-function at $y=\pm d/2$. 
380: We imposed vanishing boundary conditions. 
381: Fortunately, the extremely large dimensionless barrier also renders our 
382: results very insensitive to this choice. 
383: 
384: \section{two vortices}
385: Consider two straight parallel vortices inside the platelet, at locations $(x_i,y_i)$.  
386: The Gibbs free energy per unit length is:
387: \be V(x_1-x_2;y_1,y_2)={\phi_0\over 4\pi}h(x_1-x_2;y_1,y_2)+V_1(y_1)+V_2(y_2).\label{Vtot}\ee
388: By translational invariance in the $x$-direction, $V$ depends only on 
389: the difference of the $x$-coordinates of the two vortices:
390: \be x\equiv x_1-x_2.\ee
391: Here $h(x,y_1,y_2)$ is given in Eq. (\ref{hI}) and $V_1(y)$ by Eq. (\ref{V_1}). 
392: The first term in Eq. (\ref{Vtot}) represents the interaction of one 
393: vortex with the other one and with the images of the other one.  The second 
394: and third terms represent the interaction of each vortex with its own images 
395: and with the screened external field. 
396: $V(x,y_1,y_2)$ has a deep minimum at $y_1=y_2=0$ for large $|x|\gg \lambda_a$. $V(x,0,0)$ 
397: has a large peak centred at $x=0$ with a weak, logarithmic divergence right at $x=0$. 
398: This logarithmic divergence should be cut off at scales of order $\xi$; however 
399: this cut off has essentially no effect on the scattering length, as we shall see.   
400: 
401: Again we may study the thermodynamics of two wiggling vortices in the platelet by mapping 
402: onto a quantum mechanics model.  We make the fundamental assumption 
403: that the vortices only bend on long length scales (compared to $\lambda_a$)
404: and that we may approximate the vortex-vortex (and vortex-image vortex) interaction 
405: by an ``instantaneous'' one at a fixed value of $\tau =z$. 
406: The Boltzmann sum is now 
407: over the configuration of two vortices and we must include the vortex-vortex 
408: interaction in the free energy. Again identifying the free energy 
409: with the imaginary time action, we see that the corresponding quantum Hamiltonian is:
410: \be H = -{1\over 2}\sum_{i=1}^2\left[{1\over m_a}\left({d\over dy_i}\right)^2+
411: {1\over m_c}\left({d\over dx_i}\right)^2\right] +{V(x,y_1,y_2)\over T}.\label{SE2D}\ee
412: Here $m_a$ is given in Eq. (\ref{ma}) and
413: \be m_c\equiv {\tilde \epsilon_c\over T}\ee
414: where $\tilde \epsilon_c$ is given in Eq. (\ref{epc}). 
415: In the quantum analogue, the vortices obey Bose statistics\cite{Nelson} and consequently 
416: the two-body wave-function must be symmetric: even under $x\to -x$. The asymptotic 
417: behavior of the low energy wave-functions at $|x|\gg \lambda_a$ is given by:
418: \be \psi (x,y_1,y_2)\to f(y_1)f(y_2)\sin [k|x|-\delta (k)]\ee
419: where $f(y)$ is the ground state wave-function for a single vortex, discussed in Sec. II. 
420: The energy of this scattering state is:
421: \be E=2E_1+{k^2\over 2\mu}\ee
422: where $E_1$ is the ground state energy for a single vortex, discussed in the previous section 
423: and the reduced mass which governs the relative motion is
424: \be \mu = m_c/2.\ee
425: The scattering length is defined by Eq. (\ref{da}). 
426: 
427: It turns out that, due to the large barrier near $x=0$, for small $y_i$, the scattering length is 
428: determined almost completely by the large $x$ asymptotic form of the potential, until $d$ gets very large compared to $\lambda_c$. 
429: The $x$-dependent part of the exact potential can be written as a sum of exponentials:
430: \be V(x;y_1,y_2) = \frac{\phi_0^2}{4\pi\lambda_a^2d}\sum_{n=1}^{\infty} f\left( n,\frac{y_1}{d},\frac{y_2}{d}\right) 
431: \frac{e^{-|x|\sqrt{1+(n\pi\lambda_c/d)^2}/\lambda_a}}{\sqrt{1+(n\pi\lambda_c/d)^2}/\lambda_a} +V_1(y_1)+V_1(y_2).\ee
432: where we define
433: \be f(n,y_1/d,y_2/d) \equiv \left\{\begin{array}{rl} \cos(n\pi y_1/d)\cos(n\pi y_2/d) & 
434: \text{if $n$ is odd} \\ \sin(n\pi y_1/d)\sin(n\pi y_2/d) & \text{if $n$ is even} \end{array}. \right. \ee
435: At large $|x|$ we may approximate the sum by the first term only:
436: \be V \approx \frac{\phi_0^2\lambda }{4\pi\lambda_a^2d}e^{-|x|/\lambda}\cos(\frac{\pi y_1}{d})\cos(\frac{\pi y_2}{d})
437: +V_1(y_1)+V_1(y_2) \label{Vlx}\ee
438: where we have defined, for convenience, a reduced value of $\lambda_a$:
439: \be \lambda \equiv \lambda_a\left[1 + \left(\frac{\pi\lambda_c}{d}\right)^2\right]^{-1/2}. \label{lambda}\ee
440: For this approximation to hold, we need that the first term dominates all other terms. In the one-dimensional limit (i.e. $y_1=y_2=0$) the condition for large $x$ is
441: \be \exp\left(\frac{|x|}{\lambda_a}\frac{4\pi^2\lambda_c^2}{d^2}\right) \gg 1. \label{com} \ee
442: Note that, unlike the direct vortex-vortex interaction, this potential has a simple exponential 
443: dependence on $x$ at large $|x|$ albeit with a reduced penetration depth. 
444: 
445: Due to the large barriers in $V_1(y)$ we expect the low energy scattering states 
446: to be confined to $y_i\approx 0$. 
447: We first calculate $a$ assuming that the vortices stay exactly in the middle of the 
448: slab, $y_i=0$ throughout the scattering process. We return to a further discussion 
449: of why this is reasonable at the end of this section. 
450: This reduces the problem to a one-dimensional quantum mechanics model with Hamiltonian:
451: \be H=-{1\over 2\mu}{d^2\over dx^2}+{V(x)\over T}\ee
452: where
453: \be V(x)=V(x;0,0)\to \frac{\phi_0^2\lambda}{4\pi\lambda_a^2d}e^{-|x|/\lambda}.\ee
454: We look for parity even solutions of this Schroedinger equation with asymptotic 
455: behavior $\psi (x)\to \sin [k|x|-\delta (k)]$ with $\delta (k)\to ak$ as $k\to 0$. 
456: Note that in the small $k$ limit, $\psi (x)\to \sin [k(|x|-a)]$ for 
457: $x\gg \lambda$. Then, if we consider an intermediate range of $x$:
458: \be \lambda \ll |x|\ll 1/k,\ee
459: we may approximate the wave-function by a linear form:
460: \be \psi (x)\propto |x|-a.\label{x-a}\ee
461: Thus, to find the scattering length we need to simply solve the zero energy Schroedinger equation:
462: \be \left[ -{1\over 2\mu}{d^2\over dx^2}+{V(x)\over T}\right]\psi =0.\label{0}\ee
463: Note that this reduces the eigenvalue problem to 
464: a simple initial value problem.  We simply impose the initial conditions:
465: \bea {d\psi\over dx}(0)&=&0\nn
466: \psi (0)&=&1\eea
467: and solve the zero energy equation.
468: The asymptotic behavior of the solution at $|x|\gg \lambda$ is given by Eq. (\ref{x-a}) 
469: which determines the scattering length, $a$.
470: 
471: It is convenient to introduce a dimensionless length variable:
472: \be \tilde x\equiv x/\lambda \ee
473: in terms of which the Schroedinger equation becomes, at large $|\tilde x|$:
474: \be \left[ -{d^2\over d\tilde x^2}+V_{0x}e^{-\tilde x}\right]\psi (\tilde x)=2\mu \lambda^2E\psi (\tilde x).\label{Su}\ee
475: Here the dimensionless number which measures the strength of the repulsive potential is:
476: \be V_{0x}={\tilde \epsilon_c \phi_0^2\lambda^3\over 4\pi \lambda_a^2dT^2}=
477: {1\over 2\pi}\left({\phi_0^2\over 4\pi\lambda_aT}\right)^2
478: \ln (\lambda_c/\xi_a){\lambda_c/d\over [1+(\pi \lambda_c/d)^2]^{3/2}}. \label{V0x}\ee
479: Using our estimates of the parameters in Eq. (\ref{par}) with $T=T_c$ and $d=10\lambda_c$ we find:
480: \be V_{0x}=1.07 \times 10^8.\ee
481: Note that $V_{0x}\propto 1/d$ at $d\gg \lambda_c$. Importantly $V_{0x}\gg 1$ when 
482: $d$ is of order $\lambda_c$ and remains large out to extremely large 
483: values of $d/\lambda_c$. The largeness of $V_{0x}$ leads to a large scattering length, allows 
484: for an unusual semi-classical solution approximation and also helps to justify setting $y_i=0$ as we shall see below. 
485: It is of course, possible to solve the Schroedinger equation numerically 
486: for specified values of the parameters.  However, the largeness of $V_{0x}$ and 
487: $V_{0y}$ creates numerical difficulties for standard algorithms, when 
488: one attempts to solve the full 3-dimensional problem, including the $y_i$. 
489: In this case it is much easier, and more transparent, to use the semi-classical approximation. 
490: 
491: Our semi-classical approximation for the scattering length at $V_{0x}\gg 1$ begins with the observation that the classical 
492: turning point for 2 particles approaching each other with a small relative momentum, $k$ 
493: occurs at $\tilde x\gg 1$. Therefore $a$ is determined almost completely by the large $\tilde x$ 
494: form of the potential in Eq. (\ref{Vlx}).
495: Using the large $\tilde x$ form of the potential, we may 
496: solve the one-dimensional Schroedinger equation exactly. To do this we change variables to:
497: \be u = 2\sqrt{V_{0x}}e^{-|\tilde x|/2}.\ee
498: The zero energy Schroedinger equation, (\ref{0}), simplifies to
499: \be u^2\psi'' + u\psi' - u^2\psi = 0 \ee
500: where the primes denote differentiation with respect to $u$. This is the zeroth order modified Bessel's differential equation, and 
501: the general solution is given in terms of the modified Bessel functions:
502: \bea \psi &=& c_1 I_0(u) + c_2 K_0(u)  \nn
503: &=& c_1 I_0(2\sqrt{V_{0x}}e^{-|x|/2\lambda}) + c_2K_0(2\sqrt{V_{0x}}e^{-|x|/2\lambda}). \eea
504: For large $u$ (i.e. $\tilde x << \ln V_0$) we have
505: \be \psi \approx \frac{1}{\sqrt{2\pi u}} \left(c_1 e^u + c_2 \pi e^{-u} \right), \ee
506: or writing in terms of the original variables and putting back factors of $\lambda$, we have
507: \be \psi \approx \frac{1}{\sqrt{4\pi}V_{0x}^{1/4}e^{-|x|/4\lambda}} 
508: \left(c_1 e^{2\sqrt{V_{0x}}e^{-|x|/2\lambda}} + c_2 \pi e^{-2\sqrt{V_{0x}}e^{-|x|/2\lambda}} \right) \label{large} \ee
509: On the other hand, for small $u$ (i.e. $\tilde x >> \ln V_0$) we have 
510: \be \psi \approx c_1 - c_2 (\ln(u/2) + \gamma) = c_1 + c_2\left(\frac{|\tilde x|}{2} - \frac{1}{2}\ln V_{0x} - \gamma\right) \ee
511: where $\gamma \approx 0.5772$ is the Euler-Mascheroni constant. Note that the last formula can be written as
512: \be \psi \approx \frac{c_2}{2\lambda} (|x| - a) \ee
513: where the scattering length $a$ is (putting back factors of $\lambda$)
514: \be a = \lambda (\ln V_{0x} + 2\gamma - 2c_1/c_2). \label{scatteringlength} \ee
515: To determine $c_1/c_2$ we match our solution in the region $1\ll |\tilde x|\ll \ln V_{0x}$ to the WKB solution which 
516:  works for $|x|$ not too large, in the region where $V(x)$ is large. 
517: The even WKB wave function for $E = 0$ is given by \be \psi(x) = A \left\{ \exp\left[\int_0^{x} \sqrt{V'(x')} dx'
518: \right] + 
519: \exp\left[-\int_0^{x} \sqrt{V'(x')} dx'\right] \right\}\ee
520: where
521: \be V'(x)=2\mu V(x;0,0)\ee
522: and $V$ is the exact potential of Eq. (\ref{Vtot}). (Note that we {\it do not} make any large $x$ approximation 
523: to $V$ now.) We can rewrite this as
524: \be \psi(x) = A \left\{ e^{-\alpha} \exp\left[\int_x^\infty \sqrt{V'(x')} dx'\right] + 
525: e^\alpha \exp\left[-\int_x^\infty \sqrt{V'(x')} dx'\right] \right\} \ee
526: where 
527: \be \alpha = \int_0^\infty \sqrt{V'(x)} dx. \ee
528: Now, if $x$ is large enough for our asymptotic expression $V'(x) \approx V_{0x}e^{-x/\lambda}/\lambda^2$ 
529:  to hold, 
530: then the integral can be done quite readily:
531: \be \psi(x) = A \left\{ e^{-\alpha} \exp\left[2\sqrt{V_{0x}}e^{-x/2\lambda}\right] + e^\alpha \exp\left[-2\sqrt{V_{0x}}
532: e^{-x/2\lambda} \right] \right\}. \ee
533: Comparison with Eq. (\ref{large}) thus gives
534: \be c_1/c_2 = \pi e^{-2\alpha} = \pi \exp\left[-\int_{-\infty}^\infty \sqrt{V'(x)}dx\right].
535: \label{c1c2} \ee
536: This quantity is exponentially small in the large quantity $V_{0x}$ so it is completely negligible. 
537: Note also that the logarithmic divergence of $V(x;0,0)$ at $x\to 0$ has no important effects, 
538: leaving the integral finite in Eq. (\ref{c1c2}). 
539:  The last term in Eq. (\ref{scatteringlength})essentially  vanishes, and we simply have that
540: \be a = \lambda (\ln V_{0x} + 2\gamma). \label{a}\ee
541: Asymptotically, the scattering length is linearly dependent on the logarithm of the size of the potential.
542: 
543: Interestingly, we get almost the same result for the odd wave functions, except that the sign of $c_1/c_2$ is reversed. 
544: The even channel and odd channel scattering lengths are therefore almost exactly the same. However, it is the \emph{difference}
545:  between the even and odd channel scattering lengths that determines the transmission coefficient, and it is only then 
546: that $c_1/c_2$ plays an important role.
547: 
548: We have based this approximation on the assumption that there exists a region of separation $x$, such that the approximation
549:  Eq. (\ref{large}) holds, and the WKB approximation to the wave function also holds. Typically the matching is done around 
550: the region $x \approx a$; 
551: therefore we need that (using Eq. (\ref{a}) and Eq. (\ref{com}), and noticing 
552: $\lambda \approx \lambda_a$ for large $d$)
553: \be \exp\left[\left(\frac{2\pi\lambda_c}{d}\right)^2\ln (e^{2\gamma} V_{0x})\right] \gg 1 \label{wkbcond}\ee
554: This is the condition that must be satisfied for the formula Eq. (\ref{a}) to hold.
555: \begin{figure}
556: \centerline{\includegraphics[width=7.5cm,clip]{wavex10.eps}}
557: \caption{The  wave functions in the $k\to 0$ limit comparing a precise numerical 
558: calculation (in the 1 dimensional approximation) to the semi-classical approximation 
559:  for our standard parameters and $d=10\lambda_c$. Lengths are in units of $\lambda_a$.}
560: \label{1Dwave}
561: \end{figure}
562: 
563: 
564: The wave-function calculated numerically to high precision (in the 1-dimensional approximation) 
565: and the semi-classical wave function are compared in Fig. (\ref{1Dwave})
566: for our standard parameters and $d=10\lambda_a$.As can be seen, the two wave functions give good agreement in the large-$x$ regime (with an error $< 1\%$ for $x > 18\lambda_a$). 
567: The semi-classical wave function is grossly inaccurate in the small-$x$ region ($x < 15\lambda_a$) but 
568: the wave-function is neglegible there anyway. 
569: The predicted semi-classical scattering length is $a = \lambda (\ln V_0 + 2\gamma) = 18.7398\lambda_a$. 
570: The actual scattering length of the numerically determined wave function (obtained by fitting the wave function in the large-$x$ regime to 
571: a linear function) is $18.7409\lambda_a$. The semi-classical wave function gives an error of less than $0.01\%$. 
572: In Fig. (\ref{fig:a}) we show the scattering length versus $d$, comparing our numerical results  to the semi-classical approximation
573:  (in both cases making the 1-dimensional approximation). 
574:  Most of the $d$-dependence in our semi-classical 
575: formula, Eq. (\ref{a}), arises from the $d$-dependence of the reduced penetration depth, $\lambda$, given 
576: in Eq. (\ref{lambda}). As the sample thickness decreases, the effective range of the interaction potential, $\lambda$,  
577: also decreases,  and the scattering length simply scales with it, up to logarithmic corrections coming 
578: from $V_{0x}$, defined in Eq. (\ref{V0x}). 
579: The semi-classical and numerical
580:  values agree within $1\%$ up to about $d = 22 \lambda_c$. For $d = 22 \lambda_c$, we have that (refer to Condition (\ref{wkbcond}))
581: \be \left(\frac{2\pi\lambda_c}{d}\right)^2\ln (e^{2\gamma} V_{0x}) = 1.56 \ee
582: which is already not far from unity. There will therefore be significant deviations of the true value from Eq. (\ref{a}).
583:  We could also notice that the scattering length tends towards a finite value for large $d$. This is because as the thickness of the
584:  sample grows, the image vortices move farther away from the orginal vortices until their effects become negligible.
585: 
586: \begin{figure}
587: \centerline{\includegraphics[width=7.5cm,clip]{scattering_length.eps}}
588: \caption{The scattering lengths based on a precise numerical 
589: calculation (in the 1 dimensional approximation) compared to our 
590: semi-classical approximation for our standard parameters.
591:  $d$ is in units of $\lambda_c$, the scattering length in units of $\lambda_a$.}
592: \label{fig:a}
593: \end{figure}
594: 
595: 
596: Next, we discuss the validity of our 1-dimensional approximation, setting $y_i=0$, which is 
597: justified by the fact that the single vortex wave-function, $f(y)$ is so sharply peaked near $y=0$. 
598: If we look at the shape of the potentials, $V_1(y)$, 
599: we see that for small $d$ (Fig. (\ref{Vy})) the potential has an obvious minimum in the center, and is approximately
600:  simple harmonic near the center. For larger $d$ (Fig. (\ref{V100})), however, the potential is almost negligible 
601: except for a large potential barrier close to (but not at) the edges; the potential will be qualitatively more 
602: similar to a square well. Therefore as $d$ increases, we would expect the shape of the wave function to morph 
603: from a confined Gaussian to a spread-out sinusoidal (Fig. (\ref{psi10}), Fig. (\ref{psi100})).
604: The spread of the single-vortex wave function ($\langle y^2 \rangle ^{1/2}$) is plotted in Fig. (\ref{width}). 
605: The thickness reaches $1\%$ of the platelet thickness at around $d = 35\lambda_c$, an indication that our one-dimensional 
606: approximation fails above this value. The fact that the thickness becomes linearly dependent on $d$ at 
607: large $d$ also suggests that the wave function tends to a fixed shape (a sinusoidal).
608: 
609:  A more systematic 
610: approximation to solving the full 2-body 2-dimensional Schroedinger equation of Eq. (\ref{SE2D}), would be to write:
611: \be \psi (x;y_1,y_2)\approx \psi_1(x)f(y_1)f(y_2)\ee
612: where $\psi_1(x)$ is the 1-dimensional wave-function found above 
613: and $f(y_1)$ is the single vortex wave-function. We could then improve our estimate 
614: of the 1-dimensional effective potential by using:
615: \be V(x)\approx \int dy_1dy_2|f(y_1)|^2|f(y_2)|^2V(x;y_1,y_2)\ee
616: rather than simply $V(x)\approx V(x;0,0)$. However, because 
617: $f(y)$ is so sharply peaked at $y\approx 0$ this makes a negligible difference.
618: 
619: Note that our calculation of $a$ depended essentially only on $V(x)$ in the 
620: large $x$ region, $x\gg \lambda$. Our consideration of the small $x$ region 
621: only served to determined $c_1/c_2$ which was exponentially small anyway and can 
622: simply be ignored. In this large $x$ region, Eq. (\ref{Vlx}) is 
623: a good approximation, the wave-function approximately factorizes and the large barriers in $V_1(y)$ ensure 
624: that the wave-function is strongly peaked near $y_i=0$. For smaller 
625: values of $x$ the wave-function presumably spreads out more in the 
626: $y$ direction. However, at smaller $x$ the wave-function is exponentially small 
627: anyway.
628: 
629: Finally, we consider the case where the thin direction of the 
630: YBCO platelet is the $c$ direction: y=c (and x=a, z=b). In this case the roles of $\lambda_a$ 
631: and $\lambda_c$ are switched, as are the roles of $\tilde \epsilon_a$ 
632: and $\tilde \epsilon_c$ and we get 
633: \bea V_{0y} &=&2{\tilde \epsilon_c\phi_0^2\lambda_a\over 16\pi^2T^2\lambda_c}=
634: 4\left({\phi_0^2\over 16\pi^2\lambda_aT}\right)^2\ln (\lambda_c/\xi_a) \nn
635:  V_{0x}&=&{\tilde \epsilon_a \phi_0^2\lambda^3\over 4\pi \lambda_c^2dT^2}=
636: {1\over 4\pi}\left({\phi_0^2\over 4\pi\lambda_aT}\right)^2
637: \ln (\lambda_c/\xi_a){\lambda_a/d\over [1+(\pi \lambda_a/d)^2]^{3/2}} \eea
638: where we now define:
639: \be \lambda \equiv\lambda_c\left[1+\left({\pi \lambda_a\over d}\right)^2\right]^{-1/2}.\ee
640: Apart from some unimportant factors of $2$, the formulas for $V_{0y}$ and $V_{0x}$ are the same as 
641: for the other geometry except that $d$ now appears in the dimensionless ratio 
642: $\lambda_a/d$ rather than $\lambda_c/d$. So we now conclude that 
643: the 1 dimensional approximation holds for $d\leq 20\lambda_a\approx 10 \mu$m. 
644: and the semi-classical approximation holds out to roughly the same value of $d$. 
645: 
646: \begin{figure}
647: \centerline{\includegraphics[width=7.5cm,clip]{width.eps}}
648: \caption{The spread of the single-vortex wave function, $\sqrt{<y^2>}$, plotted against $d$ using standard parameters.
649:  Both lengths are in units of $\lambda_c$.}
650: \label{width}
651: \end{figure}
652: 
653: \section{conclusions}
654: Our main result is the formula Eq. (\ref{a}) for the scattering length.  This is 
655: plotted versus the platelet thickness, $d$, for our standard parmeters, 
656: in Fig. (\ref{fig:a}). Note that $a$ is everywhere positive and $a/\lambda_a$ is 
657: everywhere quite large, having the value $a/\lambda_a\approx 19$ at $d\approx 10\lambda_c$.
658: At somewhat larger values of $d$ we expect our semi-classical approximation to 
659: the one-dimensional problem to break down and, more problematically, the 
660: one-dimensional approximation itself to start to fail. 
661: 
662: The main use of our formula for $a$ is not, of course, for studying the 
663: system with only 2 vortices, but rather for studying the thermodynamic 
664: limit of many vortices. In the dilute limit, $n_0a\ll 1$ (where $n_0$ is 
665: the vortex density per unit length),  $a$ determines the 
666: Luttinger parameter via Eq. (\ref{Lutt}). Of course the Luttinger liquid 
667: treatment of the problem assumes that it is fundamentally 1-dimensional. 
668: Our calculations here indicate that the 1-dimensional approximation should 
669: be good, at least in the 
670: dilute limit $n_0a\ll 1$, up to platelet thicknesses of order $d=10\lambda_c$ or more, 
671:  since the vortices stay very close to 
672: the centre of the platelet. Furthermore, we have determined the Luttinger 
673: parameter for this range of thicknesses and vortex densities. 
674: When $a$ is large the Luttinger parameter decreases rapidly for 
675: increasing vortex density. It was argued in (\onlinecite{Affleck}) that, 
676: at high densites, $g\ll 1$. Taken together, these results suggest a rapid monotonic drop 
677: of $g$ from $1$ with increasing density. In this case, columnar pins 
678: would be highly relevant for essentially all fields above $H_{c1}$. 
679:   Thus 
680: a promising region to look at experimentally might be very close to $H_{c1}$ with 
681: low vortex densities, $n_0\ll 1/\lambda_a$ and samples of 
682: thickness around $10\lambda_c$.
683: \acknowledgements 
684: We  would like to thank Doug Bonn, 
685: David Broun, David Nelson and Eran Sela for many helpful discussions. 
686: IA thanks Matt Choptuik and Brian Martin  for their collaboration in 
687: an earlier attack on this problem. This research is supported 
688: in part by NSERC (CL and IA) and CIfAR (IA). 
689: 
690: 
691: \begin{thebibliography}{10}
692: \bibitem{Nelson} D. R. Nelson, Phys. Rev. Lett. {\bf 60}, 1973 (1988).
693: \bibitem{Fisher} M.P.A. Fisher and D.H. Lee, Phys. Rev. {\bf B39}, 2756 (1989).
694: \bibitem{Nelson2}N. Hatano and D.R. Nelson, Phs. Rev. Lett. {\bf 77}, 570 (1996).
695: \bibitem{Hofstetter}W. Hofstetter, I. Affleck, D.R. Nelson and U. Schollw\"ock, Europhys. Lett. {\bf 66} 178 (2004).
696: \bibitem{Affleck} I. Affleck, W. Hofstetter, D.R. Nelson and U. Schollw\"ock, J Stat P10003 (2004).
697: \bibitem{Olshanii} M. Olshanii, Phys. Rev. Lett. {\bf 81}, 938 (1998).
698: \bibitem{DeGennes} For a review of London theory see P. G. De Gennes, {\it Superconductivity of 
699: Metals and Alloys} [Perseus Books, Reading, MA (1966)].
700: \bibitem{Ivley} B.I. Ivley, Yu.N. Ovchinnikov and R.S. Thompson, Phys. Rev. {\bf B44}, 7023 (1991).
701: \end{thebibliography}
702: \end{document}
703: