cond-mat0405290/prb.tex
1: %%%%%  Version 15.04.04 %%%%%
2: %\documentclass[twocolumn,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
3: %\documentclass[preprint,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
4: 
5: %\newcommand{\agt}{\mbox{\raisebox{-4pt}{$\,\buildrel>\over\sim\,$}}}
6: %\newcommand{\alt}{\mbox{\raisebox{-4pt}{$\,\buildrel<\over\sim\,$}}}
7: 
8: % Some other (several out of many) possibilities
9: %\documentclass[preprint,aps]{revtex4}
10: 
11: \documentclass[twocolumn,showpacs,preprintnumbers,
12: floatfix,amsmath,amssymb,eqsecnum]{revtex4}% Physical Review B
13: 
14: \usepackage{epsfig,psfrag}
15: %\usepackage{graphicx}   % Include figure files
16: \usepackage{dcolumn}     % Align table columns on decimal point
17: \usepackage{bm}          % bold math
18: 
19: \begin{document}
20: 
21: %\preprint{preprint not for distribution}
22: 
23: \title{Effective low-energy theory of
24: superconductivity in carbon nanotube ropes}
25: \author{A.~De~Martino and R.~Egger}
26: 
27: \affiliation{Institut f\"ur Theoretische Physik, 
28: Heinrich-Heine-Universit\"at,
29:  D-40225 D\"usseldorf, Germany}
30: 
31: \date{\today}
32: 
33: \begin{abstract}
34: We derive and analyze the low-energy theory of superconductivity in
35: carbon nanotube ropes.  A rope is modelled as an
36: array of metallic nanotubes, taking into account 
37: phonon-mediated as well as Coulomb interactions, and 
38: arbitrary Cooper pair hopping amplitudes
39: (Josephson couplings) between different tubes.
40: We use a systematic cumulant expansion to construct the
41: Ginzburg-Landau action including quantum fluctuations. 
42: The regime of validity is carefully established, and the 
43: effect of phase slips is assessed.
44: Quantum phase slips are shown to cause a depression of the critical 
45: temperature $T_c$ below the mean-field value, and a 
46: temperature-dependent resistance below $T_c$.  
47: We compare our theoretical results to recent experimental data
48: of Kasumov {\sl et al.} [Phys. Rev. B {\bf 68}, 214521 (2003)] for
49: the sub-$T_c$ resistance,
50: and find good agreement with only one free fit parameter.
51: Ropes of nanotubes therefore represent superconductors in the one-dimensional
52: few-channel limit.
53: \end{abstract}
54: 
55: \pacs{73.63.Fg, 74.78.Na, 74.25.Fy}
56: 
57: \maketitle
58: 
59: 
60: \section{Introduction}
61: 
62: Over the past decade, the unique mechanical, electrical,
63: and optical properties of carbon nanotubes, 
64: including the potential for useful technological applications,
65: have created a lot of excitement \cite{nts,nts2}.
66: While many of these properties are well understood
67: by now, the experimental observation of intrinsic 
68: \cite{kociak,kasnew,tang} and anomalously strong proximity-induced 
69: \cite{morpurgo,kasumov} superconductivity continues to pose open
70: questions to theoretical understanding. 
71: In this paper we present a theory of  
72: one-dimensional (1D) superconductivity as found in ropes of
73: carbon nanotubes \cite{kociak,kasnew} and potentially in other
74: nanowires.  Ropes are 1D materials in the sense that there is
75: only a relatively small number of
76: propagating channels (typically, $N\approx 10$ to $100$) 
77: available to electronic transport.
78: While most other 1D materials tend to become insulating at low
79: temperatures due to the Peierls transition or as a consequence
80: of electron-electron interactions, nanotubes can stay metallic
81: down to very low temperatures \cite{nts}.  
82: If the repulsive electron-electron interactions can be 
83: overcome by attractive phonon-mediated interactions, 
84: ropes of nanotubes can then exhibit a superconducting transition.
85: 
86: However, due to strong 1D fluctuations, this transition is 
87: presumably rather broad, and the question of how precisely
88: superconductivity breaks down as the number of propagating channels
89: decreases has to be answered by theory.   Experimentally, the
90: breakdown of superconductivity manifests itself 
91: as a temperature-dependent
92: resistance below the transition temperature $T_c$, which 
93: becomes more and more pronounced as the rope gets thinner \cite{kasnew}.
94: According to our theory, this resistance is caused
95: by quantum phase slips, and therefore the experimental data 
96: published in Ref.~\cite{kasnew} have in fact explored a 
97: regime of 1D superconductivity with clear evidence for quantum
98: phase slip events that had not been reached before.
99: To the best of our knowledge, nanotube ropes represent wires
100: with the smallest number of propagating channels 
101: showing intrinsic superconductivity, even when
102: compared to the amorphous MoGe wires of diameter
103: $\approx 10$~nm studied in Ref.~\cite{lau}, 
104: where still several thousand channels are available.
105: 
106: We theoretically analyze superconductivity in nanotube ropes 
107: by starting from the microscopic model of an array of
108: $N$ individual metallic single-wall nanotubes (SWNTs) without disorder, 
109: with effectively attractive on-tube interactions 
110: and inter-tube Josephson couplings.  A similar model has been
111: suggested by Gonz{\'a}lez \cite{gonzalez1,gonzalez2}.
112: In the absence of the Josephson couplings, each SWNT would then
113: correspond to a {\sl Luttinger liquid}\
114:  with interaction parameter $g_{c+}>1$,
115: where $g_{c+}=1$ marks the noninteracting limit. For simplicity, we
116: take the same $g_{c+}$ on each SWNT.
117: For example, for $(10,10)$ armchair SWNTs, assuming good screening
118: of the repulsive Coulomb interactions, phonon exchange via a breathing
119: mode (as well as optical phonon modes) leads to $g_{c+}\approx 1.3$,
120: see Ref.~\cite{ademarti}.  In the case of attractive interactions,
121: the dominant coupling mechanism between different SWNTs is then given
122: by Cooper pair hopping, while
123: single-particle hopping is drastically  suppressed by 
124: momentum conservation arguments \cite{kane,gonzalez1}.
125: The coupling among different SWNTs is thus encoded in a
126: {\sl Josephson coupling matrix} $\Lambda_{ij}$, where $i,j=1,\ldots,N$.
127: As different nanotube chiralities are randomly distributed in 
128: a rope, only $1/3$ of the SWNTs can be expected to be metallic. 
129: In general, the $\Lambda_{ij}$ matrix should therefore be
130: drawn from an appropriate random distribution.  We consider 
131: below one individual rope with a fixed (but unspecified) 
132: matrix, and derive general statements valid for
133: arbitrary $\Lambda_{ij}$.  In that sense, our theory allows to
134: capture some disorder effects, at least qualitatively.   However,
135: since typical elastic mean free paths in SWNTs exceed $1 \mu$m 
136: \cite{nts}, disorder effects within individual SWNTs are ignored completely. 
137: The above reasoning leads us to the problem of 
138: $N$ coupled strongly correlated Luttinger liquids, where 
139: the number of ``active'' chains $N \lesssim 100$ with 
140: reference to the experiments of Ref.~\cite{kasnew}.
141: This is a difficult problem that neither permits 
142: the use of classical Ginzburg-Landau (GL) theory
143: nor of the standard BCS approach,
144: in contrast to the situation encountered in,
145: e.g., wide quasi-1D organic superconductors \cite{schulz}.
146: 
147: The approach taken in this paper is sketched next.
148: After a careful derivation of the coupled-chain action in Sec.~\ref{sec2}, we
149: proceed by introducing the appropriate order parameter
150: field.  In Sec.~\ref{sec3}, we then perform a cumulant expansion in this order
151: parameter, and thereby give a 
152: microscopic derivation of the quantum GL action,  which 
153: then allows to make further progress.  We establish the
154: temperature regime where this theory is reliable, and
155: then focus on the important phase fluctuations of the
156: order parameter field. At temperatures $T$ well below a
157: mean-field transition temperature $T_c^0$, amplitude fluctuations
158: are shown to be massive, and hence the amplitude can safely 
159: be treated in mean-field theory.  The massless phase fluctuations
160: then capture the important physics, and  
161: we specify the resulting effective low-energy action, valid at 
162: temperatures well below $T_c^0$.
163: Based on this action, Sec.~\ref{sec4} explains why
164: quantum phase slips (QPSs) 
165: \cite{tinkham,zaikin1,zaikin2,blatter} are 
166: crucial for an understanding of the experimental results of 
167: Refs.~\cite{kociak,kasnew}.  First, they cause a depression of the
168: transition temperature $T_c$ below the mean-field critical 
169: temperature  $T_c^0$.  
170: Furthermore, for $T<T_c$, a finite resistance $R(T)$ due to QPSs appears, 
171: which exhibits approximate power-law scaling.
172: We determine the full temperature dependence 
173: of $R(T<T_c)$ for arbitrary rope length in Sec.~\ref{sec5}.
174: In Sec.~\ref{sec6}, we then compare these results for $R(T)$
175: to the experimental data of Ref.~\cite{kasnew}, focussing on two
176: of their samples.  Finally, Sec.~\ref{sec7} offers some concluding
177: remarks.   Throughout the paper, we put $\hbar=k_B =1$.
178: 
179: 
180: 
181: \section{Model and order parameter} \label{sec2}
182: 
183: 
184: We consider a rope consisting of $N$
185: metallic SWNTs participating in superconductivity.  
186: Experimentally, this number can be found from 
187: the residual resistance measured as offset in the
188: resistance when extrapolating down to $T=0$ 
189: \cite{kasnew}.  Due to the attached  normal electrodes
190: in any two-terminal measurement of the rope,
191: despite of the presence of superconductivity, there will always be 
192: a finite contact resistance $R_c$. 
193: Since each metallic tube contributes two conduction channels, 
194: assuming good transparency for the contacts between metallic tubes
195: and the electrodes,
196: this is given by 
197: \begin{equation}\label{contactres}
198: R_c= \frac{R_Q}{2N}, \quad R_Q=h/2e^2\simeq 12.9\, k\Omega.
199: \end{equation}
200: Extrapolation of experimental data for the resistance $R(T)$ down
201: to $T\to 0$ within the
202: superconducting regime then allows to measure $R_c$, and hence $N$.
203: Good transparency of the contacts is warranted by the sputtering technique
204: used to fabricate and contact the suspended rope samples in the experiments
205: of Refs.~\cite{kociak,kasnew}.
206: An alternative way to estimate $N$ comes from atomic force microscopy,
207: which allows to measure the apparent radius of the rope, and hence yields
208: an estimate for the total number of tubes in the rope.  On average, 1/3
209: of the tubes are metallic \cite{nts},  and one should obtain
210: the same number $N$ from this approach. 
211: Fortunately, these two ways of estimating $N$ provide
212: consistent results in most samples \cite{kasnew}. Therefore the values
213: for $N$ used below are expected to be reliable.
214: 
215: 
216: Here we  always assume that phonon exchange leads to attractive 
217: interactions overcoming the (screened) Coulomb interactions.
218: This assumption can be problematic in ultrathin ropes, where practically
219: no screening arises unless there are close-by gate electrodes. 
220: For sufficiently large rope radius, however, theoretical arguments
221: supporting this scenario have been provided in Ref.~\cite{gonzalez3}.
222: In the absence of intra-tube disorder, then
223: the appropriate low-energy theory for  an
224: individual SWNT is the Luttinger liquid (LL) 
225: model \cite{egger97,kane97,ademarti}.
226: The LL theory of SWNTs 
227: is usually formulated within the Abelian  bosonization
228: approach \cite{gogolin}.  
229: With ${\bf x}=(x,\tau)$, where $x$ is the spatial 1D 
230: coordinate along the tube, and 
231: $0\leq \tau < 1/T$ is imaginary time, 
232:  and corresponding integration measure $d{\bf x}=dx d\tau$, 
233: the action for a single SWNT is 
234: \cite{egger97,kane97,ademarti}
235: \begin{eqnarray}\label{bosac}
236: S_{\rm LL} & = &  \int d{\bf x} \sum_{a=c\pm,s\pm}
237:  \frac{v_a}{2 g_a} \left[ (\partial_\tau \varphi_{a}/v_a)^2  + 
238: (\partial_x \varphi_{a})^2 \right] \nonumber \\ 
239: &=& \int d {\bf x } \sum_{a}
240: \frac{v_a g_a}{2} \left[ (\partial_\tau \theta_{a}/v_{a})^2  +  
241: (\partial_x \theta_{a})^2 \right],
242: \end{eqnarray}
243: which we take to be the same for every SWNT. 
244: Due to the electron spin and the additional K point degeneracy present
245: in nanotubes \cite{nts}, there are four channels, $a=c+,c-,s+,s-$,
246: corresponding to the total/relative charge/spin modes \cite{egger97,kane97},
247: with associated boson fields $\varphi_{a}({\bf x})$ and
248: dual fields $\theta_a({\bf x})$ \cite{gogolin}.
249: In the $a=(c+,s-)$ channels, the second (dual) formulation
250: turns out to be  more convenient, while the first line of Eq.~(\ref{bosac})
251: is more useful for $a=(s+,c-)$. 
252: The combined effect of Coulomb and phonon-mediated
253: electron-electron interactions results in the interaction parameter $g_{c+}$,
254: where we assume $g_{c+}>1$, reflecting effectively attractive
255: interactions \cite{ademarti}.  
256: In the neutral channels, there are only very weak
257: residual interactions, and we therefore put $g_{a\neq c+}=1$.
258: Finally, the velocities $v_a$ in Eq.~(\ref{bosac})
259: are defined as $v_a=v_F/g_a$, where
260: $v_F=8\times 10^5$~m$/$sec is the Fermi velocity.
261: 
262: 
263: Next we address the question which processes trigger the strongest
264: superconducting fluctuations in a nanotube rope.  This question
265: has been addressed in Refs.~\cite{gonzalez1,gonzalez2,ademarti}, 
266: and the conclusion of these studies is that Cooper pairs predominantly
267: form on individual SWNTs rather than involving electrons on different
268: SWNTs, see, e.g., the last section in Ref.~\cite{ademarti} for a 
269: detailed discussion.  Furthermore, the dominant intra-tube 
270: fluctuations involve {\sl singlet}\ (rather than triplet) Cooper 
271: pairs.  The relevant order parameter for superconductivity is then given by 
272: \cite{egger98}
273: \begin{equation}\label{orderpar2}
274: {\cal O}({\bf x})= \sum_{r\sigma\beta} \sigma \psi_{r,\sigma,\beta} 
275: ({\bf x}) \psi_{-r,-\sigma,-\beta}({\bf x}) ,
276: \end{equation} 
277: where $\psi_{r\sigma \beta}$ denotes the electron field 
278: operator for a right- or left-moving electron ($r=\pm$) 
279: with spin $\sigma=\pm$ and K point degeneracy index $\beta=\pm$.
280: In bosonized language, this operator can be expressed as \cite{egger98}
281: \begin{eqnarray}\label{orderpar}
282: {\cal O}&=& \frac{1}{\pi a_0}
283: \cos[\sqrt{\pi} \theta_{c+}] \cos[\sqrt{\pi} \varphi_{c-}] \\
284: \nonumber &\times & 
285: \cos[\sqrt{\pi} \varphi_{s+}] \cos[\sqrt{\pi} \theta_{s-}] - 
286: (\cos\leftrightarrow \sin),
287: \end{eqnarray}
288: where we identify the  UV cutoff
289: necessary in the bosonization scheme with 
290: the graphite lattice constant, $a_0=0.24$~nm.
291: In what follows, we use the shorthand notation $\varphi_j$
292: to label all four boson fields $\varphi_a$ (or their dual fields)
293: corresponding to the $j$th SWNT, where $j=1,\ldots,N$.
294: 
295: The next step is to look at possible couplings among the
296: individual SWNTs.  In principle, three different processes
297: should be taken into
298: account, namely (i) direct Coulomb interactions,
299: (ii) Josephson couplings, and (iii) single-electron hopping.  
300: The last process is strongly
301: suppressed due to the generally different chirality of 
302: adjacent tubes \cite{kane}, and, in addition, for $g_{c+}>1$,
303: inter-SWNT Coulomb interactions are irrelevant \cite{schulz}.
304: Furthermore, as discussed in detail in Ref.~\cite{ademarti},
305: phonon-exchange mediated interactions between
306: {\sl different}\ SWNTs can always be neglected against
307: the intra-tube interactions.
308: Therefore the most relevant mechanism is Josephson
309: coupling between metallic SWNTs.  These couplings define
310: a Josephson matrix $\Lambda_{jk}$,
311: which contains the amplitudes for
312: Cooper pair hopping from the $j$th to the $k$th SWNT.
313: We put $\Lambda_{jj}=0$, and hence $\Lambda$ is a real, symmetric,
314: and traceless matrix.  It therefore has only real eigenvalues 
315: $\Lambda_\alpha$, which we take in descending order,
316: $\Lambda_1\geq \Lambda_2\geq \ldots\geq \Lambda_N$.
317: Moreover, there is at least one positive and at least one negative
318: eigenvalue.  The largest eigenvalue $\Lambda_1$ will be shown
319: to determine the mean-field critical temperature $T_c^0$ below.
320: The matrix $\Lambda$ is then expressed in the corresponding orthonormal
321: eigenbasis $|\alpha\rangle$,
322: \begin{equation}\label{lambdade}
323: \Lambda_{jk}= \sum_{\alpha} \langle j | \alpha \rangle \Lambda_\alpha
324: \langle \alpha | k \rangle ,
325: \end{equation}
326: where $\langle j | \alpha \rangle$ is the  
327: real orthogonal transformation from the basis of lattice 
328: points $\{|j\rangle\}$ to the basis
329: $\{|\alpha\rangle\}$ that diagonalizes $\Lambda$.
330: Clearly,  $\langle j | \alpha \rangle = \langle \alpha | j \rangle $.  
331: In what follows, we define $\alpha_0$ 
332: such that $\Lambda_{\alpha}>0$ for $\alpha<\alpha_0$.
333: 
334: 
335: The Euclidean action of the rope is then
336: \begin{equation}\label{ea}
337: S = \sum_{j=1}^N S_{\rm LL}[\varphi_{j}] - \sum_{jk} 
338: \Lambda_{jk}\int d{\bf x} \, {\cal O}^\ast_j {\cal O}_k^{},
339: \end{equation}
340: where ${\cal O}_j$ is the order parameter specified in Eq.~(\ref{orderpar}).
341: The action (\ref{ea}) defines the model that is 
342: studied in the remainder of our paper.  
343: For studies of closely related models, see
344: also Refs.~\cite{schulz,carr}. 
345: 
346: In order to decouple the Josephson term in Eq.~(\ref{ea}),  
347: we employ a Hubbard-Stratonovich transformation. 
348: To that purpose,
349: since the Josephson matrix has at least one negative eigenvalue,
350: we first express $\Lambda$ in its eigenbasis, see Eq.~(\ref{lambdade}).
351: The Josephson term in Eq.~(\ref{bosac}) is then rewritten as
352: \[
353: \sum_{jk}
354: {\cal O}_j^* \Lambda_{jk} {\cal O}^{}_k = \sum_{\alpha}\ 
355: {\rm sgn}(\Lambda_\alpha) 
356: | \Lambda_\alpha |  {\cal O}_\alpha^* {\cal O}^{}_{\alpha} ,
357: \]
358: where the order parameter in the $|\alpha\rangle$ basis is 
359: \begin{equation}\label{expa00}
360: {\cal O}_\alpha^{} \equiv \sum_i \langle \alpha | i 
361: \rangle {\cal O}^{}_i, \quad
362: {\cal O}_\alpha^* \equiv \sum_i {\cal O}_i^* \langle i | \alpha \rangle.
363: \end{equation}
364: By introducing a field $\Delta^{}_\alpha({\bf x})$ for each 
365: Josephson eigenmode \cite{foot1},
366: with (formally independent) complex conjugate field
367: $\Delta^\ast_\alpha$, it is now possible to 
368: perform the Hubbard-Stratonovich transformation following
369: the standard procedure \cite{nagaosa1}.
370: With integration measure ${\cal D} \Delta  = \prod_{\alpha} 
371: {\cal D}\Delta^*_\alpha  
372: {\cal D}\Delta^{}_\alpha$,  the effective action entering
373: the partition function $Z=\int {\cal D}\Delta \exp(-S_{\rm eff}[\Delta])$ 
374: reads
375: \begin{equation} 
376: \label{effac}
377: S_{\rm eff}[\Delta]= S_0[\Delta] + \int d{\bf x} \sum_{\alpha} 
378:  \Delta^\ast_\alpha  \frac{1}{|\Lambda_\alpha|} 
379: \Delta_\alpha ,
380: \end{equation}
381: where the action $S_0[\Delta]$ is formally defined via the remaining 
382: path integral over the boson fields $\varphi_{j}$,
383: \begin{eqnarray} \label{f00}
384: S_0[\Delta] &=& -\ln \int
385: \prod_{j=1}^N {\cal D}\varphi_{j} \, e^{ -\sum_j S_{\rm LL}[\varphi_{j}]} 
386: \times \nonumber \\
387: & \times & e^{- \int d{\bf x}
388: \sum_{\alpha} c_{\alpha} \left(
389: \Delta^*_\alpha {\cal O}^{}_\alpha +
390: {\cal O}^*_\alpha \Delta^{}_\alpha
391: \right)}, 
392: \end{eqnarray}
393: with $c_\alpha = 1$ for $\alpha <\alpha_0$,
394: and $c_\alpha = i$ otherwise.
395: 
396: \section{Quantum Ginzburg-Landau approach}\label{sec3}
397: 
398: \subsection{Cumulant expansion}
399: 
400: Clearly, closed analytical evaluation of the path integral
401: in Eq.~(\ref{f00}) is in general impossible. 
402: In order to make progress, approximations are necessary,
403: and in the following we shall construct and analyze the
404:  Ginzburg-Landau (GL) action 
405: \cite{nagaosa1,tinkham} for this problem.
406: It turns out to be essential to take into account
407: quantum fluctuations, i.e., the imaginary-time dependence of the
408: order parameter field $\Delta_\alpha(x,\tau)$.
409: In the standard (static) Ginzburg-Landau theory, such effects are
410: ignored.   
411: 
412: The derivation of the GL action
413: proceeds from a cumulant expansion 
414: of Eq.~(\ref{f00}) up to quartic order in the $\Delta_\alpha$.
415: This is a systematic expansion 
416: in the parameter $|\Delta|/2\pi T$ \cite{nagaosa1}, 
417: and by self-consistently computing this parameter, one
418: can determine the regime of validity of GL theory.
419: We stress that this expansion is {\sl not}\ restricted to $N\gg 1$.
420: In addition, for the long-wavelength low-energy regime of primary interest
421: here, we are entitled to perform a gradient
422: expansion.  
423: Using the single-chain correlation function
424: $G({\bf x}_{12})=\langle {\cal O}({\bf x}_1)
425: {\cal O}^\ast({\bf x}_2)\rangle$ 
426: of the operator ${\cal O}$ in Eq.~(\ref{orderpar})
427:  with respect to the free boson action
428: $S_{\rm LL}$,  and the connected four-point correlation function
429: \begin{eqnarray*}
430: G^{(4)}_c({\bf x}_1,{\bf x}_2,{\bf x}_3,{\bf x}_4) &=& 
431: \langle {\cal O}({\bf x}_1) 
432: {\cal O}({\bf x}_2) {\cal O}^*({\bf x}_3) {\cal O}^*({\bf x}_4)\rangle 
433:  \\
434: &-& \langle {\cal O}({\bf x}_1) {\cal O}^*({\bf x}_3) \rangle
435: \langle {\cal O}({\bf x}_2) {\cal O}^*({\bf x}_4)\rangle \\  
436: &-&\langle {\cal O}({\bf x}_1) {\cal O}^*({\bf x}_4) \rangle
437: \langle {\cal O}({\bf x}_2) {\cal O}^*({\bf x}_3)\rangle ,
438: \end{eqnarray*}
439: the cumulant-plus-gradient expansion 
440: up to quartic order yields for
441: the effective Lagrangian density 
442: \begin{eqnarray}\label{lagal}
443: L[\Delta] &=& \sum_{\alpha<\alpha_0}
444: \Bigl[ C\, |\partial_x \Delta^{}_\alpha|^2 + 
445: D\, |\partial_\tau \Delta^{}_\alpha |^2 \\ 
446: &+&  \left( \Lambda_\alpha^{-1} - A \right) |\Delta^{}_\alpha|^2 
447: \Bigr]  \nonumber \\ \nonumber
448: &+& B \sum_{\alpha_i<\alpha_0} 
449: f^{\alpha_1,\alpha_2}_{\alpha_3,\alpha_4}
450: \, \Delta^*_{\alpha_1}
451: \Delta^*_{\alpha_2} \Delta^{}_{\alpha_3} 
452: \Delta^{}_{\alpha_4} ,
453: \end{eqnarray}
454: where we use the notation
455: \[
456: f^{\alpha_1,\alpha_2}_{\alpha_3,\alpha_4}=\sum_i 
457: \langle \alpha_1 | i\rangle \langle \alpha_2 | i\rangle
458: \langle i | \alpha_3 \rangle
459: \langle i | \alpha_4 \rangle .
460: \]
461: The temperature-dependent positive coefficients $A, B, C, D$ are obtained as
462: \begin{eqnarray} 
463: A  &=& \int d{\bf x} \, G({\bf x}) ,\label{coefA}\\
464: \label{coefB}
465: B & = & -\frac{1}{4}\int d{\bf x}_{1} d{\bf x}_2 
466: d{\bf x}_3 \, G^{(4)}_c({\bf x}_1,{\bf x}_2,{\bf x}_3,{\bf x}_4), \\
467: C &=& \frac12 \int d{\bf x} \, x^2 G({\bf x}) ,\label{coefC}\\
468: D &=& \frac12 \int d{\bf x} \, \tau^2 G({\bf x}) .\label{coefD}
469: \end{eqnarray}
470: Due to translation invariance, the integral for $B$
471: does not depend on ${\bf x}_4$.
472: Besides temperature, these coefficients basically depend only 
473: on the important LL interaction parameter $g_{c+}$.
474: In particular, as it is discussed below, for $g_{c+} >1$, 
475: the coefficient $A$ grows as $T$ is lowered. 
476: For static and uniform configurations, modes with
477: $\alpha>\alpha_0$ never become critical.  
478: One can then safely integrate over these modes, which leads to a 
479: renormalization of the parameters governing the remaining 
480: modes. Such renormalization effects are however tiny,
481: and  thus are completely neglected in Eq.~(\ref{lagal}).
482: 
483: At this stage, it is useful to switch to
484: an order parameter field defined on the $j$th SWNT,
485: \begin{equation}\label{opontube}
486: \Delta_j=\sum_{\alpha<\alpha_0} \langle j| \alpha \rangle \Delta_\alpha .
487: \end{equation}  
488: After some algebra, the Lagrangian density (\ref{lagal}) can 
489: be written as
490: \begin{eqnarray}
491: L[\Delta]  =  \sum_{j=1}^N \Bigl[
492: C  |\partial_x \Delta_j^{}|^2 + 
493: D  |\partial_\tau \Delta_j^{}|^2 + B |\Delta_j|^4 &+& \nonumber \\
494: + \left( \Lambda_1^{-1} - A \right) |\Delta^{}_j|^2 \Bigr] 
495: + \sum_{jk}\Delta^\ast_j V_{jk} \Delta^{}_k , &{}&  
496: \label{hs2} 
497: \end{eqnarray}
498: with the real, symmetric, and positive definite matrix
499: \begin{equation}\label{vjk}
500: V_{jk}=\sum_{\alpha< \alpha_0} 
501: \langle j | \alpha \rangle  \left( \Lambda_\alpha^{-1} -
502: \Lambda_1^{-1} \right) \langle \alpha | k \rangle. 
503: \end{equation}
504: Notice that, strictly speaking,
505:  the fields $\Delta_i$ are not all independent, 
506: because we have defined them from the subset of positive modes.
507: The transformation in Eq.~(\ref{opontube}) is indeed not invertible.  
508: Nevertheless, in the following, we treat them as formally independent.
509: This only affects the precise values of the $V_{ij}$ but does
510: not qualitatively change our results.
511: The expectation value of the order parameter field
512: (\ref{orderpar}) can be expressed in terms of linear combinations
513: of the fields $\Delta_j(x,\tau)$, and hence it is indeed justified
514: to call $\Delta_j$ a  proper ``order parameter field''.
515: 
516: Equation (\ref{hs2}) specifies the full GL action, taking
517: into account quantum fluctuations and transverse modes
518: for arbitrary number $N$ of active SWNTs. 
519: In the limit $N\to\infty$, and considering only static field
520: configurations, results similar to those of Ref.~\cite{schulz} are
521: recovered.  In that limit the last term in Eq.~(\ref{hs2}) gives 
522: indeed the gradient term in the transverse direction, and one
523: obtains the standard 3D GL Lagrangian.
524: There is however an important difference, namely the starting point of 
525: Ref.~\cite{schulz} is a model of Josephson-coupled 
526: 1D superconductors, whereas we start from an array
527: of metallic chains with $g_{c+} > 1$, where the 
528: inter-chain Josephson coupling is crucial in stabilizing
529: superconductivity.
530: More similar to ours is the model investigated in Ref.~\cite{carr}.
531: However, in that paper, the metallic chains are assumed to have a spin gap,
532: which is not the case for the SWNTs in a rope in the 
533: temperature range of interest. Furthermore, the main focus 
534: in Ref.~\cite{carr} is the competition between charge density wave and
535: superconducting instabilities, whereas in our case, as discussed above,
536: the formation of a charge density wave is strongly suppressed
537: by compositional disorder, i.e., different chiralities of adjacent 
538: tubes, and we do not have to take the corresponding instability 
539: into account.
540: 
541: \subsection{Ginzburg-Landau coefficients}  
542: 
543: 
544: In order to make quantitative predictions, it is necessary to compute
545: the GL coefficients defined in Eqs.(\ref{coefA})-(\ref{coefD}).   
546: While this is possible in principle for the
547: full four-channel model (\ref{bosac}), here we will instead
548: derive the coefficients
549: for a simpler model, where the K point degeneracy is neglected.
550: This leads to an effective spin-$1/2$ Luttinger liquid action
551: with interaction parameter $g_c$ ($g_s=1$) 
552: and velocity $v_c=v_F/g_c$ ($v_s=v_F$).
553: Up to a prefactor of order unity, the respective results can be matched onto
554: each other.  This can be made explicit, e.g., for the coefficient $A$,
555: where we get from the full action (\ref{bosac}) 
556: \[
557: A= \frac{c}{v_F}
558: \left( \frac{\pi a_0 T}{v_{c+}} \right)^{(g_{c+}^{-1}-1)/2} .
559: \]
560: The proportionality constant $c$ is 
561: found to differ from $\tilde A/2\pi^2$ 
562: [see Eq.~(\ref{coefA2}) below, which follows from
563: the spin-$1/2$ description] only by a factor of order unity. 
564: In the simpler model neglecting the K point
565: degeneracy, one then needs to take 
566: \[
567: g_c^{-1}=\frac{1+g^{-1}_{c+}}{2},
568: \]
569: which gives, for $g_{c+}=1.3$, a value of $g_c \approx 1.1$.
570: This way, all exponents of the resulting power-law correlation functions
571: (which are the physically relevant quantities)
572: in the ``reduced'' model 
573: are the same as in the complete model, and only 
574: prefactors of order unity may be different 
575: for the respective GL coefficients.
576: The bosonized order parameter (\ref{orderpar}) in the
577: simpler model is then given by
578: \[
579: {\cal O}= \frac{1}{\pi a_0} \cos[\sqrt{2\pi} \varphi_s]
580: \exp[i\sqrt{2\pi}\theta_c].
581: \]
582: Using the finite-temperature correlation functions of 
583: the fields $\theta_c$ and $\varphi_s$ \cite{gogolin},
584: \begin{eqnarray*}
585: \langle \theta_c({\bf x})\theta_c({\bf 0})\rangle &=& \frac{-1}{2\pi g_c} 
586: \ln \left(\frac{v_c}{\pi a_0 T} 
587: \left |\sinh \frac{\pi T (x+iv_c\tau)}{v_c}\right|\right) , \\
588: \langle \varphi_s({\bf x}) \varphi_s({\bf 0}) \rangle &=& \frac{-1}{2\pi} 
589: \ln \left(\frac{v_F}{\pi a_0 T}
590: \left|\sinh \frac{\pi T(x+iv_F\tau)}{v_F}\right|\right), 
591: \end{eqnarray*}
592: and rescaling the integration variables $x$ and $\tau$ 
593: in Eqs.~(\ref{coefA})-(\ref{coefD}), 
594:  explicit expressions follow in the form
595: \begin{eqnarray}
596: A(T) &=& \frac{1}{2\pi^2 v_F}  ( \pi a_0 T/v_c )^{g_c^{-1}-1}
597: \tilde{A},\label{coefA2}\\ \nonumber
598: B(T) &=& \frac{a_0^2 }{32\pi^4 v_c v_F^2} 
599:  ( \pi a_0 T/v_c )^{2g_c^{-1} -4} \tilde{B}, \\ \nonumber
600: C(T) &=& \frac{a_0^2}{4\pi^2 v_F} 
601:  ( \pi a_0 T/v_c )^{g_c^{-1}-3} \tilde{C},\\
602: D(T) &=& \frac{a_0^2}{4\pi^2 v_Fv_c^2} 
603:  ( \pi a_0 T/v_c )^{g_c^{-1}-3} \tilde{D}. \nonumber
604: \end{eqnarray}
605: Dimensionless $g_c$-dependent numbers 
606: $\tilde A, \tilde B, \tilde C, \tilde D$
607: were defined as follows.  With the notation ${\bf z}=(w,u)$
608: and 
609: \[
610: \int d{\bf z}= 
611:  \int_{0}^{\pi} du \int_{-\infty}^\infty dw,
612: \]
613: we have
614: \begin{eqnarray*}
615: \tilde{A}&=& \int \frac{d{\bf z}}{ f_c({\bf z}) f_s({\bf z})}, \\
616: \tilde{C}&=& \int d{\bf z} \frac{w^2}{ f_c({\bf z}) f_s({\bf z})},\\ 
617: \tilde{D}&=& \int d{\bf z} \frac{u^2}{ f_c({\bf z}) f_s({\bf z})}, 
618: \end{eqnarray*}
619: where functions $f_{c,s}$ are introduced as
620: \begin{eqnarray*}
621: f_c({\bf z}) &=& |\sinh(w+iu)|^{1/g_c},\\
622: f_s({\bf z})  &=& |\sinh(w/g_c + iu)|.
623: \end{eqnarray*}
624: The coefficient of the quartic term in the GL functional is
625: \begin{eqnarray*}
626: & &\tilde B = \int \frac{d{\bf z}_1 d{\bf z}_2 d{\bf z}_3}
627: {f_c({\bf z}_2) f_c({\bf z}_{13})} \Biggl[   
628: \frac{4}{f_s({\bf z}_2)f_s({\bf z}_{13})}-
629:  \frac{f_c({\bf z}_1) f_c({\bf z}_{23})}
630:     {f_c({\bf z}_3)f_c({\bf z}_{12})} \\ &\times&  \left( 
631:    \frac{f_s({\bf z}_1)f_s({\bf z}_{23})} 
632: {f_s({\bf z}_2)f_s( {\bf z}_{13}) f_s({\bf z}_3)f_s({\bf z}_{12})}
633: + (1\leftrightarrow 2) + (1 \leftrightarrow 3) \right)  
634: \Biggr ] 
635: \end{eqnarray*}
636: with ${\bf z}_{ij}=(w_i-w_j, u_i-u_j)$. 
637:  The quantity $\tilde{B}$
638: is evaluated using the Monte Carlo method.
639: For $g_c=1$, we first numerically reproduced
640: the exact result $\tilde{B}=8\pi^2 \tilde{C}$ 
641: with $\tilde{C}=7\pi \zeta(3)/4$ \cite{nagaosa1}.
642: Numerical values can then be obtained for arbitrary $g_c$.
643: Numerical evaluation yields for $g_c \approx 1.1$ (corresponding to
644: $g_{c+}=1.3$) the following results:
645: \begin{equation}\label{tildes}
646: \tilde{A} \simeq 17.4 , \quad \tilde{B} \simeq 392(1) , 
647: \quad \tilde{C}\simeq 8.15 ,\quad \tilde{D}\simeq 6.97 . 
648: \end{equation}
649: 
650: \subsection{Mean-field transition temperature}
651: 
652: Since in the rope only a modest number of transverse modes are present, 
653: a natural definition of the mean-field critical temperature $T_c^0$
654: is the temperature at which the mode corresponding to 
655: the largest eigenvalue of $\Lambda$ 
656: becomes critical. {}From Eq.~(\ref{hs2}), this leads to 
657: the condition $A(T)=\Lambda_1^{-1}$,  and hence to
658:  the mean-field critical temperature
659: \begin{equation}\label{tc}
660: T_c^0 = \frac{v_c}{\pi a_0} \left( \frac{\tilde A \Lambda_1}{2\pi^2 v_F}
661: \right)^{g_c/(g_c-1)} ,
662: \end{equation}
663: which exhibits a dependence on the number $N$ of 
664: active SWNTs in the rope through $\Lambda_1$.  For large
665: $N$, the eigenvalue $\Lambda_1$ saturates, and Eq.~(\ref{tc})
666: approaches the bulk transition temperature. 
667: 
668: To provide concrete theoretical
669:  predictions for $T_c^0$ is difficult,
670: since the Josephson matrix is in general unknown, and the 
671: results for $T_c^0$ very sensitively depend on $\Lambda_1$.
672: Using estimates of Ref.~\cite{gonzalez2} and typical $N$ as reported
673: in Ref.~\cite{kasnew},  as an order-of-magnitude estimate,
674: we find  $T_c^0$ values around $0.1$ to 1~K.  When comparing 
675: to experimental results, $\Lambda_1$ can be inferred from the
676: actually measured $T_c$, which in turn provides values in 
677: reasonable agreement with theoretical expectations \cite{gonzalez1}.
678: 
679: 
680: \subsection{Low-energy theory: $T < T_c^0$}
681: 
682: In what follows, we focus on temperatures $T<T^0_c$.  
683: Then it is useful to employ an amplitude-phase representation of the
684: order parameter field,
685: \begin{equation}\label{densphas}
686: \Delta_j ({\bf x})=|\Delta_j|({\bf x}) \exp [i\phi_j({\bf x})],
687: \end{equation}
688: where the amplitudes $|\Delta_j|$ are expected to be
689: finite with a gap for fluctuations
690: around their mean-field value.  At not too low temperatures, the GL action
691: corresponding to Eq.~(\ref{hs2}) is accurate (see below), 
692: and the mean-field values
693: follow from the saddle-point equations.
694: Considering only static and uniform field configurations, 
695: we find $\phi_i\equiv \phi$, where in principle also other (frustrated)
696: configurations with $\phi_i-\phi_j=\pm \pi$ could contribute.
697: Such configurations presumably correspond to maxima of the free energy,
698: and are ignored henceforth.
699: The saddle-point equations then reduce to equations for the amplitudes alone,
700: \begin{equation}\label{gap}
701: \sum_{j} V_{ij} |\Delta_j| + (\Lambda_1^{-1}-A)|\Delta_i| +
702:  2B |\Delta_i|^3 =0 ,
703: \end{equation}
704: whose solution yields the transverse order parameter profile.
705: Numerical study of Eq.~(\ref{gap}) using a standard Newton-Raphson
706: root-finding algorithm then allows to 
707: extract the profile $\{|\Delta_j|\}$ 
708: for a given Josephson matrix $\Lambda_{ij}$.
709: We briefly discuss the solution of Eq.~(\ref{gap})
710: for the idealized model of a rope as a trigonal 
711: lattice exclusively composed of
712: $N$ metallic SWNTs, where $\Lambda_{ij}=\lambda$ for nearest
713: neighbors $(i,j)$, and $\Lambda_{ij}=0$ otherwise.  For this model,
714: Fig.~\ref{fig1} shows the  resulting average amplitude 
715: $\Delta_0=\sum_i |\Delta_i|/N$ as a function of temperature for
716: $\lambda/v_F=0.1$ and two values of $N$.
717: Since $\Delta_0/2\pi T$ is the expansion parameter entering
718: the construction of the GL functional, and it
719: remains small down to $T\approx T_c^0/2$, we
720: conclude that the GL theory is self-consistently valid 
721: in a quantitative way down to such temperature
722: scales. In our discussion below, GL theory turns 
723: out to be qualitatively useful even down to $T=0$. 
724: 
725: \begin{figure}
726: \centerline{\epsfxsize=7cm\epsfysize=6cm 
727: \epsffile{prof.eps}}
728: \caption{\label{fig1}  
729: Temperature dependence of $\Delta_0/2\pi T$ versus $T/T_c^0$ for
730: $N=31$ (open circles) and $N=253$ (filled circles).  }
731: \end{figure}
732: 
733: Fixing the amplitudes $|\Delta_j|$ at their mean-field values, and
734: neglecting the massive amplitude fluctuations around these values, 
735: the Lagrangian follows from Eq.~(\ref{hs2}) as
736: \begin{eqnarray}\label{final}
737: L&=& \sum_{j=1}^N
738:  \frac{\mu_j}{2\pi} \left[ c_s^{} (\partial_x\phi_j)^2 + c_s^{-1}
739: (\partial_\tau \phi_j)^2 \right]\\ \nonumber
740: & + & \sum_{i>j} 2V_{ij}|\Delta_i||\Delta_j| 
741: \cos(\phi_i-\phi_j) ,
742: \end{eqnarray}
743: with the Mooij-Sch\"on velocity \cite{mooij},
744: \begin{equation}\label{css}
745: c_s \equiv v_c \sqrt{\tilde{C}/\tilde{D}}, 
746: \end{equation} 
747: and dimensionless phase stiffness parameters
748: \begin{equation}\label{stiff1}
749: \mu_j= 2\pi C |\Delta_j|^2 /c_s.
750: \end{equation}
751: At this stage, electromagnetic potentials can be coupled in via standard
752: Peierls substitution rule \cite{nagaosa1}, 
753: and dissipative effects due to the electromagnetic environment
754: can be incorporated following Ref.~\cite{blatter}.
755: 
756: 
757: \section{1D action and quantum phase slips}\label{sec4}
758: 
759: \subsection{1D phase action}\label{sec41}
760: 
761: Numerical evaluation of Eq.~(\ref{gap}) 
762: shows that for $T$ well below $T_c^0$,
763: transverse fluctuations are heavily suppressed.
764: While this statement only applies to amplitude fluctuations,
765: one can argue that also the transverse phase fluctuations 
766: are strongly suppressed.  The basic argument relates to 
767: the scaling dimension [in the renormalization group (RG) sense]
768: of the operator $\cos(\phi_i-\phi_j)$, which is essentially
769: governed by the $\mu_j$. For $T$ well below $T_c^0$, the 
770: $\mu_j$ become large, and the cosine operators get strongly
771: relevant, locking the phases all together.  In the low-temperature
772: regime of main interest below, this argument allows to substantially
773: simplify Eq.~(\ref{final}).
774: Then also no detailed knowledge about the Josephson 
775: matrix is required, because the only relevant information 
776: is essentially contained in $T^0_c$.
777: 
778: Putting all phases $\phi_j=\phi$, 
779: we arrive at a standard (Gaussian) 1D superconducting phase action 
780: \cite{tinkham},
781: \begin{equation}\label{finala}
782: S=\frac{\mu}{2\pi}\int dx d\tau \left[
783: c_s^{-1}(\partial_\tau\phi)^2 + c_s (\partial_x \phi)^2 \right],
784: \end{equation} 
785: with dimensionless rigidity
786: $\mu = \sum_j \mu_j$, see Eq.~(\ref{stiff1}), 
787: and $c_s$ as given in Eq.~(\ref{css}).
788: Assuming GL theory to work even down to $T=0$ for the moment, and neglecting 
789: the $V_{ij}$-term in Eq.~(\ref{gap}), a simple analytical estimate follows
790: in the form
791: \begin{equation} \label{mu1}
792: \mu(T) = N \nu \left[ 1 - (T/T_c^0)^{(g_c-1)/g_c}\right] , 
793: \end{equation}
794: where the number $\nu$  is
795: \begin{equation} \label{defnu}
796: \nu =4\pi \tilde{A} (\tilde{C}\tilde{D})^{1/2}  / \tilde{B}.
797: \end{equation}
798: The peculiar temperature dependence of the phase stiffness in Eq.~(\ref{mu1}),
799: reflecting the underlying LL physics of the individual SWNTs, is one of 
800: the main results of this paper.
801: In the effective spin-$1/2$ description employed here,
802: using the numbers specified
803: in Eq.~(\ref{tildes}) for $g_c=1.1$ results in $\nu\approx 4$.
804: Remarkably, at
805: $T=0$,  Eq.~(\ref{mu1}) coincides, up to a prefactor of order unity,
806: with the rigidity $\bar{\mu}$ obtained from standard 
807: mean-field relations \cite{nagaosa1},
808: \[
809: \bar{\mu}= \pi^2 n_s R^2/2 m^\ast c_s = \bar{\nu} N.
810: \]
811: With the density of condensed electrons $n_s$ and rope radius $R$, 
812: this implies $\bar{\nu}\approx v_F/c_s$, which is of order unity. 
813: We therefore conclude that the GL prediction (\ref{mu1}) for $\mu(T)$ is 
814: robust and useful even outside its strict validity regime.
815: 
816: The result (\ref{mu1}) for the stiffness is central for the
817: following discussion. The value we obtain for $\nu$, however,
818: should not be taken as a very precise estimate.
819: First, it can be affected
820: by factors of order unity under a full four-channel calculation 
821: taking into account the K point degeneracy,
822: as this affects each of the numbers in Eq.~(\ref{tildes}) by a 
823: factor of order unity.  Second, uncertainties in the parameter $g_c$ 
824: will also affect $\nu$ by a factor of order unity.  Moreover, based
825: on the discussion in Ref.~\cite{zaikin2}, one expects
826: on general grounds that intra-SWNT disorder and
827: dissipative effects, both of which are not included 
828: in our model, will  effectively lead to a {\sl decrease} of the parameter
829: $\nu$ entering  Eq.~(\ref{mu1}).
830: Therefore $\nu$ is taken below as a fit
831: parameter when comparing to experimental data. 
832: Since the number of active SWNTs $N$ can be estimated 
833: from the residual resistance, 
834: and the transition temperature $T_c$, see Eq.~(\ref{tc1}) below, can be
835: determined  from the experimentally observed transition temperature, 
836:  $\nu$ is basically the only free remaining parameter. 
837: Fits of our theoretical results 
838: to experimental data are then expected to yield values for $\nu$ around 
839: $\nu\approx 1$. This is verified below in Sec.~\ref{sec6}.
840: 
841: \subsection{Phase slips}
842: 
843: In the 1D situation encountered here, superconductivity
844: can be destroyed by phase slips \cite{tinkham}. 
845: A phase slip (PS) can be visualized as a process in which
846: fluctuations locally destroy the amplitude of the superconducting 
847: order parameter, which effectively disconnects the 1D
848: superconductor into two parts. Simultaneously, the phase, 
849: being defined only up to $2\pi$, is allowed to ``slip'' by $2\pi$
850: across the region where the amplitude vanishes. 
851: This process then leads to finite dissipation in the superconducting wire
852: via the
853: Josephson effect. Depending on temperature, phase slips 
854: can be produced either by thermal or by quantum fluctuations. 
855: In the first case, which is commonly realized very near the critical 
856: temperature, we have a thermally activated 
857: phase slip (TAPS). At lower temperature, 
858: the quantum tunneling mechanism dominates, 
859: and one speaks of a quantum phase slip (QPS).
860: For a textbook description of quantum phase slips, see Ref.~\cite{chaikin}. 
861: Below we demonstrate that in superconducting ropes, only QPSs
862: are expected to play a prominent role. 
863: 
864: A QPS is a topological vortex-like excitation of the
865: superconducting phase field $\phi(x,\tau)$
866: that solves the equation of motion for
867: the action (\ref{finala}) with a 
868: singularity at the core, where superconducting order 
869: is locally destroyed and a phase cannot be defined.
870: Defining a thermal lengthscale as
871: \begin{equation}\label{lt}
872: L_T=  c_s/ \pi T  ,
873: \end{equation}
874: for rope length $L\to \infty$ and $L_T\to \infty$,
875: a QPS with core at $(x_i,\tau_i)$ and winding number $k_i=\pm 1$ 
876: (higher winding numbers are irrelevant)
877: is given by \cite{chaikin}
878: \begin{equation}\label{qps}
879: \phi(x,\tau)=k_i \arctan 
880: \left[\frac{c_s(\tau-\tau_i)}{(x-x_i)}\right] ,
881: \end{equation}
882: where the finite $L,L_T$ solution 
883: follows by conformal transformation \cite{blatter}.
884: The action of a QPS consists of two terms, one associated with the local
885: loss of condensation energy, the core action $S_c$, and  
886: the other with the vortex strain energy. % $S_{\rm el}$.
887: While a detailed computation of $S_c$ requires a microscopic 
888: description of the dynamics inside the vortex core \cite{zaikin2}, 
889: a simple qualitative argument is able to predict an order-of-magnitude
890: estimate $S_c\approx \mu/2$ \cite{chaikin}.
891: 
892: This result allows us to assess the relative contribution
893: of the TAPS and QPS mechanisms.  
894: The production rate for the creation of one
895: vortex is \cite{zaikin2} 
896: $\gamma_{\rm QPS}\approx \frac{S_c L c_s}{\kappa} \exp(-S_c)$,
897: where $\kappa$ is the core size.
898: Within exponential accuracy, comparing this formula to the respective 
899: standard TAPS rate expression \cite{tinkham},  
900: the crossover temperature from
901: TAPS- to QPS-dominated behavior is 
902: $T^*_{\rm PS}= 2\Delta F/ N\nu$,
903: with activation barrier $\Delta F$.  Using results of 
904: Ref.~\cite{carr}, we estimate the latter as 
905: $\Delta F=8\sqrt{2} R(g_c) N T_c^0/3$,
906: with dimensionless coefficient $R(g_c)$ of order unity.  
907: Finally, this implies $T^*_{\rm PS}\approx  T_c^0$.
908: Since the true transition temperature $T_c<T_c^0$, see below,
909: in the temperature regime $T<T_c$, the influence of a TAPS 
910: can safely be neglected against the QPS.
911: 
912: The generalization to many QPSs then leads to the standard picture 
913: of a Coulomb gas of charges $k_i=\pm 1$, with fugacity $y=e^{-S_c}$, total 
914: charge zero, and logarithmic interactions \cite{nagaosa1,chaikin}.  
915: The partition function  $Z=Z_G Z_V$ contains a regular factor 
916: $Z_G$ and the vortex contribution
917: \begin{equation} \label{partfncvor}
918: Z_V = \sum_{n=0}^\infty \frac{y^{2n}}{(n!)^2} 
919:  \int \frac{\prod_{m=1}^{2n} d{\bf r}_m}{(c_s\kappa^2)^{2n}}
920: \sum_{\{k \}} e^{\mu \sum_{i\neq j} k_i k_j \ln (r_{ij}/\kappa)}. 
921: \end{equation}
922: This model undergoes a Berezinski-Kosterlitz-Thouless
923: transition driven by the nucleation of vortices, here corresponding to
924: a transition from a phase $\mu>\mu^*$,  where QPSs are 
925: confined into  neutral pairs and the rope forms a 1D superconductor 
926: with finite phase stiffness 
927: and quasi-long-range order,  to a phase $\mu <\mu^*$ where
928: QPSs proliferate.  In that phase, vortices  are deconfined and  
929: destroy the phase stiffness, thereby producing normal behavior,
930: where ``normal'' does of course not imply Fermi-liquid behavior.
931: The phase boundary is located at $\mu^*=2+4\pi y \simeq 2$. 
932: The true transition temperature $T_c$ is therefore {\sl not}\ the mean-field
933: transition temperature $T_c^0$, but follows
934: from the condition $\mu(T_c)=\mu^\ast$. Putting $\mu^\ast=2$, 
935: Eq.~(\ref{mu1}) yields
936: \begin{equation}\label{tc1}
937: T_c/T_c^0 = \left[1- 2/ N\nu \right]^{g_c/(g_c-1)}.
938: \end{equation}
939: This $T_c$ depression is quite sizeable for $N\alt 100$. To give 
940: concrete numbers, taking $\nu=1$, for $N=25,50,$ and 100, the 
941: ratio $T_c/T_c^0$ equals $0.40, 0.63,$ and $0.80$, respectively.
942: QPSs also have an important and observable effect in the 
943: superconducting regime, as will be discussed in the next section.
944: 
945: \begin{figure}
946: \centerline{\epsfxsize=7cm \epsfysize=6cm
947: \epsffile{figevo.eps}}
948: \caption{\label{figevo}  
949: Temperature-dependent resistance $R(T<T_c)$ 
950: predicted by Eq.~(\ref{resis})
951: for $\nu=1$ and different $N$. The smaller is $N$, the broader is
952: the transition. From the leftmost to the rightmost curve, 
953: $N=4,7,19,37,61,91,127,169,217$.
954: }
955: \end{figure}
956: 
957: \section{Resistance below $T_c$}\label{sec5}
958: 
959: A phase slip produces finite dissipation through the Josephson
960: effect, and therefore introduces a finite resistance even
961: in the superconducting state, $T<T_c$.   The QPS-induced linear
962: resistance $R(T)=V/I$ for $T<T_c$ can be computed perturbatively in the
963: QPS fugacity $y$ \cite{zaikin1}. For that purpose, we imagine that one 
964: imposes a small current $I$ to flow through the rope.
965: The presence of QPSs implies that a voltage drop $V$ occurs,
966: which is related to the average change in phase,
967: \[
968: V=\frac{\langle \dot\phi \rangle}{2e}=
969: \frac{\pi}{e}[\Gamma(I)-\Gamma(-I)] ,
970: \]
971: where $\Gamma(\pm I)$  is the rate for a phase slip 
972: by $\pm 2\pi$ \cite{zaikin1}.  This rate can be obtained
973: following Langer \cite{langer} as
974: the imaginary part acquired by the 
975: free energy $F(I)$ under an appropriate analytic continuation,
976: \begin{equation}
977: \Gamma(\pm I)= -2 \, {\rm Im} \, F(\pm I) . 
978: \end{equation} 
979: We  only consider the contribution of a single pair of QPSs, 
980: i.e., compute $R(T)$ to second order in $y$.  
981: Expanding Eq.~(\ref{partfncvor}) to order $y^2$, 
982: the free energy at this order reads
983: \begin{equation}
984: F = - \frac{Ly^2 c_s^2}{\kappa^4} \int_0^{1/T} d\tau 
985: \int_{-L/2}^{L/2} dx \, e^{\epsilon \tau -2 \mu g_E(x,\tau)} , 
986: \end{equation}
987: where the vortex-vortex interaction $g_E(x,\tau)$ 
988: only depends on relative coordinates, and 
989: $\epsilon=\pi \hbar I/e$.  The contribution $F_G$ to  
990: the free energy due to regular configurations can be dropped, because
991: it does not acquire an imaginary part under the analytic continuation.
992: We  now perform the analytic continuation $\tau \rightarrow it$, 
993: resulting in [see Ref.~\cite{weissbook} for details]
994: \begin{equation}
995: {\rm Im} \, F = - \frac{Ly^2 c_s^2}{2\kappa^4} 
996: \int_{-L/2}^{L/2} dx \int_{-\infty}^{\infty} dt \,
997: e^{i \epsilon t -2 \mu g(x,t)} ,
998: \end{equation}
999: where $g(x,t)\equiv g_E(x,\tau\rightarrow it)$.
1000: The rate $\Gamma(\epsilon)$ then follows 
1001: for $L,L_T \gg \kappa$ but arbitrary $L/L_T$ in the form 
1002: \[
1003: \Gamma(\epsilon)=\frac{c_s^2 L y^2}{\kappa^4}
1004: \int_{-L/2}^{L/2} dx \int_{-\infty}^\infty dt \, e^{i\epsilon t
1005: - \mu [\tilde g(t+x/c_s)+ \tilde g(t-x/c_s)]} ,
1006: \]
1007: where 
1008: \[
1009: \tilde g(t) = \ln \left[(L_T/\kappa) \sinh(\pi T|t|)\right]
1010: + i (\pi/2) {\rm sgn}(t).
1011: \]
1012: Analyticity of $g_E(x,\tau)$ in the strip $0\leq\tau\leq 1/T$ 
1013: also leads to the standard detailed balance relation 
1014: \cite{weissbook},
1015: \[
1016: \Gamma(-\epsilon)=e^{- \epsilon/T} \Gamma(\epsilon) .
1017: \]
1018: In order to  explicitly evaluate the rate $\Gamma(\epsilon)$ for 
1019: arbitrary $L/L_T$, we now replace the boundaries for the $x$-integral 
1020: by a soft exponential cutoff, 
1021: switch to integration variables
1022: $t'=t-x/c_s$ and $t^{\prime \prime}=t+x/c_s$, 
1023: and use the auxiliary relation
1024: \[
1025: e^{-c_s | t^{\prime\prime}- t'|/L}= \frac{c_s}{\pi L}
1026: \int_{-\infty}^\infty ds \frac{e^{-is(t^{\prime\prime}-t')}}
1027: {s^2+(c_s/L)^2}.
1028: \]
1029: The $t',t^{\prime\prime}$ time integrations then decouple,
1030: and it is straightforward to carry them out.
1031: Finally some algebra yields the linear resistance 
1032: in the form
1033: \begin{eqnarray}\label{resis}
1034: \frac{R(T)}{R_Q} &=& \left(  \frac{\pi y \Gamma(\mu/2)}{ \Gamma(\mu/2+1/2)}
1035: \right)^2 \frac{\pi L}{2\kappa} \left(\frac{L_T}{\kappa}\right)^{3-2\mu} 
1036: \\ \nonumber
1037:  &\times &  \int_{0}^\infty du \frac{2/\pi}{1+u^2}
1038: \left|\frac{\Gamma(\mu/2+iu L_T/2L)} {\Gamma(\mu/2)} \right|^4 ,
1039: \end{eqnarray}
1040: in units of the resistance quantum $R_Q$ defined in Eq.~(\ref{contactres}).
1041: 
1042: 
1043: 
1044: For $L/L_T\gg 1$, the $u$-integral approaches unity, 
1045: and hence $R\propto T^{2\mu-3}$,
1046: while for $L/L_T\ll 1$, dimensional scaling arguments
1047: give $R\propto T^{2\mu-2}$.  In Refs.~\cite{kociak,kasnew},
1048: typical lengths were $L\approx 1~\mu$m, which puts one into the
1049: crossover regime $L_T\approx L$. 
1050: While the quoted power laws have already been reported for diffusive wires 
1051: \cite{zaikin1}, Eq.~(\ref{resis}) describes the full
1052: crossover for arbitrary $L/L_T$, and applies to
1053: strongly correlated ladder compounds such as nanotube ropes.
1054: It predicts that 
1055: the transition gets significantly 
1056: broader upon decreasing the number of tubes in the rope.
1057: This is shown in Fig.~\ref{figevo}, where the theoretical
1058: results for the resistance is
1059: plotted for various $N$ at $\nu=1$.
1060: Note that Eq.~(\ref{resis}) is a perturbative result
1061: in the fugacity, and it is expected to break down close to $T_c$,
1062: see below.
1063: In the next section we directly compare Eq.~(\ref{resis}) to 
1064: experimental data obtained by Kasumov {\sl et al.} \cite{kasnew}.
1065: 
1066: \section{Comparison to experimental data}\label{sec6}
1067: 
1068: \begin{figure}
1069: \centerline{\epsfxsize=7cm \epsfysize=6cm \epsffile{figR2.eps}}
1070: \caption{\label{figR2}  
1071: Temperature dependence of the resistance below $T_c$ for 
1072: superconducting rope 
1073: $R2$ experimentally studied in Ref.~\cite{kasnew}. 
1074: Open squares denote experimental data 
1075: (with subtracted residual resistance), 
1076: the curve is the theoretical result.
1077: }
1078: \end{figure}
1079: 
1080: Here we discuss how the prediction for the 
1081: temperature-dependent resistance $R(T)$ below
1082: $T_c$ as given in Eq.~(\ref{resis}) compares to the experimental
1083: results for $R(T)$ published in Ref.~\cite{kasnew}. 
1084: More aspects of this comparison will be given elsewhere \cite{next}.
1085: The experimental data in Ref.~\cite{kasnew} were obtained from
1086: two-terminal measurements of ropes suspended between 
1087: normal electrodes. 
1088: Due to the presence of the contacts, 
1089: the residual resistance (\ref{contactres}) survives down to $T=0$ 
1090: even when the rope exhibits a superconducting transition.
1091: Extrapolation of experimental results for $R(T)$ yields $R_c$, which then 
1092: allows to infer the number $N$ in the respective sample
1093:  from Eq.~(\ref{contactres}). 
1094: This resistance $R_c$ has to be subtracted from experimental
1095: data to allow
1096: for a comparison with Eq.~(\ref{resis}), where no contact resistance
1097:  is taken into account.
1098: 
1099: In Figs.~\ref{figR2} and \ref{figR4}, experimental
1100: resistance curves (after this subtraction)
1101: for the samples named $R2$ and $R4$ 
1102: in Ref.~\cite{kasnew} are plotted versus the prediction 
1103: of Eq.~(\ref{resis}). For sample 
1104: $R2$, we find $R_c= 74~\Omega$ 
1105: corresponding to $N_{R2}=87$, while for sample $R4$, the 
1106: subtracted resistance is $R_c=150~\Omega$, 
1107: leading to $N_{R4}=43$.  We then take these $N$ values 
1108: when computing the respective
1109: theoretical curves.
1110: The experimentally determined
1111: temperature $T^\ast$ locates the onset
1112: of the transition \cite{kasnew}, and 
1113: is identified with the true transition temperature $T_c$
1114: in Eq.~(\ref{tc1}). It is therefore also not a free parameter. 
1115: Note that thereby the eigenvalue $\Lambda_1$ of the Josephson
1116: matrix has been determined.
1117: In the absence of detailed knowledge about the
1118: Josephson matrix, it is fortunate that our result 
1119: for $R(T)/R(T_c)$ following from Eqs.~(\ref{resis}) 
1120: and (\ref{mu1}) does not require more information about
1121: $\Lambda$ besides the largest eigenvalue.
1122: Given the estimate $g_{c+}=1.3$
1123: \cite{ademarti}, the comparison of Eq.~(\ref{resis}) 
1124: to experimental data then allows only one free fit 
1125: parameter, namely $\nu$.  According to our discussion
1126: in Sec.~\ref{sec41},  the fit is expected to yield  values $\nu \approx 1$.
1127: 
1128: The best fit to the low-temperature experimental curves for $R(T)$
1129:  yields $\nu=0.75$ for sample $R2$, see Fig.~\ref{figR2},
1130: and $\nu=0.16$ for sample $R4$, respectively.  The agreement between
1131: experiment and theory is excellent for sample $R2$.  For sample $R4$,
1132: the optimal $\nu$ is slightly smaller than expected, which indicates that
1133: dissipative processes may be more important in that sample.
1134: Nevertheless, for both samples, the low-temperature
1135: resistance agrees quite well, with only one free fit parameter that
1136: is found to be of order unity as expected.
1137: Whereas the agreement between theoretical and experimental curves
1138: appears then quite satisfactory in the low-temperature region, 
1139: our predictions clearly deviate in the region near $T_c$. 
1140: This is not surprising, because
1141: our expression for $R(T)$ in Eq.~(\ref{resis}) 
1142: is perturbative in the QPS fugacity. It is then expected to
1143: break down close to $T_c$, where QPSs proliferate and the approximation
1144: of a very dilute gas of QPS pairs, on which 
1145: our calculation is based, is not valid anymore. 
1146: As a consequence, also the saturation observed 
1147: experimentally above  $T^\ast$ is not captured.
1148: 
1149: We note that it is an interesting challenge to compute the 
1150: finite resistance in the normal phase at $T_c<T<T_c^0$,
1151: where the saturation should be caused by
1152: QPS and TAPS proliferation.  For temperatures $T>T_c^0$, superconducting
1153: correlations can be neglected, and the resistance should then be dominated
1154: by phonon backscattering and disorder effects.
1155: Nevertheless, we believe that the 
1156: agreement between the theoretical resistance result (\ref{resis}) 
1157: and experimental data at low temperatures shown
1158: in Figs.~\ref{figR2} and \ref{figR4}, 
1159: given the complexity of this system,  is rather satisfactory.
1160: More importantly, this comparison 
1161: provides strong evidence for the presence of quantum phase
1162: slips in superconducting nanotube ropes.  
1163: 
1164: \section{Conclusions} \label{sec7}
1165: 
1166: According to our discussion above,
1167: the intrinsic superconductivity observed in ropes of 
1168: SNWTs \cite{kasnew,kociak} represents a 
1169: remarkable phenomenon, where it has been
1170: possible to experimentally probe the extreme 1D
1171: limit of a superconductor.
1172: In this paper, we have formulated a theory for this phenomenon,
1173: based on a model of metallic SWNTs with 
1174: attractive intra-tube interactions
1175: and arbitrary inter-tube Josephson couplings. 
1176: The analysis of this model leads to an effective Ginzburg-Landau 
1177: action, whose coefficients can be expressed in 
1178: terms of parameters entering the 
1179: microscopic description of the rope. In order
1180: to get the correct low-energy dynamics, it is crucial 
1181: to include  quantum fluctuations of the order parameter.
1182: Based on the resulting low-energy action for the phase 
1183: fluctuations, we have shown that quantum phase slips 
1184: produce a depression of the critical temperature. 
1185: More importantly, the temperature dependence of the linear resistance 
1186: experimentally observed below the transition temperature 
1187: can be accounted for by considering the underlying LL physics 
1188: and the effect of quantum phase slips.
1189: Despite some admittedly crude approximations, like the neglect of
1190: intra-tube disorder and dissipation effects inside the vortex core, 
1191: the comparison of 
1192: experimental curves and theoretical predictions, 
1193: in particular in the low-temperature region, strongly 
1194: suggests that the resistive process is indeed dominated by 
1195: quantum phase slips.
1196: Our theory also suggests that, if repulsive Coulomb interactions 
1197: can be efficiently screened off, superconductivity may survive down 
1198: to only very few transverse channels in clean nanotube ropes.  
1199: 
1200: 
1201: \begin{figure}[t]
1202: \centerline{\epsfxsize=7cm \epsfysize=6cm \epsffile{figR4.eps}}
1203: \caption{\label{figR4}  
1204: Same as Fig.~\ref{figR2}, but for sample $R4$ experimentally
1205: studied in Ref.~\cite{kasnew}.
1206: }
1207: \end{figure}
1208: 
1209: \section*{Acknowledgments}
1210: 
1211: We thank H. Bouchiat and M. Ferrier
1212: for providing the data shown in Figs.~\ref{figR2} and
1213: \ref{figR4} and for many discussions. One of us (ADM) would like 
1214: to thank them also for their warm hospitality 
1215: during an extended stay in Orsay.
1216: This work has been supported by the EU network DIENOW
1217: and the SFB-TR 12 of the DFG.
1218: 
1219: 
1220: 
1221: \begin{references}
1222: 
1223: \bibitem{nts} 
1224: C. Dekker, Physics Today {\bf 52(5)}, 22 (1999);
1225: See also reviews on nanotubes in Physics World No.~{\bf 6} (2000).
1226: 
1227: \bibitem{nts2}
1228: M. Dresselhaus, G. Dresselhaus and Ph. Avouris (Eds.),
1229: Carbon Nanotubes, Topics in Appl. Physics {\bf 80}
1230: (Springer Verlag 2001).
1231: 
1232: \bibitem{kociak}
1233: M. Kociak, A.Yu. Kasumov, S. Gueron, B. Reulet, I.I. Khodos,  
1234: Yu.B. Gorbatov, V.T. Volkov, L. Vaccarini, and H. Bouchiat,
1235: Phys. Rev. Lett. {\bf 86}, 2416 (2001).
1236: 
1237: \bibitem{kasnew}
1238: A. Kasumov, M. Kociak, M. Ferrier, R. Deblock, S. Gueron, B. Reulet, 
1239: I. Khodos, O. Stephan, and H. Bouchiat, Phys. Rev. B {\bf 68}, 214521 (2003).
1240: 
1241: \bibitem{tang}
1242: Z.K. Tang, L. Zhang, N. Wang, X.X. Zhang, G.H. Wen, G.D. Li, J.N. Wang, C.T. 
1243: Chan, and P. Sheng, Science {\bf 292}, 2462 (2001).
1244: 
1245: \bibitem{kasumov}
1246: A.Yu. Kasumov,  R. Deblock, M. Kociak, B. Reulet, H. Bouchiat, I.I. 
1247: Khodos,  Yu.B. Gorbatov, V.T. Volkov, C. Journet, and M. Burghard,
1248: Science {\bf 284}, 1508 (1999).
1249: 
1250: \bibitem{morpurgo}
1251: A.F. Morpurgo, J. Kong, C.M. Marcus, and H. Dai,
1252: Science {\bf 286}, 263 (1999).
1253: 
1254: \bibitem{lau}
1255: C.N. Lau, N. Markovic, M. Bockrath, A. Bezryadin, and M. Tinkham,
1256: Phys. Rev. Lett. {\bf 87}, 217003 (2001).
1257: 
1258: \bibitem{gonzalez1} J. Gonz{\'a}lez, 
1259: Phys. Rev. Lett. {\bf 88}, 076403 (2002).
1260: 
1261: \bibitem{gonzalez2} J. Gonz{\'a}lez, 
1262: Phys. Rev. B {\bf 67}, 014528 (2003).
1263: 
1264: \bibitem{ademarti} A.~De~Martino and R. Egger,
1265: Phys. Rev. B {\bf 67}, 235418 (2003).
1266: 
1267: \bibitem{kane} A.A. Maarouf, C.L. Kane, and E.J. Mele,
1268: Phys. Rev. B {\bf 61}, 11156 (2000).
1269: 
1270: \bibitem{schulz} H.J. Schulz and C. Bourbonnais,
1271:   Phys. Rev. B {\bf 27}, 5856 (1983).
1272: 
1273: \bibitem{tinkham} M. Tinkham, {\sl Introduction to Superconductivity}, 2nd ed.
1274: (McGraw-Hill, Inc., 1996).
1275: 
1276: \bibitem{zaikin1} 
1277:   A.D. Zaikin, D.S. Golubev,  A. van Otterlo, and G.T. Zimanyi,
1278: Phys. Rev. Lett. {\bf 78}, 1552 (1997).
1279: 
1280: \bibitem{zaikin2} D.S. Golubev and A.D. Zaikin,
1281: Phys. Rev. B {\bf 64}, 014504 (2001).
1282: 
1283: \bibitem{blatter}
1284: H.P. B{\"u}chler, V.B. Geshkenbein, and G. Blatter,
1285: Phys. Rev. Lett.  {\bf 92}, 067007 (2004).
1286: 
1287: \bibitem{gonzalez3} J. Gonz{\'a}lez, Eur. Phys. J. B {\bf 36}, 317 (2003).
1288: 
1289: \bibitem{egger97} R. Egger and A.O. Gogolin,
1290: Phys. Rev. Lett. {\bf 79}, 5082 (1997).
1291: 
1292: \bibitem{kane97}
1293:   C. Kane, L. Balents, and M.P.A. Fisher,
1294: Phys. Rev. Lett. {\bf 79}, 5086 (1997).
1295: 
1296: \bibitem{gogolin} A.O. Gogolin, A.A. Nersesyan, and A.M. Tsvelik,
1297: {\sl Bosonization and Strongly Correlated Systems} (Cambridge University 
1298: Press, 1998).
1299: 
1300: \bibitem{egger98} R. Egger and A.O. Gogolin,
1301: Eur. Phys. J B {\bf 3}, 281 (1998).
1302: 
1303: \bibitem{carr} S.T. Carr and A.M. Tsvelik,
1304: Phys. Rev. B {\bf 65}, 195121 (2002).
1305: 
1306: \bibitem{foot1}
1307: Modes with $\Lambda_\alpha=0$ have to be excluded in the
1308: transformation. All $\alpha$ 
1309: summations have to be understood in this sense. 
1310: Note that such modes never cause critical behavior in any case.
1311: 
1312: \bibitem{nagaosa1} N. Nagaosa,
1313: {\sl Quantum Field Theory in Condensed Matter Physics}
1314: (Springer Verlag, 1999).
1315: 
1316: \bibitem{mooij} 
1317:   J. E. Mooij and G. Sch{\"o}n,
1318: Phys. Rev. Lett. {\bf 55}, 114 (1985).
1319: 
1320: \bibitem{chaikin} P.M. Chaikin and T. Lubensky,
1321:   {\sl Principles of Condensed Matter Physics} 
1322: (Cambridge University Press, 2000).
1323: 
1324: \bibitem{langer}
1325: J.S. Langer, Ann. Phys. (N.Y.) {\bf 41}, 108 (1967).
1326: 
1327: \bibitem{weissbook}
1328: U. Weiss, {\sl Quantum Dissipative Systems}, 2nd. ed. 
1329: (World Scientific, Singapore, 1999).
1330: 
1331: \bibitem{next}
1332: M. Ferrier, A. De~Martino, A. Kasumov, R. Deblock, S. Gueron, R. Egger,
1333: and H. Bouchiat, invited review, submitted to Solid State Communications.
1334: 
1335: \end{references}
1336: 
1337: \end{document}
1338: