1: %% Journal: Phys. Rev. B
2: %% Title: Exchange interaction in quantum rings and wires in the
3: %% Wigner-crystal limit
4: %% Authors: Michael M. Fogler and Eugene Pivovarov
5: %% Contact address: Dr Michael M. Fogler, Department of Physics,
6: %% Mail Code 0319, University of California San Diego,
7: %% 9500 Gilman Dr, La Jolla, CA 92093-0319
8: %% Contact e-mail: mfogler@ucsd.edu
9:
10: \documentclass[prb,twocolumn,showpacs,floatfix]{revtex4}
11: \usepackage{amsmath,amsfonts,amssymb,graphicx,bm}
12:
13: \begin{document}
14:
15: \title{Exchange interaction in quantum rings and wires in the Wigner-crystal
16: limit}
17: \author{Michael M. Fogler}
18: \author{Eugene Pivovarov}
19: \affiliation{Department of Physics, University of California San Diego, La Jolla,
20: California 92093}
21: \date{\today}
22:
23: \begin{abstract}
24:
25: We present a controlled method for computing the exchange coupling in
26: correlated one-dimensional electron systems based on the relation
27: between the exchange constant and the pair-correlation function of
28: spinless electrons. This relation is valid in several independent
29: asymptotic regimes, including low electron density case, under the
30: general condition of a strong spin-charge separation. Explicit formulas
31: for the exchange constant are obtained for thin quantum rings and wires
32: with realistic Coulomb interactions by calculating the pair-correlation
33: function via a many-body instanton approach. A remarkably smooth
34: interpolation between high and low electron density results is
35: shown to be possible. These results are applicable to the case of
36: one-dimensional wires of intermediate width as well. Our method can be
37: easily generalized to other interaction laws, such as the inverse
38: distance squared one of the Calogero-Sutherland-Moser model. We
39: demonstrate excellent agreement with the known exact results for the
40: latter model and show that they are relevant for a realistic
41: experimental setup in which the bare Coulomb interaction is screened by
42: an edge of a two-dimensional electron gas.
43:
44: \end{abstract}
45:
46: \pacs{71.10.Pm, 73.21.Hb, 73.22.-f}
47: \maketitle
48:
49: %% 71.10.Pm Fermions in reduced dimensions (anyons,
50: %% composite fermions, Luttinger liquid, etc.)
51: %% 73.21.Hb Quantum wires
52: %% 73.22.-f Electronic structure of nanoscale materials:
53: %% clusters, nanoparticles, nanotubes, and nanocrystals
54: %% 75.75.+a Magnetic properties of nanostructures
55:
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57:
58: \section{Introduction}
59: \label{sec:Introduction}
60:
61: Much interest has been devoted to the spin degree of freedom in novel
62: one-dimensional (1D) conductors, both of linear shape (carbon
63: nanotubes,\cite{Saito-book} semiconductor nanowires,\cite{Huang-01}
64: conducting molecules\cite{Heath-03}) and recently, of a circular one
65: (quantum rings\cite{Lorke-00, Fuhrer-04, Bayer-03}). Parameters of
66: these systems, e.g., average distance between the electrons $a$, their
67: total number $N$, their effective mass $m$, dielectric constant
68: $\epsilon $, effective Bohr radius $a_{B}=\hbar ^{2}\epsilon /me^{2}$,
69: \textit{etc.}\/, can vary over a broad range or can be tuned
70: experimentally. This creates an intriguing opportunity of reaching the
71: Wigner-crystal (WC) limit,\cite{Schulz-93,Egger-99} $r_{s}\equiv
72: a/2a_{B} \gg 1$, where electrons arrange themselves into a nearly
73: regular lattice. According to numerical simulations,\cite{Egger-99} the
74: 1D WC is well formed already at $r_s > 4$. The corresponding electron
75: densities are easily achievable but the presence of disorder has so far
76: hindered experimental investigations of the 1D WC
77: regime.\cite{Field-90} Remarkably, the unwanted disorder can apparently
78: be substantially suppressed in the case of carbon nanotubes
79: \emph{suspended\/} above a substrate. In such devices very large $r_s$
80: with no immediately obvious disorder effects have been recently
81: demonstrated.\cite{Jarillo-Herrero-04} Due to a finite length of the
82: nanotubes used, they contained only a few electrons ($< 30$) in the $r_s
83: > 4$ regime. (Such systems are commonly referred to as Wigner
84: molecules.) These encouraging progress on the experimental side and a
85: recent revival of interest to the 1D WC on the theoretical
86: side\cite{Matveev-04} have motivated us to undertake in the present
87: paper a careful investigation of the fundamental energy scales that
88: govern the spin dynamics in 1D Wigner crystals and molecules.
89:
90: In a strict academic sense, the term ``crystal'' should not be used in
91: 1D because the phonon-like vibrations of the putative lattice are
92: infrared divergent. It is nevertheless true that at $r_{s}\gg 1$ the
93: pair correlation function (PCF)
94: %
95: \begin{equation}
96: g(x)=\frac{1}{N n}\sum_{i\neq j}\langle \delta (x_{i}-x_{j}-x)\rangle
97: \label{eqn:g_intro}
98: \end{equation}
99: %
100: is sharply peaked at integer multiples of $a$. Here $x_j$ are electron
101: coordinates. Thus, for the study of electron correlations in Wigner
102: molecules or in a group of few nearby electrons in an infinite wire the
103: WC concept is fully adequate. Once adopted, this concept
104: implies\cite{Herring-62} that the low-energy dynamics of electron spins
105: is dominated by the nearest-neighbor Heisenberg interaction of strength
106: $J$ proportional to an exponentially small probability of quantum
107: tunneling under the Coulomb barriers that separate the adjacent
108: electrons. Exchanges of more distant neighbors are penalized with a much
109: stronger tunneling suppression and can therefore be neglected.
110:
111: %%%%%%%%%%%
112: % FIG. 1
113: %
114: \begin{figure}
115: \begin{center}
116: \includegraphics[height=1.5in]{wcx_1.eps}
117: \end{center}
118: \caption{(Color online) A sketch of the instanton trajectories for
119: the six-electron Wigner molecule on a ring.}
120: \label{fig:ring}
121: \end{figure}
122: %%%%%%%%%%%
123:
124: As a result of the exponential smallness of the exchange coupling $J$,
125: the energy scales for orbital and spin dynamics are drastically
126: different. This \emph{strong\/} spin-charge separation is expected to
127: cause anomalies in many essential electron properties, e.g., ballistic
128: conductance\cite{Matveev-04} of quantum wires and persistent current of
129: quantum rings.\cite{Reimann-02} To systematically explore such
130: phenomena a reliable estimate of the exchange coupling $J$ is needed.
131: This has been an important open problem however. The smallness of $J$
132: makes it difficult to compute even by rather advanced computer
133: simulations.\cite{Reimann-02} Attempts to derive $J$ analytically (for
134: the nontrivial case $N>2$) have so far been based on the approximation
135: that neglects all degrees of freedom in the problem except the distance
136: between the two interchanging electrons.\cite{Hausler-96, Matveev-04} We
137: call this a Frozen Lattice Approximation (FLA). The accuracy of the FLA
138: is unclear because it is not justified by any small parameter. When a
139: given pair does its exchange, it sets all other electrons in motion, too
140: (Fig.~\ref{fig:ring}). Simple dimensional considerations (confirmed by
141: the calculations below) show that the maximum tunneling displacements of
142: the electrons nearby the first pair reach a fraction of $a$, i.e., they
143: are at most numerically, but not \emph{parametrically}, small. In view
144: of the aforementioned and other recent experimental\cite{Auslaender-05,
145: Claessen-02} and theoretical\cite{Cheianov-04, Fiete-04, Fiete-05} work
146: on spin-related effects in 1D conductors a controlled calculation of $J$
147: seems a timely goal. It is accomplished in this paper, where we treat
148: the spin exchange in a Wigner molecule (or a WC) as a true many-body
149: process and compute $J$ to the leading order in the large parameter
150: $r_s$. A brief account of this work has been reported in
151: Ref.~\onlinecite{Fogler-05b}. The same problem has been independently
152: studied by \textcite{Klironomos-05} and wherever they can be compared
153: our results agree.
154:
155: The paper is organized as follows. The definition of the model and our
156: main results are given in Sec.~\ref{sec:Results}. The step-by-step
157: derivation for the simplest nontrivial case of a three-electron molecule
158: is presented in Sec.~\ref{sec:Three}. The $N > 3$ case is discussed in
159: Secs.~\ref{sec:Many} and \ref{sec:Instanton}. An alternative route to
160: the derivation of $J$ is outlined in Sec.~\ref{sec:Alternative}. The
161: comparison with the known analytical results for related models is done
162: in Sec.~\ref{sec:integrable_model}. The issues relevant to the current
163: experiments are addressed at the end of Sec.~\ref{sec:integrable_model}
164: and in Sec.~\ref{sec:Domain}. The concluding remarks are in
165: Sec.~\ref{section:Conclusions}.
166:
167: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
168: \section{Model and main results}
169: \label{sec:Results}
170:
171: We assume that electrons are tightly confined in the transverse
172: dimensions on a characteristic lengthscale (``radius'' of the wire)
173: $R\ll a_{B}$. In this case the energy separation $\hbar^2 / m R^2$ of
174: the 1D subbands greatly exceeds the characteristic Coulomb interaction
175: energy $e^2 / \epsilon R$. Assuming that only the lowest subband is
176: occupied, the Hamiltonian can be projected onto the Hilbert space of
177: this subband. This turns the problem into a strictly 1D one, with the
178: Coulomb law replaced by an effective interaction $U(r)$ that tends to a
179: finite value at distances $r\ll R$. The precise form of $U$ depends on
180: the details of the confinement. For simplicity, we adopt a
181: model\cite{Fogler-05a, Usukura-05}
182: %
183: \begin{equation}
184: U(r) = \frac{e^2}{\epsilon} \frac{1}{r + R},
185: \label{eqn:U}
186: \end{equation}
187: %
188: which nevertheless correctly captures both short- and long-range
189: behavior of the interaction for \emph{any\/} realistic confinement
190: scheme and is similar to other expressions used in the
191: literature,\cite{Friesen-80, Szafran-04} see more in
192: Sec.~\ref{sec:Domain}. Most of our calculations below pertain to the
193: ring geometry where $r$ stands for the chord distance, $r = (Na/\pi
194: )\left\vert \sin (\pi x/Na)\right\vert$, $x$ being the coordinate along
195: the circumference. At large $r_s$ the electrons assume a Wigner-molecule
196: configuration where they reside at the corners of a regular polygon. The
197: effective low-energy Hamiltonian of such a state is given by
198: %
199: \begin{equation}
200: H=\frac{\hbar ^{2}}{2I}L^{2}+J\sum_{j}\mathbf{S}_{j}
201: \mathbf{S}_{j+1}+\sum_{\alpha }n_{a}\hbar \omega _{\alpha },
202: \label{eqn:H}
203: \end{equation}
204: %
205: where $L$ is the center-of-mass angular momentum, $\mathbf{S}_{j}$ are
206: electron spins, and $n_{\alpha }$ are the occupation numbers of
207: ``molecular vibrations.'' (The same
208: Hamiltonian applies to small-$N$ quantum dots.\cite{Reimann-02})
209: At large $r_{s}$, the total moment of inertia $I$ and the vibrational
210: frequencies $\omega_\alpha$ are easy to compute because they are
211: close to their classical values. Our task is to calculate $J$, which
212: is more difficult.
213:
214: We show that the asymptotically exact relation exists between $J$ and
215: the PCF $g_0(x)$ of a \emph{spin polarized\/}
216: 1D system. For an ultrathin wire, $\mathcal{L}\equiv \ln
217: ({a_{B}}/{R})\gg 1$, it is particularly simple:
218: %
219: \begin{equation}
220: J = \frac{e^2 a_B^2}{2\mathcal{L}\epsilon} g_0^{\prime \prime}(0),
221: \quad r_{s}\gg \frac{1}{\mathcal{L}}.
222: \label{eqn:J_from_g_ultrathin}
223: \end{equation}
224: %
225: By virtue of Eq.~(\ref{eqn:J_from_g_ultrathin}), the calculation of $J$
226: reduces to an easier task of computing $g_0(x)$. For large $r_{s}$ our final
227: result has the form
228: %
229: \begin{equation}
230: J=\frac{\kappa }{\left( 2r_{s}\right) ^{5/4}}\frac{\pi }{\mathcal{L}}
231: \frac{e^{2}}{\epsilon a_{B}}\exp \left( -\eta \sqrt{2r_{s}}\,\right) ,
232: \quad r_{s}\gg 1, \label{eqn:J}
233: \end{equation}
234: %
235: where the values of $\eta $ and $\kappa $ are given in
236: Table~\ref{tbl:Results} together with the prediction of the
237: FLA,\cite{Hausler-96,Matveev-04} for comparison.
238: [\textcite{Klironomos-05} independently obtained $\eta =2.79805(5)$ for
239: the case of a wire.]
240: %
241: \begin{table}[tbp]
242: \begin{ruledtabular}
243: \begin{tabular}{lcccccc}
244: $N$ &3 &4 &6 &8 &$\infty$ &$\infty$-FLA\\
245: $\eta$ &2.8009 &2.7988(2) &2.7979(2) &2.7978(2) &2.7978(2) &2.8168 \\
246: $\kappa$ &3.0448 &3.18(6) &3.26(6) &3.32(7) &3.36(7) &2.20 \\
247: \end{tabular}
248: \end{ruledtabular}
249: \caption{Results for Wigner molecules on a ring (finite $N$) and for wires
250: ($N=\infty $). }
251: \label{tbl:Results}
252: \end{table}
253: %
254:
255: In the remainder of the paper we give the derivation of the above results
256: and discuss their consequences for various experimental and theoretical
257: questions.
258:
259: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
260:
261: \section{Three-electron quantum ring}
262: \label{sec:Three}
263:
264: We start with the simplest nontrivial example: three electrons on a
265: ring. Let $0\leq x_{j}<3a$, $j=0,1,2$ be their coordinates along the
266: circumference. We will compute the exchange coupling $J$ between the
267: $j=0$ and the $j=1$ electrons. In the FLA the third electron remains at
268: rest while the other two interchange, as in a classical picture.
269: However, we will show that the consistent calculation of $J$ requires
270: taking quantum fluctuations of this third electron into account.
271:
272: The system has the total of three degrees of freedom: the center of mass
273: $x_{\text{cm}}$, the distance between the exchanging electrons $x\equiv
274: x_{1}-x_{0}$, and the distance between the third electron and the center
275: of mass, $X_{2}\equiv x_{2}-x_{\text{cm}}\ -a$. We can restrict the
276: variables $x$ and $X_{2}$ to the fundamental domain, $|x|<3a/2$,
277: $|X_{2}|<a/2$, where only the chosen two electrons can closely approach
278: each other. This allows us to ignore exchanges involving the $j=2$
279: electron, and hence, its spin. Ignoring also the irrelevant
280: center-of-mass motion, we obtain the Hamiltonian
281: %
282: \begin{equation}
283: H_{3} = -\frac{\hbar^2}{2\mu} \partial _{x}^{2}
284: - \frac{\hbar^2}{2M} \partial_{X_2}^2
285: + U_{\text{tot}}(x, X_2),
286: \label{eqn:H_3}
287: \end{equation}
288: %
289: where $\mu = m / 2$ and $M = 3 \mu$. The potential term
290: %
291: \begin{align}
292: U_{\text{tot}} & = U(x)+U_{\text{ext}},
293: \\
294: U_{\text{ext}} & = U\left[ (3/2)(X_{2}+a)-x/2\right]
295: \notag\\
296: & + U\left[ (3/2)(X_{2}+a)+x/2\right] ,
297: \notag
298: \end{align}
299: %
300: has two minima in the fundamental domain, at $x=\pm a$, $X_{2}=0$. The
301: minima are separated by a high potential barrier at $x=0$. They give
302: rise to the two lowest-energy states of the system: the spin-singlet ground
303: state, $\mathbf{S}_{0}+\mathbf{S}_{1}=0$, with an orbital wavefunction $\Phi
304: _{s}=\Phi _{s}(x,X_{2})$ and a triplet with a wavefunction $\Phi _{t}$.
305: Their energy splitting is the desired exchange coupling $J$. In close
306: analogy with Herring's classic treatment of H$_2$-molecule\cite{Herring-64}
307: (see also Ref.~\onlinecite{Landau-III}) for our $N=3$ ``molecule''
308: we obtain
309: %
310: \begin{equation}
311: J = \frac{2\hbar^2}{\mu} \int dX_{2}\left. \Phi _{1}\partial _{x}\Phi
312: _{1}\right\vert _{x=0}, \label{eqn:J_from_Phi_1}
313: \end{equation}
314: %
315: where the (normalized to unity) ``single-well'' wavefunction $\Phi _{1}(x,X_{2})$
316: is the ground-state of $H_{3}$ with a modified potential $U_{\text{tot}}\
317: \rightarrow U_{1}\equiv U_{\text{tot}}(\max \{x,0\},X_{2})$.
318: Equation~(\ref{eqn:J_from_Phi_1}) is valid\cite{Herring-64} to order $O(J^{2})$;
319: with the same accuracy,
320: %
321: \begin{equation}
322: \Phi _{s,t}=\frac{1}{\sqrt{2}}[\Phi _{1}(x,X_{2})\pm \Phi _{1}(-x,X_{2})].
323: \end{equation}
324: %
325:
326: Let us discuss the form of the single-well wavefunction
327: $\Phi _{1}(x,X_{2})$. Near its maximum at $x=a$,
328: $X_{2}=0$, it is a simple Gaussian,
329: %
330: \begin{equation}
331: \Phi _{1}\left( x,X_{2}\right) \propto \exp \left[ -\frac{(x-a)^{2}}{2l^{2}}-%
332: \frac{M}{2\hbar }\omega (a)X_{2}^{2}\right],
333: \label{eqn:Phi_1_Gauss}
334: \end{equation}
335: %
336: where $l=[\hbar ^{2}/\mu U_{\text{tot}}^{\prime \prime }(a)]^{1/4}\sim
337: r_{s}^{-1/4}a$ is the amplitude of the zero-point motion in $x$
338: ($U_{\text{tot}}$ written with a single argument is meant to be evaluated at
339: $X_{2}=0$). The quantity
340: %
341: \begin{equation}
342: \omega (x)=\left[ M^{-1}\partial _{X_{2}}^{2}U_{\text{tot}}(x)\right] ^{1/2}
343: \end{equation}
344: %
345: is real and positive in the classically forbidden region $0<x<a-l$.
346: Therefore, the tunneling barrier is the lowest and $\Phi _{1}$ is the
347: largest along the line $X_{2}=0$. Furthermore, it is easy to see that $\Phi
348: _{1}(x,X_{2})$ rapidly decays at $|X_{2}|\gtrsim l$. Since $l\ll a$,
349: the following Gaussian approximation is justified in the \emph{entire\/}
350: fundamental domain of $x$:
351: %
352: \begin{equation}
353: \Phi_{1}=\phi \left( x\right) \exp \left[ -\frac{M}{2\hbar }\Omega \left(
354: x\right) X_{2}^{2}\right].
355: \label{eqn:Phi_1_ansatz}
356: \end{equation}
357: Acting on this \emph{ansatz\/} with $H_{3}$ and neglecting the subleading
358: terms $O(l/a)$, we obtain the following equations on $\Omega(x)$ and
359: $\phi(x)$:
360: %
361: \begin{subequations}
362: \label{eqn:Omega-phi-3}
363: \begin{gather}
364: \partial _{x}\Omega = \frac{\Omega^2(x) - \omega^2(x)}
365: {\left[(2 / \mu) \Delta U_{\text{tot}}(x)\right]^{1/2}},
366: \label{eqn:Omega_3} \\
367: \left\{ \frac{\hbar^2}{2\mu} \partial _{x}^{2}-U_{\text{tot}}(x)- \frac12
368: \hbar \Omega(x) + E\right\} \phi (x)=0,
369: \label{eqn:phi_3}
370: \end{gather}
371: \end{subequations}
372: %
373: where $\Delta U_{\text{tot}}(x)\equiv U_{\text{tot}}(x)-U_{\text{tot}}(a)$.
374: We wish to comment that the basic idea of the Gaussian ansatz has been
375: used previously for computing exchange constants in $^{3}$\textsc{H}e
376: crystals,\cite{Roger-83} but the simple closed form of
377: Eqs.~(\ref{eqn:Omega-phi-3}) has not been demonstrated.
378:
379: Comparing Eqs.~(\ref{eqn:Phi_1_Gauss}) and (\ref{eqn:Phi_1_ansatz})
380: we see that near $x = a$ function $\phi(x)$ takes the form,
381: %
382: \begin{equation}
383: \phi \left( x\right) =\left[ \frac{M\Omega \left( a\right) }{\pi ^{3/2}\hbar
384: l}\right] ^{1/2}e^{-\left( x-a\right) ^{2}/2l^{2}},
385: \end{equation}
386: %
387: while $\Omega(x)$ satisfies the boundary condition $\Omega(a) = \omega
388: (a)$. Resolving the $0/0$ ambiguity in Eq.~(\ref{eqn:Omega_3}) by
389: L'H\^opital's rule we furthermore obtain
390: %
391: \begin{equation}
392: \Omega ^{\prime }\left( a\right) =\frac{\omega ^{\prime }\left( a\right) }
393: {1 + \left\{ \left( \frac{M}{16 \mu }\right) \left[ 1+U^{\prime \prime }\left(
394: a\right) /U_{\text{ext}}^{\prime \prime }\left( a\right) \right] \right\}
395: ^{1/2}}.
396: \end{equation}
397: %
398: Finally, using this formula and straightforward algebraic manipulations one
399: can fix the constant term $E$ in Eq.~(\ref{eqn:phi_3}) to be
400: %
401: \begin{equation}
402: E = U_{\text{tot}}(a)+ \frac{\hbar}{2} \left[ \omega (a)+\omega _{0}\right],
403: \quad \omega_0 \equiv \frac{\hbar}{\mu l^2}.
404: \end{equation}
405: %
406: %
407: % Indeed, self-consistency requires that the derivatives of $\Omega
408: %\left( x\right)$ are small, which implies that
409: %%
410: %\begin{equation}
411: %\frac{\Omega ^{2}-\omega ^{2}}{\Omega ^{2}}\left[ \left( 2/\mu \right)
412: %\Delta U_{\text{tot}}(x)\right] ^{-1/2}\ll \sqrt{\frac{\mu }{\hbar \Omega }}.
413: %\end{equation}
414: %%
415: %This condition is obviously satisfied in the classically forbidden region.
416: %Since $\phi ^{\prime }\left( a\right) =0$, the boundary condition $\Omega
417: %(a)=\omega (a)$ assures that this condition is met everywhere.
418: %
419: Next consider the interior of the classically forbidden region, $x \ll a$.
420: Since the tunneling barrier $\Delta U_{\text{tot}}(x)$ is large here,
421: $\Omega$ is a slow function of $x$. Taking advantage of the following
422: expression for the PCF of a spin-polarized molecule,
423: %
424: \begin{equation}
425: g_0(x)\equiv 2\!\int \prod_{j=2}^{N-1}dX_{j}\Phi _{t}^{2}\left( x,X_{2},\ldots
426: ,X_{N-1}\right),
427: \quad |x|<\frac{3a}{2},
428: \label{eqn:g_def}
429: \end{equation}
430: %
431: we find that Eq.~(\ref{eqn:J_from_Phi_1}) entails the relation
432: %
433: \begin{equation}
434: J = \frac{\hbar^2}{4\mu} \frac{\phi(0)}{\phi^\prime(0)}
435: g_0^{\prime \prime}(0).
436: \label{eqn:J_from_g_triplet}
437: \end{equation}
438: %
439: Anticipating the discussion in Sec.~\ref{sec:Many},
440: Eq.~(\ref{eqn:g_def}) is written for an arbitrary $N>2$, with the
441: notation $X_{j}\equiv x_{j}-x_{\text{cm}}\ +(N-1-2j)(a/2)$ being used;
442: the PCF is normalized as appropriate in the WC limit,
443: %
444: \begin{equation}
445: \int_{0}^{3a/2} g_0(x) dx = 1.
446: \label{eqn:g_normalization}
447: \end{equation}
448: %
449:
450: %%%%%%%%%%%
451: % FIG. 2
452: %
453: \begin{figure}
454: \begin{center}
455: \includegraphics[height=1.3in]{wcx_2a.eps}
456: \end{center}
457: \caption{The PCF of a spin-polarized system (schematically). Regions I, II,
458: and III are described in the main text.}
459: \label{fig:PCF}
460: \end{figure}
461: %%%%%%%%%%%
462:
463: The dependence of $g_0$ on $x$ that results from this formalism is
464: sketched in Fig.~\ref{fig:PCF}. Near the $x = a$ maximum (region III),
465: $g_0(x)$ is given by [cf.~Eq.~(\ref{eqn:Phi_1_Gauss})]
466: %
467: \begin{equation}
468: g_0(x)=\left( 1/\sqrt{\pi }\,l\right) \exp \left[ -(x-a)^{2}/l^{2}\right] .
469: \label{eqn:g_Gauss}
470: \end{equation}
471: %
472: In the region II the quasiclassical approximation applies. In particular, at
473: $a_{B}\ll x\ll a$ the result for $g_0(x)$ can be written in terms of the
474: tunneling action,
475: %
476: \begin{equation}
477: S_{3}(x)=\frac{1}{\hbar }\int\limits_{x}^{a}dy\left[ 2\mu \Delta U_{\text{tot%
478: }}(y)\right] ^{1/2}, \label{eqn:S_3}
479: \end{equation}
480: %
481: and the appropriate prefactor, as follows:
482: %
483: \begin{align}
484: g_0(x)& =\frac{a}{l^{2}}\left[ \frac{1}{2\pi }\frac{\Omega (a)}{\Omega (x)}%
485: \frac{\hbar \omega _{0}}{U(x)}\right] ^{1/2}e^{\xi (x)-2S_{3}(x)},
486: \label{eqn:g_II} \\
487: \xi (x)& =\int\limits_{x}^{a}dy\left\{ \frac{\omega _{0}+\Omega (a)-\Omega
488: (y)}{\left[ (2/\mu )\Delta U_{\text{tot}}(y)\right] ^{1/2}}-\frac{1}{a-y}%
489: \right\} . \label{eqn:xi}
490: \end{align}
491: %
492: In comparison, the FLA\cite{Matveev-04} amounts to replacing
493: $\Omega(x)$ by a constant in the above equations, thereby effectively
494: ignoring the quantum fluctuations of the dynamical variable $X_2$.
495: Finally, in an ultrathin wire, $\mathcal{L} \gg 1$, there is also region
496: I, $x\lesssim a_{B}$, where the quasiclassical approximation breaks
497: down. Fortunately, at such $x$, Eq.~(\ref{eqn:phi_3}) can be simplified,
498: as $\Omega (x)\simeq \Omega (0)$ and $U_{\text{tot}} (x)\simeq
499: U(x)+2U(3a/2)$. This enables us to express $\phi (x)$ and $g_0(x)$ in
500: terms of the Whittaker functions,\cite{Gradshteyn-Ryzhik} similar to
501: Ref.~\onlinecite{Fogler-05a} (see also Appendix~\ref{sec:U_sigma}).
502: Using such a representation, it is easy to show that
503: %
504: \begin{equation}
505: \phi (0)/\phi ^{\prime }(0)\simeq a_{B}/\mathcal{L},\quad r_{s}\gg 1,
506: \label{eqn:phi_log_derivative}
507: \end{equation}
508: %
509: which, combined with Eq.~(\ref{eqn:J_from_g_triplet}), yields
510: Eq.~(\ref{eqn:J_from_g_ultrathin}). With a bit
511: more algebra, one can match Eq.~(\ref{eqn:g_II}) with another formula in
512: region I, leading finally to Eq.~(\ref{eqn:J}) with $\eta $ and $\kappa $
513: given by
514: %
515: \begin{align}
516: \eta & =\frac{2S_{3}(0)}{\sqrt{2r_{s}}}=2\int\limits_{0}^{a}\frac{dx}{a}%
517: \left[ \frac{\epsilon a}{e^{2}}\,\Delta U_{\text{tot}}(x)\right] ^{1/2},
518: \label{eqn:eta_3} \\
519: \kappa & =\frac{2^{5/4}}{\sqrt{\pi }}\,e^{\xi (0)}\sqrt{\frac{\Omega (a)}
520: {\Omega (0)}}\left[ \frac{\epsilon a^{3}}{e^{2}}\,U_{\text{tot}}^{\prime
521: \prime }(a)\right] ^{3/4}. \label{eqn:kappa_3}
522: \end{align}
523: %
524: Thus, for the $N=3$ case we were able to reduce the original complicated
525: three-body eigenvalue problem to routine operations of solving an
526: \emph{ordinary\/} differential equation~(\ref{eqn:Omega_3}) and taking two
527: quadratures, Eqs.~(\ref{eqn:xi}) and (\ref{eqn:eta_3}). The resultant $\eta$
528: and $\kappa$ are listed in Table~\ref{tbl:Results}. In comparison, the FLA%
529: \cite{Matveev-04} underestimates $\kappa $ by about $50\%$. It gets $\eta$
530: correctly but only for $N=3$, see more below.
531:
532: One more important comment is in order. The antisymmetry of the total
533: fermion wavefunction imposes certain selection rules\cite{Maksym-96} for the
534: allowed values of $L$ [see Eq.~(\ref{eqn:H})] at a given total spin $S$.
535: Thus, the lowest-energy $L$ eigenstates for the two possible spin states of
536: the $N = 3$ system, $S = 1/2$ and $S = 3/2$, are, respectively, $|L| = 1$
537: and $0$. Since $J \ll \hbar^2 / I$ at large $r_s$, the ground-state of the
538: system is the $L = 0$ spin-quartet, in agreement with prior numerical work.%
539: \cite{Reimann-02, Usukura-05}
540:
541: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
542:
543: \section{Case $N>3$}
544: \label{sec:Many}
545:
546: In a system of more than three electrons, the Hamiltonian that governs the
547: important degrees of freedom, $x$ and $\mathbf{X}=(X_{2},\ldots
548: ,X_{N-1})^{\dagger }$, becomes
549: \begin{equation}
550: H_{N}=-\frac{\hbar ^{2}}{2\mu }\partial _{x}^{2}-\frac{\hbar ^{2}}{2}\left(
551: \mathbf{M}^{-1/2}\partial _{\mathbf{X}}\right) ^{\dagger }\left( \mathbf{M}%
552: ^{-1/2}\partial _{\mathbf{X}}\right) +U_{\text{tot}}, \label{eqn:H_N}
553: \end{equation}
554: where
555:
556: \begin{equation}
557: M_{ij}^{-1/2}=\frac{1}{\sqrt{m}}\frac{\delta _{ij}-\left( 1-\sqrt{2/N}%
558: \right) }{N-2}.
559: \end{equation}
560: Again, the potential energy has two minima separated by a large barrier.
561: The single-well function $\Phi _{1}(x,\mathbf{X})$ can be sought in the form
562: \begin{equation}
563: \Phi _{1}=\phi (x)\exp \left[ -\frac{1}{2\hbar }\Delta \mathbf{X}^{\dagger }%
564: \mathbf{M}^{1/2}\bm{\Omega}(x)\mathbf{M}^{1/2}\Delta \mathbf{X}\right] ,
565: \label{eqn:Phi_1}
566: \end{equation}
567: where $\Delta \mathbf{X}=\mathbf{X}-\mathbf{X}^{\ast }$. In the language of
568: the quantum tunneling theory, $\mathbf{X}^{\ast }(x)$ is the instanton
569: trajectory and $\bm{\Omega}(x)$ is a matrix that controls Gaussian
570: fluctuations around the instanton. Our goal is to compute them. The usual
571: route is to parametrize the dependence of $\mathbf{X}^{\ast }$ on $x$ in
572: terms of an ``imaginary-time'' $\tau $, in
573: which case $x(\tau )$ and $\mathbf{X}^{\ast }(\tau )$ must minimize the
574: action functional
575: \begin{equation}
576: S_{N}=\int\limits_{0}^{\infty }\frac{d\tau }{\hbar }\left[ \frac{\mu }{2}%
577: (\partial _{\tau }x)^{2}+\frac{1}{2}(\partial _{\tau }\mathbf{X})^{\dagger }%
578: \mathbf{M}\,\partial _{\tau }\mathbf{X}+\Delta U_{\text{tot}}\ \right]
579: \label{eqn:S_N}
580: \end{equation}
581: subject to the boundary conditions $x(0)=0$, $x(\infty )=a$, and $\mathbf{X}
582: (\infty )=0$. Henceforth $U_{\text{tot}}$ is always meant to be evaluated on
583: the instanton trajectory and $\Delta U_{\text{tot}}$ stands for the
584: difference of its values at a given $\tau $ and at $\tau =\infty $.
585: Repeating the steps of the derivation for the $N=3$ case, we find that
586: Eq.~(\ref{eqn:Omega_3}) for $\bm{\Omega}$ is still valid once we define
587: $\bm{\omega}$ to be a positive-definite matrix such that $\bm{\omega}^{2}=
588: \mathbf{M}^{-1/2}\bm{\Xi}\mathbf{M}^{-1/2}$, where $\bm{\Xi}$ is the matrix
589: of the second derivatives $\Xi _{ij}=\partial _{X_{i}}\partial _{X_{j}}
590: U_{\text{tot}}$. This equation has an equivalent but more elegant form in terms
591: of $\tau $:
592: \begin{equation}
593: \partial _{\tau }\bm{\Omega}=\bm{\Omega}^{2}(\tau )-\bm{\omega}^{2}(\tau ).
594: \label{eqn:Omega}
595: \end{equation}
596: Also, for practical calculations, it is convenient to take advantage of a
597: formula
598: \begin{equation}
599: \text{tr}\,\left( \mathbf{M}^{-1}\bm{\Xi}\right) +\frac{1}{\mu }
600: \frac{\partial ^{2}}{\partial x^{2}}U_{\text{tot}}\left( x\right) =
601: \frac{1}{2\mu }\sum\limits_{\substack{ i,j=0 \\ i\neq j}}^{N-1}
602: U^{\prime \prime }\left( x_{i}-x_{j}\right) .
603: \end{equation}
604:
605: As for Eqs.~(\ref{eqn:phi_3}) and (\ref{eqn:xi}), they require only the
606: replacement $\Omega \rightarrow \text{tr}\,\bm{\Omega}$,
607: \begin{equation}
608: \left\{ \frac{\hbar ^{2}}{2\mu }\partial _{x}^{2}-U_{\text{tot}}(x) -
609: \frac{\hbar }{2}\,\text{tr}\,\bm{\Omega}(x) + E\right\} \phi (x)=0,
610: \end{equation}
611: while in Eq.~(\ref{eqn:kappa_3}) one has to replace $\Omega $ with $\det
612: \bm{\Omega}$.
613:
614: Finally, instead of Eq.~(\ref{eqn:eta_3}) we have $\eta =2S_{N}/\sqrt{2r_{s}}
615: $, consistent with the notion that the main exponential dependence of $J$ is
616: always determined by the tunneling action.
617:
618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
619:
620: \section{Calculation of the instanton}
621: \label{sec:Instanton}
622:
623: A few properties of the instanton follow from general considerations. The
624: dimensional analysis of Eq.~(\ref{eqn:S_N}) yields $S_{N}\propto \sqrt{r_{s}}
625: $, so that $\eta $ is indeed just a constant. Also, from the symmetry of the
626: problem, $X_{N+1-j}(\tau )=-X_{j}(\tau )$. Thus, in the special case of
627: $N=3$, the instanton trajectory is trivial: $X_{2}\equiv 0$, i.e., the $j=2$
628: electron does not move. This is why we were able to compute $S_{3}$ in a
629: closed form, Eq.~(\ref{eqn:S_3}). For $N>3$ the situation is quite
630: different: all electrons [except $j=(N+1)/2$ for odd $N$] do move. In order
631: to investigate how important the motion of electrons distant from the
632: $j=0,1$-pair is let us consider the $N=\infty $ (\emph{quantum wire\/}) case,
633: where the far-field effects are the largest. If $X_{j}$'s were small, we
634: could expand $\Delta U_{\text{tot}}$ in Eq.~(\ref{eqn:S_N}) to the second order
635: in $X_{j}$ to obtain the harmonic action
636: \begin{equation}
637: S_{h}=\frac{1}{2}\frac{m}{\hbar }\int \frac{dk}{2\pi }\int \frac{d\omega }
638: {2\pi }\left\vert u_{k\omega }\right\vert ^{2}\left[ \omega ^{2}+\omega
639: _{p}^{2}(k)\right] , \label{eqn:action_harmonic}
640: \end{equation}
641: where $u_{k\omega }$ is the Fourier transform of $u_{j}(\tau )\equiv
642: x_{j}-x_{j}^{0}$, electron displacement from the classical equilibrium
643: position $x_{j}^{0}\equiv (j-1/2)a$, $j\in \mathbb{Z}$, and
644: %
645: \begin{equation}
646: \omega_{p}(k)\simeq s_{0} k \ln^{1/2}\left(\frac{4.15}{ka}\right),
647: \quad s_{0} \equiv \sqrt{\frac{e^2}{\epsilon \mu a}}
648: \end{equation}
649: %
650: is the plasmon dispersion in the 1D WC (the logarithmic term is due to
651: the long-range nature of the Coulomb interaction). Minimization of
652: $S_{h}$ with the specified boundary conditions yields
653: %
654: \begin{equation}
655: u_{j}(\tau ) \propto \frac{v x_j^0}{(x_j^0)^2 + v^2 \tau^2},\quad
656: v \simeq \frac{s_0}{2}
657: \ln \frac{(x_{j}^{0})^2 + s_0^2 \tau^2}{a^2}.
658: \end{equation}
659: %
660: Substituting this formula into Eq.~(\ref{eqn:action_harmonic}), we
661: find that the contributions of distant electrons to $S_{h}$ rapidly
662: decay with $|j|$. Since $u_{j}(\tau )$'s are small at large $j$ and
663: $\tau $ we expect that these $j$- and $\tau $-dependencies are rendered
664: correctly by the harmonic approximation. Thus, a fast convergence of
665: $\eta $ to its thermodynamic limit is expected as $N$ increases.
666:
667: Encouraged by this conclusion, we undertook a direct numerical minimization
668: of $S$ for the set of $N$ listed in Table~\ref{tbl:Results} using standard
669: algorithms of a popular software package \textsc{MATLAB}. The optimal
670: trajectories that we found for the case $N = 8$ are shown in
671: Fig.~\ref{fig:Instanton}. In agreement with our earlier statement, $u_j(0)$
672: reach a finite fraction of $a$. This collective electron motion lowers the
673: effective tunneling barrier and causes $\eta$ to drop below its FLA value,
674: although only by $0.7\%$, see Table~\ref{tbl:Results}. Further decrease of
675: $\eta$ as $N$ increases past $8$ is beyond the accuracy of the employed
676: minimization procedure.
677:
678: Let us now discuss the prefactor $\kappa$. In the inset of
679: Fig.~\ref{fig:Instanton} we plot $\text{tr}\,\bm{\Omega}(x)$ computed by
680: solving Eq.~(\ref{eqn:Omega}) numerically. To reduce the calculational
681: burden, we set $\mathbf{X}^\ast(\tau) \to 0$ instead of using the true
682: instanton trajectory. The error in $\kappa$ incurred thereby is $\sim
683: 2\%$ (see Sec.~\ref{sec:integrable_model}). In comparison, the FLA, where
684: $\text{tr}\,\bm{\Omega}(x) = \text{const}$, yields $\kappa$ about $50\%$
685: smaller than the correct result, similar to $N = 3$.
686:
687: %%%%%%%%%%%
688: % FIG. 3
689: %
690: \begin{figure}
691: \begin{center}
692: \includegraphics[height=1.75in]{wcx_3.eps}
693: \end{center}
694: \caption{(Color online) The instanton trajectories of $1\leq j\leq 4$
695: electrons in the $N = 8$ Wigner molecule on a ring. Inset: $\text{tr}\,
696: \bm{\Omega}(x)$. The units of $\protect\tau$ and $\bm{\Omega}$ are $\protect%
697: \sqrt{2} a / s_0$ and its inverse.}
698: \label{fig:Instanton}
699: \end{figure}
700: %%%%%%%%%%%
701:
702: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
703:
704: \section{Alternative derivation of the exchange constant}
705: \label{sec:Alternative}
706:
707: The relation~(\ref{eqn:J_from_g_ultrathin}) between the exchange
708: constant $J$ and the PCF of the spinless system --- which is one of our
709: main results --- was derived in Sec.~\ref{sec:Three} assuming $r_s \gg
710: 1$. In this section we show that in ultrathin wires, $\mathcal{L} \equiv
711: \ln (a_B / R) \gg 1$, this relation holds under a more general
712: condition: with an accuracy $O(1 / \mathcal{L})$, it remains valid as
713: long as the \emph{strong\/} spin-charge separation persists, $J \ll E_F
714: \sim \hbar^2 / m a^2$. In particular, it applies at $r_s \sim 1$, i.e.,
715: in a parametric regime which has been notoriously difficult for a
716: controlled theoretical analysis. Therefore, we think that
717: Eq.~(\ref{eqn:J_from_g_ultrathin}) amounts to some definite progress.
718: The fundamental reason for the broad domain of validity of
719: Eq.~(\ref{eqn:J_from_g_ultrathin}) is that in 1D the $1/r$ fall-off of
720: the Coulomb potential~(\ref{eqn:U}) constitutes a sufficiently rapid
721: decay. In fact, the relations of type of
722: Eq.~(\ref{eqn:J_from_g_ultrathin}) are generic for 1D models in which
723: electrons interact via the potential $U(r) \sim 1 / r^\alpha$ with
724: $\alpha \geq 1$. Known to us physical realizations of $\alpha > 1$ cases
725: include (i) $\alpha = 3$, which occurs when the bare Coulomb interaction
726: is screened by a nearby metallic plane,\cite{Fogler-05a} and (ii)
727: $\alpha = 2$, where the screening is accomplished by a half-plane, see
728: Sec.~\ref{sec:integrable_model} below.
729:
730: In order to derive Eq.~(\ref{eqn:J_from_g_ultrathin}) for the unscreened
731: Coulomb interaction ($\alpha = 1$) we make a connection with a previous
732: work of one of us\cite{Fogler-05a} where it was shown that in ultrathin
733: wires an unusual correlated regime --- Coulomb Tonks gas (CTG) ---
734: exists in the window $1 / \mathcal{L} \ll r_s \ll 1$. In the CTG, unlike
735: in a WC, the long-range tails of the Coulomb potential have no important
736: effect on the spin degree of freedom. To the leading order in $r_s$, the
737: Coulomb interaction~(\ref{eqn:U}) acts simply as a strong short-range
738: repulsion. This allows one to characterize its effect solely in terms of
739: the transmission coefficient $t(q)$ for a two-electron collision with
740: the relative momentum $q$. Based on this observation, the following
741: formula for $J$ was derived\cite{Fogler-05a}
742: %
743: \begin{equation}
744: J = \frac{\hbar^2}{m} \int \frac{d q}{2 \pi} \tilde{g}_0(q) q\,
745: \text{Im}\,t(q),\quad J \ll E_F.
746: \label{eqn:J_CTG_from_t}
747: \end{equation}
748: %
749: To see how it leads to Eq.~(\ref{eqn:J_from_g_ultrathin}) we
750: first note that for the potential~(\ref{eqn:U})
751: the transmission coefficient has the form\cite{Fogler-05a}
752: %
753: \begin{equation}
754: t(q) = \frac{i q}{i q - (m / \hbar^2) c(q)},
755: \quad q \gg \frac{1}{a_B},
756: \label{eqn:t_q}
757: \end{equation}
758: %
759: where $c(q) = 2 (e^2 / \epsilon) \ln \left(1 / 2 q R\right)$. The slow
760: (in this case, logarithmic) dependence of $c$ on $q$ is crucial here and
761: it is precisely due to the aforementioned fast decay of $U(r)$ with
762: distance. As explained in Ref.~\onlinecite{Fogler-05a}, in the CTG
763: regime the integral in Eq.~(\ref{eqn:J_CTG_from_t}) is dominated by $q
764: \sim 1 / a$; therefore, to the order $O(1 / \mathcal{L})$ one can
765: replace $c(q)$ by $c(1 / a)$ and then by $2 (e^2 / \epsilon)
766: \mathcal{L}$. Finally, it is permissible to neglect $i q$ in the
767: denominator of Eq.~(\ref{eqn:t_q}), thereupon
768: Eq.~(\ref{eqn:J_CTG_from_t}) becomes identical to
769: Eq.~(\ref{eqn:J_from_g_ultrathin}).
770:
771: Our next task is to show that Eq.~(\ref{eqn:J_from_g_ultrathin}) holds
772: not only in the two limiting case (CTG and WC) but also everywhere in
773: between. We do so with the help of a formalism that can be considered a
774: continuum analog of the Schrieffer-Wolfe transformation in the Hubbard
775: model, in particular, in its 1D version.\cite{Ogata-90}
776:
777: Let $j\in \mathbb{Z}$ label electrons in the order of increasing
778: instantaneous coordinates $x_{j}$. Then, by construction, the coupling
779: constant $J$ refers to the exchanges of nearest-neighbors, say, $j = 0$
780: and $j = 1$. This coupling is suppressed because in order to exchange,
781: the two electrons have to approach each other very closely and at that
782: moment their mutual Coulomb repulsion $U(x)$ is very strong. Let $a_c$
783: be a suitable short-range cutoff. Since we are interested
784: in the physics at the energy scale $J$, it should be possible in
785: principle to ``integrate out'' the processes at the much higher energy
786: scale $U(a_c)$. In doing so one would trade the bare Hamiltonian $H$ for
787: a renormalized one, $H + H_\sigma$, which however yields the same value
788: of the effective coupling $J$. Traditionally, this renormalization
789: program is accomplished by incrementally increasing $a_c$ from zero to
790: $a$, the natural upper limit. In particular, the
791: Hamiltonian~(\ref{eqn:H}) can be viewed as the final result of such a
792: renormalization. However, it is fully legitimate to stop the process
793: while $a_c$ is still much smaller than $a$. At this stage the charge
794: correlations at scales shorter than $a_c$ can not longer be studied
795: because the corresponding degrees of freedom are already integrated out.
796: This implies that the form of $H_\sigma$ is not unique: $H_\sigma$ may
797: differ in their structure on scales $x < a_c$ as long as the resultant
798: $J$ is unchanged. Therefore, it should be possible to choose $H_\sigma$
799: in the form
800: %
801: \begin{equation}
802: H_{\sigma} = U_{\sigma }(x)(\mathbf{S}_{0}\mathbf{S}_{1}-1/4)
803: + \Delta c \delta(x).
804: \label{eqn:H_sigma}
805: \end{equation}
806: %
807: It is especially convenient to take the limit $\Delta c \rightarrow
808: \infty$, in which electrons behave as impenetrable particles. Although
809: thus renormalized Hamiltonian forbids the pair-exchange in the real
810: space, if the choose $U_\sigma(x)$ correctly, the spin exchange rate $J$,
811: which is the only observable manifestation of an interchange of
812: identical particles, would be preserved. A familiar example of such a
813: $H_\sigma$ on a lattice is the added term
814: %
815: \begin{equation}
816: \frac{4 t^2}{U} \left(\mathbf{S}_{i}\mathbf{S}_{i+1}
817: - \frac14 n_{i}n_{i+1}\right)
818: \label{eqn:H_sigma_Hubbard}
819: \end{equation}
820: %
821: complemented with the no-double-occupancy constraint (i.e., $U \to U +
822: \Delta U = \infty$) in the theory of a large-$U$ 1D Hubbard
823: model.\cite{Ogata-90}
824:
825: The choice of $U_\sigma(x)$ in Eq.~(\ref{eqn:H_sigma}) is again not
826: unique. To readily get the result we want, we require
827: that (i) $U_{\sigma }(x)$ decays exponentially with $x$ and (ii)
828: its effect on any wavefunction $\Phi(x)$ that vanishes at $x = 0$ is
829: adequately described by the first-order perturbation theory. Under
830: such conditions the spin dynamics is captured
831: correctly to the leading order in $1 / \mathcal{L}$ as long as
832: $U_\sigma(x)$ satisfies the constraint
833: %
834: \begin{equation}
835: I_2 \equiv \int\limits_{0}^{\infty} d x\, x^2 U_\sigma(x)
836: = \frac{\hbar^2}{m} \frac{a_{B}}{\mathcal{L}}.
837: \label{eqn:U_sigma}
838: \end{equation}
839: %
840: The proof is relegated to Appendix~\ref{sec:U_sigma}. From this point
841: the derivation of Eq.~(\ref{eqn:J_from_g_ultrathin}) is straightforward.
842: After the renormalization, the infinitely strong $\delta$-function
843: included in $H_\sigma$ causes the orbital wavefunction of the
844: system to vanish at $x = 0$, so that the perturbation theory in
845: $U_\sigma$ now applies. To the leading order we can compute the
846: sought exchange constant $J$ by averaging $U_{\sigma}$ over
847: such a wavefunction or equivalently over the renormalized PCF
848: $g_{\text{ren}}(x)$:
849: %
850: \begin{equation}
851: J = \frac12 \int_{-\infty}^{\infty }U_{\sigma }(x)g_{\text{ren}}(x)dx.
852: \label{eqn:J_from_g_ren}
853: \end{equation}
854: %
855: (The factor $1/2$ accounts for the two-body nature of the interaction.)
856: Let us discuss $g_{\text{ren}}(x)$ in more detail. First of all, under the
857: conditions of a strong spin-charge separation the PCF a spinful ($g$)
858: and spinless ($g_0$) systems are always very similar. Before the
859: renormalization (for the original Hamiltonian) both functions are peaked
860: at $x \sim a$ and rapidly decrease as $x \to 0$, albeit only $g_0(x)$
861: vanishes at $x = 0$, while $g(0)$ is simply very small. Once $H_\sigma$
862: is added, $g(x)$ gets replaced by the function $g_{\text{ren}}(x)$, which
863: vanishes at $x = 0$ and thus is nearly identical to the original $g_0$
864: at \emph{all\/} $x$. The PCF $g_0(x)$ of the spinless system does not
865: change appreciably. Since $U_{\sigma }(x)$ is short-range, the integral
866: in Eq.~(\ref{eqn:J_from_g_ren}) is dominated by small $x$. Therefore, we
867: can substitute the expansion $g_{\text{ren}}(x) \simeq g_{0}(x) \simeq
868: g_{0}^{\prime \prime }(0)x^{2}/2$ in place of the full $g_{\text{ren}}(x)$.
869: Comparing the result with Eq.~(\ref{eqn:U_sigma}), we arrive at
870: Eq.~(\ref{eqn:J_from_g_ultrathin}).
871:
872: This alternative route to the derivation of
873: Eq.~(\ref{eqn:J_from_g_ultrathin}) has the advantage of being valid in a
874: broader range of $r_s$ compared to Herring's method\cite{Herring-64}
875: originally designed for the quasiclassical case $r_s \gg 1$. It may also
876: be more intuitive because the Hubbard-model
877: expression~(\ref{eqn:H_sigma_Hubbard}) is widely known. Finally, it can
878: be easily generalized to other interaction types, as discussed in the
879: next section.
880:
881: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
882:
883: \section{An integrable model with a screened Coulomb interaction}
884: \label{sec:integrable_model}
885:
886: Our instanton calculation in Sec.~\ref{sec:Many} is expected to be
887: asymptotically exact in the limit of large $r_s$. However it is a good
888: practice to verify whenever possible that any given calculation is free
889: of inadvertent arithmetic or coding mistakes. With this in mind we
890: applied our method also to the interaction law
891: %
892: \begin{equation}
893: U(r) = \frac{\hbar^2}{m} \frac{\lambda (\lambda -1)}{(r + R)^2},
894: \label{eqn:U_CSM}
895: \end{equation}
896: %
897: which is similar to the celebrated Calogero-Sutherland-Moser (CSM)
898: model,\cite{Sutherland-71} except for the short-range cutoff $R$ in
899: $U(r)$. Without this cutoff (i.e., for $R = 0$) the potential $U$ is
900: impenetrable, so that $J = 0$. With a finite cutoff a weak tunneling
901: through the potential barrier $U(r)$ is possible, which imparts the
902: system with a spin dynamics and has interesting experimental
903: implications (see Appendix~\ref{sec:CSM_law}). But to finish with
904: the theoretical part,
905: let us first focus on the original CSM model, $R = 0$. This is a good
906: test case because, on the one hand, a number of exact analytical results
907: are available here. On the other hand, the model also possesses a WC
908: regime, at $\lambda \gg 1$, so that the corresponding PCF $g_0(x)$
909: should be calculable by our method. According to the exact
910: results,\cite{Sutherland-71, Forrester-92} the PCF behaves as
911: %
912: \begin{equation}
913: g_0(x)\simeq \frac{\varkappa \sqrt{\lambda}}{a}
914: \left(C \frac{x}{a}\right)^{2\lambda },
915: \quad x\rightarrow 0.
916: \label{eqn:g_CSM_exact}
917: \end{equation}
918: %
919: For the three-electron molecule on a ring the coefficients $C$ and
920: $\varkappa$ can be computed in a straightforward manner directly
921: from the exact\cite{Sutherland-71} three-body wavefunction:
922: %
923: \begin{equation}
924: N = 3,\:\: \lambda \gg 1:\quad C = \frac{8\pi}{9 \sqrt{3}},\quad
925: \varkappa = \sqrt{\frac{8\pi}{27}}.
926: \label{eqn:C_3}
927: \end{equation}
928: %
929: For $N = \infty$ the derivation is much more
930: involved, but the result\cite{Forrester-92, Ha-94} is also known:
931: %
932: \begin{equation}
933: \frac{C}{2 \pi \lambda} = \left[
934: \frac{\Gamma^3(\lambda + 1) \sqrt{{3} / {\pi \lambda}}}
935: {\Gamma(2 \lambda + 1) \Gamma(3 \lambda + 1)}\right]^{1 / (2 \lambda)},
936: \quad \varkappa =\sqrt{\frac{\pi}{3}},
937: \label{eqn:C}
938: \end{equation}
939: %
940: where $\Gamma(z)$ is Euler gamma-function.\cite{Gradshteyn-Ryzhik}
941: In the WC limit the first formula reduces to
942: %
943: \begin{equation}
944: C = \frac{e\pi}{3 \sqrt{3}},\quad \lambda \gg 1.
945: \label{eqn:C_infty}
946: \end{equation}
947: %
948: The calculation of $g_0(x)$ by the instanton method
949: is virtually the same as in Sec.~\ref{sec:Instanton} except for two
950: minor changes. First, the quasiclassical approximation is valid down to
951: $x=0$, i.e., there is no region I in Fig.~\ref{fig:PCF}. Second, to
952: handle the logarithmic divergence of the tunneling action, inherent to
953: the CSM model, one has to impose a different boundary condition for the
954: instanton, $x(\tau =0)=z$, which makes $S_{N}$ a function of $z$
955: [similar to $S_{3}$ being a function of $x$ in Eq.~(\ref{eqn:S_3})]. At
956: small $z$ it behaves as
957: %
958: \begin{equation}
959: S_{N}\simeq \sqrt{\lambda (\lambda -1)}\,[\ln (a/z)-\Delta S].
960: \label{eqn:Delta_S}
961: \end{equation}
962: %
963: Replacing $\sqrt{\lambda (\lambda -1)}$ by $\lambda -1/2$, we find [cf.
964: Eqs.~(\ref{eqn:eta_3}) and (\ref{eqn:kappa_3})]
965: %
966: \begin{equation}
967: C=e^{\Delta S},\quad \varkappa =\frac{e^{\xi (0)-\Delta S}}{\sqrt{\pi }%
968: \left( l\sqrt{\lambda }/a\right) ^{3}}\sqrt{\frac{\det \mathbf{\Omega }(a)}
969: {\det \mathbf{\Omega }(0)}} \label{eqn:C_varkappa_CSM}
970: \end{equation}
971: %
972: (it is easy to see that $\varkappa $ is a $\lambda $-independent constant).
973:
974: Remarkably, for $N=3$ the solution of Eq.~(\ref{eqn:Omega_3}) and all the
975: necessary quadratures can be done analytically. Using the expression for
976: total potential energy
977: %
978: \begin{align}
979: U_{\text{tot}}\left( x\right) &=
980: \frac{\pi^2}{9} \frac{\hbar^2}{m} \frac{\lambda (\lambda - 1)}{a^2}
981: \notag\\
982: &\times \left[ \frac{1}{\sin^2 \left( \pi x/3a\right)}
983: + \frac{2}{\cos^2 \left( \pi x/6a\right)} - 4\right],
984: \end{align}
985: %
986: we derive (for $\lambda \gg 1$)
987: %
988: \begin{equation}
989: \begin{split}
990: \phi \left( x\right) & =\left[ 4\sin \left( \frac{\pi x}{3a}\right) \left(
991: 1+\cos \frac{\pi x}{3a}\right) \right] ^{\lambda }, \\
992: \Omega \left( x\right) & = \frac{\pi^2}{3} \frac{\hbar }{m a^2}
993: \frac{\lambda }{\cos^2 \left( \pi x/6a\right)},
994: \end{split}
995: \end{equation}
996: %
997: which allows one to recover the exact result Eq.~(\ref{eqn:C_3}).
998:
999: To check if we can also reproduce Eq.~(\ref{eqn:C_infty}) for $N =
1000: \infty$, we calculated $\Delta S$ and $\varkappa $ numerically for
1001: $4\leq N\leq 10$ and fitted them to cubic polynomials in $1/N$.
1002: Extrapolating the fits to $N=\infty $, we obtained $\exp (\Delta
1003: S)=1.6438(4)$ and $\varkappa =1.04$ compared to the exact values
1004: $1.6434$ and $1.02$, respectively. We attribute the $2\%$ discrepancy in
1005: $\varkappa $ to our choice not to use the true instanton trajectory in
1006: Eq.~(\ref{eqn:Omega}). Apparently, our method has successfully passed
1007: this test; thus, our results for the unscreened Coulomb interaction
1008: (Table~\ref{tbl:Results}) should also be reliable.
1009:
1010: Let us now consider the case with a small but non-zero cutoff, $0 < R
1011: \ll a$. As explained in Appendix~\ref{sec:CSM_law}, this model can be
1012: relevant for the existing experimental setup of \textcite{Auslaender-05}
1013: provided the electron density can be made
1014: low enough. Therefore, the calculation of $J$ has an independent
1015: significance. This calculation is again similar to that for the case
1016: of the bare Coulomb interaction, Sec.~\ref{sec:Alternative}, except that
1017: instead of Eq.~(\ref{eqn:U_sigma}) the constraint on the effective
1018: potential becomes
1019: %
1020: \begin{equation}
1021: I_{2 \lambda} \equiv \int_{0}^{\infty }dx\, x^{2\lambda}
1022: U_{\sigma }\left( x\right) = \frac{\pi}{\lambda -1} \frac{\hbar^2}{m}
1023: R^{2\lambda - 1}.
1024: \label{eqn:U_sigma_CSM}
1025: \end{equation}
1026: %
1027: (For derivation, see Appendix~\ref{sec:U_sigma}.) Consequently,
1028: Eq.~(\ref{eqn:J_from_g_ultrathin}) is replaced with
1029: %
1030: \begin{equation}
1031: J = I_{2 \lambda} a_c^{-2\lambda} g_0(a_c),\quad R \ll a_c \ll a.
1032: \label{eqn:J_from_g_CSM}
1033: \end{equation}
1034: %
1035: Since $R$ is parametrically smaller than $a$, it is permissible to
1036: substitute Eqs.~(\ref{eqn:g_CSM_exact}) and (\ref{eqn:C}) into the
1037: last formula, which yields
1038: %
1039: \begin{align}
1040: J &= \gamma \frac{\hbar^2}{m a^2}
1041: \left(\frac{R}{a}\right)^{2\lambda - 1},
1042: \label{eqn:J_CSM}\\
1043: \gamma &= \frac{1}{2 \lambda (\lambda - 1)}
1044: \frac{(2 \pi \lambda)^{2\lambda + 1} \Gamma^3(\lambda + 1)}
1045: {\Gamma(2 \lambda + 1) \Gamma(3 \lambda + 1)}.
1046: \label{eqn:gamma}
1047: \end{align}
1048: %
1049: From Eq.~(\ref{eqn:J_CSM}) one concludes that the strong spin-charge
1050: separation in the model in hand may arise for an arbitrary $\lambda > 1$
1051: provided ${R} / {a} \ll \min\{1,\,\lambda - 1\}$. In the experimental
1052: setup of \textcite{Auslaender-05}, where this
1053: model can be realized physically, we estimate $R \approx 0.7 a_B$ and
1054: $\lambda = 1.5$--$2$, see Appendix~\ref{sec:CSM_law}, so that $\gamma =
1055: 36$--$40$. Our formula predicts that as the electron density is reduced,
1056: a fairly rapid \emph{algebraic\/} fall-off Eq.~(\ref{eqn:J_CSM}) of $J$
1057: should become observable, provided the disorder effects do not
1058: intervene. The crossover from high to low-density behavior is discussed
1059: in more detail in the next section where we give arguments that the
1060: strong spin-charge separation sets in already at rather modest $r_s$.
1061:
1062: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1063:
1064: \section{Validity domain of the obtained formulas and numerical
1065: estimates}
1066: \label{sec:Domain}
1067: \subsection{Intermediate densities}
1068:
1069: Returning to the case of the unscreened Coulomb interaction, an
1070: important practical question is the relevance of the obtained formulas
1071: for a more common experimental situation of $r_s \sim 1$. Although
1072: Eq.~(\ref{eqn:J_from_g_ultrathin}) reduces this question to an easier
1073: problem of calculating the PCF of spinless fermions, at such $r_s$ it
1074: still has to be done numerically. Deferring this effort for a possible
1075: future work, we attempted to see whether one can estimate $J$ at $r_s
1076: \sim 1$ by a simple interpolation between the high-density limit
1077: $r_{s}\ll 1$, corresponding to the CTG,\cite{Fogler-05a} and the
1078: low-density WC limit, $r_s \gg 1$, studied in this work.
1079:
1080: Using Eq.~(\ref{eqn:J_CTG_from_t}) with $\tilde{g}_0(q)$ computed to the
1081: linear order in $r_s$ (per Ref.~\onlinecite{Fogler-05a}) we obtained
1082: for the CTG
1083: %%
1084: %\begin{equation}
1085: %\tilde{g}_0(q) = 2 \pi n \delta(q)
1086: % - \theta(2 \pi n - q) \frac{2 \pi n - q}{2 \pi n}
1087: % + O(r_s),
1088: %\label{eqn:g_CTG}
1089: %\end{equation}
1090: %%
1091: %where $n = {1}/{a}$ is the electron density and $\theta(z)$ is the
1092: %step-function.
1093: %
1094: \begin{equation}
1095: J = \frac{e^2}{\epsilon a_B} \frac{\pi^2}{24 r_s^3 \mathcal{L}}
1096: \left[
1097: %1 - 2 r_s \left(\frac{1}{\pi^2}+ \frac{1}{3}\right)
1098: 1 - \frac{2 (3 + \pi^2)}{3 \pi^2} r_s
1099: \right] + O\left(\frac{1}{\mathcal{L}^2}\right),
1100: \label{eqn:J_CTG}
1101: \end{equation}
1102: %
1103: which is valid for ${1}/{\mathcal{L}} \ll r_s \ll 1$. We plotted this
1104: dependence in Fig.~\ref{fig:J} together with the prediction of
1105: Eq.~(\ref{eqn:J}) for the infinite-wire WC. As one can see, a smooth
1106: interpolation between the two is entirely possible. In fact, the curves
1107: match almost seamlessly around $r_s \sim 0.9$ where a tiny gap has been
1108: left on purpose to separate them from each other. However, we must warn
1109: the reader that at such $r_s$ the high-density approximation of
1110: Eq.~(\ref{eqn:J_CTG}) certainly needs higher-order corrections (taken
1111: literally, it would give negative $J$ at $r_s > 1.15$). Similarly,
1112: Eq.~(\ref{eqn:J}) is likely to have its own non-negligible corrections
1113: at such $r_s$. Thus, one should regard Fig.~\ref{fig:J} as the
1114: indication that the crossover from the CTG to the WC occurs at $r_s =
1115: 1$--$2$. To put it another way, we expect that our Eq.~(\ref{eqn:J})
1116: becomes quantitatively accurate at $r_s \gtrsim 2$. This conclusion
1117: seems to be in agreement with the numerical result of
1118: Ref.~\onlinecite{Egger-99} that the 1D WC is well-formed already at $r_s
1119: > 4$. To get an idea of the size of $J$, we can use parameters $r_s =
1120: 4$, $a_B = 1.5$~nm, and $\epsilon = 1$, which are relevant for the
1121: experiment of \textcite{Jarillo-Herrero-04} For this set of numbers
1122: Eq.~(\ref{eqn:J}) yields
1123: %
1124: \begin{equation}
1125: J \approx 1\ \text{K}.
1126: \end{equation}
1127: %
1128: Unfortunately, the lowest temperature in that experiment was $0.3$~K, so
1129: the exchange correlations may have been strongly affected. We leave the
1130: investigation of the finite-temperature effects for a future study. We
1131: also hope that lower temperatures can be achieved in the next round of
1132: experiments, so that we would be in a better position to check our
1133: predictions.
1134:
1135: %%%%%%%%%%%
1136: % FIG. 4
1137: %
1138: \begin{figure}
1139: \begin{center}
1140: \includegraphics[height=1.75in]{wcx_4.eps}
1141: \end{center}
1142: \caption{Exchange constant in units of $e^2/\epsilon a_{B}$ as a function
1143: of $r_s$ for $a_{B}/R = 100$. The left curve is computed according to
1144: Eq.~(\ref{eqn:J_CTG}), the right one is per Eq.~(\ref{eqn:J}).}
1145: \label{fig:J}
1146: \end{figure}
1147: %%%%%%%%%%%
1148:
1149: \subsection{Thicker wires}
1150:
1151: Another question that we wish to briefly address is the behavior of the
1152: exchange constant in a wire of radius $R \gtrsim a_B$. This parametric
1153: regime is more common than that of ultrathin wires, $R \ll a_B$, and so
1154: it needs to be investigated.
1155:
1156: In regards to the WC limit, the correction to parameter $\eta$ of
1157: Eq.~(\ref{eqn:J}) due to a finite $R$ has been studied by
1158: \textcite{Klironomos-05} These authors pointed out that in the wires
1159: with radius $R\gtrsim a_{B}$ the problem of spin exchange cannot be
1160: treated as a purely 1D one from the outset. They found that the
1161: tunneling trajectories of the primary electron pair avoid the head-on
1162: collision by deviating off the wire center in the transverse direction
1163: by a certain amount $r_0$ each. For the parabolic confinement potential
1164: $U_{\text{con}}(r_\perp) = (\hbar^2 / 2 m R^2) (r_\perp / R)^2$ it is
1165: easy to find the estimate $r_0 \sim (R^4 / a_B)^{1/3}$ by balancing the
1166: Coulomb repulsion and the lateral confinement forces at the point of the
1167: closest approach. As long as $r_0 \ll a$, it remains possible to modify
1168: the interaction potential $U(x)$ to take this effect into account. [At
1169: the simplest level, it is sufficient to replace $R$ by $2 r_0$ in
1170: Eq.~(\ref{eqn:U}).] Thereafter, one can still treat the problem as
1171: 1D.\footnote{
1172: %
1173: Significant changes in the calculation would be required only in very
1174: thick wires with radius $R > 11\,a_{B}$.\cite{Klironomos-05} However, in
1175: such wires the 1D WC limit can be achieved only at $r_s \gtrsim 20$
1176: where the exchange coupling $J$ would be unmeasurably small, see
1177: Eq.~(\ref{eqn:J}). At smaller $r_s$, other physics is likely to come
1178: into play, e.g., the 1D WC may undergo a transition into a multi-row
1179: WC\cite{Piacente-04, Klironomos-xxx} or into a liquid with multiple
1180: subband occupation.
1181: %
1182: }
1183:
1184: In thick wires the relationship between the PCF and the exchange constant
1185: would differ from Eq.~(\ref{eqn:J_from_g_ultrathin}). Of
1186: primary importance here is the small-distance behavior of $U(x)$. When
1187: $r_0 \gg a_B$, the interaction potential becomes so smooth near $x = 0$
1188: that it can be approximated by a constant $U(0) \sim e^2 / 2 \epsilon
1189: r_0$. Following the method outlined in Appendix~\ref{sec:U_sigma}, one
1190: arrives at the relation
1191: %
1192: \begin{equation}
1193: J = \frac{e^2 a_B}{2 \epsilon} a_0 g_0^{\prime \prime }(x),
1194: \quad a_0 = \frac{\hbar}{\sqrt{m U(0)}} \sim (R^2 a_B)^{1/3}.
1195: \label{eqn:J_smooth}
1196: \end{equation}
1197: %
1198: We see that the factor $\mathcal{L}$ in
1199: Eq.~(\ref{eqn:J_from_g_ultrathin}) gets replaced with
1200: $a_B / a_0 \ll 1$.\footnote{
1201: %
1202: It is interesting that in the opposite limit of a short-range $U(x)$,
1203: e.g., $U(x) = c \delta(x)$, the relation between $J$ and $g_0$ is
1204: formally similar, with $a_0 = 2 \hbar^2 / m c$, see
1205: Eqs.~(\ref{eqn:J_CTG_from_t}) and (\ref{eqn:t_q}).
1206: %
1207: }
1208:
1209: In the case of intermediate-width wires, $R \sim a_B$, one can use $a_0
1210: = a_B$ for order-of-magnitude estimates; however, the exact value of the
1211: coefficient $a_0$ and therefore $\kappa$ in Eq.~(\ref{eqn:J}) would
1212: depend on the details of the confinement potential
1213: $U_{\text{con}}(\textbf{r}_\perp)$.
1214:
1215: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1216:
1217: \section{Conclusions}
1218: \label{section:Conclusions}
1219:
1220: In this paper we have developed a method for a controlled calculation of
1221: the exchange constant $J$ in correlated 1D electron systems to the
1222: leading order in the large parameter $r_s$. This method is able to
1223: reproduce a number of exact results for integrable models, which is a
1224: strong indicator of its validity. In comparison to a prior uncontrolled
1225: calculation\cite{Matveev-04} of the same quantity for the Coulomb case,
1226: our method brings a small correction of about $0.7\%$ to the coefficient
1227: in the exponential term; however, the correction to the pre-exponential
1228: factor is large, about 50\%.
1229:
1230: Our another main result --- the relation between $J$ and the
1231: pair-correlation function of spinless electrons --- is a promising
1232: ground for an attack on the regime of intermediate concentrations, $r_s
1233: \sim 1$, which has so far been difficult for a controlled analytical
1234: study. We demonstrated that a very smooth interpolation between the high
1235: and low-density asymptotic is possible based already on the results in
1236: hand (Fig.~\ref{fig:J}).
1237:
1238: Finally, we have applied our method to a number of realistic interaction
1239: potentials corresponding to different geometries of experimental setup.
1240: Apart from nearly impenetrable Coulomb interaction in ultrathin quantum
1241: wires and quantum rings, which has been the primary focus of this work,
1242: we have studied penetrable smooth interactions that take place in the
1243: wires of intermediate width and a screened $1/x^2$ interaction that is
1244: realized, e.g., when the wire is located near a metallic half-plane.
1245:
1246: These findings have direct implications for the energy spectroscopy of
1247: spin-splitting in nanoscale quantum rings\cite{Lorke-00, Fuhrer-04,
1248: Bayer-03} and 1D quantum dots.\cite{Field-90, Jarillo-Herrero-04,
1249: Auslaender-05} Some estimates were given in
1250: Secs.~\ref{sec:integrable_model} and \ref{sec:Domain} above. They should
1251: be especially accurate in the ultrathin-wire limit\cite{Fogler-05a}
1252: $\mathcal{L} = \ln (a_B / R) \gg 1$ that can be achieved in
1253: carbon-nanotube field-effect transistors with high-$k$
1254: dielectrics.\cite{Javey-02}
1255:
1256: In long 1D wires, the exchange coupling $J$ determines the velocity
1257: of the elementary spin excitations
1258: %
1259: \begin{equation}
1260: v_{\sigma } = \frac{\pi}{2} \frac{J a}{\hbar},
1261: \end{equation}
1262: %
1263: which can nowadays be measured by tunneling,\cite{Auslaender-05}
1264: angle-resolved photoemission,\cite{Claessen-02} or deduced from the
1265: enhancement of the spin susceptibility and electron specific
1266: heat.\cite{Fogler-05a} Our result for $v_{\sigma }$ reads
1267: (cf.~Table~\ref{tbl:Results})
1268: %
1269: \begin{equation}
1270: {v_{\sigma }}/{v_{F}}=17.8\,{r_{s}^{3/4}}e^{-\eta \sqrt{2r_{s}}}/
1271: {\mathcal{L}},\quad \eta =2.7978(2), \label{eqn:v_sigma_Coulomb}
1272: \end{equation}
1273: %
1274: where $v_F = (\pi /2)(\hbar / m a)$ is the Fermi velocity.
1275:
1276: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1277:
1278: \begin{acknowledgments}
1279:
1280: The support from the A.~P. Sloan Foundation and C.~\&~W. Hellman Fund is
1281: gratefully acknowledged. We thank A.~D. Klironomos, R.~R. Ramazashvili,
1282: and K.~A. Matveev for discussions.
1283:
1284: \end{acknowledgments}
1285:
1286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1287: \appendix
1288:
1289: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1290: \section{Spin-dependent part of the effective potential}
1291: \label{sec:U_sigma}
1292:
1293: In this Appendix we derive Eqs.~(\ref{eqn:U_sigma}),
1294: (\ref{eqn:U_sigma_CSM}), and (\ref{eqn:J_smooth}). Let $x \equiv
1295: x_1 - x_0$ be the relative distance between the two electrons for which
1296: we want to compute the exchange coupling. Consider the orbital part of
1297: the many-body wavefunction $\Phi$,which is a function of $x$ and $N - 1$
1298: other coordinates. In principle, $\Phi$ is different for different
1299: many-body spin states. Nonetheless, under the condition $J \ll E_F$
1300: where only nearest-neighbor exchanges are important, at small $x$
1301: function $\Phi$ depends primarily on the total spin of the given pair.
1302: The triplet-state wavefunction $\Phi_t(x)$ is odd and therefore
1303: $\Phi_t(0) = 0$. In the singlet state we have $\Phi_s(x)$ which is even
1304: and finite at $x = 0$ (cf.~Sec.~\ref{sec:Three}). However, because of
1305: the strong Coulomb repulsion $\Phi_s(0)$ is very small. At $x\ll a$ the
1306: two electrons of interest are under a large Coulomb barrier and it is
1307: legitimate to assume that $\Phi_t$ and $\Phi_s$ are functions of only
1308: $x$, the fastest variable in the problem. Our goal is to choose
1309: $U_{\sigma}(x)$ that reproduces the difference between $\Phi_{s}$ and
1310: $\Phi_{t}$ at the level of the first-order perturbation theory. This is
1311: achieved if there exists a window $a_{c}\ll x\ll a$, where the following
1312: equation is satisfied:
1313: %
1314: \begin{equation}
1315: \Phi_{t}(x) - \Phi_{s}(x) = \int\limits_{0}^{\infty }dx^{\prime}
1316: G(x,x^{\prime})U_{\sigma }(x^{\prime })\Phi_{t}(x^{\prime }).
1317: \label{eqn:U_sigma_eqn}
1318: \end{equation}
1319: %
1320: Here the Green's function $G$ is the solution of
1321: %
1322: \begin{equation}
1323: \left[ -\frac{\hbar ^{2}}{2\mu }\partial _{x}^{2}+U(x)-E\right]
1324: G(x,x^{\prime })=-\delta (x-x^{\prime }),
1325: \label{eqn:G_equation}
1326: \end{equation}
1327: %
1328: where $E\sim e^{2}/\epsilon a$ is the total energy. At the end of the
1329: calculation one can verify that the final result
1330: [Eq.~(\ref{eqn:U_sigma})] holds independently of the precise value of
1331: $E$ up to terms $O(1/\mathcal{L})$.
1332:
1333: Our next step is to use the usual representation of $G$,
1334: %
1335: \begin{equation}
1336: G(x,x^{\prime })=-\frac{2\mu }{\hbar ^{2}Q}\left\{
1337: \begin{array}{ll}
1338: u(x)v(x^{\prime }), & x>x^{\prime }, \\
1339: v(x)u(x^{\prime }), & x<x^{\prime },
1340: \end{array}
1341: \right.
1342: \end{equation}
1343: %
1344: in terms of two linearly-independent solutions of the (homogeneous)
1345: Schr\"odinger equation~(\ref{eqn:G_equation}). The solution $u(x)$
1346: initially decays exponentially with $x$ and then turns into a wave
1347: propagating towards $x=+\infty $. The other solution, $v(x)$ has a node
1348: at $x = 0$, exhibits an exponential rise and finally becomes a standing
1349: wave. The quantity $Q = u v^{\prime} - u^{\prime } v$ is their Wronskian.
1350:
1351: Wavefunctions $\Phi_{s}$ and $\Phi_{t}$ are the linear combinations of
1352: $u$ and $v$ that are even and odd in $x$, respectively. For $x$ at which
1353: Eq.~(\ref{eqn:U_sigma_eqn}) is valid, $\Phi_{t}\simeq \Phi_{s}\simeq
1354: \text{const}\times v(x)$. Using this property and some trivial algebra,
1355: we get
1356: %
1357: \begin{equation}
1358: \int\limits_{0}^{\infty} dx U_\sigma(x)v^2(x)=\frac{\hbar^2}{2\mu }%
1359: \frac{Q^{2}}{u(+0)u^{\prime }(+0)}.
1360: \label{eqn:U_sigma_II}
1361: \end{equation}
1362: %
1363: For the Coulomb case [Eq.~(\ref{eqn:U})] $u$ and $v$ can be expressed in
1364: terms of Whittaker functions\cite{Gradshteyn-Ryzhik} $W_\alpha(z)$ and
1365: $M_\alpha(z)$ (see an earlier remark in Sec.~\ref{sec:Three} and
1366: Ref.~\onlinecite{Fogler-05a}):
1367: %
1368: \label{eqn:u,v_Coulomb}
1369: \begin{equation}
1370: \begin{split}
1371: u(x)& =W_{-i\nu, 1/2}[i(x-R)/b], \\
1372: v(x)& =M_{-i\nu, 1/2}[i(x-R)/b] \\
1373: & -M_{-i\nu, 1/2}(-iR / b)u(x)/u(0),
1374: \end{split}
1375: \end{equation}
1376: %
1377: where $b = \hbar / (8\mu E)^{1/2}$ and $\nu = b / a_{B}\sim
1378: \sqrt{r_{s}}\gg 1$. Using the relations\cite{Gradshteyn-Ryzhik}
1379: $W_{-i\nu ,1/2}(0)=1/\Gamma (1+i\nu )$ and $M_{-i\nu,1/2}(iz) \simeq iz$
1380: at $z\ll 1$, we find $Q=i/\Gamma (1-i\nu) b$. Substituting these
1381: results into Eq.~(\ref{eqn:U_sigma_II}) and dropping terms that are
1382: small in the parameter $R/a_{B}\ll 1$, we recover
1383: Eq.~(\ref{eqn:U_sigma}).
1384:
1385: Similarly, for the CSM model [Eq.~(\ref{eqn:U_CSM})] $u$ and $v$ are
1386: given by
1387: %
1388: \begin{equation}
1389: \begin{split}
1390: u\left( x\right) & =\sqrt{x}H_{\lambda -1/2}^{(1)}(qx), \\
1391: v\left( x\right) & =\sqrt{x}J_{\lambda -1/2}(qx),
1392: \end{split}
1393: \end{equation}
1394: %
1395: where $q=\sqrt{m E}/\hbar$, and $H_\nu^{(1)}$, $J_\nu$ are the Hankel
1396: and the Bessel functions of the first kind,\cite{Gradshteyn-Ryzhik}
1397: respectively. Their Wronskian is given by $W\{u,v\}=1/i\pi (\lambda -
1398: 1/2)$, and their known asymptotic behavior leads in this case
1399: to the formulas
1400: %
1401: \begin{equation}
1402: \begin{split}
1403: u(R) &\simeq \Gamma (\lambda -1/2) R^{1-\lambda }/i\pi,
1404: \\
1405: u^{\prime}(R) &\simeq (1-\lambda) u(R) / R,
1406: \\
1407: v(R) & \simeq R^{\lambda }/\Gamma (\lambda +1/2),
1408: \end{split}
1409: \end{equation}
1410: %
1411: so that the constraint on $U_\sigma(x)$ is expressed by
1412: Eq.~(\ref{eqn:U_sigma_CSM}).
1413:
1414: Finally, in thick 1D wires, where $U(x) \simeq \text{const}$ at small
1415: $x$ (Sec.~\ref{sec:Domain}), the solutions of Eq.~(\ref{eqn:G_equation})
1416: are $u(x) = \exp(-x / a_0)$ and $v(x)=\sinh (x / a_0)$, where $a_0 =
1417: \hbar / \sqrt{m U(0)}$. These formulas leads to
1418: Eq.~(\ref{eqn:J_smooth}).
1419:
1420:
1421: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1422: \section{Physical realization of the Calogero-Sutherland-Moser model}
1423: \label{sec:CSM_law}
1424:
1425: The CSM interaction law can be realized experimentally if the 1D wire is
1426: positioned nearby a metallic half-plane, parallel to its edge, see
1427: Fig.~\ref{fig:CSM}. In this configuration the interaction potential has
1428: the Coulomb form [Eq.~(\ref{eqn:U})] only at small $x$. Below we show
1429: that at larger distances, $|x| \gg D$, the interaction changes to the
1430: CSM law $U(x) \simeq A / x^2$, due to the screening effect of the metal.
1431: For simplicity, let us assume that the metallic half-plane is an ideal
1432: conductor and that $x \gg R$. In this case $U(x) = V(x, D, \theta)$,
1433: where $V$ is the solution of the following electrostatic problem:
1434: %
1435: \begin{align}
1436: \nabla^2 V(\textbf{r}) &= \left(\frac{\partial^2}{\partial x^2} +
1437: \frac{1}{\rho} \frac{\partial}{\partial \rho} \rho
1438: \frac{\partial}{\partial \rho}
1439: + \frac{1}{\rho^2} \frac{\partial^2}{\partial \varphi^2}\right)
1440: V(x, \rho, \varphi)
1441: \nonumber\\
1442: &= -\frac{4 \pi e^2}{\epsilon D} \delta(x) \delta(\rho - D)
1443: \delta(\varphi - \theta),
1444: \label{eqn:Laplace_V}\\
1445: V(\textbf{r}) &= 0,\quad \varphi = \pm \pi.
1446: \label{eqn:bc_V}
1447: \end{align}
1448: %
1449: Using standard methods (cf.~Ref.~\onlinecite{Fogler-04} and references
1450: therein) one can express $V$ in terms of elliptic integrals. However,
1451: for a general $\theta$, such expressions are not very illuminating. An
1452: exception is the case $\theta = 0$, where $V$ takes the form
1453: %
1454: \begin{equation}
1455: V(\textbf{r}) = \frac{2}{\pi \epsilon} \frac{e^2}{|\textbf{r} - \textbf{D}|}
1456: \arctan \frac{\sqrt{2 D \rho \left( 1 + \cos\varphi\right) }}
1457: {|\textbf{r} - \textbf{D}|},
1458: \label{eqn:V_00}
1459: \end{equation}
1460: %
1461: where $\textbf{D} = (0, 0, D)$. This formula can be verified by the direct
1462: substitution into Eqs.~(\ref{eqn:Laplace_V}) and (\ref{eqn:bc_V}). For the
1463: intra-wire interaction law we obtain
1464: %
1465: \begin{equation}
1466: U(x) = \frac{2}{\pi \epsilon} \frac{e^2}{|x|} \arctan \frac{2D}{|x|},
1467: \label{eqn:U_00}
1468: \end{equation}
1469: %
1470: so that $A = 4 D e^2 /\pi \epsilon$ in this case. We can also show that
1471: for an arbitrary $-\pi < \theta < \pi$ the coefficient $A$ is given by
1472: %
1473: \begin{equation}
1474: A = \frac{2 D}{\pi} \frac{e^2}{\epsilon} \left( 1 + \cos\theta\right) .
1475: \label{eqn:A}
1476: \end{equation}
1477: %
1478: First we note that the Fourier transform of $U$ must have the form
1479: $\tilde{U}(q) \simeq \tilde{U}(0) - \pi A |q|$ at small $q$. On the
1480: other hand, $\tilde{U}$ is given by the generalized Fourier series
1481: %
1482: %\begin{equation}
1483: \[
1484: \tilde{U}(q) = \frac{2 e^2}{\epsilon} \sum\limits_{n = 1}^\infty
1485: \left[1 - (-1)^n \cos n \theta\right] I_{{n}/{2}} (q D)
1486: K_{{n}/{2}}(q D),
1487: \]
1488: %\label{eqn:U_q}
1489: %\end{equation}
1490: %
1491: where $I_\nu$ and $K_\nu$ are the modified Bessel functions of the first
1492: and the second kind, respectively.\cite{Gradshteyn-Ryzhik} The
1493: non-analytic $|q|$-contribution comes only from the $n = 1$ term in the
1494: series. Using the formulas\cite{Gradshteyn-Ryzhik} $I_{1/2}(z) = (2 /
1495: \pi z)^{1/2} \sinh z$ and $K_{1/2}(z) = (\pi / 2 z)^{1/2} \exp(-z)$, we
1496: arrive at Eq.~(\ref{eqn:A}).
1497:
1498: %%%%%%%%%%%
1499: % FIG. 5
1500: %
1501: \begin{figure}
1502: \begin{center}
1503: \includegraphics[height=1.75in]{wcx_5.eps}
1504: \end{center}
1505: \caption{The geometry of the electrostatic problem that
1506: gives rise to the CSM interaction law. (a) The cross-sectional view
1507: in a plane normal to the $x$-axis. The thick line symbolizes the metal
1508: half-plane, the large dot stands for the 1D wire. The drawing also
1509: indicates how the polar coordinates $\rho$, $\varphi$ of an arbitrary point
1510: $\textbf{r}$ are defined. (b) A three-dimensional view of the system.}
1511: \label{fig:CSM}
1512: \end{figure}
1513: %%%%%%%%%%%
1514:
1515: The geometry of Fig.~\ref{fig:CSM}(b) has been realized in the experiment
1516: of \textcite{Auslaender-05}, where the role of the
1517: metallic half-plane was played by a two-dimensional electron gas. In
1518: contrast to our idealized electrostatic model, the two-dimensional metal
1519: has a non-zero Thomas-Fermi screening radius $r_{\text{TF}}$. We expect that in
1520: this situation the only change in Eq.~(\ref{eqn:A}) is the replacement of
1521: $D$ by $D + r_{\text{TF}}$. This yields the following estimate for
1522: the parameter $\lambda$ in Eq.~(\ref{eqn:U_CSM}):
1523: %
1524: \begin{equation}
1525: \lambda = \frac12 + \sqrt{\frac14 + \frac{2}{\pi}
1526: \frac{D + r_{\text{TF}}}{a_B} (1 + \cos\theta)}.
1527: \label{eqn:lambda}
1528: \end{equation}
1529: %
1530: In the experimental setup of Ref.~\onlinecite{Auslaender-05} $D \sim a_B
1531: \sim r_{\text{TF}}$, so that $\lambda = 1.5$--$2$.
1532:
1533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1534:
1535: \begin{thebibliography}{99}
1536:
1537: \bibitem{Saito-book} R.~Saito, G.~Dresselhaus, and M.~S. Dresselhaus,
1538: \emph{Physical Properties of Carbon Nanotubes\/} (Imperial College Press,
1539: London, 1998).
1540:
1541: \bibitem{Huang-01} Y.~Huang, X.~Duan, Y.~Cui, L.~J. Lauhon, K.-H. Kim, and
1542: C.~M. Lieber, Science \textbf{294}, 1313 (2001).
1543:
1544: \bibitem{Heath-03} A.~Nitzan and M.~A. Ratner, Science \textbf{300}, 1384
1545: (2003).
1546:
1547: \bibitem{Lorke-00} A.~Lorke, R.~J. Luyken, A.~O. Govorov, J.~P. Kotthaus,
1548: J.~M. Garcia, and P.~M. Petroff, \prl \textbf{84}, 2223 (2000); R.~J.
1549: Warburton, C.~Sch\"aflein, D.~Haft, F.~Bickel, A.~Lorke, K.~Karrai, J.~M.
1550: Garcia, W.~Schoenfeld, and P.~M. Petroff, Nature (London) \textbf{405}, 926
1551: (2000).
1552:
1553: \bibitem{Fuhrer-04} A.~Fuhrer, T.~Ihn, K.~Ensslin, W.~Wegscheider, and
1554: M.~Bichler, \prl \textbf{93}, 176803 (2004).
1555:
1556: \bibitem{Bayer-03} M.~Bayer, M.~Korkusinski, P.~Hawrylak, T.~Gutbrod,
1557: M.~Michel, and A.~Forchel, \prl \textbf{90}, 186801 (2003).
1558:
1559: \bibitem{Schulz-93} H.~J. Schulz, \prl \textbf{71}, 1864 (1993).
1560:
1561: \bibitem{Egger-99} R.~Egger, W.~H\"ausler, C.~H. Mak, and H.~Grabert,
1562: \prl \textbf{82}, 3320 (1999).
1563:
1564: \bibitem{Field-90} S.~B. Field, M.~A. Kastner, U.~Meirav,
1565: J.~H.~F. Scott-Thomas, D.~A. Antoniadis, H~I. Smith, and S.~J. Wind,
1566: \prb \textbf{42}, 3523 (1990).
1567:
1568: \bibitem[Jarillo-Herrero \emph{et al.}(2004)]{Jarillo-Herrero-04} P.~Jarillo-Herrero,
1569: S.~Sapmaz, C.~Dekker, L.~P. Kouwenhoven, and H.~S.~J. van der Zant,
1570: Nature (London) \textbf{429}, 389 (2004).
1571:
1572: \bibitem[Matveev(2004)]{Matveev-04} K.~A. Matveev, \prb \textbf{70}, 245319
1573: (2004).
1574:
1575: \bibitem{Herring-62} C.~Herring, \rmp \textbf{34}, 631 (1962);
1576:
1577: \bibitem{Reimann-02} S.~M. Reimann and M.~Manninen, \rmp
1578: \textbf{74}, 1283 (2002); S.~Viefers, P. Koskinen, P.~Singha Deo, and
1579: M.~Manninen, Physica\ E \textbf{21}, 1 (2004).
1580:
1581: \bibitem[H\"ausler(1996)]{Hausler-96} W.~H\"ausler, Z.\ Phys.\ B
1582: \textbf{99}, 551 (1996).
1583:
1584: \bibitem[Auslaender \emph{et al.}(2005)]{Auslaender-05} O.~M. Auslaender,
1585: H.~Steinberg, A.~Yacoby, Y.~Tserkovnyak, B.~I. Halperin,
1586: K.~W. Baldwin, L.~N. Pfeiffer, and K.~W. West,
1587: Science \textbf{308}, 88 (2005),
1588: and references therein.
1589:
1590: \bibitem{Claessen-02} R.~Claessen, M.~Sing, U.~Schwingenschl\"{o}gl,
1591: P.~Blaha, M.~Dressel, and C.~S. Jacobsen, \prl \textbf{88}, 096402 (2002).
1592:
1593: \bibitem{Cheianov-04} V.~V. Cheianov and M.~B. Zvonarev,
1594: \prl \textbf{92}, 176401 (2004).
1595:
1596: \bibitem{Fiete-04} G.~A. Fiete and L.~Balents,
1597: \prl \textbf{93}, 226401 (2004).
1598:
1599: \bibitem{Fiete-05} G.~A. Fiete, J.~Qian, Ya.~Tserkovnyak, and B.~I. Halperin,
1600: \prb \textbf{72}, 045315 (2005).
1601:
1602: \bibitem[Fogler and Pivovarov(2005)]{Fogler-05b} M.~Fogler and E.~Pivovarov,
1603: \eprint{cond-mat/0504502}.
1604:
1605: \bibitem[Klironomos \emph{et al.}(2005)]{Klironomos-05} A.~D. Klironomos,
1606: R.~R. Ramazashvili, and K.~A. Matveev, \eprint{cond-mat/0504118}.
1607:
1608: \bibitem[Fogler(2005)]{Fogler-05a} M.~M. Fogler, \prb\textbf{71}, 161304(R)
1609: (2005).
1610:
1611: \bibitem{Usukura-05} J.~Usukura, Y.~Saiga, and D.~S. Hirashima, J.\ Phys.\
1612: Soc.\ Jpn. \textbf{74}, 1231 (2005), and references therein.
1613:
1614: \bibitem{Friesen-80} W.~I. Friesen and B.~Bergerson,
1615: J.\ Phys.\ C \textbf{13}, 6627 (1980).
1616:
1617: \bibitem{Szafran-04} B.~Szafran, F.~M. Peeters, S.~Bednarek, T.~Chwiej, and
1618: J.~Adamowski, \prb \textbf{70}, 035401 (2004).
1619:
1620: \bibitem{Herring-64} C.~Herring, Phys.\ Rev. \textbf{134}, A362 (1964).
1621:
1622: \bibitem{Landau-III} L.~D. Landau and E.~M. Lifshitz,
1623: \textit{Quantum Mechanics} (Pergamon, Oxford, 1977), Sec.~81.
1624:
1625: \bibitem{Roger-83} M.~Roger, J.~H. Hetherington, and J.~M. Delrieu, \rmp
1626: \textbf{55}, 1 (1983).
1627:
1628: \bibitem{Gradshteyn-Ryzhik} I.~S. Gradshteyn and I.~M. Ryzhik, \emph{Table
1629: of Integrals, Series, and Products}, 6th ed., edited by A.~Jeffrey and
1630: D.~Zwillinger (Academic, San Diego, 2000).
1631:
1632: \bibitem{Maksym-96} P.~A. Maksym, \prb \textbf{53}, 10871 (1996);
1633: P.~Koskinen, M.~Koskinen, and M.~Manninen,
1634: Eur.\ Phys.\ J.\ B \textbf{28}, 483 (2002).
1635:
1636: \bibitem{Ogata-90} M.~Ogata and H.~Shiba, \prb \textbf{41}, 2326 (1990).
1637:
1638: \bibitem{Sutherland-71} B.~Sutherland, \pra \textbf{4}, 2019 (1971);
1639: Int.\ J.\ Mod.\ Phys.\ B \textbf{11}, 355 (1997).
1640:
1641: \bibitem{Forrester-92} P.~J. Forrester, Nucl.\ Phys.\ B \textbf{388}, 671
1642: (1992).
1643:
1644: \bibitem{Ha-94} Z.~N.~C. Ha,
1645: \prl \textbf{73}, 1574 (1994).
1646:
1647: \bibitem{Javey-02} A.~Javey, H.~Kim, M.~Brink, Q.~Wang, A.~Ural, J.~Guo,
1648: P.~Mcintyre, P.~Mceuen, M.~Lundstrom, and H.~Dai, Nature Materials
1649: \textbf{1}, 241 (2002); B.~M. Kim, T.~Brintlinger, E.~Cobas, M.~S. Fuhrer,
1650: H.~Zheng, Z.~Yu, R.~Droopad, J.~Ramdani, and K.~Eisenbeiser,
1651: Appl.\ Phys.\ Lett. \textbf{84}, 1946 (2004).
1652:
1653: \bibitem{Fogler-04} M.~M. Fogler,
1654: \prb \textbf{69}, 245321 (2004); \textbf{70}, 129902(E) (2004).
1655:
1656: \bibitem{Piacente-04} G.~Piacente, I.~V. Schweigert, J.~J. Betouras,
1657: and F.~M. Peeters,
1658: \prb \textbf{69}, 045324 (2004).
1659:
1660: \bibitem{Klironomos-xxx} A.~D. Klironomos, J.~S. Meyer,
1661: and K.~A. Matveev,
1662: \eprint{cond-mat/0507387}.
1663:
1664: \end{thebibliography}
1665:
1666: \end{document}
1667: