cond-mat0311086/gf.tex
1: \documentclass[prb,aps,amsfonts,amssymb,draft,floats,twocolumn]{revtex4}
2: \usepackage{epsf}
3: %\documentclass[prb,aps,amsfonts,amssymb,draft,twocolumn]{revtex4}
4: %\usepackage{graphicx}
5: 
6: \newcommand{\be}{\begin{equation}}
7: \newcommand{\ee}{\end{equation}}
8: \newcommand{\one}{\openone}
9: \newcommand{\vS}{\mathbf{S}}
10: \newcommand{\vH}{\mathbf{H}}
11: \newcommand{\vs}{\mathbf{s}}
12: \newcommand{\vm}{\mathbf{m}}
13: \newcommand{\vsigma}{\mbox{\boldmath $\sigma$}}
14: \newcommand{\vp}{\mathbf{p}}
15: \newcommand{\vu}{\mathbf{u}}
16: \renewcommand{\vr}{\mathbf{r}}
17: \newcommand{\spacing}{\delta}
18: 
19: \begin{document}
20: 
21: 
22: 
23: 
24: \title{Electron correlations in metal nanoparticles with
25: spin-orbit scattering}
26: 
27: 
28: \author{Denis A. Gorokhov and Piet W. Brouwer}
29: 
30: 
31: \affiliation{Laboratory of Atomic and Solid State Physics,
32: Cornell University, Ithaca, NY 14853-2501, USA}
33: 
34: \begin{abstract}
35: The combined effect of electron-electron interactions
36: and spin-orbit scattering
37: in metal nanoparticles
38: can be studied by measuring splitting of electron levels in 
39: magnetic field
40: ($g$ factors) in tunneling spectroscopy
41: experiments.
42: Using random matrix theory to describe the single-electron
43: states in the metal particle, we find that even a relatively small 
44: electron-electron interaction strength
45: (ratio of exchange constant $J$ and mean level
46: spacing $\spacing$  $\simeq 0.3$) significantly
47: increases $g$-factor fluctuations for
48: not-too-strong spin-orbit scattering rates (spin-orbit
49: time $\tau_{\rm so} \gtrsim 1/\spacing$).
50: In particular, $g$-factors larger than 2 could be observed.
51: (This is a manifestation of the many-body correlation effects
52: in nanoparticles). While so far measurements 
53: only on noble metal (Cu, Ag, Au) and Al samples
54: have been done for which the effects of
55:  electron-electron interactions are negligible, 
56: we discuss   
57: the possibility of observing 
58: interaction effects
59: in nanoparticles made of other metals. 
60: 
61: \end{abstract}
62: 
63: \maketitle
64: 
65: %\vskip1.5cm
66: \newpage
67: 
68: 
69: 
70: \section{Introduction}
71: 
72: While the study of the combined effect of electron-electron
73: interactions and elastic impurity scattering in two dimensions and
74: near the metal-insulator transition in three dimensions remains one of
75: the most important problems in solid state
76: physics,  the description of electron-electron
77: interactions in disordered normal metal nanoparticles ({\em i.e.,}
78: ``zero dimensions'') has been found to be remarkably 
79: simple\cite{kn:kurland2000,kn:aleiner2002}: 
80: 
81: At a 
82: fixed
83: number of electrons and without spin-orbit scattering, the only 
84: relevant interaction term is a long-range exchange
85: interaction\cite{foot0}
86: \begin{equation}
87:   H_{\rm int} = - J \vS^2, \label{eq:HintS}
88: \end{equation}
89: that couples to the total spin $\vS$ of the
90: nanoparticle. The exchange
91: constant $J$ is closely related to one of the Fermi Liquid
92: constants of the bulk metal, and is independent of the details of the
93: impurity configuration inside the nanoparticle. The
94: interaction Hamiltonian (\ref{eq:HintS}), which is known as
95: ``universal interaction Hamiltonian'', is the only form of the
96: electron-electron interaction compatible with random matrix
97: theory.\cite{kn:aleiner2002} 
98: Random matrix theory provides a valid description 
99: of single-electron states 
100: %the fact that the spin is the only
101: %good quantum numbers describing single-electron wavefunctions; both
102: %the electron momentum and position are fully randomized 
103: as long as the
104: dimensionless conductance $g$ of the nanoparticle, which is the ratio
105: of the Thouless energy $E_{\rm Th}$ and the mean level spacing
106: $\spacing$, is large.\cite{kn:efetov1983,kn:altshuler1986} 
107: Residual interaction terms not included in
108: Eq.\ (\ref{eq:HintS})  are sample specific
109: and small in comparison to Eq.\ (\ref{eq:HintS}) by, at least, a factor
110: $1/g$.
111: 
112: In the presence of spin-orbit scattering, spin is also 
113: randomized, giving rise to both sample-to-sample fluctuations of the
114: electron-electron interaction and a suppression of the exchange 
115: interaction (\ref{eq:HintS}). Since the spin-orbit scattering rate
116: $\gamma_{\rm so} = \hbar/\tau_{\rm so}$ plays the role of a
117: ``Thouless energy'' for the spin degree of freedom, for strong 
118: spin-orbit scattering the residual exchange interaction becomes 
119: small by a factor $\gamma_{\rm so}/\spacing\gg 1$ in comparison to the 
120: interaction strength without spin-orbit scattering if spin-orbit
121: scattering. However,
122: the exchange interaction remains the dominant contribution to the
123: electron-electron interaction as long as $\gamma_{\rm so}/\spacing
124: \ll g$. 
125: 
126: In this paper, we present a detailed analysis of the combined 
127: effect of spin-orbit scattering and electron-electron interactions 
128: in the regime of moderate spin-orbit scattering, $\gamma_{\rm so} 
129: \sim \spacing$. In this parameter regime, the exchange interaction
130: is not fully suppressed, while fluctuations are of the same order 
131: as the average.\cite{kn:adam2002} This makes the
132: regime of moderate spin-orbit scattering rates qualitatively
133: different from that without spin-orbit scattering ($\gamma_{\rm so} 
134: = 0$) and that of strong spin-orbit scattering 
135: ($\gamma_{\rm so}/\spacing \gg 1$.) The parameter regime $\gamma_{\rm
136:   so} \sim \spacing$ is of interest for recent
137: experiments on metal 
138: nanoparticles,\cite{kn:salinas1999,kn:davidovic1999,kn:petta2001,%
139: kn:petta2002} in which the magnetic-field
140: dependences of many-electron levels has been measured using
141: tunneling 
142: spectroscopy.\cite{kn:ralph1995} Moreover, an analysis of the regime
143: $\gamma_{\rm so}/\spacing \sim 1$ serves as a model study of the 
144: breakdown of the ``universal interaction
145: Hamiltonian'' when the dimensionless conductance is small. A brief
146: account of some of our findings was previously published in Ref.\
147: \onlinecite{kn:gorokhov2003}.
148: 
149: Experimentally, the spin structure of electronic states can 
150: be measured through the magnetic-field dependence of steps
151: in the current-voltage characteristic of a metal nanoparticle coupled to
152: source and drain electrodes via tunneling contacts. These steps
153: occur if the applied voltage is equal to the difference of
154: the energies of many-electron levels $|N_e+1;k\rangle$ and 
155: $|N_e;l\rangle$ which have $N_e$ and $N_e+1$ electrons,
156: respectively,\cite{kn:vondelft2001}
157: \begin{equation}
158:   e V_{kl} = E_{N_e+1,k} - E_{N_e,l}.
159:   \label{eq:V}
160: \end{equation}
161: The
162: derivative $\partial V_{kl}/\partial B$ of the voltage at which a
163: step occurs to the magnetic field $B$ is parameterized 
164: through a ``$g$-factor'', 
165: \begin{equation}
166:   e \frac{\partial V_{kl}}{\partial B} = \pm \frac{1}{2} g_{kl} \mu_B,
167:   \label{eq:Vg}
168: \end{equation}
169: where $\mu_B = |e| \hbar/2mc$ is the Bohr magneton. 
170: Without spin-orbit scattering, but with electron-electron 
171: interactions, many-electron states are characterized by their total 
172: spin $S$ and by its $z$ component. 
173: %Although, in 
174: %principle, many-electron states with $S > 1/2$ exist,
175: Since tunneling spectroscopy measures transitions in which the
176: electron number changes by one, the total spin $S$ of the nanoparticle 
177: changes by $1/2$ upon addition or removal of an electron. This 
178: ``selection rule'' renders all observed $g$ factors equal to two,
179: irrespective of the spin $S_k$ and $S_l$ of the two many-electron 
180: levels participating in the transition. 
181: On the other hand, without interactions, the tunneling spectroscopy
182: $g$ factors correspond to the magnetic moment of one single-electron
183: level only, giving rise to a distribution of $g$ factors in which all
184: levels have, at most, spin 
185: $1/2$.\cite{kn:brouwer2000,kn:matveev2000,kn:adam2002b}
186: As was shown in 
187: Ref.\ \onlinecite{kn:gorokhov2003}, the combined effect of spin-orbit
188: scattering and electron-electron interactions
189: is to simultaneously suppress the spin of the many-electron
190: states and lift the ``selection rules'', causing a much wider distribution
191: of tunneling spectroscopy $g$ factors than in the non-interacting
192: case. In particular, there is a significant probability
193: to find $g$ factors larger than two if spin-orbit scattering is not
194: too strong ($\gamma_{\rm so}/\spacing \lesssim 2$). The occurrence of
195: $g$ factors larger than two is a unique signature of the interplay
196: of electron-electron interactions and spin-orbit scattering.
197: 
198: What are the main differences between $g$ factor distributions with
199: and without electron-electron interactions? In order to answer that 
200: question, we note that
201: the electron-electron interactions has two main effects in a
202: nanoparticle: to organize the many-electron 
203: states according to their total spin $S$, {\em i.e.,} to lift
204: the degeneracy between states of different $S$ but with the same
205: orbital content, and to {\em lower} the
206: energy of a many electron state with spin $S$ by the amount $JS(S+1)$,
207: increasing the abundance of high-$S$ states among low-energy
208: excited states, see Fig.\ \ref{fig:1a}. 
209: (In fact, for $J/\spacing \gtrsim 0.3$, there is a significant
210: probability that the ground state has a nontrivial spin $S > 1/2$, 
211: see, {\em e.g.,} 
212: Refs.\ \onlinecite{kn:brouwer1999b,kn:baranger2000}). It is because
213: of the combination of these two effects, together with the lifting of 
214: selection rules by spin-orbit scattering, that electron-electron
215: interactions enhance the width of the $g$-factor distribution so
216: significantly. Moreover, because the relative
217: abundance of high-$S$ states depends on the excitation energy, the
218: $g$ factor distribution will be different for transitions to an
219: excited state than for transitions to the ground states. Again, this
220: is different from the non-interacting case, where 
221: $g$ factor distributions for transitions to the ground state
222: and to excited states are equal. 
223: A third difference between the cases with and without
224: interactions is that,
225: as the selection rules are gradually broken down by spin-orbit
226: scattering, different transitions may have very different weights,
227: in contrast to the non-interacting case, for which all transitions
228: have weights within a factor of order unity from each other. (The
229: ``weight'' of the transition is the height of the corresponding
230: step in the current-voltage characteristics.)
231: 
232: In principle, one should consider contributions to the $g$ factor
233: from the orbital magnetic moments of the energy levels and from
234: the spin magnetic moment.\cite{kn:matveev2000} In this work, we 
235: consider the
236: spin contribution to the $g$ factor only and neglect the orbital
237: contribution. For the spin-orbit scattering rates $\gamma_{\rm
238:   so}/\spacing \sim 1$ we consider here,
239: this is justified if the electron motion in the metal nanoparticle is
240: diffusive with mean free path $l$ much smaller than the nanoparticle size
241: $L$.\cite{kn:matveev2000,kn:adam2002b}
242: 
243: The outline of this paper is as follows. In section II we present
244: the theoretical formalism. Since we only consider spin-orbit
245: scattering rates $\gamma_{\rm so}/\spacing \ll g$, random-matrix
246: theory can be used to describe the single-electron states. 
247: %Although
248: %electron-electron interactions mix single-electron states at
249: %different energies, we verify that the relevant states contributing
250: %to the $g$ factors are all located within an energy window of width
251: %$\ll E_{\rm Th}$. 
252: In Sec.\ III we discuss the results of numerical
253: simulations for the $g$ factor distribution for transitions from
254: the $N_e$-electron ground state to the $(N_e + 1)$-electron ground
255: state and $(N_e+1)$-electron excited states, where $N_e$ is taken
256: even. The restriction to transitions starting from the $N_e$-electron 
257: ground state is appropriate if the metal nanoparticle relaxes to the
258: $N_e$-electron ground state between tunneling events. In Sec.\ IV
259: we discuss the consequences of our findings for various metals: our
260: results depend on the ratio $J/\spacing$, which strongly depends on
261: the metal under consideration. We
262: conclude in Sec.\ V. Finally, in the appendix, we report an analytical
263: calculation of the $g$ factor distribution for weak spin-orbit
264: scattering, $\gamma_{\rm so}/\spacing \ll 1$, again paying special
265: attention to differences between transitions involving ground states
266: only and transitions to or from an excited state.
267: 
268: \begin{figure}[t]
269: \epsfxsize= 0.8\hsize
270: %\hspace{0.1\hsize}
271: \epsffile{occupation.eps}
272: %\vspace{5cm}
273: \caption{\label{fig:1a}
274: Occupation of single-electron levels for the lowest-lying
275: many-electron states with total spin $S=0$ (a) and $S=1$ (b),
276: and for $S=1/2$ (c) and $S=3/2$ (d). In the absence of
277: exchange interactions, the states with spin $S=0$ and $S=1/2$
278: are the ground states for even and odd numbers of electrons,
279: respectively. The exchange interaction compensates (part of) 
280: the kinetic energy cost of the higher spin states.
281: }
282: \end{figure}
283: 
284: %It turns out that even very ($\lambda \ll 1$) spin-orbit scattering
285: %can dramatically change the $g$-factors. For example, the results
286: %of our numerical simulations show that for $\lambda\rightarrow 0$
287: %and $J/\delta = 0.6$
288: %for a grain with an odd number of electrons 
289: %the probability to have a ground state shown in Fig.~\ref{fig:1a}~(c)
290: %is $P_{1/2} = 0.62$ and for the state shown in Fig.~\ref{fig:1a}~(d)
291: %$P_{3/2} = 0.38$. The probability of having higher spin states
292: %is negligible. Naively, one could expect that the average 
293: %ground state $g$-factor is equal to 
294: %$6P_{3/2} + 2P_{1/2}\approx 3.52$. The simulations, however, give the
295: %value $\approx 2.5$ see Section \ref{sec:simulations}. 
296: %This apparent discrepancy can be explained
297: %by the detailed analysis of the splitting of the four-fold
298: %degenerate $S=3/2$ state, see Fig.~\ref{fig:1a}~(c).
299: %The appendix contains an explicit calculation
300: %of $g$ factor distributions in the limit of very weak spin-orbit
301: %scattering $\tau_{\rm so} \gg 1/\delta$ using the many-body perturbation
302: %theory.
303: 
304: \section{Theoretical Description}
305: \label{sec:theoretical_description}
306: 
307: \subsection{Effective Hamiltonian}
308: 
309: Random matrix theory can be used to describe the single-electron
310: wavefunctions and energy levels of a metal nanoparticle. For a metal 
311: nanoparticle
312: with spin orbit scattering, the appropriate random matrix ensemble
313: interpolates between the Gaussian Orthogonal Ensemble (GOE) and the
314: Gaussian Symplectic Ensemble (GSE) of random matrix theory,
315: \begin{equation}
316:   H_0(\lambda) = H_{\rm GOE} +  H_{\rm so}(\lambda ).
317:   \label{eq:H0}
318: \end{equation}
319: Writing the spin degrees of freedom explicitly, one has
320: \begin{eqnarray}
321:   H_{\rm GOE} &=& S \otimes \one_2, \\
322:   H_{\rm so}(\lambda ) &=& %\frac{1}{2} A_0 \otimes \one_2 +
323:   \frac{i \lambda}{2 \sqrt{N}}
324:   %\frac{i}{2} 
325:   \sum_{j=1}^{3} A_j \otimes \sigma_j,
326: \end{eqnarray}
327: where $\one_2$ is the $2 \times 2$ unit matrix in spin space,
328: $\sigma_j$ is the Pauli matrix ($j=1,2,3$), $S$ is an 
329: $N \times N$ real symmetric matrices, and $A_j$ is a real
330: antisymmetric matrix ($j=1,2,3$). The elements of the matrices
331: $S$, $A_1$, $A_2$, and $A_3$ are drawn from independent
332: Gaussian distributions with zero mean and with equal variances
333: for the off-diagonal elements. The diagonal elements of $S$
334: have double variance, whereas the diagonal elements of 
335: $A_1$, $A_2$, and $A_3$ are zero because of the antisymmetry
336: constraint. In random matrix 
337: theory, the limit $N \to \infty$
338: is taken at the end of the calculation.
339: The random matrix description is valid as long as energy
340: differences of the energy levels and wavefunctions of interest
341: are small compared to the Thouless energy 
342: $E_{\rm Th}$.\cite{kn:efetov1983,kn:altshuler1986}
343: For a disordered metal nanoparticle of size $L$, mean free path $l$, and 
344: Fermi velocity $v_F$, $E_{\rm Th} \sim v_F l/L^2$.
345: 
346: The parameter $\lambda$ in Eq.\ (\ref{eq:H0}) determines the
347: strength of the spin-orbit scattering, $\lambda^2 = \pi \gamma_{\rm so}/
348: \spacing = \pi/\tau_{\rm so} \spacing$, where $\tau_{\rm so} =
349: 1/\gamma_{\rm so}$ is 
350: the spin-orbit
351: time and $\spacing$ is the mean spacing between (spin-degenerate)
352: eigenvalues of $H_{\rm GOE}$, see 
353: Refs.\cite{kn:brouwer2000,kn:matveev2000,kn:adam2002b}
354: The case $\lambda = 0$ corresponds 
355: to the absence of spin-orbit scattering, while the limit 
356: $\lambda \to \infty$ describes the situation when spin-rotation
357: symmetry is completely broken.
358: The factor $1/\sqrt{N}$ in front of $H_{\rm so}$ ensures that
359: the relation between $\lambda$ and the physical parameters 
360: $\tau_{\rm so}$ and $\spacing$ does not involve the matrix size $N$.
361: 
362: Each eigenvalue $\varepsilon_{\mu}$ of the Hamiltonian (\ref{eq:H0})
363: is doubly degenerate (Kramers degeneracy).
364: After diagonalization, the Hamiltonian $H_0$ can be written as
365: \begin{equation}
366:   H_0 = \sum_{\mu} \varepsilon_{\mu} 
367:   (\hat \psi^{\dagger}_{\mu1} 
368:   \hat \psi^{\vphantom{\dagger}}_{\mu1} +
369:   \hat \psi^{\dagger}_{\mu2} 
370:   \hat \psi^{\vphantom{\dagger}}_{\mu2}),
371:   \label{eq:H0single}
372: \end{equation}  
373: where $\hat \psi^{\dagger}_{\mu\alpha}$ and
374: $\hat \psi_{\mu\alpha}$
375: are creation and annihilation operators for an electron in the
376: state $|\mu \alpha\rangle$, where $\alpha=1,2$ labels the two
377: time-reversed states in the Kramers doublet.
378: 
379: %In the 
380: %random matrix limit, there is a simple and consistent description
381: %of the electron-electron interactions in a
382: %metal grain. As shown by Kurland, Aleiner, and 
383: %Altshuler\cite{kn:kurland2000} (see also Ref.\
384: %\onlinecite{kn:aleiner2002}),
385: %to leading order in $E_{\rm Th}/\spacing$ and for a repulsive
386: %electron-electron interaction, the only relevant 
387: %contributions to the interaction Hamiltonian are the capacitive 
388: %charging energy and the long-range exchange interaction,
389: %\begin{equation}
390: %  \hat H_{\rm int} = \frac{e^2}{2C} (\hat N_e^2 - 2 \hat N_e {\cal N}) -
391: %  J \hat \vS^2,
392: %  \label{eq:HH}
393: %\end{equation}
394: %where $C$ is the geometrical capacitance of the grain, $\hat N_e$ is 
395: %the number of electrons on the dot, ${\cal N}$ is a dimensionless 
396: %gate voltage, $J$ is the exchange interaction constant,
397: %and $\vS$ is the total spin of the grain. 
398: %%In terms of creation and
399: %%annihilation operators for electrons in the eigenstates
400: %%$|\mu\alpha\rangle$ of the single-electron Hamiltonian $H_0$, the
401: %%components of the operator $\vS$ read
402: %%\begin{eqnarray}
403: %%  \hat S_i &=& \frac{1}{2} \sum_{\mu} \sum_{\sigma_1,\sigma_2}
404: %%  \hat \psi^{\dagger}_{\mu\sigma_1} (\sigma_i)_{\sigma_1\sigma_2}
405: %%  \hat \psi_{\mu\sigma_2},\ \ i=1,2,3.
406: %%\end{eqnarray}
407: %All other contributions to the interaction Hamiltonian are smaller
408: %by powers of $\spacing/E_{\rm Th}$. 
409: 
410: Combining the single-electron Hamiltonian (\ref{eq:H0single}) and
411: the interaction Hamiltonian (\ref{eq:HintS}), one find the total
412: Hamiltonian
413: \be
414:   {\hat H} = \sum_{\mu} \varepsilon_{\mu} 
415:   (\hat \psi^{\dagger}_{\mu1} 
416:   \hat \psi^{\vphantom{\dagger}}_{\mu1} +
417:   \hat \psi^{\dagger}_{\mu2} 
418:   \hat \psi^{\vphantom{\dagger}}_{\mu2})
419:   - J {\bf {\hat S}}^2.
420: \label{hamiltonian}
421: \label{half_of_Hamiltonian}
422: \ee
423: Equation (\ref{hamiltonian}) is valid up to a charging energy that
424: depends on the electron number $N_e$ only; the charging energy plays
425: no role for in the problem we consider.
426: 
427: In the absence of spin-orbit interaction, the exchange interaction
428: (second term in Eq.\ (\ref{hamiltonian})) commutes with the
429: non-interacting Hamiltonian $H_0$ (first term in Eq.\
430: (\ref{hamiltonian})). The many-electron eigenstates are found by
431: diagonalizing $H_0$ at a fixed value of the total spin $S$ and its
432: $z$ component $S_z$. With spin-orbit interaction, however, the
433: interaction does not commute with $H_0$. 
434: This has important consequences for the
435: ground state and for the excitation spectrum of a metal nanoparticle.
436: Typically, for most normal-metal nanoparticles and for quantum dots,
437: $J$ is estimated to be in the range $0 \lesssim J/\spacing \lesssim
438: 1$, see Sec.~\ref{sec:experiment}. 
439: While this implies that the effect of the exchange interaction
440: cannot be treated in first-order perturbation theory, 
441: interaction effects only cause correlations in a small window
442: around the Fermi energy which is, in principle, available to
443: direct numerical diagonalization. Hereto, we write the operator
444: $\vS$ for the total electron spin in terms of the creation and
445: annihilation operators $\hat \psi^{\dagger}_{\mu\alpha}$ and
446: $\hat \psi_{\mu\alpha}$ of the single-electron Hamiltonian $H_0$,
447: \begin{eqnarray}
448:   \hat \vS &=& \sum_{\mu,\nu} \sum_{\alpha,\beta=1,2}
449:   \hat \psi^{\dagger}_{\mu\alpha} \vs_{\mu\alpha,\nu\beta}
450:   \hat \psi_{\nu\beta}, \label{eq:Stotal}
451: \end{eqnarray}
452: where 
453: \begin{equation}
454:   (s_i)_{\mu\alpha,\nu\beta} = \frac{1}{2}
455:   \langle \nu\beta|\sigma_i|\mu\alpha \rangle,\ \ i=1,2,3.
456: \end{equation}
457: 
458: The quantity of interest in our calculation is the magnetic-field
459: dependence of the many-electron energy levels for small magnetic
460: fields and an odd number of electrons, which is described 
461: through the $g$ factors, see Eqs.\ (\ref{eq:V}) and (\ref{eq:Vg})
462: above. The 
463: magnetic-field dependence arises both through the Zeeman coupling
464: to the electron spin and through the orbital coupling to the angular
465: momentum.\cite{kn:matveev2000,kn:adam2002b} For large diffusive
466: metal nanoparticles and for not-too-large spin-orbit scattering strengths
467: $\tau_{\rm so} \spacing \gtrsim 1$,
468: the Zeeman coupling dominates.\cite{kn:matveev2000,kn:adam2002b}
469: Since the interaction effects studied here are most important for
470: $\tau_{\rm so} \spacing \sim 1$ (see below), we neglect the orbital
471: contribution to the $g$ factors in the discussion below. 
472: For a magnetic field $\vH$ along the $z$ axis, the Zeeman coupling to the 
473: magnetic field is described by the Hamiltonian
474: \begin{eqnarray}
475:   H_{\rm Z} &=& - 2 \mu_B H S_3,
476: \end{eqnarray}
477: where the $z$-component of the total spin $S_3$ is given by Eq.\
478: (\ref{eq:Stotal}) above.
479: 
480: \subsection{Tunneling spectroscopy}
481: 
482: %The magnetic-field dependence of energy levels in a metal nanoparticle is
483: %measured using ``tunneling spectroscopy''. In fact, with
484: %tunneling spectroscopy one does not measure the magnetic-field
485: %dependence of the many-electron levels $E_{k}$ directly. Instead, one
486: %measures the energy difference\cite{kn:vondelft2001}
487: %\be
488: %  \Delta E_{kl} = E_{N_{e}+1,k} - E_{N_{e},l},
489: %  \label{eq:ElEk}
490: %  \label{delta_energy}
491: %\ee
492: %corresponding to a transition from the many-electron state 
493: %$|N_e,l\rangle$ for a grain with $N_e$ electrons
494: %to the many-electron state $|N_e+1,k\rangle$ for a grain with
495: %$N_e+1$ electrons. 
496: Following Ref.\ \onlinecite{kn:vondelft2001},
497: we assume that the conductance of the tunneling contact connecting 
498: the nanoparticle to the source reservoirs is much smaller than the
499: conductance of the contacts connecting the particle to the drain
500: reservoir, so that the current $I$ through the nanoparticle is limited by 
501: the processes where an electron tunnels {\em onto} the particle.
502: In this case, one can assume that relaxation is sufficiently fast
503: that the nanoparticle is in the $N_e$-particle ground state 
504: $|N_E,0\rangle$
505: before each
506: tunneling event. It follows that current steps occur in the 
507: current-voltage
508: characteristic or, equivalently, a peak in the nanoparticle's differential
509: conductance $\partial I/\partial V$, when the source-drain
510: voltage $e V = e V_{k0} = E_{N_{\rm e} + 1,k} - E_{N_e,0}$, see
511: Eq.\ (\ref{eq:V}).
512: For a point contact that
513: injects electrons into the nanoparticle at position $\vr$, the size of the 
514: current step is proportional to the matrix element
515: \begin{eqnarray}
516:   w_{k} &=& | \langle N_e+1, k | \hat \psi^{\dagger}_{\uparrow}(\vr) 
517:   | N_e, 0 \rangle |^2 
518:   \nonumber \\ && \mbox{} + | \langle N_e+1, k | 
519:   \hat \psi^{\dagger}_{\downarrow}(\vr) | N_e, 0 \rangle |^2,
520:   \label{weights}
521: \end{eqnarray}
522: where the creation operator $\hat \psi^{\dagger}_{\sigma}(\vr)$
523: creates an electron with spin $\sigma$ in the nanoparticle at position $\vr$.
524: In terms of the basis of single-electron states, one has
525: \begin{equation}
526:   \hat \psi^{\dagger}_{\sigma}(\vr) =
527:   \sum_{\mu} \left(
528:   \hat \psi^{\dagger}_{\mu 1}
529:   \langle \vr \sigma | \mu 1 \rangle +
530:   \hat \psi^{\dagger}_{\mu 2}
531:   \langle \vr \sigma | \mu 2 \rangle \right).
532:   \label{creation_operator}
533: \end{equation}
534: In random matrix theory, the matrix element $\langle \vr \sigma | \mu
535: \alpha \rangle$, $\sigma=\pm 1$, is replaced by an (arbitrary) spinor in 
536: the $2N$-component vector representing the state $\mu \alpha$,
537: $\alpha=1,2$. 
538: 
539: \begin{figure}[t]
540: ~\\
541: \epsfxsize= 0.98\hsize
542: \hspace{0.1\hsize}
543: \epsffile{weak_SO.eps}
544: \caption{ \label{fig:1}
545: Top panel: schematic representation of four lowest many-electron 
546: levels for a nanoparticle with an odd number of particles without
547: spin-orbit scattering. The spacings between the single-electron
548: levels are as indicated in the figure. The figure represents the case 
549: $0< \Delta_0 + \Delta_1 - 3J  < \Delta_0, \Delta_1$. Note that the
550: first excited state has spin $S=3/2$ and is fourfold degenerate.
551: Bottom panel: Evolution of energy levels of the nanoparticle for small 
552: spin-orbit scattering rates $\lambda$ and magnetic fields $H$. 
553: Spin-orbit scattering separates the quadruplet (b) into two
554: doublets with ill-defined spin. The magnetic field lifts the
555: degeneracy of all remaining doublets.}
556: \end{figure}
557: 
558: In the presence of both spin-orbit scattering and electron-electron
559: interactions, the $N_e$-electron states $|N_e,l\rangle$
560: are non-degenerate for zero magnetic field if $N_e$ is even. 
561: In that case, $\partial E_l/\partial H = 0$ at $H=0$.
562: On the other hand, $(N_e+1)$-electron states are doubly degenerate 
563: if $N_e$ is even. These states split in a magnetic field. For small 
564: magnetic fields the splitting is linear. Hence, if $N_e$ is even,
565: the $g$ factor $g_{kl}$ associated with voltage $V_{kl}$ at which 
566: a current steps occurs is directly related to the magnetic-field 
567: derivative of the $(N_e+1)$-electron level $|N_e + 1,l\rangle$,
568: \begin{equation}
569:   \frac{\partial E_{N_{\rm e} + 1,k}}{\partial B}
570:   = \pm \frac{1}{2} g_{k0} \mu_B,\ \
571:   \mbox{$N_{\rm e}$ even}.
572: \end{equation}
573: In the remainder of this paper we continue to refer to the
574: even-electron ground state as the ``$N_e$-electron ground state''
575: and to the odd-electron states as the ``$(N_e+1)$-electron states''.
576: 
577: An example showing how the various degeneracies are lifted by
578: the exchange interaction and by a magnetic field is shown in Fig.\
579: \ref{fig:1}. The top panel of the 
580: figure shows the four lowest many-electron states
581: for an odd number of electrons, for the specific case that the
582: ground state has spin $S=1/2$ and the lowest excited state has
583: $S=3/2$ in the absence of spin-orbit scattering. Without spin-orbit
584: scattering, the $S=3/2$ state is fourfold degenerate. Spin-orbit
585: scattering lifts the fourfold degeneracy of the $S=3/2$ quadruplet,
586: separating it into two doublets with ill-defined spin (lower panel
587: of Fig.\ \ref{fig:1}, center). Finally, an applied magnetic field
588: lifts the degeneracy of all doublets. For an even electron number,
589: spin-orbit scattering lifts all possible degeneracies; to first
590: order in the field, the applied magnetic field has no effect.
591: 
592: The definition that $g$ factors are derivatives of energy levels to
593: the magnetic field implies that the Zeeman energy scale is the
594: smallest nonzero energy scale in the problem.  In particular, it is
595: smaller than the spin-orbit induced splittings of high-spin states
596: (see Fig.\ \ref{fig:1}), these splittings being proportional to
597: the spin-orbit scattering rate $\lambda^2 \propto
598: 1/\tau_{\rm so}$. However, when
599: $g$ factors are calculated without spin-orbit scattering, the Zeeman
600: energy is (by definition) 
601: larger than the spin-orbit rate. We'll find below that the
602: two limits do not commute in the presence of electron-electron
603: interactions, {\em i.e.,} that $g$ factor distributions calculated in
604: the limit $\lambda \to 0+$ 
605: are different from the $g$ factors at
606: $\lambda=0$. (Without spin-orbit scattering all peaks in the
607: differential conductance split with $g$ factor $g=2$.)  It should be
608: pointed out that experiment involves finite magnetic fields, for which
609: the Zeeman energy can be larger than the spin orbit rate. For such
610: magnetic-field dependences of the levels, a different slope at zero
611: field can easily go unnoticed.
612: 
613: Although $g$ factors contain information on the {\em positions} 
614: of peaks in the tunneling spectrum of the nanoparticle, knowledge of the 
615: {\em sizes} of the
616: peaks in the differential conductance is important for a correct
617: interpretation of the results. Peak heights contain information that
618: would be formulated in terms of ``selection rules'' in the absence
619: of spin-orbit scattering. Indeed, without spin-orbit scattering, the
620: total spin $S$ and its $z$ component $S_3$ can change by $1/2$ at
621: most in a tunneling process, which limits the possible transitions
622: between $N_e$ and $N_e+1$ electron states, and, hence, the possible
623: locations of peaks in the differential conductance. It is because of
624: the selection rules that one does not observe $g$ factors larger than
625: two, despite the fact that there exist high-spin many-electron states.
626: Similarly, without interactions, the occupation of single-electron
627: levels cannot change by more than one electron, which also limits the
628: number of allowed transitions in tunneling spectroscopy.
629: With spin-orbit
630: scattering and interactions, peaks that were previously ``forbidden'' 
631: are present, in principle, although their height may be small. We
632: return to this issue in more detail in Sec.\ \ref{sec:simulations}.
633: 
634: The example in Fig.\ \ref{fig:2} may further clarify the role of 
635: selection rules. The figure shows two possible $N_e$-electron
636: ground states (left) and three $(N_e+1)$-particle excited states
637: (right).
638: Without spin-orbit scattering or without exchange interactions,
639: the states (c) and (d) can be accessed from the $N_e$-particle
640: ground state (a), but not state (e). Similarly, state (e) can
641: be accessed from ground state (b), but not (c) and (d). When
642: spin-orbit scattering and exchange interactions are both present,
643: the $N_e$-electron ground state is a superposition of the states
644: (a) and (b) and all possible transitions have a finite matrix 
645: element. Notice that, 
646: since the energy difference between states (a) and (b)
647: is typically small --- on average, equal to $\delta - 2 J$, 
648: mixing of these two states is strong for spin-orbit scattering
649: rates $\lambda \sim 1$.
650: 
651: \begin{figure}[t]
652: \epsfxsize= 0.9\hsize
653: \hspace{0.1\hsize}
654: \epsffile{various_cases.eps}
655: %\vskip-0.55cm
656: \caption{ \label{fig:2}
657: Left: Schematic representation of ground states for an even number of 
658: electrons without spin-orbit scattering. Depending on the strength
659: of the exchange interaction, the $N_e$-electron ground state may
660: have spin $S=0$ (a) or $S=1$ (b). Right: three excited
661: $(N_e+1)$-electron states.}
662: \end{figure}
663: 
664: \subsection{Matrix elements of interaction Hamiltonian}
665: 
666: For the construction of the interaction Hamiltonian in the basis of 
667: many-electron eigenstates of the Hamiltonian $H_0$, one needs 
668: explicit equations for the matrix elements of the exchange
669: interaction in that basis. Since the exchange interaction changes
670: the single-electron states of at most two electrons,
671: the only nonzero matrix elements of the exchange interaction
672: occur between states that can be written in the
673: form 
674: $\hat \psi^{\dag}_{\mu\alpha} \hat \psi^{\dag}_{\nu\beta}|F\rangle$
675: and 
676: $\hat \psi^{\dag}_{\mu'\alpha'} \hat
677: \psi^{\dag}_{\nu'\beta'}|F\rangle$,
678: where $|F\rangle$ a certain reference non-interacting state. After
679: some algebra one then finds
680: \begin{widetext}
681: \begin{eqnarray}
682:   \lefteqn{
683:   \langle F|\hat \psi_{\nu'\beta'} \hat
684:   \psi_{\mu'\alpha'}  \vS^2 
685:   \hat \psi^{\dag}_{\mu\alpha} \hat \psi^{\dag}_{\nu\beta}|F\rangle }
686:   \nonumber \\
687:   &=& \vphantom{\sum_{\gamma \in F}}
688:   \left
689:   (\delta_{{\mu\alpha},{\mu'\alpha'}}\delta_{{\nu\beta},{\nu'\beta'}}
690:   - 
691:   \delta_{{\nu\beta},{\mu'\alpha'}}\delta_{{\mu\alpha},{\nu'\beta'}}\right )
692:   \langle F| \vS^2 | F \rangle 
693: %  \nonumber \\ && \mbox{} 
694:   - 2 \left [
695: {\bf s}_{{\nu'\beta'}{\mu\alpha}}\cdot{\bf s}_{{\mu'\alpha},{\nu\beta}} -
696: {\bf s}_{{\mu'\alpha'},{\mu\alpha}}\cdot{\bf s}_{{\nu'\beta'},{\nu\beta}}
697: %\langle c |{\bf s}| b \rangle \cdot
698: %\langle d |{\bf s}| a \rangle -
699: %\langle d |{\bf s}| b \rangle \cdot
700: %\langle c |{\bf s}| a \rangle
701:   \right ]  
702:   \nonumber\\ && \mbox{} 
703:   - 2 \delta_{{\nu\beta},{\nu'\beta'}} 
704: \sum_{{\phi,\gamma}\in { F}}
705: \left [ 
706: {\bf s}_{{\mu'\alpha'},{\phi\gamma}}\cdot{\bf s}_{{\phi\gamma},{\mu\alpha}} -
707: {\bf s}_{{\mu'\alpha'},{\mu\alpha}}\cdot{\bf s}_{{\phi\gamma},{\phi\gamma}}
708: %\langle d    | {\bf s}| \phi \rangle \cdot 
709: %\langle \phi |{\bf s}| b \rangle - 
710: %\langle d    |{\bf s}| b \rangle \cdot
711: %\langle \phi |{\bf s}| \phi \rangle  
712:   \right ] 
713: %  \nonumber \\ && \mbox{} 
714:   - 2\delta_{{\mu\alpha},{\nu'\beta'}} 
715: \sum_{{\phi,\gamma}\in { F}}
716: \left [ 
717: {\bf s}_{{\mu'\alpha'},{\nu\beta}}\cdot{\bf s}_{{\phi\gamma},{\phi\gamma}} -
718: {\bf s}_{{\phi\gamma},{\nu\beta}}\cdot{\bf s}_{{\mu'\alpha'},{\phi\gamma}}
719: %\langle d    | {\bf s}| a \rangle \cdot 
720: %\langle \phi |{\bf s}| \phi \rangle - 
721: %\langle \phi    |{\bf s}| a \rangle \cdot
722: %\langle d |{\bf s}| \phi \rangle  
723:  \right ]  
724:   \nonumber\\ && \mbox{} - 
725:   2 \delta_{{\nu\beta},{\mu'\alpha'}} 
726: \sum_{{\phi,\gamma}\in { F}}
727: \left [ 
728: {\bf s}_{{\nu'\beta'},{\mu\alpha}}\cdot{\bf s}_{{\phi\gamma},{\phi\gamma}} -
729: {\bf s}_{{\phi\gamma},{\mu\alpha}}\cdot{\bf s}_{{\nu'\beta'},{\phi\gamma}}
730: %\langle c    | {\bf s}| b \rangle \cdot 
731: %\langle \phi |{\bf s}| \phi \rangle - 
732: %\langle \phi    |{\bf s}| b \rangle \cdot
733: %\langle c |{\bf s}| \phi \rangle  
734:   \right ]
735: %  \nonumber \\ && \mbox{}
736:   - 2 \delta_{{\mu\alpha},{\mu'\alpha'}} 
737: \sum_{{\phi,\gamma}\in { F}}
738: \left [ 
739: {\bf s}_{{\phi\gamma},{\nu\beta}}\cdot{\bf s}_{{\nu'\beta'},{\phi\gamma}} -
740: {\bf s}_{{\nu'\beta'},{\nu\beta}}\cdot{\bf s}_{{\phi\gamma},{\phi\gamma}}
741: %\langle \phi    | {\bf s}| a \rangle \cdot 
742: %\langle c |{\bf s}| \phi \rangle - 
743: %\langle c    |{\bf s}| a \rangle \cdot
744: %\langle \phi |{\bf s}| \phi \rangle  
745: \right ],
746: \nonumber \\
747: \label{matrix_element}
748: \end{eqnarray}
749: where
750: \be
751: \langle F | \vS^2 | F \rangle =
752:   \frac{3}{4}(N_e-2) -
753:   2 \sum_{{\phi,\gamma} \in F}
754:   \sum_{{\phi',\gamma'} \in F}
755:   \left [
756:   \vs_{\phi \gamma, \phi'\gamma'}\cdot{\bf s}_{\phi'\gamma',\phi \gamma }
757:    - 
758:   \vs_{\phi \gamma, \phi \gamma }\cdot{\bf s}_{\phi'\gamma',\phi'\gamma'} 
759:   \right ].
760: \label{diagonal_matrix_element}
761: \ee
762: \end{widetext}%
763: The summations over $\phi$ and $\gamma$ extend over all
764: single-electron states $|\phi,\gamma\rangle$, $\gamma=1,2$, that are
765: occupied in the state $|F\rangle$.
766: 
767: Similarly, matrix elements of the Zeeman energy $H_{\rm Z}$
768: are nonzero only if many-electron states differ by not more than
769: one electron, {\em i.e.,} between states of the form
770: $\hat \psi^{\dag}_{\mu\alpha}|F\rangle$ and 
771: $\hat \psi^{\dag}_{\nu\beta}|F\rangle$. For those states, one
772: needs the matrix elements
773: \be
774:   \langle F| \hat \psi_{\nu\beta} \vS
775:   \hat \psi^{\dag}_{\mu\alpha} |F\rangle
776:   =
777:   \vs_{\nu\beta,\mu\alpha} +
778:   \delta_{\nu\beta,\mu\alpha} \sum_{ \phi,\gamma \in F}
779:   \vs_{\phi\gamma,\phi\gamma}.
780: \label{magnetic_field_matrix_element}
781: \ee
782: 
783: At this point, it is important to verify the applicability of the
784: random matrix theory. In order for random matrix to apply, summations
785: over the Fermi sea should converge within a Thouless energy from the
786: Fermi level. In Eqs.\ (\ref{matrix_element}),
787: (\ref{diagonal_matrix_element}) and
788: (\ref{magnetic_field_matrix_element}), the sums of the form
789: $\sum_{\phi,\gamma \in F} \vs_{\phi\gamma,\phi\gamma}$ clearly satisfy
790: this condition, by virtue of the equality 
791: \be
792:   \vs_{\phi 1,\phi 1} + \vs_{\phi 2, \phi 2}
793:   = 0,
794: \label{identity}
795: \ee
796: which follows from the observation that the states $|\phi,1\rangle$ 
797: and $|\phi,2\rangle$ are time reversed. The sums
798: of the form $\sum_{\phi,\gamma \in F} \vs_{\mu\alpha,\phi\gamma} \cdot
799: \vs_{\phi\gamma,\nu\beta}$ also meet this condition, since the summand
800: decreases $\propto 1/(\varepsilon_F - \varepsilon_{\phi})^2$ for
801: $\phi$ well below the Fermi level and $\mu$ and $\nu$ close to the
802: Fermi level.\cite{kn:adam2002b} In the diagonal matrix element
803: (\ref{diagonal_matrix_element}) the sum over $(\phi,\gamma)$ and
804: $(\phi',\gamma')$ is logarithmically divergent as a function of the
805: Fermi energy. This, however, has no consequences for the
806: magnetic-field dependence of the many-electron states and the peak
807: heights, since the divergence is the for all matrix elements and
808: simply corresponds to the overall shift of the ground state energy. We
809: conclude that random matrix theory can be used to access the
810: many-electron ground state and the low-lying excited states.
811: 
812: \section{Results and discussion}
813: \label{sec:simulations}
814: 
815: In order to calculate $g$ factors and peak heights, we have 
816: diagonalized the Hamiltonian (\ref{hamiltonian}) numerically.
817: 
818: Our numerical procedure is as follows:
819: We first diagonalized the non-interacting Hamiltonian $H_0$. The
820: interaction is considered in a truncated basis of the many-electron
821: states, taking the 92 lowest lying many-electron eigenstates of
822: $H_0$ for a $N_e+1$ electrons ($N_e$ even), and the 76 lowest lying
823: states for $N_e$ electrons. Finally, we diagonalized
824: the truncated interaction and found the $g$ factors of the $M=8$
825: lowest-lying $(N_e+1)$-electron states, and the peak heights that
826: follow for transitions from the $N_e$-electron ground state. 
827: For the calculation of the $g$ factors we introduce a small 
828: magnetic field and calculate the magnetic field derivative
829: numerically. We
830: verified that truncating the interaction Hamiltonian at the 
831: lowest lying 92 and 76 many-electron states has no effect on the
832: final results by comparing our results to those that were obtained
833: using a smaller basis set.
834: 
835: We have investigated exchange-interaction strength ranging from
836: $J=0$ to $J = 0.6\spacing$ which is valid for most metals,
837: see Sec.~\ref{sec:experiment}.
838: The same parameter range should apply to quantum dots.
839: %As it could be seen from Fig.~\cite{periodic_table} the 
840: %exchange-interaction constant can vary significantly among
841: %various materials.
842: %Our choice for the ratio $J/\spacing$ is motivated by RPA estimates
843: %for various metals, which set $J/\spacing$ around
844: %$0.3$,\cite{kn:oreg2002} by an 
845: Analysis of the Coulomb blockade 
846: peak spacing distribution\cite{kn:usaj2003} suggests that
847: $J/\spacing$ is between $0.3$ and $0.4$
848: in large quantum dots in a GaAs/GaAlAs
849: heterostructure; whereas recent density functional
850: studies of ground state spin distributions in ballistic quantum
851: dots are compatible with the Hamiltonian (\ref{hamiltonian})
852: only if $J\approx 0.6\spacing$, see Ref.\ \onlinecite{kn:jiang2003}.
853: 
854: Important changes occur within the range of exchange interactions we
855: address here. Since the metal nanoparticle is assumed to relax to the
856: $N_e$-particle ground state between tunneling events, the
857: (statistical) properties of the $N_e$-particle ground state play a key
858: role in determining the $g$ factor distribution. Without spin-orbit
859: scattering for $J \lesssim 0.3\spacing$, there is only a small
860: probability that the $N_e$-particle ground state has spin one, and a
861: vanishing probability that the $N_e+1$-particle ground state has spin
862: $3/2$ ($N_e$ is assumed even).\cite{kn:brouwer1999b,kn:oreg2002} The
863: probability to find an $N_e$-particle ground state with $S=1$ becomes
864: appreciable for $J \gtrsim 0.3$, whereas the probability to find an
865: $(N_e+1)$-particle ground with spin $3/2$ becomes significant for $J
866: \gtrsim 0.5\spacing$ only. For the values of $J$ we consider, states
867: with spin $\ge 5/2$ do not play a role; they have been excluded
868: from the truncated many-electron basis.
869: 
870: The strength of spin-orbit scattering strength $\lambda$ is taken 
871: from $0$ to $2.8$. Although larger spin-orbit scattering strengths
872: do occur in metal nanoparticles,\cite{kn:petta2001,kn:petta2002} interaction
873: effects are small at those values of $\lambda$ and the non-interacting
874: theory of Refs.\
875: \onlinecite{kn:brouwer2000,kn:matveev2000,kn:adam2002b} works well.
876: 
877: The random matrices in our simulation are taken of size $2N=400$.
878: This ensures that the condition ${\lambda}^2 \ll N$ necessary for
879: the applicability of the random matrix (\ref{eq:H0}) is satisfied
880: for all values of $\lambda$. For $\lambda <2 $ we have taken $2N=200$
881: in the simulations. 
882: %For every value of $J$ and $\lambda$ we have
883: %performed an average over $N_s = 300$ realizations.
884: 
885: \subsection{Average $g$ factors}
886: \label{sec:threshold}
887: 
888: We have calculated ensemble averages of the $g$ factors $\langle
889: g_k\rangle$,
890: $k=0,1,\ldots,M-1$ of the $M$ lowest $(N_e+1)$-electron states.
891: Here $g_k$ is the $g$ factor corresponding to the $k$th
892: $(N_e+1)$-electron system, $k=0,1,\ldots,M-1$. The ensemble average
893: is taken over 300 realizations.
894: In Fig.~\ref{ground_state_g_factor} we show the ensemble averaged 
895: $g$ factors
896: for the ground state and the first excited state, $\langle g_0 \rangle$
897: and $\langle g_1 \rangle$, as well as the average over all
898: calculated $g$ factors $\langle \bar g \rangle$
899: \begin{equation}
900:   \bar g 
901:  = M^{-1} \sum_{k=0}^{M-1}  g_k .
902: \end{equation}
903: 
904: 
905: The first observation to be made from Fig.\
906: \ref{ground_state_g_factor} is that for $J \gtrsim 0.4\spacing$ and 
907: $\lambda \lesssim 2$ interactions 
908: lead to a significant increase in the average $g$ factor. In fact,
909: there is a significant parameter range for which $\langle \bar g 
910: \rangle > 2$. The origin of the large $g$ factors is that
911: exchange interactions lift the degeneracy with respect to the 
912: total spin $S$. Hence, with exchange interactions, many-electron
913: states with a finite spin are energetically separated from states
914: with spin $0$ or $1/2$. For the parameter range considered here,
915: the relevant nontrivial spin states have $S=3/2$ for an odd number
916: of electrons. The role of spin-orbit scattering is to lift the
917: fourfold degeneracy of the $S=3/2$ states and, for larger spin-orbit
918: strengths, to suppress the spin content of the single-electron states
919: that build the many-electron state. Let us first discuss the effect
920: of lifting the degeneracy of the $S=3/2$ state by spin orbit
921: scattering.
922: 
923: In general, spin orbit scattering splits the fourfold degenerate 
924: $S=3/2$ state into two doublets. Neglecting contributions from 
925: other many-electron states, each doublet consist of two states 
926: that can written in the form 
927: \be
928:   |\mbox{state}\rangle = \sum_{n=-3/2}^{3/2} a_{n}
929:   |3/2,n\rangle, \label{eq:doublet}
930: \ee      
931: and the time-reversed of Eq.\ (\ref{eq:doublet}), which is
932: obtained by sending $a_n \to \mbox{sign}(n) a_{-n}^*$.
933: Here $|S,S_3\rangle$ is the $(N_e+1)$-electron state with 
934: total spin $S$ and $z$ component
935: of the spin equal to $S_3$. Because the spin-orbit matrix elements
936: are random, the amplitudes $a_n$ are essentially random as well.
937: (This statement is verified in the appendix.)
938: The $g$ factor of the state (\ref{eq:doublet}) 
939: is
940: \begin{eqnarray}
941:   g^2 = \left( \sum_{n=-3/2}^{3/2} 4 n |a_n|^2 
942:   \right)^2 + \left| \sum_{n=-3/2}^{3/2} 4 |n| a_n a_{-n}
943:   \right|^2.
944: \end{eqnarray}
945: One easily verifies that this can be larger than two. 
946: %For transitions to a state of the form
947: %(\ref{eq:doublet}) to be visible in a tunneling spectroscopy 
948: %experiment, it is necessary that the corresponding peak height is
949: %nonzero. Here, again, it is important that exchange interaction and
950: %spin-orbit scattering are both present.
951: With exchange interaction but without spin-orbit scattering, 
952: there is a finite probability
953: that the $N_e$-particle ground state ($N_e$ even) has spin $S=1$.
954: In that case, it has two singly occupied orbitals and, hence, in
955: principle, a finite overlap with a state of the form
956: (\ref{eq:doublet}) after addition of an electron. With spin-orbit
957: scattering, the $N_e$-electron ground state is guaranteed to be
958: non-degenerate, so that its derivative to the magnetic field is 
959: zero. We conclude that, the $g$ factor of the state (\ref{eq:doublet}) 
960: can be larger that two, that it can correspond to a transition between
961: the $N_e$-electron ground state and an $(N_e+1)$-electron state,
962: and that the corresponding
963: conductance peak has a finite height.
964: 
965: A finite amount of spin-orbit scattering stabilizes the above arguments
966: by increasing the splittings between many-electron states that are
967: degenerate in the absence of spin-orbit coupling. On the other hand,
968: with moderate
969: spin-orbit scattering, more many-electron states are added in 
970: the doublet (\ref{eq:doublet}). This has two consequences: (1) the the
971: spin content of each of the underlying single-electron states is
972: reduced, which, eventually, leads to a suppression of $g$ factors, and
973: (2) when more many-electron states are admixed, overlaps and, hence,
974: peak heights are increased, so that the role of selection rules is
975: further diminished. In order to illustrate the value of $\lambda$
976: needed to admix different many-electron states, we note that for the
977: $N_e$-particle ground state without spin-orbit interaction, the energy
978: separation between the $S=1$ and $S=0$ ground states is $2J-\delta$ on
979: average. Hence, even for a relatively small spin-orbit scattering rate
980: $\lambda \sim 0.5$ the ground state with spin-orbit scattering will
981: have significant weight in both of these states.
982: 
983: Also notice that, unlike in the non-interacting case, the average $g$ 
984: factor
985: depends on the excitation energy of the many-electron state for 
986: $\lambda \lesssim 2$, see Fig.\ \ref{ground_state_g_factor}.
987: The origin of this dependence is that, without
988: spin-orbit scattering, the probability of that an
989: $(N_e+1)$-particle state has nontrivial spin ($S \ge 3/2$) increases
990: with the excitation energy. As discussed above, although 
991: spin-orbit scattering lifts the fourfold degeneracy of these states
992: and suppresses the spin, it is the underlying nontrivial spin
993: character persisting to finite $\lambda$
994: that gives rise to the increased $g$ factors.
995: \begin{figure}
996: \epsfxsize= 0.75\hsize
997: %\hspace{0.1\hsize}
998: \epsffile{panel_1_1.eps}
999: 
1000: \epsfxsize= 0.75\hsize
1001: %\hspace{0.1\hsize}
1002: \epsffile{panel_1_2.eps}
1003: 
1004: \epsfxsize= 0.75\hsize
1005: %\hspace{0.1\hsize}
1006: \epsffile{panel_1_3.eps}
1007: \caption{Ensemble averaged $g$-factors.
1008:  Top: average $g$ factor
1009: $\langle g_0 \rangle$ of $(N_e+1)$-particle ground state for
1010: $J = 0$ (the lowest solid curve), $J = 0.3\spacing$, $J = 0.4\spacing$,
1011: $J = 0.5\spacing$, and $J = 0.6\spacing$
1012: Middle:
1013: average $g$ factor $\langle g_1 \rangle$ of first 
1014: excited $(N_e+1)$-particle state for $J/\delta = 0.1, 0.2, 0.3, 0.4, 0.5$
1015: and 0.6.
1016: Bottom: ensemble averaged $g$
1017: factor averaged over the first $M=8$ $(N_e+1)$-electron states
1018: for  $J/\delta = 0.1, 0.2, 0.3, 0.4, 0.5$ and 0.6.
1019: } 
1020: \label{ground_state_g_factor}
1021: \end{figure}
1022: 
1023: A remarkable feature of Fig.\ \ref{ground_state_g_factor} is that
1024: the ensemble averaged $g$ factor $\langle \bar g \rangle$
1025: does not approach two in the limit
1026: $\lambda \to 0$. On the other hand, without spin-orbit scattering,
1027: all observed $g$ factors should equal $g=2$. In Sec.\ 
1028: \ref{sec:theoretical_description}
1029: we discussed why there can be a difference between $g$ factors in
1030: the limit $\lambda \to 0$ and $g$ factors calculated without
1031: spin orbit scattering, {\em i.e.,} at $\lambda=0$. In fact, in Fig.\
1032: \ref{ground_state_g_factor}, $\langle \bar g \rangle$ is 
1033: overestimated for $\lambda \to 0$, because the plain ensemble average 
1034: does not take tunneling spectroscopy peak heights or ``selection 
1035: rules'' into
1036: account: The average is taken over all $(N_e+1)$-particle states
1037: irrespective of the height of the corresponding peak in the 
1038: differential conductance. In particular for small $J$, one 
1039: expects that $(N_e+1)$-electron states with $g$ factors larger
1040: than two are likely to
1041: have small tunneling spectroscopy peak heights.
1042: 
1043: \begin{figure}
1044: \epsfxsize= 0.75\hsize
1045: %\hspace{0.1\hsize}
1046: \epsffile{panel_2_1.eps}
1047: \epsfxsize= 0.75\hsize
1048: %\hspace{0.1\hsize}
1049: %\epsffile{panel_2_2.eps}
1050: \epsffile{PRB_plot_with_inset.eps}
1051: \caption{Top: average of all calculated $g$ factors, where each 
1052: $g$ factor is weighed by its normalized peak height
1053:   (\protect\ref{eq:wg}). Bottom: average of all calculated $g$
1054: factors for which the normalized peak height is larger than $0.1$.
1055: In both panels, results are shown for $J/\spacing=0$, $0.1$, $0.2$,
1056: $0.3$, $0.4$, $0.5$, and $0.6$.
1057: For $J=0$ there is no difference with Fig.\
1058: \protect\ref{ground_state_g_factor}.
1059: Inset of lower panel: the probability for a level
1060: to be visible in the experiment, 
1061: i.e. to have a weight larger than the threshold
1062: one, see 
1063: (\protect\ref{weight_criterion}).} 
1064: \label{weighted_average_g_factors}
1065: \end{figure}
1066: 
1067: 
1068: 
1069: 
1070: 
1071: We have taken two different approaches in order to account for 
1072: ``selection rules''. First, we have replaced
1073: the ensemble average by a ``weighed'' average, in which every 
1074: $g$ factor is weighed by the normalized peak height
1075: \begin{eqnarray}
1076:   \langle \bar g \rangle_{\rm w} &=&
1077:   \left\langle \sum_{k=0}^{M-1} \tilde w_k g_k \right\rangle,
1078:   \ \
1079:   \tilde w_k = \frac{M w_k}{\sum_{k=0}^{M-1} w_k}.
1080:   \label{eq:wg}
1081: \end{eqnarray}
1082: In the second approach, we have removed all peaks with normalized
1083: weight $\tilde w_k$ below a certain threshold value,
1084: where we arbitrarily set the threshold to $\tilde w_{\rm t} =
1085: 0.1\times \max^{M}_{k=1} w_k$,
1086: \begin{equation}
1087:   \tilde w_k \to \tilde w_{k,{\rm t}} =
1088:   \left\{ \begin{array}{ll} 1 & \mbox{if $ w_k
1089:   \ge w_{\rm t}$}, \\
1090:   0 & \mbox{if $w_k < w_{\rm t}$}. \end{array} \right.
1091: \label{weight_criterion}
1092: \end{equation}
1093: In this method, the number of levels per realization depends on the 
1094: realization,
1095: \begin{equation}
1096:   M_{\rm t} = \sum_{k} \tilde w_{k,{\rm t}}
1097: \end{equation}
1098: and the average $g$-factor is determined through
1099: \begin{equation}
1100:  \langle \bar g \rangle = \left \langle \frac{1}{M_t}
1101: \sum_{k=0}^{M-1} \tilde w_{k,t} g_k\right \rangle .
1102: \end{equation}
1103: The threshold mimics the experimental reality that small peaks cannot
1104: be distinguished from the noise, and, hence, have their $g$ factors
1105: left out in the statistical analysis. While the second method is
1106: closer to the way experiments are analyzed --- all $g$ factors of
1107: conductance peaks that are observed are taken equally into account
1108: in the average --- the first method has the advantage that it does
1109: not contain the somewhat arbitrary threshold at $\tilde w_k = \tilde
1110: w_{\rm t} = 0.1$. 
1111: Both methods enforce the ``selection rules'' in the absence of
1112: spin-orbit scattering. They also give almost identical results for
1113: the average $g$ factor, as is seen from Fig.\
1114: \ref{weighted_average_g_factors} where we show the weighted
1115: ensemble average of the $g$ factors of all levels considered 
1116: $\langle \bar g \rangle$, as well as the ensemble average
1117: calculated using the ``threshold'' method.
1118: As shown by comparison of Figs.\ \ref{weighted_average_g_factors}
1119: and \ref{ground_state_g_factor}, in the limit $\lambda \to 0$,
1120: the average $g$ factors are close to two for small $J$, whereas
1121: $\langle \bar g \rangle$ is significantly higher than two for
1122: $J \gtrsim 0.3$. In the inset of the lower panel of 
1123: Figs.\ \ref{weighted_average_g_factors} we show the probability
1124: of the level to be visible, i.e. to have a peak height above
1125: the threshold. Remarkably, the curves for $J\agt 0.3\delta$ 
1126: have a maximum for moderate values of spin-orbit
1127: scattering $\lambda\simeq 1$. This has a direct physical interpretation:
1128: both for small and large $\lambda$, 
1129: approximate selection rules are in place. For small $\lambda$
1130: these selection rules represent the conservation of spin 
1131: at $\lambda = 0$, whereas for large $\lambda$ they follow from
1132: the suppression of the (exchange) interaction, which causes the remaining
1133: physics to be single-particle like.
1134: 
1135: 
1136: \subsection{fluctuations of $g$ factors}
1137: 
1138: \begin{figure}
1139: \epsfxsize= 0.75\hsize
1140: %\hspace{0.1\hsize}
1141: \epsffile{cumulative_distribution.eps}
1142: \caption{Cumulative probability distributions function for $g$-factors
1143: for the cases $\lambda = 0.7$, $J=0$ (dashed curve)
1144: $\lambda = 0.7$, $J=0.3$ (solid), and $\lambda=0.9$, $J=0.3$
1145: (dotted). Even a relatively
1146: small strength of the exchange interaction is enough in order
1147: to broaden the distribution function significantly.
1148: } 
1149: \label{probability_distribution}
1150: \end{figure}
1151: Cumulative probability distributions of $g$ factors are shown in
1152: Fig.\ \ref{probability_distribution} 
1153: for $J=0$, $\lambda=0.7$, for $J=0.3\spacing$,
1154: $\lambda=0.7$ and for $J=0.3\spacing$, $\lambda=0.9$
1155: (we have taken into account only peaks whose weights are nonzero
1156: according to the criterion (\ref{weight_criterion})). 
1157:  Comparing
1158: the two distributions at $\lambda=0.7$, one notices that the
1159: exchange interaction has little effect on the tail of the $g$ factor
1160: distribution for very small $g$ factors. However, for larger
1161: $g$ factors, the weight of the probability distribution is shifted
1162: towards larger $g$ factors, including a long tail in the region
1163: $g > 2$. Fig.\ \ref{probability_distribution} confirms the 
1164: previous observation that the effect of exchange interactions is to 
1165: increase the average $g$ factors. The spin-orbit scattering rate
1166: for the third probability distribution shown in Fig.\ 
1167: \ref{probability_distribution}, 
1168: $\lambda=0.9$, has been chosen such that the average $g$ factor
1169: $\langle {\bar g} \rangle_t \approx 1.58$
1170:  coincides
1171: with that of the case $J=0$, $\lambda=0.7$. Comparing the two
1172: probability distributions, we conclude that the interactions 
1173: still
1174: lead to a significant increase of the $g$ factor fluctuations, 
1175: including a large probability to find $g$ factors larger than two,
1176: even if the average is well below two.
1177: 
1178: \begin{figure}
1179: \epsfxsize= 0.75\hsize
1180: %\hspace{0.1\hsize}
1181: \epsffile{prob_g_larger_2_relevant.eps}
1182: 
1183: \epsfxsize= 0.75\hsize
1184: %\hspace{0.1\hsize}
1185: \epsffile{prob_g_larger_2.eps}
1186: 
1187: 
1188: \caption{
1189: Top: Probability for the $g$ factor of a level to
1190: be larger than $2.0$ for the values of the exchange constant
1191: $J=0.1\div 0.6\delta$ as a function of the spin-orbit scattering
1192: strength $\lambda$
1193: for a random visible (i.e., satisfying the threshold criterion,
1194: see  Sec.\ \ref{sec:threshold}) level.
1195: Bottom: Probability for a random level to be visible and,
1196: at the same time, have $g>2$. 
1197: for the values of the exchange constant
1198: $J=0.1\div 0.6\delta$ as a function of the spin-orbit scattering
1199: strength $\lambda$.
1200: } 
1201: \label{prob_g>2}
1202: \end{figure}
1203: 
1204: In Fig.~\ref{prob_g>2} we show the probability for a level to
1205: have a $g$-factor larger than two. In an experiment, typically
1206: $g$ factors of 5-10 consecutive levels can be
1207: measured.\cite{kn:petta2001,kn:petta2002} From Fig.\ \ref{prob_g>2}
1208: we then conclude that there is a significant probability that 
1209: one of these $g$ factors is larger than two if $J \agt 0.2 \spacing$.
1210: The bottom panel shows the probability that a level $|k\rangle$ has 
1211: a $g$ factor larger than two {\em and} a weight $\bar w_k > \bar
1212: w_{\rm t}$. As a result of the breakdown of selection rules, this 
1213: probability {\it increases} with increasing spin-orbit scattering
1214: in the region $\lambda\alt 0.5$: States
1215: which have large $g$-factors but small weights for small $\lambda$
1216: become visible for larger values of $\lambda$. The ratio of the
1217: probability shown in the bottom plot of Fig.~\ref{prob_g>2} and
1218: that in the top plot of Fig.~\ref{prob_g>2} is the probability
1219: that a random level can be resolved in the experiment, see Fig.\
1220: \ref{weighted_average_g_factors}
1221: 
1222: 
1223: In addition to addressing the full probability distribution of $g$ factors,
1224: we should consider the possibility of correlations between
1225: $g$ factors within one realization. In principle, such correlations
1226: can exist, because, within one realization, all $g$ factors 
1227: correspond to transitions from
1228: the same $N_e$-electron ground state ($N_e$ even). Although the
1229: $N_e$-electron ground state does not affect the values of all
1230: possible $g$ factors for the $(N_e+1)$-particle levels, it does
1231: affect the peak heights, and hence determines which $g$ factors
1232: possibly ``drown'' in the noise.
1233: 
1234: In order to quantify $g$ factor correlations, we have looked at
1235: the correlation function
1236: \begin{eqnarray}
1237:   C(J,\lambda) &=&
1238:   \left\langle \sum_{k \neq l} \frac{1}{M_{\rm t}^2}
1239:   (g_k - \langle \bar g \rangle)(g_l - \langle \bar g \rangle)
1240:   \right\rangle.
1241:   \label{main_formula_1}
1242: \end{eqnarray}
1243: For the calculation of the correlation function $C(J,\lambda)$
1244: we removed all $g$-factors with normalized weight $\tilde w_k$
1245: below the threshold value $\tilde w_{\rm t} = 0.1$ from the average,
1246: which means that the number levels $M_{\rm t}$ considered in the 
1247: summation becomes dependent on the actual realization.
1248: For the range of exchange interactions $J$ and spin-orbit
1249: scattering rates $\lambda$ we considered, the correlation function
1250: $C(J,\lambda)$ was nonzero, but always smaller than $0.1$. The 
1251: maximum value $C(J,\lambda) \sim 0.1$ was obtained for 
1252: $\lambda \sim 0.5$. Comparing the above difference with
1253: the typical variance of $g$-factors, see Fig.\
1254: \ref{probability_distribution}, we conclude that
1255: correlations between different levels within the same
1256: nanoparticle are not important if the number of $g$ factors measured
1257: in a single metal nanoparticle does not exceed $10$.
1258: 
1259: \subsection{Probability of nontrivial ground state}
1260: 
1261: In Fig.~\ref{probability} we show the probability that the
1262: metal nanoparticle is found in the non-interacting ground state
1263: (Fermi sea), as a function of $\lambda$ and $J$. For $N_e$-electron
1264: states, the probability to find the nanoparticle in the non-interacting
1265: ground state deviates quite significantly from 1 if $J \agt 
1266: 0.4$ and $\lambda=0$.\cite{kn:brouwer1999b,kn:baranger2000} 
1267: Upon increasing
1268: $\lambda$, the probability to be in the non-interacting
1269: ground state increases and approaches unity when the spin-orbit
1270: scattering rate exceeds the exchange interaction.
1271: \begin{figure}
1272: \epsfxsize= 0.75\hsize
1273: %\hspace{0.1\hsize}
1274: \epsffile{gs_probability.eps}
1275: \caption{Probability to be in the non-interacting ground state
1276: for $J=0.3-0.6\spacing$. The horizontal line $P=1$ corresponds
1277: to the case $J=0$.
1278: } 
1279: \label{probability}
1280: \end{figure}%
1281: 
1282: 
1283: \section{Material dependence}
1284: \label{sec:experiment}
1285: 
1286: Petta and Ralph have measured the probability
1287: distribution of $g$-factors of Cu, Au, and Ag 
1288: nanoparticles.\cite{kn:petta2001,kn:petta2002}
1289: Theoretical estimates and experimental investigations 
1290: of the exchange interaction in the noble metals show that 
1291: $J/\spacing < 0.1$ for Cu, Au, and 
1292: Ag.\cite{kn:macdonald1982,kn:vier1984}
1293: Hence, the interaction effects in the above materials are very small 
1294: and it is natural that the existing experiments can be explained
1295: quantitatively using theory for the noninteracting ($J=0$) case.
1296: Indeed, both the average and the width of the  $g$ factor probability 
1297: distribution measured in Refs.\ \onlinecite{kn:petta2001,kn:petta2002} 
1298: were found to be in good agreement with the non-interacting
1299: theory of Refs.\
1300: \onlinecite{kn:brouwer2000,kn:matveev2000} using a single fit
1301: parameter, the dimensionless spin-orbit scattering rate $\lambda$. 
1302: The spin-orbit scattering time used in the fits was in 
1303: order-of-magnitude agreement with previous measurements using
1304: weak localization.\cite{kn:bergmann1984} The observation that the
1305: width of the distribution agreed well with theory after $\lambda$
1306: has been chosen to fit the average was considered a success for 
1307: random matrix theory.\cite{foot1}
1308: 
1309: %\begin{figure}[t]
1310: %\epsfxsize= 0.75\hsize
1311: %%\hspace{0.1\hsize}
1312: %\epsffile{theory_and_experiment.eps}
1313: %\caption{
1314: %Cumulative $g$-factor distribution
1315: %for $\lambda = 0.70$, $J=0$ (solid curve), 
1316: %$\lambda = 0.85$, $J=0.2\spacing$ (points), 
1317: %$\lambda=0.90$, $J=0.3\spacing$ (dotted), 
1318: %$\lambda=1.0$,d and $J=0.4\spacing$ (dash-dot),
1319: %and $\lambda=1.15$, $J=0.6\spacing$ (dashed).
1320: %The values of $\lambda$ are chosen such that
1321: %${\langle {\bar g} \rangle}_t =1.58$ in all cases. 
1322: %}
1323: %\label{probability_distribution_exp}
1324: %\end{figure}
1325: 
1326: Significant deviations from the non-interacting theory can
1327: be expected for $J/\spacing \gtrsim 0.2$ only. Although this
1328: condition is not satisfied for the noble metals, the exchange
1329: interaction is strong enough to significantly
1330: affect the $g$ factor distribution in most other metals, see 
1331: Fig.\ \ref{periodic_table}, where a list of values of 
1332: $J/\spacing$ reported in the experimental and theoretical literature
1333: is given. The exchange interaction is particularly strong in 
1334: Sc, V, Y, Nb, Rh, and, especially, in Pd (Pd is very close to the 
1335: Stoner instability $J/\delta = 1$).
1336: 
1337: %**********************************************************************
1338: %**********************************************************************
1339: \begin{figure*}
1340: \epsfxsize=0.9\hsize
1341: %\hspace{0.05\hsize}
1342: \epsffile{periodic_table.eps}
1343: \caption{Ratio $J/\spacing$ of the exchange interaction constant
1344: and the mean level spacing.
1345: No value is listed for metals from $Z = 24$ to 28 as they are magnetically
1346: ordered as well as all lanthanides ($Z=58\dots 71$) except for
1347: Pm ($Z=61$) for which no value was found in the literature.
1348: Also, for $Z>87$ no data was found in the literature. 
1349: %However, due to the large nucleus charge, it is unlikely that effects 
1350: %of interaction
1351: %could be seen in nanoparticles made of these materials.
1352: For metals in the right-hand side of the periodic table
1353: (12th column and further) only data on their Pauli susceptibility 
1354: in the liquid form are available, see Ref. \onlinecite{kn:dupree1971}.
1355: For other metals, the data are taken from:
1356: Li, experiment;\cite{kn:vier1984} 
1357: Be, calculation of electronic 
1358:   structure;\cite{kn:wilk1978,kn:janak1977}
1359: Al, experiment\cite{kn:dunifer1977} 
1360:   (note the negative value of the exchange constant);
1361: K, experiments;\cite{kn:knecht1975,kn:randles1972}; 
1362: Ca, Y, Tc, Ba, La, Ta, W, Re, Os, Ir,
1363:   calculation of electronic structure;\cite{kn:sigalas1994}
1364: Sc, calculation of electronic 
1365:   structure;\cite{kn:macdonald1977a,kn:sigalas1994}
1366: Ti, Zr, Hf, calculation of electronic 
1367:   structure;\cite{kn:bakonyi1993,kn:sigalas1994}
1368: V, fit of theory\cite{kn:stenzel1986} and 
1369:   experiment,\cite{kn:hechtfischer1976}, and
1370:   calculation of electronic structure;\cite{kn:janak1977,kn:sigalas1994}
1371: Cu, Ag, experiment \cite{kn:vier1984} and calculation of 
1372:   electronic structure\cite{kn:macdonald1982} (note the negative 
1373:   exchange constant in the experiment for Cu);
1374: Rb, Cs, experiment;\cite{kn:knecht1975}
1375: Sr, Nb, Mo, Rh, calculation of electronic
1376:   structure;\cite{kn:janak1977,kn:sigalas1994}
1377: %Y, calculation of the electronic structure;\cite{kn:sigalas1994}
1378: %Zr, calculation of electronic 
1379: %  structure;\cite{kn:bakonyi1993,kn:sigalas1994};
1380: %Nb, calculation of electronic 
1381: %  structure;\cite{kn:janak1977,kn:sigalas1994}
1382: %Mo, calculation of electronic 
1383: %  structure;\cite{kn:janak1977,kn:sigalas1994}
1384: %Tc, calculation of electronic structure;\cite{kn:sigalas1994}
1385: %Rh, calculation of electronic 
1386: %  structure;\cite{kn:janak1977,kn:sigalas1994}
1387: Pd, fit of theory\cite{kn:stenzel1986} and
1388:   experiment;\cite{kn:manuel1963}
1389: %Ag, calculation of electronic structure\cite{kn:macdonald1982} 
1390: %  and experiment;\cite{kn:vier1984}
1391: %Cs, experiment;\cite{kn:knecht1975}
1392: %Ba, calculation of electronic structure;\cite{kn:sigalas1994}
1393: %La,  calculation of electronic structure;\cite{kn:sigalas1994}
1394: %Hf,  calculation of electronic 
1395: %  structure;\cite{kn:bakonyi1993,kn:sigalas1994}
1396: %Ta, calculation of electronic structure;\cite{kn:sigalas1994}
1397: %W,  calculation of electronic structure;\cite{kn:sigalas1994}
1398: %Re, calculation of electronic structure;\cite{kn:sigalas1994}
1399: %Os, calculation of electronic structure;\cite{kn:sigalas1994}
1400: %Ir, calculation of electronic structure;\cite{kn:sigalas1994}
1401: Pt, fit of theory\cite{kn:fradin1975} and 
1402:   experiment,\cite{kn:budworth1960} and calculation of 
1403:   electronic structure;\cite{kn:sigalas1994}
1404: Au, calculation of electronic structure.\cite{kn:macdonald1982}
1405: \label{periodic_table}
1406: }
1407: \end{figure*}
1408: %**********************************************************************
1409: %**********************************************************************
1410: 
1411: For the collection of the data shown in Fig.\ \ref{periodic_table},
1412: we used the fact that the ratio $J/\delta$ is related to the Fermi 
1413: liquid parameter $F_0^a$
1414: \be
1415:   J/\spacing = -F_0^a,
1416: \ee
1417: see Ref.\ \onlinecite{kn:oreg2002}. The Fermi liquid parameter
1418: $F_0^a$ appears in the expression for the paramagnetic susceptibility
1419: \be
1420:  \chi = \chi_0 \frac{m^*/m}{1 + F_0^a},
1421: \ee
1422: where $m^*$ is the effective electron mass, including band
1423: structure effects and interaction effects, $m$ is the free
1424: electron mass, and $\chi_0$ is the Pauli susceptibility for free
1425: electrons,\cite{kn:pines1966}
1426: \be
1427:   \chi_0 = \frac{\mu_B^2 m p_F}{\pi^2 \hbar^3}.
1428: \ee
1429: %The ratio $m^*/m$ can be found from a measurement of the
1430: %specific heat. 
1431: The parameter $1/(1+F_0^a)$ is also known as the
1432: Stoner enhancement parameter.
1433: 
1434: The dimensionless spin-orbit scattering parameter
1435: $\lambda$ increases with element's nucleus charge $Z$.
1436: In experiments of Petta and Ralph\cite{kn:petta2001} 
1437: strong spin-orbit scattering was found for Au nanoparticles of
1438: a few nm in diameter ($\lambda \sim 0.1$), whereas moderate 
1439: spin-orbit scattering strengths ($\lambda \simeq 1$) were 
1440: found in Cu and Ag nanoparticles of roughly equal size. From
1441: this, we conclude that moderate spin-orbit scattering strengths
1442: can be expected for nanoparticles with a value of $Z$ around those 
1443: for Cu or Ag. From Fig.\ \ref{periodic_table} it can be seen 
1444: that there are quite a few materials for which
1445: this is true and the criterion $J/\delta > 0.2$ is satisfied.
1446: 
1447: 
1448: 
1449: 
1450: 
1451: 
1452: 
1453: 
1454: 
1455: \section{Conclusion}
1456: \label{sec:conclusion}
1457: 
1458: In this work we investigated the combined influence of 
1459: electron-electron interactions and spin-orbit scattering on the
1460: $g$ factors of metal nanoparticles. In the presence of
1461: electron-electron interactions, $g$ factors must be attributed
1462: to (transitions between) many-electron states, instead of
1463: single-electron states. Many-electron states can have $g$ factors
1464: larger than two, although these cannot be observed by tunneling 
1465: spectroscopy because of
1466: ``selection rules'' as long as spin-rotation symmetry is present. 
1467: Spin-orbit scattering breaks spin-rotation symmetry and thus
1468: removes the ``selection rules''. While this leads to a suppression
1469: of the $g$ factors for large spin-orbit scattering rates, we find 
1470: that $g$ factors larger than two occur with significant probability
1471: if the spin-orbit scattering rate $1/\tau_{\rm so}$ is moderate, 
1472: $\tau_{\rm so} \spacing \lesssim 1$, where $\spacing$ is the mean
1473: spacing between single-electron energy levels in the nanoparticle. 
1474: We have studied the $g$ factor distribution quantitatively
1475: using random matrix theory and the universal interaction
1476: Hamiltonian.\cite{kn:kurland2000,kn:aleiner2002} In addition
1477: to a confirmation of the scenario outlined above --- occurrence
1478: of $g$ factors larger than two --- we found that interactions
1479: increase the width of the $g$ factor distribution for 
1480: $\tau_{\rm so} \spacing \lesssim 1$ and 
1481: that the $g$ factors probability distribution function 
1482: is different for transitions to the odd-electron ground state and
1483: odd-electron excited states.
1484: %The broad $g$ factor distributions we find are at odds with recent
1485: %experiments which found a narrower $g$ factor
1486: %distribution,\cite{kn:petta2001,kn:petta2002} closely resembling
1487: %the random-matrix distribution without electron-electron
1488: %interactions.\cite{kn:brouwer2000,kn:matveev2000,kn:adam2002b}
1489: The enhanced fluctuations occurring for moderate spin-orbit
1490: scattering strengths (spin-orbit scattering rate $\gamma_{\rm so}$
1491: comparable to single-electron level spacing $\spacing$) may be
1492: typical of enhanced fluctuations of interaction matrix elements
1493: that are expected to occur at
1494: the breakdown of random matrix theory (dimensionless conductance
1495: $\sim 1$).
1496: 
1497: There are $\sim 20$ metallic elements
1498: in the periodic table for which the electron-electron interactions
1499: are sufficiently strong that the phenomena described here can
1500: be measured. Existing measurements of $g$ factors in nanoparticles
1501: have been made for Al and the noble metals only; in these metals,
1502: interaction effects are weak. We hope that our findings stimulate 
1503: experiments on other metals.
1504: 
1505: In our calculations we have omitted the orbital contribution to the
1506: $g$ factors. The orbital contribution arises from the fact that
1507: single-electron wavefunctions are complex if spin-orbit scattering 
1508: is present, instead of real, so that they carry a finite current 
1509: density. For the parameter regime we consider here, $\tau_{\rm so}
1510: \spacing \gtrsim 1$, the single-electron wavefunctions are mostly
1511: real, and we expect the orbital contribution to the $g$ factor to
1512: be smaller than the spin
1513: contribution.\cite{kn:matveev2000,kn:adam2002b}
1514: %Our main finding,
1515: %a large probability for $g$ factors larger than two, is the result
1516: %of electron-electron interactions, and does not require complex
1517: %wavefunctions for its existence. 
1518: We expect that taking into account a (small) orbital contribution
1519: causes a slight increase of the average $g$ factor and a further
1520: broadening of the distribution.
1521: 
1522: \section*{Acknowledgments}
1523: We thank S.~Adam, D.~Huertas-Hernando, 
1524: A.~Kaminski, A.~H.~MacDonald, J.~Petta and D.~C.~Ralph 
1525: for helpful discussions.
1526: This work was supported by the NSF under grant 
1527: no.\ DMR 0086509 and by the Packard foundation.
1528: \begin{appendix}     
1529: 
1530: \section{$g$ factors in the weak spin-orbit scattering limit}
1531: 
1532: In this appendix, we calculate $g$ factors to lowest order in the
1533: dimensionless spin-orbit scattering rate $\lambda$. 
1534: For a state that is twofold degenerate in the absence of
1535: spin-orbit scattering, spin-orbit scattering does not affect the 
1536: spin contribution to $g$ factors up to linear order in
1537: $\lambda$.\cite{kn:sone1977} Our calculation addresses the case
1538: of a fourfold degenerate level, and shows that its $g$
1539: factor is affected to zeroth order in $\lambda$.
1540: This calculation
1541: shows explicitly that the limit $\lambda \to 0$ is singular.
1542: 
1543: \begin{figure}[t]
1544: \epsfxsize= 0.75\hsize
1545: %\hspace{0.1\hsize}
1546: \epsffile{degenerate_PT.eps}
1547: \caption{ \label{weak_SOS}
1548: An $S=3/2$ state and the three relevant types of 
1549: excited states;
1550: (a)the four-fold degenerate $S=3/2$ ground state; (b) 
1551: two-fold degenerate $S=1/2$ excited state;
1552: (c)four-fold degenerate $S=3/2$ 
1553: excited state; (d) four-fold degenerate $S=1/2$ excited state,
1554: note that this state is entangled
1555: see Eq.~(\ref{s=-1/2})   
1556: } 
1557: \end{figure}
1558: 
1559: Let us start by writing the spin-orbit Hamiltonian in terms
1560: of the basis of single-electron eigenstates of the Hamiltonian
1561: $H_{\rm GOE}$ without spin-orbit scattering, see Eq.\
1562: (\ref{eq:H0}) above,
1563: \begin{eqnarray}
1564:   H_{\rm so} &=& \frac{i \lambda}{2 \sqrt{N}}
1565:   \sum_{\mu} \left[
1566:   (A_3)_{\mu\nu} (\hat \psi^{\dagger}_{\mu\uparrow}
1567:   \hat \psi^{\vphantom{\dagger}}_{\nu\uparrow} - 
1568:   \hat \psi^{\dagger}_{\mu\downarrow}
1569:   \hat \psi^{\vphantom{\dagger}}_{\nu\downarrow}) 
1570:   \nonumber \right. \\ && \left. \mbox{} +
1571:   (A_1 + i A_2)_{\mu\nu} \hat \psi^{\dagger}_{\mu\uparrow}
1572:   \hat \psi^{\vphantom{\dagger}}_{\nu\downarrow}
1573:   \nonumber \right. \\ && \left. \mbox{} +
1574:   (A_1 - i A_2)_{\mu\nu} \hat \psi^{\dagger}_{\mu\downarrow}
1575:   \hat \psi^{\vphantom{\dagger}}_{\nu\uparrow} \right].
1576:   \label{eq:Hsoapp}
1577: \end{eqnarray}
1578: Here the Greek indices $\mu$ and $\nu$ refer to the eigenvalues
1579: of $H_{\rm GOE}$, and not to the eigenvalues of the total
1580: single-electron Hamiltonian $H_{\rm GOE} + H_{\rm so}$ as in
1581: Sec.\ \ref{sec:theoretical_description}.
1582: 
1583: In order to study the effect of spin-orbit scattering on the
1584: $g$ factor of a many-electron eigenstate of $H_{\rm GOE}$ with
1585: spin $S=3/2$, one needs to calculate matrix elements between 
1586: the four members of the quadruplet. Labeling the four members
1587: of the quadruplet by the $z$ component of the spin,
1588: $S_z = p - 5/2$, $p=1,2,3,4$, these matrix elements can be
1589: arranged in a $4 \times 4$ matrix $V$ of the form
1590: \be
1591: V = 
1592: \left (
1593: \begin{array}{cccc}
1594:   - a-d   & b        & c          & 0  \\ 
1595:   b^{*}   & - a + d  & 0          & c  \\
1596:   c^{*}   & 0        &  - a + d   & -b \\
1597:   0       & c^{*}    &  - b^{*}   & - a - d 
1598: \end{array}
1599: \right ),
1600: \label{general_form}
1601: \ee
1602: with $a$ and $d$ real numbers and $b$ and $c$ complex numbers. 
1603: The specific form of (\ref{general_form}) follows from 
1604: time-reversal symmetry and guarantees that that the eigenvalues 
1605: of $V$ are double degenerate, in accordance with Kramers' theorem.
1606: 
1607: One quickly verifies that all matrix elements of $V$ are zero
1608: to first order in $H_{\rm so}$. This is the consequence of the 
1609: fact that the matrices
1610: $A_j$, $j=1,2,3$, in Eq.\ (\ref{eq:Hsoapp}) are antisymmetric,
1611: so that the spin-orbit interaction does not mix the states
1612: with opposite spin belonging to the same energy level. 
1613: In this situation one has to calculate elements of $V$ to 
1614: second order in $H_{\rm so}$.\cite{kn:landau1977} Denoting
1615: the many-electron states with roman indices, the matrix elements
1616: between the many-electron states $n$ and $n^{\prime}$ (both
1617: taken from the same quadruplet) are given by 
1618: \be
1619:   V_{n n^{\prime}} = \sum\limits_{m} 
1620:   \frac{(H_{\rm so})_{nm} (H_{\rm so})_{mn'} }
1621:   {E_{n}^{(0)} -E_{m}^{(0)}}, 
1622: \label{perturbation_theory}
1623: \ee
1624: with $m$ is summed over all many-electron states with $E_{m}
1625: \neq E_{n}$ and $E_{n}^{(0)}$ and $E_m^{(0)}$ the corresponding 
1626: many-electron energies.
1627: 
1628: The quadruplet state is represented schematically in Fig.\
1629: \ref{weak_SOS}a.
1630: For the calculation of the splitting of a quadruplet, it is 
1631: enough to consider states $m$ of the form indicated in Figs.\
1632: Fig.~\ref{weak_SOS}b, c, and d. These are: a twofold degenerate
1633: $S=1/2$ state (Fig.\ \ref{weak_SOS}b), a fourfold degenerate $S=3/2$ 
1634: state (Fig.\ \ref{weak_SOS}c) and a
1635: fourfold degenerate $S=1/2$ state (Fig.\ \ref{weak_SOS}d).
1636: There exist two variants of the states shown in Fig.\ \ref{weak_SOS}c 
1637: and d, depending on whether an empty single-electron level is
1638: filled above the Fermi level, or a hole is created below the Fermi
1639: level. The former case is shown in the figure.
1640: Since the spin-orbit interaction is a one-particle operator,
1641: only the states of the form shown in Fig.\ \ref{weak_SOS}
1642: which differ by not more than one electron-hole excitation 
1643: are important. 
1644: 
1645: {\em Virtual excitations to two-fold degenerate $S=1/2$ state.}
1646: In this case, the transition from the state $n$ to $m$ involves a 
1647: transition of an electron from the singly-occupied level $\mu$ to 
1648: the already (singly) occupied level $\nu$, see Fig.\ 
1649: \ref{weak_SOS}a and b.
1650: Representing the members of the spin $S=3/2$ quadruplet as
1651: $|3/2,S_z\rangle$ with $S_z = -3/2$, $-1/2$, $1/2$, $3/2$, and
1652: the members of the spin $S=1/2$ doublet as $|1/2,S_z\rangle$
1653: with $S_z = 1/2$, $-1/2$, we find the following matrix elements
1654: of the spin-orbit Hamiltonian $H_{\rm so}$,
1655: \begin{eqnarray}
1656:   \langle 3/2, +3/2 | H_{\rm so} | 1/2, +1/2 \rangle 
1657:   &=& \frac{i \lambda}{\sqrt{4N}}(A_1 - i A_2)_{\mu \nu},
1658:   \nonumber \\
1659:   \langle 3/2, +3/2 | H_{\rm so} | 1/2, -1/2 \rangle &=& 0,
1660:   \nonumber \\
1661:   \langle 3/2, +1/2 | H_{\rm so} | 1/2, +1/2 \rangle &=& 
1662:   - \frac{i \lambda}{\sqrt{3N}}(A_3)_{\mu\nu}, \nonumber \\
1663:   \langle 3/2, +1/2 | H_{\rm so} | 1/2, -1/2 \rangle &=&
1664:   \frac{i}{\sqrt{12 N}} (A_1-i A_2)_{\mu\nu} \nonumber \\
1665:   \langle 3/2, -1/2 | H_{\rm so} | 1/2, +1/2 \rangle &=& 
1666:   - \frac{i}{\sqrt{12 N}} (A_1+i A_2)_{\mu\nu} \nonumber \\
1667:   \langle 3/2, -1/2 | H_{\rm so} | 1/2, -1/2 \rangle &=& 
1668:   - \frac{i \lambda}{\sqrt{3N}}(A_3)_{\mu\nu}, \nonumber \\
1669:   \langle 3/2, -3/2 | H_{\rm so} | 1/2, +1/2 \rangle &=& 0
1670:   \nonumber \\
1671:   \langle 3/2, -3/2 | H_{\rm so} | 1/2, -1/2 \rangle &=&
1672:   -\frac{i \lambda}{\sqrt{4N}}(A_1 + i A_2)_{\mu \nu}. \nonumber \\
1673: \end{eqnarray}
1674: Further, in this case the energy difference 
1675: \begin{equation}
1676:   E_m^{(0)} - E_n^{(0)} = \varepsilon_{\nu} - \varepsilon_{\mu} + 3 J.
1677: \end{equation}
1678: Substituting these matrix elements and the energy difference 
1679: into Eq.~(\ref{perturbation_theory})
1680: we find the following contributions to the elements of the
1681: matrix $V$ of Eq.\ (\ref{general_form}):
1682: \begin{eqnarray}
1683:   b_{1,\mu\nu} &=& \frac{\lambda^2
1684:     (A_1 - i A_2)_{\mu \nu}(A_3)_{\mu\nu} 
1685:     }{4 N (\varepsilon_{\nu} - \varepsilon_{\mu} + 3 J) \sqrt{3}}
1686:   \nonumber \\
1687:   c_{1,\mu\nu} &=& \frac{\lambda^2 (A_1 - i A_2)_{\mu\nu}^2 }
1688:   {4 N (\varepsilon_{\nu} - \varepsilon_{\mu} + 3 J) \sqrt{3} } \nonumber \\
1689:   d_{1,\mu\nu} &=& 
1690:   \frac{\lambda^2 (A_1)_{\mu\nu}^2 + \lambda^2 (A_2)_{\mu\nu}^2 -
1691:   2 \lambda^2 (A_3)_{\mu\nu}^2}{12 N (\varepsilon_{\nu} - 
1692:   \varepsilon_{\mu} + 3 J)}. \nonumber
1693: \end{eqnarray}
1694: (We have not listed the value of $a$ in the matrix $V$ of
1695: Eq.\ (\ref{general_form}) since this coefficient does
1696: not contribute to the $g$ factor and splitting of the $S=3/2$
1697: quadruplet.)
1698: 
1699: {\em Virtual excitations to four-fold degenerate $S=3/2$ state.}
1700: These excitations involve a transition of an electron from a singly
1701: occupied level $\mu$ to an unoccupied level $\nu$, see Fig.\
1702: \ref{weak_SOS}, or the transition of an electron from a doubly
1703: occupied level $\mu$ to a singly occupied level $\nu$.
1704: Calculating the various matrix elements as before, we find after
1705: somewhat cumbersome algebra
1706: \begin{eqnarray}
1707:   b_{2,\mu\nu} &=& - \frac{\lambda^2
1708:     (A_1 - i A_2)_{\mu \nu}(A_3)_{\mu\nu} 
1709:     }{6 N (\varepsilon_{\nu} - \varepsilon_{\mu}) \sqrt{3}}
1710:   \nonumber \\
1711:   c_{2,\mu\nu} &=& - \frac{\lambda^2 (A_1 - i A_2)_{\mu\nu}^2 }
1712:   {6 N (\varepsilon_{\nu} - \varepsilon_{\mu}) \sqrt{3} } \nonumber \\
1713:   d_{2,\mu\nu} &=& 
1714:   - \frac{\lambda^2 (A_1)_{\mu\nu}^2 + \lambda^2 (A_2)_{\mu\nu}^2 -
1715:   2 \lambda^2 (A_3)_{\mu\nu}^2}{18 N (\varepsilon_{\nu} -
1716:     \varepsilon_{\mu})}.
1717:  \nonumber
1718: \end{eqnarray}
1719: 
1720: {\em Virtual excitations to four-fold degenerate $S=1/2$ state.}
1721: As in the previous case, 
1722: these excitations involve a transition of an electron from a singly
1723: occupied level $\mu$ to an unoccupied level $\nu$, see Fig.\
1724: \ref{weak_SOS}, or the transition of an electron from a doubly
1725: occupied level $\mu$ to a singly occupied level $\nu$. The four
1726: $S=1/2$ are labeled by $S_z=\pm 1/2$ and by an additional degeneracy
1727: parameter $q = \pm 1$,
1728: \begin{eqnarray}
1729:   |1/2,+1/2,q\rangle &=& \frac{1}{\sqrt{3}}
1730:   \left (
1731:   |\uparrow    \uparrow    \downarrow \rangle +
1732:   |\uparrow    \downarrow  \uparrow   \rangle e^{ i\frac{2\pi q}{3}} 
1733:   \right. \nonumber \\ && \left. \mbox{} +
1734:   |\downarrow  \uparrow    \uparrow   \rangle e^{-i\frac{2\pi q}{3}}
1735:   \right ), \nonumber \\
1736:   |1/2,-1/2,q\rangle &=& \frac{1}{\sqrt{3}}
1737:   \left (
1738:   |\downarrow  \downarrow  \uparrow   \rangle +
1739:   |\downarrow  \uparrow    \downarrow \rangle e^{ i\frac{2\pi q}{3}}
1740:   \right. \nonumber \\ && \left. \mbox{} +
1741:   |\uparrow    \downarrow  \downarrow \rangle e^{-i\frac{2\pi q}{3}}
1742:   \right ).
1743: \label{s=-1/2}
1744: \end{eqnarray}
1745: Performing the calculations as before, we find
1746: \begin{eqnarray}
1747:   b_{3,\mu\nu} &=& \frac{\lambda^2
1748:     (A_1 - i A_2)_{\mu \nu}(A_3)_{\mu\nu} 
1749:     }{6 N (\varepsilon_{\nu} - \varepsilon_{\mu} + 3 J) \sqrt{3}}
1750:   \nonumber \\
1751:   c_{3,\mu\nu} &=& \frac{\lambda^2 (A_1 - i A_2)_{\mu\nu}^2 }
1752:   {6 N (\varepsilon_{\nu} - \varepsilon_{\mu} + 3 J) \sqrt{3} } \nonumber \\
1753:   d_{3,\mu\nu} &=& 
1754:   \frac{\lambda^2 (A_1)_{\mu\nu}^2 + \lambda^2 (A_2)_{\mu\nu}^2 -
1755:   2 \lambda^2 (A_3)_{\mu\nu}^2}{18 N (\varepsilon_{\nu} -
1756:     \varepsilon_{\mu} + 3 J)}.
1757:  \nonumber
1758: \end{eqnarray}
1759: 
1760: Denoting the set of doubly occupied single-electron levels by 
1761: ``0'', the set of singly-occupied single-electron levels by ``1''
1762: and the set of unoccupied single-electron levels by ``2'', we then
1763: sum over all virtual excitations and find
1764: \begin{eqnarray}
1765:   b &=& \sum_{\mu\neq \nu \in 1} b_{1,\mu\nu}
1766:   + \sum_{\mu \in 1} \sum_{\nu \in 2} (b_{2,\mu\nu} + b_{3,\mu\nu})
1767:   \nonumber \\ && \mbox{}
1768:   + \sum_{\mu \in 0} \sum_{\nu \in 1} (b_{2,\mu\nu} + b_{3,\mu\nu}),
1769:   \nonumber \\
1770: \end{eqnarray}
1771: and similar expressions for the coefficients $c$ and $d$ in Eq.\
1772: (\ref{general_form}).
1773: 
1774: The matrix (\ref{general_form}) can be diagonalized for all
1775: values of the parameters $a$, $b$, $c$, and $d$, and the 
1776: corresponding $g$-factors can be found exactly. After diagonalization
1777: we find that the quadruplet is split into two doublets with energy
1778: separation
1779: \begin{equation}
1780:   (\Delta E)^2 = 4 d^2 + 4 |b|^2 + 4 |c|^2.
1781: \label{splitting}
1782: \end{equation}
1783: In the $4 \times 4$ matrix notation of Eq.\ (\ref{general_form}),
1784: the Zeeman Hamiltonian reads
1785: \begin{equation}
1786:   H_{\rm Z} =
1787:   \left( \begin{array}{llll} -3 \mu_B H \\
1788:   & - \mu_B H \\
1789:   && \mu_B H \\
1790:   &&& 3 \mu_B H \end{array} \right).
1791: \end{equation}
1792: Lifting the degeneracy of the two doublets by the Zeeman energy,
1793: we find $g$ factors
1794: \begin{equation}
1795:   g = 2 \sqrt{ \frac{3 |b|^2 + (2 d \pm \Delta E)^2}
1796:   {d^2 + |b|^2 + |c|^2}},
1797: \label{g_factor_special}
1798: \end{equation}
1799: where the $\pm$ sign refers to the two doublets. This result
1800: confirms the assertion made earlier, that spin-orbit scattering
1801: affects the $g$ factors of the $S=3/2$ states to zeroth order
1802: in the spin-orbit scattering rate $\lambda$. Of course, for
1803: small $\lambda$ the energy splitting $\Delta E$ is small as
1804: well, and the $g$ factor of Eq.\ (\ref{g_factor_special}) can
1805: be observed for magnetic fields such that $\mu_B H \ll \Delta E$
1806: only, which limits the practical observability of the $g$
1807: factor (\ref{g_factor_special}) for very small spin-orbit
1808: scattering rates $\lambda \ll 1$.
1809: 
1810: In the remainder of this appendix we investigate Eq.\
1811: (\ref{g_factor_special}) for the special case that one state
1812: is very close to the $S=3/2$ state of interest and virtual
1813: excitations to that state dominate the spin-orbit matrix
1814: $V$ in Eq.\ (\ref{general_form}). One important example is
1815: the case when the ground state has spin $S=3/2$, which is
1816: expected with small probability for $J/\spacing \gtrsim 0.3$, see
1817: Fig.\ \ref{two_different_possibilities}.
1818: Indeed,
1819: the energy difference between the $S=3/2$ state and the 
1820: lowest-lying $S=1/2$ state is 
1821: $\varepsilon_{\lambda+2} - \varepsilon_{\lambda} - 3 J$, 
1822: $\lambda$ being the index of the lowest singly
1823: occupied level in the $S=3/2$ state. Typically, this energy
1824: difference is small, given the small likelihood of it being
1825: positive.
1826: 
1827: Substituting the general expressions for the coefficients
1828: $b$, $c$, and $d$ into Eq.\ (\ref{g_factor_special}) and
1829: noting that one energy denominator is much smaller than
1830: all others, we find that an $S=3/2$ quadruplet splits into
1831: two doublets with $g$ factors
1832: \begin{eqnarray}
1833:   g &=& 2\sqrt{
1834:   1+
1835:   \frac{3 (A_1^2 + A_2^2)}{A_1^2 + A_2^2 + A_3^2}}, \nonumber \\
1836:   g'  &=&
1837:   \sqrt{48 - 3 g^2}
1838:   \nonumber \\
1839:   &=& 6 \sqrt{\frac{A_3^2}{A_1^2 + A_2^2 + A_3^2}},
1840: \label{simplest_case}
1841: \end{eqnarray}
1842: where we have omitted the indices referring to the levels
1843: $\mu$ and $\nu$ involved as we deal with one excited state
1844: only.  If the quadruplet state is the ground state, $g$ 
1845: corresponds to the lower lying doublet.
1846: \begin{figure}[t]
1847: \epsfxsize= 0.7\hsize
1848: \hspace{0.1\hsize}
1849: \epsffile{two_possibilities.eps}
1850: %\vskip5cm
1851: \caption{ \label{two_different_possibilities}
1852: Two different possibilities: (a) $S=3/2$ state is the ground state
1853: with the probability $P_{3/2}$
1854: (b) $S=1/2$ state is the ground state
1855: with the probability $1-P_{3/2}$. Note that due to the sign
1856: change in the energy denominator, see Eq.~\ref{perturbation_theory}
1857: the average $g$-factors $<g>$ and $<g^{\prime}>$ are inverted.  
1858: } 
1859: \end{figure}
1860: 
1861: In the special case of an $S=3/2$ ground state, Eq.\ (\ref{simplest_case})
1862: shows that the $g$-factor takes values in the interval
1863: $[2,4]$ only. One has $g=2$ 
1864: only if $A_1=A_2=0$. The corresponding eigenstates are
1865: $|3/2,\pm 1/2\rangle$. In the opposite case $A_3=0$, one has
1866: $g=4$ in the ground state and the corresponding eigenstates
1867: are $(\sqrt{3}/{2}) |3/2, \pm 3/2\rangle
1868: - (e^{2 i \phi}/2) |3/2, \mp 1/2\rangle$,
1869: where $e^{- i \phi} = (A_2 + i A_1)/|A_2 + i A_1|$.
1870: In fact, one can find the entire probability distribution of
1871: $g$ in this case, using the fact that $A_1$, $A_2$, and $A_3$
1872: are taken from identical and independent Gaussian distributions,
1873: \begin{equation}
1874:   P(g) = \frac{1}{2 \sqrt{3}}
1875:   \frac{g}{\sqrt{16 - g^2}}. \label{eq:Pg}
1876: \end{equation}
1877: The average $g$-factor for an $S=3/2$ ground state with a 
1878: nearby $S=1/2$ state is then
1879: \begin{eqnarray}
1880: \langle g \rangle & = &
1881: \frac{4}{\sqrt{3}}\left (
1882: \frac{\pi}{2} - \arcsin{\frac{1}{2}} + \frac{\sqrt{3}}{4} 
1883: \right ) \nonumber \\
1884: &\approx & 3.418 .
1885: \end{eqnarray} 
1886: For $J=0.6 \delta$, the probability to find a ground state with
1887: $S=3/2$ is $P_{3/2} \approx 0.38$ (see Ref.\ \onlinecite{kn:oreg2002} or
1888: Fig.\ \ref{probability}). Since the $g$ factor is unaffected to
1889: first order in $\lambda$ if the ground state has spin $1/2$, we
1890: expect the true average ground state $g$ factor to be approximately
1891: equal to
1892: \be
1893:   \langle g_{0} \rangle \approx 3.418 P_{3/2} + 2 (1 - P_{3/2})
1894:   \approx 2.54.
1895: \ee
1896: This value is very close to that found in the numerical simulations
1897: of the ground state $g$-factor $\approx 2.5$, see 
1898: Fig.~\ref{ground_state_g_factor}.
1899: 
1900: The distribution of $g'$ can be found from Eqs.\
1901: (\ref{simplest_case}) and (\ref{eq:Pg}). One finds the particularly
1902: simple result
1903: \be
1904:   P(g') = \frac{1}{6},\ \ 0 < g' < 6.
1905: \ee
1906: The average $\langle g' \rangle = 3$.
1907: In the limiting cases $g' = 6$ and $g'=0$ the doubly-degenerate
1908: eigenstates 
1909: have the form $| {3/2,\pm 3/2}\rangle $ and $(1/2)|3/2, \pm 3/2\rangle
1910: - (\sqrt{3}/2)e^{2i\phi}|3/2 , \mp 1/2\rangle$, respectively. We
1911: can use these results to calculate the average $g$-factors of the
1912: lowest excited states. For the first excited state
1913: \be
1914: \langle g_1 \rangle = \langle g^{\prime}\rangle P_{3/2} + 
1915: \langle g^{\prime}\rangle
1916: \left (1 - P_{3/2}\right ) = \langle g^{\prime}\rangle = 3 ,
1917: \ee
1918: where we assumed that without spin-orbit scattering the ground
1919: state has spin $S=3/2$ or the the first excited state has spin 
1920: $S=3/2$ and is slightly below an $S=1/2$ doublet, see Fig.\
1921: \ref{two_different_possibilities}. (Hence, we neglect the 
1922: possibilities that the first excited state has spin $S=1/2$ or
1923: that the first excited state has spin $S=3/2$ and is far away
1924: from the next $S=1/2$ state. Our approximation should slightly
1925: overestimate $\langle g_1 \rangle$.)
1926: The simulations give $\langle g_1 \rangle\approx 3.00$, see 
1927: Fig.~\ref{ground_state_g_factor}. With the same
1928: approximations, we find that the average $g$ factor of the
1929: second excited many-electron state is
1930: \be
1931: \langle g_2 \rangle = 2P_{3/2} + \langle g \rangle
1932: \left (1 - P_{3/2}\right )\approx 2.87 .
1933:   \label{eq:st}
1934: \ee
1935: The result of the simulation is $2.78$, in good agreement with the
1936: estimate (\ref{eq:st}).
1937: Note that the above nontrivial distributions
1938: of $g$-factors can be observed in a small 
1939: magnetic field only ($\mu_B H \ll \lambda^2 \delta$). For larger 
1940: fields the $g$-factors are the same as in the
1941: absence of spin-orbit coupling.  
1942: 
1943: 
1944: 
1945: 
1946: 
1947: %\end{widetext}
1948: 
1949: \end{appendix}
1950: 
1951: 
1952: %\bibliography{/afs/msc.cornell.edu/home/brouwer/brouwer/grant/2003/refs,%
1953: %/afs/msc.cornell.edu/home/brouwer/brouwer/sympl/gorokhov/gffoot}
1954: 
1955: %\end{document}
1956: 
1957: \begin{thebibliography}{42}
1958: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1959: \expandafter\ifx\csname bibnamefont\endcsname\relax
1960:   \def\bibnamefont#1{#1}\fi
1961: \expandafter\ifx\csname bibfnamefont\endcsname\relax
1962:   \def\bibfnamefont#1{#1}\fi
1963: \expandafter\ifx\csname citenamefont\endcsname\relax
1964:   \def\citenamefont#1{#1}\fi
1965: \expandafter\ifx\csname url\endcsname\relax
1966:   \def\url#1{\texttt{#1}}\fi
1967: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
1968: \providecommand{\bibinfo}[2]{#2}
1969: \providecommand{\eprint}[2][]{\url{#2}}
1970: 
1971: \bibitem[{\citenamefont{Kurland et~al.}(2002)\citenamefont{Kurland, Aleiner,
1972:   and Altshuler}}]{kn:kurland2000}
1973: \bibinfo{author}{\bibfnamefont{I.~L.} \bibnamefont{Kurland}},
1974:   \bibinfo{author}{\bibfnamefont{I.~L.} \bibnamefont{Aleiner}},
1975:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~L.}
1976:   \bibnamefont{Altshuler}}, \bibinfo{journal}{Phys. Rev. B}
1977:   \textbf{\bibinfo{volume}{62}}, \bibinfo{pages}{14886} (\bibinfo{year}{2002}).
1978: 
1979: \bibitem[{\citenamefont{Aleiner et~al.}(2002)\citenamefont{Aleiner, Brouwer,
1980:   and Glazman}}]{kn:aleiner2002}
1981: \bibinfo{author}{\bibfnamefont{I.~L.} \bibnamefont{Aleiner}},
1982:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
1983:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{L.~I.}
1984:   \bibnamefont{Glazman}}, \bibinfo{journal}{Phys. Rep.}
1985:   \textbf{\bibinfo{volume}{358}}, \bibinfo{pages}{309} (\bibinfo{year}{2002}).
1986: 
1987: \bibitem[{foo({\natexlab{a}})}]{foot0}
1988: \bibinfo{note}{In principle, there is one more contribution to the interaction
1989:   Hamiltonian. This is a ``pairing interaction'', which, if attractive, causes
1990:   a superconducting instability at sufficiently low temperatures. We consider
1991:   normal metal nanoparticles for which the pairing interaction is small}.
1992: 
1993: \bibitem[{\citenamefont{Efetov}(1983)}]{kn:efetov1983}
1994: \bibinfo{author}{\bibfnamefont{K.~B.} \bibnamefont{Efetov}},
1995:   \bibinfo{journal}{Adv. Phys.} \textbf{\bibinfo{volume}{32}},
1996:   \bibinfo{pages}{53} (\bibinfo{year}{1983}).
1997: 
1998: \bibitem[{\citenamefont{Altshuler and Shklovskii}(1986)}]{kn:altshuler1986}
1999: \bibinfo{author}{\bibfnamefont{B.~L.} \bibnamefont{Altshuler}}
2000:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{B.~I.}
2001:   \bibnamefont{Shklovskii}}, \bibinfo{journal}{Sov. Phys. JETP}
2002:   \textbf{\bibinfo{volume}{64}}, \bibinfo{pages}{127} (\bibinfo{year}{1986}).
2003: 
2004: \bibitem[{\citenamefont{Adam et~al.}(2002{\natexlab{a}})\citenamefont{Adam,
2005:   Brouwer, Sethna, and Waintal}}]{kn:adam2002}
2006: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Adam}},
2007:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2008:   \bibinfo{author}{\bibfnamefont{J.~P.} \bibnamefont{Sethna}},
2009:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Waintal}},
2010:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{66}},
2011:   \bibinfo{pages}{165310} (\bibinfo{year}{2002}{\natexlab{a}}).
2012: 
2013: \bibitem[{\citenamefont{Salinas et~al.}(1999)\citenamefont{Salinas, Gu\'eron,
2014:   Ralph, Black, and Tinkham}}]{kn:salinas1999}
2015: \bibinfo{author}{\bibfnamefont{D.~G.} \bibnamefont{Salinas}},
2016:   \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Gu\'eron}},
2017:   \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Ralph}},
2018:   \bibinfo{author}{\bibfnamefont{C.~T.} \bibnamefont{Black}}, \bibnamefont{and}
2019:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tinkham}},
2020:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{60}},
2021:   \bibinfo{pages}{6137} (\bibinfo{year}{1999}).
2022: 
2023: \bibitem[{\citenamefont{Davidovic and Tinkham}(1999)}]{kn:davidovic1999}
2024: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Davidovic}} \bibnamefont{and}
2025:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tinkham}},
2026:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{83}},
2027:   \bibinfo{pages}{1644} (\bibinfo{year}{1999}).
2028: 
2029: \bibitem[{\citenamefont{Petta and Ralph}(2001)}]{kn:petta2001}
2030: \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Petta}} \bibnamefont{and}
2031:   \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Ralph}},
2032:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{87}},
2033:   \bibinfo{pages}{266801} (\bibinfo{year}{2001}).
2034: 
2035: \bibitem[{\citenamefont{Petta and Ralph}(2002)}]{kn:petta2002}
2036: \bibinfo{author}{\bibfnamefont{J.~R.} \bibnamefont{Petta}} \bibnamefont{and}
2037:   \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Ralph}},
2038:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{89}},
2039:   \bibinfo{pages}{156802} (\bibinfo{year}{2002}).
2040: 
2041: \bibitem[{\citenamefont{Ralph et~al.}(1995)\citenamefont{Ralph, Black, and
2042:   Tinkham}}]{kn:ralph1995}
2043: \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Ralph}},
2044:   \bibinfo{author}{\bibfnamefont{C.~T.} \bibnamefont{Black}}, \bibnamefont{and}
2045:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Tinkham}},
2046:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{74}},
2047:   \bibinfo{pages}{3241} (\bibinfo{year}{1995}).
2048: 
2049: \bibitem[{\citenamefont{Gorokhov and Brouwer}(2003)}]{kn:gorokhov2003}
2050: \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Gorokhov}} \bibnamefont{and}
2051:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2052:   \bibinfo{journal}{Phys. Rev. Lett} \textbf{\bibinfo{volume}{91}},
2053:   \bibinfo{pages}{186602} (\bibinfo{year}{2003}).
2054: 
2055: \bibitem[{\citenamefont{von Delft and Ralph}(2001)}]{kn:vondelft2001}
2056: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{von Delft}} \bibnamefont{and}
2057:   \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Ralph}},
2058:   \bibinfo{journal}{Phys. Rep.} \textbf{\bibinfo{volume}{345}},
2059:   \bibinfo{pages}{61} (\bibinfo{year}{2001}).
2060: 
2061: \bibitem[{\citenamefont{Brouwer et~al.}(2000)\citenamefont{Brouwer, Waintal,
2062:   and Halperin}}]{kn:brouwer2000}
2063: \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2064:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Waintal}}, \bibnamefont{and}
2065:   \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Halperin}},
2066:   \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{85}},
2067:   \bibinfo{pages}{369} (\bibinfo{year}{2000}).
2068: 
2069: \bibitem[{\citenamefont{Matveev et~al.}(2000)\citenamefont{Matveev, Glazman,
2070:   and Larkin}}]{kn:matveev2000}
2071: \bibinfo{author}{\bibfnamefont{K.~A.} \bibnamefont{Matveev}},
2072:   \bibinfo{author}{\bibfnamefont{L.~I.} \bibnamefont{Glazman}},
2073:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{A.~I.}
2074:   \bibnamefont{Larkin}}, \bibinfo{journal}{Phys. Rev. Lett.}
2075:   \textbf{\bibinfo{volume}{85}}, \bibinfo{pages}{2789} (\bibinfo{year}{2000}).
2076: 
2077: \bibitem[{\citenamefont{Adam et~al.}(2002{\natexlab{b}})\citenamefont{Adam,
2078:   Polianski, Waintal, and Brouwer}}]{kn:adam2002b}
2079: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Adam}},
2080:   \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Polianski}},
2081:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Waintal}}, \bibnamefont{and}
2082:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2083:   \bibinfo{journal}{Phys. Rev. B}  (\bibinfo{year}{2002}{\natexlab{b}}).
2084: 
2085: \bibitem[{\citenamefont{Brouwer et~al.}(1999)\citenamefont{Brouwer, Oreg, and
2086:   Halperin}}]{kn:brouwer1999b}
2087: \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2088:   \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Oreg}}, \bibnamefont{and}
2089:   \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Halperin}},
2090:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{60}},
2091:   \bibinfo{pages}{13977} (\bibinfo{year}{1999}).
2092: 
2093: \bibitem[{\citenamefont{Baranger et~al.}(2000)\citenamefont{Baranger, Ullmo,
2094:   and Glazman}}]{kn:baranger2000}
2095: \bibinfo{author}{\bibfnamefont{H.~U.} \bibnamefont{Baranger}},
2096:   \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Ullmo}}, \bibnamefont{and}
2097:   \bibinfo{author}{\bibfnamefont{L.~I.} \bibnamefont{Glazman}},
2098:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{61}},
2099:   \bibinfo{pages}{2425} (\bibinfo{year}{2000}).
2100: 
2101: \bibitem[{\citenamefont{Usaj and Baranger}(2002)}]{kn:usaj2003}
2102: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Usaj}} \bibnamefont{and}
2103:   \bibinfo{author}{\bibfnamefont{H.~U.} \bibnamefont{Baranger}},
2104:   \bibinfo{journal}{cond-mat/0211649}  (\bibinfo{year}{2002}).
2105: 
2106: \bibitem[{\citenamefont{Jiang et~al.}(2002)\citenamefont{Jiang, Baranger, and
2107:   Yang}}]{kn:jiang2003}
2108: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Jiang}},
2109:   \bibinfo{author}{\bibfnamefont{H.~U.} \bibnamefont{Baranger}},
2110:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{Yang}},
2111:   \bibinfo{journal}{cond-mat/0208146}  (\bibinfo{year}{2002}).
2112: 
2113: \bibitem[{\citenamefont{Oreg et~al.}(2002)\citenamefont{Oreg, Brouwer, Waintal,
2114:   and Halperin}}]{kn:oreg2002}
2115: \bibinfo{author}{\bibfnamefont{Y.}~\bibnamefont{Oreg}},
2116:   \bibinfo{author}{\bibfnamefont{P.~W.} \bibnamefont{Brouwer}},
2117:   \bibinfo{author}{\bibfnamefont{X.}~\bibnamefont{Waintal}}, \bibnamefont{and}
2118:   \bibinfo{author}{\bibfnamefont{B.~I.} \bibnamefont{Halperin}}, in
2119:   \emph{\bibinfo{booktitle}{Nano-Physics and Bio-Electronics: a new Odyssee}},
2120:   edited by \bibinfo{editor}{\bibfnamefont{T.}~\bibnamefont{Chakraborty}},
2121:   \bibinfo{editor}{\bibfnamefont{F.}~\bibnamefont{Peeters}}, \bibnamefont{and}
2122:   \bibinfo{editor}{\bibfnamefont{U.}~\bibnamefont{Sivan}}
2123:   (\bibinfo{publisher}{Elsevier}, \bibinfo{year}{2002}).
2124: 
2125: \bibitem[{\citenamefont{MacDonald et~al.}(1982)\citenamefont{MacDonald, Daams,
2126:   Vosko, and Koelling}}]{kn:macdonald1982}
2127: \bibinfo{author}{\bibfnamefont{A.~H.} \bibnamefont{MacDonald}},
2128:   \bibinfo{author}{\bibfnamefont{J.~M.} \bibnamefont{Daams}},
2129:   \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Vosko}}, \bibnamefont{and}
2130:   \bibinfo{author}{\bibfnamefont{D.~D.} \bibnamefont{Koelling}},
2131:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{25}},
2132:   \bibinfo{pages}{713} (\bibinfo{year}{1982}).
2133: 
2134: \bibitem[{\citenamefont{Vier et~al.}(1984)\citenamefont{Vier, Tolleth, and
2135:   Schultz}}]{kn:vier1984}
2136: \bibinfo{author}{\bibfnamefont{D.~C.} \bibnamefont{Vier}},
2137:   \bibinfo{author}{\bibfnamefont{D.~W.} \bibnamefont{Tolleth}},
2138:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Schultz}},
2139:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{29}},
2140:   \bibinfo{pages}{88} (\bibinfo{year}{1984}).
2141: 
2142: \bibitem[{\citenamefont{Bergmann}(1984)}]{kn:bergmann1984}
2143: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Bergmann}},
2144:   \bibinfo{journal}{Phys. Rep.} \textbf{\bibinfo{volume}{107}},
2145:   \bibinfo{pages}{1} (\bibinfo{year}{1984}).
2146: 
2147: \bibitem[{foo({\natexlab{b}})}]{foot1}
2148: \bibinfo{note}{In the analysis of Refs.\
2149:   \onlinecite{kn:petta2001,kn:petta2002}, the orbital contribution to the $g$
2150:   factors was omitted. Omitting the orbital contribution is questionable for
2151:   nanoparticles
2152:   in which small $g$ factors are measured ($g \sim 0.2$), since the
2153:   orbital contribution to the $g$ factor is predicted to saturate around $g
2154:   \sim 1$ for strong spin-orbit scattering in ballistic
2155:   nanoparticles.\cite{kn:matveev2000}}.
2156: 
2157: \bibitem[{\citenamefont{Dupree and Geldart}(1971)}]{kn:dupree1971}
2158: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Dupree}} \bibnamefont{and}
2159:   \bibinfo{author}{\bibfnamefont{D.~J.~W.} \bibnamefont{Geldart}},
2160:   \bibinfo{journal}{Solid State Commun.} \textbf{\bibinfo{volume}{9}},
2161:   \bibinfo{pages}{145} (\bibinfo{year}{1971}).
2162: 
2163: \bibitem[{\citenamefont{Wilk et~al.}(1978)\citenamefont{Wilk, Fehlner, and
2164:   Vosko}}]{kn:wilk1978}
2165: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Wilk}},
2166:   \bibinfo{author}{\bibfnamefont{W.~R.} \bibnamefont{Fehlner}},
2167:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Vosko}},
2168:   \bibinfo{journal}{Can.\ J.\ Phys.} \textbf{\bibinfo{volume}{56}},
2169:   \bibinfo{pages}{266} (\bibinfo{year}{1978}).
2170: 
2171: \bibitem[{\citenamefont{Janak}(1977)}]{kn:janak1977}
2172: \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{Janak}},
2173:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{16}},
2174:   \bibinfo{pages}{255} (\bibinfo{year}{1977}).
2175: 
2176: \bibitem[{\citenamefont{Dunifer et~al.}(1977)\citenamefont{Dunifer, Pattison,
2177:   and Hsu}}]{kn:dunifer1977}
2178: \bibinfo{author}{\bibfnamefont{G.~L.} \bibnamefont{Dunifer}},
2179:   \bibinfo{author}{\bibfnamefont{M.~R.} \bibnamefont{Pattison}},
2180:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.~M.} \bibnamefont{Hsu}},
2181:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{15}},
2182:   \bibinfo{pages}{315} (\bibinfo{year}{1977}).
2183: 
2184: \bibitem[{\citenamefont{Knecht}(1975)}]{kn:knecht1975}
2185: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Knecht}}, \bibinfo{journal}{J.
2186:   Low Temp. Phys.} \textbf{\bibinfo{volume}{21}}, \bibinfo{pages}{619}
2187:   (\bibinfo{year}{1975}).
2188: 
2189: \bibitem[{\citenamefont{Randles}(1972)}]{kn:randles1972}
2190: \bibinfo{author}{\bibfnamefont{D.~L.} \bibnamefont{Randles}},
2191:   \bibinfo{journal}{Proc. R. Soc. Lond. A} \textbf{\bibinfo{volume}{331}},
2192:   \bibinfo{pages}{85} (\bibinfo{year}{1972}).
2193: 
2194: \bibitem[{\citenamefont{Sigalas and
2195:   Papaconstantopoulos}(1994)}]{kn:sigalas1994}
2196: \bibinfo{author}{\bibfnamefont{M.~M.} \bibnamefont{Sigalas}} \bibnamefont{and}
2197:   \bibinfo{author}{\bibfnamefont{D.~A.} \bibnamefont{Papaconstantopoulos}},
2198:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{50}},
2199:   \bibinfo{pages}{7255} (\bibinfo{year}{1994}).
2200: 
2201: \bibitem[{\citenamefont{Mcdonald et~al.}(1977)\citenamefont{Mcdonald, Liu, and
2202:   Vosko}}]{kn:macdonald1977a}
2203: \bibinfo{author}{\bibfnamefont{A.~H.} \bibnamefont{Mcdonald}},
2204:   \bibinfo{author}{\bibfnamefont{K.~L.} \bibnamefont{Liu}}, \bibnamefont{and}
2205:   \bibinfo{author}{\bibfnamefont{S.~H.} \bibnamefont{Vosko}},
2206:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{16}},
2207:   \bibinfo{pages}{777} (\bibinfo{year}{1977}).
2208: 
2209: \bibitem[{\citenamefont{Bakonyi et~al.}(1993)\citenamefont{Bakonyi, Ebert, and
2210:   Liechtenstein}}]{kn:bakonyi1993}
2211: \bibinfo{author}{\bibfnamefont{I.}~\bibnamefont{Bakonyi}},
2212:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Ebert}}, \bibnamefont{and}
2213:   \bibinfo{author}{\bibfnamefont{A.~I.} \bibnamefont{Liechtenstein}},
2214:   \bibinfo{journal}{Phys. Rev. B} \textbf{\bibinfo{volume}{48}},
2215:   \bibinfo{pages}{7841} (\bibinfo{year}{1993}).
2216: 
2217: \bibitem[{\citenamefont{Stenzel and Winter}(1986)}]{kn:stenzel1986}
2218: \bibinfo{author}{\bibfnamefont{E.}~\bibnamefont{Stenzel}} \bibnamefont{and}
2219:   \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Winter}},
2220:   \bibinfo{journal}{J. Phys. F} \textbf{\bibinfo{volume}{16}},
2221:   \bibinfo{pages}{1789} (\bibinfo{year}{1986}).
2222: 
2223: \bibitem[{\citenamefont{Hechtfischer}(1976)}]{kn:hechtfischer1976}
2224: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Hechtfischer}},
2225:   \bibinfo{journal}{Z. Phys. B} \textbf{\bibinfo{volume}{23}},
2226:   \bibinfo{pages}{255} (\bibinfo{year}{1976}).
2227: 
2228: \bibitem[{\citenamefont{Manuel and Quinton}(1963)}]{kn:manuel1963}
2229: \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Manuel}} \bibnamefont{and}
2230:   \bibinfo{author}{\bibfnamefont{J.~M.~P.~S.} \bibnamefont{Quinton}},
2231:   \bibinfo{journal}{Proc. R. Soc. A} \textbf{\bibinfo{volume}{273}},
2232:   \bibinfo{pages}{142} (\bibinfo{year}{1963}).
2233: 
2234: \bibitem[{\citenamefont{Fradin et~al.}(1975)\citenamefont{Fradin, Koelling,
2235:   Freeman, and Watson-Yang}}]{kn:fradin1975}
2236: \bibinfo{author}{\bibfnamefont{F.~Y.} \bibnamefont{Fradin}},
2237:   \bibinfo{author}{\bibfnamefont{D.~D.} \bibnamefont{Koelling}},
2238:   \bibinfo{author}{\bibfnamefont{A.~J.} \bibnamefont{Freeman}},
2239:   \bibnamefont{and} \bibinfo{author}{\bibfnamefont{T.~J.}
2240:   \bibnamefont{Watson-Yang}}, \bibinfo{journal}{Phys. Rev. B}
2241:   \textbf{\bibinfo{volume}{12}}, \bibinfo{pages}{5570} (\bibinfo{year}{1975}).
2242: 
2243: \bibitem[{\citenamefont{Budworth et~al.}(1960)\citenamefont{Budworth, Hoare,
2244:   and Preston}}]{kn:budworth1960}
2245: \bibinfo{author}{\bibfnamefont{D.~W.} \bibnamefont{Budworth}},
2246:   \bibinfo{author}{\bibfnamefont{F.~E.} \bibnamefont{Hoare}}, \bibnamefont{and}
2247:   \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Preston}},
2248:   \bibinfo{journal}{Proc. R. Soc. London A} \textbf{\bibinfo{volume}{257}},
2249:   \bibinfo{pages}{250} (\bibinfo{year}{1960}).
2250: 
2251: \bibitem[{\citenamefont{Pines and Nozi\'eres}(1966)}]{kn:pines1966}
2252: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Pines}} \bibnamefont{and}
2253:   \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Nozi\'eres}},
2254:   \emph{\bibinfo{title}{The theory of quantum liquids}} (\bibinfo{publisher}{W.
2255:   A. Benjamin Inc., New York}, \bibinfo{year}{1966}).
2256: 
2257: \bibitem[{\citenamefont{Sone}(1977)}]{kn:sone1977}
2258: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Sone}}, \bibinfo{journal}{J.
2259:   Phys. Soc. Japan} \textbf{\bibinfo{volume}{42}}, \bibinfo{pages}{1457}
2260:   (\bibinfo{year}{1977}).
2261: 
2262: \bibitem[{\citenamefont{Landau and Lifshitz}(1977)}]{kn:landau1977}
2263: \bibinfo{author}{\bibfnamefont{L.~D.} \bibnamefont{Landau}} \bibnamefont{and}
2264:   \bibinfo{author}{\bibfnamefont{E.~M.} \bibnamefont{Lifshitz}},
2265:   \emph{\bibinfo{title}{Quantum mechanics}}, vol.~\bibinfo{volume}{3} of
2266:   \emph{\bibinfo{series}{Course in Theoretical Physic}}
2267:   (\bibinfo{publisher}{Pergamon, Oxford}, \bibinfo{year}{1977}).
2268: 
2269: \end{thebibliography}
2270: 
2271: 
2272: \end{document}
2273: 
2274: 
2275: 
2276: 
2277: 
2278: 
2279: 
2280: 
2281: 
2282: