cond-mat0305412/rmp.tex
1: %FORMAT LATEXE
2: \documentstyle[multicol,aps,rmp,amsfonts,graphics,epsfig]{revtex}
3: \begin{document}
4: \title{Saturation of electrical resistivity}                              
5: \author{O. Gunnarsson$^{(1)}$, M. Calandra$^{(2)}$ and J.E. Han$^{(3)}$}
6: \address{${}^{(1)}$ Max-Planck-Institut f\"ur Festk\"orperforschung, 
7: Postfach 800665, D-70506 Stuttgart, Germany \\
8: ${}^{(2)}$Laboratoire de Min\'eralogie-Cristallographie, case 115, 
9: 4 Place Jussieu, 75252, Paris cedex 05, France \\
10: ${}^{(3)}$Department of Physics, The Pennsylvania State University,
11: University Park, PA 16802-6300}
12: 
13: 
14: \maketitle
15: 
16: \begin{abstract}
17: Resistivity saturation is observed in many metallic systems with  
18: large resistivities, i.e., when the resistivity has reached a critical
19: value, its further increase with temperature is substantially reduced. 
20: This typically happens when the apparent mean free path is comparable 
21: to the interatomic separations - the Ioffe-Regel condition. Recently, 
22: several exceptions to this rule have been found. Here, we review 
23: experimental results and early theories of resistivity saturation.
24: We then describe more recent theoretical work, addressing cases both 
25: where the Ioffe-Regel condition is satisfied and where it is violated. 
26: In particular we show how the (semiclassical) Ioffe-Regel condition 
27: can be derived quantum-mechanically under certain assumptions about 
28: the system and why these assumptions are violated for high-$T_c$
29: cuprates and alkali-doped fullerides.
30: \end{abstract}
31: 
32: \begin{multicols}{2}
33: \tableofcontents
34: \section{Introduction}\label{sec:a}
35: The electrical resistivity, $\rho$, of metals is usually calculated in
36: the Boltzmann theory, where the electrons are treated semiclassically.
37: An electron is assumed to move with a wave vector 
38: ${\bf k}$ between scattering events, caused by phonons,  
39: other electrons, impurities or some other disorder. The average 
40: distance an electron moves between two scattering events is    
41: the mean free path $l$. Assuming a three-dimensional system and a 
42: spherical Fermi surface with one sheet, $\rho$ can be expressed 
43: in terms of $l$ as (see Appendix A) 
44: \begin{equation}\label{eq:1}
45: \rho={3\pi^2 \hbar\over e^2k_F^2 l},
46: \end{equation}
47: where $k_F$ is the Fermi wave vector. Alternatively, if the resistivity 
48: is known experimentally, an apparent mean free path can be determined 
49: from Eq.  (\ref{eq:1}). In the following, Eq. (\ref{eq:1}) will
50: be used in the latter way to determine $l$, which then only depends 
51: on the experimental resistivity and the density of conduction electrons 
52: (giving $k_F$). 
53: 
54: 
55: \begin{figure}[t]
56: \centerline{
57: \rotatebox{-90}{\resizebox{2.0in}{!}{\includegraphics{fig1.eps}} }}
58: \caption[]{\label{fig:1}Resistivity of Cu, Nb$_3$Sb (Fisk and Webb, 
59: 1976) and Nb (Abraham and Deviot, 1972). The figure also shows the the 
60: Ioffe-Regel (Ioffe and Regel, 1960) saturation resistivities of Nb$_3$Sb
61: and Nb, setting the mean free path $l$ in Eq. (\ref{eq:1}) equal to the  
62: distance between the Nb atoms. The corresponding value for Cu, 260
63: $\mu\Omega$cm, falls outside the figure. The figure illustrates that 
64: for Nb$_3$Sb and Nb the resistivity saturates roughly as predicted by 
65: the Ioffe-Regel criterion, while $\rho(T) \sim T$ for Cu at large $T$.}
66: \end{figure}
67: 
68: For a good metal like Cu, $l$ is of the order of hundreds or even 
69: thousands of \AA. In the semiclassical Boltzmann theory and for $T$ 
70: larger than a fraction of a typical phonon energy, the resistivity 
71: due to phonon scattering grows linearly with $T$. This is shown           
72: for Cu in Fig. \ref{fig:1}. The apparent mean free path is then reduced 
73: correspondingly. For a good metal, however, $l$ is much larger than the 
74: separation $d$ of the atoms, even at the melting temperature. 
75: 
76: For certain metals, in particular many transition metals and 
77: transition metal compounds, the resistivity behaves in a completely 
78: different way. This was emphasized by Fisk and Webb (1976),     
79: who studied the A15 compounds Nb$_3$Sb (see Fig. \ref{fig:1}).
80: Similar results had earlier been obtained by Woodard and Cody (1964)
81: for Nb$_3$Sn and by several other groups for other compounds. 
82: For small $T$§, $\rho(T)$ grows much faster than for Cu and $l$     
83: becomes comparable to $d$ already for $T$ of the order of several 
84: hundred K. At this point, the slope of the resistivity curve is
85: substantially reduced, and $l$ stays comparable to 
86: $d$ for experimentally accessible values of $T$. This is referred to 
87: as resistivity saturation. It describes the situation where $\rho(T)$ 
88: grows much slower than $\rho(T)\sim T$, predicted by the Boltzmann 
89: equation, but it does not necessarily mean that $\rho(T)$ becomes 
90: a constant. 
91: 
92: Ioffe and Regel (1960) pointed out that the semiclassical
93: theory makes no sense if $l<d$, and we refer to $l \gtrsim d$
94: as the Ioffe-Regel condition. Inserting $l=d$ in Eq. (\ref{eq:1}) 
95: gives the Ioffe-Regel resistivity, shown in Fig. \ref{fig:1}. 
96: In a semiclassical picture, it is natural that $l$ cannot be much 
97: smaller than $d$, since one may expect that an electron at most is
98: scattered at every atom. Saturation is then expected 
99: when $l \sim d$. A semiclassical theory, however,  breaks down when 
100: $l\sim d$, since the uncertainty in the ${\bf k}$-vector
101: of an electron is comparable to the size of the Brillouin
102: zone. A semiclassical theory cannot therefore explain 
103: saturation. Nevertheless, as discussed in Sec. \ref{sec:b}, in the 
104: 1970's and early 1980's a large number of metals were found which showed 
105: saturation when $l\sim d$, and the behavior seemed to be universal. 
106: During this time, much theoretical work was performed to explain 
107: saturation. No consensus was reached, however, and the interest 
108: turned to other problems.                  
109: 
110: \begin{figure}[t]
111: \centerline{
112: \rotatebox{-90}{\resizebox{2.3in}{!}{\includegraphics{fig2.eps}} }}
113: \caption[]{\label{fig:2}Resistivity of
114: Bi$_2$Sr$_2$Ca$_{1-x}$Y$_x$Cu$_2$O$_{8+y}$  ($T_c=30$ K) (Wang 
115: {\it et al.}, 1996ab), La$_{1.93}$Sr$_{0.07}$CuO$_4$  (Takagi 
116: {\it et al.}, 1992), Nd$_{1.84}$Ce$_{0.16}$Cu$_{4-y}$ ($T_c=22.5$ K)
117: (Hikada and Suzuki, 1989), YBa$_2$Cu$_3$O$_{6+x}$ ($T_c=60$ K)
118: (Orenstein {\it et al.}, 1990), Bi$_2$Sr$_2$Cu$_{6+y}$ ($T_c=6.5$ K)
119: (Martin {\it et al.}, 1990) and Nb$_3$Sb (Fisk and Webb, 1976). 
120: The arrow shows the Ioffe-Regel resistivity of 
121: La$_{1.93}$Sr$_{0.07}$CuO$_4$. The figure illustrates 
122: that there is no sign of saturation at the Ioffe-Regel resistivity, 
123: but in some cases perhaps at much larger resistivities. Observe the 
124: magnitude compared with Nb$_3$Sb.}
125: \end{figure}
126: 
127: The situation changed drastically in the late 1980's, when it was found 
128: that the high-$T_c$ cuprates behave very differently (Gurvitch and Fiory, 
129: 1987). Fig. \ref{fig:2} shows that the resistivity for these compounds 
130: is typically much larger than for metals as Nb$_3$Sb, 
131: satisfying the Ioffe-Regel condition. As described in Appendix A, the 
132: Ioffe-Regel resistivity for La$_{2-x}$Sr$_x$CuO$_4$ is very large, 
133: about 0.7 m$\Omega$cm, due to the small carrier concentration. 
134: Nevertheless, the resistivities of the high-$T_c$ cuprates  
135: substantially exceed the Ioffe-Regel resistivity
136: or are comparable to this resistivity without signs
137: of saturation. The Ioffe-Regel condition is therefore strongly
138: violated. Both La$_{2-x}$Sr$_x$CuO$_4$ for small $x$ and 
139: Bi$_2$Sr$_2$Ca$_{1-x}$Y$_x$Cu$_2$O$_{8+y}$ show, however, signs 
140: of saturation, although at much larger values than the Ioffe-Regel 
141: resistivity. 
142: Fig. \ref{fig:3} shows other examples of the violation of the 
143: Ioffe-Regel condition, namely Rb$_3$C$_{60}$ (Hebard {\it et al.}, 
144: 1993), La$_4$Ru$_6$O$_{19}$ (Khalifah {\it et al.}, 2001) and
145: Sr$_2$RuO$_4$ (Tyler {\it et al.}, 1998).
146: 
147: 
148: \begin{figure}[t]
149: \centerline{
150: \rotatebox{-90}{\resizebox{2.2in}{!}{\includegraphics{fig3.eps}} }}
151: \caption[]{\label{fig:3}Resistivity of Rb$_3$C$_{60}$
152: (Hebard {\it et al.}, 1993) La$_4$Ru$_6$O$_{19}$ (Khalifah {\it et al.},
153: 2001), Sr$_2$RuO$_4$ (Tyler {\it et al.}, 1998), Nb$_3$Sb (Fisk and 
154: Webb, 1976) and the Ioffe-Regel resistivity for Rb$_3$C$_{60}$. There 
155: is no sign of saturation at the Ioffe-Regel resistivity, but 
156: La$_4$Ru$_6$O$_{19}$ may saturate at a much larger resistivity. }
157: \end{figure}
158: 
159: Although there are substantial uncertainties in the absolute values 
160: of the experimental resistivity, it is, nevertheless, clear that the 
161: Ioffe-Regel condition can be violated. This shows that the semiclassical 
162: argument for saturation is not just questionable, but actually gives 
163: wrong predictions for these systems. This emphasizes the need for a 
164: theory of saturation going beyond the semiclassical treatment of the 
165: electrons. Such a theory should also explain the violation of the 
166: Ioffe-Regel condition for the systems mentioned above. The drastic 
167: change in the experimental situation over the last 15 years has led 
168: to a renewed interest in resistivity saturation.                     
169: 
170: In Sec. \ref{sec:b}, we review the experimental results, and in Sec.
171: \ref{sec:c} early theoretical work. We then briefly describe methods
172: for calculating the resistivity in Sec. \ref{sec:ca}. 
173: Much of the theoretical work has started from 
174: a Boltzmann-like approach, assuming that the scattering mechanism 
175: is a relatively weak perturbation. In Sec. \ref{sec:d}, we describe 
176: an approach starting from the opposite limit of very strong scattering,
177: assuming that the Drude peak has been completely removed. This approach 
178: is based on the f-sum rule. It leads to an approximate upper limit 
179: $\rho_{\rm sat}$ to the resistivity, which usually has a weak $T$ 
180: dependence. Resistivity saturation then happens if the initial slope 
181: of $\rho(T)$ is so large that $\rho_{\rm sat}$ is reached at small 
182: values of $T$ and if $\rho_{\rm sat}$ has a weak $T$ dependence.
183: This approach is  applied to three models with different saturation 
184: behavior. In Sec. \ref{sec:e}, we describe the treatment of a model 
185: of weakly correlated electrons in a broad band, appropriate for many 
186: transition metal compounds. This model shows saturation in agreement 
187: with the Ioffe-Regel condition. We show how this condition can be derived 
188: quantum-mechanically by assuming i) noninteracting electrons and ii) $T\ll 
189: W$, where $W$ is the band width. In Sec. \ref{sec:f} we treat high-$T_c$ 
190: cuprates, for which assumption i) about noninteracting electrons is
191: violated. Using the $t-J$ model, we find that saturation can occur, but at  
192: much larger resistivities than predicted by the Ioffe-Regel condition. 
193: Other metals violating the Ioffe-Regel condition are discussed in 
194: Sec. \ref{sec:g}. In Sec. \ref{sec:h}, we describe a model of 
195: alkali-doped C$_{60}$, which shows no saturation. This result depends on
196: the intramolecular character of the phonons and on condition ii) ($T\ll W$)
197: being violated.
198: The relation to Anderson localization and Mott's minimum
199: conductivity is discussed in Sec. \ref{sec:i}. 
200: 
201: 
202: \begin{figure}
203: \centerline{
204: \rotatebox{0}{\resizebox{1.7in}{!}{\includegraphics{fig4.eps}} }}
205: \caption[]{\label{fig:4}Resistivity of Ti$_{1-x}$Al$_x$ alloys.
206: The figure suggests that the saturation resistivity is 
207: independent of the disorder (after Mooij, 1973).}
208: \end{figure}
209:  
210: \section{Experimental results}\label{sec:b}
211: Most cases of
212: resistivity saturation have been found for transition metal compounds.
213: Mooij (1973) made one of the first observations for Ti$_{1-x}$Al$_x$ 
214: (see Fig. \ref{fig:4}) and several other alloys. In addition, he found 
215: that the positive $T$ dependence of the resistivity becomes weaker or 
216: even negative for strongly disordered systems. Similar results were found 
217: by Arko {\it et al.} (1973) and by Tsuei (1986) for other compounds. 
218: These results suggest that compositional disorder and disorder due to 
219: thermally excited phonons can have a similar effect. Other early 
220: examples were found by Fisk and Lawson (1973).
221: 
222: As discussed in the introduction, resistivity saturation was
223: found for A15 compounds (Woodard and Cody, 1964; Marchenko, 1973; 
224: Fisk and Webb, 1976; Wiesmann {\it et al.}, 1977; Gurvitch {\it et al.}, 
225: 1978). Another class of systems showing resistivity saturation 
226: is the Chevrel phases (Martin {\it et al.}, 1978; Sunandana, 
227: 1979). Early measurements gave very large values for the resistivity, 
228: probably because of sample problems, but later measurements 
229: on single crystals gave resistivities comparable to those of A15 
230: compounds (Kitazawa {\it et al.}, 1981). Fig. \ref{fig:5} shows the 
231: resistivity of the 3d and 5d transition metals. Several of these 
232: metals show clear signs of saturation, in particular $\alpha$-Mn. Other
233: cases are less clear, and, e.g., the resistivity of W even bends slightly 
234: upwards, although it is already large. As discussed at the end of this 
235: section, W is nevertheless an example of saturation, if anharmonic 
236: effects are considered. The above metals can be considered as cases 
237: where the Ioffe-Regel condition is satisfied.
238: 
239:  
240: \begin{figure}[t]
241: \centerline{
242: \rotatebox{-90}{\resizebox{1.9in}{!}{\includegraphics{fig5a.eps}} }}
243: \centerline{
244: \rotatebox{-90}{\resizebox{1.9in}{!}{\includegraphics{fig5b.eps}} }}
245: \caption[]{\label{fig:5}Resistivity of 3d (a) and 5d (b) transition metals
246: (Bass, 1982).}  
247: \end{figure}
248: 
249: We next consider the in-plane resistivity of high-$T_c$ cuprates.
250: The resistivity depends strongly on the doping of the CuO$_2$ plane. 
251: Information about the 
252: doping is given by the superconductivity transition temperature $T_c$, 
253: which is quoted below, when available. As the doping is reduced below
254: the optimum value, $T_c$ drops and $\rho$ increases. In the case 
255: of La$_{2-x}$Sr$_x$CuO$_4$  we assume that the doping is given by $x$.
256: Results for several cuprates were given in Fig. \ref{fig:2}, and we 
257: give some further examples in Table \ref{table1}. 
258: These are examples of systems where the resistivity is comparable
259: to the Ioffe-Regel resistivity already at moderate values of $T$ and 
260: without showing signs of saturation.
261: 
262: \begin{minipage}{8.6cm}
263: \begin{table}
264: \caption{Resistivity $\rho(T)$ (in m$\Omega$cm) of high-$T _c$ 
265: cuprates. The measurment temperature $T$ and the superconductivity
266: transition temperature $T_c$ are given in K.\label{table1}}  
267: \begin{tabular}{lllll}
268: Compound & $T_c$  & $T$ &$\rho(T)$  & Reference \\
269: \hline
270: HgBa$_2$Ca$_0$Cu$_1$O$_{4+x}$ & 94 & 300 & 0.5  &  Daignere 
271: {\it et al.}, 2001 \\
272: HgBa$_2$Ca$_1$Cu$_2$O$_{6+x}$ &122 & 300 & 0.3  &  Yan      
273: {\it et al.}, 1998 \\
274: HgBa$_2$Ca$_2$Cu$_3$O$_{8+x}$ & 125 & 500 & 0.6  &  Carrington
275: {\it et al.}, 1994 \\
276: HgBa$_2$Ca$_3$Cu$_4$O$_{10+x}$ & 130 & 400 & 0.5  &  L\"ohle   
277: {\it et al.}, 1996 \\
278: Tl$_2$Ba$_2$CuO$_{6+y}$        &  80 & 300 & 1.3  &  Kubo       
279: {\it et al.}, 1991\\
280: Tl$_2$Ba$_2$CuO$_{6+y}$        &  80 & 270 & 0.6  &  Duan      
281: {\it et al.}, 1991\\
282: TlSr$_2$CaCu$_2$O$_{7-y}$      &  65 & 300 & 0.5  &  Kubo      
283: {\it et al.}, 1991\\
284: Bi$_2$Sr$_2$CaCu$_2$O$_{8+y}$  & 76  & 300 & 1.2  &  Chen      
285: {\it et al.}, 1998\\
286: \end{tabular}
287: \end{table}
288: \end{minipage}
289: 
290: 
291: Bi$_2$Sr$_2$Ca$_{1-x}$Y$_x$Cu$_2$O$_{8+y}$ is of particular interest,
292: since Wang {\it et al.} (1996ab) found saturation for the most 
293: underdoped samples, e.g., $x=0.11$ ($T_c=30$ K) as shown in Fig. 
294: \ref{fig:2}. Similar results have also been obtained by Forro 
295: {\it et al.} (2002). Other measurements only showed weak signs of 
296: saturation (Mandrus {\it et al.}, 1992; Ruan {\it et al.}, 2001). 
297: In the measurements of Ruan {\it et al.} (2001), the reason may have 
298: been that their $x=0.11$ sample had a higher $T_c=56$ K than that of 
299: Wang {\it et al.} (1996ab) ($T_c=30$ K), suggesting that the former 
300: sample was less underdoped. We conclude that the Ioffe-Regel condition
301: is violated for high-$T_c$ superconductors but that some systems with
302: very large resistivities may show signs of saturation at much larger 
303: values than the Ioffe-Regel resistivity.  
304: 
305:  
306: \begin{figure}[t]
307: \centerline{
308: \rotatebox{-90}{\resizebox{2.0in}{!}{\includegraphics{fig6.eps}} }}
309: \caption[]{\label{fig:fig6}Resistivity of CaRuO$_3$ (Klein {\it et al.},
310: 1999ab), CrO$_2$ (Rodbell {\it et al.}, 1966), VO$_2$ (Allen {\it et al.}, 
311: 1993) and SrRuO$_3$ (Allen {\it et al.}, 1996).  }
312: \end{figure}
313: 
314: Apart from the high-$T_c$ cuprates, a number of transition metal compounds 
315: have been found which violate the Ioffe-Regel condition. Fig. \ref{fig:3}
316: and Fig. \ref{fig:fig6} shows some examples. Some of these metals 
317: show signs of saturation (La$_4$Ru$_6$O$_{19}$, CrO$_2$ and perhaps 
318: CaRuO$_3$), but at higher resistivities than the Ioffe-Regel resistivity,
319: while other systems show no sign of saturation. Several of 
320: theses metals are believed to be non-Fermi liquids. The data in Figs. 
321: \ref{fig:3} and \ref{fig:fig6} were obtained from single crystals, except 
322: for Rb$_3$C$_{60}$ and CaRuO$_3$. This could be a reason why the
323: resistivity in Fig. \ref{fig:fig6} is much larger for CaRuO$_3$ 
324: than for the related SrRuO$_3$. Large resistivities are also found
325: in manganites (Salamon and Jaime, 2001). These systems raise interesting
326: problems in terms of the colossal magnetoresistance, possibly 
327: important polaron effects and many phase transitions, but they are 
328: outside the scope of this paper.   
329: 
330: The resistivity of A$_3$C$_{60}$ (A= K, Rb) shows a quite substantial 
331: spread between different measurements (Hebard {\it et al.}, 1993;
332: Hou {\it et al.}, 1993; Palstra {\it et al.}, 1994; Degiorgi {\it
333: et al}, 1994). These have been performed for thin films and doped 
334: single crystals using direct and optical methods. In particular
335: the thin films may have substantial defects (e.g., grain boundaries)
336: which would increase the resistivity. The optical measurements
337: should, however, be much less sensitive to this. Indeed, optical 
338: measurements for doped single crystals (Degiorgi {\it et al.}, 1994) 
339: gave resistivities that are about a factor of two smaller than those
340: obtained from direct measurements for thin films (Palstra {\it et 
341: al.}, 1994), but comparable to (somewhat larger than) direct measurements
342: for doped single crystals (Hou {\it at al.}, 1993).  This suggests 
343: that the thin film results in Fig. \ref{fig:3} could overestimate 
344: the resistivity of Rb$_3$C$_{60}$ by a factor of two or somewhat more. 
345: This would not change the qualitative conclusion  
346: that the Ioffe-Regel condition is violated for A$_3$C$_{60}$.
347: 
348: Experiments for Rb$_3$C$_{60}$ show no signs of saturation up to
349: about 500 K (Hebard {\it et al.}, 1993). It is difficult to reach
350: higher values of $T$, due to the possibility of thermally driven 
351: rearrangements of the alkali atoms. Hou {\it et al.} (1995), using
352: a pulsed heating technique to reach 800 K, found a small change 
353: in the slope at about $T=500$ K. This was interpreted as a sign of 
354: saturation. Using the parallel resistor formula (Wiesmann {\it et al.}, 
355: 1977), they deduced the saturation resistivity $6\pm 3$ m$\Omega$cm,  
356: corresponding to the mean free path $l=1\pm0.5$ \AA. Thus there is 
357: no saturation at $l \sim d$, where $d=10$ \AA \ is the separation of
358: two C$_{60}$ molecules (see Appendix A), but it is hard to judge 
359: if there is saturation at some larger resistivity. 
360: 
361: The Boltzmann equation predicts $\rho(T)\sim T$ for large $T$
362: only if various parameters, such as the electron-phonon interaction, 
363: are independent of $T$. This assumption is sometimes strongly 
364: violated. For instance, anharmonic effects reduce the phonon 
365: frequencies of W by about 30 $\%$ at the melting 
366: point (Grimvall, 2001; Grimvall {\it et al.}, 1987; Guillermet and 
367: Grimvall, 1991). According to the Boltzmann equation, this 
368: should increase $\rho(T)$  by about a factor of two, 
369: due to an increase in the electron-phonon coupling, giving a growth
370: of $\rho(T)$ which is much faster than linear. Experimentally,
371: however, the growth is only slightly faster than linear 
372: (Fig. \ref{fig:5}b). This suggests that saturation in W is masked 
373: by the reduction of the phonon frequencies (Grimvall, 2001).
374: 
375: The resistivity is usually measured at constant pressure.
376: As $T$ is increased, the crystal expands and various parameters
377: are changed. Sundqvist and Andersson (1990) converted the resistivity
378: of YBa$_2$Cu$_3$O$_{7-y}$ to constant volume data, which is the more
379: relevant quantity for theoretical discussions. They found
380: that this lowered the $T=500$ K resistivity by about 32 $\%$.
381: Although this conversion involves many uncertainties (Sundqvist
382: and Andersson, 1990), it was concluded that the pronounced 
383: linearity in the constant pressure data is lost. Actually, the
384: corrected data show signs of saturation. A similar correction
385: of the data for Rb$_3$C$_{60}$ (Vareka and Zettl,
386: 1994) changed the approximately  quadratic $T$ dependence found 
387: for constant pressure (see Fig. \ref{fig:3}) to an approximately 
388: linear dependence for constant volume. 
389: 
390: \section{Early theoretical work}\label{sec:c}
391: Resistivity saturation was studied theoretically very intensively
392: from the middle of the 1970's to the early 1980's. In this section
393: we review some of this work as well as some later work. We focus
394: on work related to the electron-phonon scattering, 
395: while work based on electron-electron scattering (Jarrell and Pruschke, 
396: 1994; Lange and Kotliar, 1999; Parcollet and Georges, 1999; Merino 
397: and McKenzie, 2000; Gunnarsson and Han, 2000) fall outside the scope 
398: of this review. 
399: 
400: Much of the early work has been reviewed by Allen (1980ab).     
401: Initially, explanations were proposed in terms of strong variations 
402: of the electronic structure on the energy scale of $k_BT$ or 
403: unusual anharmonic effects. These effects appear, however, to be 
404: too small to explain experiment, and as ever more examples of 
405: resistivity saturation were discovered, such theories had to be 
406: abandoned (Allen, 1980ab). 
407: 
408: Wiesmann {\it et al.} (1977) found empirically that the resistivity 
409: of systems such as Nb$_3$Sb can be rather accurately described by 
410: a parallel resistor formula
411: \begin{equation}\label{eq:rb1}
412: {1\over \rho(T)}= {1\over \rho_{\rm sat}}+{1\over \rho_{\rm ideal}(T)},
413: \end{equation}
414: where $\rho_{\rm sat}$ is a saturation resistivity and $\rho_{\rm 
415: ideal}(T)$ is the resistivity of the Boltzmann equation, i.e., linear 
416: in $T$ for large $T$. 
417: 
418: Mooij (1973) observed that strong disorder can lead to a negative
419: slope of $\rho(T)$ (Fig. \ref{fig:4}). Jonson and Girvin (1979), 
420: Girvin and Jonson (1980) and Imry  (1980) argued that this is due 
421: to an incipient Anderson transition. The disorder is assumed to be 
422: so large that the system is close to an Anderson transition for $T=0$. 
423: As $T$ is increased, inelastic scattering (phonon assisted hopping)
424: leads to a loss of phase information (Lee and Ramakrishnan, 1985)
425: and the system moves away from the Anderson transition. This was argued 
426: to lead to a reduction of the resistivity, as observed by Mooij (1973) 
427: for strongly disordered systems.
428: 
429: Weger and Mott (1985) and Weger (1985) proposed a theory in terms 
430: dehybridization of the $s$- and $d$-electrons at large values of $T$.
431: Laughlin (1982) presented a theory where exchange interaction reduced 
432: the density of states and increased the Fermi velocity in such a way 
433: that resistivity saturation was obtained.
434: 
435: Cote and Meisel (1978) and Morton {\it et al.} (1978) proposed
436: that an electron is only scattered by a phonon if the mean free
437: path of the electron is longer than the phonon wave length. As $T$ 
438: is increased and $l$ becomes shorter, an increasing number of phonons 
439: become inefficient scatters. With these assumptions, resistivity 
440: saturation can be derived. Later calculations treating the electrons 
441: quantum mechanically, as described before Eq.  (\ref{eq:reb6}), do not 
442: seem to support the basic assumption in this theory (Calandra and 
443: Gunnarsson, 2002). 
444: 
445: Several theories have been based on models with one
446: electronic state per unit cell.  Christoph and Shiller (1984) 
447: studied the Fr\"ohlich Hamiltonian. Using a truncated equation 
448: of motion approach, they included some higher order terms
449: in the electron-phonon interaction. In this way they derived
450: the parallel resistor formula (Eq. (\ref{eq:rb1})). As pointed
451: out by them, their $\rho_{\rm sat}\sim 0.84$ m$\Omega$cm is substantially
452: larger than what is observed for transition metal compounds, but 
453: it is comparable to what might be expected for a one band model, 
454: as in Eq. (\ref{eq:reb8}) below ($N_d=1$). Belitz and Schirmacher 
455: (1983) studied a generalized Fr\"olich model which also included 
456: impurity scattering. They included higher order terms by using a 
457: mode coupling technique in a memory function approach. They found 
458: a rather good agreement with experiment, but a substantially smaller 
459: $\rho_{\rm sat}$ than might have been expected from Eq. (\ref{eq:reb8}) 
460: for $N_d=1$. 
461: 
462: Ron {\it et al.} (1986) treated a Kronig-Penney chain, where 
463: the positions of the $\delta$-functions vibrate.                 
464: Using the Landauer formula they found saturation in the 
465: limit of large amplitude vibrations. Their $\rho_{\rm sat}$ depends, 
466: however, on the strength of the potentials, and it therefore appears 
467: that the Ioffe-Regel condition is not explained.
468: 
469: The Boltzmann equation only includes intraband transitions. 
470: Chakraborty and Allen (1979) and Allen and Chakraborty (1981) emphasized 
471: that resistivity saturation is typically found for systems where 
472: interband transitions are important. They included these transitions
473: by generalizing the Boltzmann equation, treating terms of the 
474: order $(d/l)^0$, while the Boltzmann equation only contains 
475: terms of the of the order $(d/l)^{-1}$. They demonstrated that 
476: this leads to a parallel resistor type of formula (Eq. (\ref{eq:rb1})) 
477: and estimated the saturation resistivity to be of the right order 
478: of magnitude. Due to the complexity of the equations, it was not 
479: possible to make explicit calculations and the sign of the $\rho_{\rm 
480: sat}$ in Eq. (\ref{eq:rb1}) could not be determined. They  
481: also pointed out that higher order terms in $(d/l)$ may be important. 
482: 
483: \begin{figure}[t]
484: \centerline{
485: \rotatebox{-90}{\resizebox{!}{3.5in}{\includegraphics{fig7.eps}}}}
486: \caption[]{\label{fig:fig7}Resistivity of the C$_{60}$ model for different
487: electron-phonon couplings $\lambda$ according to QMC calculations 
488: (Gunnarsson and Han, 2000; Calandra and Gunnarsson, 2002). The phonon 
489: frequency is $\omega_{ph}=0.1$ eV.  The figure illustrates the lack of 
490: saturation. For $\lambda=0.80$ the onset of superconductivity can be seen 
491: as a sharp downturn in $\rho(T)$ as $T$ is lowered, due to superconducting 
492: fluctuations. For $\lambda= 1.06$ and 1.32, the resistivity has a negative 
493: slope for small $T$, indicating an insulating system. }
494: \end{figure}
495: 
496: Millis {\it et al.} (1999) studied a Holstein like model, where 
497: the phonons couple to the electron level energies (in the following
498: called LE coupling).
499: They found that there is a change of slope in the $\rho(T)$ curve 
500: for the case of a strong electron-phonon coupling, and associated 
501: this with resistivity saturation. They observed, however, that 
502: the resistivity does not saturate and remarked that resistivity 
503: saturation is a misnomer. Gunnarsson and Han (2000) performed a 
504: quantum Monte-Carlo (QMC) calculation for a somewhat related model 
505: of A$_3$C$_{60}$ (A= K, Rb). This model included the partly occupied, 
506: three-fold degenerate $t_{1u}$ level as well as the LE coupling to a 
507: five-fold degenerate H$_g$ phonon. We refer to this as the C$_{60}$ model. 
508: The results, shown in Fig. \ref{fig:fig7}, are somewhat similar
509: to those of Millis {\it et al.} (1999), but they were considered as an 
510: example of lack of saturation. It was later shown (Calandra and 
511: Gunnarsson, 2001) that a more realistic description of saturation
512: can be obtained by using a coupling to the hopping integrals (HI 
513: coupling) (see Sec. \ref{sec:e}).  
514: 
515: 
516: \section{Calculation of resistivity}\label{sec:ca}
517: 
518: The resistivity is usually calculated in the Boltzmann theory.
519: The electrons are treated semiclassically, and are assumed to
520: be accelerated by the electric field between scattering processes,
521: caused by impurities, phonons or other electrons. This can be considered
522: as the lowest order theory in an expansion in $1/(k_Fl)$ (Kohn and 
523: Luttinger, 1957). Both because of this and because of its prediction
524: $\rho(T)\sim T$ for large $T$, it is clear that the Boltzmann equation
525: cannot be used for describing saturation. Instead, it is 
526: necessary to treat the electrons quantum-mechanically.
527: This can be done in the Kubo formalism. For an isotropic system,
528: this requires the calculation of a current-current correlation 
529: function (Mahan, 1990)
530: \begin{equation}\label{eq:1a}
531: \pi(i\omega)=-{1\over 3N\Omega}\int_0^{\beta}d\tau e^{i\omega \tau}
532: \langle T_{\tau} {\bf \hat j}(\tau) \cdot {\bf \hat j}(0)\rangle,
533: \end{equation}
534: where $N$ is the number of atoms, $\Omega$ is the volume per atom, 
535: $\beta=1/(k_BT)$, $k_B$ is the Boltzmann constant, $T_{\tau}$ is a 
536: time-ordering operator and ${\bf j}$ is the current operator. 
537: $\pi(i\omega)$ is analytically continued to real frequencies,           
538: giving $\pi_{\rm ret} (\omega)$. The optical conductivity is then 
539: \begin{equation}\label{eq:1b}
540: \sigma(\omega)=-{{\rm Im}\pi_{\rm ret}(\omega)
541: \over \omega},
542: \end{equation} 
543: and the resistivity is $1/\sigma(0)$.
544: 
545: $\pi(i\omega)$ can be calculated for imaginary times by using  
546: a determinantal quantum Monte-Carlo (QMC) method (Blankenbecler 
547: {\it et al.}, 1981). It is then analytically continued
548: to real frequencies by using a maximum entropy method (Jarrell and 
549: Gubernatis, 1996). This approach is very useful for establishing 
550: the saturation properties of a given model. 
551: 
552: To interpret the results, 
553: a simpler method is used, where the phonons are treated semiclassically 
554: (Calandra and Gunnarsson, 2001; 2002). A large supercell with $N$ atoms 
555: and periodic boundary conditions is considered. Each 
556: phonon coordinate is given a random displacement according to a 
557: Gaussian distribution with the width determined by $T$. This set 
558: of displaced coordinates (snap shot) defines a new Hamiltonian.
559: While the original Hamiltonian represents a complicated many-body 
560: problem due to the electron-phonon interaction, the dynamics of the 
561: phonons have been eliminated in the new Hamiltonian. Since we assume 
562: no electron-electron interaction, this is then a one-particle 
563: Hamiltonian for a disordered system. The corresponding eigenstates 
564: $|l\rangle$ and eigenvalues $\varepsilon_l$ are found. The Kubo-Greenwood
565: formula (Greenwood, 1958) is then used         
566: \begin{equation}\label{eq:reb6}
567: \sigma_{xx}(\omega)={2\pi\over N\Omega \omega}\sum_{ll^{'}}|\langle l|
568: \hat j_x| l^{'}\rangle |^2(f_l-f_{l^{'}})\delta(\hbar \omega -
569: \varepsilon_{l^{'}} +\varepsilon_l),
570: \end{equation}
571: where $\hat j_x$ is the current operator in the $x$ direction, and 
572: $f_l$ is the Fermi function for the energy $\varepsilon_l$.  
573: It is essential that in this approach only the phonons are treated
574: semiclassically and that the electrons are treated 
575: quantum-mechanically. This is in contrast to the semiclassical 
576: treatment of the electrons in the Boltzmann theory, which is unable 
577: to describe saturation. 
578: 
579: In the following sections we discuss a new theory for describing
580: different classes of materials with different saturation
581: properties (Gunnarsson and Han, 2000; Calandra and Gunnarsson, 
582: 2001; 2002; 2003).
583: 
584: \begin{figure}[t]
585: \rotatebox{-90}{\resizebox{2.0in}{!}{\includegraphics{fig8.eps}}} 
586: \caption[]{\label{fig:fig8}
587: $\sigma(\omega)$ as a function of $\omega$ for the Nb$_3$Sb model in the 
588: semiclassical treatment of the phonons. The frequency has been scaled  
589: by the $T=0$ band width $W$. The figure illustrates how the Drude peak 
590: is lost as $T$ is increased.}
591: \end{figure}
592: 
593: 
594: \section{\lowercase{f}-sum rule}\label{sec:d}
595: 
596: For small $T$, the optical conductivity $\sigma(\omega)$ typically 
597: has a Drude peak, due to intraband transitions between quasiparticles 
598: and centered at $\omega=0$, and a remaining incoherent part, as 
599: illustrated in Fig. \ref{fig:fig8} for Nb$_3$Sb.  As $T$ is increased, 
600: the height of the Drude peak is reduced, and for large 
601: $T$ it disappears. This is also illustrated in Fig. \ref{fig:fig9}, 
602: which shows the experimental optical conductivity of CrO$_2$ at $T=300$ K. 
603: The arrow marks the conductivity corresponding to the saturation 
604: resistivity (see Fig. \ref{fig:fig6}). This conductivity is similar 
605: to the magnitude of the incoherent contribution (at 300 K). Again,      
606: this suggests that as $T$ is   increased, the Drude peak disappears 
607: without the incoherent part changing very much.  In the following, we 
608: therefore assume that $T$ is so large that the Drude peak has disappeared
609: (Calandra and Gunnarsson, 2002). 
610: 
611: The Boltzmann theory starts from the periodic system and treats the 
612: scattering mechanisms as small perturbations. The theory focuses on 
613: intraband transitions and the Drude peak. Here we start from the    
614: the opposite limit, where the scattering is so strong that the Drude 
615: peak disappears.  
616: 
617: To estimate the incoherent contribution to $\sigma(\omega)$, we use 
618: the f-sum rule. We focus on tight-binding models with one type 
619: of electrons, e.g., $d$ electrons in a transition metal compound. 
620: Then there are no on-site matrix elements of the current operator. 
621: The off-site matrix elements can be related to hopping matrix 
622: elements, using charge and current conservation.  For an isotropic 
623: three-dimensional system, the tight-binding version of the f-sum rule 
624: then takes the form (Maldague, 1977; Calandra and Gunnarsson, 2002)
625: \begin{equation}\label{eq:rea1}
626: \int_0^{\infty} \sigma(\omega)d \omega=
627: {\pi\over 6}{d^2e^2 \over N\Omega\hbar^2}|E_K|,
628: \end{equation}
629: where $E_K$ is the hopping (kinetic) energy, $N$ is the number 
630: of atoms and $\Omega$ is the volume per atom. 
631: For a system with just nearest neighbor hopping and fixed atoms, 
632: $d$ is the separation of the sites. If the atoms vibrate    
633: or if there is hopping between more distant neighbors, $d$ is a weighted 
634: average of neighboring distances (Calandra and Gunnarsson, 2002). 
635: 
636: 
637: \begin{figure}[t]
638: \centerline{
639: \rotatebox{-90}{\resizebox{!}{2.7in}{\includegraphics{fig9.eps}}}}
640: \caption[]{\label{fig:fig9}$\sigma(\omega)$ for CrO$_2$ according
641: to experiment (full line) at $T=300$ K (Singley {\it et al.}, 1999) and 
642: the large $T$ model in Fig. \ref{fig:fig10} (dotted line), 
643: adjusting the area to the sum rule (\ref{eq:rea1}). The experimental 
644: intensity above $\omega\sim 1.5$ eV is due to interband transitions. 
645: The arrow shows the conductivity corresponding to the saturation
646: resistivity in Fig. \ref{fig:fig6}. It suggests that for large $T$, the 
647: Drude peak in the experimental $\sigma(\omega)$ disappears.
648: }
649: \end{figure}
650: 
651: For models with only one type of electrons, e.g, $d$ electrons,     
652: $\sigma(\omega)$ at large $T$ looks schematically as in Fig. 
653: \ref{fig:fig10}. This can be compared with the calculation for Nb$_3$Sb 
654: in Fig. \ref{fig:fig8} or the expected large $T$ contribution from the 
655: $t_{2g}$ band of CrO$_2$ in Fig.  \ref{fig:fig9} (Drude peak gone),
656: excluding the interband transitions for $\hbar\omega > 1.5$ eV. For 
657: $\hbar\omega$ larger than the band width $W$, $\sigma(\omega)\approx 0$. 
658: If $\sigma(\omega)$ furthermore lacks large structures around $\omega=0$, 
659: we can approximate 
660: \begin{equation}\label{eq:rea2}
661: \sigma(\omega)= \cases{ \sigma(0)/\gamma,&if $\hbar|\omega|\le W$;\cr
662: 0, &  otherwise.\cr}
663: \end{equation}
664: Then $\int \sigma(\omega)\hbar d\omega=\sigma(0)W/\gamma$, 
665: and the f-sum rule (\ref{eq:rea1}) requires that $\sigma(0)$
666: is given by the right hand side of Eq. (\ref{eq:rea1})
667: multiplied by $\gamma\hbar/W$ (Calandra and Gunnarsson, 2002; 2003), i.e., 
668: \begin{equation}\label{eq:rea3}
669: \sigma(0)={\gamma \hbar\over W}\int_0^{\infty} 
670: \sigma(\omega)d \omega=   {\pi\gamma\over 6W}{d^2e^2 
671: \over N\Omega\hbar}|E_K|,
672: \end{equation}
673: where $\gamma\sim 2$ depends on the shape of $\sigma(\omega)$. 
674: This provides an approximate lower limit to the conductivity 
675: (upper limit to the resistivity), which may be reached 
676: if $T$ is so large that the Drude peak is negligible. The conductivity 
677: could be appreciably smaller only if there is a substantial amount of 
678: weight for $\omega>W$ or if $\sigma(\omega)$ is anomalously small for  
679: $\omega=0$. The former should not be important, since 
680: $\hbar\omega>W$ requires multiple electron-hole pair excitations. Since 
681: the current operator is a one-particle operator, these excitations  
682: should have a small weight. The latter could happen in the case of an 
683: incipient Anderson transition, but it should normally not occur at large 
684: $T$, where inelastic scattering destroys the phase information required 
685: for an Anderson localization (Lee and Ramakrishnan, 1985). The (lower) limit 
686: (\ref{eq:rea3}) to the conductivity is $T$ dependent, due to the 
687: $T$ dependence of $E_K$ and $W$. This $T$ dependence is often, but not 
688: always, rather weak.  Below we use Eq. (\ref{eq:rea3}) to discuss
689: different classes of metals.                           
690: 
691: 
692: \begin{figure}[t]
693: \centerline{
694: \resizebox{2.0in}{!}{\includegraphics{fig10.eps}} }
695: \caption[]{\label{fig:fig10}
696: Schematic picture of $\sigma(\omega)$. The average of $\sigma(\omega)$   
697: over the band width is given by $\sigma(0)/\gamma$, where $\gamma\sim 2$.
698: }
699: \end{figure}
700: 
701: \section{Weakly correlated broad band systems}\label{sec:e}
702: 
703: We first focus on metals with a large band width $W$. We therefore
704: assume that i) $W$ is so large that the electrons can be treated 
705: as noninteracting and ii) $T\ll W$. 
706: In particular, we focus on  Nb$_3$Sb, showing a pronounced saturation, 
707: and Nb, showing a much weaker saturation (see
708: Fig. \ref{fig:1}). We use a model with an $N_d$-fold degenerate orbital 
709: on each site. The orbitals on different atoms are connected by hopping 
710: matrix elements, $t_{im,jm^{'}}$, where $i$ and $j$ label the atoms and 
711: $m, m^{'} = 1, .., N_d$ the orbital. Then the electronic Hamiltonian is 
712: \begin{equation}\label{eq:reb1}
713: H^{\rm el}=\sum_{i\ne j,mm^{'}\sigma}t_{im,jm^{'}} 
714: c^{\dagger}_{im\sigma}c_{jm^{'}\sigma},
715: \end{equation}
716: Three Einstein phonons are introduced for each atom, describing its 
717: displacement in the three coordinate directions. The phonons couple 
718: to the hopping integrals, since these depend on the atomic positions. 
719: We refer to this as HI coupling and the model is referred to as the 
720: transition metal (TM) model. The model for Nb has $4d$ orbitals 
721: ($N_d=5$) on a bcc lattice. For Nb$_3$Sb we use the Nb$_3^{\ast}$ model, 
722: neglecting the Sb atoms (Pickett {\it et al.}, 1979), and putting $4d$ 
723: orbitals on the Nb sites. The models are therefore identical except 
724: for the lattice structure. For strongly ionic systems, the coupling 
725: to the level energies (LE) may also be important. 
726: 
727: 
728: Fig. \ref{fig:fig11} shows the QMC (circles) results for the resistivity
729: of Nb$_3^{\ast}$ model for a supercell with $N=36$ atoms. These results 
730: extrapolate to a large resistivity 
731: at $T=0$. On the other hand, the resistivity of the model must be zero at 
732: $T=0$. It then follows that there must be a large change of the slope 
733: of $\rho(T)$ as $T$ is increased, i.e., resistivity saturation. 
734: This model can therefore be used for analyzing saturation. Fig. 
735: \ref{fig:fig11} also shows that the semiclassical treatment of the
736: phonons is quite accurate for the TM model, and that it therefore
737: can be used for interpreting the results.     
738: 
739: To use f-sum rule of Sec. \ref{sec:d}, we calculate the hopping energy 
740: $E_K$. Since we have assumed that the electrons are noninteracting 
741: and that $T\ll W$, we obtain 
742: \begin{equation}\label{eq:reb7}
743: E_K=2N_d\int_{-W/2}^{E_F}\varepsilon
744: N(\varepsilon)d\varepsilon\equiv -2\alpha N_d W N,
745: \end{equation}
746: where $E_F$ is the Fermi energy, $N(\varepsilon)$ is the density of 
747: states per atom, orbital and spin, $N_d$ is the orbital degeneracy,
748: and $\alpha$ depends on the shape 
749: of $N(\varepsilon)$ and the filling. Typically $\alpha\sim 0.1$ for 
750: a metal close to half-filling.  The dependence on filling is moderate
751: around half-filling (Calandra and Gunnarsson, 2002), but $\alpha=0$ 
752: for an empty or full band.  Inserting Eq. (\ref{eq:reb7}) in
753: Eq. (\ref{eq:rea3}) gives an approximate lower limit to the 
754: conductivity (Calandra and Gunnarsson, 2002; 2003)
755: \begin{equation}\label{eq:reb8}
756: \sigma_{\rm sat}(0)={\pi \alpha\gamma\over 3} 
757: {d^3\over \Omega}{N_de^2\over \hbar d}.
758: \end{equation}
759: Here $\pi\alpha\gamma/3 \sim 0.2 $ depends on the details of the electronic 
760: structure, and $d^3/\Omega\sim 1$ depends on the lattice structure. The 
761: result is independent of $W$, since the $W$ in Eq. (\ref{eq:rea3}) cancels 
762: that in Eq. (\ref{eq:reb7}). The essential material parameters, $d$
763: and $N_d$ appear in $N_de^2/(\hbar d)$, which has the dimension conductivity.
764: The corresponding apparent mean free path is (Calandra and Gunnarsson,
765: 2001)
766: \begin{equation}\label{eq:reb8a}
767: l=cN_d^{1\over 3}d,
768: \end{equation}
769: where $c\sim 0.6-0.8$, which is an approximate lower limit to $l$. It 
770: then follows that $l \gtrsim d$. This provides a quantum-mechanical 
771: derivation of the Ioffe-Regel condition, assuming that the electrons 
772: can be treated as noninteracting and that $T\ll W$. For a transition metal 
773: compound, with $N_d=5$ and $d\sim 3$ \AA, this leads to a resistivity of 
774: the order of 0.1-0.2 m$\Omega$cm, in agreement with the experiments. 
775: 
776: 
777: \begin{figure}[t]
778: \centerline{
779: \resizebox{2.5in}{!}{\includegraphics{fig11.eps}} }
780: \caption[]{\label{fig:fig11}Resistivity for the Nb$_3^{\ast}$ model of
781: Nb$_3$Sb, comparing the semiclassical $\lbrack$ broken ($N=36$ atoms) 
782: and full ($N=648$) curves$\rbrack$ and QMC (circles, $N=36$) treatment 
783: of the phonons. The figure also shows the 
784: small (Eq. (\ref{eq:reb9})) and large (Eq. (\ref{eq:reb8})) $T$         
785: results. It illustrates that the resistivity of the TM model 
786: saturates at large $T$ (after Calandra and Gunnarsson, 2001).}
787: \end{figure}
788: 
789: We next consider the small $T$ behavior of $\rho(T)$. In this limit 
790: the Boltzmann theory can be used, since $l\gg d$. For $T$ larger than 
791: some fraction of $\omega_{ph}$ (Grimvall, 1981) 
792: \begin{equation}\label{eq:reb9}
793: \rho(T)= 8\pi^2{\lambda T k_B\over \hbar \Omega_{pl}^2},
794: \end{equation}
795: where $\lambda$ is the dimensionless electron-phonon coupling constant 
796: and $\Omega_{pl}$ is the plasma frequency
797: \begin{equation}\label{eq:reb10}
798: (\hbar\Omega_{pl})^2={e^2\over 3\pi^2}\sum_{\nu} \int_{Bz}d^3k \lbrack 
799: {\partial \varepsilon_{\nu}({\bf k}) \over \partial {\bf k}}\rbrack^2\delta
800: (\varepsilon_{\nu}({\bf k})-E_F).
801: \end{equation}
802: Here $\varepsilon_{\nu}({\bf k})$ is the energy of a state with
803: the band index $\nu$ and the wave vector ${\bf k}$. $\Omega_{pl}$ 
804: depends on the average Fermi velocity.
805: 
806: The straight line corresponding to Eq. (\ref{eq:reb9}) is shown 
807: in Fig. \ref{fig:fig11}. It shows how the conductivity is reduced 
808: and the resistivity is increased due to the reduction of the Drude 
809: peak height as $T$ is increased. The horizontal line in Fig. 
810: \ref{fig:fig11} corresponds to the large $T$ limit (Eq. (\ref{eq:reb8})) 
811: of the conductivity. When this horizontal line is crossed by the 
812: steep line (Eq. (\ref{eq:reb9})), the Drude peak has been so strongly 
813: reduced that it disappears under the incoherent part, and is spread 
814: out over a large energy range. As discussed above, due to the f-sum rule, 
815: the conductivity cannot normally be much further reduced at this point. 
816: 
817: Saturation requires that the two lines (Eqs. (\ref{eq:reb8}, 
818: \ref{eq:reb9})) cross in the available temperature range. For most 
819: metals, the line in Eq. (\ref{eq:reb9}) has such a small slope that 
820: the crossing would happen far above the melting point. 
821: 
822: 
823: The model of Nb shows a much weaker saturation than the Nb$_3^{\ast}$ 
824: model of Nb$_3$Sb (Calandra and Gunnarsson, 2001), as is also found 
825: experimentally (see Fig. \ref{fig:1}). This is due to a much larger 
826: $\Omega_{pl}$ for Nb (9.5 eV in the model) than for Nb$_3$Sb (3.4 eV). 
827: The slope of the line in Eq. (\ref{eq:reb9}) is therefore a factor of 
828: five larger for Nb$_3$Sb than for Nb. This leads to the much more
829: pronounced saturation for Nb$_3$Sb. The difference in $\Omega_{pl}$
830: is due to the large unit cell of Nb$_3^{\ast}$, which leads to            
831: many bands and many forbidden crossings. This results in quite flat 
832: bands and small electron velocities and a small $\Omega_{pl}$
833: (Eq. (\ref{eq:reb10})). In this context it is interesting to 
834: note that $\alpha$-Mn has the most pronounced saturation among
835: the transition metals (see Fig. \ref{fig:5}a). The unit cell of 
836: $\alpha$-Mn has 58 atoms, suggesting a very small $\Omega_{pl}$ 
837: and a very steep initial slope of $\rho(T)$. Chakraborty and Allen 
838: (1979) and Allen and Chakraborty (1981) observed that saturation 
839: usually happens for systems with important interband transitions. 
840: Such systems typically have large unit cells and small $\Omega_{pl}$. 
841: Their observation can then be understood in terms of the arguments above. 
842: 
843: For a one-band model with one $s$-state per unit cell, 
844: $\rho_{\rm sat}$ is large ($N_d=1$ in Eq. (\ref{eq:reb8}) instead of 
845: $N_d=5$ for the TM model) and 
846: $\lambda/\Omega_{pl}$ is typically small. The crossing of the two
847: lines in Eqs. (\ref{eq:reb8}, \ref{eq:reb9}) then happens for such  
848: large vibration amplitudes that a calculation is not very meaningful.
849: The main reason is that for an $s$-band model, the matrix elements 
850: do not change sign as the atoms vibrate, and very large vibration 
851: amplitudes are needed to make the scattering so strong that the Drude 
852: peak essentially disappears. It is therefore interesting 
853: to study a one-band model where the state has many strongly directed 
854: lobes. Even moderate atomic vibrations then change the signs of 
855: the matrix elements. As a model we use one of the three $t_{1u}$ 
856: orbitals of a C$_{60}$ molecule (hypothetical case of a strong 
857: crystal-field splitting). This orbital has $l=5$ and correspondingly 
858: many nodes. The resulting resistivity saturation is moderate, with 
859: the slope at large $T$ being reduced by about a factor of three to four 
860: in the semiclassical treatment of the phonons. This illustrates that 
861: interband transitions are not indispensable for resistivity saturation.  
862: 
863: The optical conductivity in Figs. \ref{fig:fig8} and \ref{fig:fig9} 
864: can be viewed as a Drude peak riding on top of a broad 
865: structure of incoherent contributions. According to the f-sum rule and 
866: the arguments above, the height and shape of the latter part does not 
867: change much for values of $T$ discussed here.  
868: We then approximately have   
869: \begin{equation}\label{eq:d17}
870: \sigma(\omega=0,T)=\sigma_{\rm Drude}(\omega=0,T)+\sigma_{\rm sat},
871: \end{equation}
872: where $\sigma_{\rm Drude}(\omega=0,T)$ is the inverse of the resistivity
873: in Eq. (\ref{eq:reb9}) and $\sigma_{\rm sat}$ is the height of the 
874: incoherent contribution (Eq. (\ref{eq:reb8})). This is the parallel resistor 
875: formula (Wiesmann {\it et al.}, 1977). 
876: 
877: \section{Strongly correlated systems. High-$T_{\lowercase{c}}$ cuprates}\label{sec:f}
878: 
879: In the derivation of the Ioffe-Regel condition, we assumed i) noninteracting
880: electrons. We now focus on the high-$T_c$ cuprates, where this assumption
881: is invalid. The transport properties in the CuO$_2$ plane are mainly 
882: determined by the antibonding band of Cu $x^2-y^2$ and O $2p$ character. 
883: It is then natural to consider one ($x^2-y^2$) orbital per site and the 
884: orbital degeneracy $N_d=1$. These orbitals are put on a two-dimensional 
885: square lattice, describing the CuO$_2$ plane. The Coulomb 
886: interaction is described by a Hubbard $U$ acting between two electrons 
887: on the same site. $U$ is assumed to be so large that states with double 
888: occupancy on a site can be projected out, leading to the $t-J$ model 
889: (Zhang and Rice, 1988).  
890: 
891: \begin{figure}[t]
892: \centerline{
893: \rotatebox{-90}{\resizebox{!}{3.0in}{\includegraphics{fig12.eps}}}}
894: \caption[]{\label{fig:fig12} 
895: In-plane optical conductivity of La$_{2-x}$Sr$_x$CuO$_4$ as a function of 
896: $\omega$ for $x=0.06$ and $x=0.10$ (Takenaka {\it et al.}, 2002). The 
897: figure illustrates how the peak at $\omega=0$ gradually goes away as $T$ 
898: is increased and that the intraband contribution to $\sigma(\omega)$ 
899: is reduced as $x$ is reduced. The sharp structures at small $\omega$ are 
900: due to phonons. Interband transitions become important for $\hbar \omega
901: \gtrsim 1$ eV. }
902: \end{figure}
903: 
904: 
905: To use the f-sum rule, we first make a simple estimate of the hopping 
906: energy. Due to the doping with $x$ holes per site, a given site is 
907: occupied by a hole or an electron with the probabilities $x$ and $(1-x)$, 
908: respectively. In the $t-J$ model, a hole can hop to a neighboring site 
909: only if the site is occupied by an electron. This suggests a hopping 
910: probability of the order of $x(1-x)$. Multiplying by the number of 
911: nearest neighbors (four) and the hopping integral $t$ gives a simple 
912: estimate of the hopping energy (Calandra and Gunnarsson, 2003)
913: \begin{equation}\label{eq:rec2}
914:  E_K =-4tx(1-x)N,
915: \end{equation}
916: where $N$ is the number of sites. An exact diagonalization calculation 
917: for the $t-J$ model, gives a similar dependence on $x$ but a
918: somewhat smaller prefactor of about 3.4 (Calandra and Gunnarsson, 2003). 
919: 
920: Typically, $x \ll 1$. From Eq. (\ref{eq:rec2}) it then follows that 
921: the right hand side of the 
922: f-sum rule (Eq. (\ref{eq:rea1})) should be approximately proportional 
923: to $x$, as is also found experimentally (Yamada {\it et al.}, 1998). 
924: The constant of proportionality, extracted from  Uchida {\it et al.} 
925: (1991), agrees with the calculation to within about 20-30 $\%$ (Calandra 
926: and Gunnarsson, 2003). 
927: 
928: Fig. \ref{fig:fig12} shows $\sigma(\omega)$ of La$_{2-x}$Sr$_x$CuO$_4$
929: for different $T$ and $x$ (Takenaka {\it et al.}, 2002).
930: The figure illustrates how the peak at $\omega=0$ gradually disappears 
931: as $T$ is increased and that for large $T$, apart from phonon structures, 
932: $\sigma(\omega)$ behave roughly as assumed in Fig. \ref{fig:fig10}. 
933: A comparison of the curves
934: for $x=0.06$ and $x=0.10$, shows that $\sigma(\omega)$ is much smaller
935: for $x=0.06$ in the energy range $\omega\lesssim 1$ eV, where the 
936: intraband contributions dominate. This illustrates the reduction of 
937: the intraband contribution to $\sigma(\omega)$ when $x$ is reduced, 
938: as assumed above. Similar results were found for 
939: La$_{2-x}$(Ca,Sr)$_x$CaCuO$_{6+\delta}$ 
940: by Wang {\it et al} (2002).
941: 
942: We now analyze the resistivity in terms of the f-sum rule (Eqs. 
943: (\ref{eq:rea1}, \ref{eq:rea3})).  For conduction in the 
944: ab-plane of a cuprate, the factor $1/6$ in Eqs. (\ref{eq:rea1},
945: \ref{eq:rea3}) should be replaced by a factor $1/4$, since the 
946: the system is quasi-two-dimensional.                            
947: Insertion of Eq. (\ref{eq:rec2}), but with the prefactor 3.4, 
948: in Eq. (\ref{eq:rea3}) gives 
949: \begin{equation}\label{eq:rec3}
950: \rho_{\rm sat}={0.07 c \over x(1-x)} \approx 
951: {0.4\over x(1-x)} \hskip0.3cm {\rm m}\Omega{\rm cm},
952: \end{equation}
953: where we have used distance $c=6.4$ \AA \ between the CuO$_2$ planes
954: appropriate for La$_{2-x}$Sr$_x$CuO$_4$. For small $x$, the saturation 
955: resistivity in Eq. (\ref{eq:rec3}) is very much larger than the Ioffe-Regel 
956: resistivity, 0.7 m$\Omega$cm. It is also much larger than the 
957: result in Sec. \ref{sec:e} for weakly correlated transition metal 
958: compounds (Eq. (\ref{eq:reb8}))
959: \begin{equation}\label{eq:rec4}
960: \rho_{\rm sat} \sim {0.14 d\over N_d}\approx {0.4 \over N_d} \hskip0.3cm 
961: {\rm m}\Omega{\rm cm},
962: \end{equation}
963: where $d\sim 3$ is expressed in \AA.  This is partly due to the 
964: degeneracy being just $N_d=1$ for the $t-J$ model but $N_d=5$ for the 
965: TM model. Furthermore, the strong correlation drastically reduces the 
966: hopping energy in the cuprates, which gives the factor $x(1-x)$ in Eq. 
967: (\ref{eq:rec3}). Finally, the large separation of the CuO$_2$ planes 
968: also increases $\rho$. 
969: 
970: Fig. \ref{fig:fig13} compares the experimental results of Takagi (1992) 
971: for La$_{2-x}$Sr$_x$CuO$_4$ with the saturation resistivity in Eq. 
972: (\ref{eq:rec3}). Due to the factor $1/x$ in Eq. (\ref{eq:rec3}), all 
973: resistivities are multiplied by the doping $x$. For small $x$, 
974: $\rho_{\rm sat}$ is then a constant. The experimental resistivity is smaller 
975: than the predicted saturation resistivity. The same conclusion is obtained 
976: for other high-$T_c$ superconductors (Calandra and Gunnarsson, 2003).
977: The experimental data do therefore not demonstrate absence of saturation. 
978: On the contrary, for La$_{2-x}$Sr$_x$CuO$_4$ with $x=0.04$ and $x=0.07$ 
979: there are signs of saturation where saturation is expected to occur.
980: It is interesting that the curves $x\rho(T)$ for $x=0.04$ and $x=0.07$ 
981: fall almost on top of each other, giving a dependence $\rho\sim 1/x$, 
982: as suggested by the arguments above. For small dopings, saturation has 
983: also been reported (Wang {\it et al.}, 1996ab) for 
984: Bi$_2$Sr$_2$Ca$_{1-x}$Y$_x$Cu$_2$O$_{8+y}$ at similar values as for
985: La$_{2-x}$Sr$_x$CuO$_4$ (see Fig. \ref{fig:2}). 
986: 
987: 
988: \begin{figure}[t]
989: \hskip0.5cm
990: \rotatebox{-90}{\resizebox{!}{3.5in}{\includegraphics{fig13.eps}} }
991: \caption[]{\label{fig:fig13} Resistivity multiplied by the doping
992: $x$ for La$_{2-x}$Sr$_x$CuO$_4$ (Takagi {\it et al.}, 
993: 1992) for $x=0.04$ (full curve), $x=0.07$ (broken curve), $x=0.15$ 
994: (dotted curve) and $x=0.34$ (chain curve). The horizontal 
995: line shows the saturation resistivity (Eq. (\ref{eq:rec3})), 
996: $x\rho_{\rm sat}$, in the limit of small $x$. The figure illustrates 
997: that the saturation resistivity is not exceeded by the experimental 
998: data and that there are signs of saturation for small $x$ roughly 
999: where saturation is expected.}
1000: \end{figure}
1001: 
1002: Above we gave an upper limit to the resistivity without 
1003: specifying a scattering mechanism. For $x=0.15$, the resistivity 
1004: in the $t-J$ model is of the right order of magnitude, suggesting
1005: that electron-electron scattering can explain the resistivity, 
1006: while for $x=0.07$ and $x=0.04$ the resistivity is much too small 
1007: compared with experiment(Calandra and Gunnarsson, 2003). This 
1008: suggests that for $x=0.04$ and $x=0.07$, there is an additional 
1009: important scattering mechanism, beyond the electron-electron scattering 
1010: mechanism in the $t-J$ model, or that our small cluster (16 sites) 
1011: cannot describe some important scattering mechanism.  
1012: This mechanism apparently increases the resistivity so rapidly for 
1013: small $T$ that the upper limit (\ref{eq:rec3}) is approached and 
1014: the resistivity shows sign of saturation. The estimate of the upper limit 
1015: (\ref{eq:rec3}) is nevertheless correct, unless the additional 
1016: scattering mechanism appreciably influences the hopping energy. 
1017: The $t-J$ model alone only gives a rather weak saturation and only 
1018: at larger $T$ than seen in Fig. \ref{fig:fig13}. 
1019: 
1020: 
1021: \section{Violation of the Ioffe-Regel condition in other compounds}
1022: \label{sec:g}
1023: 
1024: Figs. \ref{fig:3} and \ref{fig:fig6} show that the Ioffe-Regel condition is 
1025: violated for several other compounds. We first discuss CrO$_2$, which 
1026: is a ferromagnet below $T_C=390$ K. According to band structure 
1027: calculations, three bands of mainly Cr t$_{2g}$ spin up character are
1028: occupied by two electrons (Lewis {\it et al.}, 1997; Mazin {\it et al.}, 
1029: 1999). We therefore consider a three-fold degenerate band with the width 
1030: $W=2.3$ eV (Lewis {\it et al.}, 1997) and without spin degeneracy. 
1031: Fig. \ref{fig:fig9} compares the experimental $\sigma(\omega)$ at 
1032: $T=300$ K (Singley {\it et al.}, 1999) with the large $T$ theoretical 
1033: model in Fig. \ref{fig:fig10}. The theoretical area, $\hbar^2 \int
1034: \sigma(\omega)d\omega =2.7$ (eV)$^2$ (for units see Appendix A) has 
1035: been obtained from the f-sum rule (Eq. (\ref{eq:rea1})), assuming           
1036: a semi-elliptical band with $W=2.3$ eV. We estimate the experimental 
1037: t$_{2g}$ contribution to the f-sum rule to be 2.2 (eV)$^2$, i.e., somewhat 
1038: smaller. As a result, our estimate of the saturation resistivity
1039: ($\rho_{\rm sat}=0.4$ m$\Omega$cm) is somewhat smaller than the
1040: experimental result ($\sim 0.6-0.7$ m$\Omega$cm). The reason is 
1041: probably both that the area of $\sigma(\omega)$ is somewhat 
1042: overestimated and that the theoretical shape of $\sigma(\omega)$ does 
1043: not take into account a strong dip in $N(\varepsilon)$ at $E_F$ 
1044: for CrO$_2$. These considerations suggest that theory gives
1045: a qualitatively correct description of saturation for CrO$_2$. 
1046: The strong increase in $\sigma(\omega)$ for $\omega\sim 1.5$
1047: eV is due to interband transitions, which are neglected in the model.    
1048: 
1049: This treatment assumed CrO$_2$ to be ferromagnetic. If we had taken
1050: into account that CrO$_2$ is paramagnetic for $T>T_C$, the theoretical
1051: saturation resistivity would have dropped by a factor of two,
1052: due to an increase of $|E_K|$. Experimentally, however, the resistivity 
1053: changes little at $T=T_C$, suggesting that correlation effects reduce 
1054: $|E_K|$ above $T_C$ in a way similar to that of magnetic effects below 
1055: $T_C$. Actually, the Hubbard $U$ of CrO$_2$ has been estimated to 
1056: $U\approx 3$ eV (Korotin {\it et al.}, 1998), comparable to the 
1057: $t_{2g}$ band width, and suggesting appreciable correlation effects. 
1058: There are rather strong magnetic fluctuations in several of the systems 
1059: violating the Ioffe-Regel condition. The arguments above suggest that 
1060: this may reduce $|E_K|$ and increase the saturation resistivity.
1061: 
1062: Using similar assumptions (correlation reduces $|E_K|$ by a factor
1063: of two in the paramagnetic state) as for CrO$_2$, we predict a similar 
1064: saturation resistivity for VO$_2$ and a somewhat larger value
1065: (0.6 m$\Omega$cm) for CaRuO$_3$ and SrRuO$_3$. As for CrO$_2$, more 
1066: experimental data is needed to decide whether this agrees qualitatively with 
1067: experiment. It is not clear if this framework (strong correlation
1068: effects reducing $|E_K|$)  can be used to understand 
1069: Sr$_2$RuO$_4$, which shows no signs of saturation up to $T=1300$ K, and 
1070: La$_4$Ru$_6$O$_{19}$, which shows signs of saturation, but at a very 
1071: large resistivity.
1072:  
1073: \section{Very large $T$ behavior. C$_{60}$ compounds}\label{sec:h}
1074: 
1075: To derive the Ioffe-Regel condition, we assumed that ii) $T\ll W$, leading 
1076: to a $T$ independent upper limit $\rho_{\rm sat}$ to the resistivity. 
1077: We now consider $T$ to be so large that the phonons substantially 
1078: change $W$. Such effects are usually not very important, but the C$_{60}$ 
1079: model is an important exception, due to its small band width. The $T$
1080: dependence of $\rho_{\rm sat}$ can then be so strong that the concept
1081: of resistivity saturation becomes meaningless. In the following 
1082: we neglect the electron-electron interaction, although it could
1083: play a substantial role for alkali-doped fullerides. We treat
1084: the phonons in the semiclassical approximation (Sec. \ref{sec:ca}).
1085: 
1086: For very large $T$, there is a rather trivial $T$ dependence
1087: due to the Fermi temperature, $T_F$, entering in the Fermi-functions
1088: of Eq. (\ref{eq:reb6}). Let us consider the resistivity due to 
1089: static disorder. Expanding the Fermi functions in $1/T$, we
1090: obtain $\sigma(0)\sim 1/T$ and $\rho(T)\sim T$,
1091: although the scattering mechanism itself is $T$-independent. 
1092: A similar dependence also enters for the electron-phonon
1093: scattering, which tends to mask interesting differences between
1094: couplings to the level energies (LE) and hopping integrals (HI). 
1095: We therefore study the case when the Fermi functions are replaced
1096: by $\Theta$-functions in Eq. (\ref{eq:reb6}) ($T_F=0$).
1097: 
1098: 
1099: 
1100: \begin{figure}[t]
1101: \centerline{
1102: \rotatebox{-90}{\resizebox{!}{3.0in}{\includegraphics{fig14.eps}}}}
1103: \caption[]{\label{fig:fig14}Resistivity for a model of ordered C$_{60}$ 
1104: molecules, considering couplings 
1105: to the level energies (full line, LE coupl.) and to the hopping 
1106: integrals (broken line, HI coupling) in the semiclassical treatment 
1107: of the phonons. Results are also shown for $T_F=0$, replacing 
1108: the Fermi functions in Eq. (\ref{eq:reb6}) by $\Theta$-functions. 
1109: The figure shows that there is a large difference between LE and HI 
1110: coupling (after Calandra and Gunnarsson, 2002).}
1111: \end{figure}
1112: 
1113: We consider the C$_{60}$ model (see Sec. \ref{sec:c}) with either 
1114: LE or HI coupling (Calandra and Gunnarsson, 2002), using the same 
1115: coupling $\lambda$ in both cases (Fig. \ref{fig:fig14}). While the 
1116: resistivity shows no signs of saturation for the LE coupling (full curve),
1117: the model with HI coupling shows a weak saturation (broken curve),
1118: i.e.,  a moderate reduction of the slope for $T\gtrsim 0.03$ eV.
1119: This saturation becomes much more pronounced for $T_F=0$ (dotted curve). 
1120: For the TM model, the slope of $\rho(T)$ is reduced at high $T$ for both 
1121: HI and LE coupling, but the reduction is more pronounced for HI coupling.
1122: 
1123: To apply the analysis of Sec. \ref{sec:d}, we calculate $W(T)$ and 
1124: $E_K(T)$. $W(T)$ increases with $T$ for both LE and HI coupling. 
1125: In the LE case this happens because the fluctuations in the level 
1126: positions increase and in the HI case because the average of the 
1127: square of the hopping integrals increases. Since the average phonon 
1128: amplitude squared is $\langle x^2\rangle \sim T$, the second moment 
1129: $S_2(T)$ of the density of states goes as $S_2(T)=S_2(0)+aT$, where 
1130: $a$ is some constant. This leads to (Calandra and Gunnarsson, 2002)
1131: \begin{equation}\label{eq:de6}
1132: W(T)=W(0)\sqrt{1+c\lambda {k_B T\over W(T=0)}},
1133: \end{equation}
1134: The $T$ dependence comes in the form $\lambda T/W$, where 
1135: the large prefactor $c=12$ is due to the large ratio $W^2/S_2\sim 12$. 
1136: Using the considerations of Sec. \ref{sec:e}, it follows 
1137: that for the HI coupling $E_K$ varies in a similar way, and the $T$ 
1138: dependences of $E_K$ and $W$ essentially cancel (Eq. (\ref{eq:rea3})). 
1139: This is confirmed by Fig. \ref{fig:fig14} for the C$_{60}$ model 
1140: (lowest HI curve), while the cancellation is less complete for the TM 
1141: model due to a stronger $T$ dependence of $\alpha$ (Eq. (\ref{eq:reb7})) 
1142: and $\gamma$ (Eq. (\ref{eq:rea3})). With LE coupling, on the other hand, 
1143: $E_K$ is reduced as $T$ is increased. As the phonons are excited, the 
1144: level energies on different sites become different and hopping 
1145: becomes more difficult. For the C$_{60}$ model  $|E_K|\sim 1/
1146: \sqrt{1+c\lambda k_BT/W(T=0)}$ (Calandra and Gunnarsson, 2002). 
1147: For the LE coupling, the hopping energy and the band width work 
1148: together, and the upper limit to the resistivity takes the form
1149: \begin{equation}\label{eq:de11}
1150: \rho_{\rm sat}(T)={0.8\over \gamma(T)}(1+c \lambda {k_BT\over W(T=0)})
1151: \hskip0.5cm {\rm m}\Omega{\rm cm}.
1152: \end{equation}
1153: 
1154: In A$_3$C$_{60}$ (A= K, Rb) there is orientational disorder, i.e.,
1155: the C$_{60}$ molecules more or less randomly take one out of two 
1156: preferred orientations (Stephens {\it et al.}, 1991). As a result, 
1157: the Drude peak is essentially gone already for $T=0$.          
1158: The resistivity can therefore be considered
1159: as ``saturated'' already at $T=0$. The upper limit for the resistivity
1160: (Eq. (\ref{eq:de11})) has, however, such a strong $T$ dependence 
1161: that the term ``saturation'' becomes meaningless in this case
1162: (see Fig. \ref{fig:fig7}). The reason for this strong $T$ dependence 
1163: is both the small band width and the LE coupling to the intramolecular
1164: phonons. The $T$ dependence of the TM model is weaker, due to the larger 
1165: band width and to the HI coupling. Even in this case, however, $\rho(T)$ 
1166: does not become a constant because of the $T$ dependence of $\alpha$ 
1167: and $\gamma$.  
1168: 
1169: \section{Anderson metal-insulator transition and Mott's
1170: minimum conductivity.} \label{sec:i}   
1171: 
1172: In the semiclassical treatment of the phonons, the phonons cause a 
1173: static disorder, and the problem is therefore related to 
1174: conduction in disordered system. Thus the LE and HI couplings 
1175: correspond to diagonal and off-diagonal disorder, respectively. 
1176: While the disordered systems are usually studied for small $T$, 
1177: we are here interested in the large $T$ behavior. In the semiclassical 
1178: treatment of the phonons, however, apart from causing disorder, 
1179: $T$ only enters via the Fermi-functions (Eq. (\ref{eq:reb6})), which 
1180: is not important for the qualitative behavior. 
1181: 
1182: Diagonal disorder can lead to an Anderson metal-insulator
1183: transition at $T=0$ (Lee and Ramakrishnan, 1985). For 
1184: off-diagonal disorder, however, Antoniou and Economou (1977)             
1185: found that there is no metal-insulator transition if
1186: the Fermi energy is located in some finite region around
1187: the middle of the band. The semiclassical calculations
1188: agree with these results, i.e., localization is found for
1189: LE but not for HI coupling as $T$ is increased.
1190: 
1191: In the QMC calculation there is no sign of localization 
1192: for LE coupling, just a lack of saturation.  This is natural. 
1193: Localization depends sensitively on the phase factors, which 
1194: are not destroyed by the elastic scattering in a disordered 
1195: system. The phase information is, however, lost in the inelastic 
1196: scattering by phonons at finite $T$, and localization is not 
1197: expected (Lee and Ramakrishnan, 1985). These effects 
1198: is properly included in the QMC but not in the 
1199: semiclassical treatment of the phonons, and therfore localization 
1200: shows up in the semiclassical (Calandra and Gunnarsson, 2002) but 
1201: not in the QMC treatment (Gunnarsson and Han, 2000).  
1202: 
1203: Mott (1974) has argued that as the disorder increases, there is 
1204: a discontinuous transition from a metal to an insulator at $T=0$.  
1205: He therefore introduced the concept of the minimum conductivity
1206: \begin{equation}\label{eq:ree1}
1207: \sigma_{\rm min}=0.026 {e^2\over \hbar d},
1208: \end{equation}
1209: where $d$ is the nearest neighbor atomic distance. Later work showed
1210: that the transition from a metal to an insulator actually 
1211: is continuous, but that $\sigma_{\rm min}$ is still relevant  
1212: for low but nonzero temperatures (Lee and Ramakrishnan, 1985).
1213: We therefore make a comparison of $\sigma_{\rm min}$  to the 
1214: resistivity in the TM and C$_{60}$ models. Converting Eq. 
1215: (\ref{eq:ree1}) to a resistivity, we obtain
1216: \begin{equation}\label{eq:ree2}
1217: \rho_{\rm max}=1.6 d \ \ {\rm m}\Omega{\rm cm},
1218: \end{equation}
1219: where $d$ is measured in \AA.
1220: Based on experiment, Mott deduced a somewhat larger minimum conductivity
1221: for systems containing transition metal atoms, 
1222: resulting in the maximum resistivity
1223: \begin{equation}\label{eq:ree3}
1224: \rho_{\rm max}=1 \ \ {\rm m}\Omega{\rm cm}.
1225: \end{equation}
1226:    
1227: Mott derived his result for diagonal disorder. His result can most
1228: naturally be compared with our saturation resistivity for HI coupling
1229: (off-diagonal disorder), since saturation is most pronounced in this 
1230: case. The resistivity $\rho_{\rm max}$ is much larger than the saturation 
1231: resistivity obtained above (Eq. (\ref{eq:reb8}, \ref{eq:rec4})) for the 
1232: TM model with a five-fold degenerate orbital ($N_d=5$). For a fcc lattice 
1233: and a half-filled semi-elliptical band it takes the form
1234: \begin{equation}\label{eq:ree4}
1235: \rho_{\rm sat}={0.14 d \over N_d} \ \ {\rm m}\Omega{\rm cm},
1236: \end{equation}
1237: which is of the order of 0.1-0.2 m$\Omega$cm. The corresponding 
1238: conductivity is substantially larger than Mott's minimum conductivity.
1239: 
1240: \section{Conclusions}\label{sec:j}
1241: We have reviewed experiments showing resistivity saturation, i.e.,
1242: $\rho(T)$ growing more slowly than $\rho(T)\sim T$ for large $T$. 
1243: Resistivity saturation is found for several classes of metals with 
1244: large resistivities, in particular for many transition metal 
1245: compounds. Saturation often happens in such a way that the Ioffe-Regel 
1246: condition, $l\gtrsim d$, remains fulfilled. Over 
1247: the last 15 years, however, a number of metals have been found 
1248: for which the Ioffe-Regel condition is violated. Some of these 
1249: metals probably show resistivity saturation, but at a much larger
1250: value than the Ioffe-Regel resistivity, while in other cases
1251: little or no sign of saturation is seen.  
1252: 
1253: We have reviewed early theories, presented at a time
1254: when saturation appeared to be universal. Several of the theories
1255: derived saturation. These theories emphasized different
1256: mechanisms for saturation, and no consensus was reached about
1257: which mechanism dominates.             
1258: 
1259: Here we argue that it is useful to study the problem
1260: using the f-sum rule. By assuming that the (Drude) peak at 
1261: $\omega=0$ is gone and that only incoherent contributions are left, 
1262: we obtain an approximate upper limit, $\rho_{\rm sat}$, to the resistivity, 
1263: which usually has a weak $T$ dependence. Saturation then happens if 
1264: the resistivity initially grows so rapidly that $\rho_{\rm sat}$ is
1265: reached for small values of $T$ and if $\rho_{\rm sat}$ has a 
1266: weak $T$ dependence.    
1267: 
1268: We have considered three models with qualitatively different 
1269: behavior: 1) A model of weakly correlated transition metal 
1270: compounds, which shows saturation in agreement with the 
1271: Ioffe-Regel condition, 2) a model of strongly correlated 
1272: high-$T_c$ cuprates, which can give saturation but at much
1273: larger values than the Ioffe-Regel resistivity, and 3) a 
1274: model of alkali-doped C$_{60}$ compounds, which shows no 
1275: saturation. 
1276: 
1277: To derive the Ioffe-Regel condition for model 1)
1278: we assumed i) noninteraction electrons and ii) $T\ll W$.  
1279: Assumption i) is violated for model 2) and assumption ii)
1280: for model 3). The type of electron-phonon coupling, i.e., 
1281: coupling to the level energies or the hopping integrals,
1282: is also important when comparing models 1) and 3). 
1283: 
1284: We have focused on work where either the electron-phonon or the 
1285: electron-electron interaction was treated alone.  In many cases 
1286: this may be an oversimplification. For instance, the electron-electron 
1287: interaction is believed to be important for alkali-doped fullerides 
1288: (Gunnarsson, 1997), although only the electron-phonon interaction was 
1289: considered here. Generally, the electron-electron interaction should 
1290: reduce the magnitude of the hopping energy and tend to increase the 
1291: resistivity, as was found for the high-$T_c$ cuprates. The interplay 
1292: between the electron-electron and electron-phonon interactions may, 
1293: however, be more intricate, as found for metal-insulator transitions 
1294: (Han {\it et al.}, 2000) and superconductivity (Han {\it et al.}, 2003) 
1295: in alkali-doped fullerides.
1296: 
1297: 
1298: \appendix
1299: \section{Mean-free path}
1300: Since the mean-free path and the Ioffe-Regel resistivity play
1301: an important role in the discussion, we give details     
1302: of how these quantities were obtained. The conductivity tensor 
1303: can be written as (Ashcroft and Mermin, 1976)
1304: \begin{equation}\label{eq:2}
1305: {\bf \sigma}=e^2\sum_{\nu}\int {d^3k\over 4\pi^3}\tau_{\nu}({\bf k})
1306: {\bf v}_{\nu}({\bf k}){\bf v}_{\nu}({\bf k})\lbrack -{\partial f\over 
1307: \partial \varepsilon}\rbrack_{\varepsilon=\varepsilon_{\nu}({\bf k})},
1308: \end{equation}
1309: where $f$ is the fermi function, $\tau_{\nu}({\bf k}$) is the relaxation 
1310: time and ${\bf v}_{\nu}({\bf k})$ is the velocity for a state with band 
1311: index $\nu$ and wave vector ${\bf k}$. We assume a 
1312: three-dimensional isotropic system and a spherical Fermi surface 
1313: with one sheet. We furthermore assume that $\tau$ is independent of
1314: $\nu$ and ${\bf k}$. Together with $v_F=\hbar k_F/m$, where $v_F$ is the 
1315: Fermi velocity, and $l=\tau v_F$, this leads to Eq. (1).  Assuming that 
1316: there are $M$ sheets, we find a moderate reduction of the apparent 
1317: mean free path from Eq. (\ref{eq:1}) by a factor $M^{1/3}$. 
1318: For a quasi-two-dimensional system, like the high-$T_c$ cuprates, 
1319: we instead assume a cylindrical Fermi surface with the height 
1320: $2\pi/c$ and radius $k_F$, where $c$ is the average separation 
1321: of the CuO$_2$ planes. Based on photoemission results (Ino {\it et al.}, 
1322: 1999; Yoshida {\it et al.}, 2001), we assume a ``large'' Fermi surface 
1323: containing roughly one electron or hole (more precisely $1\pm x$ carriers, 
1324: where $x$ is the doping). This leads to $k_F=\sqrt{2\pi}/a$, 
1325: where $a$ is the lattice parameter of the CuO$_2$ plane. Then
1326: \begin{equation}\label{eq:3}
1327: \rho_{2d}={2\pi \hbar c\over e^2k_F l}.
1328: \end{equation} 
1329: Assuming the Ioffe-Regel condition $l=a$, we find           
1330: \begin{equation}\label{eq:4}
1331: \rho_{2d}^{Ioffe}=0.055 (c/a_0) \hskip0.5cm {\rm m}\Omega{\rm cm},
1332: \end{equation}
1333: where $a_0=0.529$ \AA \ is the Bohr radius and we have used the
1334: conversion $\hbar a_0/e^2=0.022$ m$\Omega$cm. Assuming $c=6.4$ \AA, 
1335: appropriate for La$_{2-x}$Sr$_x$CuO$_4$, we obtain 
1336: $\rho_{2d}^{Ioffe}=0.7$ m$\Omega$cm. If we instead had assumed 
1337: a ``small'' Fermi surface, containing $x$ carriers, the resistivity
1338: would have been $0.7/\sqrt{x}$. 
1339: Even in this case, the experimental
1340: resistivity of La$_{2-x}$Sr$_x$CuO$_4$ in Fig. \ref{fig:2} exceeds 
1341: the Ioffe-Regel resistivity for small $x$.   
1342: 
1343: For A$_3$C$_{60}$ (A= K, Rb),
1344: the Ioffe-Regel resistivity was calculated by assuming that $l$ 
1345: is the separation of two C$_{60}$ molecules. Scattering inside the 
1346: molecule is not possible at small and intermediate $T$, since this 
1347: would involve scattering into states that are at least 10000 K higher 
1348: in energy. Therefore the intermolecular separation is the appropriate 
1349: length scale.
1350: 
1351: \parindent -10pt
1352: \vskip0.4cm
1353: {\bf REFERENCES}
1354: \vskip0.4cm
1355: 
1356: Abraham, J.M., and B. Deviot, 1972, J. Less-Common Metals {\bf 29}, 311.
1357: 
1358: Allen, P.B., 1980a, in {\it Superconductivity in d- and
1359: f-Band Metals} H. Suhl and M.B. Maple, Eds. (Academic, New York)
1360: p. 291.
1361: 
1362: Allen, P.B., 1980b, in {\it Physics of Transition Metals}, P. Rhodes, 
1363: Ed. (Inst. Phys. Conf. Ser. No. 55) p. 425.
1364: 
1365: Allen, P.B., and B. Chakraborty, 1981, Phys. Rev. B {\bf 23}, 4815.
1366: 
1367: Allen, P.B., R.M. Wentzcovitch, W.W. Schulz, and P.C. Canfield, 1993,
1368: Phys. Rev. B {\bf 48}, 4359.
1369: 
1370: Allen, P.B., H. Berger, O. Chauvet, L. Forro, T. Jarlborg, A. Junod,
1371: B. Revaz, and G. Santi, 1996, Phys. Rev. B {\bf 53}, 4393.
1372: 
1373: Antoniou, P.D., and E.N. Economou, 1977, Phys. Rev. B {\bf 16}, 3768.
1374: 
1375: Arko, A.J., F.Y. Fradin, and M.B. Brodsky, 1973, Phys. Rev. B {\bf 8}, 4104.
1376: 
1377: Ashcroft, N.W., and N.D. Mermin, 1976 {\it Solid State Physics}
1378: (Holt, Rinehart and Winston, New York), p. 259. 
1379: 
1380: Bass, J., 1982, in {\it Landolt-B\"ornstein: Numerical data and functional 
1381: relationships in science and technology}, New Series III/15a, edited 
1382: by K.-H. Hellwege and J.L. Olsen, (Springer, Berlin), p. 1. 
1383: 
1384: Belitz, D., and W. Schirmacher, 1983, J. Phys. C: Solid State Phys. 
1385: {\bf 16}, 913.  
1386: 
1387: 
1388: Blankenbecler, R., D.J. Scalapino, and R.L. Sugar, 1981, Phys. Rev. D 
1389: {\bf 24}, 2278.
1390: 
1391: Calandra, M., and O. Gunnarsson, 2001, Phys. Rev. Lett. {\bf 87}, 266601.
1392: 
1393: Calandra, M., and O. Gunnarsson, 2002, Phys. Rev. B {\bf 66}, 205105.
1394: 
1395: Calandra, M., and O. Gunnarsson, 2003, Europhys. Lett. {\bf 61},88.
1396: 
1397: Carrington, A., D. Colson, Y. Dumont, C. Ayache, A. Bertinotti, 
1398: and J.F. Marucco, 1994, Physica C {\bf 234}, 1. 
1399: 
1400: Chakraborty, B., and P.B. Allen, 1979, Phys. Rev. Lett. {\bf 42}, 736.
1401: 
1402: Chen, X.H., M. Yu, K.Q. Ruan, S.Y. Li, Z. Gui, G.C. Zhang, and
1403: L.Z. Cao, 1998, Phys. Rev. B {\bf 58}, 14219.
1404: 
1405: Christoph, V., and W. Schiller, 1984, J.Phys. F: Met. Phys. {\bf 14}, 1173.
1406: 
1407: Cote, P.J., and L.V. Meisel, 1978, Phys. Rev. Lett. {\bf 40}, 1586.
1408: 
1409: 
1410: Daignere, A., A. Wahl, V. Hardy, and A. Maignan, 2001, Physica C 
1411: {\bf 349}, 189.
1412: 
1413: Degiorgi, L., B. Briceno, M.S. Fuhrer, A. Zettl,
1414: and P. Wachter, 1994, Nature {\bf 369}, 541.
1415: 
1416: 
1417: Duan, H.M., R.M. Yandrofski, T.S. Kaplan, B. Dlugosch, J.H. Wang, 
1418: and A.M. Hermann, 1991, Physica C {\bf 185-189}, 1283.
1419: 
1420: Fisk, Z., and G.W. Webb, 1976, Phys. Rev. Lett.  {\bf 36}, 1084. 
1421: 
1422: Fisk, Z., and A.C. Lawson, 1973, Solid State Commun. {\bf 13}, 277.
1423: 
1424: Forro, L., 2002 (priv. comm.).
1425: 
1426: Girvin, S.M., and M. Jonson, 1980, Phys. Rev. B {\bf 22}, 3583.
1427: 
1428: Greenwood, D.A., 1958, Proc. Phys. Soc. {\bf 71}, 585.
1429: 
1430: Grimvall, G., 1981, {\it The electron-phonon interaction 
1431: in metals}, North-Holland (Amsterdam) pp. 210-223.
1432: 
1433: Grimvall, G., 2001, private commun.
1434: 
1435: Grimvall, G., M. Thiessen, and A.F. Guillermet, 1987, Phys. Rev. B
1436: {\bf 36}, 7816.
1437: 
1438: Guillermet, A.F. and G. Grimvall, 1991, Phys. Rev. B {\bf 44}, 4332.
1439: 
1440: Gunnarsson, O., 1997, Rev. Mod. Phys. {\bf 69}, 575.
1441: 
1442: Gunnarsson, O., and J.E. Han, 2000, Nature  {\bf 405}, 1027.
1443: 
1444: Gurvitch, M., A.K. Ghosh, B.L. Gyorffy, H. Lutz, O.F. Kammerer,
1445: J.S. Rosner, and M. Strongin, 1978, Phys. Rev. Lett. {\bf 41}, 1616.
1446:  
1447: Gurvitch, M., and A.T. Fiory, 1987, Phys. Rev. Lett. {\bf 59}, 1337.
1448: 
1449: Han, J.E., E. Koch, and O. Gunnarsson, 2000, Phys. Rev. Lett. 
1450: {\bf 84}, 1276.
1451: 
1452: Han, J.E., O. Gunnarsson, and V.H. Crespi, 2003, Phys. Rev. Lett. 
1453: {\bf 90}, 167006.
1454: 
1455: Hidaka, Y., and M. Suzuki, 1989, Nature {\bf 338}, 635.
1456: 
1457: Hou, J.G.,  V.H. Crespi, X.-D. Xiang, W.A. Vareka, G. Briceno, 
1458: A. Zettl, and M.L. Cohen, 1993, Solid State Commun. {\bf 86}, 643.
1459: 
1460: Hou, J.G., L. Lu, V.H. Crespi, X.-D. Xiang, A. Zettl, and M.L. Cohen, 
1461: 1995, Solid State Commun. {\bf 93}, 973. 
1462: 
1463: Imry, Y., 1980, Phys. Rev. Lett. {\bf 44}, 469.
1464: 
1465: Ino, A., C. Kim, T. Mizokawa, Z.-X. Shen,
1466: A. Fujimori, M. Takaba, K. Tamasaku, H. Eisaki, and S. Uchida, 1999,
1467: J. Phys. Soc. Jpn {\bf 68}, 1496. 
1468: 
1469: Ioffe, A.F., and A.R. Regel, 1960, Prog. Semicond. {\bf 4}, 237.
1470: 
1471: Jarrell, M., and J.E. Gubernatis,  1996, Phys. Rep. {\bf 269}, 133.
1472: 
1473: Jarrell, M., and Th. Pruschke, 1994, Phys. Rev. B {\bf 49}, 1458.
1474: 
1475: Jonson, M., and S.M. Girvin, 1979, Phys. Rev. Lett. {\bf 43}, 1447.
1476: 
1477: Hebard, A.F., T.T.M. Palstra, R.C. Haddon, and R.M. Fleming, 1993,
1478: Phys. Rev. B {\bf 48}, 9945.
1479: 
1480: Khalifah, P., K.D. Nelson, R. Jin, Z.Q. Mao, Y. Liu, Q. Huang,
1481: X.P.A. Gao, A.P. Ramirez, and R.J. Cava, 2001, Nature {\bf 411}, 669.
1482: 
1483: Kitazawa, K., T. Matsuura, S. Tanaka, 1981, in {\it Ternary 
1484: Superconductors}, edited by G.K. Shenoy, B.D. Dunlap, and F.Y. Fradin,
1485: Elsevier, (North-Holland, New York), p. 83.
1486: 
1487: Klein, L., L. Antognazza, T.H. Geballe, M.R. Beasley, and 
1488: A. Kapitulnik, 1999a, Phys. Rev. B {\bf 60}, 1448. 
1489: 
1490: Klein, L., L. Antognazza, T.H. Geballe, M.R. Beasley, and 
1491: A. Kapitulnik, 1999b, Physica B {\bf 259-261}, 431.
1492:  
1493: Kohn, W., and J.M. Luttinger, 1957, Phys. Rev. {\bf 108}, 590.
1494: 
1495: Korotin, M.A., V.I. Anisimov, D.I Khomskii, and G.A. Sawatzky, 1998,
1496: Phys. Rev. Lett. {\bf 80}, 4305.
1497: 
1498: Kubo, Y., Y. Shimakawa, T. Manako, T. Kondo, and H. Igarashi, 1991,
1499: Physica C {\bf 185-189}, 1253.
1500: 
1501: Lange, E., and G. Kotliar, 1999, Phys. Rev. B {\bf 59}, 1800.
1502: 
1503: Laughlin, R.B., 1982, Phys. Rev. B {\bf 26}, 3479.
1504: 
1505: Lee, P.A., and T.V. Ramakrishnan, 1985, Rev. Mod.  Phys. {\bf 57}, 287.
1506: 
1507: Lewis, S.P., P.B. Allen, and T. Sasaki, 1997, Phys. Rev. B {\bf 55}, 10253.
1508: 
1509: L\"ohle, J., J. Karpinski, A. Morawski, and P. Wachter, 1996, Physica C
1510: {\bf 266}, 104.
1511: 
1512: Mahan, G.D., 1990, {\it Many-Particle Physics}, Plenum (New York), p. 651.
1513: 
1514: Mandrus, D., L. Forro, C. Kendziora, and L. Mihaly, 1992,
1515: Phys. Rev. B {\bf 45}, 12640.
1516: 
1517: 
1518: Marchenko, V.A., 1973, Sov. Phys.-Solid Sate {\bf 15}, 1261.
1519: 
1520: Maldague, P.F., 1977, Phys. Rev. B {\bf 16}, 2437.
1521:  
1522: Martin, R., K.R. Mountfield, and L. Corruccini, 1978, J. de Phys. 
1523: (Paris) {\bf 39}, C6-371 (1978).
1524:  
1525: Martin, S., A.T. Fiory, R.M. Fleming, L.F. Schneemeyer, and J.V. 
1526: Waszczak, 1990, Phys. Rev. B {\bf 41}, 846.
1527: 
1528: Mazin, I.I., D.J. Singh, and C. Ambrosch-Draxl, 1999, J. Appl. Phys.
1529: {\bf 85}, 6220.
1530: 
1531: Merino, J., and R.H. McKenzie, 2000, Phys. Rev. B {\bf 61}, 7996.
1532: 
1533: Millis, A.J., J. Hu, and S. Das Sarma, 1999, Phys. Rev.
1534: Lett. {\bf 82}, 2354.
1535: 
1536: Mooij, J.H., 1973 phys. stat sol. (a) {\bf 17}, 521.
1537: 
1538: Mott, N.F., 1974, {\it Metal-insulator transitions}, Taylor $\&$ Francis 
1539: (London).
1540: 
1541: Morton, N., B.W. James, and G.H. Wostenholm, 1978, Cryogenics {\bf 18}, 131.
1542: 
1543: Orenstein, J., G.A. Thomas, A.J. Millis, S.L. Cooper,
1544: D.H. Rapkine, T. Timusk, L.F. Schneemeyer, and J.V. Waszczak, 1990,
1545: Phys. Rev. B {\bf 42}, 6342.
1546: 
1547: Parcollet, O., and A. Georges, 1999, Phys. Rev. B {\bf 59}, 5341.
1548: 
1549: Palstra, T.T.M., A.F. Hebard, R.C. Haddon, and P.B. Littlewood, 1994,
1550: Phys. Rev. B {\bf 50}, 3462.
1551: 
1552: Pickett, W.E., K.M. Ho, and M.L. Cohen, 1979, Phys. Rev. B {\bf 19}, 1734.
1553: 
1554: Rodbell, D.S., J.M. Lommel, and R.C. DeVries, 1966, J. Phys. Soc. Japan
1555: {\bf 21}, 2430.
1556: 
1557: Ron, A., B. Shapiro, and M. Weger, 1986, Phil. Mag. B {\bf 54}, 553.
1558: 
1559: Ruan, K.Q., Q. Cao, S.Y. Li, G.G. Qian, C.Y. Wang, X.H. Chen, L.Z. Cao,
1560: 2001, Physica C {\bf 351}, 402.
1561:  
1562: Salamon, M.B. and M. Jaime, 2001, Rev. Mod. Phys, {\bf 73}, 583. 
1563: 
1564: Singley, E.J., C.P. Weber, D.N. Basov, A. Barry, and J.M.D. Coey, 1999,
1565: Phys. Rev. {\bf 60}, 4126.
1566: 
1567: Stephens, P.W., L. Mihaly, P.L. Lee, R.L. Whetten, S.-M. Huang, 
1568: R. Kaner, F. Deiderichs, and K. Holczer, 1991, Nature {\bf 351}, 632.
1569: 
1570: 
1571: Sunandana, C.S., 1979, J. Phys. C: Solid State Phys. {\bf 12}, L165.
1572:  
1573: Sundqvist, B., and B.M. Andersson, 1990, Solid State Commun. {\bf 76}, 1019.
1574: 
1575: Takagi, H., B. Batlogg, H.L. Kao, J. Kwo, R.J. Cava, J.J. Krajewski, 
1576: and W.F. Peck, Jr., 1992, Phys. Rev. Lett. {\bf 69}, 2975.
1577: 
1578: Takenaka, K, R. Shiozaki, S. Okuyama, J. Nohara, A. Osuka, 
1579: Y. Takayanagi, and S. Sugai, 2002, Phys. Rev. B {\bf 65}, 092405.
1580: 
1581: Tsuei, C.C., 1986, Phys. Rev. Lett. {\bf 57}, 1943.
1582:  
1583: Tyler, A.W., A.P. Mackenzie, S. NishiZaki, and Y. Maeno, 1998,
1584: Phys. Rev. B {\bf 58}, R10107.
1585: 
1586: Uchida, S., T. Ido, H. Takagi, T. Arima, Y. Tokura, and S. Tajima, 1991,
1587: Phys. Rev. B {\bf 43}, 7942.
1588: 
1589: Vareka, W.A., and A. Zettl, 1994, Phys. Rev. Lett. {\bf 72}, 4121.
1590: 
1591: Wang, N.L., C. Geibel, and F. Steglich, 1996a, Physica C {\bf 262},  231.
1592: 
1593: Wang, N.L., B. Buschinger, C. Geibel, and F. Steglich, 1996b, 
1594: Phys. Rev. B {\bf 54}, 7445.
1595: 
1596: Wang, N.L., P. Zheng, T. Feng, C.D. Gu, C.C. Homes, J.M. Tranquada, 
1597: B.D. Gaulin, and T. Timusk, 2002, cond-mat/0211224.
1598: 
1599: Weger, M., and N.F. Mott, 1985, J. Phys. C: Solid State Phys.
1600: {\bf 18}, L201.
1601: 
1602: Weger, M., 1985, Phil Mag. B {\bf 52}, 701.
1603: 
1604: Wiesmann, H., M. Gurvitch, H. Lutz, A. Ghosh, B. Schwartz, M. Strongin, 
1605: P.B. Allen, and J.W. Halley, 1977, Phys. Rev. Lett. {\bf 38}, 782.
1606: 
1607: 
1608: Woodard, D.W., and G.D. Cody, 1964, Phys. Rev. {\bf 1964}, A166.
1609: 
1610: Zhang, F.C., and T.M. Rice, 1988, Phys. Rev. B {\bf 37}, 3759.
1611: 
1612: Yamada, K.,  C.H. Lee, K. Kurahashi, J. Wada, S. Wakimoto, S. Ueki,
1613: H. Kimura, Y. Endoh, S. Hosoya, G. Shirane, R.J. Birgeneau,
1614: M. Greven, M.A. Kastner, and Y.J. Kim, 1998, Phys.  Rev. B {\bf 57}, 6165.
1615: 
1616: 
1617: Yan, S.L., Y.Y. Xie, J.Z. Wu, T. Aytug, A.A. Gapud, B.W. Kang,
1618: L. Fang, M. He, S.C. Tidrow, K.W. Kirchner, J.R. Liu, and W.K. Chu,
1619: 1998, Appl. Phys. Lett. {\bf 73}, 2989. 
1620: 
1621: 
1622: Yoshida, T., X.J. Zhou, M. Nakamura, S.A. Kellar, P.V. Bogdanov, 
1623: E.D. Lu, A. Lanzara, Z. Hussain, A. Ino, T. Mizokawa, A. Fujimori, 
1624: H. Eisaki, C. Kim, Z.-X. Shen, T. Kakeshita, and S. Uchida, 2001, 
1625: Phys. Rev. B {\bf 63}, 220501(R).
1626: 
1627: \end{multicols}
1628: \end{document}
1629: 
1630: