0809.2202/mw7.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %                                                                    %
3: %          beta''-(ET)2SF5CH2CF2SO3                               %
4: %                     dc and mw                                   %
5: %
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %-----------------------------------------------------
8: %
9: %  We would like to resubmit the manuscript
10: %  for publication in European Physical Journal B.
11: %
12: %------------------------------------------------------
13: %   Corresponding author:
14: %
15: %   Martin Dressel
16: %   1.Physikalisches Institut
17: %   Universitat Stuttgart
18: %   Pfaffenwaldring 57
19: %   D-70550 Stuttgart
20: %   Germany
21: %
22: %   Phone +49-711-685-64946
23: %   FAX   +49-711-685-64886
24: %   e-mail dressel@pi1.physik.uni-stuttgart.de
25: %
26: %----------------------------------------------------
27: \def\cm{cm$^{-1}$}
28: \def\CHCF{$\beta^{\prime\prime}$-(BEDT\--TTF)$_2$\-SF$_5$\-CH$_2$\-CF$_2$\-SO$_3$}
29: \def\d8CHCF{$\beta^{\prime\prime}$-(d$_8$-BEDT\--TTF)$_2$\-SF$_5$\-CH$_2$CF$_2$\-SO$_3$}
30: \def\CHF{$\beta^{\prime\prime}$-(BEDT\--TTF)$_2$\-SF$_5$\-CHF\-SO$_3$}
31: \documentclass[epj]{svjour}
32: \usepackage{graphics}
33: \begin{document}
34: 
35: \title{DC and high-frequency conductivity of the organic metals
36: $\beta^{\prime\prime}$-(BEDT-TTF)$_2$SF$_5R$\,SO$_3$ ($R$ = CH$_2$CF$_2$ and CHF)
37: %\CHCF\ and \CHF
38: }
39: \author{M. Glied\inst{1}
40: \and S. Yasin\inst{1} \and S. Kaiser\inst{1} \and N.
41: Drichko\inst{1} \and M. Dressel\inst{1}
42: \thanks{email: dressel@pi1.physik.uni-stuttgart.de}
43: \and J. Wosnitza\inst{2}
44: \and J.A.  Schlueter\inst{3} \and G.L. Gard\inst{4} }
45: 
46: \institute{1. Physikalisches Institut, Universit\"at Stuttgart,
47: Pfaffenwaldring 57, D-70550 Stuttgart, Germany \and
48: Dresden High-Magnetic Field Laboratory, Forschungszentrum Dresden-Rossendorf, D-01314 Dresden, Germany \and
49: Material Science Division, Argonne National Laboratory, Argonne,
50: Illinois 60439-4831, U.S.A.
51: \and
52: Department of Chemistry, Portland
53: State University, Portland, Oregon 97207-0751, U.S.A.}
54: 
55: 
56: 
57: \date{Received: \today}
58: 
59: \abstract{The temperature dependences of the electric-transport
60: properties of the two-dimensional organic conductors  \CHCF, \d8CHCF,
61: and \CHF\ are measured by dc methods in and perpendicular to the
62: highly-conducting plane. Microwave measurements are performed at
63: 24 and 33.5~GHz to probe the high-frequency behavior
64: from room temperature down to 2~K.
65: Superconductivity is observed in \CHCF\ and its deuterated
66: analogue. Although all the compounds remain metallic down to
67: low-temperatures, they are close to a charge-order transition.
68: This leads to deviations from a simple Drude behavior of the
69: optical conductivity which
70: become obvious already in the microwave range. In \CHCF, for
71: instance, charge  fluctuations cause an increase in microwave
72: resistivity for $T<20$~K which is not detected in dc measurements.
73: \CHF\ exhibits a simple metallic behavior at all
74: frequencies. In the dc transport, however, we observe
75: indications of localization  in the perpendicular direction.
76: %
77: \PACS{
78:       {71.10.Hf} {Non-Fermi-liquid ground states} \and
79:       {71.30.+h} {electron phase diagrams and phase transitions in model systems, Metal-insulator transitions and other electronic
80:       transitions}
81: \and
82:       {74.70.Kn} {Organic superconductors}
83:       }
84: }
85: %
86: \titlerunning{High-frequency conductivity of the organic metals}
87: 
88: \maketitle
89:   %
90: 
91: \section{Introduction}
92: Superconductivity has been discovered in organic crystals almost
93: three decades ago; nevertheless they are still considered as novel
94: materials with exotic properties because many questions remain
95: unanswered in spite of the enormous progress achieved over these
96: years \cite{Ishiguro98}. It became clear that the normal-state
97: properties are distinct from those of conventional metals and
98: severely influence the superconducting ground state. Several issues of
99: crucial importance for the understanding of low-dimensional
100: electron systems have been identified that are currently
101: investigated in few distinct families of organic conductors,
102: which serve as prototypes \cite{ChemicalReviews}.
103: 
104: The $\kappa$-phase BEDT-TTF salts (where BEDT-TTF or ET stands for
105: bis-(ethyl\-ene\-di\-thio)\-te\-tra\-thia\-ful\-va\-lene), for
106: instance, are two-dimensional metals with a half-filled conduction
107: band and a superconducting transition temperature $T_c$ of almost
108: 13~K. The system is close to a metal-insulator transition driven
109: by electronic correlations $U$. At room temperature the transport
110: reflects more a semiconducting behavior, often described as a bad
111: metal, and only below 100~K coherent transport starts to emerge
112: with a narrow Drude-like contribution developing as the
113: temperature is reduced further
114: \cite{Kornelsen90,Faltermeier07,Dumm09}. By now there is a good
115: agreement between experiment and theory \cite{Merino08} that the
116: $\kappa$-salts represent a prime example of a band-width controlled Mott
117: metal-insulator transition in two dimensions, and the correlated carriers in
118: the narrow zero-frequency contribution of the optical conductivity
119: exhibit Fermi-liquid behavior.
120: \begin{figure}[b]
121: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{pattern2.eps}}
122: \caption{
123: \label{fig:pattern}
124: The BEDT-TTF molecules are arranged in different patterns for
125: the various phases.}
126: \end{figure}
127: 
128: The $\alpha$-phase BEDT-TTF compounds are quarter-filled systems,
129: which are subject to charge order depending on the ratio of
130: nearest-neighbor interaction $V$ to bandwidth $W$
131: \cite{Drichko07}: while $\alpha$-(BEDT-TTF)$_2$I$_3$ becomes
132: completely insulating below a sharp metal-insulator transition at
133: 135~K \cite{Bender84,Dressel92a},
134: $\alpha$-(BEDT-TTF)$_2$\-(NH$_4$)\-Hg(SCN)$_4$ is an ambient
135: pressure superconductor with $T_c\approx 1$~K \cite{Wang90}. If in
136: the anions (NH)$_4$ is substituted by Tl, Rb, or K,
137: superconductivity is suppressed but metallic properties are
138: conserved all the way down to lowest temperatures. Accordingly,
139: the optical conductivity exhibits a narrow Drude-like component
140: within a charge-order pseudogap
141: \cite{Dressel92b,Dressel03,Drichko06}. It was proposed that in
142: these systems superconductivity is mediated by charge-order
143: fluctuations \cite{Merino01}, and recent experiments
144: \cite{Drichko06,Merino06} provide strong evidence pointing in this
145: direction.
146: 
147: 
148: In order to clarify how general this proposal is, other
149: quarter-filled compounds have to be investigated that also show
150: charge order, like the $\theta$ or $\beta^{\prime\prime}$ phases
151: of the BEDT-TTF salts.
152: In
153: the case of the $\beta^{\prime\prime}$ phase, there is only one
154: kind of stack with two BEDT-TTF molecules per unit cell \cite{Mori98a} as depicted in Fig.~\ref{fig:pattern}. The
155: planar BEDT-TTF molecules are loosely stacked along the $a$ axis
156: with strong orbital overlaps between adjacent stacks ($b$
157: direction). The highly conducting $ab$ sheets are separated by the
158: anions with a much lower conductivity in $c$ direction.
159: 
160: Recently optical investigations on
161: $\beta^{\prime\prime}$-(BEDT-TTF)\-(TCNQ) have been published
162: \cite{Uruichi06} which indicate a charge-order transition around
163: 170~K. The interplay of charge order and superconductivity was
164: studied by magnetotransport experiments \cite{Bangura05} in a
165: series of compounds based on
166: $\beta^{\prime\prime}$-(BEDT-TTF)$_4$[(H$_3$O)$M$(C$_2$O$_4$)$_3$].
167: Here we will concentrate on the family
168: $\beta^{\prime\prime}$-(BEDT-TTF)$_2$\-SF$_5\-R\-$\,SO$_3$.
169: 
170: \CHCF\ is the first fully organic superconductor of the BEDT-TTF
171: family \cite{Geiser96}. If the anions are slightly modified by
172: replacing $R$ = CH$_2$CF$_2$ with CHF in
173: $\beta^{\prime\prime}$-(BEDT-TTF)$_2$SF$_5R$\,SO$_3$, the material
174: remains metallic down to low temperatures but does not
175: superconduct. Changing the anions to SF$_5$CHFCF$_2$SO$_3$ causes
176: a metal to insulator transition near 180~K, while crystals with
177: $R$ = CF$_2$ and CH$_2$ are insulators in the entire temperature
178: range \cite{Ward00,Jones00}. The question addressed here is the
179: nature of the charge transport, the effects of electronic
180: correlations, and the influence of charge fluctuations on
181: the electronic properties in the metallic
182: $\beta^{\prime\prime}$-(BEDT-TTF)$_2$SF$_5R$\,SO$_3$ salts. To
183: this end, we have performed dc and microwave measurements of
184: the in-plane resistivity as a function of temperature on the
185: organic superconductors \CHCF\ and deuterated analog \d8CHCF, as
186: well as the metallic compound \CHF.
187: 
188: \section{Experimental Details}
189: \label{sec:experimental} Single crystals of \CHCF, \d8CHCF, and
190: \CHF\ have been grown via electrochemical techniques in an H cell
191: at Argonne National Laboratory as described in
192: Ref.~\cite{Geiser96}. The crystals have typical dimensions of
193: $0.5\times 1.5 \times 0.25$~mm$^3$ and form plates with a large
194: face containing the conducting ($ab$) plane. We have performed
195: temperature-dependent measurements of the dc resistivity of these
196: crystals within the plane and perpendicular to it ($c$ direction)
197: using the standard four-point method with a typical current of
198: 50~$\mu$A. The contacts were made by evaporating 50~\AA\ thick
199: gold pads on the crystal, then 15~$\mu$m gold wires were pasted on
200: each pad with a small amount of carbon paint; in some cases the
201: leads were directly put onto the crystal. The samples were slowly
202: (0.2 - 0.5~K/min) cooled down to avoid cracks and ensure thermal
203: equilibrium. Temperature-dependent measurements were conducted in
204: a He exchange gas cryostat down to 2~K; in general data were
205: acquired upon cooling and warming.
206: 
207: It is a well-known problem that the dc transport measurements of
208: highly-anisotropic samples are hampered by contributions from
209: other directions. This is particularly important when the
210: resistivity is measured along the highly-conducting chains (or
211: layers), because there is no guarantee that the same chain (or
212: layer) is contacted by the four leads. Any interruption, stack
213: fault, kink, step, terraces, or other discontinuity causes
214: contributions of the interchain (or interlayer) resistivity. The
215: second problem common to organic but also inorganic crystals is
216: how to achieve low-resistance contacts which inject the current
217: homogeneously but do not cause cracks and non-linearities. For
218: that reason contactless microwave methods have been developed by
219: Schegolev {\it et al.} for investigating the electronic transport
220: in small and highly anisotropic crystals
221: \cite{Schegolev72,Helberg96}. It should be noted, that dc
222: measurements probe the voltage drop for a fixed electrical current
223: and thus are sensitive to any additional resistance added by
224: interchanging the conducting chains (or layers), for instance;
225: i.e.\ they measure the highest resistance for a given conductance
226: path. In contrast, microwave methods are sensitive to the current
227: induced by the electric field and thus probe the highest
228: conduction, weighted by the respective volume fraction, of course.
229: An interruption of a metallic sample by cracks, for instance, does
230: not influence the microwave properties since the high frequencies
231: easily bridge small gaps.
232: 
233: The microwave conductivity was obtained by the cavity-perturbation
234: method utilizing two different cylindrical copper cavities which
235: resonate in the TE$_{011}$ mode at 24 and 33.5~GHz, respectively.
236: They are fed by an Agilent E8257D analog signal generator via
237: suitable rectangular wave\-guides and operate in the transmission mode. The
238: coupling is about 10\%\ and done through two holes in the
239: sidewalls. The crystals are positioned in the maximum of the
240: electric field placed onto a quartz substrate (0.07~mm thick) in
241: such a way that the electric field is parallel to the $b$ axis.
242: The samples were cooled down slowly (0.2~K per minute)
243: from 300~K to 2~K by coupling to the liquid helium
244: bath with the help of low-pressure He exchange gas and by
245: utilizing a regulated heater. The stability is better than 10~mK
246: \cite{Dressel05}. By recording the center frequency $f$ and the
247: width $\Gamma$ (FWHM) of the resonance curve as a function of
248: temperature and comparing them to the corresponding parameters of
249: an empty cavity ($f_0$ and $\Gamma_0$), the complex electrodynamic
250: properties of the sample, like the surface impedance, the
251: conductivity and the dielectric constant, can be determined by
252: cavity-perturbation theory; further details on microwave
253: measurements and the data analysis are extensively discussed in
254: Ref.~\cite{Klein93,DresselGruner02}. Complementary ESR experiments have been performed down to $T=5$~K on \CHF\ using a Bruker X-band spectrometer.
255: 
256: \section{Results and Discussion}
257: \subsection{Organic Superconductor Close to Charge Ordering: \CHCF}
258: \label{sec:CHCF}
259: \begin{figure}
260: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{CH2CF2_dc.eps}}
261: \caption{
262: \label{fig:CH2CF2_dc}
263: Temperature-dependent dc resistivity of \CHCF\ parallel and perpendicular to the highly conducting plane, i.e.\ parallel to the $a$ axis and parallel to the $c$ axis. The superconducting transition is observed at $T_c=5.9$~K.}
264: \end{figure}
265: In Fig.~\ref{fig:CH2CF2_dc} the in-plane and out-of-plane dc
266: resistivity of \CHCF\ is plotted as a function of temperature. The
267: absolute value of $\rho_a=2.6~{\rm \Omega cm}$ at room temperature is comparable to
268: the one reported in literature \cite{Dong99,remark2}. Previous
269: experiments \cite{Beckmann98,Su99,Hagel03,Hagel07} found a
270: metallic temperature dependence of the $c$-axis resistivity very
271: similar to our data, except for a higher absolute value. The room-temperature anisotropy within one crystal is ${\rho_c}/{\rho_a}
272: \approx 140$. Analyzing the asymmetry factor of the Dysonian line
273: shape obtained in X-band ESR experiments, Wang {\it et al.}
274: \cite{Wang99} estimated an in-plane anisotropy $\sigma_{\rm
275: max}/\sigma_{\rm min}$ of 1.35, with the maximum conductivity in
276: $b$ direction and $\sigma_{\rm min}$ basically parallel to the
277: stacks \cite{remark1}. This behavior is confirmed by optical
278: measurements \cite{Dong99,Kaiser08} and expresses the fact that
279: the orbital overlap between the stacks is stronger than along the stacks.
280: The resistivity ratio $R(300~{\rm K})/R(6~{\rm K})$ is
281: approximately 50 and 150 for the $a$ axis and 150 for interlayer transport.
282: At low temperatures, $\rho(T)$ follows a linear
283: behavior. Hagel {\em et al.}
284: \cite{Hagel07} report a $T^2$ behavior below 10~K which becomes linear above; the temperature region of the quadratic dependence could be extended upon applying external pressure.
285: The critical temperature of the superconducting
286: transition $T_c=5.9$~K  is
287: determined by the resistivity drop in both directions; its
288: comparably high value \cite{Geiser96,Ward00} evidences the
289: excellent quality of the single crystals.  As shown in the insets,
290: the width of the superconducting transition is $\Delta T_c \approx
291: 1$~K.
292: 
293: It is interesting to note that for both directions the resistivity
294: $\rho(T)$ increases more or less linearly with temperature up to
295: approximately 125~K. It was theoretically predicted \cite{Merino06,Merino03} and for various $\alpha$-salts experimentally observed \cite{Drichko06} that the $T^2$ dependence of a regular metal changes to a linear temperature dependence of the scattering rate when charge-order fluctuations become important.
296: For $\rho_c(T)$ we may identify a second
297: change in slope around $T=200$~K. From optical experiments, we see first indications that the infrared-active vibrational features $\nu_{27}({\rm B_{1u}})$ split already
298: around 200~K \cite{Drichko08}; at $T=150$~K two Raman modes are clearly distinguished,
299: which evidences charge disproportionation \cite{Kaiser08}. Other indications of  charge fluctuations come from the electronic bands.
300: 
301: 
302: The microwave properties obtained at $f=24.0$ and 33.5 GHz along
303: the $b$ axis are presented in Fig.~\ref{fig:CH2CF2_24+33}; the
304: data are normalized to the room-temperature value because the
305: uncertainty in absolute values is large (and can exceed a factor
306: of 10) due to errors in the depolarization factor caused by the
307: irregular shape of the samples. Nevertheless, from our data we
308: roughly estimate $\rho(300~{\rm K}) \approx 30~{\rm m \Omega\, cm}$ which is about one order of magnitude less than the dc values.
309: Starting at ambient temperature, the resistivity drops continuously
310: down to $T=30$~K as expected for a metal. This change is larger for
311: lower frequencies. A similar tendency is in general observed in
312: unconventional metals and indicates a strong frequency dependence
313: even in the microwave range.
314: 
315: 
316: In both microwave measurements, at 24 and 33.5~GHz, the
317: resistivity exhibits a minimum around $T=20$ to 30~K and turns up
318: by approximately 20 to 30\%\ when cooled down further; this
319: feature is robust. We relate the resistivity upturn to
320: charge-order fluctuations which may be even responsible for
321: superconductivity. From optical experiments (infrared and Raman
322: vibrational spectroscopy) it is known that \CHCF\ is subject to
323: charge ordering around 125-150~K \cite{Kaiser08}. From that we
324: would expect that a pseudogap opens which causes a reduction of
325: the density of states and hence a lower conductivity. However, the
326: coherent-particle zero-frequency transport is not affected; and
327: the low-frequency optical measurements also give clear evidence
328: for a Drude-like peak. Alternatively, the increased microwave
329: resistivity could be caused by enhanced scattering on charge
330: fluctuations. ESR experiments do not indicate any magnetic order
331: prior to the superconducting transition \cite{Wang99}.
332: 
333: The superconducting transition is observed at $T_c=5.0$~K for the
334: 24~GHz experiment and around 4.5~K in the 33.5~GHz data. The lower
335: value compared to the dc measurement can be explained by the weak
336: thermal coupling of the sample in the microwave cavity.
337: \begin{figure}
338: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{CH2CF2_24+33.eps}}
339: \caption{ \label{fig:CH2CF2_24+33} Temperature-dependent in-plane
340: resisitivity of \CHCF\ measured at 24 and 33.5~GHz normalized to
341: their room-temperature value. The inset shows the low-temperature
342: resistivity. For these specimen the resistivity increases below
343: approximately 30~K, before the superconducting transition is
344: observed.}
345: \end{figure}
346: 
347: Fermi-surface studies of \CHCF\ by Shubnikov-de Hass oscillations
348: see no indications of one-dimensional sheets;
349: they further give evidence that the two-dimensional part of the Fermi surface
350: contains only 5\%\ of the first Brillouin zone \cite{Beckmann98}.
351: This implies that the charge carriers available for dc transport
352: are significantly reduced, leading to a small spectral weight of
353: the Drude-like optical conductivity. From our high-frequency
354: measurements we can now conclude that already at 20 to 30~GHz the
355: roll-off in the frequency-dependent conductivity
356: is taken effect. The non-monotonous temperature behavior
357: infers charge fluctuations being responsible for this
358: behavior.
359: 
360: \subsection{Organic Superconductor: \d8CHCF}
361: \label{sec:d8CHCF} The dc-transport properties of the deuterated
362: organic superconductor \d8CHCF\ are displayed in
363: Fig.~\ref{fig:d8CH2CF2_dc}; the current is either directed within
364: the  planes or perpendicular to them, as indicated. For the
365: highly-conducting $b$ direction a simple metallic response is
366: observed with a resistivity ratio $R(300~{\rm K})/R(6~{\rm K}) =
367: 100$. The exponent $\alpha$ in the power law $\rho(T)\propto
368: T^{\alpha}$ is slightly below 2. Measurements of magnetic quantum
369: oscillations evidenced virtually no difference in the
370: low-temperature band-structure parameters compared to the
371: hydrogenated sister compound. At $T_c=5.6$~K the onset of the
372: superconducting transition is observed which is about 1~K broad; a
373: somewhat lower critical temperature was reported previously based
374: on ac-susceptibility measurements \cite{Schlueter01}. In the $c$
375: direction the resistivity is very large indicating the pronounced
376: two-dimensionality of the material. Cooling down from ambient
377: temperature the resistivity increases first, goes through a
378: maximum around $T=200$~K and then continuous to drop down to
379: lowest temperatures in a linear fashion, before it exhibits
380: superconductivity \cite{remark3}.
381: \begin{figure}
382: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{d8CH2CF2_dc.eps}}
383: \caption{ \label{fig:d8CH2CF2_dc} Temperature-dependent
384: dc-resistivity of \d8CHCF\ parallel and perpendicular to the
385: highly conducting plane, i.e.\ parallel to the $b$ axis and
386: parallel to the $c$ axis. The superconducting transition occurs at
387: $T_c=5.6$~K.}
388: \end{figure}
389: 
390: Fig.~\ref{fig:d8CH2CF2_24+33} shows the microwave resistivity of
391: \d8CHCF\ measured along the highly conducting $b$ direction at
392: $24$ and 33.5 GHz. At room temperature the absolute values are
393: approximately $0.1~\rm \Omega$cm; i.e.\ they are comparable to the
394: dc resistivity. While the resistivity ratio is similar in both
395: experiments [$\rho(300~{\rm K})/\rho(6~{\rm K})$ $\approx 10$], at
396: $f=24$~GHz we observe a dramatic change at temperatures above
397: 200~K while the higher frequency data drop gradually in a more or
398: less quadratic temperature behavior. This confirms the general
399: behavior already mentioned for \CHCF, that chang\-es in
400: resistivity with temperature become smeared out as the measurement
401: frequency increases. The transition to the superconducting phase
402: is detected at 5.6~K. The microwave resistivity flattens out below
403: 50~K and exhibits only a very weak temperature dependence. But
404: most important, there are no indications of a resistivity upturn
405: as observed in the hydrogenated analogue (cf.\
406: Fig.~\ref{fig:CH2CF2_24+33}). With other words, since the
407: deuterated crystals remain metallic even in their microwave
408: properties, they seem to be less susceptible to charge
409: fluctuations.
410: \begin{figure}
411: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{d8CH2CF2_24+33.eps}}
412: \caption{ \label{fig:d8CH2CF2_24+33} Temperature-dependent
413: resistivity of \d8CHCF\ measured at 24 and 33.5~GHz. The critical
414: temperature for the superconducting phase is $T_c=5.6$~K. The
415: solid line indicates a fit to a quadratic behavior. The inset
416: exhibits a low-temperature blow-up of the 24-GHz data.}
417: \end{figure}
418: 
419: A careful investigation \cite{Schlueter01} of the isotope shift in
420: a large number of BEDT-TTF samples revealed that for \CHCF\ the
421: deuteration causes an increase of the superconducting transition
422: temperature by about 0.3~K. From pressure-dependent measurements
423: \cite{Hagel03,Sadewasser97} we know that the superconducting
424: transition temperature decrease with pressure by a rate of -1.3
425: K/bar. This is due to the weaker influence of the effective
426: intersite Coulomb repulsion $V/t$ as $t$ increases with pressure;
427: the systems becomes more metallic \cite{remark5}. The increase of
428: $T_c$ would therefore infer that \d8CHCF\ is closer to the
429: charge-order transition. However, due to the known
430: sample-to-sample variation, we cannot make any final statement in
431: this regard and have to leave this discrepancy unresolved.
432: 
433: 
434: 
435: \subsection{Organic Metal: \CHF}
436: \label{sec:CHF}
437: Replacing the (SF$_5$CH$_2$CF$_2$SO$_3)^-$ anions in the organic crystals by (SF$_5$CHFSO$_3)^-$ leads to a non-superconducting compound.
438: \begin{figure}
439: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{CHF_dc1.eps}}
440: \caption{
441: \label{fig:CHF_dc}
442: Temperature-dependent dc-resistivity of \CHF\ parallel and perpendicular to the highly conducting plane; i.e.\ the along $b$ and $c$ axes, respectively. For most samples, a metallic behavior is observed down to low temperatures. Below approximately 50~K, indications of a metal-to-insulator transition are seen along the $c$ direction.}
443: \end{figure}
444: From room temperature down to approximately 50~K, the electrical
445: resistivity of \CHF\ exhibits a metallic temperature dependence in
446: both directions, parallel and perpendicular to the highly
447: conducting ($ab$) plane. The temperature dependences are presented
448: in Fig~\ref{fig:CHF_dc}. Along the $b$ axis, the absolute value
449: $\rho_b(300~{\rm K}) = 2.3~{\rm \Omega cm}$ is comparable to the
450: resistivity measured for \CHCF; the ratio $\rho_b(300~{\rm
451: K})/\rho_b(30~{\rm K}) = 15.3$. Below 50~K the resistivity seems
452: to saturate. A rise of the low-temperature resistivity was
453: reported by Ward {\it et al.} \cite{Ward00,Jones00} without
454: providing an explanation. They also mention that the spin
455: susceptibility slightly increases below 12~K which indicates that
456: electron localization takes place at very low temperatures. We
457: cannot confirm this finding from X-band ESR measurements on our
458: samples, as demonstrated in Fig.~\ref{fig:CHF_esr}. In the
459: perpendicular direction, $\rho_c(T)$ also shows a decrease in
460: resistivity when cooled down, as plotted in Fig~\ref{fig:CHF_dc}b.
461: The room-temperature anisotropy $\rho_c/\rho_b$ of \CHF\ is
462: approximately 200, i.e.\ somewhat higher than for \CHCF. For
463: $T<50$~K an insulating behavior is observed with a rapid increase
464: in resistivity well above the room-temperature value; no simple
465: thermally activated behavior can be determined; nevertheless the
466: initial slope corresponds to an activation energy $E_a=25$~meV,
467: when fitted by $\rho_c(T)\propto \exp\{E_a/k_B T\}$.
468: \begin{figure}
469: \centering\resizebox{0.45\textwidth}{!}{\includegraphics*{CHF_esr.eps}}
470: \caption{ \label{fig:CHF_esr} ESR data obtained from
471: temperature-dependent X-band measurements on two different \CHF\
472: samples. The panels show the linewidth $\Delta H$, the $g$-value
473: and the spin susceptibility $\chi_s(T)$ for Sample 1 (left column)
474: and Sample 2 (right column).}
475: \end{figure}
476: 
477: \begin{figure}
478: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{CHF_dc2.eps}}
479: \caption{ \label{fig:CHF_dc2} Some of the \CHF\ samples exhibit an
480: insulating behavior in $\rho(T)$. The resistivity increases to a
481: maximum around 70~K, below which it decreases with a kink at
482: 16~K.}
483: \end{figure}
484: It should be noted that for some of the samples (two out of eight)
485: a different temperature behavior of the in-plane resistivity was
486: measured which resembles previous reports \cite{Ward00}. Starting
487: from approximately the same am\-bi\-ent-temperature value,
488: $\rho_b(T)$ decreases slightly when the sample is cooled down to
489: 250~K. Then the electrical resistivity becomes semiconducting and
490: gradually increases by about a factor of 4 all the way down to
491: 70~K where a broad maximum if found (Fig~\ref{fig:CHF_dc2}). While
492: Ward {\em et al.} \cite{Ward00} could identify an activated
493: behavior between $T=100$ and 300~K with $E_a=56$~meV, the increase
494: we observe does not correspond to a single activation energy; only
495: between 150 and 200~K the behavior may be described by
496: $\rho_b(T)\propto \exp\{E_a/k_B T\}$ with $E_a=26$~meV. The drop
497: in $\rho_b(T)$ observed at lower temperatures exhibits a kink
498: around 16~K which might be an indication of a phase transition. At
499: this point we do not have a definite explanation, but the absence
500: of any change in the ESR data around this temperature
501: (Fig.~\ref{fig:CHF_esr}) is an argument against magnetic order
502: \cite{remark4}.
503: 
504: Microwave measurements of \CHF\ in the highly-conducting plane
505: reveal a metallic behavior down to lowest temperatures
506: (Fig.~\ref{fig:CHF_24+33}). No indication of charge order could be
507: detected. Again, for $f=33.5$~GHz the temperature dependence of
508: the resistivity changes more gradually (almost linearly), while
509: for 24~GHz we observe a fast drop below room temperature with only
510: little variation for $T<150$~K. At low temperature we did not find
511: an upturn in resistivity as present in the dc data (cf.\
512: Fig.~\ref{fig:CHF_dc}). This basically implies that the
513: semiconducting behavior  shows up only for $\omega\rightarrow 0$
514: while in ac-transport and optical experiments the charge
515: localization can be overcome.
516: 
517: It should be pointed out that for the superconducting crystals \CHCF\ the tendency was opposite: there charge fluctuations led to a rise in high-frequency resistivity upon cooling below $T=20$~K prior to the superconducting transition (see Sec.~\ref{sec:CHCF} and in particular Fig.~\ref{fig:CH2CF2_24+33}) while the dc resistivity remains purely metallic. In the present case of \CHF, we feel that the strongly insulating behavior observed in the perpendicular direction (Fig.~\ref{fig:CHF_dc}) also affects $\rho_b$ because due to unavoidable crystal imperfections (e.g.\ cracks) the current path might have to change planes. This is not the case for  microwave measurements which are not sensitive to these imperfections and probe the electrical conduction more directly as discussed in Sec.~\ref{sec:experimental}.
518: \begin{figure}
519: \centering\resizebox{0.4\textwidth}{!}{\includegraphics*{CHF_24+33.eps}}
520: \caption{
521: \label{fig:CHF_24+33}
522: Temperature-dependent resisitivity of \CHF\ measured at 24 and 33.5~GHz. The data are normalized to the respective room-temperature value.}
523: \end{figure}
524: 
525: \section{Conclusions}
526: A comprehensive investigation of transport properties of different
527: $\beta^{\prime\prime}$-phase organic conductors and
528: superconductors has been presented, utilizing dc and microwave
529: methods at different frequencies (24 and 33.5~GHz) from room
530: temperature down to 2~K. At ambient conditions the in-plane dc
531: conductivity is of the order $1~({\rm \Omega cm})^{-1}$; the
532: an\-iso\-tropy $\sigma_{\parallel}/\sigma_{\perp}$ exceeds $10^2$.
533: Hence, the systems are highly anisotropic two-dimensional metals.
534: All compounds exhibit a more or less metallic behavior upon
535: cooling with \CHCF\ and the deuterated analog \d8CHCF\ becoming superconducting around
536: 5~K. For \CHCF\ clear signatures of charge fluctuations are
537: present for $T<30$~K. While they do not influence the dc
538: properties,  they lead to an increase of the high-frequency
539: resistivity. These effects are not seen in the deuterated
540: analog, maybe due to the better metallic behavior.
541: 
542: Considering the frequency dependence of the conductivity as
543: derived from our experiments, the general behavior certainly
544: deviates from that of simple metals. We observe a more gradual
545: temperature dependence for higher frequencies, indicating a
546: significant frequency dependence of the electrodynamic response
547: even in the microwave range. Since this happens at elevated
548: temperatures ($T>100$~K), we can rule out effects due to
549: conventional phonon scattering. Instead we are inclined to trace
550: it back to the peculiar bandstructure and correlation effects. The
551: difference in frequency dependence is strongest in \CHF\ and
552: weakest in \CHCF. Our investigations provide evidence that the
553: vicinity to charge order and the presence of charge
554: fluctuations strongly alter the dynamical response of the
555: electronic system.
556: 
557: \section*{Acknowlegements}
558: The dc measurements were done with the support of E.\ Rose. We would
559: like to thank G. Untereiner for the careful sample preparation.
560: The work at Stuttgart was supported by the Deutsche
561: Forschungsgemeinschaft (DFG). The crystal growth at Argonne
562: National Laboratory was performed under the auspices of the Office
563: of Basic Energy Sciences, Division of Material Sciences of the
564: U.S. Department of Energy, Contract No.\ DE-AC02-06CH11357.
565: 
566: 
567: \begin{thebibliography}{99}
568: \bibitem{Ishiguro98}
569: T. Ishiguro, K. Yamaji, and G. Saito, {\em Organic Superconductors}, 2nd edition, Vol.~88 of {\em  Springer Series in Solid-State Sciences} (Springer-Verlag, Berlin, 1998).
570: \bibitem{ChemicalReviews}See special issue of Chemical Reviews {\bf 104}, No. 11 (2004); and in particular M. Dressel and N. Drichko, Chem. Rev. {\bf 104}, 5689 (2004).
571: 
572: 
573: \bibitem{Kornelsen90}K. Kornelsen, J.E. Eldridge, H.H. Wang, and J.M. Williams, Solid State Commun. {\bf 76}, 1009 (1990);
574: J.~E. Eldridge, K. Kornelsen, H.~H. Wang, J.~M. Williams, A.~V.~D. Crouch, and D.~M. Watkins, Solid State Commun. {\bf 79},  583  (1991); K. Kornelsen, J.~E. Eldridge, H.~H. Wang, and J.~M. Williams, Phys. Rev. B {\bf 44}, 5235 (1992).
575: \bibitem{Faltermeier07}D. Faltermeier, J. Barz, M. Dumm, M. Dressel, N. Drichko, B. Petrov, V. Semkin,  R. Vlasova, C. Meziere, P. Batail, Phys. Rev. B {\bf 76}, 165113 (2007).
576: \bibitem{Dumm09}M. Dumm, D. Faltermeier, N. Drichko, M. Dressel, J. Merino, R. McKenzie, C. Meziere, and P. Batail, to be published.
577: \bibitem{Merino08}J. Merino, M. Dumm, N. Drichko, M. Dressel, and R. McKenzie, Phys. Rev. Lett. {\bf 100}, 086404
578: (2008).
579: 
580: 
581: \bibitem{Drichko07}
582: M. Dressel,  N. Drichko, and J. Merino,
583: Physica B {\bf 359-361}, 454 (2005);
584:  N. Drichko, M. Dumm, D. Faltermeier, M. Dressel, J.
585: Merino, and A. Greco, Physica C {\bf 460-462}, 125 (2007).
586: 
587: \bibitem{Bender84}
588: K. Bender, I, Hennig, D. Schweitzer, K. Dietz, H. Endres, and H.J. Keller, Mol. Cryst. Liq. Cryst. {\bf 108}, 359 (1984).
589: \bibitem{Dressel92a}
590: M. Dressel, G. Gr\"uner, J.P. Pouget, A. Breining, D. Schweitzer,
591: J. Phys. I (France) {\bf 4}, 579 (1994).
592: 
593: \bibitem{Wang90}H.H. Wang, K.D. Carlson, U. Geiser, W.K. Kwok, M.D. Vashon, J.E. Thompson, N.F. Larsen, G.D. McCabe, R.S. Hulscher, and J.M. Williams, Physica C {\bf 166}, 57 (1990).
594: \bibitem{Dressel92b}
595: M. Dressel, J.E. Eldridge, H.H. Wang, U. Geiser, and J.M. Williams,
596: Synth. Met. {\bf 52}, 201 (1992).
597: 
598: \bibitem{Dressel03}M. Dressel, N. Drichko, J. Schlueter, and J. Merino,
599: Phys. Rev. Lett. {\bf 90}, 167002 (2003).
600: \bibitem{Drichko06}N. Drichko, M. Dressel, C.A. Kuntscher, A. Pashkin, A. Greco, J. Merino, and J. Schlueter, Phys. Rev. B {\bf 74}, 235121 (2006).
601: \bibitem{Merino01} J. Merino and R.H. McKenzie, Phys. Rev. Lett. {\bf 87}, 237002 (2001).
602: \bibitem{Merino06}J. Merino, A. Greco, N. Drichko, M. Dressel,
603: Phys. Rev. Lett. {\bf 96}, 216402 (2006).
604: \bibitem{Mori98a}
605: T. Mori, Bull. Chem. Soc. Jpn. {\bf 71}, 2509 (1998); T. Mori, H.
606: Mori,  and S. Tanaka, Bull. Chem. Soc. Jpn. {\bf 72}, 179 (1999);
607: T. Mori, Bull. Chem. Soc. Jpn. {\bf 72}, 2011 (1999); T. Mori,
608: Bull. Chem. Soc. Jpn. {\bf 73}, 2243 (2000).
609: 
610: \bibitem{Uruichi06}
611: M. Uruichi, K. Yakushi, H.M. Yamamoto, and R. Kato, J. Phys. Soc.
612: Jpn. {\bf 75}, 074720 (2006).
613: \bibitem{Bangura05}
614: A.F. Bangura, A.I. Coldea, J. Singleton, A. Ardavan, A.
615: Akutsu-Sato, H. Akutsu, S.S. Turner, P. Day, T. Yamamoto, and K.
616: Yakushi, Phys. Rev. B {\bf 72}, 014543 (2005).
617: 
618: \bibitem{Geiser96}
619: U. Geiser, J.A. Schlueter, H.H. Wang, A.M. Kini, J.M. Williams, P.P. Sche, J.I. Zakowicz, M.L. VanZile, and J.D. Dudek, J. Am. Chem. Soc. {\bf 118}, 9996 (1996).
620: 
621: \bibitem{Ward00}
622: B.H. Ward, J.A. Schlueter, U. Geiser, H.H. Wang,
623: E. Morales, J.P. Parakka, S.Y. Thomas, J.M. Williams,
624: P.G. Nixon, R.W. Winter, G.L. Gard,
625: H.-J. Koo, M.-H. Whangbo,
626: Chem. Mater. {\bf 12}, 343 (2000).
627: \bibitem{Jones00}
628: B.R. Jones, I. Olejniczak, J. Dong, J.M. Pigos, Z.T. Zhu, A.D. Gerlach, J.L. Musfeldt, H.-J. Koo, M.-H. Whangbo, J.A. Schlueter, B.H. Ward, E. Morales, A.M. Kini, R.W. Winter, J. Mohtasham, and G.L. Gard,
629: Chem. Mater. {\bf 12}, 2490 (2000).
630: 
631: 
632: \bibitem{Schegolev72}
633: L.I. Buravov and I.F. Schegolev, Instr. Exp. Techn. (USSR) {\bf 14}, 528 (1971); I.F. Schegolev, phys. stat. sol. {\bf 12}, 9 (1972).
634: \bibitem{Helberg96}
635: H.W. Helberg and M. Dressel
636: J. Phys. I (France) {\bf 6}, 1683 (1996);
637: M. Dressel, O. Klein, S. Donovan, and G. Gr\"uner,
638: Ferroelectrics {\bf 176}, 285 (1996).
639: \bibitem{Dressel05}
640:  M. Dressel, K. Petukhov, B. Salameh, P. Zornoza, T.
641: Giamarchi, Phys. Rev. B {\bf 71}, 075104 (2005).
642: \bibitem{Klein93} O.~Klein, S.~Donovan, M.~Dressel, and G.~Gr\"{u}ner,
643: {\it Int.~J. Infrared and Millimeter Waves}, {\bf 14}, 2423
644: (1993); S.~Donovan, O.~Klein, M.~Dressel, K.~Holczer, and
645: G.~Gr\"{u}ner, {\it Int.~J. Infrared and Millimeter Waves}, {\bf
646: 14}, 2459 (1993); M.~Dressel, O.~Klein, S.~Donovan, and
647: G.~Gr\"{u}ner, {\it Int.~J. Infrared and Millimeter Waves}, {\bf
648: 14}, 2489 (1993).
649: \bibitem{DresselGruner02}
650: M. Dressel and G. Gr\"uner, {\it Electrodynamics of Solids}
651: (Cambridge University Press, Cambridge, 2002).
652: 
653: 
654: \bibitem{Dong99}
655: J.~Dong, J.L. Musfeldt, J.A. Schlueter, J.M. Williams, P.G. Nixon, R.W. Winter, and G.L. Gard,
656:  Phys. Rev. B {\bf 60}  4342 (1999).
657: \bibitem{remark2}Dong {\it et al.} \cite{Dong99} cite an absolute value of $6~{\rm \Omega cm}$; Geiser {\it et al.} \cite{Geiser96} mention that the resistivity increases a factor of 10 upon cooling to 100~K before it drops in a metallic fashion. There seems to be a considerable
658: sample dependence. In particular the in-plane transport strongly depends on crystal imperfections.
659: \bibitem{Beckmann98}
660: D. Beckmann, S. Wanka, J. Wosnitza, J.A. Schlueter, J.M. Williams,
661: P.G. Nixon, R.W. Winter, G.L. Gard, J. Ren, and M.H.  Whangbo,
662: Eur. Phys. J. B {\bf 1}, 295 (1998).
663: \bibitem{Su99}
664: X. Su, F. Zuo, J.A. Schlueter, J.M. Williams, P.G. Nixon, R.W. Winter, and G.L. Gard, Phys. Rev. B {\bf 59}, 4376 (1999).
665: \bibitem{Hagel03}J. Hagel, J. Wosnitza, C. Pfleiderer, J.A. Schlueter, J. Mohtasham, and G.L. Gard, Phys. Rev. B {\bf 68},
666: 104504 (2003).
667: \bibitem{Hagel07}J. Hagel, O. Ignatchik, J. Wosnitza, C. Pfleiderer, H. Davis, R. Winter,  and G.L. Grand, Physica C {\bf 460-462}, 639 (2007).
668: \bibitem{Wang99}
669: H.H. Wang, M.L. Van Zile, J.A. Schlueter, U. Geiser, A.M. Kini, P.P. Sche, H.-J. Koo, M.-H. Whangbo, P.G. Nixon, R.W. Winter, and G.L. Gard, J. Phys. Chem. {\bf 103}, 5493 (1999).
670: \bibitem{remark1}The crystal symmetry is triclinic, with the room-temperature unit-cell parameters $a = 9.260$~\AA, $b=11.625$~\AA, $c=17.572$~\AA, $\alpha = 94.69^{\circ}$, $\beta=91.70^{\circ}$, $\gamma=103.10^{\circ}$, $Z=2$, the space group is $P{\bar 1}$.
671: \bibitem{Kaiser08}
672: S. Kaiser, N. Drichko, M. Dressel, J.A. Schlueter, and A. Greco, to be
673: published.
674: \bibitem{Merino03}J. Merino, A. Greco, R.H. McKenzie, and M. Calandra, Phys. Rev. B {\bf 68}, 245121 (2003).
675: 
676: \bibitem{Drichko08}N. Drichko, S. Kaiser,
677: Y. Sun, C. Clauss, M. Dressel, H. Mori,
678: J. Schlueter, E.I. Zhyliaeva, S.A. Torunova, and R.N. Lyubovskaya, Physica B (in press)
679: 
680: \bibitem{Schlueter01}
681: J.A. Schlueter, A.M. Kini, B.H. Ward, U. Geiser, H.H. Wang, J. Mohtasham, R.W. Winter, and G.L. Gard, Physica C {\bf 351}, 261
682: (2001).
683: 
684: \bibitem{remark3}Since the sample exhibited cracks upon cooling, some residual resistance remains in the
685: superconducting phase.
686: \bibitem{Sadewasser97}S. Sadewasser, C. Looney, J.S. Schilling, J.A. Schlueter, J.M. Williams, P.G. Nixon, R.W. Winter, and G.L. Gard, Solid State Commun. {\bf 104}, 571 (1997).
687: \bibitem{remark5}The insulating state found at even higher pressure \cite{Hagel03} is probably caused by some structural change and not due to electronic correlations.
688: \bibitem{remark4}The linewidth of approximately 25~Oe is
689: comparable to the room temperature value reported by Wang {\em et
690: al.} \cite{Wang99} for \CHCF; also the temperature dependence is
691: comparable down to 10~K. The same holds for the $g$ factor and the
692: temperature dependence of the spin susceptibility.
693: \end{thebibliography}
694: \end{document}
695: