This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Charge imbalance in λ\lambda-(BETS)2GaCl4 and their interplay with superconductivity

Olga Iakutkina 1. Physikalisches Institut, Universität Stuttgart, 70569 Stuttgart, Germany    Ece Uykur 1. Physikalisches Institut, Universität Stuttgart, 70569 Stuttgart, Germany    Takuya Kobayashi Graduate School of Science and Engineering, Saitama University, Saitama, 338-8570, Japan    Atsushi Kawamoto Department of Physics, Graduate School of Science, Hokkaido University, Sapporo 060-0810, Japan    Martin Dressel 1. Physikalisches Institut, Universität Stuttgart, 70569 Stuttgart, Germany    Yohei Saito [email protected] 1. Physikalisches Institut, Universität Stuttgart, 70569 Stuttgart, Germany
Abstract

The two-dimensional organic superconductor λ\lambda-(BETS)2GaCl4 exhibits pronounced charge fluctuations below T150T\approx 150 K, in contrast to the sibling compound κ\kappa-(BETS)2GaCl4 that remains metallic down to milli-Kelvin. Infrared spectroscopy reveals only minor splitting in the vibrational features of the latter compound, common to other strongly dimerized κ\kappa-salts. When the organic molecules are arranged in the λ\lambda-pattern, however, a strong vibrational ν27(b1u)\nu_{27}(b_{1u}) mode is present, that forms a narrow doublet. Most important, when cooling λ\lambda-(BETS)2GaCl4 below 150 K, two weak side modes appear due to charge disproportionation that amounts to 2δ=0.14e2\delta=0.14e. In analogy to the β′′\beta^{\prime\prime}-type organic conductors, we propose that charge fluctuations play an important role in emerging of unconventional superconductivity in λ\lambda-(BETS)2GaCl4 at Tc=4.7T_{c}=4.7 K. We discuss the possibility of a charge-density-wave that coexists with the proposed spin-density-wave state.

.

I Introduction

Unconventional superconductivity in strongly correlated systems such as heavy fermionic systems [1], transition metal oxides [2, 3], organic charge-transfer salts [4, 5] is a major topic in condensed matter physics. Great interest is particularly drawn by the interplay between superconductivity and some ordered state, let it be magnetic or charge order. Concerning the pairing mechanism in cuprates, for instance, the discussion goes about the importance of spin fluctuations, originating from the antiferromagnetic Mott insulating state, and charge fluctuations, associated with dynamical stripes [6, 7, 8, 9].

There is a far-reaching similarity between two-dimensional organic conductors and cuprates [10], including the fact that the superconducting state is found at the border of metal and Mott insulator [11]. The molecular systems actually constitute textbook examples of Mott insulators because the charge-transfer band is located well below the Hubbard bands split by U0.5U\approx 0.5-1 eV. Commonly these salts crystallize in D2XD_{2}X stoichiometry, with DD the charge donor and XX the monovalent acceptor. This implies three-quarter-filled conduction bands as long as the stacks are homogeneous; the formation of dimers, however, results in half-filled bands. In the first case, inter-site Coulomb repulsion VV governs the relevant physics, giving rise to a charge-ordered insulating ground state; the latter case the genuine Mott system, completely characterized by the strength of UU with respect to the bandwidth WW.

For half-filled Mott insulators, spin fluctuations are rather common, possibly related to superconductivity. The κ\kappa-(BEDT-TTF)2X compounds are prime examples of antiferromagnetic Mott insulators located next to superconductivity, often with some coexistence regions [12, 13, 14, 15]. Charge-ordered materials, on the other hand, remain non-magnetic at all temperatures; hence any effect of spin fluctuations on the superconducting ground state next to the charge-ordered phase can be rules out. Merino and McKenzie suggested [16] that instead superconductivity is mediated by charge fluctuations here. Subsequently, the importance of charge degrees of freedom for these superconducting compounds was proven experimentally [17, *Kaiser12, 19, *PhysRevB.99.155144].

Refer to caption
Figure 1: (a) Looking along acac-planes, the two-dimensional charge-transfer salts form different dimer patterns. In the κ\kappa-phase the dimers are rotated with respect to each other, while the λ\lambda-pattern is organized in stacks with two dimers, i.e. four molecules per unit cell. (b) The BETS donor layers are separated by layers of GaCl4{}_{4}^{-} anions. The organic molecules are slightly tilted in alternating directions in case of κ\kappa-BETS, while uniform in case of λ\lambda-BETS. (c) Most common are the donor molecules ET = BEDT-TTF, i.e. bis(ethylenedithio)tetrathiafulvalene), but sulfur can be substituted by selenium, leading to STF = BEDT-STF, and BETS = BEDT-TSF, i.e. bis(ethylenedithio)tetraselenafulvalene). The more extended selenium orbitals cause a larger bandwidth favoring metallic conductivity.

Fig. 1(a) illustrates that in the κ\kappa-type salts the organic molecules are arranged in distinct pairs, while dimerization is less obvious for the λ\lambda-pattern. The interesting question now arises whether here antiferromagnetic order is present in those compounds, how important spin fluctuations are, and whether charge fluctuations or even charge order can be observed. In this context, the series λ\lambda-D2D_{2}GaCl4 is of particular interest, as the electronic bandwidth WW is varied by modifying the organic molecule DD = BEDT-TTF, BEDT-STF or BETS [Fig. 1(c)]. The strongest effect of correlations, characterized by U/WU/W, occurs in λ\lambda-(BEDT-TTF)2GaCl4: it resembles a Mott insulator with antiferromagnetic order at TN=13T_{N}=13 K [21, 22, 23]. The intermediate compound λ\lambda-(BEDT-STF)2GaCl4 does not exhibit magnetic order down to lowest temperatures but rather unconventional behavior assigned to a disordered quantum spin state [24]. λ\lambda-(BETS)2GaCl4, on the other side, remains metallic all the way down to the superconducting transition at Tc=4.7T_{c}=4.7 K. The picture is confirmed by pressure-dependent studies [21, 22].

The fact that superconductivity does not occur next to the antiferromagnetically ordered phase calls for further investigations. 13C-NMR measurements of the superconductor λ\lambda-(BETS)2GaCl4 revealed, that the behavior of 1/T1T1/T_{1}T at high temperature above 55 K is dominated by antiferromagnetic fluctuations similar to what is observed in the κ\kappa-type salts such as κ\kappa-(BEDT-TTF)2Cu[N(CN)2]Br. Below 30 K, 1/T1T1/T_{1}T becomes temperature independent and ρ(T)T2\rho(T)\propto T^{2}; evidencing a Fermi-liquid response. In addition, another increase in spin-lattice relaxation rate of λ\lambda-(BETS)2GaCl4 was observed below 10 K and interpreted as sign of Fermi surface nesting probably related to spin-density wave formation [25]. More light is shed on this issue by considering the series λ\lambda-(BETS)2GaBrxCl4-x, where the Br substitution acts as negative pressure, driving the system insulating with rising xx [26, 27]. 13C-NMR spectroscopy infers that spin degree of freedom plays an important role in the pairing mechanism of λ\lambda-type organic superconductors.

On the other hand, 77Se-NMR measurement of λ\lambda-(BETS)2FeCl4 and λ\lambda-(BETS)2GaCl4 showed linewidth broadening at low temperature that is assigned to charge disproportionation [28, 29], which can also contribute to the mechanism of superconductivity. In order to elucidate this point, comprehensive NMR investigations have been performed on λ\lambda-(BETS)2GaCl4 probing different isotopes 69Ga and 71Ga in order to separate contributions of spin and charge dynamics [30]. While there is no enhancement of charge fluctuations below T=150T=150 K, spin fluctuations become dominant at low temperature. Obviously NMR spectroscopy is more sensitive to spin than to charge, calling for a detailed investigation of charge distribution in λ\lambda-(BETS)2GaCl4, for instance by probing the electronic charge directly.

For that reason we elucidate the issue of charge disproportionation by infrared spectroscopy with the focus on the range of charge sensitive molecular vibrations of λ\lambda-(BETS)2GaCl4. We compare the findings in this paper with those on κ\kappa-(BETS)2GaCl4, which is a strongly dimerized compound that behaves metallic down to liquid-helium temperatures with no indications of strong correlations and magnetic contributions [26, 31, 32].

II Materials and Methods

Single crystals of the quasi-two-dimensional organic superconductor λ\lambda-(BETS)2GaCl4 (abbreviated as λ\lambda-BETS) and a metal κ\kappa-(BETS)2GaCl4 (κ\kappa-BETS hereafter) were grown by standard electrochemical methods [33, 34]. In these salts, BETS donor molecules are dimerized, forming a triangular lattice in the κ\kappa-type, and for the λ\lambda-type they are arranged on a square lattice, as illustrated in Fig. 1(a). κ\kappa-BETS crystalizes in orthorhombic Pnma space group. Layers of donor molecules tilted in opposite directions with respect to each other, leading to a doubling of the unit cell along the bb-axis. The unit cell contains four dimmers, and two BETS molecules in a dimer are connected by a center of symmetry [Fig. 1(b)]. λ\lambda-BETS has lower symmetry and belong to the triclinic P1¯\bar{1} space group. In contrast to κ\kappa-BETS, here the donor layers of BETS molecules are identical. As a result, it has two dimmers per unit cell as shown in Fig. 1(b). An interesting difference between both BETS-salts is that in λ\lambda-BETS two molecules forming a dimer are not identical, but have a slightly different length of C=C double bonds.

For both compounds the temperature dependence of the dc resistivity was measured along the most conducting axis using the conventional four-contacts method. The applied current was limited in the range of 100-200 μ\muA; the resistivity data were recorded while warming up from 3 to 295 K.

Optical reflectivity measurements were performed with the light polarized perpendicular to the highly-conducting BETS planes, cf. Fig. 1(b). In this direction we can detect the molecular vibration ν27(b1u)\nu_{27}(b_{1u}) which is the most sensitive local probe for the charge per molecule [35, 36, 37]. For this purpose, a Hyperion infrared microscope was attached to a Bruker Vertex 80v Fourier-transform infrared spectrometer. The data were collected in the spectral range 500-8000 cm1{\rm cm}^{-1} between T=295T=295 and 9 K. To perform the Kramers-Kronig transformation, a constant extrapolation was used for frequencies below 500 cm1{\rm cm}^{-1}  and the standard ω4\omega^{-4} decay was utilized as high-frequency extrapolation [38].

III Results

Refer to caption
Figure 2: (a) Double logarithmic plot of the relative resistivity ρ(T)\rho(T)/ρ(295K)\rho(295\,K) of λ\lambda-(BETS)2GaCl4 (blue dots) and κ\kappa-(BETS)2GaCl4 (red circles) measured as a function of temperature along the most conducting axes. The maximum in ρ(T)\rho(T) for λ\lambda-(BETS)2GaCl4 can be best determined from the zero-crossing of the derivative dρ\rho/dTT at TmaxT_{\rm max} plotted in panel (b) using arbitrary units. The superconducting transition occurs at Tc=4.7T_{c}=4.7 K as illustrated in panel (c).

In Fig. 2(a) the temperature dependence of the electrical resistivity ρ(T)\rho(T) of κ\kappa-(BETS)2GaCl4 and λ\lambda-(BETS)2GaCl4 is plotted, normalized to the room temperature value ρ(T=295K)\rho(T=295{\rm K}). The behavior of the two salts is rather different. κ\kappa-(BETS)2GaCl4 remains metallic in the entire temperature range: ρ(T)\rho(T) drops monotonously with minor changes in slope at elevated temperatures. For the λ\lambda-salt a pronounced maximum is observed at Tmax=91T_{\rm max}=91 K [Fig. 2(b)]; while below approximately 40 K, the ρ(T)T2\rho(T)\propto T^{2} dependence indicates a Fermi-liquid behavior. The compound becomes superconducting at Tc=4.7T_{c}=4.7 K (onset at 4.8 K) as shown in Fig. 2(c). These results are consistent with previous publications [34].

In order to explore possible charge distribution in organic charge-transfer salts, vibrational spectroscopy is the most sensitive and convenient method. The three modes, ν2(ag)\nu_{2}(a_{g}), ν3(ag)\nu_{3}(a_{g}) and ν27(b1u)\nu_{27}(b_{1u}), mainly involving C=C stretching, can be utilized to determine the electronic charge per molecule. The two totally symmetric aga_{g} modes should usually not be infrared active, but due to electron molecular-vibration (emv) coupling, they can be observed by infrared spectroscopy in the conducting plane. As a result of the coupling to the electronic background, the position of the ν2(ag)\nu_{2}(a_{g}) and ν3(ag)\nu_{3}(a_{g}) infrared feature is shifted – compared to the Raman mode – and depends on the midinfrared charge-transfer band. Hence, this way a precise determination of the charge distribution can become pretty tricky. The antisymmetric molecular vibration ν27(b1u)\nu_{27}(b_{1u}), on the other hand, is infrared active and – in first approximation – its frequency depends linearly on the molecular charge [36]; this makes it perfectly suited for evaluating the charge per molecule. Due to the stacking of the donor molecules in these salts, it can be best observed perpendicular to conducting BETS layers.

Refer to caption
Figure 3: Out-of-plane optical conductivity of (a) κ\kappa-(BETS)2GaCl4 and (b) λ\lambda-(BETS)2GaCl4 in the spectral range of the ν27(b1u)\nu_{27}(b_{1u}) mode measured in the temperature range from 295295 to 9 K. For clarity reasons, the curves are shifted with respect to each other. While for κ\kappa-(BETS)2GaCl4 two ν27\nu_{27}-related peaks are observed at low temperatures, which are gradually smeared out upon warming up; there are two additional satellite features (\star) identified in the λ\lambda-salts, which appear only below approximately 150 K.

Fig. 3 displays the optical conductivity of κ\kappa- and λ\lambda-(BETS)2GaCl4 in the spectral range of the ν27(b1u)\nu_{27}(b_{1u}) mode, recorded at different temperatures, as indicated. Compared to the BEDT-TTF salts, this molecular vibration shows up slightly lower in frequency because four of the eight sulfur atoms are substituted by heavier selenium atoms in BETS [cf. Fig. 1(c)] [39]. At room temperature, a broad peak is observed around 1453 cm1{\rm cm}^{-1} in κ\kappa-(BETS)2GaCl4, which is assigned to ν27(b1u)\nu_{27}(b_{1u}). When the temperature is lowered, the feature becomes narrower and below 200 K two peaks can be clearly distinguished. The behavior found in the λ\lambda-compound is rather different: here the ν27(b1u)\nu_{27}(b_{1u}) mode appears as a doublet with frequencies 1454 and 1458 cm1{\rm cm}^{-1} already at room temperature. Upon cooling a typical narrowing and blue shift occurs, but the separation of the two peaks does not change in the entire temperature range. We explain the presence of these two distinct peaks by the lower crystal symmetry. Most important, however, are the two side peaks at 1451 and 1471 cm1{\rm cm}^{-1} that appear next to this doublet when the temperature is reduced below T=150T=150 K, i.e. when the resistivity ρ(T)\rho(T) starts to develop its maximum reached at Tmax=91T_{\max}=91 K (Fig. 2).

The ν28(b1u)\nu_{28}(b_{1u}) mode, which is related to ethylene-end groups, is present in both compounds. It shows up as a small peak around 1400 cm1{\rm cm}^{-1} in κ\kappa-(BETS)2GaCl4, and as a doublet around 1405-1410 cm1{\rm cm}^{-1} in the λ\lambda-analogue due to the lower P1¯\bar{1} symmetry of the latter salt. It is interesting to note that the peak seen around 1423-1425 cm1{\rm cm}^{-1} is very similar to one observed in the quantum spin liquid κ\kappa-(BEDT-TTF)2Cu2(CN)3 [40]. The tiny features near 1480 cm1{\rm cm}^{-1} in both compounds might correspond to ν2(ag)\nu_{2}(a_{g}). No temperature dependence was observed for these modes nor for the shoulder at 1442 cm1{\rm cm}^{-1}.

IV Discussion

IV.1 Charge sensitive ν27(b1u)\nu_{27}(b_{1u}) mode

Fig. 4(a) illustrates the significantly different shape of the vibrational modes in the κ\kappa- and λ\lambda-salts by comparing the T=9T=9 K spectra in the same graph. While for λ\lambda-(BETS)2GaCl4 the modes can be fitted by simple Lorentz functions, the excitation in κ\kappa-(BETS)2GaCl4 are surprisingly asymmetric indicating some influence of the electronic background. Typically no sizeable conductivity occurs in the out-of-plane direction, however, for κ\kappa-type salts the donor molecules are tilted within the plane, leading to some appreciable coupling to the in-plane conductivity. In order to account for the interaction of the vibrational modes with the electronic background, the Fano model should be applied in the case of in κ\kappa-(BETS)2GaCl4 [40].

Refer to caption
Figure 4: Analysis of the vibrational feature ν27(b1u)\nu_{27}(b_{1u}). (a) Comparison of the low-temperature (T=9T=9 K) optical conductivity of κ\kappa-(BETS)2GaCl4 and λ\lambda-(BETS)2GaCl4 in the range of the charge-sensitive ν27(b1u)\nu_{27}(b_{1u}) modes. For κ\kappa-(BETS)2GaCl4 (red) a splitting of approximately 9 cm1{\rm cm}^{-1} occurs due to the double-layer structure. In the case of λ\lambda-(BETS)2GaCl4 (blue), the pronounced doublet at 1459.1 and 1462.7 cm1{\rm cm}^{-1} results from crystallographically distinct molecules; the static separation is temperature independent. In addition two broad modes appear at 1453.5 and 1473 cm1{\rm cm}^{-1} due to fluctuating charge. (b) Temperature evolution of peak positions. The red crosses correspond to the modes of κ\kappa-(BETS)2GaCl4. The blue solid dots represent the fluctuating modes of λ\lambda-(BETS)2GaCl4, while the open blue circles are related to the static modes of this compound. (c) Temperature dependent charge imbalance 2δ2\delta for λ\lambda-(BETS)2GaCl4, calculated from the resonance frequencies of panel (b). (d) Linewidth of all modes plotted as a function of temperature on a logarithmic scale.

In general, a doubling of charge sensitive modes is taken as evidence for inequivalent bonds due to donor molecules containing different amount of charge 2δ=ρrichρpoor2\delta=\rho_{\rm rich}-\rho_{\rm poor}; the observed peak separation Δν27\Delta\nu_{27} allows us to evaluate the charge imbalance 2δ=Δν27/(140cm1/e)2\delta=\Delta\nu_{27}/(140~{}{\rm cm}^{-1}/e), assuming the same relation as reported for the ν27(b1u)\nu_{27}(b_{1u}) mode of BEDT-TTF salts [36]. Despite the fact that the spectrum is shifted to lower frequencies by almost 10 cm1{\rm cm}^{-1} due to the heavier selenium atoms, our finding can be compared with the observations Sedlmeier et al. made on several κ\kappa-phase BEDT-TTF compounds [40]. Such minor satellite peaks occur in κ\kappa-(BEDT-TTF)2Cu[N(CN)2]Br and κ\kappa-(BEDT-TTF)2Cu[N(CN)2]Cl and were ascribed to Davydov-like splitting; Maksimuk et.al. gave a detailed description of this phenomenon including symmetry considerations and selection rules [41]. The situation is similar for κ\kappa-BETS, where the unit cell is doubled due to alternating tilting of the BETS molecules within the planes, as depinced in Fig. 1(b). We conclude therefore that the two maxima separated by Δν=9\Delta\nu=9 cm1{\rm cm}^{-1} are due to structural reasons. This splitting is well resolved at low temperatures but can be followed all the way up to room temperature where it is smaller by about 30%. We exclude the presence of static or dynamic charge disproportionation at any temperature.

The case of λ\lambda-(BETS)2GaCl4 is more complex because there are two rather sharp central peaks, which are distinct by only 3.6 cm1{\rm cm}^{-1}. We assign the doublet to the ν27(b1u)\nu_{27}(b_{1u}) mode of the BETS molecules bearing +0.5e+0.5~{}e charge. Due to the low symmetry of the λ\lambda-compound, there are two inequivalent molecules in the unit cell, resulting in a doubling of this vibrational feature; the unequal bonds correspond to a static charge imbalance of only 0.02e0.02e. No variation with temperature is observed, except of the common thermal smearing above 100 K, demonstrated in Fig. 4(d). The surprising observation is that in addition to this central doublet, two side bands appear below T150T\approx 150 K, right when the resistivity behavior changes from metallic to semiconducting upon cooling (Fig. 2). The emergence of these features is attributed to the noticeable reorganization of charge and will be discussed in all detail in the following.

From Figs. 3(b) and 4(d) we see that the width of central and side peaks in λ\lambda-(BETS)2GaCl4 differs significantly. While the central modes are very narrow, with more than 5 cm1{\rm cm}^{-1} linewidth, the side bands are quite broad, even at low temperatures. We suggest that the satellite peaks are due to a dynamic charge imbalance of 2δfluct=0.14e2\delta_{\rm fluct}=0.14e emerging only for T<150T<150 K. As a matter of fact, each satellite peak should contain a doublet, as well 111There are some indications of a doublet at the high-frequency satellite, but these features are too weak to make a strong point.; however our data do not allow such an assignment reliably due to the broad features. The splitting of the sidebands was calculated using the average frequency of the central doublet. In addition, we should note that from 77Se-NMR measurements on λ\lambda-(BETS)2GaCl4 a charge distribution 2δ=0.15e2\delta=0.15e was concluded considering the low-temperature Knight shift [29], rather similar to the charge disproportionation we found. It is interesting to compare our observations with those reported for the superconductor β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CH2CF2SO3 where at low temperature broad fluctuation modes coexist with localized charges, seen in well separated sharp peaks [43, *girlando2012spectroscopic].

Finally, we want to point out that the ν27(b1u)\nu_{27}(b_{1u}) molecular vibrations involve the inner C=C double bonds. These should not be significantly affected by the freezing-out of the ethylene end-groups motion, that was suggested as a source of charge imbalance by Kobayashi et al. [30]. Varying the cooling rate is a common approach for investigating freezing effects. Yakushi et al. report that the width of the ν2(ag)\nu_{2}(a_{g}) Raman mode is affected by the cooling rate in κ\kappa-(BEDT-TTF)2Cu2(CN)3 and relate that to disorder of ethylene end-groups at low temperature [45]. However, we could not identify any dependence of the charge sensitive ν27\nu_{27}(b1ub_{1u}) infrared mode on the cooling rate [46]. In addition, the ν28\nu_{28} mode, which involves the terminal CH2-groups, is also seen up to room temperature (Fig. 3) with no strong variation around 100 K. Since in the present compound the ethylene end-groups are closely linked to the GaCl4{}_{4}^{-} anions, investigation of infrared active ν3\nu_{3} vibrational mode of the tetrahedral GaCl4{}_{4}^{-} anions (located in the far-infrared range) is highly appreciable to shed light on the influence of ethylene motion on physical properties of λ\lambda-(BETS)2GaCl4. More specific studies in this regard are in progress.

As far as the mode strength is concerned, the intensity of vibrational modes generally increases with electronic charge, corresponding to a shift to lower frequencies, as shown by quantum mechanical calculations [36, *girlando2012spectroscopic]. Indeed, the peak observed in λ\lambda-(BETS)2GaCl4 at 14531453 cm1{\rm cm}^{-1} has a much higher intensity compared to one at 15731573 cm1{\rm cm}^{-1}.

We have seen in Fig. 2 that the κ\kappa-phase salt is a metal at all temperatures; therefore it is not surprising to find only broad vibrational modes even in the perpendicular direction. The non-monotonous resistivity ρ(T)\rho(T) for λ\lambda-(BETS)2GaCl4 that goes through a maximum around Tmax=91T_{\rm max}=91 K upon cooling, indicates that electronic charges are prone to localization, because electronic correlation are more important [31]. The presence of superconductivity right at the metal-insulator transition in the phase diagram raises the question about possible mechanisms.

Now we want to draw the attention to the series of β′′\beta^{\prime\prime}-(BEDT-TTF)2XX salts, that was subject to extended investigations for many years [47, *Schlueter01, 43, 15], including comprehensive optical studies that focussed on the interplay of charge order and superconductivity. By variation of the all-organic anions XX, completely insulating, charge-ordered, superconducting and metallic systems can be achieved. In the latter compound, β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CHFCF2SO3, only a very low degree of charge fluctuations is seen, resulting a very broad feature in the optical conductivity. The superconductor β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CH2CF2SO3, on the other hand, exhibits well distinct peaks corresponding to fluctuating charges [17, 19, *PhysRevB.99.155144]; very much fostering the idea of charge fluctuations as glue for superconductivity.

For a more quantitative analysis, the vibrational features were fitted by one Fano function each in case of κ\kappa-type; for the λ\lambda-type salt one Lorentzian each suffices . The temperature evolution of the peak positions and the calculated charge imbalance 2δ2\delta is plotted in Fig.4(b) and (c), respectively. It is surprising that 2δ2\delta does not change significantly upon cooling for any of the features. This behavior is completely different to quasi-one-dimensional (TMTTF)2XX compounds where the charge imbalance increases below TCOT_{\rm CO} in a mean-field manner [49, 50, 51, 52]. The present behavior instead resembles the finding in the superconducting β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CH2CF2SO3 and metallic β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CHFSO3 [20].

The temperature dependence of the linewidth is displayed in Fig.4(d) for all modes considered. The peaks of the κ\kappa-salt are mainly affected by the electronic background and thermal broadening. A similar behavior is observed for the narrow static modes in λ\lambda-(BETS)2GaCl4. The temperature evolution of the sidebands is much smaller, they remain rather broad even at low temperatures. In the following we elaborate on the possible origin of charge fluctuations below T=150T=150 K.

IV.2 Interplay of broken symmetry ground states in spin and charge sectors

In λ\lambda-type salts the Fermi surface comprises cylindrical portions with one-dimensional flat sections [53, 54], thus nesting effects and possible instabilities are expected. Indeed, the spin-density-wave (SDW) state accompanied by the metal-insulator transition was recently observed in λ\lambda-(BETS)2GaBrxCl4-x series with x=0.75x=0.75 [27]. No doubt that this state of itinerant antiferromagnetism certainly affects electronic properties of adjacent λ\lambda-(BETS)2GaCl4, where the increase of 1/T1T1/T_{1}T just above the superconducting transition was ascribed to the SDW fluctuations [25], and as a result, magnetic fluctuations were considered as a pairing mechanism of superconductivity.

Our current infrared measurements reveal that in λ\lambda-(BETS)2GaCl4 not only magnetic fluctuations but also charge fluctuations are present at low temperature. In this case, the simple Hubbard model is insufficient to describe the underlying physics, instead both on-site Coulomb repulsion UU and inter-site interaction VV has to be takes into account. The properties of a square lattice at half-filling by using the extended Hubbard model, infer coexistence of spin and charge density waves (SDW and CDW) near the superconducting state when V=U/4V=U/4 [55]. Along this lines, we may now explain how superconductivity arises when both charge- and spin-fluctuations are present.

For the one-dimensional Bechgaard salt (TMTSF)2PF6, for instance, several groups have established the low-temperature SDW phase next superconductivity [56, 57, 58, 59]. X-ray scattering experiments revealed features associated with a CDW state even for temperatures below TSDWT_{\rm SDW} [60, 61]; thus Clay, Mazumdar and collaborators concluded that charge and spin density waves coexist [62, 63]. Since for (TMTSF)2PF6 and λ\lambda-(BETS)2GaBr0.75Cl3.25 a qualitatively similar situation is observed, we expect that in the latter compound a CDW state can coexist with a SDW ground state. This implies that charge fluctuations and spin fluctuations may occur next to and even cause the superconducting state. Interestingly, several of the strongly dimerized κ\kappa-type salts with dominant magnetic fluctuations exhibit superconducting transition temperatures of 10 K and more, while those systems with no or weak dimerization TcT_{c} barely exceeds 5 K. We might speculate that in the present case of λ\lambda-BETS the charge imbalance prevents a larger TcT_{c}. Further investigations of the neighboring insulating state are required to eventually prove this suggestion. Pressure-dependent studies could be an option to tune the systems continuously albeit rather challenging. Beside a few examples in the field of quantum spin liquids [64, 65, *Rosslhuber21], here we want to mention the dimer Mott insulator β\beta^{\prime}-(BEDT-TTF)2ICl2 where superconductivity under pressure is discussed near spin- and charge-density-wave phases [67].

Finally, let us recall the series α\alpha-(BEDT-TTF)2MMHg(SCN)4, with MM = K, Tl, Rb, NH4, which is a typical quarter-filled system with clear indications of charge order. Interestingly, MM = NH4 is superconducting while for the other compounds various density-wave scenarios have been suggested because the electronic structure contains quasi-one-dimensional bands at EFE_{F} [68, 35, 69, *Dressel05, *Drichko06, *Merino06, 73, *Christ00, *Andres05, 76]. As the system is at the verge of a charge-ordered phase, the effect of charge fluctuations on superconductivity is evident and should become subject of further investigations, including the recently synthesized modified compounds which also show density wave states [77, *Ohnuma19].

At this point, we see that both charge and spin fluctuations may be present at low temperature in λ\lambda-(BETS)2GaCl4, but their influence on superconductivity is not completely clear. In the case of the half-filled κ\kappa-(BEDT-TTF)2X salts, where V=0V=0 and only UU is present, charge fluctuations were found to be insignificant. For the quarter-filled β′′\beta^{\prime\prime}-type systems, where VV is strong, a link between charge fluctuations and superconductivity were suggested theoretically and proven experimentally [16, 17, 19]. This is consistent with the phase diagram proposed by Onari et al. [55]. When both UU and VV are present, it is difficult to say whether spin- and charge- fluctuations compete, whether they enhance each other, and in which way their interplay leads to superconductivity. Further investigations are needed to draw firm conclusions. Nevertheless, the present study clearly shows that when going from the bare metal κ\kappa-(BETS)2GaCl4 to the superconductor λ\lambda-(BETS)2GaCl4, the influence of the charge degrees of freedom increases. The λ\lambda-type salts are another example for the idea that charge fluctuations affect superconductivity.

V Conclusions

In summary, infrared optical spectroscopy was carried out in the range of the charge-sensitive ν27(b1u)\nu_{27}(b_{1u}) mode in order to learn about the interplay of charge order and superconductivity in the two-dimensional organic conductor λ\lambda-(BETS)2GaCl4, using κ\kappa-(BETS)2GaCl4 as a reference. In the latter compound, alternating layers and metallic conductivity results in Fano-shaped slightly distinct modes. When we go to the superconducting λ\lambda-(BETS)2GaCl4, we observe slight splitting of the ν27\nu_{27} mode due to the lower symmetry structure. Most important, two side modes appear upon cooling which are assigned to a dynamical charge imbalance of 2δ0.14e2\delta\approx 0.14e. We suggest a situation similar to α\alpha- and β′′\beta^{\prime\prime}-type salts, where charge fluctuations are closely linked to superconductivity and contribute to their pairing mechanism. In accord with theoretical proposals, we suggest that these charge fluctuations originate in a dynamic charge-density-wave state, which coexists with the dynamic spin-density-wave state evidenced by NMR measurements. A complete picture should eventually explain the properties of charge-ordered α\alpha- β\beta^{\prime}- and β′′\beta^{\prime\prime}-compounds together with the observations of λ\lambda-(BETS)2GaCl4.

VI Acknowledgements

We would like to thank K. Tsuji and also G. Untereiner for her crucial technical support. The work was supported by the Deutsche Forschungsgemeinschaft via DR228/39-3.

References

  • White et al. [2015] B. White, J. Thompson, and M. Maple, Unconventional superconductivity in heavy-fermion compounds, Physica C 514, 246 (2015).
  • Plakida [2010] N. Plakida, High-temperature cuprate superconductors, Springer Series of Solid-State Science, Vol. 166 (Springer-Verlag, Berlin, 2010).
  • Yamauchi et al. [2005] T. Yamauchi, M. Isobe, and Y. Ueda, Charge order and superconductivity in vanadium oxides, Solid State Sci. 7, 874 (2005).
  • Lebed [2008] A. G. Lebed, The physics of organic superconductors and conductors (Springer, Berlin, 2008).
  • Wosnitza [2019] J. Wosnitza, Superconductivity of organic charge-transfer salts, J. Low Temp. Phys. 197, 250 (2019).
  • Moriya and Ueda [2000] T. Moriya and K. Ueda, Spin fluctuations and high temperature superconductivity, Adv. Phys. 49, 555 (2000).
  • Misawa and Imada [2014] T. Misawa and M. Imada, Origin of high-Tc{T}_{c} superconductivity in doped Hubbard models and their extensions: Roles of uniform charge fluctuations, Phys. Rev. B 90, 115137 (2014).
  • Vojta et al. [2006] M. Vojta, T. Vojta, and R. K. Kaul, Spin excitations in fluctuating stripe phases of doped cuprate superconductors, Phys. Rev. Lett. 97, 097001 (2006).
  • Torchinsky et al. [2013] D. H. Torchinsky, F. Mahmood, A. T. Bollinger, I. Božović, and N. Gedik, Fluctuating charge-density waves in a cuprate superconductor, Nature Mat. 12, 387 (2013).
  • McKenzie [1997] R. H. McKenzie, Similarities between organic and cuprate superconductors, Science 278, 820 (1997).
  • Imada et al. [1998] M. Imada, A. Fujimori, and Y. Tokura, Metal-insulator transitions, Rev. Mod. Phys. 70, 1039 (1998).
  • Kuroki and Aoki [1999] K. Kuroki and H. Aoki, Superconductivity and spin correlation in organic conductors: A quantum Monte Carlo study, Phys. Rev. B 60, 3060 (1999).
  • Kondo and Moriya [1998] H. Kondo and T. Moriya, Spin fluctuation-induced superconductivity in organic compounds, J. Phys. Soc. Jpn. 67, 3695 (1998).
  • Lefebvre et al. [2000] S. Lefebvre, P. Wzietek, S. Brown, C. Bourbonnais, D. Jérome, C. Mézière, M. Fourmigué, and P. Batail, Mott transition, antiferromagnetism, and unconventional superconductivity in layered organic superconductors, Phys. Rev. Lett. 85, 5420 (2000).
  • Dressel and Tomić [2020] M. Dressel and S. Tomić, Molecular quantum materials: electronic phases and charge dynamics in two-dimensional organic solids, Adv. Phys. 69, 1 (2020).
  • Merino and McKenzie [2001] J. Merino and R. H. McKenzie, Superconductivity mediated by charge fluctuations in layered molecular crystals, Phys. Rev. Lett. 87, 237002 (2001).
  • Kaiser et al. [2010] S. Kaiser, M. Dressel, Y. Sun, A. Greco, J. J.A.Schlueter, G. Gard, and N. Drichko, Bandwidth tuning triggers interplay of charge order and superconductivity in two-dimensional organic materials, Phys. Rev. Lett. 105, 206402 (2010).
  • Kaiser et al. [2012] S. Kaiser, S. Yasin, N. Drichko, M. Dressel, T. Rõõm, D. Hüvonen, U. Nagel, G. L. Gard, and J. A. Schlueter, Optical investigations of the superconducting energy gap in β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CH2CF2SO3phys. stat. sol. (b) 249, 985 (2012).
  • Pustogow et al. [2019a] A. Pustogow, Y. Saito, A. Rohwer, J. A. Schlueter, and M. Dressel, Coexistence of charge order and superconductivity in β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CH2CF2SO3Phys. Rev. B 99, 140509 (2019a).
  • Pustogow et al. [2019b] A. Pustogow, K. Treptow, A. Rohwer, Y. Saito, M. Sanz Alonso, A. Löhle, J. A. Schlueter, and M. Dressel, Charge order in β′′\beta^{\prime\prime}-phase BEDT-TTF salts, Phys. Rev. B 99, 155144 (2019b).
  • Mori et al. [2001] H. Mori, T. Okano, M. Kamiya, M. Haemori, H. Suzuki, S. Tanaka, Y. Nishio, K. Kajita, and H. Moriyama, Bandwidth and band filling control in organic conductors, Physica C 357-360, 103 (2001).
  • Minamidate et al. [2015] T. Minamidate, Y. Oka, H. Shindo, T. Yamazaki, N. Matsunaga, K. Nomura, and A. Kawamoto, Superconducting phase in λ\lambda-(BEDT-STF)2GaCl4 at high pressures, J. Phys. Soc. Jpn. 84, 063704 (2015).
  • Saito et al. [2018] Y. Saito, S. Fukuoka, T. Kobayashi, A. Kawamoto, and H. Mori, Antiferromagnetic ordering in organic conductor λ\lambda-(BEDT-TTF)2GaCl4 probed by 13C NMR, J. Phys. Soc. Jpn. 87, 013707 (2018).
  • Saito et al. [2019] Y. Saito, H. Nakamura, M. Sawada, T. Yamazaki, S. Fukuoka, N. Matsunaga, K. Nomura, M. Dressel, and A. Kawamoto, Disordered quantum spin state in the stripe lattice system consisting of triangular and square tilings (2019), arXiv:1910.09963 .
  • Kobayashi and Kawamoto [2017] T. Kobayashi and A. Kawamoto, Evidence of antiferromagnetic fluctuation in the unconventional superconductor λ\lambda-(BETS)2GaCl4 by 13C NMR, Phys. Rev. B 96, 125115 (2017).
  • Tanaka et al. [1999] H. Tanaka, A. Kobayashi, A. Sato, H. Akutsu, and H. Kobayashi, Chemical control of electrical properties and phase diagram of a series of λ\lambda-type BETS superconductors, λ\lambda-(BETS)2GaBrxCl4-xJ. Am. Chem. Soc. 121, 760 (1999).
  • Kobayashi et al. [2020a] T. Kobayashi, T. Ishikawa, A. Ohnuma, M. Sawada, N. Matsunaga, H. Uehara, and A. Kawamoto, Spin-density wave in the vicinity of superconducting state in λ\lambda-(BETS)2GaBrxCl4-x probed by 13C NMR spectroscopy, Phys. Rev. Research 2, 023075 (2020a).
  • Hiraki et al. [2007] K.-I. Hiraki, H. Mayaffre, M. Horvatić, C. Berthier, S. Uji, T. Yamaguchi, H. Tanaka, A. Kobayashi, H. Kobayashi, and T. Takahashi, 77Se NMR evidence for the Jaccarino–Peter mechanism in the field induced superconductor, λ\lambda-(BETS)2FeCl4J. Phys. Soc. Jpn. 76, 124708 (2007).
  • Hiraki et al. [2010] K.-I. Hiraki, M. Kitahara, T. Takahashi, H. Mayaffre, M. Horvatić, C. Berthier, S. Uji, H. Tanaka, B. Zhou, A. Kobayashi, and H. Kobayashi, Evidence of charge disproportionation in λ\lambda-type BETS based organic superconductors, J. Phys. Soc. Jpn. 79, 074711 (2010).
  • Kobayashi et al. [2020b] T. Kobayashi, K. Tsuji, A. Ohnuma, and A. Kawamoto, Selective observation of spin and charge dynamics in an organic superconductor λ\lambda-(BETS)2GaCl4 using 69,71Ga NMR measurements, Phys. Rev. B 102, 235131 (2020b).
  • Pesotskii et al. [1999] S. Pesotskii, R. Lyubovskii, N. Kushch, M. V. Kartsovnik, W. Biberacher, K. Andres, H. Kobayashi, and A. Kobayashi, de Haas-van Alphen oscillations and angular magnetoresistance oscillations in the organic metal κ\kappa-(BETS)2GaCl4JETP 88, 114–117 (1999).
  • Pratt et al. [2003] F. Pratt, S. Blundell, I. Marshall, T. Lancaster, S. Lee, A. Drew, U. Divakar, H. Matsui, and N. Toyota, Polyhedron 22, 2307 (2003).
  • Montgomery et al. [1994] L. Montgomery, T. Burgin, J. Huffman, J. Ren, and M.-H. Whangbo, Synthesis, superconductivity, x-ray structure and electronic band structure of λ\lambda-(BETS)2GaCl4Physica C 219, 490 (1994).
  • Kobayashi et al. [1993] A. Kobayashi, T. Udagawa, H. Tomita, T. Naito, and H. Kobayashi, New organic metals based on BETS compounds with MXMX4{}_{4}^{-} anions (BETS = bis(ethylenedithio)tetraselenafulvalene MM = Ga, Fe, In; XX = Cl, Br), Chem. Lett. 22, 2179 (1993).
  • Dressel and Drichko [2004] M. Dressel and N. Drichko, Optical properties of two-dimensional organic conductors: Signatures of charge ordering and correlation effects, Chem. Rev. 104, 5689 (2004).
  • Girlando [2011] A. Girlando, Charge sensitive vibrations and electron-molecular vibration coupling in bis(ethylenedithio)-tetrathiafulvalene (BEDT-TTF), J. Phys. Chem. C 115, 19371 (2011).
  • Painelli and Girlando [1986] A. Painelli and A. Girlando, Electron–molecular vibration (e–mv) coupling in charge‐transfer compounds and its consequences on the optical spectra: A theoretical framework, J. of Chem. Phys. 84, 5655 (1986).
  • Dressel and Grüner [2002] M. Dressel and G. Grüner, Electrodynamics of Solids (Cambridge University Press, Cambridge, 2002).
  • Olejniczak et al. [1996] I. Olejniczak, A. Graja, N. Kushch, P. Cassoux, and H. Kobayashi, Polarized ir reflectance studies of the organic conductor κ\kappa-(BETS)2FeCl4J. Phys. I France 6, 1631 (1996).
  • Sedlmeier et al. [2012] K. Sedlmeier, S. Elsässer, D. Neubauer, R. Beyer, D. Wu, T. Ivek, S. Tomić, J. A. Schlueter, and M. Dressel, Absence of charge order in the dimerized κ\kappa-phase BEDT-TTF salts, Phys. Rev. B 86, 245103 (2012).
  • Maksimuk et al. [2001] M. Maksimuk, K. Yakushi, H. Taniguchi, K. Kanoda, and A. Kawamoto, The C=C stretching vibrations of κ\kappa-(BEDT–TTF)2Cu[N(CN)2]Br and its isotope analogues, Journal of the Physical Society of Japan 70, 3728 (2001).
  • Note [1] There are some indications of a doublet at the high-frequency satellite, but these features are too weak to make a strong point.
  • Girlando et al. [2014] A. Girlando, M. Masino, J. A. Schlueter, N. Drichko, S. Kaiser, and M. Dressel, Charge-order fluctuations and superconductivity in two-dimensional organic metals, Phys. Rev. B 89, 174503 (2014).
  • Girlando et al. [2012] A. Girlando, M. Masino, S. Kaiser, Y. Sun, N. Drichko, M. Dressel, and H. Mori, Spectroscopic characterization of charge order fluctuations in BEDT-TTF metals and superconductors, phys. stat. sol. (b) 249, 953 (2012).
  • Foury-Leylekian et al. [2018] P. Foury-Leylekian, V. Ilakovac, V. Balédent, P. Fertey, A. Arakcheeva, O. Milat, D. Petermann, G. Guillier, K. Miyagawa, K. Kanoda, P. Alemany, E. Canadell, S. Tomic, and J.-P. Pouget, (BEDT-TTF)2Cu2(CN)3 spin liquid: Beyond the average structure, Crystals 8, 528 (2018).
  • Sanz Alonso [2016] M. Sanz Alonso, Effect of disorder on the optical propertiesof organic quantum spin-liquid compounds (2016).
  • Ward et al. [2000] B. H. Ward, J. A. Schlueter, U. Geiser, H. H. Wang, E. Morales, J. P. Parakka, S. Y. Thomas, J. M. Williams, P. G. Nixon, R. W. Winter, G. L. Gard, H.-J. Koo, and M.-H. Whangbo, Comparison of the crystal and electronic structures of three 2:1 salts of the organic donor molecule BEDT-TTF with pentafluorothiomethylsulfonate anions SF5CH2SO3{}_{3}^{-}, SF5CHFSO3{}_{3}^{-}, and SF5CF2SO3{}_{3}^{-}Chem. Mater. 12, 343 (2000).
  • Schlueter et al. [2001] J. A. Schlueter, B. H. Ward, U. Geiser, H. H. Wang, A. M. Kini, J. Parakka, E. Morales, H.-J. Koo, M.-H. Whangbo, R. W. Winter, J. Mohtasham, and G. L. Gard, Crystal structure, physical properties and electronic structure of a new organic conductor β′′\beta^{\prime\prime}-(BEDT-TTF)2SF5CHFCF2SO3J. Mater. Chem. 11, 2008 (2001).
  • Dressel et al. [2012] M. Dressel, M. Dumm, T. Knoblauch, and M. Masino, Comprehensive optical investigations of charge order in organic chain compounds (TMTTF)2XXCrystals 2, 528 (2012).
  • Knoblauch and Dressel [2012] T. Knoblauch and M. Dressel, Charge disproportionation in (TMTTF)2XX (XX = PF6, AsF6 and SbF6) investigated by infrared spectroscopy, phys. stat. sol. (c) 9, 1158 (2012).
  • Rösslhuber et al. [2018] R. Rösslhuber, E. Rose, T. Ivek, A. Pustogow, T. Breier, M. Geiger, K. Schrem, G. Untereiner, and M. Dressel, Structural and electronic properties of (TMTTF)2XX salts with tetrahedral anions, Crystals 8, 121 (2018).
  • Pustogow et al. [2016] A. Pustogow, T. Peterseim, S. Kolatschek, L. Engel, and M. Dressel, Electronic correlations versus lattice interactions: Interplay of charge and anion orders in (TMTTF)2XXPhys. Rev. B 94, 195125 (2016).
  • Kobayashi et al. [1996] H. Kobayashi, H. Tomita, T. Naito, A. Kobayashi, F. Sakai, T. Watanabe, and P. Cassoux, New BETS conductors with magnetic anions (BETS = bis(ethylenedithio)tetraselenafulvalene), J. Am. Chem. Soc. 118, 368 (1996).
  • Mielke et al. [2001] C. Mielke, J. Singleton, M.-S. Nam, N. Harrison, C. C. Agosta, B. Fravel, and L. K. Montgomery, Superconducting properties and fermi-surface topology of the quasi-two-dimensional organic superconductor λ\lambda-(BETS)2GaCl4 (BETS = bis(ethylene-dithio)tetraselenafulvalene), J. Phys.: Condens. Mat. 13, 8325 (2001).
  • Onari et al. [2004] S. Onari, R. Arita, K. Kuroki, and H. Aoki, Phase diagram of the two-dimensional extended hubbard model: Phase transitions between different pairing symmetries when charge and spin fluctuations coexist, Phys. Rev. B 70, 094523 (2004).
  • Vuletić et al. [2002] T. Vuletić, P. Auban-Senzier, C. Pasquier, D. Jćrome, M. Héritier, S. Tomić, and K. Bechgaard, Coexistence of superconductivity and spin density wave orderings in the organic superconductor (TMTSF)2PF6, Eur. Phys. J. B 25, 319 (2002).
  • Nagata et al. [2013] S. Nagata, M. Misawa, Y. Ihara, and A. Kawamoto, Commensurability of the spin-density-wave state of (TMTSF)2PF6(\mathrm{TMTSF})_{2}\mathrm{PF}_{6} observed by 13C-NMR, Phys. Rev. Lett. 110, 167001 (2013).
  • Gerasimenko et al. [2014] Y. A. Gerasimenko, S. V. Sanduleanu, V. A. Prudkoglyad, A. V. Kornilov, J. Yamada, J. S. Qualls, and V. M. Pudalov, Coexistence of superconductivity and spin-density wave in (TMTSF)2ClO4: Spatial structure of the two-phase state, Phys. Rev. B 89, 054518 (2014).
  • Narayanan et al. [2014] A. Narayanan, A. Kiswandhi, D. Graf, J. Brooks, and P. Chaikin, Coexistence of spin density waves and superconductivity in (TMTSF)2PF6Phys. Rev. Lett. 112, 146402 (2014).
  • Pouget, J. P. and Ravy, S. [1996] Pouget, J. P. and Ravy, S., Structural aspects of the bechgaard salts and related compounds, J. Phys. I France 6, 1501 (1996).
  • Pouget and Ravy [1997] J. Pouget and S. Ravy, X-ray evidence of charge density wave modulations in the magnetic phases of (TMTSF)2PF6 and (TMTTF)2Br, Synt. Met. 85, 1523 (1997).
  • Mazumdar et al. [1999] S. Mazumdar, S. Ramasesha, R. Torsten Clay, and D. K. Campbell, Theory of coexisting charge- and spin-density waves in (TMTTF)2Br, (TMTSF)2PF6, and α\alpha-(BEDT-TTF)2MMHg(SCN)4Phys. Rev. Lett. 82, 1522 (1999).
  • Clay and Mazumdar [2019] R. Clay and S. Mazumdar, From charge- and spin-ordering to superconductivity in the organic charge-transfer solids, Phys. Rep. 788, 1 (2019).
  • Li et al. [2019] W. Li, A. Pustogow, R. Kato, and M. Dressel, Transition of a pristine mott insulator to a correlated fermi liquid: Pressure-dependent optical investigations of a quantum spin liquid, Phys. Rev. B 99, 115137 (2019).
  • Pustogow et al. [2021] A. Pustogow, R. Rösslhuber, Y. Tan, E. Uykur, A. Böhme, M. Wenzel, Y. Saito, A. Löhle, R. Hübner, A. Kawamoto, J. A. Schlueter, V. Dobrosavljević, and M. Dressel, Low-temperature dielectric anomaly arising from electronic phase separation at the mott insulator-metal transition, npj Quantum Materials 6, 9 (2021).
  • Rösslhuber et al. [2021] R. Rösslhuber, A. Pustogow, E. Uykur, A. Böhme, A. Löhle, R. Hübner, J. A. Schlueter, Y. Tan, V. Dobrosavljević, and M. Dressel, Phase coexistence at the first-order mott transition revealed by pressure-dependent dielectric spectroscopy of κ\kappa-(BEDT-TTF)2Cu2(CN)3Phys. Rev. B 103, 125111 (2021).
  • Hashimoto et al. [2015] K. Hashimoto, R. Kobayashi, H. Okamura, H. Taniguchi, Y. Ikemoto, T. Moriwaki, S. Iguchi, M. Naka, S. Ishihara, and T. Sasaki, Emergence of charge degrees of freedom under high pressure in the organic dimer-mott insulator β\beta^{\prime}-(BEDT-TTF)2ICl2Phys. Rev. B 92, 085149 (2015).
  • Mori et al. [1990] H. Mori, S. Tanaka, M. Oshima, G. Saito, T. Mori, Y. Maruyama, and H. Inokuchi, Crystal and electronic structures of (BEDT–TTF)2[MMHg(SCN)4](MM=K and NH4), Bull. Chem. Soc. Jpn. 63, 2183 (1990).
  • Dressel et al. [2003] M. Dressel, N. Drichko, J. Schlueter, and J. Merino, Proximity of the layered organic conductors α\alpha-(BEDT-TTF)2MMHg(SCN)4 (MM=K, NH4) to a  charge-ordering transition, Phys. Rev. Lett. 90, 167002 (2003).
  • Dressel et al. [2005] M. Dressel, N. Drichko, and J. Merino, Evidence of charge ordering in the electronic spectra of two-dimensional organic conductors, Physica B 359-361, 454 (2005).
  • Drichko et al. [2006] N. Drichko, M. Dressel, C. A. Kuntscher, A. Pashkin, A. Greco, J. Merino, and J. Schlueter, Electronic properties of correlated metals in the vicinity of a charge-order transition: Optical spectroscopy of α\alpha-(BEDT-TTF)2MMHg(SCN)4 (MM=NH4, Rb, Tl), Phys. Rev. B 74, 235121 (2006).
  • Merino et al. [2006] J. Merino, A. Greco, N. Drichko, and M. Dressel, Non-fermi liquid behavior in nearly charge ordered layered metals, Phys. Rev. Lett. 96, 216402 (2006).
  • Kartsovnik et al. [1993] M. V. Kartsovnik, A. E. Kovalev, and N. D. Kushch, Magnetotransport investigation of the low-temperature state of transition (BEDT-TTF)2TIHg(SCN)4: Evidence for a Peierls-type transition, J. Phys. I France 3, 1187 (1993).
  • Christ et al. [2000] P. Christ, W. Biberacher, M. V. Kartsovnik, E. Steep, E. Balthes, H. Weiss, and H. Müller, Magnetic field-temperature phase diagram of the organic conductor α\alpha-(BEDT-TTF)2KHg(SCN)4JETP Lett. 71, 303–306 (2000).
  • Andres et al. [2005] D. Andres, M. V. Kartsovnik, W. Biberacher, K. Neumaier, E. Schuberth, and H. Müller, Superconductivity in the charge-density-wave state of the organic metal α\alpha-(BEDT-TTF)2KHg(SCN)4Phys. Rev. B 72, 174513 (2005).
  • Kawai and Kawamoto [2008] T. Kawai and A. Kawamoto, Charge disproportionation and inversion symmetry breaking in organic conductor α(BEDT-TTF)2RbHg(SCN)4\alpha\text{$-$}{(\text{BEDT-TTF})}_{2}\text{RbHg}{(\text{SCN})}_{4}Phys. Rev. B 78, 165119 (2008).
  • Ohnuma et al. [2018] A. Ohnuma, H. Taniguchi, Y. Takahashi, and A. Kawamoto, Density wave state in new α\alpha-type organic conductor, α\alpha-(BEDT-TTF)2MMHg(XXCN)4 (MM = NH4, XX = Se): A key material for universal phase diagram of XX = S and XX = Se systems, J. Phys. Chem. C 122, 24321 (2018).
  • Ohnuma and Kawamoto [2019] A. Ohnuma and A. Kawamoto, 13NMR study of charge fluctuations in α\alpha-(BEDT-TTF)MX salts (X=alkali metal; X=S, Se), Phys. Rev. B 100, 245107 (2019).