Subscriber access provided by Massachusetts Institute of Technology
J. Phys. Soc. Jpn. 87, 114702 (2018) [8 Pages]
FULL PAPERS

Detailed X-band Studies of the π–d Molecular Conductor λ-(BETS)2FeCl4: Observation of Anomalous Angular Dependence of the g-value

+ Affiliations
1Department of Condensed Matter Physics, Graduate School of Science, Hokkaido University, Sapporo 060-0810, Japan2Condensed Molecular Materials Laboratory, RIKEN Cluster for Pioneering Research, RIKEN, Wako, Saitama 351-0198, Japan

The antiferromagnetic insulating ground state of the π–d molecular conductor λ-(BETS)2FeCl4 has been under intense debate for the last decades. One of the difficulties in studying this system comes from the needle-shape of its single crystal where crystallographic a*- and b*-axes are difficult to identify. We performed electron spin resonance (ESR) measurements of λ-(BETS)2FeCl4 with precise angular dependence of the g-value to study the relation between the principal axes of the g-tensor and the crystallographic axes. In the paramagnetic metal phase, the angular dependence of the g-value shows a characteristic “canine-teeth” structure owing to the strong π–d interaction. This structure develops rapidly below 150 K suggesting the enhancement of the π–d spin correlation. The principal axes of the g-value are related to the crystal axes, and a simple two-step method using ESR measurements, (i) find the g-value maximum in the a*b*-plane, and (ii) find the linewidth maximum in the b*c-plane, was examined to find the easy-axis of the antiferromagnetic state. Using the two-step method, we found that antiferromagnetic resonance of the easy-axis appears below 11 K, which is higher than the metal–insulator transition temperature.

©2018 The Physical Society of Japan
1. Introduction

Molecular conductor λ-(BETS)2FeCl4, where BETS is bis(ethylenedithio)tetraselena-fulvalene, is one of the most well-known πd system since it shows a superconductivity at high magnetic field.1,2) The system consists of conducting π-electrons (\(S_{\pi} = 1/2\)) from the BETS layers and localized magnetic d-electrons (\(S_{d} = 5/2\)) from the FeCl4 anions. It is expected that there is a strong exchange interaction between the π- and d-electrons due to many short cation-anion contacts, namely, Se/S⋯Cl contacts which are shorter than the corresponding van der Waals contacts as partly shown in Fig. 1.3,4) Although the field-induced superconducting phase that appears above 17 T is well understood by the Jaccarino–Peter compensation effect,1,2,57) the antiferromagnetic insulating (AFI) ground state is under strong debate for the past few decades.8)


Figure 1. (Color online) Crystal structure of λ-(BETS)2FeCl4 projected to the \(a^{*}b^{*}\)-plane in the reciprocal lattice. The black, orange, yellow, red, and green balls represent the C, Se, S, Fe, and Cl atoms, respectively. The thick dotted lines are short Se/S⋯Cl contacts that are numbered in accordance with Ref. 4. The blue solid line is the \((1\bar{1}0)\) plane, and the angle θ represents the direction of the applied magnetic field B measured from \(a^{*}-b^{*}\). The \(a'\)- and \(b'\)-axes are projection of the a- and b-axes to the \(a^{*}b^{*}\)-plane.

In the high temperature phase, λ-(BETS)2FeCl4 is metallic and paramagnetic. Accompanied with an antiferromagnetic long-range order, λ-(BETS)2FeCl4 shows a metal–insulator (MI) transition at \(T_{\text{MI}} = 8.3\) K.3,9) It was initially proposed that this MI transition of π-electrons is caused by the antiferromagnetic long-range order of d-electrons.9) However, Akiba et al. found a huge excess specific heat below \(T_{\text{MI}}\), and its origin was explained in terms of the “paramagnetic Fe” model, where d-electrons remain paramagnetic and only π-electrons become antiferromagnetic and insulating below \(T_{\text{MI}}\).10) Recent transport, magnetic torque and magnetocaloric studies by Sugiura et al. seem to agree with the “paramagnetic Fe” model.1113)

In contrast to the aforementioned macroscopic measurements, electron spin resonance (ESR) measurement is one of the most sensitive microscopic measurements, which directly observes the electron spins. ESR was also used to control the field-induced superconducting state in λ-(BETS)2FexGa\(_{1-x}\)Cl4.14) Previous ESR studies of λ-(BETS)2FeCl4 indicated no observation of electron paramagnetic resonance (EPR) below \(T_{\text{MI}}\).8,9,15,16) Instead, a transition from EPR to the antiferromagnetic resonance (AFMR) was clearly observed below \(T_{\text{MI}}\), which is in contradiction to the “paramagnetic Fe” model. This lack of EPR observation below \(T_{\text{MI}}\) is critical for the “paramagnetic Fe” model, but it is never considered seriously in the literature that supports the model.1013,1720) Meanwhile, it is also fair to mention that there are some disagreements in the previous EPR studies above \(T_{\text{MI}}\).3,9,21) A review paper about this mysterious AFI phase can be found elsewhere.8)

We suspect that one of the reasons of the disagreement in the EPR studies originates from the needle-shape and the triclinic unit cell of the λ-(BETS)2FeCl4 single crystal. Although the needle-axis is easily assigned to the c-axis, we have noticed from X-ray diffraction analyses that, for the most cases, the largest surface on the needle-like crystal is parallel to neither a- nor b-axis. This induces a critical problem in studies which use aligned polycrystalline samples, since the a- and b-axes can be confused. It is also crucial for studying the AFI phase since the easy-axis is tilted at an about 30° angle from the c-axis to the \(b^{*}\)-axis, where the symbol \(*\) expresses the reciprocal lattice.22) Hence, it is extremely important to assign beforehand the crystal axes of the single crystal by the X-ray diffraction method.

From the ESR research point of view, λ-(BETS)2FeCl4 is also an interesting system. In general, the anisotropy of the g-value, deduced from the angular dependence of EPR, can be expressed by a rotational ellipsoid reflecting the symmetry of the molecular orbital where the unpaired spin is situated. Hence, a fairly large anisotropy of the g-value is expected for EPR originating from the d-electrons while a smaller anisotropy is expected for the π-electrons. In the meantime, the πd interaction of this system is an indirect exchange between the π-electron on the BETS molecule and the d-electron of the Fe via a non-magnetic Cl. However, there are 7 short Se/S⋯Cl contacts between various BETS molecules and the FeCl4 anion, which lead to complex exchange paths with different exchange energies between the FeCl4 and several BETS molecules inside the unit cell.3,4) It is not trivial how the g-anisotropy will be affected by such πd exchange couplings with several exchange paths. Kawamata and co-authors reported the angle- and temperature-dependence of EPR for λ-(BETS)2FeCl4 using a single crystal.21) Although they observed a huge anisotropy of the g-value in the conducting \(a^{*}c\)-plane at 20 K, the range of the magnetic field direction covered by them is narrow for the triclinic system that shows many cation-anion contacts with various directions, and the obtained results are not so clear due to the broad EPR linewidth in the conducting plane. Therefore, it is interesting and worthwhile to study more precisely the EPR behavior of the λ-(BETS)2FeCl4 single crystal.

In this paper, we report detailed EPR studies of the λ-(BETS)2FeCl4 single crystal using X-band ESR. We found that the angular dependence of EPR in λ-(BETS)2FeCl4 shows anomalous behavior owing to the strong πd interaction. Although the anisotropy of the g-value at room temperature can be expressed by a typical ellipsoid, two distinct minima of the g-value are observed in the low temperature region when the magnetic field is applied along specific Se/S⋯Cl contacts. Such anomalous angular dependence of EPR indicates that π and d-electrons are strongly coupled especially at low temperature.

Moreover, it was found that the \(a^{*}\)- and \(b^{*}\)-axes of λ-(BETS)2FeCl4 correspond to the minimum and the maximum of the g-value in the \(a^{*}b^{*}\)-plane, respectively, and the easy-axis of the AFI state is along the direction with the maximum EPR linewidth in the \(b^{*}c\)-plane. Hence, it is possible to determine the easy-axis of the AFI state by a two-step method using ESR, (i) find the maximum of the g-value in the \(a^{*}b^{*}\)-plane (i.e., determination of the \(b^{*}\)-axis), and then (ii) find the linewidth maximum in the \(b^{*}c\)-plane. The antiferromagnetic easy-axis, where the easy-axis mode of AFMR and the spin-flop resonance are both observed, is inclined at about 30° from the c-axis toward the \(b^{*}\)-axis in consistency with the previous torque measurements.22) Our results suggest that the principal crystal axes of λ-(BETS)2FeCl4 and the easy-axis of the AFI state are easily defined from the ESR measurement, which facilitates future studies of the AFI ground state in λ-(BETS)2FeCl4.

2. Experimental Methods

Needle-shaped single crystals of λ-(BETS)2FeCl4 were prepared by electrochemical oxidation of BETS in a mixed solvent of 90% chlorobenzene and 10% ethanol with the tetraethylammonium salt of FeCl4 as a supporting electrolyte under argon atmosphere. Single crystal X-ray diffraction data were collected using a Rigaku HyPix-6000 AFC system with monochromated Mo \(K\alpha\) radiation down to 85 K. The temperature was controlled by a nitrogen gas cooling system (DX-CS190LD, Japan Thermal Engineering) with cooling rate of 1.5 K/min. The crystal structures were solved and refined using SHELXT 2015/5 and SHELXL-2017/1.23,24) Below 20 K, single crystal X-ray diffraction data were collected by a Rigaku UltraX-6-E Imaging Plate system with monochromated Mo \(K\alpha\) radiation. The single crystal was mounted in the cryostat, cooled by a GM refrigerator, with cooling rate of 1.0 K/min. The crystal structures were solved by direct method (SIR97) and refined using SHEXL97.25,26) The crystal has a triclinic unit cell with space group \(P\bar{1}\). The crystal data at room temperature and 7 K (italic values in parentheses) are as follows; \(a = 16.193\) (15.906) Å, \(b = 18.577\) (18.395) Å, \(c = 6.602\) (6.536) Å, \(\alpha = 98.425\) (98.652)°, \(\beta = 96.627\) (95.706)°, \(\gamma = 112.545\) (112.045)°, \(V = 1782.14\) (1726.84) Å3.

We performed ESR measurements using a conventional X-band ESR system (JEOL JES-RE3X, 9–10 GHz). The magnetic field was swept linearly over the range between 50 and 550 mT for EPR, and between 550 and 1350 mT for AFMR measurements. The measurements were performed in the temperature range between 4.3 K and room temperature using a liquid-helium flow cryostat for the X-band ESR system (Oxford Instruments). A thermocouple was used as a temperature sensor. The initial cooling rate was about 3 K/min. A quartz rod was used as a sample holder, and the quartz rod and the sample was inserted in the cylindrical cavity with TE011 mode. The magnetic field was applied horizontally to the quartz rod, and the angular dependence of ESR was measured by rotating the quartz rod. The error range was about few degrees.

The crystal plane of the sample was checked beforehand by the X-ray diffraction method, and the sample was set on the quartz rod using a silicon grease as shown in Figs. 2(a) and 2(b). We chose a single crystal where the largest surface of the needle-shaped crystal corresponds to the \((1\bar{1}0)\) plane. Hence, we define \(\theta = 0\)° as the direction parallel to \(a^{*}-b^{*}\) for the \(a^{*}b^{*}\)-plane rotation [see Fig. 2(a)]. Note that \(a^{*}\)-/\(b^{*}\)-axis of the reciprocal lattice is perpendicular to the bc-/\(ac\)-plane of the real lattice, respectively. The sample configuration for \(B\parallel\text{$b^{*}c$-plane}\) is shown in Fig. 2(b). We mounted the same sample on the triangular prism-shaped polyethylene jig for applying the magnetic field parallel to the \(b^{*}c\)-plane. For this rotation, the direction of the magnetic field is defined by the angle ϕ, where \(\phi = 0\)° is set to the \(-b^{*}\)-direction. Therefore, the angles \(\phi = 90\) and 180° correspond to the c-axis and the \(b^{*}\)-axis, respectively.


Figure 2. (Color online) Sample configurations in measurements for (a) \(B\parallel\text{$a^{*}b^{*}$-plane}\) and (b) \(B\parallel\text{$b^{*}c$-plane}\). The same sample was mounted on the flat surface of the quartz rod. For each configuration, the cross-section of the sample parallel to the rotation plane of the magnetic field is schematically drawn on the top of the rod. The magnetic field is parallel to \(a^{*}-b^{*}\) at \(\theta = 0\)°, and \(-b^{*}\) at \(\phi = 0\)°, respectively.

3. ESR Results
\(B\parallel \text{$a^{*}b^{*}$-plane}\)

If there is no exchange interaction between the π- and d-electrons, two EPR signals from each electron, \(S_{\pi} = 1/2\) and \(S_{d} = 5/2\), should be observed separately. However, if a strong interaction exists, the local field is averaged out due to the fast exchange, and the spectra should merge to a single line. This is known as the “exchange narrowing”.27,28) As reported in the previous ESR study by Kawamata et al., we observed only a single EPR line in the temperature range between \(T_{\text{MI}}\) and room temperature.21) This suggests that the π- and d-electrons are strongly coupled in this system. For the case of λ-(BETS)2FeCl4, it is also considered that the contribution of the Fe spins in the EPR signal should be dominant in the high temperature phase since the magnetic moment of Fe3+ is larger and the π-electrons are itinerant.8)

The angular dependences of the EPR signal in the \(a^{*}b^{*}\)-plane are obtained for several temperatures, and its g-value and EPR peak-to-peak linewidth are presented in Figs. 3(a) and 3(b), respectively. The raw EPR spectra and the methods of analysis can be found in the supplemental material.29) The relation between the angle θ and the crystal axes can be found in Figs. 1 and 2. As shown as a long-dashed curve in Fig. 3(a), a typical angular dependence of the g-value, which mostly follows the \({g_{\text{eff}} =\sqrt{\smash{g_{\text{max}}^{2}\cos^{2}\theta + g_{\text{min}}^{2}\sin^{2}\theta}\mathstrut}}\) dependence, is observed at 300 K. The maximum and minimum of the g-value are \(g_{\text{max}}\sim 2.05\) and \(g_{\text{min}}\sim 2.04\), and the principal axes of the g-tensor correspond to the \(b'\)- and \(a'\)-axes, respectively. Note that the \(a'\)- and the \(b'\)-axes are defined as the projection of the a- and the b-axes onto the \(a^{*}b^{*}\)-plane, respectively (see also Fig. 1). The principal axes are related to the ligand field of the tetrahedral FeCl4 anion where \(g_{\text{max}}\) is along the Fe–Cl bond parallel to the \(b'\)-axis (see Fig. 1). Moreover, the ESR signal of λ-(BETS)2GaCl4, which consists of only π-electrons, is \(g=1.99{\text{--}}2.02\).21) Thus, the g-factor of the BETS molecule is different with the observed g-value. These results confirm that the EPR signal at 300 K comes mainly from the Fe spins.


Figure 3. (Color online) Angular dependence of (a) the g-value and (b) the peak-to-peak linewidth from the EPR in the \(a^{*}b^{*}\)-plane for 10, 30, 100, 150, 230, 270, 280, and 300 K. The solid lines are eye guide. The long-dashed line for 300 K is the fitting from a typical g-tensor (see text for detail). The definitions of the crystal axes in relation with the angle θ are shown in Figs. 1 and 2.

As the temperature is lowered, a “canine teeth” structure, where two minima of the g-values are observed around \(\theta = 30\) and 90°, starts to develop below 270 K [Fig. 3(a)]. The two minima at 10 K are about \(g_{\text{min}}\sim 1.98\) and 1.99 for \(\theta = 30\) and 90°, respectively. Moreover, the maximum of the g-value, \(g_{\text{max}}\sim 2.05\) around \(\theta = 130\)° at 300 K, increases drastically to 2.11 at 30 K, and the maximum shifts towards \(\theta = 150\)° at 10 K as shown in Fig. 3(a). This suggests the principal axis of \(g_{\text{max}}\) tilts from the \(b'\)-axis toward the \(b^{*}\)-axis by decreasing the temperature. As described later, the tilting of the principal axis might be due to the development of the πd correlation or the deformation of the FeCl4 anion with decreasing temperature. In Fig. 3(b), the maximum and the minimum of the EPR linewidth are found nearby the \(a'\)- and \(b'\)-axes, respectively. The linewidth along the \(a'\)-axis, which is parallel to the conducting plane, becomes extremely broad at 10 K as shown in Fig. 3(b). Such kinds of anomalous EPR behavior (i.e., teeth structure and linewidth broadening) suggest that the πd interaction develops with lowering temperature. The details will be discussed later.

As shown in Fig. 4, the temperature dependence of the EPR signal for \(\theta=20\)°, which is around the g-value minimum in Fig. 3(a), also shows a distinct feature. The g-value starts to decrease below 150 K, and a sudden increase is observed below 12 K. In turn, the linewidth gradually increases by lowering temperature, and a steep linewidth broadening is observed below 12 K. We suppose the gradual decrease of the g-value and the gradual linewidth broadening between 12 and 150 K are due to the development of the spin correlation between π- and d-electrons. In contrast, the drastic g-shift and the linewidth broadening below 12 K are typical of EPR above the Néel temperature, which are due to the development of the antiferromagnetic spin correlation. Let us also note that, consistent with previous studies, the EPR intensity starts to diminish around 12 K and disappears in the AFI phase.8,9,15,16,29)


Figure 4. (Color online) Temperature dependence of the g-value and the peak-to-peak linewidth at \(\theta = 20\)°. The solid lines are eye-guide.

In general, microscopic information surrounding the magnetic ions can be obtained from the angular dependence of the g-value since the g-tensor reflects the symmetry of the ligand field, and its angular dependence behaves as \(g_{\text{eff}} =\sqrt{\smash{g_{\text{max}}^{2}\cos^{2}\theta + g_{\text{min}}^{2}\sin^{2}\theta}\mathstrut}\) (i.e., rotational ellipsoid). Therefore, the observation of the “canine teeth” structure in the angular dependence of the g-value in the low temperature region is quite unconventional.

The two-minima of the teeth structure are observed around \(\theta=30\) and 90°. The direction of the minimum at \(\theta = 30\)° appears to be along the projection of two cation-anion contacts, S⋯Cl (3.836 [3.696] Å, #2) and S⋯Cl (3.650 [3.526] Å, #4) contacts, in the \(a^{*}b^{*}\)-plane which are shown as purple and orange dotted lines in Fig. 1, respectively. The values in the parentheses are their inter-atomic distances at room temperature (italic values for 7 K), and the label of the corresponding cation-anion contact in Fig. 1. Note that the labels of the cation-anion contacts are according to the paper by Mori et al.4) In this paper, they estimated the exchange interactions of these πd couplings from the crystal structure at room temperature, and its values are 0.39 K for #2 and 0.50 K for #4.4) Meanwhile, the minimum at \(\theta=90\)° mostly corresponds to the projection of the S⋯Cl (3.647 [3.553] Å, #1) and Se⋯Cl (3.539 [3.465] Å, #6) contacts shown as green and blue dotted lines in Fig. 1, respectively. The estimated exchange couplings for these contacts are 1.44 and 14.14 K, respectively.4) Let us note that there is no g-value anomaly when the magnetic field is along the S⋯Cl (3.424 [3.389] Å, #5) contact, and contacts #3 and #7 are omitted for clarity in Fig. 1 since they are the second smallest and the smallest πd couplings, respectively.4) Hence, it is interesting that the two-minima of the g-value are observed along the projection of cation-anion contacts with noticeable exchange interaction.

Let us now discuss what makes the “canine teeth” structure in the low temperature region. In the transition metal coordination compounds, it is well known that the “covalent” character of the metal–ligand bond appears in the ligands such as halides, oxides, and sulfides.27) This covalency of the metal–ligand bond can induce the nephelauxetic effect, where the name “nephelauxetic” comes from the Greek meaning “cloud-expanding”. In such a case, the spin density of the central metal ion spread out to the surrounding non-magnetic ions. The expansion of the spin density from the central ion has two main effects: (i) the decrease of the spin–orbit coupling and (ii) the increase of interaction with the surrounding atoms. The former affects the g-value to be close to the g-value of the free spin (i.e., \(g=2\)), and the latter broadens the EPR linewidth since the contribution of the super hyperfine coupling increases.27) The latter effect also contributes to the super exchange interaction if adjacent magnetic ions are present.30,31) The ligand field of isolated FeCl4 anion was studied both theoretically and experimentally, and it was shown that the spin–orbit coupling is reduced by the nephelauxetic effect, and the expansion of the Fe3+ spin density affects the g-value and the single ion anisotropy.32,33)

However, the asymmetric anisotropy of the g-value with the “canine teeth” structure observed in λ-(BETS)2FeCl4 is quite different from the one observed in the previous studies which shows a typical g-anisotropy.32,33) We suppose the difference comes from the magnetic interactions surrounding the FeCl4 anion since the previous studies examined isolated anions.32,33) Considering that (i) the two-minima of the g-value are observed along the Se/S⋯Cl contacts with noticeable interactions, and (ii) its g-value is close to \(g=2\) and linewidth maximum is also observed (i.e., suggesting the reduction of the spin–orbit coupling), the “canine teeth” structure might be a result of the nephelauxetic effect where the spin density of the central metal ion is expanded toward the p-orbital of the Cl which has significant exchange path with the BETS molecule. The elongated spin density of the central metal ion along the indirect exchange path is previously observed in IrCl62− by neutron diffraction.31) Hence, the observation of the “canine teeth” structure with \(g\sim 2\) at \(\theta=30\) and 90° might indicate that the πd interaction is strong along these directions.

It is also interesting that the “canine teeth” structure develops with decreasing temperature as shown in Figs. 3(a) and 4. The decrease of the g-value towards \(g\sim 2\) suggests that the spin–orbit coupling gradually reduces and the spin density of Fe3+ expands with decreasing the temperature. If the above argument is correct, there should be some increase of the πd correlation with decreasing the temperature. Hence, we verified whether the inter-atomic distance of the Se/S⋯Cl contacts and the bond angles of the FeCl4 anion are affected by temperature. The inter-atomic distances of typical Se/S⋯Cl contacts and the Cl–Fe–Cl bonds angles as a function of temperature are shown in Figs. 5(a) and 5(b), respectively. The inset in Fig. 5(a) is the zoomed crystal structure around the FeCl4 anion with the typical Se/S⋯Cl contacts. In Fig. 5(a), the S⋯Cl contacts #3 and #7 are not shown again for clarity since they are the two smallest πd couplings. The Se/S⋯Cl contacts is normalized with the one at room temperature.4) It is found that inter-atomic distances of the S⋯Cl contacts #2 and #4 become much shorter than other Se/S⋯Cl contacts with decreasing temperature as shown in Fig. 5(a). It suggests that the πd couplings #2 and #4 are more enhanced as the temperature decreases. Let us remind that the S⋯Cl contacts #2 and #4 are related with the g-value minimum observed around \(\theta=30\)°. On the other hand, the S⋯Cl contact #1 and Se⋯Cl contact #6, which are related to the g-value minimum at \(\theta=90\)°, show a modest contraction of the inter-atomic distance in Fig. 5(a). However, the bond angles of the FeCl4 anion show significant changes in the Cl(2)–Fe–Cl(1) and Cl(2)–Fe–Cl(4) bond angles with temperature as shown in Fig. 5(b). The changes of these two bond angles suggest that the Fe–Cl(2) bond is pointing toward the Se(7) of the BETS molecule with decreasing temperature. Indeed, the bond angle of Fe–Cl(2)–Se(7) changes from 160.2° at 300 K to 162.5° at 7 K, enhancing the πd coupling #6. Therefore, we conclude the enhancement of the “canine teeth” structure in the low temperature region is owing to the development of the πd coupling in λ-(BETS)2FeCl4.


Figure 5. (Color online) Temperature dependence of (a) the ratio of the inter-atomic distance of the typical Se/S⋯Cl contacts compared with the distance at room temperature, and (b) the Cl–Fe–Cl bond angles in the FeCl4 anion. The inset in (a) shows the typical Se/S⋯Cl contacts surrounding the FeCl4 anion.

To summarize this subsection, a characteristic angular dependence of g-value, which has a “canine teeth” structure, has been observed in λ-(BETS)2FeCl4 below 270 K. The structure is enhanced especially below 150 K. The g-value minima are the results of the asymmetric nephelauxetic effect where the spin density expands along the direction of significant indirect exchange paths, and the development of the “canine teeth” structure is related to the enhancement of the πd exchange couplings as a consequence of the deformation of the tetrahedral coordination in FeCl4 and the development of several significant S(Se)⋯Cl⋯Fe exchange paths with decreasing temperature.

\(B\parallel \text{$b^{*}c$-plane}\)

The results of the measurements for \(B\parallel\text{$a^{*}b^{*}$-plane}\) in the previous section indicate that the crystal axes can be identified by the minimum and the maximum of the g-values within the \(a^{*}b^{*}\)-plane. The \(b^{*}\)-axis corresponds to the maximum of the g-value, and the \(a^{*}\)-axis to one of the minima of the g-value at about 120° from the \(b^{*}\)-axis [see Fig. 3(a)]. Hence, the \(b^{*}\)-axis is defined from the g-value in the \(a^{*}b^{*}\)-plane rotation, and it is possible to perform the EPR measurement in the \(b^{*}c\)-plane by using a triangular prism jig as shown in Fig. 2(b). The angle ϕ indicates the direction of the applied magnetic field, and the sample was set so that the directions of \(\phi = 90\) and 180° are along the c- and \(b^{*}\)-axes, respectively. The direction of the magnetic field in the crystal structure is shown in Fig. 6(a).


Figure 6. (Color online) (a) Crystal structure of λ-(BETS)2FeCl4 and the direction of the magnetic field projected to the \(b^{*}c\)-plane. The thick dotted lines are the same short Se/S⋯Cl contacts as Fig. 2. (b) Angular dependence of the g-value and linewidth within the \(b^{*}c\)-plane at 10 K. The red box is the area where AFMR is observed at 4.5 K [see Fig. 7(a)].

Figure 6(b) shows angular dependence of the g-value and the linewidth within the \(b^{*}c\)-plane measured at 10 K. In this case, an asymmetric canine teeth structure of the g-value is observed. The g-value minima are 2.01 and 2.03 at \(\phi = 30\) and 100°, respectively. Similarly to the results of the \(a^{*}b^{*}\)-plane rotation, a broad EPR linewidth is observed around the g-value minima (i.e., suggesting the nephelauxetic effect). Many projections of the πd contacts, such as Se⋯Cl (#6, blue dotted line), S⋯Cl (#1, green dotted line), and S⋯Cl (#2, purple dotted line) contacts, correspond to the minimum at \(\phi = 30\)° [see Fig. 6(a)]. They are the same contacts that contributed to the g-value minima in the \(a^{*}b^{*}\)-plane. On the other hand, there is no corresponding projection of the Se/S⋯Cl contact around \(\phi=100\)°, where the g-value dip is about 2.03, and no minimum of the g-value is observed along the projection of the S⋯Cl contact (#4, orange dotted line), where the projection is around \(\phi=150\)°. In contrast to the \(a^{*}b^{*}\)-plane rotation which shows two minima with \(g\sim 2\), the minimum value for \(\phi=100\)° is around \(g\sim 2.03\). The difference of the effective g-value from \(g=2\) is due to the effect of the spin–orbit coupling, therefore, it seems that the spin–orbit coupling is still effective at \(\phi=100\)° (\(g\sim 2.03\)) in comparison with \(\phi=30\)° (\(g\sim 2.01\)). Considering that there is no projection of the Se/S⋯Cl contact and the g-value is about 2.03 at \(\phi=100\)°, the minimum of the g-value at \(\phi=100\)° might not be due to the cation-anion interaction but due to the minimum of the g-tensor reflecting the ligand field of FeCl4 anion. Furthermore, the lack of g-value anomaly along \(\phi=150\)° might suggest the absence of the nephelauxetic effect for the S⋯Cl contact #4. Hence, the S⋯Cl contact #2 might be the main origin of the g-value minimum at \(\theta=30\)° in the \(a^{*}b^{*}\)-plane. It is not clear why there is no sign of the nephelauxetic effect for the S⋯Cl contact #4 despite that the exchange coupling is finite, 0.50 K at room temperature.4)

It is known from the magnetic torque measurement by Sasaki et al. that the easy-axis of the antiferromagnetic state forms an angle of about 30° with respect to the c-axis and an angle of about 120° with respect to the \(b^{*}\)-axis within the \(b^{*}c\)-plane.22) In our configuration, this angle corresponds to \(\phi = 60\)° where a linewidth maximum of the EPR line is observed as shown in Fig. 6(b). In general, the maximum/minimum of the EPR linewidth is related with the magnetic anisotropy, and it is understandable that the antiferromagnetic easy-axis appears at the linewidth maximum. Therefore, the easy-axis mode of AFMR should be observed if the magnetic field is applied nearby \(\phi = 60\)° below \(T_{\text{MI}}\) in the AFI phase.

The angular dependence of the ESR spectrum in the high field region (550–1350 mT) at 4.5 K, namely in the AFI phase, is presented in Fig. 7(a). For X-band ESR, EPR is observed around 330 mT for the temperature above \(T_{\text{MI}}\) (i.e., paramagnetic metal phase), however, no EPR is observed at 4.5 K as mentioned above. Instead, two ESR signals are observed around 950 and 1150 mT at \(\phi = 60\)°, and more clear and stronger signals are observed for \(\phi = 65\)° as shown in Fig. 7(a). At \(\phi = 55\)° or 70°, the ESR intensity rapidly decreases, and the two ESR lines merge to a single line. Then, ESR completely disappears at \(\phi = 50\) and 80°. This suggests that the two ESR signals only appear for the magnetic field orientation between \(\phi = 60\) and 65°. The temperature dependence of the ESR spectrum at \(\phi = 65\)° is presented in Fig. 7(b). Although the higher ESR signal does not shift with temperature, the lower ESR signal shifts to lower field as the temperature increases. Meanwhile, the intensity for both ESR signals decreases with increasing temperature. Note that the resonance field for the higher ESR signal, 1150 mT, corresponds to the spin-flop field as previously reported.9,16,22)


Figure 7. (Color online) (a) ESR spectra in the high field region for \(\phi = 50\), 55, 60, 65, 70, and 80° at 4.5 K. (b) Temperature dependence of ESR spectrum at \(\phi = 65\)°. The low- and high-field ESR signals are attributed to the easy-axis mode of AFMR and the spin-flop resonance, respectively. The broken arrow is an eye-guide for the resonance shift of the easy-axis mode with temperature.

From above-mentioned typical behavior, these two ESR lines observed in the AFI phase are ascribed to the easy-axis mode of AFMR and the spin-flop resonance. The behavior of AFMR lines is consistent with previous studies.9,15,16) Since the easy-axis mode and the spin-flop resonance were observed, our results suggest that the antiferromagnetic easy-axis corresponds to \(\phi = 60{\text{--}}65\)°. This result, where the antiferromagnetic easy-axis is about 30° from the c-axis to the \(b^{*}\)-axis, is consistent with the magnetic torque measurements.22) Interestingly, this angle area of \(\phi = 60{\text{--}}65\)° corresponds to the area where the maximum of the EPR linewidth is observed at 10 K as shown in the solid red box in Fig. 6(b). Hence, the antiferromagnetic easy-axis can be assigned from the maximum of the EPR linewidth in the \(b^{*}c\)-plane rotation.

It is also worth to mention that the spin-flop resonance of AFMR is observed below 11 K, which is higher than the temperature of the MI transition, \(T_{\text{MI}}\). The intensity of AFMR is rather small at 11 K, suggesting the antiferromagnetic transition is only partial, and the intensity is enhanced as temperature decreases. Furthermore, as it can be seen in the spin-flop resonance of 4.5 and 7 K in Fig. 7(b), a fine double peak structure of the spin-flop resonance, two identical signals with different linewidth, is observed below \(T_{\text{MI}}\). The double peak structure is also observed in the easy-axis mode of AFMR at 4.5 K in Fig. 7(b). This double peak structure suggests that two kinds of antiferromagnetic spin states exist. It should be noted that splitting of the sextets is observed from 8.3 K down to 3.2 K in the 57Fe Mössbauer spectra, which indicates two different environments for the Fe sites.34) Although only one Fe site is crystallographically independent down to 7 K, we think the fine double peak structure of AFMR is related to the magnetic splitting observed in the Mössbauer measurements.3) The two different environments might be due to the metastable state of the π-electrons as discussed in the next section.

4. Conclusions

We performed detailed angular and temperature dependence of EPR in the paramagnetic metal phase of λ-(BETS)2FeCl4 using X-band ESR. Peculiar angular dependence of the g-value with the canine teeth structure was observed in the \(a^{*}b^{*}\)-plane rotation below 270 K. We found that the minima of the g-value correspond to the projection of the typical cation-anion contacts with noticeable πd exchange interaction.

Thanks to this characteristic canine teeth structure of the g-value, crystal axes such as \(a^{*}\)- and \(b^{*}\)-axes can be easily found from the ESR measurements. It can be an alternative method to determine the crystal axes other than the X-ray diffraction analysis. Moreover, we have presented a useful and straightforward two-step method to determine the antiferromagnetic easy-axis, (i) find the g-value maximum in the \(a^{*}b^{*}\)-plane (i.e., \(b^{*}\)-axis), and (ii) find the linewidth maximum in the \(b^{*}c\)-plane. We think this two-step method using ESR will facilitate future studies of the AFI ground state in λ-(BETS)2FeCl4.

The origin of the canine teeth structure is ascribed to the nephelauxetic effect of the FeCl4 anion where the expansion of the spin density occurs along the prominent πd exchange paths. Our crystal structure analysis also revealed that the inter-atomic S⋯Cl distances and the deformation of the FeCl4 anion play an important role on the development of the canine teeth structure. We suppose such change of the inter-atomic distances and the deformation of the anions with temperature affects the πd exchange couplings, and its exchange interactions develop by decreasing temperature. Although the local displacement of the molecules and the asymmetric deformation of the anions with temperature were observed, we could not totally determine what makes the drastic decrease of the g-value below 150 K as shown in Fig. 4. Furthermore, when the field is applied along projections of a few S⋯Cl contacts, such as #4 and #5, the g-value did not show any anomalies. Hence, a more detailed theoretical analysis of πd exchange couplings at low temperature might be worth to study.

Moreover, the EPR intensity start to diminish below 14 K, and the signal is completely lost at low temperature in the AFI phase (see the supplemental material).29) Instead, AFMR is observed below 11 K, and its intensity grows as temperature decreases. Hence, a slow transition from EPR to AFMR is observed in a wide temperature range around \(T_{\text{MI}}=8.3\) K. On the other hand, the lack of the EPR signal in the low temperature region of the AFI phase, especially below 6 K, suggests that there is no paramagnetic ground state. Hence, it is clear from our point of view that the “paramagnetic Fe” model needs some reconsideration.1013,1720)

The insulating mechanism of λ-(BETS)2FeCl4 at \(T_{\text{MI}}\) is also a point of debate. Two types of insulating mechanisms, charge-driven and spin-driven insulating mechanisms, were proposed. In the charge-driven case, the π-electrons make the insulating state by the Mott transition or charge ordering, while in the spin-driven case, the Fe spins become antiferromagnetic at first and opens a gap at the Fermi surface (see review in Ref. 8). Interestingly, the spin-flop resonance is observed above \(T_{\text{MI}}\) in Fig. 7(b). Although the AFMR signal is small, there is a sign of the antiferromagnetic state even above \(T_{\text{MI}}\). Hence, the magnetic long-range order seems to trigger the MI transition favoring the spin-driven MI transition scenario proposed by Brossard et al.9)

Finally, we would like to briefly comment on the observation of the excess specific heat observed below \(T_{\text{MI}}\) since our results favor the spin-driven insulating mechanism.10) The excess specific heat was first discovered by Negishi et al. where they attributed it to the slow magnetic transition of the short-range antiferromagnetic ordering of the Fe ions.35) Meanwhile, Akiba et al. claimed that this excess specific heat can be explained with the paramagnetic Fe model.10) As it turned out, the first explanation better fits our results. As mentioned above, a small spin-flop resonance (i.e., AFMR) is observed above \(T_{\text{MI}}\) in our ESR measurements. Hence, the antiferromagnetic long-range order of the Fe3+ spins partially occurs above \(T_{\text{MI}}\). This is only partial due to the competition between the πd coupling (i.e., RKKY interaction) and the Kondo effect. The former supports the magnetic order, but in the latter case, conducting π-electrons screen the magnetic moment of the Fe3+ ion, and it would work against the long-range ordering. The screening of the \(3d\) magnetic moment should be especially effective since its spin density seems to largely expand inside the anion. The antiferromagnetic domain gradually grows with decreasing temperature as seen in the temperature dependence of AFMR in Fig. 7(b), and the major magnetic transition occurs at \(T_{\text{MI}}\) to open the gap mostly at the Fermi surface. However, the competition remains below \(T_{\text{MI}}\), and the π-electrons are not fully localized. Such a metastable state of the π-electron below \(T_{\text{MI}}\) is observed as inhomogeneous domain structure, huge anomaly in the high-frequency response, non-linear transport properties, and colossal magnetodielectricity.8,15,3638) It is clear from our study that the system has the strong πd interaction. Therefore, the metastable nature of the π-electron should affect the πd network, and slow magnetic transition should be observed around \(T_{\text{MI}}\). Actually, such slow dynamics of the magnetic transition is observed in Mössbauer measurement.34) Furthermore, the observation of the fine double peak structure of AFMR in Fig. 7(b) is due to two different environments of the Fe sites that are surrounded by either localized or itinerant π-electrons owing to its metastable state. We suppose the observation of excess specific heat below \(T_{\text{MI}}\) is due to the metastable state and slow magnetic transition in the AFI phase. Precise AFMR studies of the AFI phase are in progress, and the results are consistent with the above scenario. The results will be published elsewhere.39)

Acknowledgment

This work is partially supported by the Grant-in-Aid for Scientific Research (S) (No. 16H06346) and the Grant-in-Aid for Scientific Research (C) (No. 16K04882). Y.O. acknowledges H. Shimahara, N. Matsunaga, A. Kawamoto, and K. Hiraki for fruitful discussions.


References

  • 1 S. Uji, H. Shinagawa, T. Yakabe, Y. Terai, M. Tokumoto, A. Kobayashi, H. Tanaka, and H. Kobayashi, Nature 410, 908 (2001). 10.1038/35073531 CrossrefGoogle Scholar
  • 2 S. Uji and J. S. Brooks, J. Phys. Soc. Jpn. 75, 051014 (2006). 10.1143/JPSJ.75.051014 LinkGoogle Scholar
  • 3 H. Kobayashi, H. Tomita, T. Naito, A. Kobayashi, F. Sakai, T. Watanabe, and P. Cassoux, J. Am. Chem. Soc. 118, 368 (1996). 10.1021/ja9523350 CrossrefGoogle Scholar
  • 4 T. Mori and M. Katsuhara, J. Phys. Soc. Jpn. 71, 826 (2002). 10.1143/JPSJ.71.826 LinkGoogle Scholar
  • 5 V. Jaccarino and M. Peter, Phys. Rev. Lett. 9, 290 (1962). 10.1103/PhysRevLett.9.290 CrossrefGoogle Scholar
  • 6 S. Fujiyama, M. Takigawa, J. Kikuchi, H.-B. Cui, H. Fujiwara, and H. Kobayashi, Phys. Rev. Lett. 96, 217001 (2006). 10.1103/PhysRevLett.96.217001 CrossrefGoogle Scholar
  • 7 K. Hiraki, H. Mayaffre, M. Horvatic, C. Berthier, S. Uji, T. Yamaguchi, H. Tanaka, A. Kobayashi, H. Kobayashi, and T. Takahashi, J. Phys. Soc. Jpn. 76, 124708 (2007). 10.1143/JPSJ.76.124708 LinkGoogle Scholar
  • 8 Y. Oshima, H.-B. Cui, and R. Kato, Magnetochemistry 3, 10 (2017). 10.3390/magnetochemistry3010010 CrossrefGoogle Scholar
  • 9 L. Brossard, R. Clerac, C. Coulon, M. Tokumoto, T. Ziman, D. K. Petrov, V. N. Laukhin, M. J. Naughton, A. Audouard, F. Goze, A. Kobayashi, H. Kobayashi, and P. Cassoux, Eur. Phys. J. B 1, 439 (1998). 10.1007/s100510050207 CrossrefGoogle Scholar
  • 10 H. Akiba, S. Nakano, Y. Nishio, K. Kajita, B. Zhou, A. Kobayashi, and H. Kobayashi, J. Phys. Soc. Jpn. 78, 033601 (2009). 10.1143/JPSJ.78.033601 LinkGoogle Scholar
  • 11 S. Sugiura, K. Shimada, N. Tajima, Y. Nishio, T. Terashima, T. Isono, A. Kobayashi, B. Zhou, R. Kato, and S. Uji, J. Phys. Soc. Jpn. 85, 064703 (2016). 10.7566/JPSJ.85.064703 LinkGoogle Scholar
  • 12 S. Sugiura, K. Shimada, N. Tajima, Y. Nishio, T. Terashima, T. Isono, R. Kato, and S. Uji, J. Phys. Soc. Jpn. 86, 014702 (2017). 10.7566/JPSJ.86.014702 LinkGoogle Scholar
  • 13 S. Sugiura, K. Shimada, N. Tajima, Y. Nishio, T. Terashima, T. Isono, R. Kato, B. Zhou, and S. Uji, J. Phys. Soc. Jpn. 87, 044601 (2018). 10.7566/JPSJ.87.044601 LinkGoogle Scholar
  • 14 Y. Oshima, H. Nojiri, S. Uji, J. S. Brooks, T. Tokumoto, H.-B. Cui, R. Kato, A. Kobayashi, and H. Kobayashi, Phys. Rev. B 86, 024525 (2012). 10.1103/PhysRevB.86.024525 CrossrefGoogle Scholar
  • 15 I. Rutel, S. Okubo, J. S. Brooks, H. Kobayashi, A. Kobayashi, and T. Tanaka, Phys. Rev. B 68, 144435 (2003). 10.1103/PhysRevB.68.144435 CrossrefGoogle Scholar
  • 16 T. Suzuki, H. Matsui, H. Tsuchiya, E. Negishi, K. Koyama, and N. Toyota, Phys. Rev. B 67, 020408(R) (2003). 10.1103/PhysRevB.67.020408 CrossrefGoogle Scholar
  • 17 H. Akiba, K. Nobori, K. Shimada, Y. Nishio, K. Kajita, B. Zhou, A. Kobayashi, and H. Kobayashi, J. Phys. Soc. Jpn. 80, 063601 (2011). 10.1143/JPSJ.80.063601 LinkGoogle Scholar
  • 18 H. Akiba, H. Sugawara, K. Nobori, K. Shimada, N. Tajima, Y. Nishio, K. Kajita, B. Zhou, A. Kobayashi, and H. Kobayashi, J. Phys. Soc. Jpn. 81, 053601 (2012). 10.1143/JPSJ.81.053601 LinkGoogle Scholar
  • 19 K. Shimada, H. Akiba, N. Tajima, K. Kajita, Y. Nishio, R. Kato, A. Kobayashi, and H. Kobayashi, JPS Conf. Proc. 1, 012110 (2014). 10.7566/JPSCP.1.012110 LinkGoogle Scholar
  • 20 K. Shimada, N. Tajima, K. Kajita, and Y. Nishio, J. Phys. Soc. Jpn. 85, 023601 (2016). 10.7566/JPSJ.85.023601 LinkGoogle Scholar
  • 21 S. Kawamata, T. Kizawa, T. Suzuki, E. Negishi, H. Matsui, N. Toyota, and T. Ishida, J. Phys. Soc. Jpn. 75, 104715 (2006). 10.1143/JPSJ.75.104715 LinkGoogle Scholar
  • 22 T. Sasaki, H. Uozaki, S. Endo, and N. Toyota, Synth. Met. 120, 759 (2001). 10.1016/S0379-6779(00)00775-X CrossrefGoogle Scholar
  • 23 G. M. Sheldrick, Acta Crystallogr., Sect. A A71, 3 (2015). 10.1107/S2053273314026370 CrossrefGoogle Scholar
  • 24 G. M. Sheldrick, Acta Crystallogr., Sect. C C71, 3 (2015). 10.1107/S2053229614024218 CrossrefGoogle Scholar
  • 25 A. Altomare, M. Burla, M. Camalli, G. Cascarano, C. Giacovazzo, A. Guagliardi, A. Moliterni, G. Polidori, and R. Spagna, J. Appl. Crystallogr. 32, 115 (1999). 10.1107/S0021889898007717 CrossrefGoogle Scholar
  • 26 G. M. Sheldrick, SHELXL97. Program for the Refinement of Crystal Structures (University of Göttingen, Germany, 1997). Google Scholar
  • 27 A. Abragam and B. Bleaney, Electron Paramagnetic Resonance of Transition Ions (Dover, New York, 1986). Google Scholar
  • 28 A. Bencini and D. Gatteschi, Electron Paramagnetic Resonance of Exchange Coupled Systems (Springer, Berlin/Heidelberg, 1990). CrossrefGoogle Scholar
  • 29 (Supplemental Material) The typical EPR spectra and the integrated intensity of EPR are provided online. Google Scholar
  • 30 J. Owen and J. H. M. Thornley, Rep. Prog. Phys. 29, 675 (1966). 10.1088/0034-4885/29/2/306 CrossrefGoogle Scholar
  • 31 B. C. Tofield, J. Phys. Colloq. 37, C6-539 (1976). 10.1051/jphyscol:19766115 CrossrefGoogle Scholar
  • 32 J. C. Deaton, M. S. Gebhard, and E. I. Solomon, Inorg. Chem. 28, 877 (1989). 10.1021/ic00304a016 CrossrefGoogle Scholar
  • 33 F. Neese and E. I. Solomon, Inorg. Chem. 37, 6568 (1998). 10.1021/ic980948i CrossrefGoogle Scholar
  • 34 J. C. Waerenborgh, S. Rabaça, M. Almeida, E. B. Lopes, A. Kobayashi, B. Zhou, and J. S. Brooks, Phys. Rev. B 81, 060413(R) (2010). 10.1103/PhysRevB.81.060413 CrossrefGoogle Scholar
  • 35 E. Negishi, H. Uozaki, Y. Ishizaki, H. Tsuchiya, S. Endo, Y. Abe, H. Matsui, and N. Toyota, Synth. Met. 133–134, 555 (2003). 10.1016/S0379-6779(02)00344-2 CrossrefGoogle Scholar
  • 36 S. Komiyama, M. Watanabe, Y. Noda, E. Negishi, and N. Toyota, J. Phys. Soc. Jpn. 73, 2385 (2004). 10.1143/JPSJ.73.2385 LinkGoogle Scholar
  • 37 N. Toyota, Y. Abe, H. Matsui, E. Negishi, Y. Ishizaki, H. Tsuchiya, and H. Uozaki, Phys. Rev. B 66, 033201 (2002). 10.1103/PhysRevB.66.033201 CrossrefGoogle Scholar
  • 38 E. Negishi, T. Kuwabara, S. Komiyama, M. Watanabe, Y. Noda, T. Mori, H. Matsui, and N. Toyota, Phys. Rev. B 71, 012416 (2005). 10.1103/PhysRevB.71.012416 CrossrefGoogle Scholar
  • 39 T. H. Lee, Y. Oshima, H.-B. Cui, and R. Kato, unpublished. Google Scholar

Cited by

View all 10 citing articles

full accessEffects of Anion Substitution on Unconventional Antiferromagnetic Ground State of Molecular πd System λ-(BEDT-STF)2FexGa1−xCl4

043702, 10.7566/JPSJ.90.043702

full accessThermodynamic Investigation on Antiferromagnetic Ordered State of the Molecular πd System λ-(BEDT-STF)2FeCl4

073704, 10.7566/JPSJ.89.073704

full accessScaling Theory for Critical Phenomena in Coupled Magnetic Systems —Critical Exponents of πd Antiferromagnets—

043001, 10.7566/JPSJ.89.043001

full accessCritical Phenomena in Coupled Magnetic Systems —Application to the Organic Antiferromagnet λ-(BETS)2FeCl4

043001, 10.7566/JPSJ.88.043001