Next Article in Journal / Special Issue
Equivalence of Electron-Vibration Interaction and Charge-Induced Force Variations: A New O(1) Approach to an Old Problem
Previous Article in Journal
Synthesis and Crystal Structures of the Quaternary Zintl Phases RbNa8Ga3Pn6 (Pn = P, As) and Na10NbGaAs6
Previous Article in Special Issue
A New BEDT-TTF-Based Organic Charge Transfer Salt with a New Anionic Strong Acceptor, N,N'-Disulfo-1,4-benzoquinonediimine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Properties of Mn2+ and Π-Electron Spin Systems Probed by 1H and 13C NMR in the Organic Conductor κ-(BETS)2Mn[N(CN)2]3

1
Institute of Solid State Physics, Russian Academy of Sciences, Chernogolovka, Moscow region, 142432, Russia
2
Institute of Physical and Chemical Research (RIKEN), Hirosawa, Wako, Saitama 351-0198, Japan
3
Institute of Problems of Chemical Physics, Russian Academy of Sciences, Chernogolovka, Moscow region, 142432, Russia
*
Author to whom correspondence should be addressed.
Crystals 2012, 2(2), 224-235; https://doi.org/10.3390/cryst2020224
Submission received: 7 March 2012 / Revised: 30 March 2012 / Accepted: 31 March 2012 / Published: 12 April 2012
(This article belongs to the Special Issue Molecular Conductors)

Abstract

:
Properties of the spin systems of the localized 3d Mn2+ ions and the conduction π electrons in quasi-two-dimensional organic conductor κ-(BETS)2Mn[N(CN)2]3 were accessed using 1H and 13C NMR in order to find their relation to the metal-insulator transition which occurs at 23 K. The transition of the system into the insulating state is shown to be followed by localization of the π spins into a long-range ordered staggered structure of AF type. In contrast, the 3d Mn2+ electron spin moments form a disordered tilted structure, which may signify their trend to AF order, frustrated geometrically by the triangular arrangement of Mn in the anion layer. This result suggests that the MI transition in κ-(BETS)2Mn[N(CN)2]3 is not the consequence of the interactions within the Mn2+ spins but due to the interactions within the π-electron system itself. Vice versa, it is more likely that the disordered tilted structure of the Mn2+ spins is induced by the ordered π-spins via the π-d interaction.
PACS Classification:
74.70.Kn; 71.30.+h; 76.60.Jx; 75.30.-m

1. Introduction

NMR is a versatile tool for local-level investigation of magnetic properties in complex molecular systems. This technique uses nuclear spins to spy on the electron spin system by monitoring the local magnetic field created by electron spins at the nucleus site. Since NMR frequency is nucleus-specific, the experiment tuned to particular nucleus probes the electron spin system at the site of the corresponding atom within the crystal structure. In complex compounds incorporating more than one electron spin system (e.g., itinerant and localized electrons) such information helps to discriminate between the events developing within the electron subsystems and determine the interrelation between them.
Figure 1. (a) Side view (along Crystals 02 00224 i001) of crystal structure of κ-(BETS)2Mn[N(CN)2]3; (b) BETS molecule with 13C in positions of the central C = C bond; (c) Top view of schematic crystal structure of the conducting BETS layer. The circled pairs of lines represent (BETS)2 dimers; (d) Side view of BETS dimer with site definition of the central C = C carbons; (e) Top view of the anion layer.
Figure 1. (a) Side view (along Crystals 02 00224 i001) of crystal structure of κ-(BETS)2Mn[N(CN)2]3; (b) BETS molecule with 13C in positions of the central C = C bond; (c) Top view of schematic crystal structure of the conducting BETS layer. The circled pairs of lines represent (BETS)2 dimers; (d) Side view of BETS dimer with site definition of the central C = C carbons; (e) Top view of the anion layer.
Crystals 02 00224 g001
The quasi-two-dimensional organic conductor κ-(BETS)2Mn[N(CN)2]3 (BETS = C10S4Se4H8, bis-(ethylenedithio)tetraselenafulvalene) is an example of such compound (Figure 1). The system is metallic down to TMI ~ 23K where it undergoes the metal-insulator (MI) transition [1,2]. Its conductivity is associated with π-electrons confined within the layers of dimerised organic BETS molecules, sandwiched between the insulating polymeric Mn[N(CN)2]3 anion layers, while the bulk magnetization is determined by Mn2+ (J = S = 5/2, L = 0) ions of the anion layers [3]. The static susceptibility in the metallic state is isotropic and obeys accurately the Curie-Weiss law. Below TMI it becomes slightly smaller than the Curie-Weiss value, χcw (T), so that at T = 2K it makes 80%-85% of χCW, still essentially isotropic. This trend suggests that below TMI the Mn2+ system tends towards antiferromagnetic (AF) order, which is however frustrated geometrically by triangular arrangement of Mn in the anion layer (Figure 1e). In the akin compound λ-(BETS)2FeCl4 with rectangular network of FeCl4 units, the susceptibility governed by Fe3+ moments demonstrates below TMI = 8K the distinctive features characteristic of a uniaxial Néel-type antiferromagnet with TN = TMI, which gave grounds to nominate the AF ordering within Fe3+ ions as driving force of the MI transition [4]. One can assume therefore that the AF interaction within the Mn2+ d-electron spins in the title compound may in a similar way cause localization of the conduction π electrons at TMI due to strong π-d coupling presumably present in the system [5].
On the other hand, the MI transition is not only specific to quasi-two-dimensional organic conductors with magnetic anions: there are known non-magnetic compounds where the insulating state is attributed to strong electron correlations within the half-filled conducting band leading to Mott instability and electron localization [6]. Moreover, in some of them the electrons localize into the long-range AF structure (e.g., κ-(ET)2Cu[N(CN)2]Cl [7], β΄-(ET)2X (X = ICl2, AuCl2) [8], where ET is isostructural to BETS but with sulphur on Se sites). Besides, recent experiments on λ-(BETS)2FeCl4 [9,10] have doubted the AF order in the Fe3+ spin system and its responsibility for the MI transition, stating that it is the π-electron system that orders antiferromagnetically and produces the MI transition, while the Fe3+ d-electrons remain in the paramagnetic state.
There is therefore a question if the magnetic interactions within Mn2+ moments in κ-(BETS)2Mn[N(CN)2]3 are responsible for the MI transition, or it is a domestic affair of the conduction electrons as in some non-magnetic quasi-two-dimensional systems. To address this issue we have performed 1H and 13C NMR measurements on κ-(BETS)2Mn[N(CN)2]3 single crystals. For 13C NMR, the crystal with BETS molecules containing >99% of 13C isotope in the central C = C bond (Figure 1b) was used. Hydrogen atoms belonging to the ethylene groups at the terminals of BETS molecules (Figure 1a) are relatively close to the Mn sites and therefore are expected to provide information about the Mn2+ subsystem through dipolar fields induced by Mn2+ moments at hydrogen sites. 13C NMR on the carbons from the central C = C bond of the BETS molecule is known to be effective in probing the conduction π-electron spin system [11]. The two experiments combined together promise the complete description of the magnetic properties of the system.

2 Results and Discussion

2.1. 1H NMR

Hydrogen sites located at the terminals of BETS molecules (Figure 1a) have negligibly small hyperfine coupling with conduction electrons [12], therefore the NMR frequency shift should be determined by the dipolar fields from the 3d Mn2+ ion electron spin moments (Sd = 5/2,g ≈ 2). For the arbitrary orientation of the magnetic field, the 1H NMR spectrum will count 16 peaks: there are 8 inequivalent crystallographic hydrogen sites and two magnetically inequivalent orientations of the BETS dimer (Figure 1c). The BETS molecules within the dimer are inversion-symmetric to each other, thus magnetically equivalent.
Figure 2 shows 1H NMR spectrum in κ-(BETS)2Mn[N(CN)2]3 at T = 74 K in field H0 = 7 T oriented at θ = 22°from Crystals 02 00224 i017towards Crystals 02 00224 i018 direction. The spectrum is shown with respect to 1v0 = 1γH0, where 1γ = 42.5759 MHz/T is the proton gyromagnetic ratio. For the chosen geometry Crystals 02 00224 i022, the two different orientations of the BETS dimer shown in Figure 1c become magnetically equivalent which reduces the total number of peaks to 8. Arrows in Figure 2 indicate the 8 peaks.
Figure 2. 1H NMR spectrum in κ-(BETS)2Mn[N(CN)2]3 at T = 74 K in field H0 = 7 T oriented at θ = 22°from Crystals 02 00224 i017towards Crystals 02 00224 i023 Crystals 02 00224 i024direction. The spectrum is shown with respect to 1v0 = 1γH0 .
Figure 2. 1H NMR spectrum in κ-(BETS)2Mn[N(CN)2]3 at T = 74 K in field H0 = 7 T oriented at θ = 22°from Crystals 02 00224 i017towards Crystals 02 00224 i023 Crystals 02 00224 i024direction. The spectrum is shown with respect to 1v0 = 1γH0 .
Crystals 02 00224 g002
Figure 3 presents the angular evolution of 1H NMR peak positions in κ-(BETS)2Mn[N(CN)2]3 measured at 74 K in field H0=7 T. The field is in the (a*c)-plane, and the polar angle, θ, is reckoned from Crystals 02 00224 i017 direction towards Crystals 02 00224 i023.
Figure 3. Angular evolution of 1H NMR spectrum in the (a*c) plane measured at 74 K in field H0=7 T. Circles are the measured peak positions. Lines are model calculations using Equation 1.
Figure 3. Angular evolution of 1H NMR spectrum in the (a*c) plane measured at 74 K in field H0=7 T. Circles are the measured peak positions. Lines are model calculations using Equation 1.
Crystals 02 00224 g003
To model the observed proton spectrum one needs to sum up the dipolar fields, hdip, created at the nucleus site by Mn2+ electronic spins, and take into account the sample geometry resulting in a demagnetizing field, hD, and the Lorentz field, hLor, which is the mean field induced at the nucleus site by the bulk of the sample located outside the dipolar summation sphere [13].
For the magnetic field in the (a*c)-plane, Crystals 02 00224 i031we model the spectrum as
Crystals 02 00224 i032
Crystals 02 00224 i033
Crystals 02 00224 i034
Here, γn = 1γ; μMn is the thermal average of the Mn2+ magnetic moment projection on the field direction, riis the length of the position vector from the proton site to the Mn site i, αi is the angle between this vector and the field direction, VMn is the unit cell volume per Mn2+ ion, and N = N cos2θ + Nsin2θ is the demagnetization factor.
Since the DC magnetization in κ-(BETS)2Mn[N(CN)2]3 is determined by Mn2+ magnetic moments [3], we put in Equation 3 μMn = M/N A, where M is the measured DC magnetization per mole (3800 emu/mol at H0= 7 T, T = 74 K) and NA is Avogadro’s number. For our very thin-plate sample we assume the demagnetization factors N = 1, N = 0. Crystallographic positions of Mn and H atoms required to calculate Equation 1b are available online from the CCDC library as mentioned in the Experimental Section. The sum in Equation 1a has been taken over ~ 200 Mn sites within ±20 Å to provide reasonable convergence.
The peak positions calculated using Equation 1 for each of the 8 inequivalent hydrogen sites are shown in Figure 3 by solid lines. The agreement between the calculated and the measured peak positions is clearly reasonable despite the absence of any fit parameters used in the calculations. This indicates that 1H NMR frequency shift is determined by the dipolar fields from Mn2+ at the hydrogen site, at least at this field and temperature.
To check if this is true for all temperatures, we compare the temperature dependences of the DC magnetization and the 1H NMR peak positions. Figure 4a,b show temperature dependences of, respectively, the molar DC magnetization normalized to the applied field, M/N, and the normalized frequency shift of the lowest-frequency peak in the spectrum measured in Ha* geometry (θ = 0 in Figure 3), for the magnetic fields 1.4 and 7 T.
Figure 4. (a) Molar magnetization vs. temperature in H = 1.4 and 7 T; (b) Position of the lowest-frequency peak in 1H NMR spectrum in Crystals 02 00224 i053geometry vs. temperature in H = 1.4 and 7 T; (c) Position of the peak for the temperature range 4-150 K, in function of the DC magnetization per Mn ion, μMn = M/NA, measured at the same temperatures and fields and expressed in terms of the Bohr magneton, μB. Open red circles and closed black circles correspond to the fields 1.4 and 7 T, respectively.
Figure 4. (a) Molar magnetization vs. temperature in H = 1.4 and 7 T; (b) Position of the lowest-frequency peak in 1H NMR spectrum in Crystals 02 00224 i053geometry vs. temperature in H = 1.4 and 7 T; (c) Position of the peak for the temperature range 4-150 K, in function of the DC magnetization per Mn ion, μMn = M/NA, measured at the same temperatures and fields and expressed in terms of the Bohr magneton, μB. Open red circles and closed black circles correspond to the fields 1.4 and 7 T, respectively.
Crystals 02 00224 g004
The right panel of Figure 4 depicts the plot of the proton frequency shift, v - 1v0, in function of the magnetization measured at the same temperatures and fields. The linearity of the data in Figure 4 demonstrates that for the whole temperature and field ranges covered in the experiment, the positions of 1H NMR spectrum peaks are determined by the magnetic subsystem associated with Mn2+ moments, and can be utilized as Hall probes of the Mn2+ dipolar fields at hydrogen sites.
Figure 5a shows the evolution of 1H NMR spectrum (H0 = 1.4T parallel to Crystals 02 00224 i017) with temperature. As can be seen, the frequency span of the spectrum increases with decreasing temperature while its shape is maintained down to ~20 K. At lower temperature the peaks broaden rapidly, which is more pronounced on the right-hand side of the spectrum. Figure 5b depicts the temperature dependence of the half width at half hight, Crystals 02 00224 i059, for the leftmost and the rightmost peaks in the spectrum. As one can see in Figure 5, the linewidth is relatively flat above TMI ≈ 23 K (especially for the leftmost peak) and increases sharply below this temperature. Figure 5c demonstrates the plot of the linewidths in function of the DC magnetization. One can see in Figure 5c a crossover in the linewidth behavior at TMI which is observed as an upturn from the linear M-dependence (shown by solid lines) obeyed at higher temperatures.
Figure 5. (a) Temperature evolution of 1H NMR spectrum measured at H0 = 1.4T parallel to Crystals 02 00224 i017; (b) Temperature dependence of the half-linewidth of the leftmost (circles) and the rightmost (squares) peaks; (c) The leftmost and the rightmost peak half-linewidths in function of the DC magnetization. Solid lines extrapolate the linewidth behavior above 25 K to low temperatures. Dashed lines with numbers on top mark measurement temperatures.
Figure 5. (a) Temperature evolution of 1H NMR spectrum measured at H0 = 1.4T parallel to Crystals 02 00224 i017; (b) Temperature dependence of the half-linewidth of the leftmost (circles) and the rightmost (squares) peaks; (c) The leftmost and the rightmost peak half-linewidths in function of the DC magnetization. Solid lines extrapolate the linewidth behavior above 25 K to low temperatures. Dashed lines with numbers on top mark measurement temperatures.
Crystals 02 00224 g005
Since the 1H NMR peak positions are highly anisotropic as can be seen in Figure 3, there is a number of trivial reasons for the peaks to be broadened, including sample imperfections on the macroscopic and local levels, as well as minor misalignment of the magnetic field from Crystals 02 00224 i017 direction, as discussed in [14]. However, all of them should make the linewidth as linear in the magnetization as the peak position itself is (Figure 4c). The upturn of the linewidth M-dependence below TMI ≈ 23 K (Figure 5c) indicates the onset of yet another broadening mechanism. Provided that the crystal lattice is intact, this should be related with local-level scatter of the dipolar fields from Mn2+ at the hydrogen site, for example, a variable from site to site tilt of the Mn2+ static moments. This site variation of the tilt is apparently random or incommensurate with the crystal lattice, because otherwise a splitting of 1H NMR peaks would take place instead of the featureless broadening observed in the experiment (Figure 5a). The tilt of the Mn2+ moments emerging below TMI, together with the deviation of the magnetization from the Curie-Weiss behavior [3], indicates apparently the tendency of the Mn2+ spin system towards the AF order. Frustrated geometrically by triangular arrangement of Mn in the anion layer (Figure 1e), Mn2+ system is resolved into a disordered (like spin-glass) or an incommensurate spin structure.

2.2. 13C NMR

The 13C NMR spectra were measured in κ-(BETS)2Mn[N(CN)2]3 in the external field H = 7 T aligned perpendicular to Crystals 02 00224 i001 at 45°to Crystals 02 00224 i017. For this experimental geometry, the metallic-state spectrum in theory is represented by four peaks arising from the two magnetically different orientations of the BETS dimers (Figure 1c) and nonequivalent(“inner” and “outer”) carbon sites of the central C=C bond within the dimer (Figure 1d) [15]. The dipolar interaction between 13C spins in the C=C bond, which in general provides another factor of two to the number of peaks, is nearly zero for this field orientation.
Figure 6. 13C NMR spectra measured in the external field H = 7 T aligned perpendicular to Crystals 02 00224 i001 at 45° to Crystals 02 00224 i017. The spectra are shown with respect to 13v0 = 74.946 MHz. (a) The spectrum at T = 50 K. Blue solid lines indicate positions of the resonance peaks calculated with the shift tensor of κ-(ET)2Cu[N(CN)2]Br [15]; (b) The evolution of the 13C NMR spectrum with temperature.
Figure 6. 13C NMR spectra measured in the external field H = 7 T aligned perpendicular to Crystals 02 00224 i001 at 45° to Crystals 02 00224 i017. The spectra are shown with respect to 13v0 = 74.946 MHz. (a) The spectrum at T = 50 K. Blue solid lines indicate positions of the resonance peaks calculated with the shift tensor of κ-(ET)2Cu[N(CN)2]Br [15]; (b) The evolution of the 13C NMR spectrum with temperature.
Crystals 02 00224 g006
Unlike the magnetic ion-free quasi-two-dimensional ET-based conductors where the NMR peaks from magnetically nonequivalent 13C sites are usually well resolved, 13C spectrum in the metallic state of κ-(BETS)2Mn[N(CN)2]3 for this field orientation is represented by a single featureless Gaussian-shaped line. The left panel in Figure 6 represents the spectrum taken at T = 50 K. Using the 13C NMR shift tensors obtained for κ-(ET)2Cu[N(CN)2]Br [15] (which is nearly the same as for κ-(ET)2Cu[N(CN)2]Cl [16]), for the given experimental geometry one expects a κ-(BETS)2Mn[N(CN)2]3 sample to produce the resonance peaks at 2.7, 10.7, 22.7, and 40.8 kHz (with respect to 13v0 = 74.946 MHz). Dipolar fields from Mn2+ moments provide additional shifts to the resonance frequencies, which can be calculated following Equations 3 using the values of Mn2+ moments known from the bulk magnetization measurements [3]. At T = 50 K the calculated dipolar fields from Mn2+ move the listed peaks to positions at 0.3, 7.7, 19, and 38.2 kHz. The calculated peak positions are shown by vertical lines in Figure 6a. One can see that the spectrum measured at 50 K covers fairly well the range of the calculated peak positions indicating that the 13C shift tensor in the title compound is not much different from that in κ-(ET)2Cu[N(CN)2]Br.
Evidently, the peaks from individual carbon sites in the metallic state of κ-(BETS)2Mn[N(CN)2]3 are vastly broadened that merges them into a single line. Some broadening of 13C NMR peaks has been noticed in κ-(ET)2Cu[N(CN)2]Br and a number of reasons has been recruited to explain it [15], including spacial variation of the π-electron spin density due to precursors to Anderson localization or spin-density wave. For the sample reported here, the linewidth is expected broader than in non-magnetic compounds due to the presence of the anisotropic dipolar fields from Mn2+. For example, the same mechanism that creates the 1H NMR linewidth (Figure 5b) will be responsible for ~10 kHz linewidth of the 13C peaks at carbon sites at 50 K. Anyway, more detailed analysis of the possible broadening mechanisms in the metallic state is beyond the scope of this communication.
The right panel in Figure 6 demonstrates the evolution of the 13C NMR spectrum with temperature. The single peak characteristic of the spectrum in the metallic state above T = 23K ≈ TMI develops below this temperature into a broad symmetric pattern counting 5 pronounced peaks. Below T = 15 K the spectrum spans the range of nearly ± 1 MHz, which is huge compared to the spectrum width in the metallic state. This cannot result from the dipolar fields created by Mn2+: calculations show that fully polarized Mn2+ can provide a dipolar shift ranging from -12.5 to -19 kHz (depending on the carbon site) in 7 T field. Therefore the spectrum in the insulating state evidences enhancement of the electron spin density at carbon sites due to electron spins localized on the dimers of the BETS molecules. Moreover, several pronounced peaks are visible in the low-temperature spectrum, which infers a commensurate order of the localized spins. Finally, the symmetric shape of the spectrum indicates the staggered order, because antiparallel components of the staggered electron spins ( Crystals 02 00224 i069) produce opposite local fields at carbon sites i and j: Crystals 02 00224 i070, where Crystals 02 00224 i071 and Crystals 02 00224 i072 are the nuclear and electron spin operators, respectively, and A is the hyperfine tensor. In turn, this signifies the AF exchange interaction between the localized spins. The staggered component of the spin magnetization lies apparently somewhere in the plane perpendicular to the magnetic field, since the anisotropic AF exchange term is usually much smaller than 7 T, the external field of this experiment.
The frequency range of the 13C spectrum in κ-(BETS)2Mn[N(CN)2]3 at T = 5 K (Figure 6b) is of the same order of magnitude (~±1 MHz) as observed in the AF state of κ-(ET)2Cu[N(CN)2]Cl [16] and β΄-(ET)2ICl2 [17]. The magnitude of the electron spin magnetization of 0.5 μB and 1 μB per dimer, respectively, has been reported for these two compounds. Therefore in κ-(BETS)2Mn[N(CN)2]3 this value should be within the same range since the hyperfine tensor here is expected to be similar, as the measurements in the metallic phase suggest.
More detailed and quantitative information about the spin structure of the localized πelectrons can hardly be derived at the moment. To do this one needs to know more or less exactly the hyperfine tensors for the central carbons in κ-(BETS)2 Mn[N(CN)2]3, which in turn requires the value of the πspin susceptibility never reported so far (but probably accessible via ESR).

3 Experimental

The crystal structure of κ-(BETS)2Mn[N(CN)2]3 is monoclinic with the space group P21/c and the lattice constants at 88 K a = 19.428 Å, b = 8.379 Å, c = 11.869 Å, β = 92.67°, and V = 1930.1 Å3, with two formula units per unit cell. The conducting layers formed by BETS dimers in the (bc) plane are sandwiched between the polymeric Mn[N(CN)2]3 anion layers in the a direction (Figure 1a). The crystal structure is available at Cambridge Crystallographic Data Centre [18], CCDC 775974-775977, and has been discussed in details in earlier communications [1,2].
Crystals for 1H NMR (with all natural-abundant carbon) were obtained by electrochemical oxidation of BETS in the mixture of solvents, 1,1,2-trichloroethane/ethanol (10:1 v/v), in the presence of Mn[N(CN)2]2 as electrolyte [1,2]. The 13C-enriched samples were synthesized by electrochemical oxidation of 13C-labeled BETS in benzonitrile/ethanol (10:1 v/v) in the presence of Ph4PMn[N(CN)2]3 complex salt as electrolyte following the procedure reported in Reference [19]. 13C-enriched (>99%) BETS (Figure 1b) was synthesized with the use of 13C-enriched triphosgene (bis(trichloromethyl-13C3)carbonate) according to the method described in literature [20].
The single crystals used in the 1H NMR experiments had the dimensions a* × b × c ~ 0.05 × 3 × 1 mm3. The dimensions of the crystal for 13C NMR were ~0.05 × 3 × 3 mm3. Crystallographic orientations of the crystals were X-ray defined. NMR spectra were acquired using standard spin-echo sequence with π-pulse length ≤2.5-3 μs. To cover broad spectra, Fourier-transforms of the acquired spin-echoes were collected at 150 kHz intervals and summed up.

4 Conclusions

We performed 1H and 13C NMR measurements in κ-(BETS)2Mn[N(CN)2]3 to find the relation between the MI transition and the properties of the spin systems of the localized 3d Mn2+ ions and the conduction πelectrons. We found that transition of the system into the insulating state is followed by localization of the πspins into a long-range ordered staggered structure of AF type. It should be emphasized here that this structure does not necessarily signify the conventional AF state but can be a field-induced effect in the presence of the Dzyaloshinskii-Moriya interaction, as it happens in κ-(ET)2Cu[N(CN)2]Cl [21].
In contrast, the Mn2+ spins do not order below TMI but are likely to form a disordered tilted structure, which may signify their trend to AF order, frustrated geometrically by the triangular arrangement of Mn in the anion layer.
Our findings show that the MI transition in κ-(BETS)2Mn[N(CN)2]3 is not the consequence of the interactions within the 3d Mn2+ electron spin system transferred to the conduction spin system by the π-d coupling, but is a result of the interactions within the π-electron system itself as in some magnetic ion-free quasi-two-dimensional charge-transfer salts. Vice versa, the observed formation of the disordered tilted structure in the Mn2+ electron spin system could be induced by the π-spin ordering via the π-d interaction. It has been reported in [3] that the susceptibility of κ-(BETS)2Mn[N(CN)2]3 determined by the Mn2+ spin system depends on temperature as 1/ (T+θ)with θ ≈ 5.5 K. That means that the Mn2+ subsystem should remain paramagnetic down to T~θ, while in fact it deviates from the paramagnetic state at higher temperature TMI ~23 K, suggesting the influence of the spin-ordered localization of π-electrons.

Acknowledgements

The authors gratefully acknowledge fruitful suggestions from S. E. Brown, assistance in the questions of crystallography from S. S. Khasanov, and technical support from N. A. Belov. This work was supported by the RFBR grants 10-02-01202 and 11-02-91338-DFG (DFG grant Bi 340/3-1), Russian State Contract 14.740.11.0911, and the Program of Russian Academy of Sciences. This work was partially supported by Grant-in-Aid for Scientific Research (S) (No. 22224006) from the Japan Society for the Promotion of Science (JSPS).

References

  1. Zverev, V.N.; Kartsovnik, M.V.; Biberacher, W.; Khasanov, S.S.; Shibaeva, R.P.; Ouahab, L.; Toupet, L.; Kushch, N.D.; Yagubskii, E.B.; Canadell, E. Temperature-pressure phase diagram and electronic properties of the organic metal κ-(BETS)2Mn[N(CN)2]3[N(CN)2]3. Phys. Rev. B 2010, 82, 155123:1–155123:9. [Google Scholar]
  2. Kushch, N.D.; Yagubskii, E.B.; Kartsovnik, M.V.; Buravov, L.I.; Dubrovskii, A.D.; Chekhlov, A.N.; Biberacher, W. Donor BETS based bifunctional superconductor with polymeric dicyanamidomanganate(II) anion layer: κ-(BETS)2Mn[N(CN)2]3[N(CN)2]3. J. Am. Chem. Soc. 2008, 130, 7238–7240. [Google Scholar]
  3. Vyaselev, O.M.; Kartsovnik, M.V.; Biberacher, W.; Zorina, L.V.; Kushch, N.D.; Yagubskii, E.B. Magnetic transformations in the organic conductor κ-(BETS)2Mn[N(CN)2]3[N(CN)2]3at the metal-insulator transition. Phys. Rev. B 2011, 83, 094425:1–094425:6. [Google Scholar]
  4. Brossard, L.; Clerac, R.; Coulon, C.; Tokumoto, M.; Ziman, T.; Petrov, D.K.; Laukhin, V.N.; Naughton, M.J.; Audouard, A.; Goze, F.; et al. Interplay between chains of S = 5/2localised spins and two-dimensional sheets of organic donors in the synthetically built magnetic multilayer λ-(BETS)2FeCl4. Eur. Phys. J. B 1998, 1, 439–452. [Google Scholar] [CrossRef]
  5. Uji, S.; Brooks, J.S. Magnetic-field-induced superconductivity in organic conductors. J. Phys. Soc. Jpn. 2006, 75, 051014:1–051014:10. [Google Scholar]
  6. Kagawa, F.; Miyagawa, K.; Kanoda, K. Magnetic Mott criticality in a κ-type organic salt probed by NMR. Nat. Phys. 2009, 5, 880–884. [Google Scholar] [CrossRef]
  7. Welp, U.; Fleshler, S.; Kwok, W.K.; Crabtree, G.W.; Carlson, K.D.; Wang, H.H.; Geiser, U.; Williams, J.M.; Hitsman, V.M. Weak ferromagnetism in κ-(ET)2Cu[N(CN)2]Cl, where (ET) is bis-(ethylenedithio)tetrathiafulvalene. Phys. Rev. Lett. 1992, 69, 840–843. [Google Scholar] [CrossRef]
  8. Yoneyama, N.; Miyazaki, A.; Enoki, T.; Saito, G. Magnetic properties of (BEDT-TTF)2X with localized spins. Synth. Met. 1997, 86, 2029–2030. [Google Scholar] [CrossRef]
  9. Akiba, H.; Nakano, S.; Nishio, Y.; Kajita, K.; Zhou, B.; Kobayashi, A.; Kobayashi, H. Mysterious paramagnetic states of Fe 3d spin in antiferromagnetic insulator of κ-BETS2FeCl4system. J. Phys. Soc. Jpn. 2009, 78, 033601:1–033601:3. [Google Scholar]
  10. Waerenborgh, J.C.; Rabaça, S.; Almeida, M.; Lopes, E.B.; Kobayashi, A.; Zhou, B.; Brooks, J.S. Mössbauer spectroscopy and magnetic transition of κ-(BETS)2FeCl4. Phys. Rev. B 2010, 81, 060413:1–060413:4. [Google Scholar]
  11. Klutz, T.; Hennig, I.; Haeberlen, U.; Schweitzer, D. Knight shift tensors and π-spin densities in the organic metals αt-(BEDT-TTF)2I3and (BEDT-TTF)2Cu(NCS)2. Appl. Mag. Res. 1991, 2, 441–463. [Google Scholar] [CrossRef]
  12. Kanoda, K. Recent progress in NMR studies on organic conductors. Hyperfine Interact. 1997, 104, 235–249. [Google Scholar] [CrossRef]
  13. Uritskii, M.; Irkhin, V. Calculation of dipole fields at interstices of rare-earth metals. Phys. Solid State 2000, 42, 399–405, Fizika Tverdogo Tela 2000, 42, 390-396.. [Google Scholar] [CrossRef]
  14. Vyaselev, O.M.; Kushch, N.D.; Yagubskii, E.B. Proton NMR study of the organic metal κ-(BETS)2Mn[N(CN)2]3[N(CN)2]3. J. Exp. Theor. Phys. 2011, 113, 835–841, Zh. Eksp. Teor. Fiz. 2011, 140, 961-967.. [Google Scholar] [CrossRef]
  15. de Soto, S.M.; Slichter, C.P.; Kini, A.M.; Wang, H.H.; Geiser, U.; Williams, J.M. 13C NMR studies of the normal and superconducting states of the organic superconductor κ-(ET)2Cu[N(CN)2]Br. Phys. Rev. B 1995, 52, 10364–10368. [Google Scholar] [CrossRef]
  16. Smith, D.F.; de Soto, S.M.; Slichter, C.P.; Schlueter, J.A.; Kini, A.M.; Daugherty, R.G. Dzialoshinskii-Moriya interaction in the organic superconductor κ-(BEDT-TTF)2Cu(N[CN]2)Cl. Phys. Rev. B 2003, 68, 024512:1–024512:9. [Google Scholar]
  17. Eto, Y.; Kawamoto, A. Antiferromagnetic phase in β΄-(BEDT-TTF)2ICl2 under pressure as seen via 13C NMR. Phys. Rev. B 2010, 81, 020512:1–020512:4. [Google Scholar]
  18. Cambridge Crystallographic Data Centre. Available online: http://www.ccdc.cam.ac.uk/data_request/cif (accessed on 10 April 2012).
  19. Kushch, N.D.; Buravov, L.I.; Chekhlov, A.N.; Spitsina, N.G.; Kushch, P.P.; Yagubskii, E.B.; Herdtweck, E.; Kobayashi, A. The first BETS radical cation salts with dicyanamide anion: Crystal growth, structure and conductivity study. J. Solid State Chem. 2011, 184, 3074–3079. [Google Scholar]
  20. Kato, R.; Kobayashi, H.; Kobayashi, A. Synthesis and properties of bis(ethylenedithio) tetraselenafulvalene (BEDT-TSeF) compounds. Synth. Met. 1991, 42, 2093–2096. [Google Scholar] [CrossRef]
  21. Kagawa, F.; Kurosaki, Y.; Miyagawa, K.; Kanoda, K. Field-induced staggered magnetic moment in the quasi-two-dimensional organic Mott insulator κ-(BEDT-TTF)2Cu[N(CN)2]Cl. Phys. Rev. B 2008, 78, 184402–1. [Google Scholar]

Share and Cite

MDPI and ACS Style

Vyaselev, O.M.; Kato, R.; Yamamoto, H.M.; Kobayashi, M.; Zorina, L.V.; Simonov, S.V.; Kushch, N.D.; Yagubskii, E.B. Properties of Mn2+ and Π-Electron Spin Systems Probed by 1H and 13C NMR in the Organic Conductor κ-(BETS)2Mn[N(CN)2]3. Crystals 2012, 2, 224-235. https://doi.org/10.3390/cryst2020224

AMA Style

Vyaselev OM, Kato R, Yamamoto HM, Kobayashi M, Zorina LV, Simonov SV, Kushch ND, Yagubskii EB. Properties of Mn2+ and Π-Electron Spin Systems Probed by 1H and 13C NMR in the Organic Conductor κ-(BETS)2Mn[N(CN)2]3. Crystals. 2012; 2(2):224-235. https://doi.org/10.3390/cryst2020224

Chicago/Turabian Style

Vyaselev, Oleg M., Reizo Kato, Hiroshi M. Yamamoto, Megumi Kobayashi, Leokadiya V. Zorina, Sergey V. Simonov, Nataliya D. Kushch, and Eduard B. Yagubskii. 2012. "Properties of Mn2+ and Π-Electron Spin Systems Probed by 1H and 13C NMR in the Organic Conductor κ-(BETS)2Mn[N(CN)2]3" Crystals 2, no. 2: 224-235. https://doi.org/10.3390/cryst2020224

Article Metrics

Back to TopTop