Large Diamagnetism and Electromagnetic Duality in Two-dimensional Dirac Electron System
Abstract
A Dirac electron system in solids mimics relativistic quantum physics that is compatible with Maxwell’s equations, with which we anticipate unified electromagnetic responses. We find a large orbital diamagnetism only along the interplane direction and a nearly temperature-independent electrical conductivity of the order of per plane for the new 2D Dirac organic conductor, -(BETS)2I3, where BETS is bis(ethylenedithio)tetraselenafulvalene. Unlike conventional electrons in solids whose nonrelativistic effects bifurcate electric and magnetic responses, the observed orbital diamagnetism scales with the electrical conductivity in a wide temperature range. This demonstrates that an electromagnetic duality that is valid only within the relativistic framework is revived in solids.
Dirac electron systems (DESs) such as bismuth and graphene can be described by the Dirac equation and provide a platform to realize physical properties rooted in relativistic quantum physics [1, 2]. One prominent property of DESs is their large orbital diamagnetism, which reaches a maximum when the chemical potential is in the mass gap, unlike Landau diamagnetism in metals. This orbital diamagnetism, which is theoretically argued to originate from the interband effect of magnetic fields, is observed in three-dimensional (3D) DESs including bismuth and antiperovskites [3, 4, 5, 6]. This mechanism also applies to two-dimensional (2D) systems. The diamagnetism was observed for mass-produced graphene flakes [7]. Here the random orientation of the flakes prevented separating the orbital diamagnetism agreeable with theory. Organic conductors have recently been found to realize 2D DESs with a bulk form such as -(BEDT-TTF)2I3 (BEDT-TTF = bis(ethylene)dithiotetrathiafulvalene) [8, 9, 10]; however, this is realized only under high pressure, limiting magnetic experiments and making it difficult to obtain the absolute value of the susceptibility using SQUID magnetometers.
The electric responses of 3D and 2D DESs show a sharp contrast. The uniform permittivity of bismuth is enhanced in accordance with its orbital diamagnetism [11, 12]. On the other hand, graphene has no enhancement in the permittivity, but rather shows exotic quantized optical conductance and minimum dc conductivity through Klein tunneling [2, 13, 14]. The organic conductor -(BEDT-TTF)2I3 also shows temperature-independent conductivity on the order of per sheet [15].
These magnetic and electric responses of DESs can be viewed as parallel to quantum electrodynamics (QED), a relativistic quantum field theory, in which two responses are unified due to the Lorentz covariance (space–time symmetry). Indeed, for 3D DESs, the large orbital diamagnetism and the enhanced permittivity can be explained by charge renormalization in a unified way, demonstrating an electromagnetic duality specified by the space–time symmetry of the Dirac equation [16]. In contrast to 3D DESs, permittivity enhancement due to charge renormalization is absent in 2D DESs [17], although they do exhibit a quantized conductance. The dependence of the charge renormalization on the dimensionality of the system raises the fundamental question of the existence and nature of the universal phenomena in DESs irrespective of the system dimension. Therefore, the two principal goals of the study of 2D DESs are to determine the behavior of the orbital diamagnetism and its relationship with quantized electric responses, and to clarify whether the two responses can be described by a unified theory. Observation of the orbital diamagnetism would resolve these questions, and in order to obtain absolute values of the magnetic susceptibility, a bulk-form single crystal at ambient pressure would be ideal.

In this Letter, we demonstrate the magnetic and transport properties of a newly identified 2D DES organic conductor with a bulk form at ambient pressure, -(BETS)2I3 with a strongly anisotropic magnetic susceptibility . We discriminated a large orbital diamagnetism () from a spin susceptibility () by changing the field direction. shows quantitative agreement with the theory for K, where the dc-conductivity () per sheet is independent of temperature with the value of . The scales with in a wide temperature range, showing an electromagnetic duality specified by the space–time symmetry in DESs, corresponding to the Lorentz covariance in QED.
-(BETS)2I3 is composed of bis(ethylenedithio)tetraselenafulvalene (BETS) molecules that contain Se atoms (Fig. 1 (a)) [18, 19]. The structure is isomorphous with -(BEDT-TTF)2I3, as shown in Fig. 1 (b). The molecular orbital of BETS is spatially larger than that of BEDT-TTF, which yields uncorrelated electron characteristics and prevents the instabilities toward charge ordering or excitonic orders observed in -(BEDT-TTF)2I3 [20]. This noninteracting character enables us to extract the ideal physics of DES through theoretical and quantitative analysis.
The full relativistic first-principles calculation provides a Dirac-like linear dispersion with a mass gap of meV (Fig. 1 (c)) and an effective “speed of light” of m/s [21, 22, 23, 24]. The resistivity above 50 K is nearly independent of temperature, as has also been observed in the high-pressure massless Dirac phase of -(BEDT-TTF)2I3 (Fig. 1 (d)). Here, Dirac electrons compensate for the temperature dependence of the mobility and the density of states [8], resulting in the temperature-independent resistivity corresponding to quantum sheet resistance, m cm, where Å is the interplane distance using the value of the lattice constant along .
The resistivity increases upon cooling below 50 K without a phase transition, consistent with the mass gap but does not follow an activated temperature dependence. The Hall coefficient () is small above 50 K, and its sign changes at K from high-temperature positive (hole-like) values to low-temperature negative (electron-like) values, as shown in Fig. 1 (e). This indicates that the Fermi energy is in the mass gap but shifts slightly with temperature. At 150 K, the Fermi energy will be exactly at the midpoint of the gap.
We show in Fig. 2 (a) the magnetic susceptibilities for and (interplane direction). The susceptibility , which nearly agrees with , decreases linearly upon cooling below 150 K, consistent with Dirac-type linear dispersion. A possible in-plane anisotropy of originating from the tilting of the Dirac cone is negligible; therefore, we can consider that solely depends on the density of states and the electronic correlation is negligible. A detailed formula for the spin contribution to is given below using , the spin susceptibility of 2D nonrelativistic electron gas with the interplane distance .
(1) |
where is the electron mass, is the mass gap, and . in SI units, where is the vacuum permeability. The chemical potential is set to be zero for simplicity (see Ref. [21] for general ). In Fig. 2 (b), we plot using K, which provides with the ab initio value of , as well as the experimental , and find that this simple formula quantitatively reproduce the experiments.

The magnetic susceptibility perpendicular to the conducting plane, , is strongly suppressed and shows negative values below 150 K, which indicates an orbital diamagnetism, , that emerges only along the -direction. This diamagnetic was theoretically predicted for 2D DESs [25, 26]. In contrast to spins, which are conserved, the orbital currents that generate are not conserved. In general, the susceptibility of a non-conserved quantity has a contribution from high-energy bands, , which is independent of temperature and irrelevant to the Dirac band. We estimated by fitting for K with (see below Eq. (3)). We plot in Fig. 2 (b) the temperature dependent component, .
For free electrons, a detailed formula for in the presence of a mass gap is given by [27]
(2) |
where the chemical potential is set to be zero (see Ref. [21] for general ). In Fig. 2 (b), we plot experimental and calculated obtained using Eq. (2) with the same parameters as those for . We find quantitative agreements between the experimental and theoretical values as well as those for in the wide temperature range of K. Equation (2) shows a crossover at and is approximated as
(3) |
so that = const is expected for .
The observed uncorrelated character of the 2D DES for K and the deviation of from Eq. (2) are microscopically supported by 13C NMR. High-temperature Korringa-like for K is significantly reduced following with for K as shown in Fig. 2 (c), which indicates a linear dispersion. Note that the observed is five times smaller than that of -(BEDT-TTF)2I3, showing that the Dirac electrons in -(BETS)2I3 are relatively free of one-body renormalization of Coulomb repulsions [20]. The increase in below 20 K indicates other emergent relaxation mechanisms. Fig. 2 (d) depicts the anisotropies of , which we expect to be temperature-independent when the spin contribution dominates . The reduction of the anisotropy below 100 K, the onset temperature of the DES, coincides with that of . Since 13C does not couple with the electric field gradient, the most plausible source of the relaxation at low temperatures is the fluctuation of the orbital currents, which contributes to as . Recent theories point out that dominates in 3D Weyl materials [28, 29], but predict for clean 2D DESs, which does not reproduce the experiments below 20 K [29]. Later, we will discuss a potential mechanism for the deviation related to .
The correspondence between DES and QED relates Eq. (3) to the exotic quantized electric property in 2D DESs. In parallel to the Lorentz covariance in QED, we can show a duality between electric and magnetic responses in DES [21, 30, 31]. For and , the static magnetic susceptibility is given exactly as
(4) |
where and are the speed of light and vacuum permittivity, respectively [21]. Here, is the dynamical electrical conductivity, which originates only from interband electron–hole excitations. Thus, this duality relation indicates that dynamical vacuum fluctuations (the creation and annihilation of virtual electron–hole pairs), or the interband effect across the mass gap, necessarily generate the orbital diamagnetism .
A dimensional analysis gives for the massless limit of in the dimensions. In three dimensions (), Eq. (4) leads to a logarithmic divergence in for , which corresponds to the well-known ultraviolet divergence in the charge renormalization of QED [16]. In two dimensions (), on the other hand, there is no charge renormalization [17]. The large diamagnetism in Eq. (3) is therefore free from charge renormalization but closely linked to the -independent electrical conductivity for , where it takes a universal value of (quantized optical conductance) [32, 14]. (In Table 1, we summarize and as well as the permittivity for 3D and 2D DESs.) More precisely, using a detailed formula for [32], we find that the duality relation, Eq. (4), expresses in terms of the universal constant as
(5) |
3D | |||
---|---|---|---|
2D | 1 |
It is noteworthy that the conductivity unit can be rewritten using the susceptibility unit as , where is the impedance of free space and is the Compton wavelength, leading to the equivalence of Eqs. (3) and (5). This equivalence shows that scales with the universal electric conductance even for finite temperatures.
The dc conductivity ( is of the order of ) is difficult to determine theoretically, depending on the characteristics of the disorder [33, 34, 35, 36]. is naively given as [37] for 2D massless Dirac electrons but remains under debate for . The experimentally obtained values of ’s for organic DES, in contrast, are independent of temperature both for -(BETS)2I3 and -(BEDT-TTF)2I3; the values are approximately equal to km-1, corresponding to [15].
We plot and , normalized by and , respectively, in Fig. 3, and find that these electromagnetic responses are scaled in a wide temperature range, as anticipated from Eq. (5). The observed electromagnetic duality manifests itself in the correspondence with the Lorentz covariance in QED. The interband effect across the mass gap in the presence of electromagnetic fields characterizes the physical properties.

We now discuss potential sources for the deviation of from Eq. (2), although Eq. (1) does reproduce below 40 K. The most plausible source is a disorder in real materials, which raises a new problem related to the interaction of disorder and orbital currents in DESs. We found that the function fits the , as shown in Fig. 2 (c). The logarithmic increase upon cooling below 20 K does not originate from the electronic correlation, which enhances for the whole temperature range, and the crossing of the curves of -(BETS)2I3 and -(BEDT-TTF)2I3 at K suggests disorder effects on . A similar moderate increase of is observed for a 3D Weyl system [38], which has been theoretically analyzed considering the effects of impurities or temperature-dependent chemical potential to [39, 40]. Likewise, an observed increase in for a noninteracting 3D DES with meV, Bi0.9Sb0.1 [41], is closely related to our observation of the moderate increase in below 20 K. A related phenomenon is also observed for the transport properties: namely, unconventional negative magnetoresistance with a field dependence of the form [21]. These deviations from an ideal 2D DES are observed solely for the orbital-related properties at low temperatures, which suggests a new problem in disordered orbital physics in DESs. Surprisingly, despite the deviation of from Eq. (2) below 50 K, approximately scales with , thus maintaining the electromagnetic duality even at low temperatures where the effect of a disorder becomes crucial as shown in Fig. 3. This suggests a possible relationship between the effects of disorder on and , which results in a less disturbed electromagnetic duality.
In summary, we identified the organic conductor, -(BETS)2I3, as a 2D DES at ambient pressure through electric and magnetic measurements of , , , and of 13C NMR. The latter two magnetic responses show negligible electronic correlation, enabling us to study an ideal characteristics of DES. We found orbital diamagnetism () only along the interplane direction. We demonstrate that the equation holds approximately for K and , and that small shifts from the gapless DES are well reproduced by the theory using a unique parameter, , the mass gap for the DES. We found a unified electromagnetic responses in which scales with in a wide temperature range, as shown in Fig. 3, consistent with an electromagnetic duality that is valid only within the relativistic framework.
Acknowledgements.
We are grateful to H. Fukuyama, H. Sawa, T. Morinari, Y. Fuseya, H. Matsuura, I. Tateishi, and S. Ozaki for fruitful discussions. This work was supported by Grants-in-Aid for Scientific Research (20K03870, 21K03426, 18K03482, 18H01162, 19K21860, and 16H06346) from JSPS.References
- Wolff [1964] P. Wolff, Matrix elements and selection rules for the two-band model of bismuth, Journal of Physics and Chemistry of Solids 25, 1057 (1964).
- Novoselov et al. [2005] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov, Two-dimensional gas of massless Dirac fermions in graphene, Nature 438, 197 (2005).
- Shoenberg et al. [1936] D. Shoenberg, M. Z. Uddin, and E. Rutherford, The magnetic properties of bismuth I Dependence of susceptibility on temperature and addition of other elements, Proceedings of the Royal Society of London. Series A - Mathematical and Physical Sciences 156, 687 (1936).
- Wehrli [1968] L. Wehrli, Die magnetische Suszeptibilität von Bi und Bi-Sb-Legierungen, Physik der kondensierten Materie 8, 87 (1968).
- Fukuyama and Kubo [1970] H. Fukuyama and R. Kubo, Interband Effects on Magnetic Susceptibility. II. Diamagnetism of Bismuth, Journal of the Physical Society of Japan 28, 570 (1970).
- Suetsugu et al. [2021] S. Suetsugu, K. Kitagawa, T. Kariyado, A. W. Rost, J. Nuss, C. Mühle, M. Ogata, and H. Takagi, Giant orbital diamagnetism of three-dimensional Dirac electrons in antiperovskite, Phys. Rev. B 103, 115117 (2021).
- Li et al. [2015] Z. Li, L. Chen, S. Meng, L. Guo, J. Huang, Y. Liu, W. Wang, and X. Chen, Field and temperature dependence of intrinsic diamagnetism in graphene: Theory and experiment, Phys. Rev. B 91, 094429 (2015).
- Tajima et al. [2000] N. Tajima, M. Tamura, Y. Nishio, K. Kajita, and Y. Iye, Transport Property of an Organic Conductor -(BEDT-TTF)2I3 under High Pressure - Discovery of a Novel Type of Conductor -, Journal of the Physical Society of Japan 69, 543 (2000).
- Katayama et al. [2006] S. Katayama, A. Kobayashi, and Y. Suzumura, Pressure-Induced Zero-Gap Semiconducting State in Organic Conductor -(BEDT-TTF)2I3 Salt, Journal of the Physical Society of Japan 75, 054705 (2006).
- Hirata et al. [2016] M. Hirata, K. Ishikawa, K. Miyagawa, M. Tamura, C. Berthier, D. Basko, A. Kobayashi, G. Matsuno, and K. Kanoda, Observation of an anisotropic Dirac cone reshaping and ferrimagnetic spin polarization in an organic conductor, Nature Communications 7, 12666 (2016).
- Édel’Man [1975] V. S. Édel’Man, Investigation of bismuth in a quantizing field, Soviet Journal of Experimental and Theoretical Physics 41, 125 (1975).
- Boyle and Brailsford [1960] W. S. Boyle and A. D. Brailsford, Far Infrared Studies of Bismuth, Phys. Rev. 120, 1943 (1960).
- Katsnelson et al. [2006] M. I. Katsnelson, K. S. Novoselov, and A. K. Geim, Chiral tunnelling and the Klein paradox in graphene, Nature Physics 2, 620 (2006).
- Li et al. [2008] Z. Q. Li, E. A. Henriksen, Z. Jiang, Z. Hao, M. C. Martin, P. Kim, H. L. Stormer, and D. N. Basov, Dirac charge dynamics in graphene by infrared spectroscopy, Nature Physics 4, 532 (2008).
- Tajima et al. [2007] N. Tajima, S. Sugawara, M. Tamura, R. Kato, Y. Nishio, and K. Kajita, Transport properties of massless Dirac fermions in an organic conductor -(BEDT-TTF)2I3 under pressure, Europhysics Letters (EPL) 80, 47002 (2007).
- Maebashi et al. [2017] H. Maebashi, M. Ogata, and H. Fukuyama, Lorentz Covariance of Dirac Electrons in Solids: Dielectric and Diamagnetic Properties, Journal of the Physical Society of Japan 86, 083702 (2017).
- González et al. [1994] J. González, F. Guinea, and M. Vozmediano, Non-Fermi liquid behavior of electrons in the half-filled honeycomb lattice (A renormalization group approach), Nuclear Physics B 424, 595 (1994).
- Inokuchi et al. [1995] M. Inokuchi, H. Tajima, A. Kobayashi, T. Ohta, H. Kuroda, R. Kato, T. Naito, and H. Kobayashi, Electrical and Optical Properties of -(BETS)2I3 and -(BEDT-STF)2I3, Bulletin of the Chemical Society of Japan 68, 547 (1995).
- Hiraki et al. [2011] K.-i. Hiraki, S. Harada, K. Arai, Y. Takano, T. Takahashi, N. Tajima, R. Kato, and T. Naito, Local Spin Susceptibility of -D2I3 (D = bis(ethylendithio)tetraselenafulvalene (BETS) and bis(ethylendithio)dithiadiselenafulvalene (BEDT-STF)) Studied by 77Se NMR, Journal of the Physical Society of Japan 80, 014715 (2011).
- Hirata et al. [2017] M. Hirata, K. Ishikawa, G. Matsuno, A. Kobayashi, K. Miyagawa, M. Tamura, C. Berthier, and K. Kanoda, Anomalous spin correlations and excitonic instability of interacting 2D Weyl fermions, Science 358, 1403 (2017).
- [21] see Supplemental Material.
- Kitou et al. [2021] S. Kitou, T. Tsumuraya, H. Sawahata, F. Ishii, K.-i. Hiraki, T. Nakamura, N. Katayama, and H. Sawa, Ambient-pressure Dirac electron system in the quasi-two-dimensional molecular conductor , Phys. Rev. B 103, 035135 (2021).
- Perdew et al. [1996] J. P. Perdew, K. Burke, and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett. 77, 3865 (1996).
- Wimmer et al. [1981] E. Wimmer, H. Krakauer, M. Weinert, and A. J. Freeman, Full-potential self-consistent linearized-augmented-plane-wave method for calculating the electronic structure of molecules and surfaces: O2 molecule, Phys. Rev. B 24, 864 (1981).
- McClure [1956] J. W. McClure, Diamagnetism of Graphite, Phys. Rev. 104, 666 (1956).
- Fukuyama [2007] H. Fukuyama, Anomalous Orbital Magnetism and Hall Effect of Massless Fermions in Two Dimension, Journal of the Physical Society of Japan 76, 043711 (2007).
- Koshino and Ando [2011] M. Koshino and T. Ando, Singular orbital magnetism of graphene, Solid State Communications 151, 1054 (2011).
- Dóra and Simon [2009] B. Dóra and F. Simon, Unusual Hyperfine Interaction of Dirac Electrons and NMR Spectroscopy in Graphene, Phys. Rev. Lett. 102, 197602 (2009).
- Maebashi et al. [2019] H. Maebashi, T. Hirosawa, M. Ogata, and H. Fukuyama, Nuclear magnetic relaxation and Knight shift due to orbital interaction in Dirac electron systems, Journal of Physics and Chemistry of Solids 128, 138 (2019).
- Peskin and Schroeder [1995] M. E. Peskin and D. V. Schroeder, An Introduction to Quantum Field Theory (Addison-Wesley, Reading, MA, 1995).
- Giuliani and Vignale [2005] G. F. Giuliani and G. Vignale, Quantum Theory of the Electron Liquid (Cambridge University Press, Cambridge, U.K., 2005).
- Gusynin et al. [2006] V. P. Gusynin, S. G. Sharapov, and J. P. Carbotte, Unusual Microwave Response of Dirac Quasiparticles in Graphene, Phys. Rev. Lett. 96, 256802 (2006).
- Shon and Ando [1998] N. H. Shon and T. Ando, Quantum Transport in Two-Dimensional Graphite System, Journal of the Physical Society of Japan 67, 2421 (1998).
- Adam et al. [2007] S. Adam, E. H. Hwang, V. M. Galitski, and S. Das Sarma, A self-consistent theory for graphene transport, Proceedings of the National Academy of Sciences 104, 18392 (2007).
- Ostrovsky et al. [2006] P. M. Ostrovsky, I. V. Gornyi, and A. D. Mirlin, Electron transport in disordered graphene, Phys. Rev. B 74, 235443 (2006).
- Ostrovsky et al. [2007] P. M. Ostrovsky, I. V. Gornyi, and A. D. Mirlin, Quantum Criticality and Minimal Conductivity in Graphene with Long-Range Disorder, Phys. Rev. Lett. 98, 256801 (2007).
- Ando et al. [2002] T. Ando, Y. Zheng, and H. Suzuura, Dynamical Conductivity and Zero-Mode Anomaly in Honeycomb Lattices, Journal of the Physical Society of Japan 71, 1318 (2002).
- Yasuoka et al. [2017] H. Yasuoka, T. Kubo, Y. Kishimoto, D. Kasinathan, M. Schmidt, B. Yan, Y. Zhang, H. Tou, C. Felser, A. Mackenzie, and et al., Emergent Weyl Fermion Excitations in TaP Explored by Ta181 Quadrupole Resonance, Physical Review Letters 118, 10.1103/physrevlett.118.236403 (2017).
- Hirosawa et al. [2020] T. Hirosawa, H. Maebashi, and M. Ogata, Nuclear spin relaxation rate near the disorder-driven quantum critical point in Weyl fermion systems, Phys. Rev. B 101, 155103 (2020).
- Okvátovity et al. [2019] Z. Okvátovity, H. Yasuoka, M. Baenitz, F. Simon, and B. Dóra, Nuclear spin-lattice relaxation time in TaP and the Knight shift of Weyl semimetals, Phys. Rev. B 99, 115107 (2019).
- MacFarlane et al. [2014] W. A. MacFarlane, C. B. L. Tschense, T. Buck, K. H. Chow, D. L. Cortie, A. N. Hariwal, R. F. Kiefl, D. Koumoulis, C. D. P. Levy, I. McKenzie, F. H. McGee, G. D. Morris, M. R. Pearson, Q. Song, D. Wang, Y. S. Hor, and R. J. Cava, -detected NMR of in Bi, Sb, and the topological insulator , Phys. Rev. B 90, 214422 (2014).