Next Article in Journal / Special Issue
Digital Quantum Simulation of Scalar Yukawa Coupling
Previous Article in Journal
Ring and Linear Structures of CdTe Clusters
Previous Article in Special Issue
A Dyson Brownian Motion Model for Weak Measurements in Chaotic Quantum Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quantum Theory of Lee–Naughton–Lebed’s Angular Effect in Strong Electric Fields

Physics Department, University of Arizona, 1118 E. 4th Street, Tucson, AZ 85721, USA
Quantum Rep. 2024, 6(3), 359-365; https://doi.org/10.3390/quantum6030023
Submission received: 24 May 2024 / Revised: 7 July 2024 / Accepted: 9 July 2024 / Published: 17 July 2024
(This article belongs to the Special Issue Exclusive Feature Papers of Quantum Reports in 2024–2025)

Abstract

:
Some time ago, Kobayashi et al. experimentally studied the so-called Lee–Naughton–Lebed’s (LNL) angular effect in strong electric fields [Kobayashi, K.; Saito, M.; Omichi E.; Osada, T. Phys. Rev. Lett. 2006, 96, 126601]. They found that strong electric fields split the LNL conductivity maxima in an α -(ET)2-based organic conductor and hypothetically introduced the corresponding equation for conductivity. In this paper, for the first time, we suggest the quantum mechanical theory of the LNL angular oscillations in moderately strong electric fields. In particular, we demonstrate that the approximate theoretical formula obtained by us well describes the above mentioned experiments.

1. Introduction

It is well known that organic conductors having quasi-one-dimensional (Q1D) pieces of the Fermi surfaces (FSs) demonstrate unique magnetic properties due to the Bragg reflections of moving electrons from the Brillouin zone boundaries in moderate and strong magnetic fields [1,2,3,4,5]. Among them are the Field-Induced Spin(Charge)-Density-Wave (FIS(C)DW) phase diagrams [3,4,5,6,7,8,9,10,11,12,13,14,15], 3D Quantum Hall Effect (3D QHE) [14,15,16], the so-called Lebed’s Magic Angles (LMAs) [17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40], the Lee–Naughton–Lebed’s (LNL) angular oscillations [41,42,43,44,45,46,47], and some others. Note the LMA phenomena [17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40] seem to be very complicated and, in most cases, possess some non-Fermi liquid (FL) properties [1,27,29], whereas the FIS(C)DW, 3D QHE, and LNL phenomena have been successfully explained in the framework of the Landau FL approach [1,2]. In particular, the LNL phenomenon has been successfully theoretically explained in Refs. [48,49,50,51,52,53,54]. Indeed, in Refs. [48,49,50,51,52,53,54], a layered Q1D conductor with the electron spectrum was considered,
ϵ 0 ± ( p ) = ± v F ( p x p F ) + 2 t b cos ( p y b * / ) + 2 t c cos ( p z c * / ) .
[Note that in Equation (1), the first term represents electron free motion along the conducting chains on the right (+) and left (−) sheets of the Q1D FS, with p F and v F being the Fermi momentum and Fermi velocity, correspondingly. The second and the third term correspond to the hopping of electrons in the perpendicular axes, b * and c * ( p F v F t b t c ); p is the total electron momentum; p x is its component along conducting chains, whereas p y and p z are electron momentum components along b * and c * axes, correspondingly; is the Planck constant.] The Q1D conductor is placed in the following inclined magnetic field,
H = H ( sin θ cos ϕ , sin θ sin ϕ , cos θ ) ,
and in electric field E along z direction (see Figure 1). In the quasi-classical approximation, the following expression for the LNL conductivity was derived by several methods:
σ z z ( H , θ , ϕ ) = σ z z ( 0 ) N = J N 2 [ ω c * ( θ , ϕ ) / ω b ( θ ) ] 1 + τ 2 [ ω c ( θ , ϕ ) N ω b ( θ ) ] 2 ,
where σ z z ( 0 ) is conductivity at H = 0 , and J N ( x ) is the Bessel function of the N-th order. Note that in Equation (3), the so-called electron cyclotron frequencies can be expressed as [48,49,50,51,52,53,54]:
ω b ( θ ) = | e | v F H b * cos θ c , ω c ( θ , ϕ ) = | e | v F H c * sin θ sin ϕ c ,
ω c * ( θ , ϕ ) = | e | v y 0 H c * sin θ cos ϕ c , v y 0 = 2 t b b * ,
where e is the electron charge, and c is the speed of light. More recently, Kobayashi et al. [55] experimentally studied the LNL phenomenon in rather strong electric fields and found that the strong electric field splits the LNL maxima of conductivity (3). What is also important is that they suggested a hypothetical theoretical formula which described the above mentioned experimental splitting.
The goal of our paper is to derive the quasi-classical expression for conductivity in moderately strong electric and strong magnetic fields which describes the experimentally observed splitting of the LNL maxima of conductivity [55]. In particular, we show that our equation has a limited area of applicability and is not applicable in very strong electric fields.

2. Materials and Methods

First, let us perform the quasi-classical Peierls substitution [56,57] for motion along the conducting chains in Equation (1), in the absence of both magnetic and electric fields.
ϵ ^ 0 ± ( x , p y , p z ) = i v F d d x + 2 t b cos ( p y b * / ) + 2 t c cos ( p z c * / ) .
The solution of the corresponding Schrödinger equation is
Ψ 0 ± ( x , p y , p z ) = exp ( ± i ϵ x v F ) exp [ i 2 t b x v F cos ( p y b * ) ] exp [ i 2 t c x v F cos ( p z c * ) ] ,
where energy ϵ is counted from the Fermi level, ϵ F = p F v F .
Then, we introduce the electric field applied along the least conducting z axis as a small perturbation to the Hamiltonian (6),
δ ϵ ^ ( z ) = e E z ,
and perform one more quasi-classical Peierls substitution [56,57]:
δ ϵ ^ ( p z ) = e E z = i e E d d p z .
In this case, the application of the perturbation (9) to the free electron wave function (7) gives
δ ϵ ^ ( p z ) Ψ 0 ± ( x , p y , p z ) = ± e E x v F 2 t c c * sin ( p z c * / ) Ψ 0 ± ( x , p y , p z ) .
It is easy to prove that, for not extremely strong electric fields, the total Hamiltonian in the electric field can be written as
ϵ ^ ± ( x , p y , p z ) = i v F d d x + 2 t b cos ( p y b * / ) + 2 t c cos ( p z c * e E c * x v F ) .
Here, we introduce the magnetic field (2) in the electron Hamiltonian and the electron velocity operator along z axis. For further development, it is convenient to choose the vector potential of the magnetic field in the following form:
A = ( 0 , x cos θ , x sin θ sin ϕ + y sin θ cos ϕ ) H .
To define the corresponding electron wave functions for the case, where t b t c , as shown in Ref. [51], it is necessary to take into account only two first terms in Hamiltonian (11) and to perform in the second term the following quasi-classical Peierls substitution,
p y p y e c A y .
In this case, wave function in the mixed ( x , p y ) representation obeys the following Schrödinger equation [3,51]:
( i v F d d x + 2 t b cos [ p y b * ω b ( θ ) x v F ] ) Φ ϵ ± ( x , p y ) = ϵ Φ ϵ ± ( x , p y ) ,
where the two wave functions (7) and (14) are related by the following equation:
Ψ ϵ ± ( x , p y ) = exp ( ± i p F x / ) Φ ϵ ± ( x , p y ) .
It is important that Equation (14) can be exactly solved,
Φ ϵ ± ( x , p y ) = exp ( ± i ϵ v F x ) exp { ± 2 i t b ω b ( θ ) ( sin [ p y b * ω b ( θ ) x v F ] sin [ p y b * ] ) } .
Let us apply the quasi-classical Peierls substitution to energy dependence (11) to the momentum component along z axis:
ϵ ^ z ± ( x , y , p z ) = 2 t c cos ( p z c * e E c * x v F ) 2 t c cos [ p z c * e E c * x v F + ω c ( θ , ϕ ) x v F ω c * ( θ , ϕ ) y v y 0 ] .
Taking into account that, in the quasi-classical approximation,
v ^ z ± ( x , y , p z ) = d [ ϵ ^ z ± ( x , y , p z ) ] / d p z , y = i ( d / d p y ) ,
it is possible to write the velocity component operator along z axis in the following form:
v ^ z ± ( x , y , p z ) = 2 t c c * sin [ p z c * e E c * x v F + ω c ( θ , ϕ ) x v F i ω c * ( θ , ϕ ) ( d / d p y ) v y 0 ] .
In Equation (19), for further development, we introduce
ω c ± ( θ , ϕ ) = ω c ( θ , ϕ ) e E c * / .
It is important that wave functions (16) are eigenfunctions of the velocity operator along z axis (19), (20) with the following eigenvalues:
v ^ z ± ( x , y , p z ) Φ ϵ ± ( x , p y ) = 2 t c c * sin { p z c * + ω c ± ( θ , ϕ ) x v F ± ω c * ( θ , ϕ ) ω b ( θ ) × ( cos [ p y b * ω b ( θ ) x v F ] cos [ p y b * ] ) } Φ ϵ ± ( x , p y ) .

3. Results

Let us make use of the Kubo formula for conductivity [51,58]. We can do this because the electron wave functions (16) and the eigenvalues of velocity operators (21) are known. The total conductivity along z axis can be represented as a summation of the following two contributions: one from the right sheet of the FS (1) and another from the left sheet,
σ z z ( H , θ , ϕ ) = σ z z + ( H , θ , ϕ ) + σ z z ( H , θ , ϕ ) .
By means of the Kubo formalism [51,58], we obtain
σ z z ± ( H , θ , ϕ ) π π d ( p y b * ) 0 d x exp ( x v F τ ) × cos { ω c ± ( θ , ϕ ) x v F ± ω c * ( θ , ϕ ) ω b ( θ ) ( cos [ p y b * ω b ( θ ) x v F ] cos [ p y b * ] ) } ,
where τ is an electron relaxation time. The complicated double integration in Equation (23) can be simplified using definitions of the Bessel functions of the N-th order, J N ( x ) [51,59],
σ z z ± ( H , θ , ϕ ) = σ z z ( 0 ) 2 N = J N 2 [ ω c * ( θ , ϕ ) / ω b ( θ ) ] 1 + τ 2 [ ω c ± ( θ , ϕ ) N ω b ( θ ) ] 2 ,
where σ z z ( 0 ) —conductivity along z axis in low electric fields in the absence of the magnetic field. If we make use of Equation (22), we finally obtain for the total conductivity in moderately strong electric fields in the presence of the inclined magnetic field (2) the following:
σ z z ( H , θ , ϕ ) = σ z z ( 0 ) 2 N = { J N 2 [ ω c * ( θ , ϕ ) / ω b ( θ ) ] 1 + τ 2 [ ω c + ( θ , ϕ ) N ω b ( θ ) ] 2 + J N 2 [ ω c * ( θ , ϕ ) / ω b ( θ ) ] 1 + τ 2 [ ω c ( θ , ϕ ) N ω b ( θ ) ] 2 } .

4. Discussion

We stress that Equation (25) is the main result of our paper, whereas in Ref. [55], this equation was just guessed. Moreover, we have shown that it is not exact and has to be used for not too high (i.e., moderately high) electric fields. Indeed, let us discuss its applicability. We recall that we have derived Equation (25) using some approximation: we have suggested that we can use Equation (11), instead of Equation (10). It is easy to prove that this can be done under the condition that
| e | E c * x 0 v F 1 ,
where x 0 is characteristic length where the integral (23) converges. Since, as follows from (23), x 0 v F τ , the condition (26) can be written as
| e | E c * / τ .
If we take the lowest experimentally used electric field, V 0 = E d = 2 V, d = 0.2 mm [55], / τ = 2 K and c * 2 nm [1], we obtain the inequality (27) in the form
0.25 K 2 K ,
which shows that, at lowest voltages, the analysis in [55] is correct, whereas at higher experimental voltages like V 0 = 20 V [55], Equation (25) must be used with great caution, since Equation (27) gives quantities of the same orders of magnitudes for the left side and for the right one.
Let us briefly discuss one important consequence of Equation (25)—the splitting of the LNL maxima of conductivity in moderately strong electric fields [55]. In the limit of zero electric field at the following typical experimental conditions, where
ω b ( θ ) τ 1 , ω c ( θ , ϕ ) τ 1 ,
the maxima of conductivity, as follows from Equation (3), appear at
ω c ( θ , ϕ ) = N ω b ( θ ) ,
where N is an arbitrary integer. Under the experimental condition (29), Equation (25) splits each maximum into two ones, which are defined by the following equations
ω c 1 , 2 ( θ , ϕ ) = N ω b ( θ ) ω E , ω E = e E c * / .
The effect of splitting was experimentally observed in Ref. [55]. Our analysis of the applicability of Equation (25), as we discussed above, has shown that Equation (25) is valid for lower experimentally used voltages, V 0 2 V, and become controversial at higher ones, V 0 20 V.
It is interesting that the obtained results are general for all families of the Q1D conductors. The splitting of the LNL maxima of conductivity appears due to the fact that the Lorentz force changes its sign between the left and right pieces of the Q1D FS due to the change in the electron velocity sign, whereas the electric force does not change its sign. Therefore, it is instructive to analyze novel Equations (26) and (27) in the typical type of Q1D conductors like the (TMTSF)2X conductors. Indeed, the LNL oscillations are best studied in (TMTSF)2PF6, where c * = 1.36 nm [1] and / τ 1 K [43]. As seen from Equation (27), the splitting of the LNL maxima of conductivity has to be observed in the same electric field range as they were observed in the α -(ET)2-based conductor by Kobayashi et al. [55]. The obvious experimental problem is how to avoid the overheating of the (TMTSF)2PF6 sample. As for the (TMTSF)2ClO4 conductor, we have to be careful and use the Peierls substitution method at magnetic fields which are lower than the so-called magnetic breakdown field [1].

Funding

This research received no external funding.

Data Availability Statement

Available data is contained within the article.

Acknowledgments

We are thankful to N.N. Bagmet (Lebed) for numerous and fruitful discussions.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Lebed, A.G. (Ed.) The Physics of Organic Superconductors and Conductors; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  2. Ishiguro, T.; Yamaji, K.; Saito, G. Organic Superconductors, 2nd ed.Springer: Berlin/Heidelberg, Germany, 1998. [Google Scholar]
  3. Gor’kov, L.P.; Lebed, A.G. On the stability of the quasi-one-dimensional metallic phase in magnetic fields against the spin density wave formation. J. Phys. Lett. 1984, 45, L-433. [Google Scholar]
  4. Heritier, M.; Montambaux, G.; Lederer, P. Stability of the spin density wave phases in (TMTSF) 2ClO4: Quantized nesting effect. J. Phys. Lett. 1984, 45, 943–952. [Google Scholar] [CrossRef]
  5. Chaikin, P.M. Magnetic-field-induced transition in quasi-two-dimensional systems. Phys. Rev. 1985, B31, 4770. [Google Scholar] [CrossRef] [PubMed]
  6. Chaikin, P.M.; Choi, M.-Y.; Kwak, J.F.; Brooks, J.S.; Martin, K.P.; Naughton, M.J.; Engler, E.M.; Greene, R.L. Tetramethyltetraselenafulvalenium Perchlorate, (TMTSF)2ClO4, in High Magnetic Fields. Phys. Rev. Lett. 1983, 51, 2333. [Google Scholar] [CrossRef]
  7. Ribault, M.; Jerome, D.; Tuchendler, J.; Weyl, C.; Bechgaard, K. Low-field and anomalous high-field Hall effect in (TMTSF)2ClO4. J. Phys. Lett. 1983, 44, 953–961. [Google Scholar] [CrossRef]
  8. Lebed, A.G. Field-induced spin-density-wave phases in quasi-one-dimensional conductors: Theory versus experiments. Phys. Rev. Lett. 2002, 88, 177001. [Google Scholar] [CrossRef]
  9. Zanchi, D.; Bjelis, A.; Montambaux, G. Phase diagram for charge-density waves in a magnetic field. Phys. Rev. 1996, B53, 1240. [Google Scholar] [CrossRef] [PubMed]
  10. Qualls, J.S.; Balicas, L.; Brooks, J.S.; Harrison, N.; Montgomery, L.K.; Tokumoto, M. Competition between Pauli and orbital effects in a charge-density-wave system. Phys. Rev. 2000, B62, 10008. [Google Scholar] [CrossRef]
  11. Andres, D.; Kartsovnik, M.V.; Biberacher, W.; Weiss, H.; Balthes, E.; Muller, H.; Kushch, N. Orbital effect of a magnetic field on the low-temperature state in the organic metal α–(BEDT–TTF) 2 KHg (SCN) 4. Phys. Rev. 2001, B64, 161104(R). [Google Scholar] [CrossRef]
  12. Andres, D.; Kartsovnik, M.V.; Grigoriev, P.D.; Biberacher, W.; Muller, H. Orbital quantization in the high-magnetic-field state of a charge-density-wave system. Phys. Rev. 2003, B68, 201101. [Google Scholar] [CrossRef]
  13. Lebed, A.G. Theory of magnetic field-induced charge-density-wave phases. JETP Lett. 2003, 78, 138–142. [Google Scholar] [CrossRef]
  14. Hannahs, S.T.; Brooks, J.S.; Kang, W.; Chiang, L.Y.; Chaikin, P.M. Quantum Hall effect in a bulk crystal. Phys. Rev. Lett. 1989, 63, 1988. [Google Scholar] [CrossRef]
  15. Cooper, J.R.; Kang, W.; Auban, P.; Montambaux, G.; Jerome, D.; Bechgaard, K. Quantized Hall effect and a new field-induced phase transition in the organic superconductor (TMTSF)2PF6. Phys. Rev. Lett. 1989, 63, 1984. [Google Scholar] [CrossRef]
  16. Yakovenko, V.M. Quantum Hall effect in quasi-one-dimensional conductors. Phys. Rev. 1991, B43, 11353. [Google Scholar] [CrossRef]
  17. Lebed, A.G.; Bak, P. Theory of unusual anisotropy of magnetoresistance in organic superconductors. Phys. Rev. Lett. 1989, 63, 1315. [Google Scholar] [CrossRef]
  18. Naughton, M.J.; Chung, O.H.; Chiang, L.Y.; Brooks, J.S. MRS-Symposia Proceedings. Mater. Res. Soc. Symp. Proc. 1990, 173, 257. [Google Scholar] [CrossRef]
  19. Osada, T.; Kawasumi, A.; Kagoshima, S.; Miura, N.; Saito, G. Commensurability effect of magnetoresistance anisotropy in the quasi-one-dimensional conductor tetramethyltetraselenafulvalenium perchlorate, (TMTSF)2ClO4. Phys. Rev. Lett. 1991, 66, 1525. [Google Scholar] [CrossRef]
  20. Boebinger, G.S.; Montambaux, G.; Kaplan, M.L.; Haddon, R.C.; Chichester, S.V.; Chiang, L.Y. Anomalous magnetoresistance anisotropy in metallic and spin-density-wave phases of the quasi-one-dimensional organic conductor (TMTSF)2ClO4. Phys. Rev. Lett. 1990, 64, 591. [Google Scholar] [CrossRef]
  21. Naughton, M.J.; Chung, O.H.; Chaparala, M.; Bu, X.; Coppens, P. Commensurate fine structure in angular-dependent studies of (TMTSF)2ClO4. Phys. Rev. Lett. 1991, 67, 3712. [Google Scholar] [CrossRef]
  22. Kang, W.; Hannahs, S.T.; Chaikin, P.M. Lebed’s magic angle effects in (TMTSF)2PF6. Phys. Rev. Lett. 1992, 69, 2827. [Google Scholar] [CrossRef]
  23. Kartsovnik, M.V.; Kovalev, A.E.; Laukhin, V.N.; Pesotskii, S.E. Giant angular magnetoresistance oscillations in (BEDT-TTF) 2 TlHg (SCN) 4: The warped plane Fermi surface effect. J. Phys. I 1992, 2, 223–228. [Google Scholar] [CrossRef]
  24. Kartsovnik, M.V.; Kovalev, A.E.; Kushch, N.D. Magnetotransport investigation of the low-temperature state of transition (BEDT-TTF) 2TIHg (SCN) 4: Evidence for a Peierls-type transition. J. Phys. I 1993, 3, 1187–1199. [Google Scholar] [CrossRef]
  25. Benhia, K.; Ribault, M.; Lenior, C. Lebed resonance effects in the metallic and spin-density-wave phases of (TMTSF) 2PF6. Europhys. Lett. 1994, 25, 285. [Google Scholar]
  26. Kartsovnik, M.V.; Kovalev, A.E.; Laukhin, V.N.; Ito, H.; Ishiguro, T.; Kushch, N.D.; Anzai, H.; Saito, G. Agnetoresistance anisotropy in the organic superconductor κ-(BEDT-TTF) 2Cu (NCS)2. Synth. Met. 1995, 70, 819–820. [Google Scholar] [CrossRef]
  27. Chashechkina, E.I.; Chaikin, P.M. Magic Angles and the Ground States in (TMTSF)2PF6. Phys. Rev. Lett. 1998, 80, 2181. [Google Scholar] [CrossRef]
  28. Osada, T.; Nose, H.; Kuraguchi, M. Angular dependent phenomena in low-dimensional conductors under high magnetic fields. Phys. B Condens. Matter 2001, 294–295, 402–407. [Google Scholar] [CrossRef]
  29. Chashechkina, E.I.; Chaikin, P.M. Simple fit for magic-angle magnetoresistance in (TMTSF)2PF6. Phys. Rev. 2002, B65, 012405. [Google Scholar] [CrossRef]
  30. Kang, H.; Jo, Y.J.; Uji, S.; Kang, W. Evidence for coherent interchain electron transport in quasi-one-dimensional molecular conductors. Phys. Rev. 2003, B68, 132508. [Google Scholar] [CrossRef]
  31. Kang, H.; Jo, Y.J.; Kang, W. Pressure dependence of the angular magnetoresistance of (TMTSF)2PF6. Phys. Rev. 2004, B69, 033103. [Google Scholar] [CrossRef]
  32. Ito, H.; Suzuki, D.; Yokochi, Y.; Kuroda, S.; Umemiya, M.; Miyasaka, H.; Sugiura, K.-I.; Yamashita, M.; Tajima, H. Quasi-one-dimensional electronic structure of (DMET)2CuCl2. Phys. Rev. 2005, B71, 212503. [Google Scholar] [CrossRef]
  33. Kartsovnik, M.V.; Andres, D.; Simonov, S.V.; Biberacher, W.; Sheikin, I.; Kushch, N.D.; Miller, H. Angle-Dependent Magnetoresistance in the Weakly Incoherent Interlayer Transport Regime in a Layered Organic Conductor. Phys. Rev. Lett. 2006, 96, 166601. [Google Scholar] [CrossRef] [PubMed]
  34. Takahashi, S.; Betancur-Rodiguez, A.; Hill, S.; Takasaki, S.; Yamada, J.; Anzai, H. Are lebed’s magic angles truly magic? J. Low Temp. Phys. 2007, 142, 311–314. [Google Scholar] [CrossRef]
  35. Kang, W.; Osada, T.; Jo, Y.J.; Kang, H. Interlayer magnetoresistance of quasi-one-dimensional layered organic conductors. Phys. Rev. Lett. 2007, 99, 017002. [Google Scholar] [CrossRef] [PubMed]
  36. Kang, W. Absence of magic-angle effects in the intralayer resistance Rxx of the quasi-one-dimensional organic conductor (TMTSF)2ClO4. Phys. Rev. 2007, B76, 193103. [Google Scholar] [CrossRef]
  37. Bangura, A.F.; Goddard, P.A.; Singleton, J.; Tozer, S.W.; Coldea, A.I.; Ardavan, A.; McDonald, R.D.; Blundell, S.J.; Schlueter, J.A. Angle-dependent magnetoresistance oscillations due to magnetic breakdown orbits. Phys. Rev. 2007, B76, 0525010. [Google Scholar] [CrossRef]
  38. Kang, W.; Chung, O.-H. Quasi-one-dimensional Fermi surface of (TMTSF)2NO3. Phys. Rev. 2009, B79, 045115. [Google Scholar] [CrossRef]
  39. Graf, D.; Brooks, J.S.; Choi, E.S.; Almeida, M.; Henriques, R.T.; Dias, J.C.; Uji, S. Geometrical and orbital effects in a quasi-one-dimensional conductor. Phys. Rev. 2009, B80, 155104. [Google Scholar] [CrossRef]
  40. Kang, W.; Jo, Y.J.; Noh, D.Y.; Son, K.Y.; Chung, O.-H. Stereoscopic study of angle-dependent interlayer magnetoresistance in the organic conductor κ-(BEDT-TTF)2Cu(NCS)2. Phys. Rev. 2009, B80, 155102. [Google Scholar] [CrossRef]
  41. Naughton, M.J.; Lee, I.J.; Chaikin, P.M.; Danner, G.M. Critical fields and magnetoresistance in the molecular superconductors (TMTSF) 2X. Synth. Metals 1997, 85, 1481–1485. [Google Scholar] [CrossRef]
  42. Yoshino, H.; Saito, K.; Nishikawa, H.; Kikuchi, K.; Kobayashi, K.; Ikemoto, I. Fine structure of in-plane angular effect of magnetoresistance of (DMET) 2I3. J. Phys. Soc. Jpn. 1997, 66, 2248–2251. [Google Scholar] [CrossRef]
  43. Lee, I.J.; Naughton, M.J. Effective electrons and angular oscillations in quasi-one-dimensional conductors. Phys. Rev. 1998, B57, 7423. [Google Scholar] [CrossRef]
  44. Lee, I.J.; Naughton, M.J. Metallic state in (TMTSF)2PF6 at low pressure. Phys. Rev. 1998, B58, R13343. [Google Scholar] [CrossRef]
  45. Lebed, A.G.; Ha, H.-I.; Naughton, M.J. Angular magnetoresistance oscillations in organic conductors. Phys. Rev. 2005, B71, 132504. [Google Scholar] [CrossRef]
  46. Ha, H.I.; Lebed, A.G.; Naughton, M.J. Interference effects due to commensurate electron trajectories and topological crossovers in (TMTSF)2ClO4. Phys. Rev. 2006, B73, 033107. [Google Scholar] [CrossRef]
  47. Wu, S.; Lebed, A.G. Unification theory of angular magnetoresistance oscillations in quasi-one-dimensional conductors. Phys. Rev. 2010, B82, 075123. [Google Scholar] [CrossRef]
  48. McKenzie, R.H.; Moses, P. Periodic orbit resonances in layered metals in tilted magnetic fields. Phys. Rev. 1999, B60, R11241. [Google Scholar] [CrossRef]
  49. Lebed, A.G.; Naughton, M.J. Fermi surface interference effects and angular magnetic oscillations in Q1D conductors. J. Phys. IV 2002, 12, 369–372. [Google Scholar] [CrossRef]
  50. Osada, T. Resonant tunneling tuned by magnetic field orientations in anisotropic multilayer systems. Phys. E Low-Dimens. Syst. Nanostructures 2002, E12, 272–275. [Google Scholar] [CrossRef]
  51. Lebed, A.G.; Naughton, M.J. Interference commensurate oscillations in quasi-one-dimensional conductors. Phys. Rev. Lett. 2003, 91, 187003. [Google Scholar] [CrossRef]
  52. Osada, T.; Kuraguchi, M. General quantum picture for magnetoresistance angular effects in quasi-one-dimensional conductors. Synth. Met. 2003, 133–134, 75–77. [Google Scholar] [CrossRef]
  53. Banerjee, A.; Yakovenko, V.M. Angular magnetoresistance oscillations in quasi-one-dimensional organic conductors in the presence of a crystal superstructure. Phys. Rev. 2008, B78, 125404. [Google Scholar] [CrossRef]
  54. Cooper, B.K.; Yakovenko, V.M. Interlayer Aharonov-Bohm Interference in Tilted Magnetic Fields in Quasi-One-Dimensional Organic Conductors. Phys. Rev. Lett. 2006, 96, 037001. [Google Scholar] [CrossRef] [PubMed]
  55. Kobayashi, K.; Saito, M.; Omichi, E.; Osada, T. Electric-Field Effect on the Angle-Dependent Magnetotransport Properties of Quasi-One-Dimensional Conductors. Phys. Rev. Lett. 2006, 96, 126601. [Google Scholar] [CrossRef] [PubMed]
  56. Abrikosov, A.A. Fundamentals of Theory of Metals; Elsevier Science: Amsterdam, The Netherland, 1988. [Google Scholar]
  57. LIfshits, I.M.; Azbel, M.Y.; Kaganov, M.I. Electron Theory of Metals; Consultants Bureau: New York, NY, USA, 1973. [Google Scholar]
  58. Grosso, G.; Parravichini, G.P. Solid State Physics; Academic Press: New York, USA, 2000. [Google Scholar]
  59. Gradshteyn, L.S.; Ryzhik, I.M. Tables of Integrals, Series, and Products, 5th ed.; Academic Press, Inc.: London, UK, 1994. [Google Scholar]
Figure 1. Definition of the azimuthal angle θ and polar angle ϕ for the typical Lee–Naughton–Lebed’s experiment, where z is the least conducting axis.
Figure 1. Definition of the azimuthal angle θ and polar angle ϕ for the typical Lee–Naughton–Lebed’s experiment, where z is the least conducting axis.
Quantumrep 06 00023 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lebed, A.G. Quantum Theory of Lee–Naughton–Lebed’s Angular Effect in Strong Electric Fields. Quantum Rep. 2024, 6, 359-365. https://doi.org/10.3390/quantum6030023

AMA Style

Lebed AG. Quantum Theory of Lee–Naughton–Lebed’s Angular Effect in Strong Electric Fields. Quantum Reports. 2024; 6(3):359-365. https://doi.org/10.3390/quantum6030023

Chicago/Turabian Style

Lebed, Andrei G. 2024. "Quantum Theory of Lee–Naughton–Lebed’s Angular Effect in Strong Electric Fields" Quantum Reports 6, no. 3: 359-365. https://doi.org/10.3390/quantum6030023

Article Metrics

Back to TopTop