This work is concerned with the derivation of the Green’s function for Landau-quantized carriers in the Group-VI dichalcogenides. In the spatially homogeneous case, the Green’s function is separated into a Peierls phase factor and a translationally invariant part which is determined in a closed form integral representation involving only elementary functions. The latter is expanded in an eigenfunction series of Laguerre polynomials. These results for the retarded Green’s function are presented in both position and momentum representations, and yet another closed form representation is derived in circular coordinates in terms of the Bessel wave function of the second kind (not to be confused with the Bessel function). The case of a quantum wire is also addressed, representing the quantum wire in terms of a model one-dimensional δ(x)-potential profile. This retarded Green’s function for propagation directly along the wire is determined exactly in terms of the corresponding Green’s function for the system without the δ(x)-potential, and the Landau quantized eigenenergy dispersion relation is examined. The thermodynamic Green’s function for the dichalcogenide carriers in a normal magnetic field is formulated here in terms of its spectral weight, and its solution is presented in a momentum/integral representation involving only elementary functions, which is subsequently expanded in Laguerre eigenfunctions and presented in both momentum and position representations.

Starting with the discovery of the exceptional electronic transport and detection properties of graphene, there has grown up an extremely broad and powerful research effort directed at understanding and developing the basic science, engineering and production as such Dirac-like materials, having a relativistic-type energy spectrum. These materials include silicene, topological insulators and now group-VI dichalcogenides. As an agent for probing the properties and also inducing new physical features, the magnetic field has always had an important role. Such studies of the effects of the magnetic field in nonrelativistic carrier propagation and collective modes date back to the 1960’s and 1970’s for three and two-dimensional quantum plasmas.1,2 The relativistic Green’s function in a magnetic field was most elegantly first determined by Schwinger,3 and interesting variants for particular Dirac-like materials have been advanced by Rusin and Zawadski,4 as well as the author.5 This paper is focused on yet another important representation for the Landau-quantized Green’s function for the group-VI dichalcogenides, particularly in direct time representation, whose behavior is more transparent in the low field limit in comparison with that of the eigenfunction expansion. Moreover, it facilitates the analysis of the role of a quantum wire, modeled here by a one-dimensional δ(x)-potential profile, enabling the explicit determination of the wire Green’s function and its Landau quantized Dirac-like spectrum.

As in the case of graphene, the group-VI dichalcogenides have two Dirac points, K and K, at which relativistic-type spectra prevail for low energies. However, they also exhibit an energy gap and strong spin-orbit coupling, which must be taken into account. An appropriate kp model Hamiltonian is block diagonal, involving four 2×2 blocks (each block for a specific choice of spin and valley), and the individual 2×2 blocks are quite similar to that of graphene. Specifically, for the case at hand, the hamiltonian is given by

(1)

where inclusion of the valley index ν represents two of the four 2x2 blocks, and

(2)

is the canonical momentum (=c=1). A is the vector potential of the normal magnetic field B, σ=(σx,σy,σz) are the Pauli spin matrices, γv=γsign(ν) with ν=±1 as valley index and γ is determined by the tight binding hopping parameter and lattice spacing. Furthermore, a spin index sz=±1 enters into energy shifts as Esz=szνλ2 and g=Δ2Esz, where λ is the spin splitting and Δ is the energy gap without spin splitting.

Addressing the issue of gauge, it is useful to note that for a uniform constant magnetic field, the Green’s function equation [I2D is the 2D unit matrix]

(3)

involves Hamiltonian dependence on the vector potential A (and magnetic field B directly through spin), H(,A,B), which may be gauge transformed so that it is supplanted by A(r)12B×(rr), with the corresponding gauge transformation of the Green’s function producing a Peierls phase factor1,3C(r,r)(R=rr;T=tt)

(4)

and for the case of retardation, G(r,r;t,t)0 for t<t.

In this, C(r,r) embodies all dependence on (r+r) if there is no other spatial inhomogeneity, ϕ(r) is an arbitrary gauge function and G(R,T) is translationally invariant in both position space and time, obeying the modified Green’s function equation

(5)

In addition to G(R,T) being translationally invariant, it is independent of gauge. It will be analyzed here in rectangular position-time coordinates and in circular coordinates, also in momentum representation.

In Section II, the retarded Green’s function for group-VI dichalcogenides in a magnetic field is explicitly determined in a closed form integral representation in terms of elementary functions in both position and momentum representations; also its eigenfunction representation is derived. In Section III, the Green’s function for this system with a superposed model quantum wire δ(x)-potential is also determined explicitly for propagation directly along the wire, and its energy spectrum is examined. Furthermore, the thermodynamic Green’s function and spectral weight for the Landau-quantized dichalcogenides is discussed in Section IV. In all cases we exhibit the Green’s function in both position and momentum representations. A summary is presented in Section V. Details of solution, integration and the dispersion relation are included in three appendices.

Considering the Hamiltonian of Eq. (1) jointly with Eqns. (4) and (5), the 2×2 matrix Green’s function equation for G(rr;tt) near the K,K points of the group-VI dichalcogenides may be written as

(6)

or, in frequency representation (note that we discuss the retarded Green’s function here, ωω+i0+):

(7)

where

(8)

The 4 elements of the Green’s function matrix equation (Eq. 7) are given by

(9)
(10)
(11)
(12)

Employing Eqns.(10) and (12) to eliminate G21 and G12 from Eqns.(9) and (11) for G11 and G22, respectively, we have

(13)

where G1122=G11 or G22 corresponding to the upper or lower of alternative ±, signs elsewhere in the equation.

On the basis of Eq. (8) and the identity

(14)

we obtain

(15)

This result may be further simplified by noting that the operator LZ,

(16)

vanishes in application to the Green’s function,

(17)

because of conservation of the z-component of angular momentum. (The creation of a particle at (x,t) provides angular momentum which is subsequently eliminated with its annihilation at (x, t); this is readily verified by expansion of G(x,t;x,t) in angular momentum eigenfunctions.) Moreover taking the arbitrary gauge function ϕ0, LZ commutes with C(x,x) so Eq. (17) reads as

(18)

Correspondingly, we have

(19)

or

(20)

with the definitions

(21)

The solution may be written in a “time”-τ representation as ( Appendix A)1 

(22)

and noting that the τ-integrand is the generator of Laguerre polynomials Ln, we obtain6 (notation: γν1νγ with 1ν=±1 for KKDirac points)

(23)

In the absence of the energy shifts presently under consideration, Dirac-like Landau-quantized eigenenergies of the form for graphene are given by 𝜖n±2=(2n+11ν)γ2eB, so we write the energy denominators on the right of Eq.(23) as

(24)

Correspondingly, the eigenenergy spectrum for the case at hand is given by Dn±=0, whence (=1)

(25)

With the results of Eq.(23) for G11 and G22, the determination of G21 and G12 can be readily carried out using Eqns. (10) and (12), which may be summarized as

(26)

Using the identity7 

(Lnm(z) are generalized Laguerre polynomials) we have

(27)

Thus, G21(R;ω) is given by

(28)

where

(29)

with a corresponding adjoint relation for G12(R,ω) (involving the interchange of positional matrix elements rr so that RR, as well as complex conjugation).

For many purposes8 it is preferable to use momentum p-representation rather than position representation. Fourier transforming Eq.(22), we obtain Rp,  Appendix A)

(30)

It is also of interest to examine Eq.(20) in circular coordinates (R,𝜃):

(31)

(the factor of 2 in parentheses on the right compensates for the semi-infinite range of R=[0]). Actually, Eq.(23) exhibits G1122(R,ω)=G1122(R,ω) as a function of radius R alone, to the exclusion of angle 𝜃. Consequently, 2G1122(R,ω)/𝜃20and the angular average 02πd𝜃2π of Eq.(31) yields

(32)

This result has the form of a Bessel Wave Equation of order 0 for all R>0 (with δ(R)=0), and its homogeneous solutions are linear combinations of Bessel Wave Functions9 Z1(R) of the first kind and Z2(R) of the second kind. For an infinitesimal integration range of R[0,𝜖], it behaves as

(33)

so cancellation of the common 1R-factor and integration 0𝜖dR yields

(34)

which corresponds to

(35)

This behavior for small R𝜖 is indicative of the Z2 second Frobenius solution of the Bessel Wave Equation of order 0, whose leading term is ln𝜖. Accordingly, we indentify this solution

(36)

as an alternative to Eq.(30).

This section is focused on Green’s function dynamics of a quantum wire modelled by a 1D Dirac delta function δ(x)-potential well profile superposed on carriers in a 2D group-VI dichalcogenide material subject to a normal quantizing magnetic field. This additional potential (β<0 is the product of the potential well depth and its width)

(37)

modifies the hamiltonian in Eq.(1) above, and the associated matrix Green’s function GW-equation can be written in integral form as (ω-representation)

(38)

or (G(r1,r2;ω) is the Green’s function in the absence of the wire; suppress ω)

(39)

As in the case of G, gauge considerations mandate the presence of a Peierls phase factor in GW:

(40)

However, GW(r1,r2;ω) lacks translational invariance in the x-direction because of the δ(x)-wire potential. This complicates the full determination of GW for off-axis dynamics, but if propagation is restricted to the wire itself, x0 (x1 = x2 = 0), we have

(41)

Furthermore, G(0,y1;0,y3)=G(0,y1;0,y3) since C(r1,r3)1 for x1 = x3 = 0; and both G and GW are both manifestly translationally invariant in the y-direction. Accordingly, this integral equation is readily solved using a single Fourier transformation in the y-direction,

(42)

with the result (on y-axis, x0)

(43)

for propagation along the wire. A similar result obtains for propagation strictly parallel to the wire since the Peierls phase factor is unity when r1 and r2 are parallel, C(r1,r2)=exp(ie2r1B×r2)1. Alternatively expressed in py-representation Eq.(41) may be expressed in a more symmetric form using GW(py) from Eq.(43), as

(44)

The infinite-space Green’s function involved in the description of propagation along the wire, G(py)G(py,x0), is given in terms of its elements as follows:

For the diagonal elements,

(45)

where Dn± was defined in Eq.(24). Expanding Ln(mωc[X2+Y2]/2) in products of Hermite polynomials with X = 0 yields10 

(46)

and the Y-integral can now be carried out using related Hermite polynomials Hen(x)2n/2Hn(x/2), with the result ( Appendix B)11 

(47)

where we have defined

(48)

In regard to the off-diagonal elements, their relation to the diagonal elements in py-momentum representation (with X0) may be written as

(49)

or

(50)

The wire energy spectrum in the case of a thin wire (small β) requires Dn± to also be small–of order β– in order to satisfy the determinental relation

(51)

This is to say that a root determined by Dn+O(β) involves βG11(py)O(β0)=O(1); but Dn and Dn+ differ by order O(β0)=O(1) so Dn is correspondingly of order Dn=O(β0)=O(1) and βG22(py)O(β) is negligible. Similar considerations apply when ”+” and ”−” are interchanged along with the interchange of G11 and G22, and they also confirm the negligibility of β2G11G22G21G12 for small βAppendix C). In this case, an energy root ω is sufficiently close to the nearest pole position so that its location is determined by that pole alone, with the result (ωωn for the nth root; note that ± indicates an additional proliferation of modes, some of which may be extraneous):

(52)

or

(53)

The thermodynamic magnetic field matrix Green’s function for the group-VI dichalcogenide materials also satisfies Eqns.( 1-19) above, and it will be distinguished by an overhead bar, bold letter G¯,

(54)

with C(r,r) as the usual Peierls phase factor of Eq.(4). This thermal Green’s function G¯(r,r;T) in direct time representation has nonvanishing parts G¯>(r,r;T) and G¯<(r,r;T) for bothT>0 and T<0, respectively η+(T)=0,1forT<0,T>0, respectively, is the Heaviside unit step function):

(55)

A spectral weight matrix A(T) operating in the same spectral pseudospin subspace as G¯ may be defined in direct time representation as

and, equivalently,

(56)

with

(57)

Since both G¯>(R;T) and G¯<(R;T) satisfy the homogenous counterpart of the Green’s function equation (absent the driving term δ(2)(R)δ(T) with T 0), so does the spectral weight function A(R,T). The latter is subject to the sum rule (due to the canonical commutation/anticommutation relations)1,12,13

(58)

where the spatial delta function mandates that r=rinC(r,r)C(r,r)1.

In frequency representation, the spectral weight function (matrix) provides the ”greater” (G¯>) and ”lesser” (G¯<) parts of the thermodynamic Green’s function, corresponding to t>t and t<t, respectively, as12,13

(59)

where f(ω) is the Fermi (or Bose) statistical distribution function (± here refers to bosons (upper sign) or Fermions (lower sign)). As it is often preferable to work in p-momentum representation, we consider the A11(p,T) and A22(p,T) elements of A(p,T), which must satisfy the homogenous counterpart of the Green’s function equation

(60)

subject to the sum rule Eq.(58) in momentum representation

(61)

The solution of Eqns. (60) and (61) is given by1 

(62)

and from Eq.(59) with Rpsett¯=Ωc±T/2 ,

(63)

It is immediately clear that the periodic cos(t¯) and tan(t¯) functions in Eqns.(62, 63) may be written in a Fourier series, producing a row of frequency δ-functions in ω-representation, which will place the eigenmode energies in the arguments of the statistical f(ω) -functions in Eq.(63). This may be seen from yet another point of view, since the spectral weight matrix function A(r,r;ω) can also be determined from the retarded Green’s function G(r,r;ω) discussed above as13 

(64)

Employing Eq.(23) for G(R,ω+i0+) we obtain the diagonal elements of A1122(r,r;ω) as

(65)

Again, the off-diagonal elements can be readily obtained using Eqns.(10,12).

Furthermore, in momentum representation A1122is given by (Eq. (23); Eq. (A.2.5))

(66)

In dealing with the position/momentum representations of the thermal Green’s function and spectral weight as 2×2 pseudospin matrices, it is should be borne in mind that the important spectral weight function trace A in diagonal energy representation is given by the collection of all energy state contributions, and since the trace is invariant with respect to change of basis/representation, it is equally well given in the pseudospin position/momentum matrix representations by the trace of Eqns.(65,66) whence

(67)

excluding the off-diagonal elements. With this, the trace of Eq.(59) provides the associated thermal Green’s function trace, subject to the addition of the other spin/valley contributions.

This work has addressed the matrix Green’s function G for group-VI dichalcogenide materials in the presence of a normal, quantizing magnetic field. After seperating the Peierls phase factor, the translationally invariant part of the retarded Green’s function G is derived in position/frequency representation in a closed form integral representation involving only elementary functions, Eq.(22) (diagonal elements, from which the off-diagonal elements are easily determined by Eqns.(10,12)): This integral representation, Eq.(22), is also expanded in an eigenfunction series of Laguerre polynomials, explicitly exhibiting both the diagonal (Eq.(23)) and off-diagonal (Eqns.(28,29)) elements as functions of position and frequency and identifying the energy eigenvalues (Eq.(25)). It is also of interest to present the retarded dichalcogenide Green’s function (diagonal elements) in direct time T-representation (as opposed to the τ -”time” transformation variable of convenience employed in Eq.(22) as conjugate to Ω): This is given by Fourier ω-integration of Eq.(23) as

(68)

As momentum representation is of great importance in more complicated calculations, the closed form integral representation of G is also determined as a function of momentum and frequency (Eq.(30)). Yet another representation of G is derived in circular coordinates in terms of the second Frobenius solution of the Bessel wave equation of order 0 in Eq.(36).

This work has also addressed the presence of a quantum wire impressed on the dichalcogenides by a model one-dimensional δ(x)potential profile. The associated retarded Green’s function integral equation for this system is solved exactly for propagation directly along the wire (Eqns.( 44-50)) in terms of the corresponding Green’s function for the system without the wire δ(x)potential, and the eigenenergy dispersion relation is determined (Eq.(51)) and solved approximately (Eq.(53)).

We have also formulated the thermodynamic Green’s function (matrix) for the group-VI dichalcogenides in terms of its spectral weight, exhibiting its solution in a momentum/frequency integral representation in terms of elementary functions alone (Eq.(63)); moreover, using a Laguerre polynomial expansion, it is also presented in both position/frequency and momentum/frequency representations (Eqns.(65,66)). (Important mathematical issues relating to the solution and integrations are presented in  Appendices A and  B, respectively.)

In all representations considered here, for both the homogeneous infinite system and the quantum wire model as well, the principal affect of the magnetic field is the splintering of the energy spectrum into a proliferation of discrete Landau-quantized energy states, as was the case for graphene and a quantum dot model study done previously.5 Of course, the determination of the Green’s functions in these cases also provides detailed information about the pertinent carrier propagation characteristics, including the Peierls phase factor and orbit curvature, as well as the energy eigenstate spectra.

Prof. M.L. Glasser has been helpful in resolving some mathematical issues that arose in the course of this work.

For convenience, we write Eq.(20) in momentum representation1Rp,

(A1)

and returning to a ”time”-τ representation we have (τ is a ”time” variable conjugate to Ω under Fourier transformation)

(A2)

Here, Ω is viewed as the conjugate frequency variable and

(A3)

G1122(px,py,τ) obeys the homogeneous counterpart of Eq.(B) for τ>0(δ(τ)=0) and the homogeneous solution has the form1 (k1,k2 are constants)

(A4)

Integrating Eq.(B) from τ=0 to τ=0+, the initial condition for retardation requires that G1122(px,py;τ=0+)=i, whence we obtain

(A5)

and

(A6)

Fourier transformation to position representation yields (px,pyX,Y)

(A7)

Here, we note that inverting the Fourier transform from τrepresentation to ω-representation, G1122(R,Ω), returns us to the original physical quantity represented by the function G1122(R,Ω)=G1122(R,ω)(although mathematicians would write it differently).

A. The integral I1=dYeipyYexpeB4Y2H2keB2Y:

Introducing the related Hermite polynomials Hen(x)=2n/zHnx2, and changing the scale of the integration variable x¯¯=eB2Y, we have

(B1)

This integral is evaluated in reference 11 as11 

(B2)

B. The integral I2=dzReipReeB4R2LneB2X2+Y2:

The Laguerre function may be expanded in Hermite polynomials separating the variables of integration as10 

(B3)

whereupon the double integral separates completely:

(B4)

These one-dimensional integrals can be evaluated using Eqns.(B1, 2) above and the resulting sum of products of Hermite polynomials can be re-assembled into the Laguerre polynomial using Eq.(B3), with the result

(B5)

The determinental dispersion relation takes the form

(C1)

which may be written out more fully for small β as

(C2)

For β small, then a denominator Dn+ or Dn must be order O(β) to obtain a root. However, Dn+ and Dn differ by order O(β0)1, so if one is of order O(β) then the other is of order O(β0)1, and the term associated with the latter is smaller than that of the former because its denominator is larger by a factor of β. Extending these considerations to the last term on the right involving the denominator product Dn+DnO(β) shows that the last term is of order O(β2/β)=O(β), and it can be neglected for small β.

1.
N.
Horing
,
Annals of Physics (NY)
31
,
1
(
1965
);
2.
N. J. M.
Horing
and
M.
Yildiz
,
Annals of Physics (NY)
97
,
216
(
1976
).
3.
4.
T. M.
Rusin
and
W.
Zawadzki
,
J. Phys. A
44
,
105201
(
2011
);
arXiv:1612.03944v1 [cond-mat.mes-hall] (
12 Dec 2016
).
5.
N. J. M.
Horing
and
S. Y.
Liu
,
J. Phys. A
42
,
225301
(
2009
).
6.
A.
Erdelyi
, et al. ”
Higher Transcendental Functions, Bateman Manuscript Project
”, Vol. 2, p.
189 #17
(
1953
).
7.
A.
Erdelyi
, et al. ”
Higher Transcendental Functions, Bateman Manuscript Project
, Vol. 2, p.
189 #15
(
1953
).
8.
T.
Ando
,
J. Phys. Soc. Japan
74
,
777
(
2005
).
9.
P.
Moon
and
E.
Spencer
, ”
Field Theory Handbook
”, 2nd Ed.,
Springer Verlag
(
1971
), pp
171
, also p.154 bottom.
10.
A.
Erdelyi
, et al. ”
Higher Transcendental Functions, Bateman Manuscript Project
” Vol. 2, p.
195#32
; p.
193 #15
.
11.
A.
Erdelyi
, et al. ”
Tables of Integral Transforms, Bateman Manuscript Project
”, Vol. 1, p.
39#7
(
1954
).
12.
P. C.
Martin
and
J.
Schwinger
,
Phys. Rev.
115
,
1342
(
1959
).
13.
N. J. M.
Horing
, ”
Quantum Statistical Field Theory
”,
Oxford University Press
(
2017
, in press).