You are on page 1of 9

Superlattices and Microstructures 85 (2015) 401–409

Contents lists available at ScienceDirect

Superlattices and Microstructures


journal homepage: www.elsevier.com/locate/superlattices

Spin–orbit interaction and magnetic field effects on the energy


dispersion of double quantum wire
Y. Karaaslan a,⇑, B. Gisi a, S. Sakiroglu b, E. Kasapoglu c, H. Sari c, I. Sokmen b
a _
Physics Department, Graduate School of Natural and Applied Sciences, Dokuz Eylül University, 35390 Izmir, Turkey
b _
Physics Department, Faculty of Science, Dokuz Eylül University, 35390 Izmir, Turkey
c
Physics Department, Faculty of Science, Cumhuriyet University, 58140 Sivas, Turkey

a r t i c l e i n f o a b s t r a c t

Article history: We investigate the effects of Rashba spin–orbit interaction on the electronic energy
Received 24 April 2015 dispersion and zero-temperature ballistic conductance of double quantum wire under
Received in revised form 1 June 2015 the influence of perpendicular magnetic field. The wire system is represented by a
Accepted 4 June 2015
symmetric, double quartic-well confinement potential. Numerical results reveal that
Available online 5 June 2015
competing effects between spin–orbit interaction and magnetic field modify strongly the
energy subband structure and introduce anticrossings between subbands. Moreover, it is
Keywords:
found that the conductance character of a quasi-one-dimensional ballistic conductor,
Double quantum wire
Anharmonic potential
which is closely related to the energy-dispersion spectrum, is very sensitive to the shape
Rashba spin–orbit interaction of potential profile, magnetic field and Rashba spin–orbit coupling.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction

The examination of electronic and optical properties of low-dimensional semiconductor structures formed by restricting
the motion of charge carriers in the bulk structures, offers significant contributions to the evolving product technology of
today. In the last two decades, with the advances in fabrication technology low-dimensional double semiconductor
structures as well as single ones (such as quantum wells, quantum wires and quantum dots) are produced [1].
The one-dimensional double well potential (called as quadratic anharmonic double well oscillator, double minimum
problem) in physics and chemistry is a well-known problem that have been studied for many years. This potential has been
used to explain the many physical phenomena such as tunneling and doublet splitting by using quantum mechanical or
semi-classical methods [2–7]. In these studies, usually WKB approximation, variational perturbation method etc. have been
used for calculation of energy eigenvalues. In addition to these, vibrational-energy eigenvalues, eigenfunctions, transition
moments, and absorption intensities have been investigated in detail with model of H bonds for symmetric and asymmetric
double well potentials in the study given with Ref. [3]. Furthermore, there are also many studies about the inverse structure
of ammonia molecule ðNH3 Þ with this potential function in the literature [2,8].
Many interesting characteristics has been reported in double well and double-wire system. In the double-well potential
system studies exhibit that if the two wells are sufficiently separated from each other, the lower energy eigenvalues are clo-
sely bunched as a couple [9–11]. Similar to this occurrence was obtained also in the study of double quantum wire structure
(if the potential barrier is height enough) in Ref. [12]. Several theoretical [12–14] and experimental [15,16] studies investi-
gated the effects of external fields on the electronic and transport properties of coupled double quantum wire structure

⇑ Corresponding author.
E-mail address: yenalkaraaslan@gmail.com (Y. Karaaslan).

http://dx.doi.org/10.1016/j.spmi.2015.06.002
0749-6036/Ó 2015 Elsevier Ltd. All rights reserved.
402 Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409

which is created with the separation of two identical narrow quantum wires by a potential barrier. In particular, Lyo and his
colleagues have investigated mobility, conductivity and magnetoresistance of the coupled double quantum wires subjected
to magnetic field [17,18]. Morever, Korepov and Liberman have searched theoretically different physical properties (such as
transport, spectrum of impurity states etc.) of two parallel quantum wires coupled by tunneling in an external perpendicular
magnetic field [19,20]. Recently, magnetotransport properties of window-coupled double quantum wire have been investi-
gated by Tang et al. [21,22].
On the other hand, the electron spin as well as the electron charge started to taken into account for information
processing in the examinations of low-dimensional semiconductor structures. New devices based on manipulation
and control of electron spin have the characteristics of more powerful, smaller and faster than those of existing
[23–27]. The spin orientation in a semiconductor structure can generally be adjusted by the Rashba spin–orbit coupling
(SOC) effect which arises due to the asymmetry associated with the confinement potential [28]. The effect of Rashba
spin–orbit coupling can be varied by gate electric field and magnitude of coupling parameter can be determined from
the characteristic beating pattern of the Shubnikov-de Haas oscillations in the magnetoresistence of two-dimensional
electron gas [29,30].
Several theoretical and experimental works investigated the effects of external fields and spin–orbit coupling on the
electronic and transport properties in low-dimensional systems [31–42]. Mireles and Kirczenow [43] researched the ballistic
spin-transport properties of quasi-one-dimensional wires in the presence of the Rashba spin–orbit interaction, Governale
and Zülicke [44] presented the effect of Rashba spin–orbit coupling on the band structure, spin accumulation and transport
properties of quantum wires. Influence of a perpendicular magnetic fields on the spectral, spin and transport properties
of semiconductor quantum wires with Rashba SOC has been investigated in Ref. [31,45,46]. Serra and coworkers reported
the effects of in-plane magnetic field [47] whereas Su et al. focused on an inhomogeneous magnetic field [48] effects in
quantum wires considering SOC.
Despite there are several works reporting the effects of magnetic field on electronic energy spectrum and conductance in
tunnel-coupled (double) quantum wires [17], to our knowledge, less attention has been paid on the influence of spin–orbit
coupling in double quantum wires. In the present work we will focus energy spectra and ballistic conductance quantum wire
defined by a non-parabolic confinement potential, subjected to an external perpendicular magnetic field and taking into
account the Rashba spin–orbit interaction. The organization of this work is as follows. Firstly in Section 2, we briefly describe
the system and present the methodology used throughout our study. Section 3 is devoted to a summary of numerical results
which is followed by a short concluding section.

2. Theoretical model and formalism

We consider a quasi-one-dimensional double quantum wire subjected to a perpendicular magnetic field as shown in
Fig. 1. The system is imagined to be confined by a double-well potential given as
 
1 l2 2
VðxÞ ¼ k x2  ; ð1Þ
4 k
where l and k are positive, adjustable structural parameters controlling the height of the barrier between wells and the
width of wells. The double-quantum wire system is chosen on the xy-plane therefore the electrons can move freely along
the y-direction. The orientation of applied magnetic field is assumed to be in the z-direction, ~
B ¼ ð0; 0; BÞ. Thus, the vector
potential corresponding to this field can be chosen as ~ A ¼ ð0; Bx; 0Þ in Landau gauge. Under these conditions, the
single-particle Hamiltonian for a quasi-one-dimensional double-quantum wire taking into account the Rashba spin–orbit
interaction is
1 h 2 i 1
H¼ p þ ðpy þ eBxÞ2 r0 þ VðxÞr0 þ g  lB Brz þ HR : ð2Þ
2m x 2

Fig. 1. (color online) Schematic representation of double quantum wire system.


Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409 403

Here, the first term is kinetic contribution where m is the effective mass of electron, px and py are components of electron
momentum, e is the electron charge and r0 is 2  2 unit matrix. Second term is confining potential contribution whereas
the third term is Zeeman contribution with g  being effective Lande-g factor, lB Bohr magneton and rz standing for the
z-component of Pauli spin matrix. HR term describes Rashba spin–orbit interaction that is given by

aR  
HR ¼ rx ðpy þ eBxÞ  ry px : ð3Þ
h

Here aR is the Rashba spin–orbit coupling factor that can be varied by gate electric field, and rx and ry are the x- and
y-components of Pauli spin matrix, respectively.
Due to translational invariance along the y-direction, the energy eigenfunctions of Hamiltonian can be to written in terms
of plane-waves as

Wðx; yÞ ¼ /ðxÞ expðiky yÞ; ð4Þ

where ky , the wave numbers of the plane-wave, is good quantum number of the system. Therefore, Hamiltonian
ðH ¼ HI þ HII þ HR Þ can be written as:
" #
h2 d2 1  2 2
h2 k2y x20 1
HI ¼  þ m x ð x  x 0 Þ r 0 þ r0 þ g  lB Brz ;
2m dx2 2 2m x2 2

1  1 1l 4
HII ¼  m x20 þ l2 x2 þ kx4 þ r0 ;
2 4 4 k

eB d
HR ¼ aR rx ðky þ xÞ þ iry : ð5Þ
h dx
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Variables defined are as follows: x0 stands for harmonic oscillator, x ¼ x20 þ x2c is the effective oscillator and with
xc ¼ eB=m is the cyclotron frequency, respectively. x0 ¼ l0 xco x2ow j0 is the guiding center coordinate with
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
l0 ¼ h=m x0 (being the characteristic length of the harmonic oscillator with frequency x0 ), xow ¼ x0 =x; xco ¼ xc =x0 ,
and j0 ¼ l0 ky .
The set of eigenfunctions of HI Hamiltonian obtained from the Schrödinger equation HI /nr ðxÞ ¼ EInr /nr ðxÞ is given as
follows:
   2 !
1 x  x0 1 x  x0
/nr ðxÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffi ffi H n p ffiffiffiffiffiffiffiffiffi exp  p ffiffiffiffiffiffiffiffiffi vr ; ð6Þ
l0 pxow 2n n! l0 xow 2 l0 xow
 
with n ¼ 0; 1; 2; . . . ; r ¼ ; Hn ðxÞ are the Hermite polynomials of integer order n; vr are spinor functions for vþ ¼ 10
 
0
spin-up and v ¼ spin-down in the z-direction. The energy eigenvalues corresponding to HI are
1
 
EIn 1 1 x2 DZ
¼ nþ þ ow j20  :
hx0 xow 2 2 2

In this expression, DZ ¼ g  lB B=hx0 describes the Zeeman energy splitting contribution. In this work, we chose l0 , the
characteristic length of the harmonic oscillator, as length scale and correspondingly  hx0 as energy scale. By expanding
P
/ðxÞ ¼ n;r an;r /n;r ðxÞ in terms of eigenfunctions of HI ; /n;r ðxÞ, Schrödinger equation corresponding to Eq. (5) takes the form
X 0 X 0
ðEInr  EÞanr þ ðHII Þrr
nm amr0 þ ðHR Þrr
nm amr0 ¼ 0: ð7Þ
m;r0 ¼r m;r0 –r

0
In the above equation, matrix elements of the second term on the left ðHII Þrr
nm ¼ h/nr jHII j/mr0 i are expressed as
follows:
0
ðHII Þrr
nm
¼ K m4 dn;m4 þ K m3 dn;m3 þ K m2 dn;m2 þ K m1 dn;m1 þ K 0 dn;m þ K p1 dn;mþ1 þ K p2 dn;mþ2 þ K p3 dn;mþ3 þ K p4 dn;mþ4 : ð8Þ
hx0

We should note that, in the each term dr;r0 factor exists, but it is not written here. In addition, the K factors in Eq. (8) are given
as the following:
404 Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409

1 ~ 2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K m4 ¼ kxow ðn þ 4Þðn þ 3Þðn þ 2Þðn þ 1Þ;
16
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K m3 ¼ ~kx~0 xow 2 xow ðn þ 3Þðn þ 2Þðn þ 1Þ;
4    
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3  
K m2 ¼ xow ðn þ 2Þðn þ 1Þ k~ 3x~0 2 þ xow n þ  1þl ~2 ;
4 2
  
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3  
K m1 ¼ x~0 2 xow ðn þ 1Þ k~ x~0 2 þ xow ðn þ 1Þ  1 þ l ~2 ;
2 2
      
1 l~ 4 1 ~ 3 2 2 1 1  1
K 0 ¼ ~k x~0 4 þ 3xow x~0 2 ð2n þ 1Þ þ þ k xow n þ n þ ~ 2 x~0 2 þ xow n þ
 1þl ;
4 ~k2 4 2 2 2 2
  
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 3  
K p1 ¼ x~0 2 xow n k~ x~0 2 þ xow n  1 þ l ~2 ;
2 2
   
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 1  
~
K p2 ¼ xow nðn  1Þ k 3x~0 2 þ xow n   1þl ~2 ;
4 2
1 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K p3 ¼ ~kx~0 xow 2 xow nðn  1Þðn  2Þ;
4
1 ~ 2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
K p4 ¼ kxow nðn  1Þðn  2Þðn  3Þ;
16
lffi
with x~0 ¼ xl00 , l~ ¼ pffiffiffiffi
m x0
k ¼ khx20 2 .
and ~
ðm x Þ 0
0
The matrix elements of Rashba Hamiltonian ðHR Þrr
nm ¼ h/nr jHR j/mr0 i in Eq. (7) are given by
qffiffiffiffiffiffiffiffiffi
ðHR Þ
nn

~ R j0 ð1  x2 x2 Þ ;

¼ 2D co ow
hx0
sffiffiffiffiffiffiffiffiffi" rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi rffiffiffi #
ðHR Þ 2D ~R ðn þ 1Þ n
nm
¼ ðxco xow  1Þ dn;m1 þ ðxco xow  1Þ dn;mþ1 ; ð9Þ
hx0 xow 2 2
R  2
~ R ¼ DSO with DR ¼ m a2R .
with D h x0
 SO 2
h
The energy dispersion relations and eigenfunctions in double-quantum wire with Rashba spin–orbit interaction can be
obtained by diagonalization of Eq. (7) for different values of an applied external perpendicular magnetic field. Obtained
results are presented in the next section.

3. Results

3.1. Energy dispersion

In our calculations, we chose the characteristic length l0 corresponding to confinement potential as a length scale and
consequently energy scale becomes to be hx0 . The parameters used in our numerical work are adapted as: the effective mass
of electron is m ¼ 0:03m0 (as in InAs) with m0 being the free-electron mass, the effective Lande-g factor g  ¼ 15 [35,36].
Moreover, we admit that the system is exposed to the confining potential with hx0 ¼ 2meV. For the sake of simplicity, one of
~ ¼ 1.
structural parameters controlling the shape of the double-well potential is fixed, k
In order to show the effect of structural parameters, we begin with the study of eigenenergy as a function of ky l0 in zero
magnetic field and Rashba contribution. In the absence of spin–orbit coupling the single-electron spectrum is split into sub-
bands having quadratic dispersion in wavevector ky l0 . It is clear from Fig. 2 that the energy band levels exhibit degeneracy in
pairs by starting from the lowest with increasing values of l ~ . It is worthwhile to note that for higher values of parameter l
~,
the system can be regarded as weakly-coupled quantum wires which results in parabolic dispersion curves formed by the
overlap of individual wires. This observation is consistent with the literature [12,9–11].
R 2
 =2m aR . Competing effects between
Turning on the spin–orbit effect introduces spin-related characteristic length lSO ¼ h
R
confinement and Rashba contribution allows us to define three different regimes: weak for cases lSO  l0 , moderate for
R R
lSO  l0 and strong for lSO < l0 . As seen in Fig. 3(a) and (b), Rashba spin splitting removes degeneracy except ky l0 ¼ 0 and
causes wavevector-shifted dispersion curves. Increment in Rashba spin–orbit coupling contribution (DRSO =
hx0 – 0) results
in deviation from the parabolic structure in energy bands and anticrossings between neighboring subbands emerge. We also
should note that, for strong SOC regime (Fig. 3(c)) camelback-like structure in the energy dispersion diagrams around
ky l0 ¼ 0 appears. Dashed curves in the figure denote the case of DRSO =
hx0 ¼ 0.
Investigating the effect of an external perpendicular magnetic field (xc =x0 – 0) in the absence of the Rashba SOC is
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
important. Comparison of the characteristic length of the harmonic oscillator with the magnetic length (lc ¼ h  =m xc ),
allows us to distinguish two regimes for magnetic field: weak (lc  l0 ) and strong (lc < l0 ) regime. Fig. 4 shows that exitance
Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409 405

Fig. 2. (color online) The energy dispersion relations of double quantum wire in the absence of magnetic field and Rashba spin–orbit coupling, (a) dashed
lines stand for l
~ ¼ 0:5 and solid lines for l
~ ¼ 1:5, (b) dashed lines stand for l
~ ¼ 2:0 and solid lines for l
~ ¼ 3:0.

Fig. 3. (color online) The energy band structure of double quantum wire in the absence of magnetic field, (a) for weak regime DRSO =
hx0 ¼ 0:01, (b) for
moderate regime DRSO =hx0 ¼ 0:1, (c) for strong regime DRSO =
hx0 ¼ 0:5.

Fig. 4. (color online) The energy spectrum of double quantum wire subjected to perpendicular magnetic field without considering of Rashba spin–orbit
coupling contribution. Variation in case of l
~ ¼ 1 (a) at xc =x0 ¼ 0:5, (b) at xc =x0 ¼ 2 and in case of l
~ ¼ 2:5 the corresponding figures are (c) and (d).

of magnetic field lifts the degeneracy in each energy band by the Zeeman effect while parabolic shape of subbands preserves
at weak strengths of the field. With further increment of xc =x0 – 0 deviation from parabolicity is leaps out. Dashed curves
in the figure denote the case of zero-magnetic field for DRSO =
hx0 ¼ 0. Separation between spin branches enhances with
406 Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409

increasing values of the magnetic field, and more flattened structures with distortion of the parabolic structure around
ky l0 ¼ 0 stand out particularly in lower energy levels. Larger values of structural parameter l ~ cause significant variation
on the energy dispersion that is depicted in Fig. 4(c) and (d). Narrow energy gaps at the oscillatory crossing points of two
sets of dispersion spectra are becomes evident. Meanwhile separation between the two sets enhances with magnetic field
can be seen.
To elucidate the interplay between SO interaction and uniform magnetic field, in Fig. 5 we present energy level spectrum
of an electron in double quantum wire as a function of xc =x0 at ky l0 ¼ 0. When only the contribution of the magnetic field is
considered, we can impart that the applying of perpendicular field results in enhancement in confinement potential with
subband separation while lifting degeneracy in each subband. Simultaneous contribution of magnetic field and Rashba SO
interaction gives rise to varying remarkable separation in spin-up and spin-down branches of the energy states. Higher mag-
netic field values lead to a complicated behavior in the energy spectrum, in comparison with zero field case, while causing
anticrossings between the subbands.
It is useful at this stage to inquire to what extent the interplay between spin–orbit interaction and magnetic field affect
the system. In Fig. 6 we plot the variation of energy spectrum for three different values of magnetic field by using a fixed
value of Rashba SO coupling parameter (DRSO = hx0 ¼ 0:5). Even in case of weak magnetic field (Fig. 6(a)) spin degeneracy
at ky l0 ¼ 0 point is removed, camel-back shape of the lower subbands is preserved whereas strong anticrossings/crossings
in higher energy levels are clearly seen (Fig. 6(b) and (c)). With an increment in magnetic field, subband structure undergo
substantial modifications by becoming to a parabolic-like form. In this case external field dominates over the Rashba SO
interaction and eventuates in significant softening in anticrossings. Distinctive impact of the adjustable structural parameter
l~ is seen in Fig. 6(d). Magnetic field removes the degeneracy giving rise to a rather complicated spectra.
Electronic structure and transport properties of low-dimensional systems can be altered by the size and shape of confin-
~ which plays role (in conjunction with ~
ing potential. Therefore, in Fig. 7 we investigate the effect of structural parameter l k
that is fixed to 1) in controlling the shape of the confinement potential. When perpendicular magnetic field and Rashba SO

Fig. 5. (color online) The energy dispersion at ky l0 ¼ 0 as a function of xc =x0 , (a) for DRSO =
hx0 ¼ 0:0, (b) for DRSO =
hx0 ¼ 0:5.

Fig. 6. (color online) The energy band spectrum belonging to double quantum wire structure, (a) for xc =x0 ¼ 0:1 and DRSO = hx0 ¼ 0:5, (b) for xc =x0 ¼ 0:3
and DRSO =
hx0 ¼ 0:5, (c) for xc =x0 ¼ 1:0 and DRSO =
hx0 ¼ 0:5, and (d) for the same parameters as in (c) but in case of l
~ ¼ 2:5.
Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409 407

~ , (a) for xc =x0 ¼ 0:0 and DRSO =


Fig. 7. (color online) The energy dispersion at ky l0 ¼ 0 as a function of l hx0 ¼ 0:0, (b) for xc =x0 ¼ 1:0 and DRSO =
hx0 ¼ 0:0, (c)
for xc =x0 ¼ 1:0 and DRSO =
hx0 ¼ 0:5.

coupling is chosen as zero, degenerate states at ky l0 are observed and with an increasing value of l
~ this degeneracy becomes
stronger. Turning on the magnetic field results in separation of spin-up and spin-down states. However, for higher values of
l~ degeneracy is manifested again. Stronger values of l~ cause ascent in bump at the center of the wire which consequently
results in weak-coupling of the dispersion relations of non-parabolically confined quantum wire. We should note that sep-
aration distance of the two sets of dispersion curves is proportional to the strength of the applied magnetic field.

3.2. The conductivity

In this part of the paper, we aim to calculate the influence of the interplay between spin–orbit interaction and magnetic
field on the zero-temperature conductance. Conductivity of a ballistic nanostructure connected to two macroscopic electron
reservoirs, by neglecting the electron–electron interaction and considering zero-temperature, can be calculated using
Landauer–Büttiker formalism [49,50],
X
G ¼ G0 T aa0 ; ð10Þ
aa0

where T aa0 is the transition probability from jai state to ja0 i state and G0 ¼ 2e2 =h is the conductance quantization. We follow
the strategy of Pershin and colleagues [51,52] wherein a small bias (lL > lR ) is applied between the contacts (having lL and
lR chemical potentials) at the ends of quantum wire. Under the assumption of non-scattering and unity for the transition
probability, the conductance of such a system can be calculated as follows:
e2 XX ðn;sÞ n;s
G¼ b f ðEi Þ: ð11Þ
h n;s i i
th
Here n and s represent the level of state and the level of spin, respectively while En;s
i is the energy of i extremum point in the
ðn;sÞ
related energy subband and f ðEn;s
i Þ is the Fermi–Dirac distribution function. corresponds to 1 for a maximum and þ1
bi
for a minimum point in the energy subband labeled with n and s. Thereby, we expect that the minimum (maximum) point in
the energy bands to increase (reduce) the conductance.

Fig. 8. (color online) The conductance as a function of the Fermi energy, (a) for xc =x0 ¼ 0:0 and DRSO =
hx0 ¼ 0:1, (b) for xc =x0 ¼ 0:5 and DRSO =
hx0 ¼ 0:0, (c)
for xc =x0 ¼ 0:1 and DRSO =
hx0 ¼ 0:5.
408 Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409

In Fig. 8 ballistic conductance of double quantum wire is depicted as a function of Fermi energy. When only a Rashba SO
coupling of DRSO ¼ 0:1
hx0 is considered, the conductance exhibits an increasing stepwise profile with integer conductance
plateaus for both values of structural parameter l
~ . As seen from this figure, each minimum point in the energy subbands
given in Fig. 3(b) provides increment to the conductivity of e2 =h. Fig. 8(b) gives information of conductivity corresponding
to energy diagram in Fig. 4(a) considering only magnetic field in the absence of Rashba spin–orbit interaction. Emergency of
weak peaks and dips on quantization steps which display monotonic increasing feature is standing out. This situation is more
pronounced for higher value of l ~ parameter. When both magnetic field and spin–orbit interaction are taken into account,
possession of camel-back structure in the energy subbands with extreme points more than one, induces a narrow dips in
conductance steps (Fig. 8(c)). These phenomena can be attributed to the energy dispersion of the electron which can be
tuned by the spin–orbit coupling and the magnetic field.

4. Conclusion

In this study, we have investigated the combined effect of Rashba spin–orbit interaction and an external magnetic field
applied perpendicularly to a quantum wire. The electronic energy spectrum and the ballistic conductance of the wire
described by a double-well anharmonic potential are calculated. Our numerical results reveal that the energy subband
dispersion can be modulated by structural parameters of the confining potential, strengths of magnetic field and spin–orbit
coupling. Moreover, we conclude that the ballistic magnetoconductance oscillations are very sensitive to spin–orbit-induced
modifications in the wire subband structure and to magnetic field. Results of the present work may assist in understanding
the transport and optical properties of quasi-one-dimensional structures.

Acknowledgements

One of the authors (Y. Karaaslan) would like to thank the Scientific and Technological Research Council of Turkey
_
(TÜBITAK _
BIDEB 2211) for partial financial support. This work has been completed at the Dokuz Eylül University,
Graduate School of Natural and Applied Sciences and is the subject of the forthcoming Ph.D. Thesis of Yenal Karaaslan.

References

[1] J.H. Davies, The Physics of Low-Dimensional Semiconductors: An Introduction, Cambridge University Press, Cambridge, England, 1998.
[2] D.M. Dennison, G.E. Uhlenbeck, Phys. Rev. 41 (1932) 313.
[3] R.L. Somorjai, D.F. Horning, J. Chem. Phys. 36 (1962) 1980.
[4] S.B. Yuste, A.M. Sánchez, Phys. Rev. A 48 (1993) 3478.
[5] M.R.M. Witwit, J. Comput. Phys. 123 (1996) 369.
[6] C.R. Handy, Phys. Rev. A 46 (1992) 1663.
[7] S.K. Lyo, Phys. Rev. B 50 (1994) 4965.
[8] M.F. Manning, J. Chem. Phys. 3 (1935) 136.
[9] K. Banerjee, S.P. Bhatnagar, Phys. Rev. D 18 (1968) 4767.
[10] R. Balsa, M. Plo, J.G. Esteve, A.F. Pacheco, Phys. Rev. D 28 (1983) 1945.
[11] R.J.W. Hodgson, Y.P. Varshni, J. Phys. A: Math. Gen. 22 (1989) 61.
[12] J.-R. Shi, B.-Y. Gu, Phys. Rev. B 55 (1997) 9941.
[13] S.K. Lyo, J. Phys.: Condens. Matter 8 (1996) L703.
[14] D.-W. Wang, E.G. Mishchenko, E. Demler, Phys. Rev. Lett. 95 (2005) 086802.
[15] J.S. Moon, M.A. Blount, J.A. Simmons, J.R. Wendt, S.K. Lyo, J.L. Reno, Phys. Rev. B 60 (1999) 11530.
[16] M. Yamamoto, Y. Tokura, Y. Hirayama, M. Stopa, S. Tarucha, AIP Conf. Proc. 772 (2005) 25.
[17] S.K. Lyo, D. Huang, Phys. Rev. B 64 (2001) 115320.
[18] D. Huang, S.K. Lyo, K.J. Thomas, M. Pepper, Phys. Rev. B 77 (2008) 085320.
[19] S.V. Korepov, M.A. Liberman, Physica B 322 (2002) 92.
[20] S.C. Arapan, S.V. Korepov, M.A. Liberman, B. Johansson, Phys. Rev. B 67 (2003) 115328.
[21] V. Gudmundsson, C.-S. Tang, Phys. Rev. B 74 (2006) 125302.
[22] N.R. Abdullah, C.-S. Tang, V. Gudmundsson, Phys. Rev. B 82 (2010) 195325.
[23] S. Datta, B. Das, Appl. Phys. Lett. 56 (1990) 665.
[24] L. Zhang, P. Brusheim, H.Q. Xu, Phys. Rev. B 72 (2005) 045347.
[25] X.F. Wang, Phys. Rev. B 69 (2004) 035302.
[26] A. Dargys, Acta Phys. Pol. A 113 (2008) 937.
[27] M. Governale, U. Zlicke, Solit State Com. 131 (2004) 581.
[28] E.I. Rashba, Sov. Phys. Solid State 2 (1960) 1109.
[29] J. Luo, H. Munekata, F.F. Fang, P.J. Stiles, Phys. Rev. B 38 (1988) 10142;;
J. Luo, H. Munekata, F.F. Fang, P.J. Stiles, Phys. Rev. B 41 (1989) 7685.
[30] Th. Schäpers, J. Knobbe, V.A. Guzenko, Phys. Rev. B 69 (2004) 235323.
[31] A.V. Moroz, C.H.W. Barnes, Phys. Rev. B 60 (1999) 14272;;
A.V. Moroz, C.H.W. Barnes, Phys. Rev. B 61 (2000) R2464.
[32] S. Zhang, R. Liang, E. Zhang, L. Zhang, Y. Liu, Phys. Rev. B 73 (2006) 155316.
[33] F. Malet, M. Pi, M. Barranco, L. Serra, E. Lipparini, Phys. Rev. B 76 (2007) 115306.
[34] T.-Y. Zhang, W. Zhao, X.-M. Liu, J. Phys.: Condens. Matter 21 (2009) 335501.
[35] S. Bandyopadhyay, S. Pramanik, M. Cahay, Superlattices Microstruct. 35 (2004) 67;;
S. Bandyopadhyay, S. Pramanik, M. Cahay, Phys. Rev. B 76 (2007) 155325.
[36] P. Upadhyaya, S. Pramanik, S. Bandyopadhyay, Phys. Rev. B 77 (2008) 155439;
S. Gujarathi, K.M. Alam, S. Pramanik, Phys. Rev. B 85 (2012) 045413.
[37] M. Kumar, S. Lahon, P. Kumar Jha, M. Mohan, Superlattices Microstruct. 57 (2013) 11.
Y. Karaaslan et al. / Superlattices and Microstructures 85 (2015) 401–409 409

[38] S. Sarikurt, S. Sakiroglu, K. Akgungor, I. Sokmen, Chin. Phys. B 23 (2014) 017102.


[39] R. Khordad, J. Luminescence 134 (2013) 201.
[40] F. Malet, P. Martí, M. Baaranco, L. Serra, E. Lipparini, Phys. Rev. B 76 (2007) 115306.
[41] X. Fu, Z. Chen, F. Zhong, G. Zhou, Physica B 405 (2010) 4785.
[42] P.M. Krstajić, M. Pagano, P. Vasilopoulos, Physica E 43 (2011) 893.
[43] F. Mireles, G. Kirczenow, Phys. Rev. B 64 (2001) 024426.
[44] M. Governale, U. Zülicke, Phys. Rev. B 66 (2002) 073311;;
M. Governale, Solid State Commun. 131 (2004) 581.
[45] S. Debald, B. Kramer, Phys. Rev. B 71 (2005) 115322.
[46] J. Knobbe, Th. Schäpers, Phys. Rev. B 71 (2005) 035311.
[47] L. Serra, D. Sánchez, R. López, Phys. Rev. B 72 (2005) 235309.
[48] H. Su, B.-Y. Gub, Phys. Lett. A 341 (2005) 198.
[49] R. Landauer, IBM J. Res. Dev. 1 (1957) 223;
R. Landauer, Phys. Lett. 85A (1981) 91.
[50] M. Büttiker, Y. Imry, R. Landauer, S. Pinhas, Phys. Rev. B 31 (1985) 6207.
[51] Y.V. Pershin, J.A. Nesteroff, V. Privman, Phys. Rev. B 69 (2004) 121306(R).
[52] J.A. Nesteroff, Y.V. Pershin, V. Privman, IEEE Trans. Nanotech. 4 (2005) 141.

You might also like