paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1501.03486 | 2 | 1501 | 2015-01-16T10:15:42 | A spin-wave logic gate based on a width-modulated dynamic magnonic crystal | [
"cond-mat.mes-hall"
] | An electric current controlled spin-wave logic gate based on a width-modulated dynamic magnonic crystal is realized. The device utilizes a spin-wave waveguide fabricated from a single-crystal Yttrium Iron Garnet film and two conducting wires attached to the film surface. Application of electric currents to the wires provides a means for dynamic control of the effective geometry of the waveguide and results in a suppression of the magnonic band gap. The performance of the magnonic crystal as an AND logic gate is demonstrated. | cond-mat.mes-hall | cond-mat | A spin-wave logic gate based on a width-modulated dynamic magnonic crystal
A.A. Nikitin1,2,3*, A.B. Ustinov1,3, A.A. Semenov1, A.V. Chumak2, A.A. Serga2, V.I. Vasyuchka2,
E. Lähderanta4, B.A. Kalinikos1, and B. Hillebrands2
1) Department of Physical Electronics and Technology, St. Petersburg Electrotechnical University, St. Petersburg,
197376 Russia
2) Fachbereich Physik and Landesforschungszentrum OPTIMAS, Technische Universität Kaiserslautern,
Kaiserslautern, 67663 Germany
3) Department of Mathematics and Physics, Lappeenranta University of Technology, Lappeenranta, 53850 Finland
An electric current controlled spin-wave logic gate based on a width-modulated dynamic magnonic crystal is
realized. The device utilizes a spin-wave waveguide fabricated from a single-crystal Yttrium Iron Garnet film and
two conducting wires attached to the film surface. Application of electric currents to the wires provides a means for
dynamic control of the effective geometry of the waveguide and results in a suppression of the magnonic band gap.
The performance of the magnonic crystal as an AND logic gate is demonstrated.
* and.a.nikitin@gmail.com
In recent years artificially patterned magnetic media,
magnonic crystals (MCs), have commanded increased interest
(see reviews [1-4] and literature therein). One of the important
features of MCs is the presence of band gaps in the spin-wave
spectrum, i.e. frequency bands in which the propagation of
spin waves is forbidden [5-7]. The dispersion management
introduced by an artificial periodicity of the magnetic film
allows for the observation of a variety of different linear and
nonlinear spin-wave phenomena. Among
the
formation of band gaps [8-11] and the generation of gap
solitons [12, 13]. Magnonic crystals are also promising for
practical applications, particularly for phase shifters [14] and
generators [15], and they have also been suggested for use as
temperature or magnetic field sensors [16, 17].
them are
A very fascinating type of MCs is the dynamic magnonic
crystal (DMC) [18-20]. This is a wave guiding structure with
rapidly switchable periodic properties. Up to now, it has been
realized in the form of a YIG film with a meander type
conductor placed on its surface. An electric current in the
conductor produces a periodic spatial modulation of the bias
magnetic field [18]. Due to this property, the dynamic
for unique signal processing
magnonic crystal allows
functions, such as all-linear
frequency
conversion [19], and signal storage [20].
reversal,
time
logic functionality can also be realized via control of the spin-
wave amplitude in the interferometer arms [23, 26]. The logic
gate presented here does not utilize any interferometer
circuitry; instead, it consists of one WMDMC, which controls
the spin-wave amplitude. This function is realized through
control of the sensitivity of the propagating spin wave to the
edge modulation of the magnetic-film waveguide structure.
A schematic view of the WMDMC is shown in Fig. 1. The
device was fabricated using an epitaxial Yttrium Iron Garnet
(YIG) film. For the experimental investigation, the YIG-film
waveguide was cut from a high-quality single-crystal YIG film
of 8.5 µm thickness grown on 500 µm thick Gadolinium
Gallium Garnet (GGG) substrate by liquid-phase epitaxy. The
film has pinned surface spins that lead to standing spin-wave
resonances and appearance of additional dips in the amplitude-
frequency characteristic [27]. Periodic sinusoidal width
modulation of the YIG-film waveguide was made by chemical
etching of the film. The spatial modulation period Tm and the
peak-to-peak amplitude 2Am (see Fig. 1) were 400 m and
200 m, respectively (see Fig. 1). The length of the modulated
In this Letter, we report on the realization of a width-
modulated dynamic magnonic crystal (WMDMC). In the
designed structure, in contrast to the aforementioned DMC,
electric currents control the effective geometry of the magnetic
film waveguide for spin waves propagating therein. Moreover,
we demonstrate the application of the WMDMC as a spin-
wave logic gate.
It should be noted that so far spin-wave logic gates have
only been realized using a spin-wave interferometer geometry
[21-26]. For example, elements providing controllable phase
shifts for spin waves propagating in the arms of a Mach-
Zehnder interferometer were included in each arm. The logic
operations were based on changing the phase of the spin-wave
by π with an external current signal coded as a logical ‘1’. The
Fig. 1. Sketch of the width-modulated dynamic magnonic crystal.
The ports A and B represent the logical inputs. The currents I1 and
I2 supplied to the ports correspond to a logical ‘1’ if their values
equal IMBG. Zero currents correspond to a logical ‘0’. The output
spin-wave signal represents the logical output.
(a)
(b)
(c)
(d)
1570 Oe
Hext
1600 Oe
Fig. 2. Distributions of the bias magnetic field (top) and transmitted
spin-wave power versus frequency for
the width-modulated
dynamic magnonic crystal (bottom) for the cases when the electric
current is not applied to the wires (a); the current IMBG is applied to
only one of the wires (b,c); and the current IMBG is applied to both
wires (d). The value of the current IMBG = 1500 mA.
the conditions
for excitation of
area was 4 mm, while the width of the unmodulated section of
the YIG-film waveguide was 1.5 mm. A spatially uniform bias
magnetic field was applied across the YIG waveguide in order
to provide
surface
magnetostatic spin waves [28]. These waves were exited and
detected by the microwave strip-line antennas placed at equal
distances from both ends of the modulated YIG-film area
10 mm away from each other. Two gold wires of 50 m in
diameter were placed on the YIG-film surface along the
modulated edges, as shown in Fig. 1. These wires were used
for the dynamic control of the magnonic crystal band gap by
supplying the electric currents.
Let us consider qualitatively the formation of a band gap in
the width-modulated magnonic crystal. Generally, a wave
vector of traveling spin waves exited by the input antenna in
the YIG film has two components, which are in-plane. The
first component is longitudinal along the direction from the
input to the output antenna while the second component is
transverse. As it is physically clear, the value of the transverse
component is dependent on the width of the waveguide. This
effect leads to the periodic change in the wave guiding
properties of the width-modulated YIG film and provides the
magnonic band gap (MBG) in the amplitude vs. frequency
characteristic of the investigated MC. The number of MBGs is
defined by the type of the width modulation. For example, in
the case of rectangular modulation there are several MBGs. As
soon as the width modulation is sinusoidal, as is the case in the
investigated structure (Fig. 1), there is only one MBG in the
power vs. frequency characteristic (see Fig. 2(a)) [29].
the borders of
The physical mechanism underlying the control of the
magnonic crystal band gap by the electric currents applied to
the wires can be understood as follows. Positive currents I1 and
I2 applied to both wires create an additional negative Oersted
field with respect to the applied static bias field. Therefore, the
spatial distribution of the resulting magnetic field has two
minima at
the width modulation. The
distributions of the bias magnetic field are shown in Fig. 2 as
color maps. Provided the depth of the minima is sufficiently
large, they screen the YIG-film modulated edges due to a
decrease in the magnetic field inside the YIG waveguide. In
other words, the existence of these minima cause the width
modulation of the waveguide to become “invisible” to spin
waves. This effect leads to the disappearance of the band gap
at f0 = 6.6 GHz, as is shown in Fig. 2(d).
We will now present the dynamic properties of the device
structure realized upon application of the controlling dc
electric currents. The experiments were carried out in the
following manner. A microwave signal applied to the input
antenna excited a surface magnetostatic spin wave. After
passing through the magnonic crystal the spin wave was
detected by the output antenna (Fig. 1). The power of the
received spin waves was measured as a function of their carrier
frequency. A typical dependence measured for H = 1600 Oe
with no current in the edge wires is shown in Fig. 2(a). As can
be seen in the graph, the center frequency of the band gap was
f0 = 6.6 GHz.
The dependence of the power of the output signal vs.
frequency obtained under two equal controlling currents IMBG
of 1500 mA is shown in Fig. 2(d). Note that the data clearly
demonstrate (in comparison to Fig. 2(a)) a suppression of the
band gap at 6.6 GHz. If the dc current was applied to one of
the wires then only one of the modulated edges was
deactivated. In this case, the input spin wave was still scattered
on its path through the waveguide. In this situation, the band
gap appears as shown in Figs 2(b, c). The difference between
Fig. 2(b) and Fig. 2(c) could be explained by a possible
asymmetry in the distribution of the spin-wave amplitude
along the cross-section of the YIG-film waveguide. In this
case, there is a difference in the sensitivity of the propagating
wave to the boundary conditions on the different waveguide
edges.
Dependences of the spin-wave transmission as a function
of the currents I1 and I2 measured at the frequency f0 =
6.6 GHz is shown in Fig. 3. As is visible, an increase in both
currents leads to a substantial increase in the spin-wave
transmission due to a reduction in the influence of the
modulated edges. At the same time, there is a relatively weak
influence with only one dc current applied.
6.66.76.80.1110Magnonic crystal band gap I1=I2= 0 mA Power (a. u.)Frequency (GHz) f0=6.6 GHz6.66.76.80.1110 Magnonic crystal band gap I1= IMBG I2= 0 mAFrequency (GHz)Power (a. u.) 6.66.76.80.1110 Magnonic crystal band gap I2= IMBG I1= 0 mAFrequency (GHz) Power (a. u.)6.66.76.80.1110 I1=I2=IMBG Power (a. u.)Frequency (GHz)
Fig. 3 Dependence of the spin wave transmission on the value of
the dc current at frequency f0=6.6 GHz. Squares: both currents I1
and I2 are changed; triangles: the current I1 is changed and I2=0;
circles: I1=0 and I2 is changed.
A possible application of the WMDMC for realization of
an AND logic gate will now be explored. The operating
principle of the AND logic gate is based on controlling the
bias magnetic field distribution along the sinusoidal borders of
the YIG-film waveguide through a change in the electric
currents I1 and I2. The logic ‘0’ is represented by I = 0 mA and
a logic ‘1’ – by I = IMBG = 1500 mA, which is enough to
"switch off" the width modulation on one side of the
WMDMC. The microwave pulses at the input antenna
represent clock pulses. The microwave signal at the output
antenna serves as the logic signal output. The operational
frequency corresponds to the central frequency of the band
gap, for example f0 = 6.6 GHz. Therefore, the presence of the
band gap (i.e., low power level at the output) represents a logic
‘0’ and the absence of the band gap (i.e., high power at the
output, Pout = P1) represents a logic ‘1’.
traces of
signal
demonstrating the performance of the logic gate are shown in
Fig. 4. It is observed that the logic ‘0’ appears at the output
port in three situations: for two logic ‘0s’ applied to the input
ports A and B (Fig. 4(a)) and for the combinations of the logic
‘0’ and the logic ‘1’ applied to the inputs (Fig. 4(b, c)). In the
case where a logic ‘1’ is applied simultaneously to both inputs,
the signal at the output port corresponds to a logic ‘1’.
Typical oscilloscope
the output
In conclusion, width-modulated magnonic crystals may
find a variety of applications. For example, they can be
employed for the development of dynamic magnonic crystals
and spin-wave logic gates, as demonstrated in this paper.
Another example could be microwave notch filters with a stop-
band frequency corresponding to the magnonic band gap.
Advantages of the dynamic magnonic crystal presented above
in comparison to the ones previously developed [18-20] are the
potentially
for
miniaturization. Indeed, the current control is provided here by
short conductors, which have an inductance much less than the
meander type wire reported in [18]. The use of nanostructured
width-modulated waveguides made of Permalloy films, as in
Ref. [9], will allow for a significant reduction of the size of the
dynamic magnonic crystal and the logic circuit presented here.
fast performance and
the possibility
Fig. 4 Waveforms of the output signal of the AND logic gate.
Microwave signal frequency f0=6.6 GHz. Currents I1 and I2 serve
as logic inputs; logic ‘0’ corresponds to zero current, logic ‘1’- to
IMBG=1500 mA. High level of the microwave output signal
corresponds to a logic ‘1’, and low level to a logic ‘0’.
The work was supported in part by the Russian Science
Foundation (Grant 14-12-01296), the Ministry of Education
and Science of Russian Federation, the Academy of Finland,
EU-FET grant
the Deutsche
Forschungsgemeinschaft.
InSpin 612759, and by
1. A. A. Serga, A. V. Chumak, and B. Hillebrands, J. Phys. D: Appl.
Phys. 43, 264002 (2010).
2. V. V. Kruglyak, S. O. Demokritov, and D. Grundler, J. Phys. D:
Appl. Phys. 43, 264001 (2010).
3. B. Lenk, H. Ulrichs, F. Garbs, and M. Münzenberg, Phys. Rep.
507, 107 (2011).
4. M. Krawczyk and D. Grundler, J. Phys.: Condens. Matter 26,
123202 (2014).
5. J. W. Kos, M. Krawczyk, Yu. S. Dadoenkova, N. N. Dadoenkova,
and I. L. Lyubchanskii, J. Appl. Phys. 115, 174311 (2014).
6. H. T. Nguyen and M. G. Cottam, J. Appl. Phys. 111, 07D122
(2012).
7. M. Mruczkiewicz, M. Krawczyk, V. K. Sakharov, Yu. V.
Khivintsev, Yu. A. Filimonov, and S. A. Nikitov, J. Appl. Phys.
113, 093908 (2013).
8. K.-S. Lee, D.-S. Han, and S.-K. Kim, Phys. Rev. Lett. 102,
127202 (2009).
9. A. V. Chumak et al., Appl. Phys. Lett. 95, 262508 (2009).
10. M. Arikan, Y. Au, G. Vasile, S. Ingvarsson, and V. V. Kruglyak,
J. Phys. D: Appl. Phys. 46, 135003 (2013).
11. A. B. Ustinov and B. A. Kalinikos, Tech. Phys. Lett. 40, 568
(2014).
12. A. B. Ustinov, B. A. Kalinikos, V. E. Demidov, and S. O.
Demokritov, Phys. Rev. B 81, 180406(R) (2010).
13. S. V. Grishin, Y. P. Sharaevskii, S. A. Nikitov, E. N. Beginin, and
S. E. Sheshukova, IEEE Trans. Mag. 47, 3716 (2011).
14. Y. Zhu, K. H. Chi, and C. S. Tsai, Appl. Phys. Lett. 105, 022411
(2014).
0300600900120015000.000.250.500.751.00 I1=I2= I I1= I, I2= 0 I1= 0, I2= INormalized transmission (a. u.)Current I (mA)15. S. V. Grishin, E. N. Beginin, M. A. Morozova, Yu. P.
Sharaevskii, and S. A. Nikitov, J. Appl. Phys. 115, 053908
(2014).
16. M. Inoue, A. Baryshev, H. Takagi, P. B. Lim, K. Hatafuku,
J. Noda, and K. Togo, Appl. Phys. Lett. 98, 132511 (2011).
17. R. G. Kryshtal and A. V. Medved, Appl. Phys. Lett. 100, 192410
(2012).
18. A. V. Chumak, T. Neumann, A. A. Serga, B. Hillebrands, and M.
P. Kostylev J. Phys. D: Appl. Phys. 42, 205005 (2009).
19. A. V. Chumak, V. S. Tiberkevich, A. D. Karenowska, A. A.
Serga, J. F. Gregg, A. N. Slavin, B. Hillebrands, Nat. Commun. 1,
141 (2010).
20. A. D. Karenowska, J. F. Gregg, V. S. Tiberkevich, A. N. Slavin,
A. V. Chumak, A. A. Serga, and B. Hillebrands, Phys. Rev. Lett.
108, 015505 (2012).
21. A. B. Ustinov and B. A. Kalinikos, Tech. Phys. Lett. 27, 403
(2001).
22. M. P. Kostylev, A. A. Serga, T. Schneider, B. Leven, and B.
Hillebrands, Appl. Phys. Lett. 87, 153501 (2005).
23. T. Schneider, A. A. Serga, B. Leven, B. Hillebrands, R. L.
Stamps, and M. P. Kostylev, Appl. Phys. Lett. 92, 022505 (2008).
24. A. Khitun, M. Bao, and K. L. Wang, J. Phys. D: Appl. Phys. 43,
264005 (2010).
25. Y. Au, M. Dvornik, O. Dmytriiev, and V. V. Kruglyak, Appl.
Phys. Lett. 100, 172408 (2012).
26. A. V. Chumak, A. A. Serga, and B. Hillebrands, Nat. Commun. 5,
4700 (2014).
27. B. A. Kalinikos, A. N. Slavin, J. Phys. C: Solid State Phys. 19,
7013 (1986).
28. D. D. Stancil and A. Prabhakar, Spin Waves: Theory and
Applications, Springer, 2009.
29. F. Ciubotaru, A. V. Chumak, N. Yu. Grigoryeva, A. A. Serga, and
B. Hillebrands, J. Phys. D: Appl. Phys. 45, 255002 (2012).
|
0908.4485 | 2 | 0908 | 2010-03-15T10:35:23 | Discontinuous Euler instability in nanoelectromechanical systems | [
"cond-mat.mes-hall"
] | We investigate nanoelectromechanical systems near mechanical instabilities. We show that quite generally, the interaction between the electronic and the vibronic degrees of freedom can be accounted for essentially exactly when the instability is continuous. We apply our general framework to the Euler buckling instability and find that the interaction between electronic and vibronic degrees of freedom qualitatively affects the mechanical instability, turning it into a discontinuous one in close analogy with tricritical points in the Landau theory of phase transitions. | cond-mat.mes-hall | cond-mat |
Discontinuous Euler instability in nanoelectromechanical systems
Guillaume Weick,1, 2 Fabio Pistolesi,3, 4 Eros Mariani,1, 5 and Felix von Oppen1
1Dahlem Center for Complex Quantum Systems and Fachbereich Physik,
Freie Universitat Berlin, D-14195 Berlin, Germany
2IPCMS (UMR 7504), CNRS and Universit´e de Strasbourg, F-67034 Strasbourg, France
3CPMOH (UMR 5798), CNRS and Universit´e de Bordeaux I, F-33405 Talence, France
4LPMMC (UMR 5493), CNRS and Universit´e Joseph Fourier, F-38042 Grenoble, France
5School of Physics, University of Exeter, Stocker Road, Exeter, EX4 4QL, UK
(Dated: November 1, 2018)
We investigate nanoelectromechanical systems near mechanical instabilities. We show that quite
generally, the interaction between the electronic and the vibronic degrees of freedom can be ac-
counted for essentially exactly when the instability is continuous. We apply our general framework
to the Euler buckling instability and find that the interaction between electronic and vibronic de-
grees of freedom qualitatively affects the mechanical instability, turning it into a discontinuous one
in close analogy with tricritical points in the Landau theory of phase transitions.
PACS numbers: 73.63.-b, 85.85.+j, 63.22.Gh
I.
INTRODUTION
The buckling of an elastic rod by a longitudinal com-
pression force F applied to its two ends constitutes the
paradigm of a mechanical
instability, called buckling
instability.1 It was first studied by Euler in 1744 while in-
vestigating the maximal load that a column can sustain.2
As long as F stays below a critical force Fc, the rod re-
mains straight, while for F > Fc it buckles, as sketched in
Fig. 1a-b. The transition between the two states is con-
tinuous and the frequency of the fundamental bending
mode vanishes at the instability.
There has been much recent interest in exploring buck-
ling instabilities in nanomechanical systems.
In the
quest to understand the remarkable mechanical prop-
erties of nanotubes,3 -- 5 there have been observations of
compressive buckling instabilities in this system.6 The
Euler buckling instability has been observed in SiO2
nanobeams and shown to obey continuum elasticity
theory.7 There are also close relations with the recently
observed wrinkling8 and possibly with the rippling9 of
suspended graphene samples. Theoretical works have
studied the quantum properties of nanobeams near the
Euler instability,10 -- 13 proposing this system to explore
zero-point fluctuations of a mechanical mode11 or to serve
as a mechanical qubit.13
In this work, we study the interaction of current flow
with the vibrational motion near such continuous me-
chanical
instabilities which constitutes a fundamental
issue of nanoelectromechanics.14 Remarkably, we find
that under quite general conditions, this problem ad-
mits an essentially exact solution due to the continu-
ity of the instability and the consequent vanishing of
the vibronic frequency at the transition ("critical slow-
ing down").
In fact, the vanishing of the frequency
implies that the mechanical motion becomes slow com-
pared to the electronic dynamics and an appropriate non-
equilibrium Born-Oppenheimer (NEBO) approximation
becomes asymptotically exact near the transition. Here,
FIG. 1: (color online) Sketch of a nanobeam (a) in the flat
state and (b) the buckled state with two equivalent metastable
positions of the rod (solid and dashed lines). An equivalent
circuit of the embedded SET is shown in (c).
we illustrate our general framework by applying it to the
nanoelectromechanics of the Euler instability.
We find that the interplay of electronic transport and
the mechanical instability causes significant qualitative
changes both in the nature of the buckling and in the
transport properties.
In leading order, the NEBO ap-
proximation yields a current-induced conservative force
acting on the vibronic mode. At this order, our principal
conclusion is that the coupling to the electronic dynamics
can change the nature of the buckling instability from a
continuous to a discontinuous transition which is closely
analogous to tricritical behavior in the Landau theory
of phase transitions. Including in addition the fluctua-
tions of the current-induced force as well as the corre-
sponding dissipation leads to Langevin dynamics of the
vibrational mode which becomes important in the vicin-
ity of the discontinuous transition. Employing the same
NEBO limit to deduce the electronic current, we find that
the buckling instability induces a current blockade over
a wide range of parameters. This is a manifestation of
the Franck-Condon blockade15 -- 17 whenever the buckling
instability remains continuous but is caused by a novel
tricritical blockade when the instability is discontinuous.
The emergence of a current blockade in the buckled state
suggests that our setup could, in principle, serve as a
mechanically-controlled switching device.
II. MODEL
Close to the Euler instability, the frequency of the fun-
damental bending mode of the beam approaches zero,
while all higher modes have a finite frequency.1 This al-
lows us to retain only the fundamental mode of amplitude
X (see Fig. 1a-b) and following previous studies,10 -- 12 we
reduce the vibrational Hamiltonian to18
Hvib =
P 2
2m
+
mω2
2
X 2 +
α
4
X 4,
(1)
which is closely analogous to the Landau theory of contin-
uous phase transitions. In Eq. (1), P is the momentum
conjugate to X and m denotes an effective mass. The
mode frequency ω2 ∼ 1 − F/Fc changes sign when F
reaches a critical force Fc. Global stability then requires
a quartic term with α > 0. Thus, for F < Fc (ω2 > 0),
X = 0 is the only stable minimum and the beam remains
straight. For F > Fc (ω2 < 0), the beam buckles into
one of the two minima at ±X+ = ±(cid:112)−mω2/α.
The vibronic mode of the nanobeam interacts with an
embedded metallic single-electron transistor (SET), con-
sisting of a small metallic island coupled to source, drain,
and gate electrode (Fig. 1c). We assume that the SET
operates in the Coulomb blockade regime19 and that bias
and gate voltage are tuned to the vicinity of the conduct-
ing region between SET states with, say, zero and one
excess electron. The electronic degrees of freedom cou-
ple to the vibronic motion through the occupation n of
excess electrons on the metallic island. Specifically, we
assume that the electron-vibron coupling does not break
the underlying parity symmetry of the vibronic dynam-
ics under X → −X. This follows naturally when the
coupling emerges from the electron-phonon coupling in-
trinsic to the nanobeam20 and implies that the coupling
depends only on even powers of the vibronic mode coor-
dinate X. The dominant coupling is quadratic in X,
with a coupling constant g > 0.20 When there is a signif-
icant contribution to the electron-vibron coupling origi-
nating from the electrostatic dot-gate interaction, we en-
vision a symmetric gate setup consistent with Eq. (2).
In the presence of the vibronic dynamics X(t), the elec-
tronic occupation n(X, t) of the island is described by the
Boltzmann-Langevin equation21
= {n, Hvib} + Γ+(1 − n) − Γ−n + δJ+ − δJ−.
(3)
dn
dt
This equation assumes that the bias is large compared
to temperature so that tunneling is effectively unidi-
rectional and the relevant tunneling rates Γ± for tun-
neling onto (+) and off (−) the island are given by
Γ± = R−1(V /2 ± ¯Vg)Θ(V /2 ± ¯Vg). Here, R denotes
the tunneling resistances (R (cid:29) h/e2) between island and
leads, V is the bias voltage, Θ(x) denotes the Heaviside
2
step function, and we set = e = 1. Since both gate
voltage (via the capacitances C and Cg in Fig. 1c) and vi-
bronic deformations couple to the excess charge n on the
island, the effective gate voltage ¯Vg = Vg − gX 2/2 com-
bines the gate potential Vg (measured from the degen-
eracy point between the states with zero and one excess
electron) and the vibron-induced shift of the electronic
energy described by Hc. The stochastic Poisson nature
of electronic tunneling is accounted for by including the
Langevin sources δJ± with correlators (cid:104)δJ+(t)δJ+(t(cid:48))(cid:105) =
Γ+(1 − n)δ(t − t(cid:48)) and (cid:104)δJ−(t)δJ−(t(cid:48))(cid:105) = Γ−nδ(t − t(cid:48)).
The vibronic dynamics enters Eq. (3) through the Pois-
son bracket {n, Hvib}.
III. STABILITY ANALYSIS
We are now in a position to investigate the influence
of the electronic dynamics on the vibronic motion. Near
the instability, the vibrational dynamics becomes slow
compared to the electronic tunneling dynamics. As has
recently been shown,22,23 the effect of the current on
the vibrational motion can then be described within a
NEBO approximation in which the vibrational motion is
subject to a current-induced force −gXn(X, t) originat-
ing in the electron-vibron interaction (2). This current-
induced force involves both a time-averaged and conser-
vative force as well as fluctuating and frictional forces,
resulting in Langevin dynamics of the vibronic degree of
freedom.
In lowest order, the Langevin dynamics only involves
the conservative force which emerges from the average
occupation n0(X) in the absence of fluctuations (δJ± (cid:39)
0) and vibronic dynamics ({n, Hvib} (cid:39) 0). In this limit,
Eq. (3) reduces to the usual rate equation of a metallic
SET so that19
1,
1
2
0,
vg(x) > v/2,
+
vg(x)
v
, −v/2 (cid:54) vg(x) (cid:54) v/2,
vg(x) < −v/2
(4)
with v > 0 and vg(x) = vg − x2/2. Here and below,
we employ dimensionless variables by introducing char-
acteristic scales E0 = g2/α of energy, l0 = (cid:112)g/α of
length, and ω0 = (cid:112)g/m of frequency (or time t) from
a comparison of the quartic vibron potential in Hvib and
the electron-vibron coupling Hc. Specifically, we intro-
duce the reduced variables x = X/l0, p = P/mω0l0,
τ = ω0t, v = V /E0, vg = Vg/E0, and r = Rω0/E0. In
terms of these variables, we can also write Hvib + Hc =
E0[p2/2 + (− + n)x2/2 + x4/4] in terms of a reduced
compressional force = −mω2/g.
The current-induced force −xn0(x) has dramatic ef-
fects on the Euler instability, as follows from a stability
analysis of the vibrational motion. The (meta)stable po-
sitions of the nanobeam are obtained by setting the effec-
tive force feff (x) = x − x3 − xn0(x) to zero. Our results
Hc =
g
2
X 2 n,
(2)
n0(x) =
3
FIG. 2: (color online) (Meta)stable (solid blue lines) and unstable (dashed red lines) positions of the nanobeam vs. scaled force
for (a) vg < v/2, v < 1/2, (b) vg < v/2, v > 1/2, (c) vg > v/2, v < 1/2 (for + > 1; a similar plot holds for + < 1), (d)
vg > v/2, v > 1/2, as indicated in the vg -- v plane in (e). The dotted blue line is the result without electron-vibron coupling.
+ = − 1, and x2− = ( − )/(1 − 1/2v). Grey indicates conducting regions.
Notation: ± = 2vg ± v, = 1/2 + vg/v, x2
+ = , x2
are summarized in the stability diagrams in Fig. 2. The
most striking results of this analysis are: (i) The current
flow renormalizes the critical force required for buckling
towards larger values.
(ii) At low biases, the buckled
state can appear via a discontinuous transition.
These results can be understood most directly in terms
of the potential veff (x) associated with feff (x). Focusing
on the current-carrying region (shown in grey in Fig. 2
and delineated by max{0, −} < x2 < + with ± =
2vg ± v), we find
veff (x) =
1
2
− +
v + 2vg
2v
x2 +
1
4
1 − 1
2v
x4.
(5)
(cid:18)
(cid:19)
(cid:18)
(cid:19)
The quadratic term shows that the current indeed stabi-
lizes the unbuckled state, renormalizing the critical force
to = 1/2 + vg/v when − < 0 < + (Fig. 2a-b). Re-
markably, however, the current-induced contribution to
the quartic term is negative at small x2 and thus desta-
bilizes the unbuckled state. According to Eq. (5), the
quartic term in the current-induced potential becomes in-
creasingly significant as the bias voltage v decreases and
we find that the overall prefactor of the quartic term be-
comes negative when v < 1/2.24 It is important to note
that this does not imply a globally unstable potential
since the current-induced force contributes only for small
x2. A sign reversal of the quartic term is also a famil-
iar occurrence in the Landau theory of tricritical points
which connect between second- and first-order transition
lines.25 In close analogy, the sign reversal of the quartic
term in the effective potential (5) signals a discontinuous
Euler instability which reverts to a continuous transition
at biases v > 1/2 where the prefactor of the quartic term
remains positive.
Specifically, when v > 1/2 (Fig. 2b,d), the current-
induced potential renormalizes the parameters of the vi-
bronic Hamiltonian but leaves the quartic term posi-
tive. This modifies how the position of the minimum
depends on the applied force in the conducting region
max{0, −} < x2 < +, but the Euler instability remains
continuous. When v < 1/2, the equilibrium position
at finite x becomes unstable within the entire current-
carrying region. This leads to a discontinuous Euler tran-
sition when − < 0 < + (Fig. 2a) and to multistability
in the region − < x2 < + when − > 0 (Fig. 2c).26
At the level of the stability analysis, we can also ob-
tain the current I by evaluating the rate-equation result19
RI(x)/V = 1/4 − [vg(x)/v]2 at the position of the most
stable minimum. Corresponding results in the vg -- v
plane are shown in Fig. 3a-f for various values of the
applied force . By comparison with the Coulomb di-
amond in the absence of the electron-vibron coupling
(dotted lines in Fig. 3), we see that the Euler instability
leads to a current blockade over a significant parameter
range. For v > 1/2, the blockade is a manifestation of
the Franck-Condon blockade,15,17 caused by the induced
linear electron-vibron coupling when expanding Eq. (2)
about the buckled state.
In contrast, for v < 1/2, the current blockade is a
direct consequence of the discontinuous Euler instabil-
FIG. 3: (color online) Conductance G = RI/V in the vg -- v
plane for applied force (a) (cid:54) 0, (b) = 0.25, (c,g) = 0.5,
(d) = 0.75, (e) = 1, (f,h) = 1.25, within (a-f) stability
analysis and (g-h) full Langevin dynamics (r = γe = T =
0.01). Color scale: G = 0 → 1/4 from dark blue to white.
Dotted lines delineate the Coulomb diamond for g = 0.
4
ity. We have seen above that in this regime, the buckled
state becomes unstable throughout the entire current-
carrying region. As a result, the current-induced force
will always drive the system out of the current-carrying
region, explaining the current blockade. An intriguing
feature of this novel tricritical current blockade is the
curved boundary of the apparent Coulomb-blockade dia-
mond (Fig. 3), a behavior which is actually observed in
nanoelectromechanical systems.
Planck equation (6). Numerical results for the scaled lin-
ear conductance G = RI/V are shown in Fig. 3g-h, using
the same parameters as in Fig. 3c,f. We observe that the
fluctuations reduce the size of the blockaded region and
blur the edges of the conducting regions as the system
can explore more conducting states in phase space. Nev-
ertheless, the conclusions of the stability analysis clearly
remain valid qualitatively.
IV. LANGEVIN DYNAMICS
V. CONCLUSION
To investigate the robustness of the stability analysis
against fluctuations, we turn to the complete vibronic
Langevin dynamics x + γ(x) x = feff(x) + ξ(τ ). The
fluctuating force ξ(τ ) is generated by fluctuations of the
electronic occupation and the frictional force −γ(x) x by
the delayed response of the electrons to the vibronic
dynamics. To compute γ(x) and ξ(τ ), we solve Eq.
(3) including the vibronic dynamics and the Langevin
sources. Writing separate equations for average and
fluctuations of the occupation by setting n = ¯n + δn,
we see that the leading correction to ¯n arises from the
Poisson bracket, yielding ¯n = n0 −
X∂X n0. At
the same time, the fluctuations δn obey the correla-
tor (cid:104)δn(t)δn(t(cid:48))(cid:105) =
Insert-
ing these results into the expression for the current-
induced force −gXn and employing reduced units, we
find (cid:104)ξ(τ )ξ(τ(cid:48))(cid:105) = D(x)δ(τ−τ(cid:48)) with diffusion and damp-
ing coefficients D(x) = 2rx2n0(1 − n0)/v and γ(x) =
−rx∂xn0/v, respectively. Finally, we can pass from the
Langevin to the equivalent Fokker-Planck equation22,23
for the probability P(x, p, τ ) that the nanobeam is at po-
sition x and momentum p at time τ ,
Γ++Γ− n0(1 − n0)δ(t − t(cid:48)).
2
1
Γ++Γ−
We have presented a general approach to the interplay
between continuous mechanical instabilities and current
flow in nanoelectromechanical systems, and have applied
our general framework to the Euler buckling instabil-
ity. The current flow modifies the nature of the buck-
ling instability from a continuous to a tricritical transi-
tion. Likewise, the instability induces a novel tricritical
current blockade at low bias. Our nonequilibrium Born-
Oppenheimer approach generalizes not only to other con-
tinuous mechanical instabilities, but also to other systems
such as semiconductor quantum dots or single-molecule
junctions with a discrete electronic spectrum, to other
types of electron-vibron coupling,28 and to further trans-
port characteristics (e.g., current noise).
Our proposed setup can be realized experimentally by
clamping, e.g., a suspended carbon nanotube and apply-
ing a force to atomic precision either using a break junc-
tion or an atomic force microscope. Indeed, several recent
experiments show that the electron-vibron coupling is
surprisingly strong in suspended carbon nanotube quan-
tum dots.4,5,16
∂τP = −p∂xP − feff∂pP + γ∂p(pP) +
pP.
∂2
(6)
Acknowledgments
D
2
The current I =(cid:82) dxdpPst(x, p)I(x) is now obtained
Note that the diffusion and damping coefficients are non-
vanishing only in the conducting region.27
from the stationary solution ∂τPst = 0 of the Fokker-
We acknowledge financial support through Sfb 658 of
the DFG (GW, EM, FvO) and ANR contract JCJC-
036 NEMESIS (FP). FvO enjoyed the hospitality of the
KITP (NSF PHY05-51164).
1 L. D. Landau and E. M. Lifshitz, Theory of Elasticity
(Pergamon Press, Oxford, 1970).
2 L. Euler, in Leonhard Euler's Elastic Curves, translated
and annotated by W. A. Oldfather, C. A. Ellis, and D. M.
Brown, reprinted from ISIS, No. 58 XX(1), 1744 (Saint
Catherine Press, Bruges).
3 P. Poncharal et al., Science 283, 1513 (1999).
4 G. A. Steele et al., Science 325, 1103 (2009).
5 B. Lassagne et al., Science 325, 1107 (2009).
6 M. R. Falvo et al., Nature 389, 582 (1997).
7 S. M. Carr and M. N. Wybourne, Appl. Phys. Lett. 82,
709 (2003).
9 J. C. Meyer et al., Nature 446, 60 (2007).
10 S. M. Carr, W. E. Lawrence, and M. N. Wybourne, Phys.
Rev. B 64, 220101(R) (2001).
11 P. Werner and W. Zwerger, Europhys. Lett. 65, 158 (2004).
12 V. Peano and M. Thorwart, New J. Phys. 8, 21 (2006).
13 S. Savel'ev, X. Hu, and F. Nori, New J. Phys. 8, 105 (2006).
14 H. G. Craighead, Science 290, 1532 (2000); M. L. Roukes,
Phys. World 14, 25 (2001).
15 J. Koch and F. von Oppen, Phys. Rev. Lett. 94, 206804
(2005).
16 R. Leturcq et al., Nature Phys. 5, 327 (2009).
17 F. Pistolesi and S. Labarthe, Phys. Rev. B 76, 165317
8 W. Bao et al., Nature Nanotech. 4, 562 (2009).
(2007).
18 The rotation of the plane of the buckled nanobeam is as-
sumed massive due to clamped boundary conditions.
19 See, e.g., Chap. 3 in T. Dittrich et al., Quantum Transport
and Dissipation (Wiley-VCH, Weinheim, 1998).
20 E. Mariani and F. von Oppen, Phys. Rev. B 80, 155411
(2009).
21 See Ya. M. Blanter and M. Buttiker, Phys. Rep. 336, 1
(2000) for a review of the Boltzmann-Langevin method.
22 Ya. M. Blanter, O. Usmani, and Yu. V. Nazarov, Phys.
Rev. Lett. 93, 136802 (2004); ibid. 94, 049904(E) (2005).
23 D. Mozyrsky, M. B. Hastings, and I. Martin, Phys. Rev. B
73, 035104 (2006).
24 We note that the singularity at small v is cut off for bias
voltages of the order of temperature or level broadening.
5
25 See, e.g., P. M. Chaikin and T. C. Lubensky, Principles of
Condensed Matter Physics (Cambridge University Press,
Cambridge, 1995).
26 In a quite different context, a discontinuous Euler insta-
bility has also been predicted in: S. Savel'ev and F. Nori,
Phys. Rev. B 70, 214415 (2004).
27 In some cases, a stable numerical solution of Eq. (6) re-
quires a small extrinsic damping γe and temperature T .
28 E.g., a small symmetry-breaking coupling linear in X leads
to a tricritical point in an external field, a purely linear
coupling to a 2nd order transition in a field, G. Weick et
al., unpublished.
|
1007.3144 | 1 | 1007 | 2010-07-19T13:24:01 | Polarization dependence of coherent phonon generation and detection in highly-aligned single-walled carbon nanotubes | [
"cond-mat.mes-hall"
] | We have investigated the polarization dependence of the generation and detection of radial breathing mode (RBM) coherent phonons (CP) in highly-aligned single-walled carbon nanotubes. Using polarization-dependent pump-probe differential-transmission spectroscopy, we measured RBM CPs as a function of angle for two different geometries. In Type I geometry, the pump and probe polarizations were fixed, and the sample orientation was rotated, whereas, in Type II geometry, the probe polarization and sample orientation were fixed, and the pump polarization was rotated. In both geometries, we observed a very nearly complete quenching of the RBM CPs when the pump polarization was perpendicular to the nanotubes. For both Type I and II geometries, we have developed a microscopic theoretical model to simulate CP generation and detection as a function of polarization angle and found that the CP signal decreases as the angle goes from 0 degrees (parallel to the tube) to 90 degrees (perpendicular to the tube). We compare theory with experiment in detail for RBM CPs created by pumping at the E44 optical transition in an ensemble of single-walled carbon nanotubes with a diameter distribution centered around 3 nm, taking into account realistic band structure and imperfect nanotube alignment in the sample. | cond-mat.mes-hall | cond-mat |
Polarization dependence of coherent phonon generation and detection in
highly-aligned single-walled carbon nanotubes
L. G. Booshehri,1, 2 C. L. Pint,2, 3, 4 G. D. Sanders,5 L. Ren,1, 2 C. Sun,1, 2 E. H. H´aroz,1, 2
J.-H. Kim,6 K.-J. Yee,6 Y.-S. Lim,7 R. H. Hauge,2, 4 C. J. Stanton,5 and J. Kono1, 2, 3, ∗
1Department of Electrical and Computer Engineering, Rice University, Houston, Texas 77005, USA
2The Richard E. Smalley Institute for Nanoscale Science and Technology, Rice University, Houston, Texas 77005
3Department of Physics and Astronomy, Rice University, Houston, Texas 77005, USA
4Department of Chemistry, Rice University, Houston, Texas 77005, USA
5Department of Physics, University of Florida, Box 118440, Gainesville, Florida 32611-8440
6Department of Physics, Chungnam National University, Daejeon, 305-764, Republic of Korea
7Department of Applied Physics, Konkuk University, Chungju, Chungbuk, 380-701, Republic of Korea
(Dated: January 5, 2018)
We have investigated the polarization dependence of the generation and detection of radial breath-
ing mode (RBM) coherent phonons (CP) in highly-aligned single-walled carbon nanotubes. Using
polarization-dependent pump-probe differential-transmission spectroscopy, we measured RBM CPs
as a function of angle for two different geometries. In Type I geometry, the pump and probe po-
larizations were fixed, and the sample orientation was rotated, whereas, in Type II geometry, the
probe polarization and sample orientation were fixed, and the pump polarization was rotated. In
both geometries, we observed a very nearly complete quenching of the RBM CPs when the pump
polarization was perpendicular to the nanotubes. For both Type I and II geometries, we have de-
veloped a microscopic theoretical model to simulate CP generation and detection as a function of
polarization angle and found that the CP signal decreases as the angle goes from 0◦ (parallel to
the tube) to 90◦ (perpendicular to the tube). We compare theory with experiment in detail for
RBM CPs created by pumping at the E44 optical transition in an ensemble of single-walled carbon
nanotubes with a diameter distribution centered around 3 nm, taking into account realistic band
structure and imperfect nanotube alignment in the sample.
PACS numbers: 78.67.Ch, 63.22.+m, 73.22.-f, 78.67.-n
I.
INTRODUCTION
The one-dimensionality of single-walled carbon nan-
otubes (SWNTs) is attractive from both fundamental
and applied points of view, where the 1D confinement
of electrons and phonons results in unique anisotropic
electric, magnetic, mechanical, and optical properties.1,2
Individualized SWNTs, both single-tube and in ensem-
ble samples, have shown anisotropy with polarized Ra-
man scattering and optical absorption measurements
where maximum signals result when the nanotube axis is
aligned parallel to the polarization of incident light.3–8
Additionally, due to strong anisotropic magnetic sus-
ceptibilities, both semiconducting and metallic SWNTs
align well within an external magnetic field, and with the
added properties of the Aharonov-Bohm effect, the elec-
tronic band structure of SWNTs respond anisotropically
with the strength of a tube-threading magnetic flux.9–16
Such optical and magnetic anisotropy is also expected
with bulk samples, but detailed measurements showing
extreme anisotropy have been lacking.17
Here we investigate anisotropic optical and vibrational
properties of a bulk film of highly-aligned SWNTs us-
ing polarization-dependent coherent phonon (CP) spec-
troscopy. CP spectroscopy is an ultrafast pump-probe
technique that complements CW Raman spectroscopy,
and although both techniques provide information about
electron-phonon coupling, CP spectroscopy avoids the
common disadvantages of Raman spectroscopy that in-
clude detection of Rayleigh scattering and photolumi-
nescence and the broadening and blending of peak
features.18 Such advantages are useful when investigat-
ing SWNTs, where a large majority of samples include
ensembles dispersed in various environments and their
optical properties are obscured by the collection of vary-
ing species of nanotubes.
Recent CP studies on SWNTs have produced direct
observation of CP oscillations of both the radial breath-
ing mode (RBM) and G-band phonons, in addition to
their phase information and dephasing times.18–24 Fur-
thermore, when pulse-shaping techniques are combined
with CP spectroscopy, predesigned trains of femtosec-
ond pulses selectively excite RBM CPs of a specific
chirality, avoiding inhomogeneous broadening from the
ensemble.23 However, it is important to note that pre-
vious CP studies investigated randomly-aligned SWNT
samples, and as the quasi-1D nature of SWNTs leads to
optical anisotropy that is dominant in the polarization
dependence of PL, absorption, and Raman scattering,
CP studies on aligned SWNTs are necessary.
Below, we present results of a detailed experimental
and theoretical study of CPs in highly-aligned SWNT
thin films and found a very strong polarization anisotropy
of the RBM as a function of angle. In particular, we ob-
served a very nearly complete quenching of the RBM
when the optical polarization is perpendicular to the
e
c
n
a
b
r
o
s
b
A
e
c
n
a
b
r
o
s
b
A
3.0
2.5
2.0
1.5
1.0
0.5
1.0
0.8
0.6
0.4
0.2
0.0
(a)
A//
A⊥
1.6
2.4
2.0
Energy (eV)
2.8
3.2
(b)
1
2
A//
A⊥
4
3
Energy (meV)
5
6
7
2
!
Sapphire
Type I
V
V
(a)
SWNTs
H
pump probe
(b)
SWNTs
H
!
FIG. 1: (color online) Absorption spectra of aligned single-
walled carbon nanotubes in the (a) near-infrared to visible
and (b) far-infrared (or terahertz) ranges for parallel (Ak)
and perpendicular (A⊥) polarization.
tubes. We have developed a theoretical model to un-
derstand this extreme anisotropy, including band struc-
ture, interband optical transition elements, ultrafast car-
rier and phonon dynamics, and imperfect nanotube align-
ment in the sample. Fitting our results to theory, we also
calculated the nematic order parameter, S,15–17 i.e., the
degree of alignment of SWNTs in the sample.
II. SAMPLES AND EXPERIMENTAL
METHODS
We investigated CPs through degenerate pump-probe
spectroscopy measurements on highly-aligned SWNT
thin films transferred onto sapphire substrates. Pat-
terned, vertically aligned SWNT arrays grown by chemi-
cal vapor deposition were subsequently etched with H2O
vapor to allow transffer to sapphire via a dry contact
transfer printing technique, which forms horizontally-
aligned SWNT thin films.25,26 The resulting film pro-
duces aligned nanotubes of the same length, with a di-
ameter distribution centered around 3 nm.26 Figure 1(a)
shows polarization-dependent absorption spectra of a
typical aligned sample in the visible to near-infrared
range, while Fig. 1(b) shows absorption spectra in the
far-infrared or terahertz (THz) range. The polarization
Sapphire
pump probe
Type II
FIG. 2: (color online) Scanning electron microscopy image of
aligned single-walled carbon nanotubes, and the two exper-
imental configurations employed in the current pump-probe
spectroscopy work.
(a) Type I: pump and probe polariza-
tions are fixed and sample orientation is rotated. (b) Type
II: probe and sample orientations are fixed and pump polar-
ization is rotated.
anisotropy is obvious in both (a) and (b), but extreme
in the THz regime (b), as there is virtually zero ab-
sorption when the sample is perpendicular to the THz
polarization.17 Therefore, with such an aligned sample,
our polarization measurements can be extended to in-
clude the sample as an additional rotation parameter,
compared to our earlier study of CP polarization depen-
dence in randomly-aligned SWNTs.21
Using a mode-locked Ti:Sapphire laser with ∼80 fs
pulse width that is shorter than the period of excited
CP oscillations and ∼40 mW average pump power, the
laser was tuned to central wavelength of 850 nm (1.46 eV)
to predominantly excite the E44 interband transitions of
the SWNTs in the sample. A shaker-delay system and
balance detector was employed for fast-scan, real-time
observation of CPs. Extraction of CP oscillations from
raw pump-probe time-domain signal were performed as
previously described.18 As depicted in Fig. 2, two types
of polarization measurements were investigated. Type I
configuration maintained the same polarization for the
60
50
40
30
20
10
6
−
0
1
×
/
T
T
∆
0
0.5
(a) Type I
q = 0°
(b) Type II
f = 0°
15°
30°
45°
60°
90°
15°
30°
45°
60°
90°
)
s
t
i
.
n
u
b
r
a
(
y
t
i
s
n
e
t
n
I
1.5
1.0
Time (ps)
2.0
0.5
1.0
1.5
Time (ps)
2.0
3
Type II
λ = 850 nm
f = 0°
30°
45°
60°
90°
FIG. 3: (color online) Experimental differential transmission
data in the time domain showing coherent phonon oscillations
of the radial breathing mode for different polarization angles
in (a) Type I and (b) Type II configurations (see Fig. 1). The
traces are vertically offset for clarity.
pump and probe, while the alignment axis of the sample
was rotated. Type II configuration maintained the same
orientation for the probe and sample, while the pump
polarization was rotated. For Type II measurements, a
half-wave plate provided 90-degree rotation of the pump.
All measurements were performed at room temperature.
III. EXPERIMENTAL RESULTS
Results of our polarization-dependent ultrafast pump-
probe differential transmission measurements on the
aligned SWNT film for both Type I and Type II ori-
entations are shown in Fig. 3. Here, it is seen that the
typical differential transmission amplitude of the RBM
CP oscillations is on the order of 10−6, with a CP de-
cay time of ∼1.5 ps. This short decay time, compared
to our earlier studies on individually-suspended SWNTs
in solution,18,21,23 is expected with our sample of bun-
dled SWNTs. As the polarization angle is rotated, we
see a strong polarization anisotropy of the RBM CPs as
a function of angle, where the strongest oscillations are
observed at 0◦ while the signal is very nearly completely
quenched at 90◦ for both Type I and Type II geometries.
Figure 4 shows CP spectra for different polarization
angles for Type II configuration. To produce these spec-
tra in the frequency domain, we calculated the Fourier
transform (FT) of the time-domain CP oscillations. The
frequencies of the CP oscillations have a wide range, from
50 to 200 cm−1, which is consistent with the large di-
ameter distribution of nanotubes within the sample.26
The polarization anisotropy is clearly observed in the CP
0
100
200
Frequency (cm-1)
300
FIG. 4: (color online) Coherent phonon spectra for different
polarization angles in Type II configuration obtained through
Fourier transform of the time-domain data in Fig. 3(b). The
traces are vertically offset for clarity.
)
s
t
i
n
u
.
b
r
a
(
y
t
i
s
n
e
n
t
I
t
d
e
a
r
g
e
t
n
I
1.0
0.8
0.6
0.4
0.2
0.0
0
(a) Type I
(b) Type II
30
60
q (°)
90
0
30
60
f (°)
90
FIG. 5: (color online) Spectrally integrated coherent phonon
intensity as a function of angle, measured in (a) Type I and
(b) Type II configurations.
spectra. Figures 5(a) and 5(b) plot the integrated CP in-
tensity of the FT for all excited nanotubes as a function
of θ [in Fig. 5(a)] and φ [in Fig. 5(b)], where θ (φ) is the
angle of rotation for the Type I (Type II) configuration.
IV. THEORY
We have developed a microscopic theory for the gener-
ation of coherent phonons in single-walled carbon nan-
otubes and their detection in coherent phonon spec-
troscopy experiments. Our microscopic theory is de-
scribed in detail in Ref. 22, so we only summarize the
main points here and indicate how our earlier work has
been extended to include polarization-dependent coher-
ent phonon spectroscopy.
We treat the π and π∗ electronic states in (n,m) car-
bon nanotubes using a third-nearest-neighbor extended
tight binding (ETB) formalism developed by Porezag et
al.27 for carbon compounds. Using a density functional
based parametrization, Porezag et al. derived analytic
expressions for the Hamiltonian and overlap matrix ele-
ments that depend only on the C-C bond lengths. Using
the ETB formalism, we obtain tight-binding wave func-
tions and electronic energy levels Esµ(k) where s = c, v
labels the conduction and valence bands, µ labels the cut-
ting lines, and k is the one-dimensional nanotube Bril-
louin zone.1 By exploiting the screw symmetry of the
nanotube, we can block-diagonalize the electronic Hamil-
tonian and overlap matrices into 2×2 blocks, one for each
cutting line, as described in Appendix A of Ref. 22.
We treat nanotube lattice dynamics using a seven-
parameter valence-force-field model described in Ap-
pendix B of Ref. 22. Exploiting the nanotube screw sym-
metry, the dynamical matrix can be block-diagonalized
into 6 × 6 blocks, one for each cutting line, which can
be solved for the phonon displacement vectors and dis-
persion relations ω2
βν(q). In the the dispersion relations,
β = 1,··· , 6 labels the phonon modes, ν is an angu-
lar momentum quantum number, and q is the phonon
wave vector in the one-dimensional nanotube Brillouin
zone. Following the work of Jiang et al.28 and Lobo and
Martins,29 we include four types of force field potentials,
i.e., the bond stretching, in-plane bond bending, out-of-
plane bond bending, and bond twisting potentials. Our
force-field potential energies are invariant under rigid ro-
tations and translations (force constant sum rule), and,
as a result, our phonon model correctly predicts the dis-
persion relation for the long-wavelength flexure modes.22
The electron-phonon interaction is treated in a second-
quantized formalism where the electron-phonon inter-
action matrix elements are evaluated in Appendix C
of Ref. 22.
In calculating electron-phonon matrix ele-
ments, we use 2pz graphene atomic wave functions and
screened atomic potentials obtained from an ab initio
calculation.28
We obtain equations of motion for coherent phonon
amplitudes using the Heisenberg equations as described
by Kuznetsov and Stanton.30 We assume that the opti-
cal pulse and photoexcited carrier distributions are dis-
tributed uniformly over the nanotube. In this case, only
the ν = q = 0 phonon modes are excited. To excite
RBM coherent phonons, the laser pulse must be short
in comparison with the RBM phonon oscillation period.
4
For coherent RBM phonons, the coherent phonon ampli-
tude is proportional to the tube diameter D(t). The tube
diameter, being proportional to the coherent phonon am-
plitude, satisfies a driven oscillator equation
∂2D(t)
∂t2 + ω2D(t) = S(t)
(1)
where ω is the angular frequency of the ν = q = 0 RBM
phonon mode. The driving function S(t) for RBM co-
herent phonons is given by
S(t) ∝ −Xsµk
Msµ(k)(cid:2)fsµ(k, t) − f 0
sµ(k)(cid:3)
(2)
where Msµ(k) is the driving function kernal for RBM co-
herent phonons defined in Eq. (12) of Ref. 22 and fsµ(k, t)
and f 0
sµ(k) are the time-dependent and initial equilibrium
carrier distribution functions, respectively.
From Eq. (2) we see that the driving function depends
on the photoexcited carrier distribution functions. We
treat photoexcitation of carriers in a Boltzmann equa-
tion formalism and obtain the photogeneration rate us-
ing Fermi's golden rule. In Ref. 22 we only considered
linearly-polarized laser pulses with the electric polariza-
tion vector parallel to the nanotube axis. For parallel po-
larization, optical transitions only occur between states
with the same cutting line index µ, and the resulting
photogeneration rate is given by Eq. (13) in Ref. 22.
In the present work, we consider linearly-polarized
pump and probe beams in which the electric polariza-
tion vector makes a finite angle with the nanotube axis.
If ǫ is the unit electric polarization vector for the pump,
then the photogeneration rate is now given by
∂fsµ(k)
∂t
=
(cid:12)(cid:12)(cid:12)(cid:12)gen
8π2e2 u(t)
n2
g (ω)2 (cid:18) 2
×(cid:16)fs′µ′ (k, t) − fsµ(k, t)(cid:17) δ(cid:16)∆Eµµ′
m0(cid:19)Xs′µ′ ǫ · ~P µµ′
ss′ (k) − ω(cid:17)
ss′ (k)2
(3)
where ∆Eµµ′
ss′ (k) = Esµ(k) − Es′µ′ (k) are the k-
dependent transition energies, ω is the pump photon
energy, u(t) is the time-dependent energy density of the
pump pulse, e is the electronic charge, m0 is the free
electron mass, and ng is the index of refraction in the
surrounding medium. To account for spectral broaden-
ing of the laser pulses, the delta function in Eq. (3) was
replaced with a Lorentzian lineshape31
δ(∆E − ω) →
Γp/(2π)
(∆E − ω)2 + (Γp/2)2
(4)
In Eq. (3), ~P µµ′
ss′ (k) are the dipole-allowed optical ma-
trix elements between the initial and final states where µ
and µ′ are initial and final state cutting lines. For polar-
ization parallel to the tube axis (taken to be z), µ′ = µ
and z · ~P µµ
ss′ (k) with the right hand side being
defined in Eq. (14) of Ref. 22. For linear polarization
ss′ (k) = P µ
parallel to the x axis, µ′ = µ ± 1 and in the notation of
Ref. 22, we have
ss′
x · ~P µ,µ±1
×XrJ
(k) =
√2m0
1
2 Xr′
C ∗
r′ (s′, µ ± 1, k)
Cr(s, µ, k) eiφJ(k,µ) (cid:0)Mx(r′, rJ) ± iMy(r′, rJ)(cid:1)
Similarly, for linear polarization along the y axis, we have
(5)
(a)
E44
)
1
-
m
c
4
0
1
(
(38,0)
E77
)
V
e
(
y
g
r
e
n
E
1
0
-1
y · ~P µ,µ±1
ss′
(k) = ∓ i(cid:0)x · ~P µ,µ±1
ss′
(k)(cid:1)
(6)
-0.4 -0.2
0.0
0.2
0.4
8
6
4
2
0
5
(b)
00
150
300
450
600
900
E44
2
1
0
3
Photon Energy (eV)
E44
(d)
RBM
10.91 meV
Type II
(38,0)
0
30
60
90
Angle (deg.)
In Eqs. (5) and (6), the atomic dipole matrix element
vectors (which can be evaluated analytically) are given
by
M(r′, rJ) = Z dr ϕ∗
r′ 0(r − Rr′ 0) ∇ ϕrJ(r − RrJ)
(7)
where the 2pz orbitals ϕrJ are defined in Eq. (C3) of
Ref. 22.
In our experiments, a probe pulse is used to mea-
sure the time-varying absorption coefficient. In our the-
ory, the time-varying optical properties measured by the
probe pulse are obtained from the imaginary part of the
dielectric function. For a linearly polarized probe pulse,
we have
)
s
t
i
n
u
.
b
r
a
(
r
e
w
o
P
k ( /T)
E44
E77
(c)
)
s
t
i
n
u
.
b
r
a
(
r
e
w
o
P
d
e
a
r
g
e
n
t
t
I
00
150
300
450
600
900
2
1
Pump Energy (eV)
3
ε2(ω, t) =
8π2e2
At(ω)2 (cid:18) 2
m0(cid:19) Xss′µµ′Z dk
π ǫ · ~P µµ′
ss′ (k)2
×(cid:16)fsµ(k, t) − fs′µ′ (k, t)(cid:17) δ(cid:16)∆Eµµ′
ss′ (k) − ω(cid:17)
(8)
where At = π(dt/2)2 is the cross-sectional area of
the tube and dt is the equilibrium nanotube diameter.
The dirac delta function is replaced by a broadened
Lorentzian of the form shown in Eq. (4). The photon
energy of the probe pulse photons is ω, and ǫ is the
probe unit polarization vector. In our experiments, the
photon energy ω is the same for both pump and probe,
while the polarization vector ǫ may be different for pump
and probe. To obtain the computed coherent phonon
spectrum, we take the power spectrum of the computed
differential transmission signal after background subtrac-
tion using the Lomb periodogram algorithm described in
Ref. 32.
V. COMPARISON OF THEORY AND
EXPERIMENT
To compare experiment with theory, we performed
simulations of polarization-dependent CP spectroscopy
on a (38,0) zigzag nanotube. This is a mod-2 semicon-
ducting tube with a diameter of 3.01 nm. Our sample
contains an ensemble of nanotubes with diameters cen-
tered around 3 nm,26 and we expect that (38,0) tubes
FIG. 6: (color online) (a) Theoretical bandstructure for (38,0)
nanotubes showing the strong E44 and E77 transitions for a
polarization vector parallel to the tube. The wavevector k is
expressed in units of π/T where T = 4.31 A is the length of
the translational unit cell. (b) Computed absorption spectra
for polarization angles varying from 0◦ (parallel to tube) to
90◦ (perpendicular to tube). (c) Coherent phonon spectra as
a function of pump energy and pump polarization angle φ for
RBM coherent phonons (ω = 10.91 meV). Coherent phonon
spectra are Fourier transforms of computed time dependent
differential transmission for Type II pump-probe experiments.
(d) Integrated power (black dots) obtained by taking the area
under the computed E44 peaks in lower left panel. We fit (red
curve) the theoretical calculations to A cosp(φ) where A and
p = 4.017 are fitting parameters.
will contribute strongly to the CP signal for the ensem-
ble since (i) the (38,0) tube has a diameter in the center of
the measured diameter distribution and (ii) mod-2 zigzag
tubes tend to have very strong intrinsic CP signals.18,22,23
Our theoretical results for the (38,0) nanotube are
shown in Fig. 6. Our computed electronic π band struc-
ture for the (38,0) tube is shown in Fig. 6(a). The bands
are doubly degenerate with two distinct cutting lines cor-
responding to each band. The π valence bands have neg-
ative energy, and the π∗ conduction bands have posi-
tive energy. For z-polarized light, the allowed E44 and
E77 transitions (selection rule ∆µ = 0) giving rise to the
strongest CP signals are indicated by vertical arrows.
The absorption coefficient for (38,0) nanotubes as a
function of linearly-polarized photon energy is shown in
Fig. 6(b) for polarization angles varying from 0◦ (parallel
to tube) to 90◦ (perpendicular to tube).
For the (38,0) nanotube, the RBM phonons have a
computed energy ω = 10.91 meV and an oscillation
period of 379 fs. We simulate the generation of RBM
coherent phonons by pumping with a 50 fs pump pulse,
which is much shorter than the RBM oscillation period.
Polarization-dependent CP spectra for RBM coherent
phonons are shown in Fig. 6(c) for pump polarization
angles φ ranging from 0◦ to 90◦. In these simulations,
the probe polarization is kept fixed parallel to the tube
axis (Type II geometry). As can be seen in the figure, the
CP signal is maximum when the pump is polarized paral-
lel to the tube. As the pump photon energy is varied, we
excite RBM coherent phonons by successively photoex-
citing carriers in different bands. The strongest RBM
CP signals are obtained by pumping at the E44 and E77
transitions. As the pump polarization angle increases,
the RBM CP signal decreases until it is finally quenched
when the pump polarization is perpendicular to the tube.
By taking the area under the RBM CP signal curves in
the vicinity of the computed E44 transition energy near
1.48 eV, we obtain the integrated power as a function of
pump polarization angle shown as black dots in Fig. 6(d).
The integrated CP power can be well fit with a fitting
function of the form A cosp(θ), where A and p = 4.017
are fitting parameters.
The polarization dependence of the RBM CP inte-
grated power in Fig. 6(d) can be understood by examin-
ing Fig. 7. In Fig. 7(b) we plot the coherent RBM diame-
ter oscillations as a function of the pump polarization an-
gle θ. Fitting the theoretical results to A cosp(θ), we ob-
tain p = 2.11 as indicated in the figure. For the RBM di-
ameter oscillations we get p ∼ 2 which makes sense. The
driving function S(t) is proportional to the photogener-
ated carrier density as seen in Eq. (2) and this in turn is
proportional to the squared optical matrix element which
has a cos2(θ) angular dependence3. In Fig. 7(a) we plot
the amplitude of the time-dependent differential trans-
mission oscillations measured by the probe beam after
background subtraction. For the differential transmission
signal, we get p ∼ 2 for Type II and p ∼ 4 for Type I.
This also makes sense. In Type II experiments the probe
polarization is parallel to the tube axis and the pump po-
larization angle θ is varied. In this case the signal should
be proportional to the amplitude of the diameter oscil-
lations. In Type I experiments on the other hand, the
pump and probe polarizations are parallel and we should
get an extra factor of cos2(φ) to account for anisotropy
of the probe beam absorption. In our CP spectroscopy
measurements, we extract the power spectrum of the dif-
ferential transmission oscillations. The CP power spec-
trum should be proportional to the square of the am-
plitude of the differential transmission signal which we
anticipate will give us a cos4(φ) polarization dependence
6
(a)
1.64
(b)
3.65
Type I
Type II
(38,0)
E44
2.11
1.0
0.5
0.0
0.04
0.02
)
.
b
r
a
(
e
d
u
t
i
l
p
m
A
T
T
/
)
A
(
e
d
u
t
i
l
p
m
A
M
B
R
0.00
-90
-60
RBM
10.91 meV
-30
30
0
Angle (degrees)
60
90
FIG. 7: (color online) In (b) we plot the amplitude of RBM
coherent phonon oscillations as a function of the pump polar-
ization angle (θ or φ) for a (38,0) nanotube photoexcited by
a 50 fs pump at the theoretical E44 transition. In (a) we plot
the amplitude of the resulting differential transmission signal
for Type I and II experiments as a function of angles θ and φ
respectively. Our results are fit to A cosp(θ) [A cosp(φ)] where
numerical values of the best fit p are indicated in the figure.
for Type II CP spectroscopy experiments and a cos8(θ)
dependence for Type I CP spectroscopy experiments.
The experimental integrated CP power is obtained by
taking the Fourier transform power spectrum of the time-
dependent differential transmission data shown in Fig. 3
and then computing the area under the E44 peak in the
resulting spectrum. The results for our Type I and Type
II experiments are shown in Fig. 8, where the experi-
mental data points are plotted as downward-pointing red
triangles and the corresponding theoretical predictions
are shown as upward-pointing black triangles. In over-
laying the experimental and theoretical data points, we
subtracted a background from the experimental data to
obtain complete quenching at 90◦. We then rescaled ex-
perimental and theoretical data (both in arbitrary units)
so that the integrated power at 0◦ is equal to unity. We
then fit experimental and theoretical data to functions
of the form A cosp(θ) [A cosp(φ)]. The best fit functions
for our experimental and theoretical results are shown in
Fig. 8 as solid red and dashed black lines, respectively.
For the Type II experiments, where the probe po-
larization is fixed parallel to the average tube orienta-
tion, we get decent agreement between theory and ex-
periment with best fit parameters p = 4.017 for theory
and p = 3.630 for experiment. For the Type I exper-
iments, where pump and probe polarization vectors are
parallel to each other, there is a discrepancy between the-
ory and experiment. As can be seen in Fig. 8(a), the best
)
s
t
i
n
u
.
b
r
a
(
r
e
w
o
P
d
e
t
a
r
g
e
t
n
I
1.0
0.8
0.6
0.4
0.2
E44
(a)
Type I
pump
probe
e
b
u
t
Type II
e
b
u
t
e
b
o
r
p
(b)
pump
8.308
Theory
(38,0)
4.435
Expt
3.630
Expt
4.017
Theory
(38,0)
0.0
0
30
60
30
Angle (degrees)
0
60
90
7
1.0
0.8
0.6
0.4
0.2
0.0
0
30
Standard Deviation
Angle degrees)
0
10
20
30
(a)
Type I
A cos8(
(b)
Type II
A cos4(
60
0
30
Angle (degrees)
60
90
)
s
t
i
n
u
.
b
r
a
(
r
e
w
o
P
d
e
t
a
r
g
e
t
n
I
FIG. 8:
(color online) Integrated RBM coherent phonon
power as a function of (a) θ for Type I and (b) φ for Type
II experiments pumping close to the E44 transition. Experi-
mental data points are red downward pointing triangles fit by
a solid red line. Theoretical data points for a (38,0) nanotube
are black upward pointing triangles and are fit by a black
dashed line. The fitting functions are of the form A cosp(θ)
and A cosp(φ) where the numerical values of the best fit p are
indicated in the figure. The experimental curves have under-
gone background subtraction to obtain complete quenching
at an angle of 90◦ and then rescaled so the integrated power
at an angle of 0◦ is equal to unity.
fit parameters are p = 8.308 for theory and p = 4.435 for
experiment. The experimental fits in Fig. 8 would imply
that the CP intensity scales roughly as cos4(θ) indepen-
dently of the probe polarization. This seems unlikely
given the anisotropy of the absorption coefficient in car-
bon nanotubes. In fact, from our theory, we would expect
p ∼ 4 for Type II and p ∼ 8 for Type I.
We believe that the cause of this discrepancy is most
likely due to misalignment effects. Our sample consists
of an ensemble of nanotubes lying horizontally on a sap-
phire substrate. While the tubes are highly-aligned, they
are not perfectly aligned. To consider the effects of mis-
alignment on the experimentally-measured integrated CP
intensity in a Type I experiment, we assume that the tube
alignment angles on the transferred film are described by
a Gaussian distribution function with a small standard
deviation ∆θ. If the angle between the pump polariza-
tion vector and the ensemble averaged tube axis is θ, then
the angles ϑ between the pump polarization vector and
the axes of each tube in the ensemble are described by a
Gaussian distribution
P (ϑ, θ, ∆θ) =
1
√2π(∆θ)
exp(cid:18)−
(ϑ − θ)2
2(∆θ)2 (cid:19)
(9)
If the integrated RBM coherent phonon power for each
nanotube in the ensemble is given by A cosp(ϑ), the en-
semble averaged integrated power Icp(θ, ∆θ) is obtained
FIG. 9: (color online) Effects of tube misalignment on inte-
grated coherent phonon power for an ensemble of nanotubes
with tube axis orientation angles following a Gaussian distri-
bution. (a) Type I fitting function A cosp(θ) with p = 8 and
(b) Type II fitting function A cosp(φ) with p = 4.
by taking the ensemble average over ϑ
Icp(θ, ∆θ) = A Z ∞
−∞
dϑ P (ϑ, θ, ∆θ) cosp(ϑ)
(10)
In Eq. (10), we extended the integration limits on ϑ to
infinity in the limit of small ∆θ.
If p is a positive in-
teger, Icp(θ, ∆θ) can be found analytically. Otherwise,
it can be evaluated numerically. In the case of Type II
experiments, the ensemble averaged integrated power is
obtained by replacing θ with φ in Eqs. (9) and (10).
The effects of nanotube misalignment on the integrated
CP power fitting function for an ensemble of Gaussian
misaligned tubes is illustrated in Fig. 9, assuming p =
8 for Type I and p = 4 for Type II experiments. As
the standard deviations ∆θ [∆φ] increase, quenching of
the CP intensity is reduced and Icp(θ, ∆θ) [Icp(φ, ∆φ)]
become flatter.
Using this model, we are able to get reasonable fits to
the experimentally measured CP intensity in both Type I
and Type II experiments. We assume a fitting function of
the form A cosp(θ + ∆θ) + B [A cosp(φ + ∆φ) + B], where
A and B are background subtraction and rescaling pa-
rameters and ∆θ [∆φ] is a random Gaussian distributed
misalignment angle (standard deviation ∆θ [∆φ]).
In
our fitting procedure, we set p = 8 for Type I and p = 4
for Type II and use the same standard deviation for the
tube misalignment angles in fitting both Type I and II
data. The results of our fitting procedure are shown
in Fig. 10, where the best fits for Type I and Type
II geometries are shown as dashed black and solid red
lines, respectively. Our best fit standard deviation is
∆θ = ∆φ = 18.7◦, which implies that the tube align-
ment angles on the sapphire substrate are 0◦ ± 9.35◦.
)
s
t
i
n
u
.
b
r
a
(
r
e
w
o
P
d
e
t
a
r
g
e
t
n
I
1.0
0.8
0.6
0.4
0.2
0.0
0
p = 8
Type I
pu m p
pro b e
e
b
u
t
p = 4
Type II
pu m p
e
b
o
r
p
e
b
u
t
Type I
Type II
30
E44
Experiment
90
60
120
Angle (degrees)
150
180
FIG. 10: (color online) Experimental integrated RBM coher-
ent phonon power for the E44 transition as a function of θ for
Type I (black upward pointing triangles) and φ for Type II
(red downward pointing triangles) experiments. Type I exper-
iments are fit to A cosp(θ+∆θ)+B where p = 8 (black dashed
line) and Type II experiments are fit to A cosp(φ+∆φ)+B for
p = 4 (solid red line). A and B are background subtraction
and rescaling parameters and the standard deviations for the
random tube axis misalignment ∆θ and ∆φ are restricted to
be the same for both Type I and II.
For small ∆θ (measured in radians) the nematic order
parameter is S = exp(−2 (∆θ)2) = 0.81.
Although a value of S = 0.81 indicates a strongly
aligned sample, it is not perfectly aligned (S = 1). This
is unlike previous work in the THz regime with these
highly-aligned samples, where the nematic order param-
eter was calculated to be exactly S = 1.17 It is clear that
there is a wavelength dependence of the nematic order
parameter, but we believe this can be explained quali-
tatively: any slight misalignment in the sample would
go undetected in the THz regime, as the wavelength is
much larger than our visible wavelengths that can de-
tect the misalignments. This could explain the differ-
ence in calculated nematic order parameter between our
8
CP measurements and previous THz measurements, but
nonetheless, regardless of the calculated differences with
wavelength, our calculated nematic order parameter is
still quite large and we can confidently confirm the high
degree of alignment of the sample with our CP measure-
ments.
VI. CONCLUSION
In summary, we
investigated the polarization
anisotropy of coherent phonon dynamics in highly-
aligned single-walled carbon nanotubes and measured
RBM coherent phonons as a function of polarization
angle. We saw a very nearly complete quenching of
the RBM for both geometries and extended our results
to determine the degree of alignment of the sample.
Comparing our
results with theory, we performed
simulations of polarization-dependent CP spectroscopy
on a (38,0) zigzag nanotubes and also found a decrease
in CP signal as optical polarization varies from parallel
to perpendicular to the nanotube axis. Using those
simulated results, we also theoretically determined a
cos8(θ) dependence for Type I CP spectroscopy experi-
ments and a cos4(φ) polarization dependence for Type II
CP spectroscopy experiments.
Including misalignment
effects to our fitting, we finally determined the nematic
order parameter of our sample to be 0.81.
Acknowledgments
This work was supported by the National Science
Foundation under grant numbers DMR-0325474, OISE-
0530220, and DMR-0706313, the Robert A. Welch foun-
dation under grant number C-1509, and the Office of
Naval Research (ONR) under contract number 00075094.
Y.-S. Lim and K.-J. Yee are supported by a Korea Science
and Engineering Foundation (KOSEF) grant funded by
the Korean Government (Most) (R01-2007-000-20651-0).
We acknowledge useful discussions with Andrew Rinzler
at the University of Florida.
∗ kono@rice.edu; www.ece.rice.edu/~kono; corresponding
author.
1 R. Saito, G. Dresselhaus, and M. S. Dresselhaus, Physi-
cal properties of carbon nanotubes (World Scientific, Sin-
gapore, 2003).
2 J.-C. Charlier, X. Blase, and S. Roche, Rev. Mod. Phys.
79, 643 (2007).
M. S. S. Dantas, M. A. Pimenta, A. M. Rao, R. Saito,
C. Liu, and H. M. Cheng, Phys. Rev. Lett. 85, 2617 (2000).
6 Z. M. Li, Z. K. Tang, H. J. Liu, N. Wang, C. T. Chan,
R. Saito, S. Okada, G. F. Li, J. S. Chen, N. Nagasawa,
et al., Phys. Rev. Lett. 87, 127401 (2001).
7 A. Hartschuh, H. N. Pedrosa, L. Novotny, and T. D.
Krauss, Science 301, 1354 (2003).
3 H. H. Gommans, J. W. Alldredge, H. Tashiro, J. Park,
J. Magnuson, and A. G. Rinzler, J. Appl. Phys. 88, 2509
(2000).
4 G. S. Duesberg, I. Loa, M. Burghard, K. Syassen, and
S. Roth, Phys. Rev. Lett. 85, 5436 (2000).
8 M. F. Islam, D. E. Milkie, C. L. Kane, A. G. Yodh, and
J. M. Kikkawa, Phys. Rev. Lett. 93, 037404 (2004).
9 H. Ajiki and T. Ando, J. Phys. Soc. Jpn. 62, 2470 (1993).
10 J. P. Lu, Phys. Rev. Lett. 74, 1123 (1995).
11 M. A. L. Marques, M. d'Avezac, and F. Mauri, Phys. Rev.
5 A. Jorio, G. Dresselhaus, M. S. Dresselhaus, M. Souza,
B 73, 125433 (2006).
12 S. Zaric, G. N. Ostojic, J. Kono, J. Shaver, V. C. Moore,
R. H. Hauge, R. E. Smalley, and X. Wei, Nano Lett. 4,
2219 (2004).
13 S. Zaric, G. N. Ostojic, J. Kono, J. Shaver, V. C. Moore,
M. S. Strano, R. H. Hauge, R. E. Smalley, and X. Wei,
Science 304, 1129 (2004).
14 M. F. Islam, D. E. Milkie, O. N. Torrens, A. G. Yodh, and
J. M. Kikkawa, Phys. Rev. B 71, 201401 (2005).
15 J. Shaver, A. N. G. Parra-Vasquez, S. Hansel, O. Portugall,
C. H. Mielke, M. von Ortenberg, R. H. Hauge, M. Pasquali,
and J. Kono, ACS Nano 3, 131 (2009).
16 T. A. Searles, Y. Imanaka, T. Takamasu, H. Ajiki,
J. A. Fagan, E. K. Hobbie, and J. Kono, arXiv (2010),
1001.0524v1.
17 L. Ren, C. L. Pint, L. G. Booshehri, W. D. Rice, X. Wang,
D. J. Hilton, K. Takeya, I. Kawayama, M. Tonouchi, R. H.
Hauge, et al., Nano Lett. 9, 2610 (2009).
18 Y.-S. Lim, K.-J. Yee, J.-H. Kim, E. H. H´aroz, J. Shaver,
J. Kono, S. K. Doorn, R. H. Hauge, and R. E. Smalley,
Nano Lett. 6, 2696 (2006).
19 A. Gambetta, C. Manzoni, E. Menna, M. Meneghetti,
G. Cerullo, G. Lanzani, S. Tretiak, A. Piryatinski, A. Sax-
ena, R. L. Martin, et al., Nature Physics 2, 515 (2006).
20 K. Kato, K. Ishioka, M. Kitajima, J. Tang, R. Saito, and
H. Petek, Nano Lett. 8, 3102 (2008).
21 J.-H. Kim, J. Park, B. Y. Lee, D. Lee, K.-J. Yee, Y.-S.
Lim, L. G. Booshehri, E. H. H´aroz, J. Kono, and S.-H.
Baik, J. Appl. Phys. 105, 103506 (2009).
9
22 G. D. Sanders, C. J. Stanton, J.-H. Kim, K.-J. Yee, Y.-S.
Lim, E. H. H´aroz, L. G. Booshehri, J. Kono, and R. Saito,
Phys. Rev. B 79, 205434 (2009).
23 J.-H. Kim, K.-J. Han, N.-J. Kim, K.-J. Yee, Y.-S. Lim,
G. D. Sanders, C. J. Stanton, L. G. Booshehri, E. H. H´aroz,
and J. Kono, Phys. Rev. Lett. 102, 037402 (2009).
24 Y.-S. Lim, J.-G. Ahn, J.-H. Kim, K.-J. Yee, T. Joo, S.-H.
Baik, E. H. H´aroz, L. G. Booshehri, and J. Kono, ACS
Nano (2010), published online on May 13, 2010.
25 C. L. Pint, Y.-Q. Xu, M. Pasquali, and R. H. Hauge, ACS
Nano 2, 1871 (2008).
26 C. L. Pint, Y.-Q. Xu, S. Moghazy, T. Cherukuri, N. T.
Alvarez, E. H. H´aroz, S. Mahzooni, S. K. Doorn, J. Kono,
M. Pasquali, et al., ACS Nano 4, 1131 (2010).
27 D. Porezag, T. Frauenheim, T. Kohler, G. Seifert, and
R. Kaschner, Phys. Rev. B 51, 12947 (1995).
28 J.-W. Jiang, H. Tang, B.-S. Wang, and Z.-B. Su, Phys.
Rev. B 73, 235434 (2006).
29 C. Lobo and J. L. Martins, Z. Phys. D 39, 159 (1997).
30 A. V. Kuznetsov and C. J. Stanton, Phys. Rev. Lett. 73,
3243 (1994).
31 S. L. Chuang, Physics of Optoelectronic Devices (Wiley,
New York, 1995).
32 W. H. Press, S. A. Teukolsky, W. T. Vetterling, and B. P.
Flannery, Numerical Recipes (Cambridge University Press,
New York, 1992).
|
1607.04802 | 2 | 1607 | 2016-09-07T10:45:25 | Heat production and error probability relation in Landauer reset at effective temperature | [
"cond-mat.mes-hall",
"cond-mat.stat-mech"
] | The erasure of a classical bit of information is a dissipative process. The minimum heat produced during this operation has been theorized by Rolf Landauer in 1961 to be equal to $k_B T \ln 2$ and takes the name of Landauer limit, Landauer reset or Landauer principle. Despite its fundamental importance, the Landauer limit remained untested experimentally for more than fifty years until recently when it has been tested using colloidal particles and magnetic dots. Experimental measurements on different devices, like micro-mechanical systems or nano-electronic devices are still missing. Here we show the results obtained in performing the Landauer reset operation in a micro-mechanical system, operated at an effective temperature. The measured heat exchange is in accordance with the theory reaching values close to the expected limit. The data obtained for the heat production is then correlated to the probability of error in accomplishing the reset operation. | cond-mat.mes-hall | cond-mat |
Heat production and error probability relation in
Landauer reset at effective temperature
Igor Neri1,2,*,+ and Miquel L ´opez-Su´arez1,!,+
1NiPS Laboratory, Dipartimento di Fisica e Geologia, Universit`a degli Studi di Perugia, 06123 Perugia, Italy
2INFN Sezione di Perugia, via Pascoli, 06123 Perugia, Italy
*igor.neri@nipslab.org
!miquel.lopez@nipslab.org
+these authors contributed equally to this work
ABSTRACT
The erasure of a classical bit of information is a dissipative process. The minimum heat produced during this operation has
been theorized by Rolf Landauer in 1961 to be equal to kBT ln 2 and takes the name of Landauer limit, Landauer reset or
Landauer principle. Despite its fundamental importance, the Landauer limit remained untested experimentally for more than
fifty years until recently when it has been tested using colloidal particles and magnetic dots. Experimental measurements
on different devices, like micro-mechanical systems or nano-electronic devices are still missing. Here we show the results
obtained in performing the Landauer reset operation in a micro-mechanical system, operated at an effective temperature. The
measured heat exchange is in accordance with the theory reaching values close to the expected limit. The data obtained for
the heat production is then correlated to the probability of error in accomplishing the reset operation.
Introduction
The minimum energy required to reset one bit of information represents one of the fundamental limits of computation arising
when one bit of information is erased or destroyed. Starting with a system encoding two possible states with the same
probability and finishing the procedure with only one possible state, the variation of entropy in the system is equal to the
difference of entropy between the final state and the initial one, D S = kB ln 1 − kB ln 2 = −kB ln 2. On average this reduction
of entropy has to be accompanied with an increase of heat in the surrounding environment in order to not violate the second
law of thermodynamics, QL ≥ kBT ln 2. This limit takes the name of Landauer limit and was theorized by Rolf Landauer in
the 60's1, and remained untested for over fifty years. The development of computational methods allowed to evaluate tiny
amounts of heat exchanged and recent technical advances in micro- and nano-fabrication made possible to recently test the
validity of the Landauer principle. In particular, the first experimental verification of the Landauer principle has been carried
out by B´erut et al. using a colloidal particle trapped in optical tweezers2. More recently Jun et al. presented similar results
considering a colloidal particle in a feedback trap3. Finally, the last experimental verification of the Landauer limit has been
performed by Hong et al. in a nanomagnetic system4. In this case information is encoded in the magnetization state of the
system while an external magnetic field is used in order to flip between two preferred magnetization states. The evaluation
of the heat produced during the reset operation was demonstrated to be compatible with the Landauer limit. Even if there is
no doubt nowadays on the validity of the Landauer principle, the test on micro-mechanical systems is still missing. We have
recently shown that it is possible to use electro-mechanical devices to accomplish basic5 and complex logic operations6 with
an arbitrarily low energy expenditure. Therefore the possibility to use this class of devices for memory storage completes the
logic architecture for a complete computing device.
In the following we show the measurement on heat production when performing the reset operation in a novel memory unit
based on a bistable mechanical cantilever at effective temperature. The results are in good agreement with the Landauer limit.
The dependence of the dissipation with the error rate is also investigated showing a trend in accordance with the theory10.
Results
Bistable micro-mechanical system
The mechanical system used to perform the experiment is depicted in Figure 1 (a). A triangular micro-cantilever, 200µm long,
is used to encode one bit on information. In order to obtain two stable states two magnets with opposite magnetization are
placed on the tip of the cantilever and on a movable stage facing the cantilever. In this way, depending on the distance between
the magnets, d, and the relative lateral alignment, D x, it is possible to induce bistability on the system. Figure 1 (b) shows
the potential energy as a function of d reconstructed from the probability density function of the position of the cantilever
at equilibrium, r (x,d) = A exp(−U(x,d)/kBT ), which implies that U(x,d) = −kBT lnr (x,d) + U0
2. When the magnets are
far away the effect of the repulsive force is negligible, the system is then monostable and can be approximated to a linear
system. Decreasing the distance the repulsive force between magnets tends to soften the system up to the point where two
stable positions appear. The effect of reducing even more d is to enlarge the separation of the rest states and to increase the
potential barrier separating these two wells. Eventually, when the distance between the magnets is small enough, the system
remains trapped in one well for a period of time larger than the relaxation time of the system. Logic states are encoded in
the position of the cantilever tip: logic 0 for x < 0 and logic 1 for x > 0. The proposed system presents intrinsic dissipative
processes that depend on the maximum displacement of the cantilever tip6. The minimum heat produced when performing a
physical transformation of the system is proportional to x2
max. In our setup it is not possible to reduce the separation between
the two potential wells to a value that bounds the heat produced by intrinsic dissipation below the Landauer limit. Increasing
the effective temperature increases the value for the Landauer limit making possible to have negligible intrinsic dissipation.
A piezoelectric shaker is used to excite the structure with a band limited white Gaussian noise to mimic the effect of an
arbitrary temperature. In the present experiment the white noise is limited to 50kHz, well above the resonance frequency of
the free cantilever ( f0=5.3kHz). The dependence of the effective temperature with the root-mean-squared voltage supplied to
the shaker is reported in Figure 1 (c). The red dot, corresponding to an effective temperature of Teff = 5 × 107 K, highlights
the condition considered in the present case. The solid line represent the expected trend where T (cid:181) V 2
7. The effective
rms
temperature has been estimated computing the power spectral density (PSD) of the system at various piezoelectric noise
excitation voltages. The obtained curves have been fitted with Lorenzian curves taking as reference for the calibration the one
at room temperature (T =300K). The other curves have been used to extract the only varying parameter Teff, corresponding to
the effective temperature of the system under external excitation. Finally two electrostatic probes, placed one on the left and
the other on the right of the cantilever, are used to apply a negative and positive forces respectively. When a voltage different
to zero is applied on one probe the cantilever feels an attractive electrostatic force toward the probe due to the polarization of
the cantilever itself. The voltage on the probes, the distance between the magnets and their time evolution are used to specify
the protocols used in order to change the bit stored in the system as described in the following subsection.
Reset protocol
Varying the magnets distance d the barrier separating the two stable wells varies from the minimum value B0 for d=3.65 a.u.
to Bmax for d=2.8 a.u. Applying a voltage on one probe corresponds to apply a force toward the probe itself. Thus a voltage
on the left probe corresponds to a negative force and a voltage on the right probe corresponds to a positive force. Details of
the force calibration are presented in methods section. In Figure 2 (a) the protocol followed to reset the bit to the state 0 and
1 is presented. The procedure is similar to the ones presented in Refs. 8 and 2. Initially the barrier separating the two stable
states is removed moving the magnet away (red curve in the first panel) making the system monostable. Once the barrier is
removed we apply a negative (positive) force to reset the bit status to 0 (1) applying a finite voltage VL (VR) on the left (right)
probe. This is represented by the magenta curve in Figure 2 (a). Once the force is applied we restore the barrier to its original
value. Finally, we remove the lateral force finishing in the original parameters configuration. At the end of the operation, if
there are no errors the cantilever position encodes the desired bit of information. In order to be sure to perform the operation
starting from both initial states, we mimic a statistical ensemble where the initial probability is 50% to start in the left well
and 50% in the right well. A trace of the cantilever tip position, x, is shown in Figure 2 (a) (black line), where the dashed
blue lines represent the two stable positions once the barrier is restored. When the barrier is removed the system goes from
the local prepared state to an undefined state with a free expansion, the entropy on the system thus increases in a irreversible
manner8, 9. This increment is related to uncontrollable transitions from one well to the other once the barrier height is close to
kBT . These large excursions of the cantilever can be seen in the time series of position. Figure 2 (b) shows a representation
of the time evolution of the potential energy during the set of the bit to the logic state 1. In a first step the barrier is removed
allowing the system to oscillate in a monostable potential landscape. Then the potential is slightly tiled and when the barrier
is restored the system is confined in the desired state. After these stages the bit is set to the state 1. In order to have a reliable
measure of the heat produced during the considered operations, reset protocols are repeated for 800 times in order to have a
large enough statistic.
Heat vs error probability
In the optimal case the initial configuration is a mixed logical 0 and 1 where both states have the same probability while the
final configuration is the selected state with a 100% probability. This corresponds to an entropy variation of D S = −kB ln(2)
and a minimum heat produced of Q ≥ −TD S = QL. It has been shown that the bound QL = −kBT ln(2) applies only for
11. In
symmetric potentials. Considering asymmetries on the system the minimum produced heat can be lowered below QL
our setup the system is slightly asymmetric and we have evaluated the variation of entropy from the initial to the final state
2/8
from the probability density function of the tip position, r (x), being D SG = −0.61kB D SS = −0.68kB for the Gibbs and
Shannon entropy respectively, both close to −kB ln(2).
If we consider the possibility to commit errors during the reset operation the heat produced becomes a function of the
probability of success10:
Q(Ps) ≥ kBT [ln(2) + Ps ln(Ps) + (1 − Ps)ln(1 − Ps)]
(1)
where Ps is the probability of success or success rate. When Ps is 0.5 no reset operation is performed and thus there is no
minimum heat to be produced during the operation3, 10.
In Figure 3 (a) we present the average heat produced for the reset operation as function of the lateral alignment of the
counter magnet D x. When the system is aligned closely to perfection (i.e., D x ≈ 0) we estimate a heat production slightly
above kBT and below two times QL. Asymmetrizing the potential, by means of setting D x 6= 0, the heat produced tends to
decrease reaching values this time below QL. However, in this conditions the error rate in performing the reset operation have
a major role, in fact in this configuration the probability of success, Ps, decrease rapidly. This is represented by the color map
of dots in Figure 3 (a), where green represents higher success rate while blue represents a higher probability of error. In Figure
3 (b) the success rate of the reset operation is reported as function of the lateral alignment. Solid violet circles represent the
overall success rate while red and black symbols represent the error rate for resetting to 1 or to 0 respectively. Circles are used
to report the error probability for the same initial and final state while crosses are used for 0 to 1 and 1 to 0 transitions. For
instance let us consider the case where D x < 0: the counter magnet is moved towards the right and as a consequence the 0
state is more favorable respect the 1 state. From Figure 3 (b) we can see that for D x < 0 the probability of resetting toward 0
is almost 100% while the probability of resetting toward 1 decreases rapidly reaching values below 50%. The same behavior
is present in the case D x > 0, where the counter magnet is moved to the left, where the state 1 is more favorable.
We can now correlate the heat produced to the probability of success for resetting to 0 and 1 as presented in Figure 3 (c).
Dashed lines represent the Landauer limit for a 100% of success rate (≈ 0.7kBT ). While in both cases the heat produced is
above the Landauer limit, in the reset to 0 case the obtained values are very close to QL. As expected, decreasing the success
rate the obtained values goes below the Landauer limit for both cases accordingly to Equation 1.
As it is well known the adiabatic limit in presence of dissipation mechanisms like viscous damping can be only reached if the
operation is performed slowly when these mechanisms are negligible. We increased the protocol time for the reset operation
from 0.25s up to 3.5s. The results are presented in Figure 3 (d). Increasing the protocol time decreases the heat production
reaching values well below the Landauer limit. Notice that in these cases where Q < QL the Ps is well below 1 since the
system has more time to relax and therefore tends to thermalize before the reset operation is correctly performed. In fact for
protocols lasting more than 1 s the success rate is below 75%.
Discussion
We have measured the intrinsic minimum energy dissipation during the reset of one bit of information in a micro-mechanical
system. We have considered a completely different physical system respect to the existing literature, i.e., micro-electro-
mechanical system. To achieve these results we have performed the experiment at an effective temperature of 5 × 107 K
in order to make the intrinsic dissipation of the mechanical structure negligible respect to the thermodynamic contribution. In
these conditions we have reached values of heat produces consistent with the Landauer limit approaching it closely. Moreover
we presented experimental data relating the minimum heat produced with the probability of success of resetting one bit of
information. Nowadays where there is a lot of attention on micro-electro-mechanical systems able to perform computation
at arbitrary low energy, the achieved results have a significant importance in the development of new computing paradigms
based on systems different from the well established CMOS technology.
Methods
Setup preparation and calibration
The micro-cantilever used is a commercial atomic force microscopy (AFM) probe (NanoWorld PNP-TR-TL12). It is long
200µm, with a nominal stiffness k=0.08Nm−1 and a nominal resonace frequency of 17kHz. A fragment of NdFeB (neodymium)
magnet is attached to the cantilever tip with bi-component epoxy resin. To set the magnetization to a known direction the sys-
tem is heated up to 670K, above its Curie temperature13 in the presence of a strong external magnetic field with the desired
orientation. With this additional mass the resonance frequency decreases to 5.3kHz. The quality factor of the system has been
estimated from the power spectral density of the displacement, x, fitted with a Lorenzian curve, giving a value of Qf = 320.
The deflection of the cantilever, x, is determined by means of an AFM-like laser optical lever. A small bend of the cantilever
provokes the deflection of a laser beam incident to the cantilever tip that can be detected with a two quadrants photo detector.
3/8
The laser beam is focused on the cantilever tip with an optical lens (focal length f =50mm). For small deflections the response
of the photo detector remains linear, thus x = rxD VPD, where VPD is the voltage difference generated by the two quadrants of
the photo detector. In order to determine rx we look at the frequency response of the system as in6 under thermal excitation.
The relation between the measured voltage and the expected displacement gives rx =1.8365 × 10−5m V−1. The system is
placed in a vacuum chamber and isolated from seismic vibrations to maximize the signal-to-noise ratio. All measurements
were performed at pressure P =4.7 × 10−2 mbar.
Effective temperature estimation
As the system is modeled as a harmonic oscillator with one degree of freedom, according to the Equipartition Theorem the
thermal energy present in the system is simply related to the cantilever fluctuations as 1
2 khx2i. According to Parseval's
0 G(w )2 dw where G(w ) stands for the PSD of the system. This takes the form of a Lorenzian function
2 kBT = 1
theorem hx2i =R
G(w ) =s 4kBT
Qfkw 0vuut
(cid:16)1 −
w 2
w 2
0(cid:17)2
1
+ 1
Q2
f
w 2
w 2
0
where w = 2p
only free parameter, T , from the measured PSDs to the expected function G(w ).
f . Once the system has been calibrated at room temperature we estimated the effective temperatures fitting the
Force calibration
A set of two electrodes (see Fig. 1(a)) is used in order to polarize the cantilever producing a bend on the mechanical structure.
Electrostatic forces depend on the voltage applied to the electrodes, i.e. VL and VR for the left and right electrode respectively.
Since the restoring force of the cantilever can be expressed as Fk = −kx, the relation between applied voltage to the probe and
the force acting on the cantilever has been estimated in static conditions assuming Fk = Fel. The relation between the applied
voltages, VL and VR, and the electrostatic force Fel is then fitted with a 9th degree polynomial.
Heat production evaluation
The work performed on the system along a given trajectory x(t) is given by the integral14, 15:
W =Z
0
t p
M(cid:229)
k=1
¶ U(x,lll )
¶l
k
¶l
k
¶ t
dt
(2)
where t p is the protocol time duration, U(x,lll ) is the total potential energy of the system and lll
is a vector containing all the
M control parameters. In our case we have two controls parameter, the voltage applied to the piezoelectric stage to control
the energy barrier, and the electrostatic forces. To obtain the heat produced Q we have to consider the variation in internal
energy, D E. Ideally the potential energy at both the bottom wells is the same, however considering asymmetries on the system
D E can be different from zero. The evaluation of the total energy variation D E is obtained from the reconstructed potential
energy, U, and the variation of the kinetic energy. The latter quantity is however negligible even for the shortest t p, where the
total kinetic energy variation is one order of magnitude lower than QL (1.4 × 10−17J versus 5.3 × 10−16 J). Finally the heat
produced is obtained from Q = W − D E.
References
1. Landauer, R. Irreversibility and heat generation in the computing process. IBM journal of research and development 5,
183 -- 1911961.
2. B´erut, A. et al. Experimental verification of Landauer's principle linking information and thermodynamics. Nature 483,
187 -- 189 (2012).
3. Jun, Y., Gavrilov, M. & Bechhoefer, J. High-precision test of Landauer's principle in a feedback trap. Physical review
letters 113, 190601 (2014).
4. Hong, J., Lambson, B., Dhuey, S. & Bokor, J. Experimental test of Landauer's principle in single-bit operations on
nanomagnetic memory bits. Science Advances 2 (2016).
5. Lopez-Suarez, M., Neri, I. & Gammaitoni, L. Operating micromechanical logic gates below kBT: Physical vs logical
reversibility. In Energy Efficient Electronic Systems (E3S), 2015 Fourth Berkeley Symposium on, 1 -- 2 (IEEE, 2015).
6. Lopez-Suarez, M., Neri, I. & Gammaitoni, L. Sub kBT micro electromechanical irreversible logic gate. Nature Commu-
nication 7 12068 (2016).
4/8
¥
7. Venstra, W. J., Westra, H. J. & van der Zant, H. S. Stochastic switching of cantilever motion. Nature communications 4
(2013).
8. Bennett, C. H. The thermodynamics of computation -- a review. International Journal of Theoretical Physics 21, 905 -- 940
(1982).
9. Crooks, Gavin E. Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences.
Physical Review E 60.3 2721 (1999).
10. Gammaitoni, L., Chiuchi´u, D., Madami, M. & Carlotti, G. Towards zero-power ICT. Nanotechnology 26, 222001 (2015).
11. Sagawa, Takahiro. Thermodynamic and logical reversibilities revisited. Journal of Statistical Mechanics: Theory and
Experiment 2014.3 P03025 (2014)
12. NanoWorld - AFM tip - PNP-TR-TL - Pyrex-Nitride. http://www.nanoworld.com/pyrex-nitride-triangular-silicon-nitride-
tipless-cantilever-afm-tip-pnp-tr-tl. Accessed: 2016-05-30.
13. Ma, B. et al. Recent development in bonded NdFeB magnets. Journal of magnetism and magnetic materials 239, 418 -- 423
(2002).
14. Seifert, U. Stochastic thermodynamics, fluctuation theorems and molecular machines. Reports on Progress in Physics 75,
126001 (2012).
15. Douarche, F., Ciliberto, S., Petrosyan, A. & Rabbiosi, I. An experimental test of the Jarzynski equality in a mechanical
experiment. EPL (Europhysics Letters) 70, 593 (2005).
Acknowledgements
The authors gratefully acknowledge useful discussion with L. Gammaitoni. The authors acknowledge financial support from
the European Commission (FPVII, Grant agreement no: 318287, LANDAUER and Grant agreement no: 611004, ICT- En-
ergy).
Author contributions statement
M.L.S. and I.N. designed the experiment, performed the measurements, analyzed the measured data and contributed to the
writing of the manuscript.
Data availability statement
The data that support the findings of this study are available from the corresponding author upon request.
Additional information
Competing financial interests The authors declare no competing financial interests.
5/8
Figure 1. Schematic of the whole system and measurement setup. Lateral view of the whole system and measurement setup.
Two magnets with opposite magnetic orientations are used to induce bistability in the system. Two electrodes are used to
apply electrostatic forces on the mechanical structure: VL and VR to force the cantilever to bend to the left (negative x) and to
the right (positive x) respectively. The magnetic interaction can be engineered by changing geometric parameters such as d
and D x. (b) Color-map of the reconstructed potential energy as function of the distance between the magnets, d. The distance
is expressed in arbitrary units proportional to the voltage applied to the piezoelectric stage. Decreasing the distance between
the magnets the potential energy softens and eventually two stable states appear. (c) Dependence of the effective temperature,
Teff, with the root mean square of the white Gaussian voltage applied to the piezoelectric shaker. The red dot represent the
condition accounted for the experimental data presented.
6/8
Figure 2. Reset protocol. (a) Protocol used to perform the reset operation. In order to account for all the possible transitions
we considered the reset to 1 (first two columns) and reset to 0 (last two columnss) starting from both 0 and 1 states. The first
row depict the protocol used for removing the barrier. Once the barrier is removed a lateral force is applied (second row).
The resulting displacement of the cantilever tip is represented in the third row. Once all the forces are removed and the
barrier is restored the cantilever tip encodes the final state. (b) Schematic time evolution of the potential energy and state of
the system for the case presented in the first column of panel (a). The operation starts from a double well potential and in the
first step of the protocol the barrier is removed. During the next step the potential is tilted and the barrier is restored to its
initial value. Finally, the lateral force is removed recovering the initial condition where the barrier between wells is at its
maximum and the electrostatic forces are equal to zero.
7/8
Figure 3. Produced heat and probability of success for the reset operation. (a) Average heat produced during the reset
operation as function of the lateral alignment D x. For D x < 0 the counter magnet is moved to the right and the 0 state (x < 0)
is favorable. Accordingly, for D x > 0 the 1 state is more favorable. Introducing an asymmetry on the potential Q decreases,
which is accounted to the probability of success, Ps, that tends to decrease (Ps is encoded in the color map). (b) Success rate
of the reset operation as function of lateral alignment. Solid violet circles represent the overall success rate while black and
red symbols account for the success rate resetting to 0 and 1 state respectively. The maximum overall success rate is present
when the system is almost symmetric, D x ≈ 0. (c) Relation between success rate and heat dissipated. Red circles correspond
to the resetting to 1 case while black ones correspond to the resetting to 0. (d) Dependence of Q with the protocol time
duration, t p. As t p is increased the effects of frictional phenomena becomes negligible and the produced heat should
approach the thermodynamic limit. However, for large t p the reset operation fails giving a wrong logic output. In these cases,
where the error probability is high, the produced heat is clearly below the Landauer limit. Inset shows the obtained relation
between error probability (1-Ps) and produced heat. The data are compatible with the minimum energy required for a given
error probability as predicted by Eq. 1, represented by dashed line.
8/8
|
1109.0619 | 1 | 1109 | 2011-09-03T13:19:56 | Magnetic Fields Effects on the Electronic Conduction Properties of Molecular Ring Structures | [
"cond-mat.mes-hall"
] | While mesoscopic conducting loops are sensitive to external magnetic fields, as seen by observations of the Aharonov-Bohm (AB) effect in such structures, the field needed to observe the AB periodicity in small molecular rings is unrealistically large. The present study aims to identify conditions under which magnetic field dependence can be observed in electronic conduction through such molecules. We consider molecular ring structures modeled both within the tight-binding (H\"uckel) model and as continuous rings. In fact, much of the observed qualitative behavior can be rationalized in terms of a much simpler two-state model. Dephasing in these models is affected by two common tools: the B\"uttiker probe method and coherence damping within a density matrix formulation. We show that current through a benzene ring can be controlled by moderate fields provided that (a) conduction must be dominated by degenerate (in the free molecule) molecular electronic resonances, associated with multiple pathways as is often the case with ring molecules; (b) molecular-leads electronic coupling must is weak so as to affect relatively distinct conduction resonances; (c) molecular binding to the leads must be asymmetric (e.g., for benzene, connection in the meta or ortho, but not para, configurations) and, (d) dephasing has to be small. Under these conditions, considerable sensitivity to an imposed magnetic field normal to the molecular ring plane is found in benzene and other aromatic molecules. Interestingly, in symmetric junctions (e.g. para connected benzene) a large sensitivity of the transmission coefficient to magnetic field is not reflected in the current-voltage characteristic. Although sensitivity to magnetic field is suppressed by dephasing, quantitative estimates indicate that magnetic field control can be observed under realistic condition. | cond-mat.mes-hall | cond-mat | Magnetic Fields Effects on the Electronic Conduction Properties of
Molecular Ring Structures
Dhurba Rai, Oded Hod and Abraham Nitzan
School of Chemistry, Tel Aviv University, Tel Aviv 69978, Israel.
Abstract
While mesoscopic conducting loops are sensitive to external magnetic fields, as is
pronouncedly exemplified by observations of the Aharonov-Bohm (AB) effect in such
structures, the small radius of molecular rings implies that the field needed to observe the AB
periodicity is unrealistically large. In this paper we study the effect of magnetic field on
electronic transport in molecular conduction junctions involving ring molecules, aiming to
identify conditions where magnetic field dependence can be realistically observed. We
consider electronic conduction of molecular ring structures modeled both within the tight-
binding (Hückel) model and as continuous rings. We also show that much of the qualitative
behavior of conduction in these models can be rationalized in terms of a much simpler
junction model based on a two-state molecular bridge. Dephasing in these models is affected
by two common tools: the Büttiker probe method and coherence damping within a density
matrix formulation. We show that current through benzene ring can be controlled by
moderate fields provided that several conditions are satisfied: (a) conduction must be
dominated by degenerate (in the free molecule) molecular electronic resonances, associated
with multiple pathways as is often the case with ring molecules; (b) molecular-leads
electronic coupling must be weak so as to affect relatively distinct conduction resonances; (c)
molecular binding to the leads must be asymmetric (e.g., for benzene, connection in the meta
or ortho, but not para, configurations) and, (d) dephasing has to be small. When these
conditions are satisfied, considerable sensitivity to an imposed magnetic field normal to the
molecular ring plane is found in benzene and other aromatic molecules. Interestingly, in
symmetric junctions (e.g. para connected benzene) the transmission coefficient can show
sensitivity to magnetic field that is not reflected in the current-voltage characteristic. The
analog of this behavior is also found in the continuous ring and the two level models.
1
Although sensitivity to magnetic field is suppressed by dephasing, quantitative estimates
indicate that magnetic field control can be observed in suitable molecular conduction
junctions.
2
1. Introduction
that comprise
junctions
through molecular
transmission
Controlling electron
molecular ring structures by magnetic fields is considered challenging because the required
field strengths are believed to be unrealistically high, of the order of the Aharonov-Bohm
410 Tesla for typical molecular rings.1, 2. In contrast, it was demonstrated
period of ~
theoretically3-6 that the conductance of a nano-size ring can be significantly modulated by
relatively moderate magnetic fields (< 50 Tesla). This large sensitivity to an external
magnetic field results from the presence of sharp resonances that are possible only for low
coupling between the molecular-bridge and the metal-contacts.
Recent studies of electronic conduction through molecular ring structures such as
benzene, biphenyl, azulene, naphthalene and anthracene as well as carbon nanotubes, by us7
and others8-11 have shown that although the net current through these molecules at low metal-
molecule coupling is relatively small, induced circular currents can be considerable. The
magnitude of such a voltage driven ring current depends significantly on the metal-molecule
coupling strength, while its very existence depends on junction geometry, specifically on the
location and symmetry of the molecule-metal contact along the circumference of the ring.
The magnetic field associated with such a circular current at the center of a molecular ring
can be quite significant, for instance, we have found ~ 0.23 Tesla at 2 Volt bias in a tight-
binding model of a meta-connected benzene bridge.7 Possible exploitations of such high local
magnetic fields at the molecular level could come through the realization of carbon nanotubes
as molecular solenoids,11-13 or by controlling the alignment of spin orientation of magnetic
ions embedded in the bridge as suggested in Ref. 14.
Voltage driven circular currents in molecular ring structures are similar in nature to
the persistent currents induced in isolated mesoscopic rings that are threaded by magnetic
fluxes. Both phenomena originate from splitting of degenerate ring electronic states
characterized by opposite orbital angular momenta, either by the magnetic fields or by the
voltage bias even in the absence of external magnetic field.15-19 20 Theoretical studies have
shown that magnetic flux induced persistent currents can be controlled by external
radiation.21, 22 Also, theory indicates that ring currents can be induced by polarized light,23-25
twisted light,26 and other optical coherent control methodologies,27-29 Such optically induced
circular currents can be effectively controlled by externally applied magnetic fields,30 or,
conversely, can be used to control the local magnetic field at the ring,31-33 thereby opening the
possibility to control the orientation of an impurity spin at the ring center.
3
In a recent preliminary study,34 we have reconsidered the possibility of controlling
electrical conduction characteristics of molecular ring structures by static uniform moderate
magnetic fields, following the lead of Refs. 3-6, which indicate that large sensitivity to
magnetic fields may be found in junctions where (a) electronic state degeneracy leads to
interference that can be suitably tuned by the magnetic field, and (b) weak molecule-metal
coupling results in sharp transmission resonances. We have shown that for certain geometries
and under specified conditions magnetic field effects on molecular ring conduction can
become observable. In the present study we analyze in detail the role of molecular geometry
and dephasing on the transport properties of molecular rings such as benzene, biphenyl and
anthracene, subject to external magnetic fields by means of tight-binding (Hückel) molecular
ring models as well as a scattering matrix approach to transmission through continuum loops.
The close similarity between the results obtained from these different models indicates their
generic nature, and provides evidence to the integrity of results obtained for the effect of a
magnetic field in the limited-basis tight binding model (the London approximation). In
addition we show that the essential physics underlying the observed magnetic field
dependence of conduction can be obtained already from a suitably constructed two-state
model. The important role of interference processes implies that the system behavior will
strongly depend on the junction geometry that determines the transmission pathways and on
the effect of dephasing (decoherence) processes resulting from thermal motions in the
junction. In this regard we note that an earlier study35 indicates that increasing electron-
vibration coupling in the junction may lead to sharper resonance structure in the current
dependence on the magnetic field, and therefore to larger sensitivity of conduction to such
field. This however is not simply related to pure dephasing as the effect discusses by Ref. 35
increases at lower temperatures.
The structure of the paper is as follows: in section 2, we present three different models
for electron transmission under the influence of an external magnetic field: (i) A tight-binding
(Hückel) model, which is analyzed using the steady state approach described in our earlier
work7, 36, 37 and (ii) continuum ring model analyzed using scattering theory6. We also describe
(iii) a simple 2-level model that captures the essence of the observed behavior. Section 3
presents results from our model calculations that describe (a) the magnetic field effect on the
transmission probability and the current voltage characteristics of simple molecular ring
junctions of various symmetries; (b) comparison between the tight-binding, continuum and
two-level models and (c) the effect of structure and geometry on the dependence of junction
4
transport properties on the applied magnetic field. The effect of dephasing processes on these
behaviors is discussed in Section 4. Section 5 summarizes our main results and discusses
their implications for further study.
2. Models and Methods
In this section we describe the models used in this work to analyze magnetic field effects on
electron transport through molecular rings. The tight-binding (Hückel) model seems to be the
most suitable for a simple description of molecular transport, however we will see in the
following sections that the main characteristics of the transport behavior are found also in the
continuous ring model. In fact, much of the physics is already contained in an ever simpler
model based on a two-level bridge. These models are described below.
(a) The tight-Binding Model: Scattering theory on a 1-d lattice
We consider a molecular junction formed by a ring molecule bridging the metal leads
(L, R) through two chosen sites on the ring. The molecule is described by a tight-binding
(Hückel) model with on-site energies Mα and nearest-neighbor interactions Mβ . The metal
contacts (L, R) are modeled as infinite 1-dimensional tight-binding periodic arrays of atoms
with lattice constant a and on-site energies and nearest-neighbor coupling matrix elements
Kα and
R LK β K
, respectively. A sketch of such a molecular junction is shown in Fig.
)
,
(
1
Fig. 1. A tight-binding model for current conduction through a molecular ring structure connected to
two 1-dimensional metal leads L and R, at bias voltage V. A static uniform magnetic field B
is
applied perpendicular to the molecular plane.
In the site representation, this TB Hamiltonian is given as
HH
H
H
V
V
,
LM
RM
L
R
M
(1)
5
where
K
1
2
m
1
2
m
H
K
2
p
β
K
p
α
(2)
n
1
e A
nn
nn
1
M R LK
n
;
,
,
, where m and e are the electron mass and charge. We
K n
K n
V
Mm R LKn
nm
mn
β
;
;
,
,
(3)
KM
LM
where n is an orthogonal set of atomic orbitals centered at atomic sites n
). The indices K = L, R and M correspond to the subspaces of the left (L),
M R LK n
(
,
,
right (R) and molecule (M), respectively.
The model (1)-(3) is supplemented by an additional magnetic field B
applied to the
junction, so that the kinetic energy operator of an electron is modified according to
2
assume that the field is uniform, applied in the bridge region only and, for the planar ring-
molecules considered, perpendicular to the molecular plane. In the standard London
approximation38 one (a) represents this field by a vector potential in the symmetric gauge,
, and (b) modifies the finite basis of field-free atomic orbitals
,
1
( )A r
B r
n
r
n
2
where r is measured from the center of atom n, so as to account for the phase difference
between wavefunctions centered on different atomic sites,
A r
i e
) n
(
/
r
r
r
. This leads to a tight binding (Hückel) Hamiltonian with coupling between
A r
A
n
n
atomic sites given by
i e
A
A
r
/
*
dr
r e
r V
n
m
(5)
n
mn
m
For nearest neighbor interactions, r in the exponent is approximately replaced by
/ 2
r
r
, leading to
n
m
where
(4)
n
e
n
n
M e
M
mn
i
mn
n m M
; ,
,
where
mn
A A
m
n
e
B
r
m
r
n
r
m
r
n
e
4
r
r
m n
2
r
n
r
m
B
e
2
and the molecular Hamiltonian is now given by
(6)
(7)
6
)
2
2
)
K
.
K
H
M
mn
i
mn
e
)
r
m
(
E
)
e
(
E
(8)
(
E)
(
E
)
K
i
mn
n n
4
K
m n
n m
(i 2)
Note that
position vectors
2
B
0
(
M
M
n m M
n M
,
is twice the area of the triangle spanned by the vectors, so
r
n
B is the magnetic flux through the triangle spanned by the
, where
mn r
r
/h e
,
is the flux quantum.
, and 0
To evaluate the transmission coefficient associated with this setup one could use the
non-equilibrium Green function method, but here we follow the scattering method used in
Refs. 36 and 37 that gives an easier access to bond currents and, in its density matrix version,
can be generalized to account (approximately) for dephasing processes. In this approach, a
wire or a network of wires described by a tight-binding Hamiltonian is made to carry current
by injecting electrons at one of its sites (source site). At the same time absorption is affected
at the “end sites” of other wires by adding the self-energy term
E
(
)
i
K
2
2
to the site energy. This “absorption” represents the infinite extent of the wire and makes it
possible to describe dynamics in an infinite system by a finite system calculation. For
n
electrons injected at energy E, The steady state wavefunctions are written in the site
representation in the form
C E t n
,
n
n
The orbital coefficients or amplitudes
t
n
nC E obtained by solving the Schrödinger equation
(
)
for steady state situation provide the particle current between any two nearest-neighbor sites
(n, m) on wire K as
2
K
which, in the presence of magnetic field (applied in the molecular region), becomes
2
K
The transmission probability through any exit segment or bond is calculated as the ratio of
inter-site current (bond current) to the incoming particle current, i.e.,
K
J
E
(
)
nm
E
J
(
)
In
C C
Im(
n
)
n m
( ,
)
)
n m
( ,
)
K M
C C
n m
C n
n
(12)
(11)
(10)
(13)
Im(
e
(9)
i E
/
K
(
E
)
J
K
nm
J
K
nm
K
nm
(
E
)
(
E
)
E t
,
e
.
,
m
nmi
.
K
7
.
j
n
j
I n
j
(14)
N segments of nj bonds (
) on which the current has been determined to be Ij, the
Under certain conditions, a bond current in a biased molecular ring structure may exceed the
cI has developed in the ring. In a
net transport current – a signature that a circular current
previous paper,7 we have defined the circular current as the sole source of (bias induced)
magnetic flux through the ring. For a ring comprising n identical bonds, divided by nodes into
N
j n
1
circular current in a ring is given by7
1
n
j
It was also useful to define the circular transmission coefficient7 as the ratio of circular
current to the incoming current as
EJ
(
)
c
J
E
(
)
In
note that this number can be larger than 1.
For a finite bias voltage, the net bond current between any two nearest neighbor sites
in a two-terminal junction is obtained from the Landauer formula
)(E
f K
RLK μ K
where
and
are the Fermi functions and chemical potential of the left
)
,
(
and right leads, respectively. In the calculations reported below, unless otherwise stated, we
have taken the leads temperature to be zero, assumed that the potential bias falls on the metal-
eV
FE
and
molecule interfaces and considered symmetric potential drop, that is
/ 2
L
FE
e
R
))
dE
(
E
) (
f E
(
R
f E
(
L
)
(16)
(15)
V
(
)
/ 2
.
(
E
)
eV
I
mn
c
,
,
I
c
nm
(b) Scattering in the continuous ring model
Next, we consider the continuous ring model, in which the junction comprises a one-
dimensional (1D) conducting ring connecting between two 1D leads (Fig. 2a). The symmetry
of the scattering process is imposed through the angle γ between the leads, 39 and the
properties of the ring-lead contact are embedded in the imposed junction scattering
amplitudes (Fig. 2b) as detailed below.
8
(b)
(a)
y
U2
U1
L1
D1
L2
D2
R1
R2
x
Fig. 2: Schematic representation of the scattering model. Panel (a): assignment of the wave amplitudes on the
different parts of the system. Panel (b): junction scattering amplitudes.
We assume that the electrons can be represented by plane waves traveling along the
ikl
ikl
A e
A e
where l is the electron path length (on the ring l R
,
ring with the form
2
1
where is the angle traversed by the electron, and R – the ring radius) and 1,2A are the
amplitudes of the respective waves (In Fig. 2a these amplitudes are denoted L, U, D and R in
different segments of the ring and the leads). We use the standard notation where
negative(positive) k values represent (counter-)clockwise propagating waves. The ring-lead
coupling is modeled by assigning scattering amplitudes at the ring-leads junctions40-46 as
2c
is the probability of an electron approaching the junction from the lead
shown in Fig. 2b.
to be back scattered into the lead, is the probability of an electron approaching the junction
2a is the probability of an electron approaching the junction
from the lead to mount the ring,
from one of the arms of the ring to be back scattered into the same arm, and
2b is the
probability of an electron approaching the junction from one of the arms of the ring to be
transmitted into the other arm. Based on current conservation considerations the scattering
matrix can be shown to be unitary such that all scattering amplitudes (taken to be real42, 47, 48)
can be folded into a single parameter which we choose to be such that:
1
1
2
2
1 2
(17)
c
1
1
c
;
b
;
a
c
9
It is now possible (Appendix A) to write scattering equations for both junctions taking into
account the spatial and magnetic phases accumulated by the electrons while traveling along
the arms of the ring. Focusing on the scattering process associated with an incident wave
coming on the right lead with amplitude R1, that is, taking the incoming amplitude on the left
lead to vanish, L1=0, these equations can be used to relate all amplitudes in the ring segment
and the leads to the incoming amplitude R1. This leads (see Appendix A) to the transmission
probability in the form
2
(18a)
k
, B
L
2
R
1
1
2
1
2
4
kR
1 cos 2
cos 2
kR
2 sin
kR
2
sin
kR
B
cos 2
0
(18b)
2
2
c
2
1
kR
4 cos 2
2
a
cos 2
kR
2
b
C
2
kR
cos 2
2
a
cos 2
kR
2
b
B
cos 2
0
B
cos 2
0
(18c)
where
S
S
B S B S
B
z
, S
is a vector perpendicular to the ring's surface such that
2
/ e
R
. We note that when the above expression
and, as above, 0
2
reduces to the standard expression for symmetrically connected rings.49
(c) A two State Model
In the weak leads-ring coupling limit, the width of the doubly-degenerate energy
levels on the ring is considerably smaller than the inter-level spacing between states of
different angular momentum. Therefore, at low bias voltages, we can safely assume that
electronic transport takes place mainly through a couple of degenerate levels close to the
Fermi energy of the leads and model the transport physics using the simplified two-level
model shown in Fig. 3. To assign the relevant model parameters, consider an external
zB
threading a molecular ring of radius R that lies in the XY plane.
magnetic field
B
,0,0
The energy levels of an electron moving otherwise freely on the ring are given by (see
Appendix B):
10
2
2
2
1
R
mE
B
m
0
,2,1,0
is the angular
where we use atomic units unless otherwise stated. Here,
m
quantum number (for the isolated ring, the quantum number m relates to the wave vector k
of the previous section via
) and the flux quantum is
. In the basis of the
2
kR
m
0
corresponding eigenstates of the isolated ring, the molecular ring Hamiltonian MH is given
m
)
(
MH , that correspond to states whose
by a repeated sequence of diagonal 2x2 blocks,
(19)
degeneracy is split by the field:
)
2
2
2
2
0
m
H
m
(
M
1
R
B
0
m B
z
2
The full transport problem of a molecular ring connecting between two metal leads can thus
be replaced by the simplified model involving only the two levels characterized by a given
,j KV
K L R
j
m , shown in Fig. 3. Here
,
1, 2 ;
denotes the coupling between molecular
;
m B
z
2
. (20)
B
0
0
E
2
E
1
0
1
R
m
2
0
2
2
2
level j and the lead K, and
E
1/ 2
1
R
2
2
2
m
B
0
2
m B
z
2
V2L
E2
V2R
(21)
V1L
V1R
E1
Figure 3: Schematic representation of the two-level model. The leads are represented in this figure by
tight binding chains.
transmission coefficient
The corresponding
a
r
E G E
E G E
M
M
where the broadening function of lead K (=L,R) is given by
a
r
K
K
the Landauer
is given by
formula
i
(23a)
(22)
Tr
E
E
E
E
K
R
L
11
the retarded Greens function of the molecule is calculated as
1
†
E
a
r
G E
G E
M
M
And the self-energy of lead K is represented as
EI H
E
r
R
r
L
M
(23b)
†
2
K
K
kk
r
K
a
K
E
E
(23c)
r
0
G E
K
V
K
1,
*
V V
K
2,
1,
*
V V
K
K
1,
2,
2
V
2,
0r
KG E being the retarded Green's function of the isolated lead K. Eq. (23c) is written under
the assumption of short range interaction between ring and lead, whereby the ring is coupled
to the nearest neighbor lead site denoted by the index k. We could have used here the explicit
Newns-Anderson model for a 1-dimensional tight binding lead for which is given by Eqs.
(9), however, because we will be using the 2-level model as a generic simple model to gain
physical insight into the nature of our results it is enough to make the simplest wide-band
r
r
0
0
G E
G E
R
L
kk
kk
electronic states and we assume identical leads. On the other hand, the magnetic field
dependent transport properties are determined by the choice of lead-ring coupling elements
V
that enter Eq. (23c). Aiming to capture this dependence, we assume that
lead
ring
where ρ is the density of lead
i
approximation for which
im
Ve
reflects the phase of the wave function on the ring at the positions of the leads-ring junction.
Referring to Fig. 2a and setting the angle at which the right lead is attached to the ring to be
0 , the left lead is attached at the respective angle γ (γ is π, 2π/3 and π/3 for the para,
0
meta and ortho configurations, respectively). Hence we take
im
im
im
Ve
Ve
V V
Ve
V
V V
V
;
;
;
0
0
R
R
L
L
1,
2,
1,
2,
where V is the coupling strength. Consequently, Eq. (23c) yields the retarded self-energies
associated with the right and left leads in the forms
2 1 1
1 1
where m is related to the imposed magnetic field and to the electron energy through Eq.(21).
/L R E
It is now possible to obtain explicit expressions for the broadening matrices
and
EGEG
the retarded and advanced molecular Green's functions
from which the
r
a
,
M
M
transmission probability can be calculated using Eq. (22). A long but straightforward
calculation (see Appendix B for a detailed derivation) leads to
1
im
2
i V
i V
r
R E
(25)
(24)
;
im
E
r
L
1
e
e
2
2
12
zE B
,
1
2
1
2
E E
1
2
2
E E
1
E E
2
cos 2
m
E E
2
2
(26a)
(26b)
2
2
E E
1
2
E E
1
E E
2
2
2
2
E E
1
E E
2
2
cos
m
E E
2
2
sin
m
4
(26c)
2
2 V
where the dependence on
3. Results and discussion
zB originates from Eq.(21).
(26d)
As discussed above, current conduction through ring structures is inherently
associated with interfering transmission pathways50 that may be conveniently described in
terms of degenerate eigenstates of the isolated ring. The corresponding degenerate states can
be represented in terms of rotating, clockwise and counter-clockwise, Bloch states on the
ring. Indeed, it is the tuning of relative phases of these states by magnetic field that
potentially provides magnetic field control of the ring transmission properties. This implies
several important aspects of the resulting behavior: First, transport will be affected by
interference (and consequently most amenable to magnetic field control) in energy regimes
dominated by such degenerate states. Second, strong interference effects and large sensitivity
to magnetic field are expected when these states are associated with sharp transmission
resonances, i.e., for sufficiently weak metal-molecule coupling. Third, the symmetry of a
given junction geometry strongly affects the interference pattern and hence dictates the
transport properties. Fourth, these phenomena will be strongly affected by dephasing
processes. In what follows we will see different manifestations of these statements.
We study single-molecule junctions consisting of molecular ring structures such as
benzene, biphenyl and anthracene connecting metal leads. The results presented below focus
on the response of these systems to an externally applied static uniform magnetic field in
terms of modification in their electronic transport characteristics. The molecular junction is
emulated by the tight-binding (Hückel) molecular Hamiltonian and 1-dimensional tight
binding leads as presented above. For the latter we take zero on-site energies, i.e.,
K K L R
eV6
(
, )
0
and nearest neighbor coupling
that corresponds to a
L
R
metallic band of width 24 eV. The zero bias Fermi energies of these contacts are set to
13
M
eV5.2
eV5.1
and
for all
α M
0FE
. For the molecular structure we take
nearest-neighbor atom pairs. 51
Consider first a simple benzene ring that can couple to the metal leads in para, meta
and ortho configurations (Fig. 1). In the free molecule the highest occupied molecular orbitals
(HOMOs) and the lowest unoccupied molecular orbitals (LUMOs) constitute pairs of doubly
degenerate orbitals which, with our choice of molecular parameters and energy origin, are
eV
eV
4
positioned at
, respectively. Upon connecting to the
and
1
M
M
M
M
metal leads these levels get broadened and, more importantly, their degeneracy split. For
sufficiently weak metal-molecule coupling these split levels constitute sharp transmission
resonances at the corresponding energies.
It should be emphasized that degeneracy splitting in benzene affected by imposing
perturbations at some atomic positions does not by itself specify the nature of the new
eigenstates. An important property of the resonances obtained when the ring is connected to
infinite leads at the meta or ortho positions (i.e. scattering resonances characterized by
scattering boundary conditions, that is, incoming in one lead and outgoing in the other(s)), is
that they can be shown to maintain the character of circulating Bloch eigenstates of the
isolated ring. The corresponding transmission resonances are therefore associated with
considerable circular current in the benzene ring.7 Consequently, in the meta and ortho-
connected configurations, large circular currents are found when bias and gate potentials are
such that one of the split resonances dominates. In contrast, in the para connected ring, one of
the split eigenstates turns out to have a node at one of the para positions and, consequently,
does not contribute to transmission, while the other is characterized by zero net circular
current as could be expected from symmetry.
This is shown in the upper panel of Fig. 4, where the bias voltage is set to V = 2 V.
This brings the upper Fermi energy to the vicinity of the LUMO pair: In the meta and para
connected benzenes one of the split resonances is below and the other above this energy.52
The metal-molecule coupling is taken βLM = βRM = 0.05 eV, low enough so these resonances
remain well separated. For these parameters the net current through the junction is of order ~
nA, while the circular current in both the meta and ortho configurations is three orders of
magnitudes larger, yielding 0.23 Tesla for the induced magnetic field in the ring center in
both cases. Note that the direction of the circular current and the ensuing magnetic field is
opposite in the meta and ortho configurations. See Ref. 7 for more details.
14
Fig. 4. Internal current distribution in (a) para (b) meta and (c) ortho-connected benzene ring,
eV under voltage bias of V 2 V. The upper panel shows the
0.05
KM
connected to leads with
bond currents calculated in the absence of an external magnetic field, while the lower panel
corresponds to the presence of a magnetic field, B = -2T (negative B corresponds to a field pointing
down into the plane). The arrows along bonds represent bond currents with magnitudes proportional
to the corresponding arrows lengths. The encircled dot (cross) in the meta (ortho) structures in the
upper panel denote the directions of the magnetic field induced by the circular current: out of (into)
the molecular plane.
When an external magnetic field is switched on, the bond-current map changes. In
these and the following calculations the external magnetic field is taken in the Z direction,
perpendicular to the molecular XY (also page) plane. Positive field direction is taken to be
outwards, towards the reader and a positive circular current is taken to be in the
counterclockwise direction. This field generates an additional circular, so called persistent,
current that can reinforce or suppress the voltage driven circular current. Thus, a negative
magnetic field (direction into the paper plane) generates a current in the anticlockwise
V
2 V
direction that adds to the circular current in the meta connected ring (at
) and subtract
from it in the ortho-connected structure, as seen in the lower panels of Fig. 4. Of course, the
interplay between the voltage driven and field induced circular current depends on the
voltage range considered. For example, in the meta-connected ring the voltage driven circular
V
2 V
current reverses its direction above
and a negative field induced persistent current
will add to it destructively as in the ortho case. This implies that at any finite bias
I B
I
B
(
)
(
)
c
.
c
15
Fig. 5. Circular current as a function of bias voltage in the range (1.994 to 2.006 Volt) in a
in the presence of external magnetic field, B = 0, +/- 1T
meta-connected benzene for
eV05.0
KM
and +/- 5T. The inset depicts the case for
for applied field B = 0, +/- 100 T.
eV5.0
KM
Fig. 5 shows another aspect of this effect. Here the circular current in a meta-
eV (K = L, R) is shown as a
0.05, 0.5
connected benzene ring connected to leads with
KM
function of voltage for different applied magnetic fields. Again it is seen that
I B
(
c
c
in the presence of bias voltage. It is also seen (inset) that the sensitivity to magnetic field is
strongly reduced when the molecule-lead coupling becomes stronger.
B
I
(
)
)
16
2V V plotted as function of
Fig. 6. The circular current in a meta-connected benzene ring biased at
magnetic field applied perpendicular to the ring. The three cases shown correspond to different
0.05
0.10
eV, dashed line (red):
eV and
molecule-lead coupling. Full line (black):
KM
KM
eV (in the inset). The magnetic field induced by the voltage driven
indB
KM
current is practically the same in all cases,
dotted line (blue):
0.5
0.23
T.
It is of interest to ask, what is the external magnetic field that will annihilate the
circular current in a voltage driven molecule? One may naively expect that this is just the
field equal in magnitude and opposite in direction to that induced by the circular current so
that the two fields annihilate each other, however Fig. 6 shows that the magnetic field needed
to stop the circular current is considerably larger than the magnetic field produced by that
0.05, 0.1, 0.5
current, and generally depends on the molecule-lead coupling. For
eV we
KM
find this field to be 0.98, 3.92 and 96.40 tesla, respectively, at the voltage bias employed (2
V). On the other hand we find that the magnetic field induced by the circular current,
indB
tesla at the same voltage bias, almost independent of the molecule-lead
0.23
coupling.53 This non-trivial behavior results from the fact that the application of the external
magnetic field does not simply oppose the circular-current induced magnetic field but also
alters the electronic structure of the ring and strongly influences the interference pattern of
coherent electrons mounting the ring thus influencing the resulting induced magnetic field
itself.
17
Fig. 7 The I-V Characteristics of a junction comprising para- (upper panel) and meta- (lower panel)
connected benzene coupled to the leads with coupling element 0.05 eV, evaluated for different
magnetic field strengths B normal to the ring. The results obtained for different magnetic fields for the
para system are essentially indistinguishable from each other. The inset in the upper panel shows a
close-up on the V = 2 V neighborhood that shows the consequence of the split degeneracy in the para-
connected junction. The inset in the lower panel shows the I-V behavior in the meta configuration for
molecule-metal coupling 0.5 eV, which is essentially field independent in the same range of magnetic
field strengths. Results for the ortho-connected molecule are qualitatively similar to those shown for
the meta configuration.
Of more practical implications is the question whether the junction transport
properties can be affected by an externally applied magnetic field. As already mentioned,
E
previous studies1, 2 seem to indicate that while the transmission
may be affected by an
external magnetic field, the integrated transmission that yields current-voltage characteristics,
is not sensitive to this field. The main reason for this observation is that at realistic magnetic
18
the
field values the splitting between the degenerate levels of the ring is of the order of meVs
(see Fig. 9). At high leads-ring coupling this splitting is much smaller than the width of the
E
curves. At the low coupling limit,
two levels and is therefore hardly seen even in the
E
curve clearly shows the magnetic field induced level splitting but in order to
observe the magnetic field effect in the I(V) curves, the Fermi integration window (see Eq.
(16)) should include only one of the split levels. This, however, requires bias and gate voltage
precision smaller than the level splitting, as well as very low temperatures.
One may conclude that despite the ability to control the magnetic field sensitivity of
the transmission probability through molecular rings via the leads-ring coupling,5, 6 practical
measurements of the I(V) curves will hardly show any magnetic-field effect. This can be
clearly seen in the upper panel of Fig. 7 where the current-voltage relationship of a
symmetrically (para-) connected benzene ring is found to be robust against the external field.
Only when zooming into the current step region (inset of the upper panel of Fig. 7) one finds
a small shoulder resulting from the level splitting at a finite magnetic field value.
While this conclusion is true for the symmetric junction, in the asymmetrically
connected junction a different behavior is observed. In the lower panel of Fig. 7 we present
the I(V) curves of the meta-connected ring for different magnetic field intensities. Despite the
fact that the level splitting is similar to that of the symmetric junction, the current-voltage
characteristics show pronounced sensitivity towards the magnetic field.
19
around E=1eV through a junction comprising para- (upper
Fig. 8 Transmission probability
E
panel) and meta- (lower panel) connected benzene coupled to the leads with coupling matrix element
0.05 eV, evaluated for different magnetic field strengths B normal to the plane of the ring. Again,
results for the ortho-connected molecule are qualitatively similar to those shown for the meta
configuration.
To get further insight into the origin of this behavior, we show in Fig. 8 the
( E
)
transmission functions
that constitute the input to the I-V results of the preceding
figure. Shown are the transmission functions for para- (upper panel) and meta- (lower panel)
connected benzene rings under different perpendicular magnetic fields in the vicinity of the
eV
doubly degenerate LUMO energy of the isolated molecule,
1
. We see that the
M
M
transmission function depends on the magnetic field in both the symmetric and the
asymmetric junctions. Consider first the para connected junction and denote the split states in
this configurations by
2 . As noted above, only one of these, say
1 and
1 , contributes to
the transmission and the system is characterized by a single transmission resonance. In terms
of the two counter-rotating Bloch states that are eigenfunctions of the isolated ring, this
resonance is a linear superposition in which the corresponding paths add constructively. The
20
the eigenstates become (as B0)
, implying that (a) the single transmission
2 of zero transmission corresponds to the superposition in which they interfere
other state
destructively to give a node in one of the para positions. In the presence of an applied field B
1/ 2
2
2
1
0B splits into two peaks of equal intensities, and (b) the total area under the
peak at
transmission function remains unchanged. This leads to an I-V characteristics that does not
depend on the magnetic field (Fig. 7a) except in the very narrow voltage region where the
Fermi step goes through the split peak (Fig. 7a inset). As discussed above, such a sharp Fermi
step requires very low temperatures (~1 K) to be resolved.
In contrast, in the meta- and ortho- connected junctions, the asymmetric coupling to
the leads results in the appearance of two transmission peaks to appear already in the absence
of external magnetic field. As B is increased, this splitting reduces up to a certain magnetic
field intensity (for instance, +/-2 tesla in the meta-configuration for 0.05 eV molecule-lead
coupling) where it vanishes (level crossing) engendering constructive interference at this field
value. This can be understood as phase adjustment of the interfering electron waves by field,
causing them to interfere constructively until full resonant transmission is reached.
Interestingly, in this regime of magnetic field strength not only the splitting but also the area
under the transmission function is field dependent. As the field intensity increases from zero
the total area under the peaks increases until the peaks become fully separated and then the
area remains constant. This is clearly manifested in the field dependence of the current-
voltage characteristic shown in Fig. 7b, suggesting that in the asymmetric case the I-V field
dependence should be experimentally accessible in the low leads-ring coupling regime. Note
that Fig. 7b shows results obtained at 0K, however the results obtained at 300K are almost
indistinguishable.
A more general view of this behavior is seen in Fig. 9, which shows, for the para- and
junctions,
meta- connected
levels at
the benzene energy
the evolution of
E
eV
1
as a function of molecule-leads coupling (left side of figures) and
M
M
magnetic field (right) as expressed by the transmission function. It is seen that level
degeneracy is lifted by the molecule lead coupling in the meta structure, but not in the para
structure. Increasing the magnetic field for a given (0.05 eV) molecule-lead coupling splits
the levels in the para case, but brings them together first in the meta case, as discussed above.
21
E
plane at B = 0 (left of the vertical
K L R
,
Fig. 9. Transmission probability maps in the
KM
dashed line) and in the E B plane for
0.05 eV
KM
junctions comprising para- (upper panel) and meta-connected (lower panel) benzene molecules. Color
) to red (
).
code varies from deep blue (
1
0
(right of the vertical dashed line) for
Two additional observations should be pointed out. First, another regime of field
dependence takes place at very high fields, where shift of energy levels makes more levels to
appear within the Fermi window between μL and μR. This happened at unrealistically large
fields
( Ec
~ 1000T . Second, in contrast to the circular transmission coefficient
, the
)
BI
BI
( E
(
)
)
)
(
total transmission coefficient
is not affected by the field direction,
.
Qualitatively similar results as described above are obtained for other ring-containing
molecular systems. For example, the isolated biphenyl molecule, which comprises two
coupled benzene rings, possesses two 2-fold degenerate orbitals, viz., HOMO-1 and
eV
and
4
LUMO+1 that for our choice of parameters are positioned at
M
M
, respectively. A biphenyl molecule connected to leads at positions 6,10 (see
M
M
Fig. 10) can be considered as two coupled benzene rings in para configuration, and
consequently we expect that its I-V characteristic is insensitive to an imposed weak magnetic
field. Indeed, a recent work54 finds that to affect transport in this configuration by magnetic
0.5 which, as discussed, is unrealistic for such a small
field requires flux of order
0
molecular structure. On the other hand, the diagonally connected biphenyl shown in Fig. 10a
eV
1
22
can be considered as two coupled benzene rings, each in meta configuration. For such a
structure weakly coupled to leads, we find again high sensitivity of the transmission (Fig.
10b) and the I-V curve (Fig. 10c) to the magnetic field. It should be noted that sensitivity to
magnetic field is manifested when the molecular levels at 1eV (degenerate in the free
molecule) enter the Fermi window at voltage bias 2V. The current rise at
and
V52.0V
3.58V is due to resonant transmission through non-degenerate energy levels at ~ 0.26 eV
(LUMO) and 1.79 eV (LUMO+2), respectively, and is not affected by the field.
Fig. 10. (a) Field free internal current distribution in a diagonally connected biphenyl molecule at a
bias voltage of 2V showing the circular currents in the absence of external magnetic field for metal-
molecule coupling of 0.005 eV. (b) The transmission probability displayed against the electron energy
around 1 eV at different field strengths. (c) Magnetic field effects on the I-V characteristic for B
in
the range 0…15 Tesla.
23
Fig. 11. (a) Field free internal current distribution in a diagonally connected anthracene molecule at a
bias voltage of 2V showing the circular currents in absence of external magnetic field for metal-
molecule coupling of 0.005 eV. (b) The transmission probability displayed against the electron energy
around 1 eV at different magnetic field strengths. (c) Magnetic field effects on the I-V characteristic
for
Tesla.
B
20
Similar results for a junction with diagonally connected anthracene bridge are shown
in Fig. 11. In the voltage range shown, sensitivity to a weak magnetic field is associated with
eV
(LUMO+1) and
1
levels at
the doubly degenerate anthracene
M
M
2.03
eV
(LUMO+2) (responsible for the current rise at V = 2 and 4.06 eV,
2
M
M
respectively). The other current steps seen in Fig. 11c are due to non-degenerate levels and
are not sensitive to the field. As in the previous cases discussed, this behavior depends on the
symmetry of the molecular junction. For example, in agreement with Ref. 55, no sensitivity of
the I-V characteristic to weak fields is found for contacts placed in the (2,6) positions (see
Fig. 11a) although the transmission function itself does depend on the field in a way
reminiscent to the para-connected benzene junction.
Finally, we note that naphthalene does not have orbital degeneracy and indeed no I-V
sensitivity to weak magnetic field is found (although circular currents are induced) in
junction models based on this structure. Conduction through this molecule is found to be
affected only by unrealistic strong fields of order ~ 1000 tesla. These observations can be
summarized by stating that sensitivity of the I-V behavior to relatively weak external
magnetic fields is a generic phenomenon in many ring molecular structures characterized by
24
weak molecule-lead coupling, however its manifestation depends on details of the electronic
structure of the molecule and on the junction geometry.
) for (a)
0.0005
4
.
0/
~ 10
B
in a continuum model (
Fig. 12. Transmission probability around
1kR
2 / 3
(meta) at very small fractions of the flux quantum
(para) and (b)
It is interesting to compare the results shown above for the field affected electronic
transmission through molecular structures to the equivalent transmission problem in the
continuum ring model described in Section 2b. Fig. 12 depicts the transmission probability as
function of the dimensionless parameter kR calculated from Eq. (18), using the reflection
0/B of the order of 10-4. The cases
parameter
at different flux ratios,
0.0005
2 / 3
and
correspond to the para and meta connected rings, respectively. Estimating the
kR correspond to a free electron energy of
benzene ring radius as R~0.13 nm, we find that
1
4
0/
B
of the results displayed in Fig. 12 and those of Fig. 8 is obvious, showing that the complex
corresponds to B of order 10 tesla. The close similarity
the order ~ 2eV, while
10
25
nature of the magnetic field dependent transmission probability through junctions involving
rings of various geometries is intimately related to the symmetry of the wavefunctions
obtained by a simple model of a particle on a ring.
This further justifies the use of the simplified two-level model introduced in Section
2c which relates the generic nature of the magnetic field dependence described above to the
phase dependent coupling coefficients and magnetically controlled wave interferences on the
ring. In Fig. 13, we plot the transmission probability as function of energy obtained by the
two-level model expression (Eq. (26)) for two system geometries under various magnetic
field intensities.
0
0
180
120
Fig. 13. Transmission probability as a function of energy obtained using the two-level model, Eq. (26
), for meta (left panel) and para (right panel) connected rings. We use m=2, ring's radius= 1.0 nm,
V=0.1 eV, =0.05 eV-1 and let the x-axis origin follow the average position of the two levels in the
presence of the magnetic field. The ortho-connected ring results are identical to those obtained for the
meta configuration.
As can be seen, the two-level model fully captures all the features appearing in the
transmission probability curves obtained by both the tight-binding Hamiltonian and the
continuum scattering description. Here, as well, for the symmetrically connected ring the
magnetic field serves to split the energy levels while conserving the total area under the
transmission peaks, whereas for the para and ortho configurations the integrated transmission
probability grows with the magnetic field. This equivalence between the three approaches
(tight binding Hamiltonian, scattering model, and two-level model) sheds light on the origin
of the different behavior of the transmission probability between the three system geometries.
26
As suggested by the two-level model, the differences between the three geometries enter via
the coupling integrals between the ring and the leads taken to be proportional to the phase of
the wave function at the locations of the junctions. Coupling of one of the leads to a nodal
point of the wave function will cause destructive interferences which may be lifted by the
application of the external magnetic field. Since different geometries couple the leads with
different phases the response of the transmission probability towards the magnetic field is
altered.
4. Effect of dephasing
Since much of the effects discussed above result from interference between
transmission pathways, it is expected that dephasing processes will have a strong effect on
these observations. Such effects were studied by several authors in this context using the
Büttiker probe method56 whose application predominates the field of junction transport.
Because this phenomenological method is based on a rather artificial process of replacing
coherently transmitted electrons by electrons with indeterminate phase, we chose to compare
such results with those obtained from another phenomenological model in which the
dynamics of the density matrix of the molecular bridge incorporates damping of non-diagonal
elements as done in the Bloch or Redfield equations describing relaxation in a multilevel
system.
In both the Büttiker probe and the density matrix approaches it is possible to affect
dephasing locally, i.e. at any given site of the tight-binding bridge that represents our
molecule, using the following procedures:
In the Büttiker probe method56, 57 the dephasing rate at such a site, j, is determined by
its coupling to an external thermal electron reservoir, J, with chemical potential set such that
no net current flows through the corresponding contact. We describe the probe by the same
tight-binding metal model, Eq. (2), and the same energetic parameters (site energy and
intersite coupling) as our source and drain leads. The dephasing rate is determined by the
,j J between molecular site j and the site of the probe J that is coupled to it. In the
coupling
calculations reported below we take all these coupling parameters between molecular sites
, for all j and J. Within the model, one calculates the
and probes to be the same,
,j J
MB
K K E
'
,
probes) and obtains the effective transmission function between source and drain in the form
between any two leads (
transmission functions
for N
1, ...
N
L R J
,
,
K K
,
'
27
eff
RL
RL
N
J J
,
' 1
RJ
W
JJ
'
J L
'
, where
W
1
JJ
'
JJ
'
1
JJ
'
1
JJ
JJ
'
(δ is
the Kronecker delta function) and
R
JJ
1
K R L J
,
,
'
J
JK
is the reflection function in
,
m
.
n
(
E
)
(
E
)
(1
(27)
nm
]
nm
n and
where
)
nm
nm
probe J.
In the density matrix description36, 37 one looks, for a given injection current in the
source wire, for the steady state solution of the Liouville equation for the density matrix of
the inner system
i H
1
1
[
nm
2
2
n is the self-energy, Eq. (9), representing the effect of an infinite lead coupled to site
, where nM is 1 if site n is on the molecule and is zero
(1 / 2)
nm
mM
nM
otherwise. Here, the parameter η (taken to be the same for all ring sites) represents the
dephasing rate. The resulting steady state solution is then used to evaluate the current on any
bond segment as well as the transmission coefficient.36, 37
Figures 14 compares results from these calculations in the absence of a magnetic
field. Here dephasing is affected on all sites of the molecular (benzene) ring and the
transmission coefficient is plotted against electron energy for the para, meta and ortho
connected benzene molecules for different values of the dephasing parameters. We note in
passing that our calculations using the Büttiker probe method are practically identical to those
obtained by Dey et al.58 when the same junction parameters are used. On this level of
presentation the main effect of dephasing is seen to be broadening of the transmission peaks.
We note that the two different phenomenological models of dephasing give qualitatively
similar results.
28
Fig. 14. Transmission probability as a function of energy in the presence of dephasing: Top, middle
and bottom figures correspond to para-, meta- and ortho-connected benzene molecules. Left: Results
obtained using the Büttiker probe model with the indicated coupling parameter BM . Right: results
obtained by the density-matrix calculation with the indicated dephasing rate η. The molecule-leads
coupling is 1 eV. The other model parameters are as given in the second paragraph of Section 3
(parameters of the probe leads are the same as for the source and drain leads).
Figures 15 (using the Büttiker probe method) and 16 (density matrix model) focus on
1E eV. The broadening effect caused by dephasing can be
the transmission resonance near
clearly seen both in the strong (upper panels) and weak (middle panels) leads-ring coupling
regimes. In addition the effect of dephasing on eliminating interference characteristics is
apparent. This is most pronouncedly manifested in the integrated transmission (lower panels).
For low bias and temperature, broadening alone makes the integral smaller at any finite E
29
because part of the integrand exits the narrow integration window. Furthermore, in the para
connected molecule, the integrated transmission goes down also because of the destruction of
constructive interference. Conversely, in the meta configuration it goes up (Fig. 15) or
considerably more weakly down (Fig. 16) because the destructive interference is eliminated.
Fig. 15. Dependence of the transmission resonance at 1 eV on dephasing calculated with the Büttiker
probe method. The line style and color representing different values of BM and given in the framed
inset in panel 1a correspond to all panels. Panels 1a,b,c show results for para-connected benzene
while panels 2a,b,c correspond to the meta-connected molecule. In the a and b panels the molecule-
K L R
1
,
KM eV and 0.05 eV, respectively (
source/drain couplings are
). In the weak molecule-
0
BM (no dephasing) lines are scaled down by a
source/drain coupling case (panels 1b, 2b) the
factor of 100 (i.e. multiplied by 1/100) in the para case and by a factor of 2 in the meta case, in order
E
a
to fit on the given scale. The c panels show the integral about the resonance,
for the
dE
E
30
weak coupling (0.05 eV; as in panel s b) case plotted against E, where
the 1eV resonance but well above the lower transmission resonance.
a
e
1 V
is placed well below
Fig. 16. Same as Fig. 15, where the effect of dephasing is obtained from the density matrix approach.
The line style and color representing different values of and given in the framed inset in panel 1a
0 lines are scaled down by the multiplicative factor
correspond to all panels. In the b panels, the
2 10
4 10
2
4
in the para case, and
of
in the meta connected molecule in order to fit into the
4 10
4
, i.e. the peak transmission in
scale used here. Note that the vertical scale itself goes up to
the para-connected molecule without dephasing is 1.
In Figures 7 and 8 above we have demonstrated the possibility of magnetic field
control of the I-V characteristic in the case of meta (and ortho) connected benzene. Figures
17-18 show the effect of dephasing on this dependence. Figures 17a is the analog of Fig. 7b,
31
) meta-connected junction with dephasing implemented by the Büttiker probe
Fig. 17. (a) Magnetic field dependence of the current near the 2V step (associated with the
E
1eV
; analog of Fig. 7b) calculated for a weakly coupled
transmission resonance at
0.05 eV
KM
0.1eV
BM
method (
I
I B
0
0
plotted respectively against the dephasing parameters BM ( Büttiker probe
method) and (density matrix method). The inset in 17c displays the results of the main figure on a
different scale, emphasizing the observation (also seen in 17a) of deviations from B symmetry at
intermediate dephasing rates.
). (b) and (c) The same magnetic field dependence expressed by the ratio
(
I
32
where the I-V characteristic for the meta-connected benzene under different magnetic fields
normal to the ring plane is shown, now in the presence dephasing (Büttiker probe method,
0.1eV
). Similar results are obtained when dephasing is introduced in the density
BM
matrix method. We see that the field dependence of the current step strongly diminishes in
the presence of dephasing. This is shown more explicitly in Figures 17b,c , where the
difference between the current evaluated at V = 2.2 V for
0B and
B
is plotted
15 tesla
against the dephasing parameter.
More insight about the origin of this behavior is obtained from Fig. 18, which is the
analog of Fig. 8 calculated in the presence of dephasing. The primary effect of dephasing in
the range considered is seen to be the destruction of interference effects, which strongly
diminishes the difference between the transmission properties of the para and meta connected
junctions as well as the dependence on an imposed magnetic field.
Fig. 18. Transmission probability for para (1a, b) and meta (2a, b) weakly connected
0.05 eV
KM
(
) benzene as a function of electron energy in the presence of dephasing affected
through Büttiker probes (panels a) (
eV
001.0
BM
) at B = 0, -5 and -15 Tesla.
(
eV1.0
) and density matrix method (panels b)
33
The significance of the dynamic destruction of phase in the quick erasure of the
magnetic field effect on the conduction properties of asymmetrically connected benzene
molecules can be gauged against other effects that can potentially affect this sensitivity. First,
our calculations have disregarded the implications of the voltage distribution across the
molecular junction. On the simplest level of description this can appear in the voltage
division between the two contact, expressed by a factor ξ and a voltage V such that
and
V
V
V
V
are the potential drops on the two molecule-lead contacts
1
while V is the potential drop on the molecule itself. All the calculations described above
0V . We have established that our results are not sensitive
and
where done with
0.5
to the choice of . An example of the effect of V is shown in Fig. 19, which extends the
calculations displayed in Fig. 8b (meta-connected benzene) to the case shown in panel (a)
where some of the site energies are changed,
. We see that the magnetic field
effect on the transmission decreases with increasing and practically disappears for
eV. This emphasizes the need to carry such experiments under low bias
0.05
conditions, implying that need to align the molecular spectrum with respect to the lead Fermi
energy with a gate potential.
34
Fig. 19. The transmission function
for the meta-connected Benzene, same as Fig. 8b,
E
with some site energies shifted as shown in panel (a). In panels (b), (c), (d) is taken 0.01,
0.02 and 0.05 eV, respectively.
Secondly, keeping in mind that the required slow dephasing implies low temperature,
it is important to note that the effect seen in Figs. 17, 18 arises from the dynamic destruction
of phase, and is not reproduced merely by raising the electronic temperature of the leads. This
is seen in Fig. 20, which shows the effect of leads electronic temperature on the observed
magnetic field dependence of the conduction through a meta-connected benzene junction
near the molecular resonance at 1 eV (bias voltage 2 V) in the absence of dephasing. Fig. 20a
is similar to Fig. 7b except that the Fermi distributions in the electrodes were taken at 300K.
Fig. 20b displays the current at V = 2.2 V through this junction, plotted against the electrodes
temperature for B = 0 and 15 tesla. Only weak electrode temperature effect is seen at the
realistic temperature range considered.
Fig. 20. (a) Current vs. bias voltage for a weakly (
0.05 eV
KM
different imposed magnetic fields perpendicular to the molecular plane (same as Fig. 7b) calculated at
T = 300K. (b) The current through the same junction at V = 2.2 V, displayed as a function of
temperature for B = 0 and 15 T.
) meta-connected benzene for
Finally, It is interesting to note that in the presence of dephasing, deviations from
Onsager symmetry under reversal of field direction, B
B , are observed (as seen in Fig.
17). Such deviations were discussed in previous work in different contexts, including
coupling to a thermal environment.59-62 We defer further discussion of this issue to a later
publication.
35
5. Concluding Remarks
In this paper we have addressed the issue of magnetic field effect on electronic
transport in molecular conduction junctions. Observations of the Aharonov-Bohm effect in
mesoscopic conducting loops could suggest that molecular junctions comprising molecular
ring structures as bridges will show similar effects, however the small radius of molecular
rings implies that the field needed to observe the AB periodicity is unrealistically large. Still,
we have found that strong magnetic field effects can be seen under the following conditions:
(a) The molecular resonance associated with its conduction behavior is at least doubly
degenerate, as is often the case in molecular ring structures; (b) the molecule - lead coupling
is weak, implying relatively distinct conduction resonances, (c) asymmetric junction structure
and (d) small dephasing (implying low temperature) so as maintain coherence between
multiple conduction pathways. Interestingly, in weakly connected symmetric junctions (e.g.
para connected benzene) the transmission coefficient can show sensitivity to magnetic field,
however it is found that the integrated transmission is field independent, so that this
sensitivity is not reflected in the current-voltage characteristic.
For the organic structures we have used the within the tight-binding (Hückel) model
modified for the presence of magnetic field using the London approximation. Qualitatively
similar results were obtained from the analog model of a continuous ring, showing that the
qualitative effect studied depends mostly on the strength and symmetry of the molecule-lead
coupling. We have also shown that much of the qualitative behavior of conduction in these
models can be rationalized in terms of a much simpler junction model based on a two-state
molecular bridge.
When the conditions outlined above are satisfied, strong dependence of conduction on
the imposed magnetic field can be found. The effect of dephasing processes on this
observation are studied using two different phenomenological models: the Büttiker probe and
phenomenological coherence damping imposed on the Liouville equation for the molecular
density matrix. Both treatments are approximate: The approximate nature of the density
matrix approach stems from the fact that dephasing was affected by damping non-diagonal
elements of the density matrix in the local site representation while assuming that the
transmission energy remained well defined.36, 37 The Büttiker probe approach is limited to
linear response and cannot be rigorously applied to threshold phenomena in the current-
36
voltage dependence. Still, the fact that these two different approaches gave qualitatively
similar results in all cases studied, provide some assurance about their validity. Both models
show strong suppression of the sensitivity of conduction to the imposed magnetic field.
Next, consider the implications of the conditions outlined above to experimental
considerations. Conditions (a) and (c) can be met by making a proper choice of molecular
bridge and the positions of linker groups. Condition (b) of weak molecule-lead coupling does
not imply weak molecule-lead bonding, only that the resonance states that involve multiple
pathways through the ring (or counter propagating wavefunctions in the ring) are weakly
coupled to the metal electrodes. This can be achieved by connecting molecular ring to leads
via saturated alkane chains.
Condition (d), the requirement for small dephasing, is inherent in all experiments
trying to observe interference phenomena in molecular junctions, and implies the need to
work at relatively low temperatures. The Büttiker-probe procedure is not related directly to a
physical process, so it is hard to assess the experimental implication of the coupling MNV . The
equivalent analysis in terms of the coherence damping rate η does provides an estimate: For
the reasonable molecular parameters chosen in our calculations, Fig. 17c shows that magnetic
field effects are suppressed when this damping rate exceeds ~ 0.001 eV, that is, dephasing
times of the order of 1 ps. Recent observations in different systems63 have shown that
molecular electronic coherence can persist on such timescales even at room temperatures.
This suggests that the magnetic field effects discussed in this paper may be observables.
Finally, we have observed magnetic asymmetry (under reversal of field direction) in
the presence of dephasing. This observation and its repercussions will be discussed
elsewhere.
Acknowledgment
We thank Joe Imry for helpful discussions. The research of A.N. is supported by the Israel
Science Foundation, the Israel-US Binational Science Foundation, the European Science
Council (FP7 /ERC grant no. 226628) and the Israel – Niedersachsen Research Fund. O.H.
acknowledges the support of the Israel Science Foundation under Grant No. 1313/08, the
support of the Center for Nanoscience and Nanotechnology at Tel-Aviv University, and the
Lise Meitner-Minerva Center for Computational Quantum Chemistry. D. R. Acknowledges a
Fellowship received from the Tel Aviv University nanotechnology Center.
37
Appendix A: Scattering model for asymmetric nanoscale junctions containing
molecular rings and magnetic fields.
We consider the setup presented in Fig. 2a, and write the electron wavefunction in each
ikl
ikl
A e
where l denotes a propagation distance, i.e. l R
A e
in
segment of the ring as
1
2
ring segments; is the angle traversed by the electron, and R the radius of the ring. We use
the standard notation where positive k represents counter-clockwise propagating waves and
negative values represent clockwise moving waves. In order to calculate the transmission
probability and the circular current as a function of magnetic field we assign scattering
amplitudes as shown in Fig. 2b and explained in the main text. For simplicity, we assume in
what follows that all scattering amplitudes are real. This assumption implies that no rigid
phase shifts occur upon scattering at the junctions consistent with the tight-binding model
which conserves the continuity of the wave-functions across the junctions.42, 47, 48 Focusing on
junction I (Fig. 2a) we may write the following scattering equation relating the incoming
amplitudes to the outgoing amplitudes:
c
R
2
D
2
U
1
a
b
a
b
Current conservation implies that the scattering amplitude matrix must be unitary, i.e.
2
R
1
D
1
U
2
(A.1)
I
(A.2)
c
a
b
a
b
where I is a
33 unit matrix. This provides three independent equations for the four
scattering amplitudes, leading to
1
1
2
2
Having characterized the contacts we turn back to the circular setup in Fig. 2a. For a given
wavenumber
k we can write a scattering equation similar to Eq. (A.1) for contacts I and II,
taking into account the spatial phase accumulated by the electrons while traveling along the
arms of the ring:
1 2 ;
(A.3)
1
1
a
b
;
c
c
c
38
R
2
D
2
U
1
c
a
b
a
b
R
1
i k R
2
i k R
D e
1
U e
2
(A.4)
2
L
1
i k R
A
i k R
1
2
2
(A.5)
c
L
2
D
1
U
r B
D e
2
U e
1
a
b
a
b
To model the influence of an external magnetic field threading the ring we assume that the
zB
such
magnetic field is homogeneous and perpendicular to the plane of the ring
B
,0,0
that the vector potential may be written as:
x
y
z
z
x
y
B
0 0
z
On the circle defining the ring this becomes
1
A
B R
2 z
The magnetic phase accumulated by an electron traveling along the ring is thus given by:
2
2
A dl
1
1
(A.8)
(A.7)
(A.6)
1
2
y x
,
, 0
d
B R
z
sin
yB
z
,
xB
z
sin
, cos
, 0
, 0
, 0
, cos
sin
, 0
B
z
, cos
m
1
2
2
e
1
2
e
1
2
1
2
B
0
e
. In the presence of such magnetic field, Eqs.
2
,
S
R
, and
zB S
where B
2
0
(A.4) and (A.5) are modified as follows:
R
1
B
0
i k R
D e
1
R
2
D
2
U
1
c
a
b
a
b
i k R
B
0
U e
2
2
L
2
D
1
U
2
c
a
b
a
b
L
1
B
0
i k R
2
D e
2
i k R
B
0
U e
1
(A.9)
(A.10)
39
Focusing on a scattering process with incoming electron coming from the right, we set L1 = 0,
that is, take zero incoming amplitude on the left lead. Eqs. (A.9)-(A.10) then lead to
be
0
2
1
2
2
be
ae
2
B
0
B
0
B
0
B
0
i k R
i k R
i k R
i k R
i k R
0
B
0
Inverting (A.11) yields the wavefunctions amplitudes D and U on the ring segments as a
function of the incoming amplitude R1. In particular, the results for D2 and U1 can be used in
(A.10) to yield L2 and therefore the transmission probability
0
R
1
R
1
0
D
1
D
2
U
1
U
1
B
0
(A.11)
i k R
i k R
i k R
B
0
B
0
ae
be
2
1
0
1
ae
ae
be
0
k
,
B
2
L
2
R
1
2
R
1
D e
2
i k R
B
0
2
i k R
B
0
2
U e
1
(A.12)
which finally results in Eq. (18)
Appendix B: The two-level transport model
Here we construct a two-level transport model that, for weak molecule-lead coupling,
captures the main features observed in the magnetic field dependence of electron
transmission through a molecular ring. The validity of a two-level model stems for the fact
that in this coupling limit only pairs of molecular levels, degenerate in the limit of zero
coupling, are coupled through their mutual interaction with the leads. The Hamiltonian for
system
two-level
the Hamiltonian,
considering
by
obtained
be
can
this
2
for a charged particle moving in a magnetic field. The
P qA
V r
Hamiltonian describing a free (
e ; atomic units are used
1
0V ) electron (
1 / 2
H
q
throughout) moving on a circular ring of radius R lying in the XY plane under a uniform
B
B
magnetic field oriented in the Z direction,
can be written by setting
0, 0,
z
r B
A
(B.1)
y x
,
, 0
1
2
zB
2
This leads to
40
P
A
2
A P P A
2
B
z
8
2
2
2
i
x
y
y
x
B
z
2
H
H
1
2
(B.2)
1
2
2
y
x
where the central term in the last stage may be identified as the angular momentum
component along the Z direction. In circular coordinates with the origin placed at the center
of the ring Eq. (B.2) becomes
2
2
2
R B
B
1
z
z
2
2
8
R
2
in which the last term is an additive constant. The eigenstates of this Hamiltonian can be
written in the form
1
2
with the corresponding energy eigenvalues
2
m
0, 1, 2,
(B.3)
(B.4)
ime
;
i
2
2
mE
B
0
1
R
2
B S
m
2
Here, B
and 0
respectively. Using relation (B.4) the wave functions can be rewritten as
are the magnetic flux threading the ring and the flux quantum,
(B.5)
i
B
0
2
E R
m
e
cw
1
2
Noting that Eq. (B.5) may be written in the form
2
;
ccw
i
e
1
2
2
E R
m
B
0
,
(B.6)
2
2
m
2
,
m
E
1
R
B
0
m B
z
2
the Hamiltonian of the isolated ring in the subspace of these two levels is given by Eq. (20).
When coupled to leads as in Fig. 3, the self-energy terms appearing in Eq. (23c) are given by
12
11
r
0
G
L R
/
r
0
G
L R
/
*
V
L R
1,
/
*
V
2,
0 0
0 0
V
L R
2,
/
0
0
V
L R
1,
/
0
0
r
0
G
L R
/
r
0
G
L R
/
(B.7)
r
L R
/
L R
/
E
E
E
E
E
21
22
*
V
V
L R
2,
1,
/
2
V
1,
L R
/
r
0
G
L R
/
E
11
*
V
2,
V
L R
/
1,
L R
/
V
2,
L R
/
L R
/
2
41
(B.8)
Here,
is the retarded (r) Green's function of the isolated (0) left/right (L/R) lead and
E
G r
0
RL
/
we assume short range interaction between the ring and the leads, whereby the ring is coupled
to
the nearest neighbor
leads
sites. The advanced
self-energy
is given by
†
r
RL
/
The coupling matrix elements
that appear in expression (B.8)
K L R
,
1, 2 ;
a
RL
/
E
E
;
.
j
,j KV
for the self-energy are formally calculated via
V and should therefore be
ring
lead
proportional to the phase of the wave function on the ring. In Eq. (24) we take this phase
dependence into account where the specific symmetry of the system (ortho, meta, or para) is
taken explicitly into account via the angular separation between the leads. Using the coupling
matrix elements given in Eq. (24) we obtain explicit matrix representations for the retarded
(and advanced) self-energies in the forms
r
R E
i V
2 1 1
1 1
;
r
L
E
i V
2
1
im
2
e
2
im
e
1
(B.9)
And
a
E
r
E
†
, i.e.
i V
a
R E
2 1 1
1 1
The broadening matrices Γ, Eq. (23a), and the ring Green's functions, Eq. (23b) are then
obtained in the forms
1
im
2
(B.10)
i V
;
E
a
L
1
e
e
2
2
im
R E
V
2
2 1 1
1 1
;
L
E
V
2
2
1
im
2
e
2
im
e
1
(B.11)
r
G E
M
EI H
m
M
r
L
E
r
R
E
1
2
E E
i V
2
1
im
2
i V e
m
2
cos
where we have used Eq. (20) for the Hamiltonian of the isolated ring. Inverting the matrix in
a
MG E leads to
(B.12) and in the corresponding expression for
m
cos
2
i V
2
2
i V e
2
E E
2
im
1
(B.12)
42
1
i V
2
2
E E
2
i V
2
2
i V
2
im
cos
m
2
im
2
2
V
4
m
cos
2
i V
2
4
2
cos
m
(B.13)
2
i V e
2
E E
1
r
G
M
E
E E
1
And the retarded counterpart is
†
E E
2
2
i V e
2
a
G
M
E
G
r
M
E
E E
1
i V
2
E E
2
2
i V
2
im
cos
E E
2
2
i V e
2
2
im
2
2
V
4
m
cos
2
i V
2
4
2
cos
m
1
i V
2
2
m
2
i V e
2
E E
1
(B.14)
With
the broadening and Green's functions matrix
these explicit expression for
representations, evaluating the transmission coefficient (22) becomes a lengthy but
straightforward calculation leading to the final result (26).
43
References
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
K. Walczak, Cent. Eur. J. Chem 2, 524 (2004).
S. K. Maiti, Chemical Physics 331, 254 (2007).
O. Hod, R. Baer, and E. Rabani, The Journal of Physical Chemistry B 108, 14807
(2004).
O. Hod, R. Baer, and E. Rabani, Journal of the American Chemical Society 127, 1648
(2005).
O. Hod, E. Rabani, and R. Baer, Accounts of Chemical Research 39, 109 (2006).
O. Hod, R. Baer, and E. Rabani, Journal of Physics: Condensed Matter 20, 383201
(2008), and references therein.
D. Rai, O. Hod, and A. Nitzan, Journal of Physical Chemistry C 114, 20583 (2010),
and references therein.
S. Nakanishi and M. Tsukada, Jpn. J. Appl. Phys. 37 (1998) pp. , L1400 (1998).
S. Nakanishi and M. Tsukada, Physical Review Letters 87, 126801 (2001).
M. Ernzerhof, H. Bahmann, F. Goyer, M. Zhuang, and P. Rocheleau, J. Chem. Theory
Comput. 2, 1291 (2006).
N. Tsuji, S. Takajo, and H. Aoki, Phys. Rev. B 75, 153406 (2007).
K. Tagami, M. Tsukada, W. Yasuo, T. Iwasaki, and H. Nishide, Journal of Chemical
Physics 119, 7491 (2003).
B. Wang, R. Chu, J. Wang, and H. Guo, Physical Review B 80, 235430 (2009).
K. Tagami and M. Tsukada, Current Applied Physics 3, 439 (2003).
A. M. Jayannavar and P. Singha Deo, Physical Review B 51, 10175 (1995).
K. Tagami and M. Tsukada, e-Journal of Surface Science and Nanotechnology 2, 205
(2004).
J.-L. Xiao, Y.-L. Huang, and C.-L. Li, Journal of Luminescence 119-120, 513 (2006).
Y.-l. Huang and J.-l. Xiao, Superlattices and Microstructures 41, 17 (2007).
L. G. Wang, Physica B: Condensed Matter 404, 143 (2009).
Note that splitting take place already due to the asymmetric coupling. However,
without magnetic field or externally induced current, the split states do not carry net
angular momenta.
O. Entin-Wohlman, Y. Imry, and A. Aharony, Physical Review Letters 91, 046802
(2003).
O. Entin-Wohlman, Y. Imry, and A. Aharony, Physical Review B 70, 075301 (2004).
I. Barth, J. Manz, Y. Shigeta, and K. Yagi, J. Am. Chem. Soc. 128, 7043 (2006).
I. Barth and J. Manz, Angewandte Chemie-International Edition 45, 2962 (2006).
K. Nobusada and K. Yabana, Phys. Rev. A 75, 032518 (2007).
G. F. Quinteiro and J. Berakdar, Opt. Express 17, 20465 (2009).
A. Matos-Abiague and J. Berakdar, Phys. Rev. Let. 94, 166801 (2005).
44
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
S. S. Gylfadottir, M. Nita, V. Gudmundsson, and A. Manolescu, Physica E: Low-
dimensional Systems and Nanostructures 27, 278 (2005).
A. S. Moskalenko and J. Berakdar, Physical Review B 80, 193407 (2009).
A. S. Moskalenko and et al., EPL (Europhysics Letters) 78, 57001 (2007).
Y. V. Pershin and C. Piermarocchi, Physical Review B 72, 245331 (2005).
E. Räsänen, A. Castro, J. Werschnik, A. Rubio, and E. K. U. Gross, Physical Review
Letters 98, 157404 (2007).
E. Rasanen, A. Castro, J. Werschnik, A. Rubio, and E. K. U. Gross, Physical Review
B (Condensed Matter and Materials Physics) 77, 085324 (2008).
D. Rai, O. Hod, and A. Nitzan, The Journal of Physical Chemistry Letters, 2118
(2011).
O. Hod, R. Baer, and E. Rabani, Phys. Rev. Lett. 97, 266803 (2006).
V. Ben-Moshe, A. Nitzan, S. S. Skourtis, and D. Beratan, J. Phys. Chem. C 114 8005
(2010).
V. Ben-Moshe, D. Rai, A. Nitzan, and S. S. Skourtis, J. Chem. Phys. 133, 054105
(2010).
F. London, J. Phys. Radium 8, 397 (1937).
V. A. Geyler, V. V. Demidov, and V. A. Margulis, Technical Physics 48, 661 (2003).
S. Datta, Electric transport in Mesoscopic Systems (Cambridge University Press,
Cambridge, 1995).
Y. Imry, Introduction to Mesoscopic Physics (Oxford University Press, Oxford,
2008).
M. Buttiker, Y. Imry, and M. Y. Azbel, Physical Review A 30, 1982 (1984).
Y. Gefen, Y. Imry, and M. Y. Azbel, Physical Review Letters 52, 129 (1984).
M. Cahay, S. Bandyopadhyay, and H. L. Grubin, Physical Review B 39, 12989
(1989).
C. Benjamin, S. Bandopadhyay, and A. M. Jayannavar, Solid State Communications
124, 331 (2002).
A. M. Jayannavar and C. Benjamin, Pramana-J. Phys. 59, 385 (2002).
B. Kubala and J. König, Physical Review B 67, 205303 (2003).
I. A. Ryzhkin, Phys. Solid State 41, 1901 (1999).
A. Aharony, O. Entin-Wohlman, B. I. Halperin, and Y. Imry, Physical Review B 66,
115311 (2002).
T. Hansen, G. C. Solomon, D. Q. Andrews, and M. A. Ratner, J Chem Phys 131,
194704 (2009).
the actual parameters vary slightly for the different molecules used in our
calculations. We have checked that these variations do not affect our qualitative
observations.
In the para case, one of the levels actually remains at that energy.
45
This statement holds for the voltage used (2 eV) for which the circular current goes
through a maximum. The dependence of the circular current on the molecule-leads
coupling is more pronounced at other voltages.
S. K. Maiti, Journal of Computational and Theoretical Nanoscience 6, 1561 (2009).
S. Nakanishi, R. Tamura, and M. Tsukada, Jpn. J. Appl. Phys 37, 3805 (1998).
M. Buttiker, Phys. Rev. B 33, 3020 (1986).
J. L. D’Amato and H. M. Pastawski, Phys. Rev. B 41, 7411 (1990).
M. Dey, S. K. Maiti, and S. N. Karmakar, Organic Electronics 12, 1017 (2011).
B. Szafran, M. R. Poniedziałek, and F. M. Peeters, Europhys. Let. 87, 47002 (2009).
D. Sánchez and K. Kang, Physical Review Letters 100, 036806 (2008).
C. A. Marlow, R. P. Taylor, M. Fairbanks, I. Shorubalko, and H. Linke, Physical
Review Letters 96, 116801 (2006).
G. L. J. A. Rikken and P. Wyder, Physical Review Letters 94, 016601 (2005).
A. Ishizaki, T. R. Calhoun, G. S. Schlau-Cohen, and G. R. Fleming, Physical
Chemistry Chemical Physics 12, 7319 (2010).
53
54
55
56
57
58
59
60
61
62
63
46
|
1601.03861 | 1 | 1601 | 2016-01-15T10:13:31 | Decoherence and Decay of Two-level Systems due to Non-equilibrium Quasiparticles | [
"cond-mat.mes-hall",
"cond-mat.supr-con"
] | It is frequently observed that even at very low temperatures the number of quasiparticles in superconducting materials is higher than predicted by standard BCS-theory. These quasiparticles can interact with two-level systems, such as superconducting qubits or two-level systems (TLS) in the amorphous oxide layer of a Josephson junction. This interaction leads to decay and decoherence of the TLS, with specific results, such as the time dependence, depending on the distribution of quasiparticles and the form of the interaction. We study the resulting decay laws for different experimentally relevant protocols. | cond-mat.mes-hall | cond-mat | Decoherence and Decay of Two-level Systems due to Non-equilibrium Quasiparticles
1Institut fur Theoretische Festkorperphysik, Karlsruhe Institute of Technology, D-76128 Karlsruhe, Germany
Sebastian Zanker,1 Michael Marthaler,1 and Gerd Schon1
It is frequently observed that even at very low temperatures the number of quasiparticles in
superconducting materials is higher than predicted by standard BCS-theory. These quasiparticles
can interact with two-level systems, such as superconducting qubits or two-level systems (TLS) in
the amorphous oxide layer of a Josephson junction. This interaction leads to decay and decoherence
of the TLS, with specific results, such as the time dependence, depending on the distribution of
quasiparticles and the form of the interaction. We study the resulting decay laws for different
experimentally relevant protocols.
INTRODUCTION
Superconducting quantum devices have a wide range
of applications. Due to the weak dissipation in the su-
perconducting state they are promising candidates for
building large scale quantum information systems [1, 2].
They are easily controlled and measured by electromag-
netic fields, but for the same reason they couple rather
strongly to the environment and are prone to decoher-
ence [3]. Much effort has been put into understanding
and minimizing various noise sources.
One ubiquitous source of decoherence arises from two
level systems (TLS), such as bistable defects residing in
dielectric substrates, disordered interfaces, surface ox-
ides, or inside the barriers of Josephson junctions [4].
While originally introduced to explain anomalous prop-
erties of glasses at low temperatures [5] there is evidence
that they are also an important source of decoherence
for superconducting qubits [1] or superconducting res-
onators [6]. A bath of TLS can explain the 1/f noise
which limits the performance of many devices [7], while
fluctuations of very slow TLS induce long-time parame-
ter shifts. Still, the microscopic origin of those TLS re-
mains unclear. Some potential sources are small groups
of atoms that tunnel between two stable positions, or
dangling bonds, or hydrogen defects. A better experi-
mental as well as theoretical understanding of those TLS
has been the focus of much recent work [8–12].
Recent experiments demonstrated the coherent control
of TLS, residing inside the amorphous layer of a phase
qubit’s Josephson junction, with help of this qubit [13]. It
was possible to carry out typical coherence experiments
as used in magnetic resonance or other qubit experiments
[14]. The aim of those experiments is a better under-
standing of the microscopic nature of individual TLS as
well as their respective environment responsible for TLS
state fluctuations, with the ultimate goal to reduce their
detrimental effects.
In this paper we analyze the decoherence of charged
TLS residing inside the amorphous layer of a Josephson
junction [4, 15] due to scattering and tunneling of non-
equilibrium quasiparticles in the superconducting leads.
Experiments suggest that even at low temperatures the
number of quasiparticles is large as compared to the pre-
dictions of equilibrium BCS theory [16]. Similar as ob-
served for superconducting qubits [17] or the dynamics
of Andreev bound states [18] the scattering with quasi-
particles provides an intrinsic noise source also for the
TLS. We investigate the TLS decoherence properties due
to the coupling between the TLS and the quasiparticles.
We find characteristic differences for quasiparticles which
scatter back to the same superconducting electrode and
those which tunnel across the junction. The effect of the
latter depends on the phase difference. Our results ap-
ply both for the TLS, which are the main focus of the
present paper, as well as for qubit decoherence [19]. In
fact for the latter the effect of tunneling electrons is more
pronounced.
THE MODEL
We consider the scattering of quasiparticles from a two-
level system located inside the amorphous barrier of an
aluminum oxide Josephson junction. The model Hamil-
tonian reads as
H = HTLS + Hqp + HC,
(1)
where HTLS is the TLS Hamiltonian, Hqp are the free
quasiparticle Hamiltonians of the left and right lead and
HC is the coupling between both subsystems. Since
the microscopic nature of TLS remains unclear, we use
the phenomenological TLS standard model for HTLS of
Ref. [20].
It describes the TLS as an effective charged
particle trapped in a double well potential with asymme-
try between the two potential minima and tunneling
amplitude ∆0,
2
∆0
2
σz +
σx.
HTLS =
2 ETLS σz with ETLS = (cid:112)∆2
In the TLS eigenbasis the Hamiltonian reduces to
HTLS = 1
0 + 2. Electrons
in nearby leads couple to the TLS’ electric dipole mo-
ment and induce an interaction that is well established
in the context oft metallic glasses [21]:
(2)
HC = σz V = σz
†
gkk(cid:48)c
kck(cid:48) + h.c.
.
(3)
(cid:88)
(cid:16)
kk(cid:48)
(cid:17)
6
1
0
2
n
a
J
5
1
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
1
v
1
6
8
3
0
.
1
0
6
1
:
v
i
X
r
a
Here, gkk(cid:48) is the coupling strength between electrons and
the TLS dipole and ck is an electron annihilation opera-
tor with multi-index k = {(cid:126)k, σ, α} that includes electron
momentum (cid:126)k, spin σ, and the index α = l, r of the lead
(left or right), where the electron resides.
In general
we can distinguish two processes: Electrons that tunnel
through the junction (α (cid:54)= α(cid:48)) while interacting with the
TLS and electrons, that scatter back into their original
lead (α = α(cid:48)). Because of the exponentially decaying
electron wave function inside the junction and the local-
ized character of the TLS wave function, the interaction
rapidly decreases for TLS away from the junction edges.
We expect that for most TLS scattering electrons are the
main source of TLS decoherence, while the influence of
tunneling electrons is insignificant. This is quite differ-
ent for the decoherence of a qubit, where only tunneling
electrons couple to the qubit and induce decoherence.
Furthermore, electrons that contribute to TLS decoher-
ence have energies close to the Fermi energy with mo-
mentum (cid:126)k = kF . Hence, we take the direction average
and introduce the direction-averaged coupling constant
αα(cid:48) ≡ (cid:104)g2
kk(cid:48)(cid:105). Since (cid:126)k ≈ kF the averaged coupling con-
g2
stant does not depend on energy.
The free particles in the leads are Bogoliubov quasi-
particles with mixed electron- and hole-like nature and
†
creation operators a
k and Hamiltonian
Hqp =
Their energy is Ek =(cid:112)ξ2
bcs with ξk being the elec-
tron energy in the normal state. Rewriting the coupling
Eq. (3) in terms of quasiparticle operators we find
k + ∆2
V =
gαα(cid:48)
eiϕ/2ukuk(cid:48) − e−iϕ/2vkvk(cid:48)
†
a
kak(cid:48) + h.c.
(cid:16)
(cid:88)
kk(cid:48)
(5)
with coherence factors u2
2 (1 + ξk/Ek). For
tunneling quasiparticles ϕ is the superconducting phase
difference across the junction. The phase difference ϕ
vanishes for scattering quasiparticles.
k = 1 − v2
k = 1
An important quantity in the context of decoherence
is the noise spectral density
SV (ω) =
(cid:104) V (t) V (0)(cid:105)eiωt
(6)
where the average is over the quasiparticle states. With
Eq. (5) we find
SV (ω) =4N 2
0 g2
dEdE(cid:48) ρ(E)ρ(E(cid:48))
∆bcs
1 − ∆2
BCS
EE(cid:48) cos ϕ
× {fα(E)[1 − fα(cid:48)(E(cid:48))]δ(E − E(cid:48) + ω)
+ fα(cid:48)(E(cid:48))[1 − fα(E)]δ(E(cid:48) − E + ω)}
(7)
with the density of states at the Fermi energy of the
normal state N0, the BCS density of states ρ(E) =
2π
(cid:90) dt
αα(cid:48)(cid:82) ∞
×(cid:16)
∆bcs
(cid:82) ∞
(cid:88)
k
†
kak
Eka
(4)
UI (t, t0) = Texp
−i
dt(cid:48) HC,I (t(cid:48))
,
(10)
where UI (t, t0) is the time evolution operator
(cid:20)
(cid:90) t
(cid:21)
2
E/(cid:112)E2 − ∆2
BCS and the quasiparticle distribution func-
†
tion f (Ek) = (cid:104)a
kak(cid:105). We assume that the quasiparticles
can be described with equilibrium BCS gap and density
of states, but with a non-equilibrium distribution func-
tion f (Ek). Due to the square root singularity of the
BCS density of states, the noise spectral density is log-
divergent at low frequencies for tunneling quasiparticles,
e.g. ϕ (cid:54)= 0. Scattering quasiparticles, e.g. ϕ = 0 have a
finite spectral density at low frequencies.
TLS DECOHERENCE
The time evolution of the TLS in the presence of the
quasiparticle reservoirs is best described with the help of
the reduced density matrix
ρ(t) = Tr [(t)]qp ≡
(8)
(cid:18) ρ0 ρ01
(cid:19)
ρ10 ρ1
that is obtained from the full density matrix after tracing
out quasiparticle degrees of freedom. It evolves according
to
ρ(t) = e−iHTLSt Trqp
UI (t)(0)U
†
I (t)
eiHTLSt
(9)
(cid:104)
(cid:105)
t0
and HC,I (t(cid:48)) the coupling in the interaction picture.
It is assumed that the initial density matrix, (t0) ≡
ρ(t0)ρqp(t0), factorizes into a quasiparticle and TLS com-
ponent and that initial correlations are irrelevant on ex-
perimental time scales. We can distinguish two effects
due to the quasiparticles: Decay and decoherence. The
former describes exponential decay of diagonal elements
of the TLS density matrix to their stationary state val-
ues, while the latter concerns the decay of off-diagonal
elements.
Transforming the coupling (3) into the TLS energy
basis we find two contributions σz → /ETLS σz +
∆0/ETLS σx. The off-diagonal term ∼ σx induces transi-
tions and is responsible for the decay rate Γ1. It can be
calculated in first-order perturbation theory [22],
Γ1 =
∆2
0
E2
TLS
[SV (ETLS) + SV (−ETLS)] .
(11)
The diagonal coupling ∼ σz generates pure dephasing,
determined by the low-energy part of the spectral den-
sity. Due to the strong energy dependence in this energy
range pure dephasing does not lead to a simple exponen-
tial decay law. Rather the off-diagonal elements of the
density matrix take the form
ρ10/01(t) = e±iETLSte− 1
2 Γ1te−h(t)
(12)
(cid:17)
(cid:17)
where h(t) describes the deviations from the simple ex-
ponential decay. It reduces to a linear time-dependence
only for flat spectral densities and long times. This form
of the density matrix follows from Eq. (9) and the fact
that the TLS–quasiparticle coupling is diagonal in TLS
space for pure dephasing. Due to the simple coupling we
can pull all TLS operators through the trace and arrive
at the form for the TLS density matrix given in (12) with
(cid:20)
(cid:26)
(cid:90) t
V (t(cid:48))dt(cid:48)(cid:27)
(cid:21)
e−h(t) ≡ Trqp
TC exp
i
ρqp(t0)
.
(13)
t0
The contour time-ordering operator Tc orders along a
contour from t0 to t and back again. To further evaluate
that expression we expand the exponential and introduce
an additional approximation [19]: We assume that we can
split averages over Vi operators in the form
(cid:104)V (t1)V (t2) . . . V (tn)(cid:105) =
(cid:104)V (ti)V (tj)(cid:105)···(cid:104)V (tk)V (tl)(cid:105)
(cid:89)
perm
(14)
With this approximation the quasiparticles behave simi-
lar to a Gaussian noise source [22] and we find
3
NON–EQUILIBRIUM QUASIPARTICLES
Based on the dephasing functions (16) and (15) as well
as the decay rate (11) we are ready to analyze the de-
phasing process due to quasiparticles. The spectral den-
sity (7) depends on the quasiparticle distribution func-
tion. Several experiments provide evidence that, even
at low temperatures where quasiparticles should be ex-
ponentially suppressed with the BCS gap, finite den-
sities of quasiparticles remain, estimated to be nqp ∼
10−6 · ∆BCSN0 [16]. Similar to the treatment of non-
equilibrium quasiparticles for qubit decoherence, e.g. in
Ref. [24], we assume that both, the BCS gap and the
density of states are not changed, but the distribution
function is of a non-equilibrium form. Although the ex-
act form depends on experimental details most of the
non-equilibrium quasiparticles have energies close to the
superconducting gap because of scattering with phonons
and among each other. We therefore assume that the
distribution function has a width δ above the gap, which
for a Fermi distribution is determined by temperature,
δeq ∼ kBT , but here it is treated as a parameter. In the
following we derive analytical forms for the different rates
in the experimental relevant long-time limit t (cid:29) δ−1.
hR(t) = t2
dω Sqp(ω)
sin2 (ωt/2)
(ωt/2)2
(15)
Decay
(cid:90)
(cid:90)
This specific form for pure dephasing is well established
in the context of magnetic resonance or qubit experi-
ments in a Ramsey protocol. The weighting function
g(ωt) = sin2(ωt/2)/(ωt/2)2 has a pronounced peak for
zero energy and decreases rapidly for larger ω. There-
fore, the Ramsey-type experiments are sensitive to the
spectral density at low energies.
Within the Gaussian approximation we can extend our
analysis to more sophisticated measurement protocols,
such as spin echo or more complicated refocusing tech-
niques that suppress low-frequency noise contributions.
The dephasing function h(t) for those protocols looks
very much like the Ramsey function but with different
filter functions depending on the particular pulse proto-
col [23]
h(t) = t2
dω Sqp(ω) g(ωt).
(16)
The relevant energy scale for TLS decay is the TLS
energy splitting ETLS as evident from Eq. (11). The TLS
which can be probed by a qubit have energy splittings
close to that of the qubit and thus fulfill ETLS (cid:29) δ. In
order to evaluate the spectral density in this limit we
introduce the normalized quasiparticle density
(cid:90) ∞
1
∆bcs
∆BCS
xqp =
dE ρ(E)f (E).
(18)
For typical TLS energies ∆BCS (cid:29) ETLS (cid:29) δ we can ap-
ply the ’low-energy’ approximation to evaluate the spec-
tral density at the TLS energy [17, 24]. In this limit all
quasiparticle energies in Eq. (7) can be set to ∆BCS. The
only exception is the quasiparticle energy in the divergent
BCS density of states together with the corresponding
distribution function f (E). They enter in the quasipar-
ticle density (18) and the spectral density, describing the
decay, reads as
E.g. for spin echo, which is the ’first order’ improvement
to the Ramsey experiment, the filter function is
ge(ωt) = sin4 (ωt/4) / (ωt/4)2
(17)
with a maximum slightly shifted to higher frequencies.
For typical experimental times in the range of microsec-
onds the filter function measures the spectral density at
energy equivalents of several MHz.
SV (ETLS) = 4N 2
0 g2
αα(cid:48)∆BCS(xqp,α + xqp,α(cid:48))
× ρ(ETLS + ∆BCS)
1 −
∆BCS
∆BCS + ETLS
cos ϕ
,
(19)
while ∼ S(−ETLS) and the resulting excitation rate is
much smaller. This form for the high-energy spectral
density and thus decay rate Γ1 is well established in the
context of qubit decay due to quasiparticles [16, 25].
(cid:18)
(cid:19)
Ramsey and Spin Echo Dephasing
To calculate the Ramsey dephasing rate (15) we need
an approximation for the low-energy spectral density. We
proceed as in Ref. [19] and split the spectral density into
a regular and a divergent part. For low energies, the
former is flat and can be considered constant, while the
latter is log divergent, Sdiv ∼ (1 − cos ϕ) log(δ/ω). We
find the Ramsey dephasing function
hR(t) = Γ∗
2 t + πN 2
0 g2
αα(cid:48)[fα(∆BCS) + fα(cid:48)(∆BCS)]
×[1 − cos(ϕ)] [γe − 1 + log(4δ · t)] t
(20)
with the Euler constant γe and the pure dephasing rate
Γ∗
2 = 8πN 2
0 g2
αα(cid:48)
[fα(E) + fα(cid:48)(E)]dE .
(21)
(cid:90) ∞
∆BCS
For scattering quasiparticles we have cos ϕ = 1 and the
divergent contribution vanishes. Thus, scattering parti-
cles induce simple exponential dephasing ∼ e−Γ∗
2 t with
rate Γ∗
2. On the other hand, the dephasing effect of tun-
neling quasiparticles is dominated by the second term in
(20) stemming from the divergent contribution.
The spin echo protocol filters out low energies, but
we observe that the relevant energy scales are still much
smaller than the width of the quasiparticles. Thus the
calculation proceeds similar to the calculation for Ram-
sey dephasing, with the result
he(t) = Γ∗
0 g2
2 t + πN 2
×(1 − cos(ϕ))
αα(cid:48)[fα(∆BCS) + fα(cid:48)(∆BCS)]
[γe − 1 + log(δ · t)] t.
(22)
1
2
Since the non-divergent part of the spectral density is
almost flat for the relevant frequency scales, spin echo
does not improve coherence and, similar to white-noise-
induced dephasing, the pure dephasing rate Γ∗
2 is the
same for both protocols, spin echo and Ramsey. Thus,
for scattering electrons there is no measurable difference
between both experimental protocols.aeternus20!0
On the other hand, for tunneling quasiparticles the
second term comes into play and dominates dephasing.
In this limit the ratio between spin echo and Ramsey is
lim
t→∞
he(t)
hr(t)
=
1
2
.
(23)
This improvement in dephasing time due to spin echo is
typical for noise with divergent spectral density at small
energies and could be measured in an experiment.
CONCLUSION
We analyzed the decoherence of TLS located in disor-
dered systems in vicinity to superconducting leads, es-
pecially inside the amorphous layer of a Josephson junc-
tion. We distinguish in our analysis between scattering
4
quasiparticles, that cause decoherence for all the TLS
mentioned above, and quasiparticles that tunnel through
the junction. If there exists a phase difference between
the superconducting electrodes, the latter exhibit a log-
divergent spectral density for low energies leading to in-
creased and time-dependent dephasing rates while the
difference is negligible for the TLS decay. We further
showed that the spin echo technique reduces the TLS de-
coherence rate due to tunneling particles, while it has
little effect for scattering quasiparticles. The results ob-
tained for tunneling quasiparticles, arising form a diver-
gent spectral density, apply to single-junction qubits, and
they are sensitive to refocusing techniques. This opens
possibilities to analyze the quasiparticle-environment of
a qubit.
ACKNOWLEDGMENTS
We thank A. Bilmes, J. Lisenfeld and A. Shnirman for
many useful discussions during the work on this paper.
This work was supported by the German-Israeli Founda-
tion for Scientific Research and Development (GIF).
[1] Barends R., Kelly J., Megrant A., Veitia A., Sank D., Jef-
frey E., White T. C., Mutus J., Fowler A. G., Campbell
B., Chen Y., Chen Z., Chiaro B., Dunsworth A., Neill
C., O’Malley P., Roushan P., Vainsencher A., Wenner J.,
Korotkov A. N., Cleland A. N., and Martinis John M.,
Nature 508, 500 (2014).
[2] M. H. Devoret and R. J. Schoelkopf, Science 339, 1169
(2013).
[3] A. Shnirman, Y. Makhlin, and G. Schon, Physica Scripta
2002, 147 (2002).
[4] R. W. Simmonds, K. M. Lang, D. A. Hite, S. Nam, D. P.
Pappas, and J. M. Martinis, Phys. Rev. Lett. 93, 077003
(2004).
[5] P. Esquinazi, Tunneling Systems in Amorphous and
Crystalline Solids (Springer, 1998).
[6] L. Faoro and L. B. Ioffe, Phys. Rev. Lett. 109, 157005
(2012).
[7] J. Schriefl, Y. Makhlin, A. Shnirman, and G. Schon,
New Journal of Physics 8, 1 (2006).
[8] T. C. DuBois, S. P. Russo, and J. H. Cole, New Journal
of Physics 17, 023017 (2015).
[9] W. A. Phillips, Reports on Progress in Physics 50, 1657
(1987).
[10] A. M. Holder, K. D. Osborn, C. J. Lobb, and C. B.
Musgrave, Phys. Rev. Lett. 111, 065901 (2013).
[11] D. Gunnarsson, J.-M. Pirkkalainen, J. Li, G. S.
Paraoanu, P. Hakonen, M. Sillanp, and M. Prunnila, Su-
perconductor Science and Technology 26, 085010 (2013).
[12] J. M. Martinis, K. B. Cooper, R. McDermott, M. Stef-
fen, M. Ansmann, K. D. Osborn, K. Cicak, S. Oh, D. P.
Pappas, R. W. Simmonds, and C. C. Yu, Phys. Rev.
Lett. 95, 210503 (2005).
5
[13] J. Lisenfeld, C. Muller, J. H. Cole, P. Bushev,
A. Lukashenko, A. Shnirman, and A. V. Ustinov, Phys.
Rev. Lett. 105, 230504 (2010).
[14] J. Lisenfeld, A. Bilmes, S. Matityahu, S. Zanker,
M. Marthaler, G. Weiss, A. Shnirman, M. Schechter, and
A. Ustinov, “Decoherence of individual two-level systems
in dependence of their strain-tuned asymmetry energy,”
(2015), unpublished.
[15] G. J. Grabovskij, T. Peichl, J. Lisenfeld, G. Weiss, and
A. V. Ustinov, Science 338, 232 (2012).
[16] J. M. Martinis, M. Ansmann, and J. Aumentado, Phys.
Rev. Lett. 103, 097002 (2009).
[17] J. Leppakangas and M. Marthaler, Phys. Rev. B 85,
144503 (2012).
[18] D. G. Olivares, A. L. Yeyati, L. Bretheau, i. m. c. O.
and C. Urbina, Phys. Rev. B 89,
Girit, H. Pothier,
104504 (2014).
[19] S. Zanker and M. Marthaler, Phys. Rev. B 91, 174504
(2015).
[20] W. Phillips, Journal of Low Temperature Physics 11, 757
(1973).
[21] J. L. Black and B. L. Gyorffy, Phys. Rev. Lett. 41, 1595
(1978).
[22] G. Ithier, E. Collin, P. Joyez, P. J. Meeson, D. Vion,
D. Esteve, F. Chiarello, A. Shnirman, Y. Makhlin,
J. Schriefl,
and G. Schon, Phys. Rev. B 72, 134519
(2005).
[23] Bylander Jonas, Gustavsson Simon, Yan Fei, Yoshi-
hara Fumiki, Harrabi Khalil, Fitch George, Cory David
G., Nakamura Yasunobu, Tsai Jaw-Shen,
and Oliver
William D., Nat Phys 7, 565 (2011).
[24] G. Catelani, S. E. Nigg, S. M. Girvin, R. J. Schoelkopf,
and L. I. Glazman, Phys. Rev. B 86, 184514 (2012).
[25] G. Catelani, R. J. Schoelkopf, M. H. Devoret, and L. I.
Glazman, Phys. Rev. B 84, 064517 (2011).
|
1205.2029 | 1 | 1205 | 2012-05-09T16:40:11 | Mode bifurcation on the rythmic motion of a micro-droplet under stationary DC electric field | [
"cond-mat.mes-hall",
"physics.bio-ph"
] | Accompanied by the development of microfabrication techniques, such as MEMS and micro-TAS, there has been increasing interests on the methodology to generate a desired motion on a micro object in a solution environment. It is well know that the principle to create an electric motor in a macroscopic scale is not applicable to a micro system because of the enhancement of sticky interaction and higher viscosity in micrometer sized system. On the other hand, living organisms generate various motions under isothermal condition in a well-regulated manner. Despite the past intensive studies, the underlying mechanism of the biological molecular motors has not been unveiled yet. Under such development stage of science and technology at the present, we are performing the study toward real-world modeling on the emergence of regular motion in micro system. In this paper, we report a novel experimental system on the generation of regular movement, such as periodic go-back motion and circular motion, for a micro object under DC electric field. | cond-mat.mes-hall | cond-mat | Mode bifurcation on the rythmic motion of a micro-droplet under stationary DC
electric field
Tomo Kurimura,1, a) Masahiro Takinoue,2 Masatoshi Ichikawa,1 and Kenichi Yoshikawa3
1)Department of Physics, Graduate School of Science, Kyoto University,
Kitashirakawa-Oiwake-cho, Sakyo-ku, Kyoto 606-8502 Japan
2)Interdisciplinary Graduate School of Science and Engineering,
Tokyo Institute of Technology, 4259 Nagatsuta-cho, Midoriku, Yokohama,
Kanagawa 226-8503, Japan
3)Faculty of Biological and Medical Sciences, Doshisha Univ.1-3 Tataramiyakodani,
Kyotanabe, Kyoto 610-0394, Japan
(Dated: May 2012)
2
1
0
2
y
a
M
9
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
1
v
9
2
0
2
.
5
0
2
1
:
v
i
X
r
a
a)Electronic mail: kurimura@chem.scphys.kyoto-u.ac.jp
1
I.
INTRODUCTION
Accompanied by the development of microtechnology, such as MEMS and µTAS, there
is increasing interest on the methodology to realize a desired motion of a micro object in a
solution environment. It is well known that the principle to create an electric motor in a
macro system is not applicable to micro system because of the enhanced sticky interaction
and higher viscosity in micrometer sized system. On the other hand, living organisms
generate various motions on microscopic scale under isothermal condition. Despite the
past intensive studies1,2, the underlying mechanism of the biological molecular motors has
not been fully unveiled yet. Under such development status of science and technology on
micro-motor sat the present, we report a simple motoring system which work smoothly in a
microscpic scale.
Recently, we found that rhythmic motion is generated for an aqueous droplet in an
oil phase under DC voltage on the order of 50 - 100 V. We have already reported some
experiments and models for a w/o droplet under DC electric field.3,4 There are some reports
about experiments of w/o droplets moving under electrical field5,6, bouncing and being
absorbed on a surface between water and oil7, deforming and spliting8,9. Manipulating this
kind of droplet, which is interesting as the model of the cell10 -- 12, the micro-sized reactor,
by optical tweezers13, by micro channel14,15. And manipulating the cells or micro objects by
electrical field has been attempted16, to know manipulating this kind of droplets in detail
will help this in the future. In the present article, we will show that rhythmic motion on
micro-droplet is induced under the DC potential on the order of several volts. We will
also propose a simple mathematical model to reproduce the rhythmic motion and mode
bifurcation.
II. EXPERIMENTAL
A schematic illustration of the experimental setup is given in FIG.1. A water droplet
was suspended in mineral oil on a glass slide, and constant voltage was applied to the
droplet using cone-shaped tungsten electrodes. Droplet motion was observed using an optical
microscope (KEYENCE , Japan).
The w/o droplet was generated using a vortex mixer as follows. We prepared mineral
2
oil including surfactant: 10µm surfactant, dioleylphosphatidylcholine (DOPC) (Japan), was
solved in mineral oil (Nacalai Tesque, Japan) by 90 min sonication at 50 ◦C. 2µl ultrapure
water (Millipore, Japan) was added to 100 µl of the prepared mineral oil, and then agitated
by a vortex mixer for approximately 3 s.
(cid:80)(cid:67)(cid:75)(cid:70)(cid:68)(cid:85)(cid:74)(cid:87)(cid:70)(cid:1)(cid:77)(cid:70)(cid:79)(cid:84)(cid:1)
(cid:80)(cid:71)(cid:1)(cid:78)(cid:74)(cid:68)(cid:83)(cid:80)(cid:84)(cid:68)(cid:80)(cid:81)(cid:70)
(cid:78)(cid:74)(cid:79)(cid:70)(cid:83)(cid:66)(cid:77)(cid:1)(cid:80)(cid:74)(cid:77)
(cid:88)(cid:66)(cid:85)(cid:70)(cid:83)(cid:1)(cid:69)(cid:83)(cid:80)(cid:81)(cid:77)(cid:70)(cid:85)
(cid:55)
(cid:45)
(cid:72)(cid:77)(cid:66)(cid:84)(cid:84)(cid:1)(cid:81)(cid:77)(cid:66)(cid:85)(cid:70)
FIG. 1. Schematic representation of the experimental setup. Mineral oil containing water droplets
was placed on a glass slide and couple of electrodes was situated inside the oil phase. V: Applied
DC voltage. L: distance between the electrodes.
III. RESULTS
FIG.2 exemplifies the motion of a droplet under DC electric field, indecating the occur-
rence of the periodic go-back motion between the electrodes accompanied by the increase
of the electical potential. In the experiments, we observed two following types of behavior:
oscillatory and stationary. These behaviors switch each other depending on the applied volt-
age. When the distance between two electrodes was 213µm [FIG.2(a)], the droplet started
the motion with the applied voltage above 16.3 V. When the distance between two electrodes
was 141µm [FIG.2(b)], the droplet started moving with the applied voltage above 13.7 V.
FIG.3 shows the diagram of the mode of droplet behavior depending on the applied
voltage with the size of the droplet is -- -. When the distance of two electrodes is below
approximately 70µm , droplets are sticked to an electrode (adhered). The diagram indecates
3
that the threshold of applied voltage is roughly propotional to the distance between two
electrodes.
(a) L = 213µm
(b) L = 141µm
V = 16.0V
V = 16.3V
t=0 [s]
0
1
]
s
[
t
2
3
V = 13.0V
V = 13.7V
0
1
2
3
100μm
FIG. 2. Spatio-temporal diagram on the motion of a droplet with the diameter of 34 m at (a)
L=213m, (b) L=141m.@Bifurcation from the stationary state into an oscillatory state is induced
by the increase of the applied voltage.
IV. DISCUSSION
We propose a model to describe the oscillatory-stationary motion of w/o droplets. In
an eqation of motion at a micrometer scale, a viscosity term is more dominant than inertia
term because the Reynolds number, Re, is rather small; Re = ρvd/η ∼ 10−9 ≪ 1, where
ρ(∼ 103kg/m3) and η(∼ 103Pas) are the density and viscosity of the mineral oil, respectively,
and v(∼ 10−4m/s) and d(∼ 10−5m) are the velocity and diameter of the water droplet.
Therefore, an over-damped eqation of motion under the constant electric field, E, is given
4
25
20
]
V
[
V
15
10
5
0
0
Diameter of the droplet
20µm
34µm
50µm
d
Oscillatory
Adhered
2-(b)
2-(a)
Stationary
50
100
150
200
250
L [µm]
FIG. 3. Phase diagram for the mode bifurcation between rhythmic motion and stationary state as
obderved for the droplets with different diameter, where each point represent the threshold value
on the bifurcation. The two arrows are correspond to the mode bifurcation as shown in FIG.2
by
k x = qE + α∇E2
(1)
where k(= 6πηd ∼ 10−7kg/s) is a coefficient of viscosity resistance, and k x represents the
viscosity resistance for a moving droplet with diameter d and velocity x. qE and α∇E2
indicate an electric force and a dielectric force acting on the droplet with charge q and
polarizability α (Il)17.
Here we assume that the time-dependent rate od the charge, q, is described as
q = −βǫx3 −
q
t0
(2)
where β is the proportionality coefficient, and ǫ is the constrant is proportional to the
magnitude of the electrical field. The first term of this means the time-dependence of charge
is in proportion to the number of lines of electric force. The second term means the charge
leak, and t0 is the relaxing time.
We would like to consider the condition where the droplet stays on the same position
between two electrodes. When the electric field can be written as E = (Ex, Ey), The force
on the second term in the right hand of Eq. (1) is caused by the number of the lines of
electric force penetrating the droplet. Comparing with the size of droplet, the change of Ex
5
along x axis can be neglected. By considering the symmetry of the system, we simply adapt
that the change of E2
y along x axis is written as
E2
y = ǫ (cid:0)−(x + 1)2(x − 1)2 + 2(cid:1) .
Then the x component of eq.(1) is given as
k x =
∼=
qEx + α∂xE2
qEx + α∂xE2
y
=qEx − 4αǫx(x + 1)(x − 1)
x = Ex
k q − 4αǫ
k x(x + 1)(x − 1)
(3)
(4)
For simplicity, we introduce the following parameters: Ex/k = a, 4αǫ/k = e and keβ/4α =
γ. Then Eq.(2) and Eq.(4) can be written as
x = − ex(x + 1)(x − 1) + αq
q =
− γex3 − q/t0
(5)
(6)
FIG.4 shows the result of numeric calculation with these equations, where the time and
space scales ,T and X, are arbitrary. The change of the distance between the electorodes
corresponds to the change of the magnitude of the electric field. For example, if L becomes
larger, a and e become smaller. In FIG.4, a and e in (b) are larger than those in (a). The
frequency of the back-and-force motion of the droplet is faster when the electric field between
the electrodes becomes stronger. Thus, our numerical model reproduces the essential aspect
of the rhythmic motion of a droplet under DC voltage.
V. ACKNOWLEDGEMENTS
REFERENCES
1Y. Hiratsuka, M. Miyata, T. Tada, and T. Q. P. Uyeda, Proc. Natl. Acad. Sci. U.S.A. 103,
13618 (2006).
2M. G. L. van den Heuvel and C. Dekker, Science 317, 333 (2007).
3Hase, et al., PRE 74, 046301(2006).
4Takinoue, et al. Appl. Phys. Lett. 96, 104105 (2010)
6
(a)
0
5
10
15
τ
20
25
30
35
40
(b)
0
5
10
15
τ
20
25
30
35
40
-0.8
-0.4
0
X
0.4
0.8
-1 -0.8
-0.4
0.4
0.8 1
0
X
FIG. 4. Numerical results on the Spatio-Temporal diagram with eqs. (5) and (6), the common
parameters are t0 = 0.5 and γ = 0.1. The parameters a and e are changed at T = 15. Initial state
is the same in both graphs; a = 100 and e = 2. After T = 15 in (a), a = 200 and e = 4. In (b),
a = 250 and e = 5.
(a) L = 213µm
(b) L = 141µm
0
1
t
2
3
-1
-0.5
0
0.5
1
-1
-0.5
0
0.5
1
x/L
x/L
FIG. 5. Experimental results on the Spatio-Temporal diagram
7
5T. Mochizuki, Y. Mori, and N. Kaji, AIChE J. 36, 1039 (1990).
6Y. Jung, H. Oh, and I. Kang, J. Colloid Interface Sci. 322, 617 (2008).
7W. D. Ristenpart, J. C. Bird, A. Belmonte, F. Dollar, and H. A. Stone, Nature (London)
461, 377 (2009).
8J. S. Eow, M. Ghadiri, and A. Sharif, Colloids Surf., A 225, 193 (2003).
9S. Teh, R. Lin, L. Hung, and A. Lee, Lab Chip 8, 198 (2008).
10A. V. Pietrini and P. L. Luisi, ChemBioChem 5, 1055 (2004).
11D. S. Tawfik and A. D. Griffiths, Nat. Biotechnol. 16, 652 (1998).
12M. Hase and K. Yoshikawa, J. Chem. Phys. 124, 104903 (2006).
13S. Katsura, A. Yamaguchi, H. Inami, S. Matsuura, K. Hirano, and A. Mizuno, Elec-
trophoresis 22, 289 (2001).
14D. R. Link, E. Grasland-Mongrain, A. Duri, F. Sarrazin, Z. Cheng, G. Cristobal, M.
Marquez, and D. A. Weitz, Angew. Chem. Int. Ed. 45, 2556 (2006).
15J. Atencia and D. J. Beebe, Nature (London) 437, 648 (2005).
16J. Voldman, Annu. Rev. Biomed. Eng. 8, 425 (2006).
17T. B. Jones, Electromechanics of Particles (Cambridge University Press, New York, 1995).
8
|
1811.03229 | 6 | 1811 | 2019-07-02T08:23:13 | Temperature dependence of side-jump spin Hall conductivity | [
"cond-mat.mes-hall"
] | In the conventional paradigm of the spin Hall effect, the side-jump conductivity due to electron-phonon scattering is regarded to be temperature independent. To the contrary, we draw the distinction that, while this side-jump conductivity is temperature independent in the classical equipartition regime where the longitudinal resistivity is linear in temperature, it is temperature dependent below the equipartition regime. The mechanism resulting in this temperature dependence differs from the familiar one of the longitudinal resistivity. In the concrete example of Pt, we show that the change of the spin Hall conductivity with temperature can be as high as 50%. Experimentally accessible high-purity Pt is proposed to be suitable for observing this prominent variation below 80 K. | cond-mat.mes-hall | cond-mat | Temperature dependence of side-jump spin Hall conductivity
Cong Xiao,1, ∗ Yi Liu,2 Zhe Yuan,2, † Shengyuan A. Yang,3 and Qian Niu1
1Department of Physics, The University of Texas at Austin, Austin, Texas 78712, USA
2The Center for Advanced Quantum Studies and Department of Physics,
Beijing Normal University, 100875 Beijing, China
3Research Laboratory for Quantum Materials, Singapore University of Technology and Design, Singapore 487372, Singapore
In the conventional paradigm of the spin Hall effect, the side-jump conductivity due to electron-
phonon scattering is regarded to be temperature independent. To the contrary, we draw the distinc-
tion that, while this side-jump conductivity is temperature independent in the classical equipartition
regime where the longitudinal resistivity is linear in temperature, it is temperature dependent below
the equipartition regime. The mechanism resulting in this temperature dependence differs from the
familiar one of the longitudinal resistivity. In the concrete example of Pt, we show that the change
of the spin Hall conductivity with temperature can be as high as 50%. Experimentally accessible
high-purity Pt is proposed to be suitable for observing this prominent variation below 80 K.
The spin Hall effect refers to a transverse spin current
in response to an external electric field [1]. In strongly
spin-orbit-coupled electronic systems such as 4d and 5d
transition metals [2 -- 10] and Weyl semimetals [11], the
spin Hall conductivity due solely to the geometry of Bloch
bands, the so-called spin Berry curvature, has attracted
much attention. Besides, there is a scattering induced
mechanism called side-jump, whose contribution to the
spin Hall conductivity turns out to be of zeroth order
of scattering time and independent of the density of a
given type of impurities [1]. Furthermore, the side-jump
spin Hall conductivity arising from the electron-phonon
scattering is conventionally regarded to be temperature
(T ) independent although the phonon density varies with
T [12, 13].
In this work we draw the distinction that, while the
electron-phonon scattering induced side-jump spin Hall
conductivity is T -independent in the classical equipar-
tition regime where the longitudinal resistivity ρ is lin-
ear in T , it is T -dependent at temperatures below the
equipartition regime. This character distinguishes side-
jump from the geometric contribution, and provides a
new mechanism for T -dependent spin Hall conductivi-
ties in high-purity experimental samples. An intuitive
picture is proposed for the T -dependence of the side-
jump conductivity, which differs from ρ that is always
T -dependent. Moreover, our first-principles calculation
demonstrates a prominent T -variation of the spin Hall
conductivity in experimentally accessible high-purity Pt
below 80 K.
We consider strongly spin-orbit coupled multiband sys-
tems. The Fermi energy and the interband-splitting
around the Fermi level are assumed to be much larger
than the room temperature, thus the thermal smearing
of Fermi surface is negligible. Aiming to provide semi-
quantitative and intuitive understanding, the electron-
phonon scattering is approximated by a single-electron
elastic process, which can be called the "quasi-static
approximation". In calculating the resistivity resulting
from phonon scattering, this approximation produces not
only the correct low-T power law (ρ ∼ T 5 for three-
dimensional isotropic single-Fermi-surface systems) [14]
but also the values that are quantitatively comparable
with experimental data [15]. When applied to the side-
jump transport, the high-T and low-T asymptotic be-
haviors are grasped in this approximation. Quantita-
tive deviations appearing in the intermediate tempera-
ture regime are not essential for the present purpose.
The side-jump was originally proposed as the side-way
shift in opposite transverse directions for the carriers with
different spins, when they are scattered by spin-orbit ac-
tive impurities [12, 16]. This picture works well in sys-
tems with weak spin-orbit coupling [17 -- 19], where the
spin-orbit-induced band splitting is smeared by disorder
broadening [1]. Whereas in strongly spin-orbit-coupled
Bloch bands of current interest, the side-jump contribu-
tion arises microscopically from the scattering-induced
band-off-diagonal elements of the out-of-equilibrium den-
sity matrix [20 -- 23]. This corresponds to in the Boltz-
mann transport formalism the dressing of Bloch states
by interband virtual scattering processes involving off-
shell states away from the Fermi surface [24].
Transport
formalism involving off-shell states. -- In
weakly disordered crystals perturbed by an weak external
electric field E, the expectation value of an observable A
(assumed to be a vector without loss of generality) reads
(cid:88)
(cid:104)A(cid:105) =
Alfl
(1)
l
in the Boltzmann transport formalism, where fl is the oc-
cupation function of the carrier state marked by l = (η, k)
with η the band index and k the crystal momentum, Al
is the quantum mechanical average on state l. The car-
rier state is the Bloch state dressed by interband virtual
processes induced by both the electric field and scattering
[24]. In the linear response and weak scattering regime,
these two dressing effects are independent [24]:
Al = A0
l + Abc
l + Asj
l ,
(2)
9
1
0
2
l
u
J
2
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
6
v
9
2
2
3
0
.
1
1
8
1
:
v
i
X
r
a
where
(cid:17)
(cid:16)
Abc
l
e
EαΩA
αβ (ηk)
=
β
(3)
2
arises from the electric-field induced dressing, with
(cid:88)
η(cid:48)(cid:48)(cid:54)=η
vηη(cid:48)(cid:48)
α
(k) Aη(cid:48)(cid:48)η
(cid:0)ηk − η(cid:48)(cid:48)k
β
(cid:1)2
(k)
,
(4)
αβ (ηk) = −22 Im
ΩA
(cid:17)
and(cid:16)
Asj
l
β
η(cid:48)k(cid:48)
= −2π
(cid:88)
Wkk(cid:48)δ(cid:0)ηk − η(cid:48)k(cid:48)(cid:1)
(cid:88)
− (cid:88)
η(cid:48)k(cid:48) − η(cid:48)(cid:48)k(cid:48)
η(cid:48)(cid:48)(cid:54)=η(cid:48)
(cid:104)uη(cid:48)(cid:48)kuη(cid:48)k(cid:48)(cid:105)(cid:104)uη(cid:48)k(cid:48)uηk(cid:105)Aηη(cid:48)(cid:48)
× Im
β
ηk − η(cid:48)(cid:48)k
η(cid:48)(cid:48)(cid:54)=η
(cid:104)uηkuη(cid:48)k(cid:48)(cid:105)(cid:104)uη(cid:48)(cid:48)k(cid:48)uηk(cid:105)Aη(cid:48)η(cid:48)(cid:48)
β
(k)
(cid:0)k(cid:48)(cid:1)
.
(5)
originates from the scattering-induced dressing. Equa-
tion (5) is diagrammatically represented in Fig. 1. The
summation over repeated spatial indices α, β is implied
hereafter. Here Aη(cid:48)(cid:48)η
(k) ≡ (cid:104)uη(cid:48)(cid:48)kAβuηk(cid:105) with uηk(cid:105)
the periodic part of the Bloch state. For impurities
, with ni the impurity density and V o
Wkk(cid:48) = ni
kk(cid:48)
the plane-wave part of the matrix element of the impurity
potential. For electron-phonon scattering
(cid:12)(cid:12)V o
kk(cid:48)(cid:12)(cid:12)2
β
(cid:12)(cid:12)U o
kk(cid:48)(cid:12)(cid:12)2
Wkk(cid:48) =
2Nq
V
,
(6)
where U o
k(cid:48)k is the plane-wave part of the electron-phonon
matrix element, Nq is the Bose occupation function of
phonons (q is the wave-vector of phonons with energy
ωq), V is the volume (area in two-dimension) of the
system, and the factor 2 accounts for the absorption and
emission of phonons.
When calculating the electric current A = ev, ΩA
αβ and
vsj
l are the Berry curvature and "side-jump velocity" [22,
24, 25], respectively. When calculating the spin current
A = j, ΩA
αβ is the so-called spin Berry curvature [11],
whereas Asj
l provides the spin-current counterpart of the
side-jump velocity [24, 26].
The occupation function of the carrier states is de-
composed, around the Fermi distribution f 0
l , into fl =
f 0
l + g2s
Its out-of-equilibrium part satisfies
the linearized steady-state Boltzmann equations eE ·
l ∂f 0
v0
l /∂l = −(cid:80)
ll(cid:48)(cid:0)g2s
l + ga
l .
l(cid:48) w2s
l(cid:48) (cid:1) and
(cid:88)
l − g2s
eE · vsj
l ∂f 0
l /∂l =
w2s
ll(cid:48) (ga
l − ga
l(cid:48))
(7)
l(cid:48)
Wkk(cid:48)(cid:12)(cid:12)(cid:104)ulul(cid:48)(cid:105)(cid:12)(cid:12)2
tering rate, v0
in the presence of weak scalar disorder [22]. w2s
ll(cid:48) = 2π
δ (l − l(cid:48)) is the lowest-Born-order scat-
l is the usual band velocity.
FIG. 1. Graphical representation of Eq. (5), where the off-
shell states away from the Fermi surface are marked by red
arrows. Wkk(cid:48) is represented by the disorder line connected
with two interaction vertexes.
(cid:17)
(cid:16)
(cid:88)
SH = (cid:80)
l
Collecting the above ingredients, the spin Hall current
is [27, 28] jSH = jbc
SH + jsj
SH + jad
SH. The first two terms
jbc
SH =
jbc
l
l , and jsj
f 0
SH =
jsj
l
g2s
l
(8)
(cid:17)
(cid:88)
(cid:16)
l
l j0
l ga
l
AH and jad
l . Whereas jad
arise from off-shell-states induced corrections to the semi-
classical value of j0
incorpo-
rates the nonequilibrium occupation function modified
by off-shell states, since ga
l appears as a response to the
generation term proportional to the "side-jump velocity"
vsj
[25]. In calculating the anomalous Hall (AH) current
l
[21] A = ev, jsj
AH are related to the transverse
(side-way) and longitudinal components of vsj
l , respec-
tively [25]. Thereby their sum is also often referred to
as the side-jump contribution in the literature on the
anomalous Hall effect [21, 22, 25]. Given this conven-
tion, we also include the jad
SH contribution, although jad
AH
has nothing to do with the original concept of side-jump,
and the microscopic theory [20, 23] indeed shows that ga
l
is not related to the interband elements of the out-of-
equilibrium density matrix. In fact, in two-dimensional
nonmagnetic models for the spin Hall effect, such as the
two-dimensional electronic systems with Rashba, cubic
Rashba and Dresselhaus spin-orbit couplings [29 -- 34], the
spin current operator (for out-of-plane spin component)
has only interband matrix elements, i.e., j0
l = 0, thus jad
SH
does not appear at all [27, 28].
Phonon-induced T-dependence of spin Hall conductiv-
ity. -- In order to show the T -dependence of the phonon-
induced side-jump spin Hall conductivity, we prove that
its values in the low-T and high-T limits can be different.
In the low-T limit, Wkk(cid:48) for phonon scattering is highly
peaked at vanishing scattering angle, and the on-shell
scattering can only be the intraband transition, hence in
Eq. (5) η(cid:48) = η, and k(cid:48) is very close to k. We then ex-
pand the integrand up to the first order of(cid:0)k(cid:48) − k(cid:1), get-
(cid:16)
(cid:17)
= (cid:80)
ting
Asj
l
β
k(cid:48) w2s
l(cid:48)lΩA
αβ (ηk)(cid:0)k(cid:48) − k(cid:1)
α, with w2s
l(cid:48)l =
𝐴𝛽𝜂′𝜂′′η′𝐤′η𝐤η′′𝐤′η𝐤𝐴𝛽𝜂′′𝜂′η′′𝐤′η′𝐤′η𝐤η𝐤𝐴𝛽𝜂′′𝜂η′′𝐤η𝐤η𝐤η𝐤𝐴𝛽𝜂𝜂′′η′′𝐤η′𝐤′η′𝐤′η𝐤k(cid:48) w2s
curvature) yields ga
scattering-angle limit. Concurrently, vsj
2π Wkk(cid:48)δ(cid:0)ηk − ηk(cid:48)(cid:1) the scattering rate in the small-
×(cid:0)k(cid:48) − k(cid:1) (Ω is the vector form of the ordinary Berry
has [28](cid:16)
(cid:17)
and(cid:16)
l =(cid:80)
l = eE·(cid:2)k × Ω (l)(cid:3) ∂f 0
(cid:88)
(cid:17)
(cid:16)
E ·(cid:2)k × Ω (l)(cid:3) ∂f 0
αβ (l) E · v0
l /∂l. Thus one
(cid:88)
l(cid:48)lΩ (l)
l /∂l,
kαΩj
l ∂f 0
(cid:17)
jsj
SH
(10)
= e
= e
(9)
β
l
l /∂l.
jad
SH
j0
l
β
β
l
Thus σSH = (jSH)x /Ey is a T -independent constant in
the low-T limit. It is clear that this constant equals that
contributed by scalar-impurities in the long-range limit,
whose Wkk(cid:48) is also highly concentrated around vanishing
scattering angle.
than kBT indicating Wkk(cid:48) = 2kBT V−1(cid:12)(cid:12)U o
In the high-T limit, the phonon energy is much smaller
/ωq,
l ∼ T 0, and
l ∼ T , and ga
l ∼ T −1, jsj
(cid:12)(cid:12)2
k(cid:48)k
then we have g2s
consequently,
ρ ∼ T, σSH ∼ T 0.
(11)
Accordingly, in practice the high-T limit is identified as
the equipartition regime with linear-in-T resistivity [14].
This regime is usually marked qualitatively by T > TD
in textbooks, with TD the Debye temperature, but can
extend practically to about T > TD/3 in Pt, Cu and Au
[15, 35], and to about T > TD/5 in Al [35]. Besides, it is
apparent that the T -independent σSH in the equipartition
regime can be different from that in the low-T limit.
k(cid:48)k
(cid:12)(cid:12)2
troduced as [36, 37] λ2 = 2V−1(cid:12)(cid:12)U o
To acquire a more transparent picture, we assume
any large-angle electron-phonon scattering can occur via
normal processes, and take the approximation of the
deformation-potential electron-acoustic phonon coupling,
for which a electron-phonon coupling constant can be in-
/ωq, hence arriv-
ing at Wkk(cid:48) = λ2kBT in the high-T regime. This Wkk(cid:48)
is uniformly distributed on the Fermi surface, just simi-
lar to Wkk(cid:48) = niV 2
i contributed by randomly distributed
zero-range scalar impurities, with Vi the strength. There-
fore, we infer that the σSH due to phonons in the high-T
equipartition regime takes the same value as that due
to zero-range scalar impurities. This speculation can
be verified by noticing that g2s,ep
niV 2
i ,
Asj
= ga,ei
,
l
where the superscripts "ep" and "ei" mean the contribu-
tions due to electron-phonon scattering and zero-range
scalar impurities, respectively.
λ2kBT = g2s,ei
i and ga,ep
l
/niV 2
/λ2kBT =
(cid:17)ep
(cid:17)ei
(cid:16)
(cid:16)
Asj
l
l
l
l
According to the above results, σSH induced by
electron-acoustic phonon scattering is T -dependent pro-
vided that the σSH induced by scalar impurity scatter-
ing in the long-range and zero-range limits are different.
3
This unexpected relation in turn provides a qualitative
picture for comprehending the T -dependence of phonon-
induced side-jump, by analogy with the recently revealed
sensitivity of the side-jump conductivity to the scatter-
ing range of impurities [38]. The accessible phase-space
of the electron-phonon scattering changes with temper-
ature, thus implies a T -dependent average momentum
transfer, i.e., effective range, of this scattering.
Note that this mechanism differs from that for the T -
dependent ρ. To directly see this point, one need just
consider the fact that in the equipartition regime σSH ∼
T 0 although ρ ∼ T is still T -dependent.
The above revealed relation facilitates judging whether
a model system has a T -dependent phonon-induced side-
jump conductivity based on the familiar knowledge about
the impurity-induced side-jump. There are models which
possess different side-jump conductivities induced by
scalar impurity-scattering in the long-range and zero-
range limits, such as the Luttinger model describing p-
type semiconductor [39] and the k-cubic Rashba model
for the two-dimensional heavy-hole gas in confined quan-
tum wells [30]. In these systems the phonon-induced side-
jump conductivities are thus T -dependent.
In the k-cubic Rashba model [33, 34], the Hamiltonian
reads
H =
2k2
2m
+ i
αR
2
(cid:16)
(cid:17)
σ+k3− − σ−k3
+
,
(12)
1
where k = k (cos φ, sin φ) is the two-dimensional wave-
vector, k± = kx ± iky, σ's are Pauli matrices with σ± =
σx ± iσy, αR is the spin-orbit coupling coefficient that
can be tuned to very large values by the gate voltage
2 {σz, vx}
[34]. The spin current operator [31, 32] x = 3
SH + jsj
has only interband components, hence jSH = jbc
SH.
Letting η = ± labels the two Rashba bands, the spin
Berry curvature is Ωj
4mαRk3 , thus σbc
SH =
x /Ey = 9e2
[30, 31], with kη the
Fermi wave-number of band η. The side-jump spin Hall
conductivity due to electron-phonon scattering in the
(cid:16)
low-T limit is given by Eq. (9) as
(cid:80)
yx (ηk) = −η 93 sin2 φ
η ηk−1
(cid:17)
(cid:0)jbc
16πmαR
(cid:1)
SH
η
2
1
4
σbc
SH.
(13)
σsj
SH =
When mαR/2 (cid:28) 1/
In the high-T regime, we have jsj
jsj
SH
√
x
/Ey =
πn [31], one has σsj
SH = 9e/32π.
l = 0 [28], leading to
σsj
SH = 0.
(14)
Since the side-jump conductivities in the high-T and low-
T limits are different, there must be a crossover in the
intermediate regime resulting in the T -dependent behav-
ior.
Temperature -- dependent spin Hall conductivity in pure
Platinum. -- To show the applicability of our theoreti-
cal ideas in real materials, we perform a first-principles
calculation to the spin Hall conductivity of pure Pt in
the range 20 -- 300 K. The minimal
interband split-
ting around the Fermi level of Pt is much larger than
300 K [9], thus the spin Berry-curvature contribution
should be T -independent up to 300 K. The tempera-
ture is modeled by populating the calculated phonon
spectra of Pt into a large supercell with its length L
along fcc [111] and 5 × 5 unit cells in the lateral di-
mensions [15, 40]. Then the transport calculation is car-
ried out using the above disordered (finite-temperature)
supercell sandwiched by two perfectly crystalline (zero-
temperature) Pt electrodes. The scattering matrix is
obtained using the so-called "wave function matching"
technique within the Landauer-Buttiker formalism [40].
The calculated total resistance of the scattering geome-
try is found to be linearly dependent on L following the
Ohm's law. By varying L in the range of 5 -- 60 nm,
we extract the resistivity at every temperature using a
linear least squares fitting for the calculated resistances.
For each L, at least 10 random configurations have been
considered to ensure both average value and standard
deviation well converged with respect to the number of
configurations. The calculated resistivity ρ is plotted in
Fig. 2(a) as a function of temperature. The spin-Hall
angle ΘSH is computed by examining the ratio of trans-
verse spin current density and longitudinal charge current
density [10]. At every temperature, we use more than 20
random configurations, each of which contains 60 nm-
long disordered Pt. Then the spin Hall conductivity is
obtained as σSH = (/e)ΘSH/ρ, shown in Fig. 2(b).
4
low the equipartition regime, in agreement with our the-
oretical prediction.
Finally, we discuss the possibility of observing the pre-
dicted effect in experiments. In high-purity metals, the
electron-electron scattering dominates over the electron-
phonon scattering in determining transport behaviors at
very low temperature. To observe our prediction, lower
characteristic temperature Tt marking the crossover from
the electron-electron dominated regime to the electron-
phonon dominated one is required, such that the inter-
mediate range from Tt to the high-T equipartition regime
is wide enough. In experimentally accessible high-purity
Pt samples with residual resistivity as small as 10−3 --
10−2 µΩ cm [41, 42], Tt can be as low as 10K, and at
T = 20 K the phonon-induced ρ is nearly one order of
magnitude larger than that contributed by the electron-
electron scattering and the residual resistivity [42]. Be-
cause in high-purity Pt, the T -linear scaling of ρ emerges
at T (cid:38) 80 K [15], the suitable range for observing the
first-principles predicted T -dependence of σSH [Fig. 2(b)]
is 20 K (cid:46) T (cid:46) 80 K. Very recently experimentalists have
been developing new techniques, with which the spin cur-
rent is generated and detected in a single transition-metal
sample, thus avoiding all the complications associated
with the interfaces and shunting effect [7, 8]. The pre-
dicted effect is expected to be observed as the quality of
Pt samples in such measurements is improved.
We thank Ming Xie, Tianlei Chai and Haodi Liu for
helpful discussions. Q.N. is supported by DOE (DE-
FG03-02ER45958, Division of Materials Science and En-
gineering) on the transport formulation in this work.
C.X. is supported by NSF (EFMA-1641101) and Welch
Foundation (F-1255). S.A.Y. is supported by Singapore
Ministry of Education AcRF Tier 2 (MOE2017-T2-2-
108). Y.L. and Z.Y are supported by the National Natu-
ral Science Foundation of China (Grants No. 61604013,
No. 61774018, and No. 11734004), the Recruitment Pro-
gram of Global Youth Experts, and the Fundamental Re-
search Funds for the Central Universities (Grants No.
2016NT10 and No. 2018EYT03).
FIG. 2. Calculated longitudinal resistivity ρ (a) and spin Hall
conductivity σSH (b) of pure Pt as a function of temperature.
The black dashed line in panel (a) illustrates the linear de-
pendence.
For T (cid:38) 80 K, a linear-in-T ρ is obtained, and σSH is
approximately a constant of 1.6 × 105 /e (Ω m)−1 [10].
Below the equipartition regime, ρ deviates from the lin-
ear T -dependence [illustrated by the black dashed line
in Fig. 2(a)], concurrently the calculated σSH increases
with decreasing temperature. At T = 20 K, σSH reaches
2.3 × 105 /e (Ω m)−1. Compared to the value at 80
K, the T -variation of σSH is as large as 50%. This T -
dependence just begins when the temperature drops be-
∗ congxiao@utexas.edu
† zyuan@bnu.edu.cn
[1] J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back,
and T. Jungwirth, Rev. Mod. Phys. 87, 1213 (2015).
[2] A. Hoffmann, IEEE Trans. Magn. 49, 5172 (2013).
[3] T. Seki, Y. Hasegawa, S. Mitani, S. Takahashi, H. Ima-
mura, S. Maekawa, J. Nitta, and K. Takanashi, Nat.
Mater. 7, 125 (2008).
[4] O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E.W. Bauer,
S. D. Bader, and A. Hoffmann, Phys. Rev. Lett. 104,
046601 (2010).
[5] L. Liu, C.-F. Pai, Y. Li, H.W. Tseng, D. C. Ralph, and
R. A. Buhrman, Science 336, 555 (2012).
[6] E. Sagasta, Y. Omori, M. Isasa, M. Gradhand, L. E.
0.1110ρ (µΩ cm)20305070100200300Temperature (K)12σSH (105 h_/e Ω-1m-1)(a)(b)5
Hueso, Y. Niimi, Y. C. Otani, and F. Casanova, Phys.
Rev. B 94, 060412(R) (2016).
[7] C. Stamm, C. Murer, M. Berritta, J. Feng, M. Gabureac,
P. M. Oppeneer, and P. Gambardella, Phys. Rev. Lett.
119, 087203 (2017).
[8] C. Chen, D. Tian, H. Zhou, D. Hou, and X. Jin, Phys.
Rev. Lett. 122, 016804 (2019).
[9] T. Tanaka, H. Kontani, M. Naito, T. Naito, D. S. Hi-
rashima, K. Yamada, and J. Inoue, Phys. Rev. B 77,
165117 (2008).
[10] L. Wang, R. J. H. Wesselink, Y. Liu, Z. Yuan, K. Xia,
and P. J. Kelly, Phys. Rev. Lett. 116, 196602 (2016).
[11] Y. Sun, Y. Zhang, C. Felser, and B. Yan, Phys. Rev.
Lett. 117, 146403 (2016).
[12] L. Berger, Phys. Rev. B 2, 4559 (1970).
[13] A. Crepieux and P. Bruno, Phys. Rev. B 64, 014416
[27] For our purpose of showing the T -dependence of the spin
Hall conductivity induced by electron-phonon scattering,
the results for jsj
SH are sufficient. For complete-
ness, we present in Ref. [28] the other spin Hall contribu-
tion of order (Wkk(cid:48) )0 from the antisymmetric part of the
third-Born-order scattering rate. This contribution also
vanishes when j0
SH and jad
l = 0.
[28] See Supplemental Material for the transport formalism
for the spin Hall conductivities of order (Wkk(cid:48) )0 and cal-
culation details of the spin Hall conductivity in the k-
cubic Rashba model, as well as discussions on the low-
T limit in the presence of both electron-phonon and
electron-impurity scattering.
[29] O. V. Dimitrova, Phys. Rev. B 71, 245327 (2005).
[30] S. Y. Liu and X. L. Lei, Phys. Rev. B 72, 155314 (2005).
[31] J. Schliemann and D. Loss, Phys. Rev. B 71, 085308
(2001).
(2005).
[14] J. M. Ziman, Principles of the Theory of Solids (Cam-
[32] B. A. Bernevig and S. C. Zhang, Phys. Rev. Lett. 95,
bridge University Press, Cambridge, 1972).
016801 (2005).
[15] Y. Liu, Z. Yuan, R. J. H. Wesselink, A. A. Starikov,
M. van Schilfgaarde, and P. J. Kelly, Phys. Rev. B 91,
220405(R) (2015).
[16] S. K. Lyo and T. Holstein, Phys. Rev. Lett. 29, 423
[33] R. Moriya, K. Sawano, Y. Hoshi, S. Masubuchi, Y. Shi-
raki, A. Wild, C. Neumann, G. Abstreiter, D. Bougeard,
T. Koga, and T. Machida, Phys. Rev. Lett. 113, 086601
(2014).
(1972).
[17] H.-A. Engel, B. I. Halperin, and E. I. Rashba, Phys. Rev.
Lett. 95, 166605 (2005).
[18] W. K. Tse and S. Das Sarma, Phys. Rev. Lett. 96, 056601
(2006).
[19] E. M. Hankiewicz and G. Vignale, Phys. Rev. B 73,
115339 (2006).
[20] W. Kohn and J. M. Luttinger, Phys. Rev. 108, 590
(1957); J. M. Luttinger, Phys. Rev. 112, 739 (1958).
[21] N. A. Sinitsyn, J. Phys.: Condens. Matter 20, 023201
(2008).
[22] N. A. Sinitsyn, Q. Niu, and A. H. MacDonald, Phys. Rev.
B 73, 075318 (2006).
[23] C. Xiao, Z. Z. Du, and Q. Niu, arXiv: 1907.00577
[24] C. Xiao and Q. Niu, Phys. Rev. B 96, 045428 (2017).
[25] N. A. Sinitsyn, A. H. MacDonald, T. Jungwirth, V. K.
Dugaev, and J. Sinova, Phys. Rev. B 75, 045315 (2007).
[26] C. Xiao, Front. Phys. 13, 137202 (2018).
[34] H. Liu, E. Marcellina, A. R. Hamilton, and D. Culcer,
Phys. Rev. Lett. 121, 087701 (2018).
[35] J. M. Ziman, Electrons and Phonons (Clarendon, Oxford,
1960).
[36] A. A. Abrikosov, L. P. Gor'kov, and I. E. Dzyaloshin-
skii, Quantum Field Theoretical Methods in Statistical
Physics (Pergamon Press, New York, 1965).
[37] J. Rammer and H. Smith, Rev. Mod. Phys. 58, 323
(1986).
[38] I. A. Ado, I. A. Dmitriev, P. M. Ostrovsky, and M. Titov,
Phys. Rev. B 96, 235148 (2017).
[39] S. Murakami, Phys. Rev. B 69, 241202(R) (2004).
[40] Anton A. Starikov, Yi Liu, Zhe Yuan and Paul J. Kelly,
Phys. Rev. B 97, 214415 (2018).
[41] G. K. White and S. B. Woods, Phil. Trans. Roy. Soc.
(London) A251, 273 (1958).
[42] R. H. Freeman, F. J. Blatt, and J. Bass, Phys. Kondens.
Materie 9, 271 (1969).
|
1106.3697 | 1 | 1106 | 2011-06-19T00:12:16 | Electronic charge and spin density distribution in a quantum ring with spin-orbit and Coulomb interactions | [
"cond-mat.mes-hall"
] | Charge and spin density distributions are studied within a nano-ring structure endowed with Rashba and Dresselhaus spin orbit coupling (SOI). For a small number of interacting electrons, in the presence of an external magnetic field, the energy spectrum of the system is calculated through an exact numerical diagonalization procedure. The eigenstates thus determined are used to estimate the charge and spin densities around the ring. We find that when more than two electrons are considered, the charge-density deformations induced by SOI are dramatically flattened by the Coulomb repulsion, while the spin density ones are amplified. | cond-mat.mes-hall | cond-mat |
Electronic charge and spin density distribution in a quantum ring
with spin-orbit and Coulomb interactions
Csaba Daday,1, 2 Andrei Manolescu,1 D. C. Marinescu,3 and Vidar Gudmundsson2
1School of Science and Engineering, Reykjavik University, Menntavegur 1, IS-101 Reykjavik, Iceland
2Science Institute, University of Iceland, Dunhaga 3, IS-107 Reykjavik, Iceland
3Department of Physics and Astronomy, Clemson University, Clemson, SC 29621, USA
Charge and spin density distributions are studied within a nano-ring structure endowed with
Rashba and Dresselhaus spin orbit coupling (SOI). For a small number of interacting electrons,
in the presence of an external magnetic field, the energy spectrum of the system is calculated
through an exact numerical diagonalization procedure. The eigenstates thus determined are used to
estimate the charge and spin densities around the ring. We find that when more than two electrons
are considered, the charge-density deformations induced by SOI are dramatically flattened by the
Coulomb repulsion, while the spin density ones are amplified.
PACS numbers: 71.70.Ej, 73.21.Hb, 71.45.Lr
I.
INTRODUCTION
The possibility of controlling the flow of the electron
spins in semiconductor structures by external electric
means through spin-orbit interaction (SOI) has domi-
nated the recent past of spintronics research. This fun-
damental principle, first explored in the Datta-Das spin
transistor configuration,1 has been guiding a sustained
effort in understanding all the phenomenological impli-
cations of this interactions on systems of electrons. The
coupling between spin and orbital motion results either
from the two-dimensional confinement (Rashba)2 or from
the inversion asymmetry of the bulk crystal structure
(Dresselhaus).3 The usual expression of the the spin-orbit
Hamiltonian HSO retains only the linear terms in the
electron momentum p = (px, py) and is given by
HSO =
α
(σxpy − σypx) +
β
(σxpx − σypy) .
(1)
The Rashba and Dresselhaus coupling constants are α
and β, respectively, while σx,y,z are the Pauli matrices. In
general, the two interactions are simultaneously present
and often have comparable strengths. While α, the cou-
pling constant of the Rashba interaction, can be modi-
fied by external electric fields induced by external gates,
the strength of the Dresselhaus SOI, β, is fixed by the
crystal structure and by the thickness of the quasi two-
dimensional electron system.4,5 In many situations of in-
terest, an additional energy scale is introduced by the
Zeeman interaction of the electron spin with an external
magnetic field, proportional to the effective gyromagnetic
factor, g∗, which depends on the material energy-band
structure. While g∗ = −0.44 is very small in GaAs, it
can be more that 100 times larger in InSb.
The interplay between the two types of SOI, which
have competing effects on the precession of the electron
spin as they rotate it in opposite directions, and the
Zeeman splitting, which minimizes the energy by align-
ing the spin parallel to the external field, determines
the ground state polarization of the electron system and
the characteristics of spin transport. The investigation
of such problems in mesoscopic rings has been pursued
intensively by several authors.6 -- 10 In particular, it was
noticed that,
in the absence of the Coulomb interac-
tion among the electrons, the interference between the
Rashba and Dresselhaus precessions relative to the or-
bital motion, leads to the creation of an inhomogeneous
spin and charge distribution around the ring.8 The charge
inhomogeneity created in this situation has a symmetric
structure with two maxima and two minima around the
ring and will be called here a charge-density deformation
(CDD). The effect of the Coulomb interaction on this
type of charge distribution has been considered for two
electrons. It was obtained that, on account of the elec-
trostatic repulsion, the two electrons become even more
localized in the potential minima associated to the CDD,
leading to an amplitude increase.9,10
In this work we obtain an estimate of the effect of the
Coulomb interaction on the charge and spin distribution
associated with N = 2, 3 and 4 electrons in a ring with
SOI coupling by an exact diagonalization procedure that
uses the configuration interaction method. Our results
indicate that when the number of electrons increases, the
mutual repulsion leads to more uniform charge distribu-
tion around the ring, generating a dramatically flattened
CDD. In contrast, the spin-density deformation (SDD) is
amplified by the Coulomb effects. This can be explained
by the appearance of a stronger repulsion between same
spin electrons, leading to more favorable spin orienta-
tions.
II. THE RING MODEL
The system of
interest in our problem is a two-
dimensional quantum ring of exterior and interior radii,
Rext and Rint respectively. The ring is placed in a per-
pendicular magnetic field B associated in the symmetric
gauge with a vector potential A = B/2(−y, x, 0). The
single-particle Hamiltonian of an electron of momentum
2
In the same basis, the Zeeman Hamiltonian is simply
diagonal in the spatial coordinates,
hkjσHZ k′j′σ′i =
1
2
T tBγ(σz)σσ′ δkk′ δjj ′ ,
(4)
where γ = g∗m∗/(2me) is the ratio between the Zeeman
gap and the cyclotron energy, me being the free electron
mass.
For the spin-orbit Hamiltonian we obtain:
hkjσHSOk′j′σ′i =
1
2
T tα"tB
rk
4Rext
(σj
r)σσ′ δkk′ δjj ′
+ it1/2
ϕ
(σj
r + σj+1
r
2
)σσ′
δkk′ δjj ′ +1 − it1/2
r
(σj
ϕ)σσ′ δkk′ +1δjj ′#
+ T tβXk,j "σj
r → (σj
ϕ)∗ and σj
ϕ → −(σj
r)∗# + h.c. ,
(5)
where tα = α/(RextT ) and tβ = β/(RextT ) are the
two types of spin-orbit relative energies, while σr(ϕ) =
σx cos ϕ + σy sin ϕ and σϕ(ϕ) = −σx sin ϕ + σy cos ϕ are
the radial and angular Pauli matrices, respectively. We
used the slightly shorter notations σj
r = σr(ϕj ) and
σj
ϕ = σϕ(ϕj) for the matrices at the particular angles
on our lattice. The Rashba spin-orbit terms are all in-
cluded in the first square bracket. The Dresselhaus terms
are very similar to the Rashba ones, being given by the
substitutions indicated in the second square bracket.
The single particle states of the noninteracting Hamil-
tonian (2), Hψa = ǫaψa, are computed as eigenvalues
and eigenvectors of the matrices (3)-(5), ψa(rk, ϕj) =
Pσ Ψa,σ(k, j)σi. where Ψa,σ(k, j) are c−numbers.
In the basis provided by {ψa} the interacting many-
body Hamiltonian is written in the second quantization
as
ǫac†
aca +
H =Xa
1
2 Xabcd
Vabcdc†
ac†
bcdcc ,
(6)
where c†
a and ca are the creation and annihilation opera-
tors on the single-particle state a. The matrix elements
of the Coulomb potential V (r − r′) = e2/(κr − r′), κ be-
ing the dielectric constant of the material, are in general
give by
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
-0.4
-0.6
-0.8
t
x
e
R
/
y
-1.0
-1.0
-0.5
0.0
x/Rext
0.5
1.0
FIG. 1:
(Color online) The discretized ring with Rint =
0.8Rext, and 10 radial × 50 angular sites. The sites are shown
with circular points. The thin dotted lines connection sites
are for guiding the eye.
p = −i∇ + eA and effective mass m∗ is written as the
sum of an orbital term HO = p2/2m∗, a Zeeman contri-
bution HZ = (1/2)g∗µBBσz and the spin-orbit coupling
given in Eq. (1). The ensuing expression,
H = HO + HZ + HSO ,
(2)
is discretized in a standard manner6,7,11 on a grid12
defined by Nr radial and Nϕ angular sites, as shown
in Fig. 1. The radial coordinate of each site is rk =
Rext − (k − 1)δr, with k = 1, 2, ..., Nr, while δr =
(Rext − Rint)/(Nr − 1) is the distance between adja-
cent sites with the same angle. Similarly, the angular
coordinate is ϕj = (j − 1)δϕ, where j = 1, 2, ..., Nϕ and
δϕ = 2π/Nϕ is the angle between consecutive sites with
the same radius. The Hilbert space is spanned by the
ket-vectors kjσi, where the first integer, k, stands for
the radial coordinate, the second one, j, for the angular
coordinate, and σ = ±1 denotes the spin projection in
the z direction.
In this basis {kjσi}, the matrix elements of the orbital
Hamiltonian are given by:
hkjσHOk′j′σ′i =
T δσσ′("tϕ + tr +
−(cid:20)tϕ + tB
i
1
2
t2
B(cid:18) rk
4Rext(cid:19)2# δkk′ δjj ′
4δϕ(cid:21) δkk′ δjj ′ +1 + trδkk′ +1δjj ′) + h.c. .
(3)
Vabcd = hψa(r)ψb(r′)V(r − r′)ψc(r)ψd(r′)i .
(7)
In the present discrete model the double scalar product
is in fact a double summation over all the lattice sites
and spin labels
T = 2/(2m∗R2
ext) is the energy unit, while Rext is the
length unit. In T units, we obtain tϕ = [Rext/(rkδϕ)]2
the angular hopping energy, tr = (Rext/δr)2 the radial
hopping energy, and tB = eB/(m∗T ) the magnetic cy-
clotron energy. (h.c. denotes the Hermitian conjugate.)
Vabcd = T tC Xkjσ
k′j ′σ′
Ψ∗
a,σ(kj)Ψ∗
b,σ′ (k′j′)
Rext
rjk − rj ′k′
(8)
× Ψc,σ(kj)Ψd,σ′(k′j′) .
(a)
(b)
(c)
T
/
y
g
r
e
n
E
174
172
170
168
166
164
1
z
S
0
-1
2
1
0
0
1
2
3
4
3
GS
ES1
ES2
ES3
6
7
8
9 10
5
tB
(9)
0
0
0
1
x
c
∆
The new energy parameter introduced by the Coulomb
repulsion is tC = e2/(κRextT ). In the above summation
over the sites, the contact terms (k = k′, j = j′) are
avoided, as their contribution vanishes in the continuous
limit.
The many-body states Φµ are found by solving the
eigenvalue problem for the Hamiltonian (6),
HΦµ = EµΦµ .
A potential solution of the equation is written in the con-
figuration interaction representation13 -- 15 as a linear com-
bination of the non-interacting system eigenstates (Slater
determinants),
cµααi ,
Φµ =Xα
1 , iα
2 , ..., iα
Ki} where iα
havePa iα
with {αi = iα
a = 0, 1 is the occu-
pation number of the single-particle state ψa and K is
the number of single-particle states considered. The oc-
cupation numbers iα
K are listed in the increasing energy
order, so ǫK is the highest energy of the single-particle
state included in the many-body basis. For any αi we
a = N , which is the number of electrons in the
ring. It is straightforward to derive the matrix elements
of Hαα′ using the action of the creation and annihila-
tion operators on the many-body basis. In practice Eq.
(9) is convergent with K for a sufficiently small number
of electrons, and sufficiently small ratio of Coulomb to
confinement energy, tC . This procedure, also known as
"exact diagonalization", does not rely on any mean field
description of the Coulomb effects, like Hartree, Hartree-
Fock, or DFT.16
To be able to carry the numerical calculations in a rea-
sonable amount of time, we choose a small ring of radii
Rext = 50 nm and Rint = 0.8Rext, containing N ≤ 4
electrons. The discretization grid has 10 radial and 50 an-
gular points (500 sites), as shown in Fig. 1. Two common
semiconductor materials used in the experimental spin-
tronics are used for the selection of the material constants
needed: InAs with m∗ = 0.023me, g∗ = −14.9, κ = 14.6,
and estimated (or possible) values for the spin-orbit in-
teractions α ≈ 20 and β ≈ 3 meVnm; InSb with m∗ =
0.014me, g∗ = −51.6, κ = 17.9, and α ≈ 50, β ≈ 30
meVnm.4,5 The relative energies which we defined are:
for InAs tα = 0.60, tβ = 0.09, tC = 2.9, γ = −0.17; for
InSb tα = 0.92, tβ = 0.55, tC = 1.5, γ = −0.36.
In
our calculations we have considered material parameters
somewhere in between these two sets: tα = 0.7, tβ =
0.3, tC = 2.2, γ = −0.2
III. RESULTS
A. Single particle calculations
In the absence of the SOI, (α = β = 0), the single
particle Hamiltonian (2) shares its eigenstates ψa with
FIG. 2: (Color online) (a) The lowest 12 energies of the single
particle states vs. the magnetic energy tB. The solid (red)
and the dashed (green) lines show the states with positive
and negative parity, respectively. (b) The expected value of
the spin projection along the z direction Sz, in units of /2,
for the first four states on the energy scale. (c) The standard
deviation ∆c of the charge distribution around the circle with
radial site index k = 6, for the first four energy states: ground
state (GS), first, second, and third exited states (ES1, ES2,
ES3). The same association of line types with states is used
in panel (b).
the z components of the angular momentum Lz and spin
Sz = σz/2. In the presence of only one type of SOI, ei-
ther α 6= 0 or β 6= 0, the Hamiltonian commutes with the
z component of the total angular momentum, Lz + Sz,
which is conserved. When both α 6= 0 and β 6= 0,
the angular momentum is no longer conserved. How-
ever, ψa continue to be eigenstates of the parity operator
P = Πσz, Π being the (three dimensional) spatial in-
version operator.
Indeed, the general Hamiltonian (2)
commutes with P, which can be easily verified by using
Πp = −pΠ and the commutation rules of the Pauli ma-
trices. So in general Pψa = sψa, and thus the parity
s = ±1 of any state a is conserved, i. e. it is indepen-
dent on the magnetic field. In particular, when α = β
and g∗ = 0, all states become parity-degenerate at any
magnetic field.8,10,17 We identify the parity of the single
particle states calculated on our discrete ring model by
looking at the relation ψa,σ(k, j) = sσψa,σ(k, ¯)) where
(k, j) and (k, ¯) are diametrically opposed sites, with an-
gular coordinates ϕ¯ = ϕj + π.
In Fig. 2 we show the single particle states energy for
0 < tB < 10 (units of T ), which corresponds to a mag-
netic field strength between 0 and 1.32 Tesla. Further
increment of the magnetic field requires an augmentation
of the number of sites on the ring in order to maintain the
discrete model as a reasonable approximation of a physi-
cally continuous ring. At zero magnetic field all states are
parity degenerate, which is just the ordinary spin degen-
eracy. The degeneracy is in general lifted by a finite mag-
netic field. There are, however, some particular values of
tB where the degeneracy persist. This situation is repre-
sented in Fig. 2(a) by all intersection points of two lines
corresponding to the two possible parities. Such intersec-
tions do not occur between states with the same parity.
Due to the spin-orbit coupling, the orbital momentum
of one state depends on the spin of the other state and
vice versa, and, consequently, states of same parity do
in fact interact and thus avoid intersections.10 Although
in Fig. 2(a) many states represented by the same line
type apparently cross each other, in reality there are al-
ways tiny gaps between them, similar to those visible at
tB ≈ 2 between the first and the second excited states
or at tB ≈ 5.5 between the first, second, and third ex-
cited states. The magnitude of these gaps depends on the
g−factor, reducing in size for a smaller g∗ parameter.
In Fig. 2(b) the evolution of the expected spin in the
z direction, Sz = hψaσzψai/2 is presented for the first
four states in the energy order. One can see how the
spin flips for states avoiding the crossing, like those with
negative parity at tB ≈ 2 (dashed and dotted lines in
Fig. 2(b)).
Only one type of SOI, either Rashba or Dresselhaus,
is sufficient to avoid the crossing of states with the same
parity, but in this case the charge and the spin densi-
ties are uniform around the ring. When both SOI types
are present the charge and spin densities become nonuni-
form. This situation is equivalent with the presence of
a potential with two maxima and two minima around
the ring, having reflection symmetry relative to the axes
y = x (or ϕ = π/4, corresponding to the crystal direction
[110]) and y = −x (or ϕ = −π/4, corresponding to the
crystal direction [1¯10]).8,10 The amplitude of the CDD
is illustrated in Fig. 2(c) where the standard deviation
(in the statistical sense) ∆c of the charge density cal-
culated around one circle on the ring, close to the mean
radius, with radial site index k = 6, is plotted for the low-
est four energy states. The density deformation occurs
on account of the two combined SOI types which lead
to spin interference and additional interaction between
states with the same parity. Consequently, the ampli-
tude of the CDD for a certain state is maximum at those
magnetic fields where the parity degeneracy is lifted (the
state avoids a crossing with another state of the same
parity). In Fig. 2(c) this is clearly seen at tB ≈ 2, for the
excited state. In this example the CDD in the ground
state is very weak. The sharp peak at tB ≈ 0.4 indicates
the existence of a narrow gap between the 4-th and 5-th
energy states that avoid crossing.
4
T
/
y
g
r
e
n
E
331
330
329
328
327
494
493
492
491
659
658
657
656
655
333
332
331
330
329
501
500
499
498
673
672
671
670
669
(a) N=2 tC=0
(c) N=3 tC=0
(e) N=4 tC=0
0
2
4
6
8
10
tB
(b) N=2 tC=2.2
(d) N=3 tC=2.2
(f) N=4 tC=2.2
0
2
4
tB
6
8
10
FIG. 3: (Color online) Energy spectra of the first 12 states
for N = 2, 3, and 4 electrons without Coulomb interaction,
tC = 0 in panels (a),(c),(e), and with Coulomb interaction,
tC = 2.2, in panels (b),(d),(f). The solid (red) and the dashed
(green) lines show the states with positive and negative parity,
respectively.
B. Many particle calculations
In the following considerations, we will include more
than one electron. In Fig. 3 we compare the energy spec-
tra for the first 12 states vs.
the magnetic energy for
N = 2, 3, and 4 electrons, without and with Coulomb
interaction. Since the Coulomb interaction is invariant
at spatial inversion (and independent on spin) the par-
ity s is also conserved in the many-body states. Spectra
drawn for tC = 0 and tC = 2.2 have similar features.
The interacting spectrum presents a shift to higher ener-
gies, on account of the additional Coulomb energy, and a
slight increase of the gaps at high magnetic fields. More-
over, the crossings and the anti-crossings (points where
the crossings were avoided) of the energy levels have a
tendency to shift slightly to higher magnetic fields. Sim-
ilarly, the gap between the ground state and the excited
states increases at high tB.
The total spin Sz for each of the first three energy
states is shown in Fig. 4. At zero magnetic field, for an
even number of electrons, here N = 2 or N = 4, the
ground state is non-degenerate and has total spin Sz =
0, i. e. the spin-up and spin-down states of individual
electrons compensate. When the field is applied, the first
spin flip in the interacting ground state, as well as the
spin saturation, occur at lower magnetic fields than in
the absence of the Coulomb repulsion. This is a result of
the mixing of spin states with the same parity produced
z
S
2
1
0
-1
-2
2
1
0
-1
3
2
1
0
-1
2
1
0
-1
GS
ES1
ES2
(a) N=2 tC=0
0 2 4 6 8 10 0 2 4 6 8 10
(b) N=2 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
-2
2
1
0
(c) N=3 tC=0
0 2 4 6 8 10 0 2 4 6 8 10
(d) N=3 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
-1
3
2
1
0
(e) N=4 tC=0
0 2 4 6 8 10 0 2 4 6 8 10
tB
-1
(f) N=4 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
tB
FIG. 4:
(Color online) The total spin projection in the z
direction, in units of /2, for the many body states with N =
2, 3, 4 electrons. Without interaction, i. e. tC = 0 in panels
(a),(c),(e), and with interaction, with tC = 2.2, in panels
(b),(d),(f). Only the first three states are shown here, the
ground state (GS), the 1-st excited states (ES1), and the 2-nd
excited state (ES2). The magnetic energy tB varies between
0 and 10 and the lines showing the excited states are shifted
to the right, for clarity.
by the interaction. For N = 3 the ground state is spin
(double) degenerate at zero field. In the presence of the
Coulomb interaction, the total spin in the ground state
and the higher state is reversed.
Similar to the case of one electron (N = 1, Fig. 2), the
charge deformation of each state is maximized for those
magnetic fields where the state has an anti-crossing (or
repulsion) with another state of the same parity. The
charge deformation parameter ∆c is shown in Fig. 5. For
N = 2 the amplitude of the CDD increases with the
Coulomb interaction. There is a simple reason for that:
the potential associated with the charge deformation has
two minima diametrically opposite on the ring and each
of the two electrons tends be localized in one of these min-
ima. The mutual Coulomb repulsion fixes the electrons in
those places better.9,10 The situation changes, however,
for N > 2. The Coulomb forces spread the electrons
differently, more or less uniformly, such that the charge
deformation created by the SOI is drastically reduced. In
other words, the associated potential is strongly screened
even by one extra electron above N = 2. This effect can
be clearly seen in Fig. 5, comparing panels (c) with (d)
and (e) with (f). The vertical scale of panels (d) and (e)
has been magnified three times, for visibility.
0
0
0
1
x
c
∆
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
3.0
2.0
1.0
0.0
(a) N=2 tC=0
GS
ES1
ES2
0 2 4 6 8 10 0 2 4 6 8 10
(c) N=3 tC=0
0 2 4 6 8 10 0 2 4 6 8 10
(e) N=4 tC=0
0 2 4 6 8 10 0 2 4 6 8 10
tB
3.0
2.0
1.0
0.0
0.8
0.6
0.4
0.2
0.0
0.8
0.6
0.4
0.2
0.0
5
(b) N=2 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
(d) N=3 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
(f) N=4 tC=2.2
0 2 4 6 8 10 0 2 4 6 8 10
tB
FIG. 5:
(Color online) The standard deviation ∆c of the
charge on the circle k = 6 around the ring, as a measure of the
amplitude of the charge deformation. Shown are the results
for the ground state (GS), the 1-st excited states (ES1), and
the 2-nd excited state (ES2), for N = 2, 3, 4 electrons with-
out without (tC = 0), and with interaction (tC = 2.2). The
amplitude of the CDD's is strongly reduced by the Coulomb
effects for N = 3, 4; notice the different scales used in the
paired panels (c),(d) and (e),(f). The magnetic energy tB
varies between 0 and 10 and the lines corresponding to the
excited states are shifted to the right, for clarity.
only generates the specific effective potential which de-
termines the CDD. In the absence of SOI (α = β = 0),
we checked that a similar screening effect occurs in the
presence of a potential that induces a charge deformation
comparable to that obtained with the SOI. It is, however,
surprising that by adding only one extra electron such a
drastic effect ensues.
Next, we investigate the effect of the Coulomb inter-
action on the spin distribution around the ring. The
standard deviation of the spin density projected along
the z direction, ∆z, is plotted in Fig. 6 where we show
the results calculated as before for the circle correspond-
ing to the sites with radial coordinate k = 6. The spin
density deformation (SDD) is actually amplified by the
Coulomb interaction for all N = 2, 3, 4. As the CDD's,
the SDD's reach their maximum at the magnetic fields
where level repulsion occurs and remains prominent even
when the gaps are very small. The Coulomb enhance-
ment is a result of the mixing of states with the same
parity, but with different spin orientation produced by
the Coulomb potential. Consequently the SDD's have in
general a richer structure than the CDD's.
In principle, the screening of the charge deformation
is not particularly related to the spin-orbit effects. SOI
Finally, in Fig. 7 we display an example of CDD and
SDD, obtained for N = 4 interacting particles. The
6
charge and spin distributions corresponding to the sec-
ond excited state and tB = 4.5 are illustrated in Figs.
5(f) and 6(f), respectively. The CDD is weak, but still it
has four visible maxima. For two electrons the CDD has
only two maxima which are along the directions x = y
or x = −y, depending on the state and on the magnetic
field, both with and without the Coulomb interaction. In
particular, for a strictly one-dimensional ring model and
N = 2, in the ground state, the maxima are always along
the line x = −y,8,9 whereas for a two-dimensional model
they can also be along x = y.10 But in general, for N > 2
electrons, screening effects may distribute the charge in
more complicated configurations. Similar profiles with
multiple local oscillation may be obtained for the spin
density, eventually becoming spin-density waves around
the ring.
IV. CONCLUSIONS
We calculated the many-body states of a system of
N = 2, 3, and 4 interacting electrons located in a ring of
finite width with Rashba and Dresselhaus spin-orbit cou-
pling, in the presence of a magnetic field perpendicular on
the surface of the ring. The Coulomb effects are fully in-
cluded in the calculation via the "exact diagonalization"
method. We obtained inhomogeneous charge densities,
or CDD's, around the ring due to the combined effect of
the two types of SOI. When the Coulomb interaction is
included the charge deformation is amplified for N = 2,
as also shown by other authors.9,10 For N > 2 we find
that the CDD is dramatically flattened out in the pres-
ence of the Coulomb interaction. We interpret the result
as a screening effect. On the contrary, the spin inhomo-
geneities, or SDD's, are amplified by Coulomb effects for
all N > 1.
2
(a) N=2 tC=0
GS
ES1
ES2
2
(b) N=2 tC=2.2
1
1
0
0
0
1
x
z
∆
0
2
1
0
2
1
0
0 2 4 6 8 10 0 2 4 6 8 10
(c) N=3 tC=0
0
2
0 2 4 6 8 10 0 2 4 6 8 10
(d) N=3 tC=2.2
1
0 2 4 6 8 10 0 2 4 6 8 10
(e) N=4 tC=0
0
2
0 2 4 6 8 10 0 2 4 6 8 10
(f) N=4 tC=2.2
1
0 2 4 6 8 10 0 2 4 6 8 10
tB
0
0 2 4 6 8 10 0 2 4 6 8 10
tB
FIG. 6: (Color online) The standard deviation ∆z of the spin
projection along the z direction on the circle k = 6 around
the ring, as a measure of the amplitude of the spin-density
wave. Like in the previous figures GS, ES1, and ES2 in the
legend indicate the ground state, the 1-st excited states, and
the 2-nd excited states, respectively. The results are shown for
N = 2, 3, 4 electrons without (tC = 0), and with interaction
(tC = 2.2). Unlike the CDD's, the SDD's are amplified by
the Coulomb interactions for all N . The magnetic energy tB
varies between 0 and 10 and the lines corresponding to the
excited states are shifted to the right, for clarity.
(a)
(b)
y
t
i
s
n
e
d
e
g
r
a
h
C
-1.0 -0.5 0.0
x/Rext
0.5
1.0
1.0
0.5
y/Rext
-1.0
0.0
-0.5
FIG. 7: (Color online) (a) The charge density for N = 4
electrons with interaction (tC = 2.2), in the second excited
state, i. e. ES2 in Fig.5(f), with magnetic energy energy tB =
4.5. (b) The corresponding total spin distribution along the
ring k = 6 where the standard deviation of the z component
is calculated and shown in Fig.6(f).
Acknowledgments
This work was supported by the Icelandic Research
Fund. Valuable discussions with Sigurdur Erlingsson,
Gunnar Thorgilsson, and Marian Nit¸a are cordially ac-
knowledged.
1 S. Datta and B. Das, Appl. Phys. Lett. 56 (1990).
2 Y. Bychkov and E. I. Rashba, JETP Lett. 39, 78 (1984).
3 G. Dresselhaus, Phys. Rev. 100, 580 (1955).
4 R. Winkler, Spin orbit coupling effects in two-dimensional
electron and hole systems (Springer-Verlag Berlin, Heidel-
berg, New York, 2003).
5 T. Ihn, Semiconductor nanostructures. Quantum states
and electronic transport (Oxford University Press, 2010).
6 J. Splettstoesser, M. Governale, and U. Zulicke, Phys. Rev.
B 68, 165341 (2003).
12 The
first
and
function
second
f (x),
generic
f ′(x)
[f (x + h) + f (x − h) − 2f (x)] /h2,
h is considered sufficiently small.
≈
are
[f (x + h) − f (x − h)] /h and
7
any
as
≈
derivatives
of
approximated
f ′′(x)
respectively, where
13 P. Hawrylak and D. Pfannkuche, Phys. Rev. Lett. 70, 485
(1993).
14 N. T. T. Nguyen and F. M. Peeters, Phys. Rev. B 83,
7 S. Souma and B. K. Nikoli´c, Phys. Rev. B 70, 195346
075419 (2011).
(2004).
8 J. S. Sheng and K. Chang, Phys. Rev. B 74, 235315 (2006).
9 Y. Liu, F. Cheng, X. J. Li, F. M. Peeters, and K. Chang,
15 N. T. T. Nguyen and F. M. Peeters, Phys. Rev. B 78,
045321 (2008).
16 D. Pfannkuche, V. Gudmundsson, and P. Maksym, Phys.
Phys. Rev. B 82, 045312 (2010).
Rev. B 47, 2244 (1993).
10 M. P. Nowak and B. Szafran, Phys. Rev. B 80, 195319
17 J. Schliemann, J. C. Egues, and D. Loss, Phys. Rev. Lett.
(2009).
11 F. E. Meijer, A. F. Morpurgo, and T. M. Klapwijk, Phys.
Rev. B 66, 033107 (2002).
90, 146801 (2003).
|
1601.07524 | 2 | 1601 | 2016-06-14T11:58:55 | Surface plasmon polaritons in topological Weyl semimetals | [
"cond-mat.mes-hall",
"cond-mat.str-el"
] | We consider theoretically surface plasmon polaritons in Weyl semimetals. These materials contain pairs of band touching points - Weyl nodes - with a chiral topological charge, which induces an optical anisotropy and anomalous transport through the chiral anomaly. We show that these effects, which are not present in ordinary metals, have a direct fundamental manifestation in the surface plasmon dispersion. The retarded Weyl surface plasmon dispersion depends on the separation of the Weyl nodes in energy and momentum space. For Weyl semimetals with broken time-reversal symmetry, the distance between the nodes acts as an effective applied magnetic field in momentum space, and the Weyl surface plasmon polariton dispersion is strikingly similar to magnetoplasmons in ordinary metals. In particular, this implies the existence of nonreciprocal surface modes. In addition, we obtain the nonretarded Weyl magnetoplasmon modes, which acquire an additional longitudinal magnetic-field dependence. These predicted surface plasmon results are observable manifestations of the chiral anomaly in Weyl semimetals and might have technological applications. | cond-mat.mes-hall | cond-mat | a
Surface plasmon polaritons in topological Weyl semimetals
Johannes Hofmann1, 2 and Sankar Das Sarma1
1Condensed Matter Theory Center and Joint Quantum Institute, Department of Physics,
University of Maryland, College Park, Maryland 20742-4111 USA
2T.C.M. Group, Cavendish Laboratory, University of Cambridge, Cambridge CB3 0HE, United Kingdom
(Dated: June 15, 2016)
We consider theoretically surface plasmon polaritons in Weyl semimetals. These materials contain
pairs of band touching points – Weyl nodes – with a chiral topological charge, which induces an
optical anisotropy and anomalous transport through the chiral anomaly. We show that these effects,
which are not present in ordinary metals, have a direct fundamental manifestation in the surface
plasmon dispersion. The retarded Weyl surface plasmon dispersion depends on the separation of
the Weyl nodes in energy and momentum space. For Weyl semimetals with broken time-reversal
symmetry, the distance between the nodes acts as an effective applied magnetic field in momentum
space, and the Weyl surface plasmon polariton dispersion is strikingly similar to magnetoplasmons
in ordinary metals.
In
addition, we obtain the nonretarded Weyl magnetoplasmon modes, which acquire an additional
longitudinal magnetic-field dependence. These predicted surface plasmon results are observable
manifestations of the chiral anomaly in Weyl semimetals and might have technological applications.
In particular, this implies the existence of nonreciprocal surface modes.
PACS numbers: 73.20.Mf, 78.68.+m, 71.20.Gj, 03.65.Vf
Surface plasmon polaritons (SPPs) are collective elec-
tromagnetic and electron-charge excitations that are con-
fined to the surface of a metal or semiconductor. They
were proposed in the 1950's [1, 2] and have been ob-
served via electron energy loss spectroscopy [3, 4] as well
as optically via surface gratings [5] or attenuated total
reflection [6]. Over the past decades, SPPs have found
widespread technological applications, for example,
in
surface microscopy [7], for biomolecular detection [8],
or lithography [9]. Because SPPs are focused to sizes
smaller than the wavelength of light, they hold promise
to realize miniaturized plasmon-based optoelectronic de-
vices, and research in creating such plasmonic devices
is flourishing [10], with the subject being dubbed "plas-
monics" or "nano plasmonics," which is a huge applied
physics field in its own right.
In this Rapid Communication, we add a fundamen-
tal physical aspect to the study of SPPs (and the field
of plasmonics), and demonstrate that the surface plas-
mon polaritons of recently discovered Weyl semimetals
(WSMs), which possess topological properties, show a
much richer (and unanticipated) structure compared to
standard SPPs in ordinary metals and semiconductors.
We find that due to the quantum anomalous electrody-
namic response of the WSM (which is their hallmark),
the retarded Weyl surface plasmon is strongly sensitive
to details of the band structure. In particular, we find
a geometry in which the SPP is nonreciprocal (i.e., the
propagation is unidirectional), even without an applied
external magnetic field. In addition, we show that the
magnetoplasmon mode displays an additional longitudi-
nal field dependence which is absent in ordinary metals.
This can serve as a direct signature of Weyl semimetals in
surface measurements. We note that the SPP physics in-
troduced in this work applies to extrinsic or doped WSM
materials with no requirement of fine-tuning the chem-
ical potential to the band touching points, making our
predictions easy to test experimentally.
An important aspect of SPPs for technological appli-
cations is their nonreciprocity, i.e., the SPPs can only
propagate in one direction [10].
In conventional met-
als, nonreciprocal modes are only possible by breaking
time-reversal symmetry in an applied external magnetic
field [49, 52]. This comes with great technological chal-
lenges since for a sizable nonreciprocity, these magnetic
fields have to be very large [10]. In this Rapid Commu-
nication, we report nonreciprocal SPPs in the pristine
WSMs that are induced by topological Weyl node sepa-
ration without any external magnetic field. This provides
an alternative route to nonreciprocal modes and could
point to interesting technological applications of WSMs
in nanoplasmonics.
Weyl semimetals contain a valence and conduction
band that touch in isolated points of the Brillouin zone
near the chemical potential µ. The minimal Hamiltonian
in the vicinity of such a Weyl node is [11, 12]
H = χvp · σ − µ,
(1)
where χ = ± is the chirality, v the Fermi velocity, p
the momentum, and σ are Pauli matrices. We consider
the generic case of an extrinsic (doped) semimetal with
positive chemical potential µ > 0 (the "Weyl metal,"
although we continue referring to them as WSM). The
spectrum of the Hamiltonian (1) is linear with disper-
sion εp = ±vp. Weyl nodes appear in pairs of oppo-
site chirality [13–15], and they can be separated by a
wave vector b in the first Brillouin zone or by an en-
ergy offset b0 in energy. The topological properties of a
Weyl semimetal are manifested in the form of a θ-term
contribution to the action Sθ = e2
4πc
(cid:82) dt(cid:82) d3r θ E · B
with θ = 2(b · r − b0t) [16–19], where e is the electron
charge, E the electric field and B the magnetic field. If
the bands are degenerate with b0 = b = 0 (the so-called
Dirac semimetal), the system does not possess topologi-
cal properties. The θ term changes the electromagnetic
response of the material in the bulk medium by altering
the constitutive relation that links the displacement field
and the electric field [20–26], which in frequency space
reads
ε∞ +
4πi
ω
σ
E +
ie2
πω
(∇θ) × E +
ie2
πcω
θ B, (2)
(cid:18)
D =
(cid:19)
where ε∞ is the static dielectric constant of the medium
and σ the conductivity. The first term in parentheses
is the standard term as in normal metals, and the last
two terms arise due to the chiral anomaly. The gradi-
ent term in θ describes the contribution of an anoma-
lous Hall current, and the last time-derivative term de-
scribes the chiral magnetic effect [19]. Weyl semimetals
have recently been reported for TaAs [27, 28], NbAs [29],
YbMnBi2 [30], and Eu2Ir2O7 [31]. In addition, semimet-
als with degenerate bands (Dirac semimetals) are re-
ported for Cd3As2 [32–34], ZrTe5 [35], and Na3Bi [36].
They are parent materials from which nontrivial topo-
logical behavior is induced by symmetry breaking, for
example, by applying an external magnetic field [35, 37–
39]. These experimental results, which do not necessar-
ily have an exclusive interpretation in terms of the chiral
anomaly [39], motivate the search for direct signatures
of Weyl semimetals that are of topological origin, i.e.,
effects that are not explained by the linear semimetal-
lic Dirac dispersion and hence are not found in Dirac
semimetals or small-gap semiconductors. We establish
that SPP carry distinctive observable features in Weyl
systems arising purely from their topological properties.
Here, we show that the topological properties of Weyl
semimetals affect the surface plasmon polariton disper-
sion. There are two main results of this work: First,
because a WSM is an optically anisotropic medium, the
surface plasmon dispersion depends on the Weyl node
separation. The effect is strongest in the retarded limit
(i.e., small wave vector), where the magnitude of the wave
vector is comparable to the bulk plasmon frequency (di-
vided by c, the velocity of light). We find that for a time-
reversal broken WSM without external magnetic field,
the SPP dispersion resembles retarded magnetoplasmon
modes in standard metals. In particular, for certain ori-
entations of the surface, we predict a nonreciprocal dis-
persion, i.e., a dispersion that depends on the sign of
the wave vector. As the second main result of this work,
we predict that the Weyl surface magnetoplasmon modes
possess an anomalous longitudinal magnetic field depen-
dence that is absent for standard metals. This effect is
caused by the anomalous magnetic field dependence of
the WSM longitudinal conductivity ("negative magne-
toresistance") [25, 40, 41], a direct consequence of the
2
chiral anomaly. In the remainder of this Rapid Commu-
nication, we derive both effects and discuss their exper-
imental implications. All the algebraic details are pro-
vided in the Supplemental Material.
Surface plasmon polaritons are solutions of Maxwell's
equations localized at the interface of two media. We con-
sider the following geometry: A Weyl semimetal fills the
positive half volume z > 0, and a vacuum for z < 0. The
WSM-vacuum interface lies in the xy plane. For simplic-
ity, we restrict the analysis to a single pair of Weyl nodes
(although our results are obviously valid for WSM with
arbitrary pairs of nodes). Since we have translational in-
variance along the interface, the SSP are parametrized
by the parallel wave vector q = (qx, qy). We search for
electric fields of the form
y, Ej
z)eiqxx+iqyye−iωte−κjz,
Ej = (Ej
x, Ej
(3)
which decay exponentially away from the boundary, i.e.,
for which Re κ > 0, and we label j = 0 on the vacuum
side and j > 0 enumerates the solutions in the WSM.
The decay constants κj are determined from a solution
of the wave equation
∇ × (∇ × E) = − 1
c2
∂2
∂t2 D,
(4)
z = D2
z and B1
z = B2
x/y = E2
x/y = B2
x/y and B1
iω∇ × E). Substituting the ansatz (3) in Eq. (4),
where on the vacuum side, we have D = E, and for the
WSM, D is given by Eq. (2) (the magnetic field in this
expression is related to the electric field by Faraday's law
B = c
we obtain a linear system of equations. The zeros of the
determinant of this system yield κj. In general, on the
WSM side, it turns out that there are two solutions of
Eq. (4) with exponentially decaying field. We demand
the continuity of the parallel components of electric and
magnetic fields (E1
x/y) and of
the perpendicular components of the displacement fields
(D1
z ). This gives four linearly inde-
pendent conditions that determine the surface plasmon
dispersion as well as the relative magnitude of the fields.
In the following, we assume that the dielectric ten-
sor does not depend on the wavelength or the posi-
tion inside the WSM. This approximation applies if the
inverse wave vector of the SPP is large compared to
the Thomas-Fermi length, which in a WSM is propor-
tional to the inverse Fermi wave vector [42, 43]. In this
case, the diagonal component of the dielectric tensor is
ε1(ω) = ε∞(1 − Ω2
3π ( µ )2 denotes the
bulk plasmon frequency [44–46] with α = e2/vε∞ being
the finestructure constant of the WSM.
ω2 ), where Ω2
p = 4α
p
We first discuss the results for the SPP of a Dirac
semimetal (for which b = 0 and b0 = 0). The SPP
solves
with κ0 = (cid:112)q2 − ω2/c2. This coincides with the con-
ε1(ω) + κ0 = 0,
(5)
ventional SPP condition in standard metals [1, 47]. The
3
FIG. 1. Surface plasmon dispersion of a Weyl semimetal with broken time-reversal symmetry for different values of (top to
bottom) (a) ωb/Ωp = 0.25, 0.5, 0.75, and 1; (b) 0.5, 1, and 1.5; and (c) 0.5. In (c), the continuous blue line denotes positive
wave numbers q > 0 and the dotted blue line q < 0. In all plots, the bulk plasmon dispersion is indicated by thin red lines.
The thin black lines denote the asymptotic light line and the nonretarded frequency, respectively, and the black dashed line
indicates the SPP dispersion of a standard Dirac material. As discussed in the main text, the black dots mark the points where
the SPP hybridizes with the bulk plasmon mode and is damped.
ω =(cid:112)ε∞/(ε∞ + 1)Ωp (horizontal thin black line), which
surface plasmon dispersion is indicated by a thick black
dashed line in Figs. 1 (a)-(c).
In the fully retarded
limit cq (cid:28) Ωp, the SPP follows the light-line ω = cq
[thin black line in Figs. 1(a)-1(c)] and turns over in
the hydrodynamic limit cq (cid:29) Ωp to a constant value
solves ε1(ω) = −1. In particular, for ε∞ = 1, this coin-
√
2 [1].
cides with the famous result by Ritchie, ω = Ωp/
However, these surface plasmon modes are different from
ordinary metals since they are purely quantum with
p ∼ αn2/3 ∼ n2/3/. Fur-
appearing explicitly [48]: Ω2
thermore, they show a sub-linear density dependence as
opposed to a linear density-dependence for Ω2
p in ordi-
nary metals. Electron interactions can introduce a loga-
rithmic correction to this scaling through charge renor-
malization [45]. We predict that the linear dispersion of a
Dirac semimetal is manifested in a nonlinear dependence
of the squared SPP mode frequency on doping density.
Note that the characteristic scaling behavior may not
only be probed by varying the doping density but also
by finite-temperature measurements [31, 45].
We now present results for WSMs with broken time-
reversal symmetry (b (cid:54)= 0) and broken parity (b0 (cid:54)= 0)
for ε∞ = 13 as measured in Eu2Ir2O7 [31]. The results
are shown in Fig. 1 for three relevant configurations: (a)
b perpendicular to the sample surface, where the surface
plasmon dispersion depends only on the magnitude of the
parallel wave vector q; (b) b parallel to the surface with
q parallel to b; and (c) b parallel to the surface with q
perpendicular to b. The chiral anomaly induces an off-
diagonal term in the dielectric tensor iε2(ω) = iε∞ωb/ω
with ωb = 2e2b/πε∞. All the analytical calculational
details are provided in the Supplemental Material. Fig-
ure 1 (a) shows the SPP mode as a function of wave vec-
tor for four values of ωb/Ωp = 0.5, 1, 1.5, and 2. The SPP
deviates from the Dirac semimetal result (black dashed
line) for intermediate wave vectors cq ∼ Ωp and departs
from the light line at smaller wave number and energy.
For comparison, we include as thin red lines the corre-
sponding bulk plasmon modes for which one of the decay
constants vanishes κ = 0 and the plasmon is no longer
confined to the surface. As is evident from the plots,
for some wave vectors, bulk and surface modes are de-
generate, while in other regions (marked by black end
points), the SPP vanishes. Here, a generalized SPP still
exists, but with a complex wave vector q, indicating a
coupling of surface and bulk modes [49]. It is interesting
to note that the geometry [Fig. 1(a)] shows signs of the
chiral anomaly, even though another characteristic signa-
ture of WSM – topological Fermi arc surface states – are
absent in this configuration. Similar features are seen
in Fig. 1(b), which is shown for three different values
ωb/Ωp = 0.5, 1, and 1.5. Both cases shown in Figs. 1(a)
and 1(b) are reciprocal, i.e., the dispersion is indepen-
dent of the sign of q. In case (c), however, the SPP dis-
persion is nonreciprocal. For positive q > 0 (blue dotted
line), there is a transition from the light line to an asymp-
totic nonretarded constant frequency. For negative q < 0,
the dispersion has a discontinuity as it merges with the
bulk plasmon mode, at which point it jumps to a higher
frequency. In particular, there exists a frequency range
where the system supports only modes with q < 0. The
nonreciprocity that we report could have interesting tech-
nological applications [50]. While nonreciprocal SSP in
normal metals require magnetic fields or impurities [51],
nonreciprocity is a fundamental intrinsic material prop-
erty of a WSM arising from its topological nature.
Strikingly, the SPP with b (cid:54)= 0 resemble, on a qual-
itative level, SPP of an ordinary metal in the presence
of an external magnetic field [49, 52, 53], even though
they have quite a different origin. Hence, the topolog-
ical contribution to the dielectric tensor, which stems
from an anomalous Hall displacement current, induces
an "anomalous surface magnetoplasmon." This is a cen-
tral result of our work.
The retarded SPP dispersion of a WSM with b0 (cid:54)= 0
(and b = 0) is shown in Fig. 2 for ωb0 /Ωp = 0.5. The
dispersion solves
(cid:20)
(cid:21)
(cid:18)
(cid:113) 1
(κ0 + κ1b)
κ1a
×
1b − q2 +
κ2
1a)
ω2
(cid:19)
c2 + κ0(q2 − κ2
ω2
c2 ε1
+ (κ1a ↔ κ1b) = 0,
(6)
2
ω2
ω4
c4 ε2
4 ω8ε4
2 + c2ω6ε1ε2
c4
/ω2, with ω2
b0
2 −
where the decay constants are κ1a/b = q2 − 1
c2 ε1± 1
2 and ε2 = 2e2b0Ωp/πcω2 =
= 2e2b0Ωp/πcε∞. There is no de-
ε∞ω2
b0
pendence on the direction of the parallel wave number
q. In addition to the changed SPP dispersion, another
observable effect would be a tilt of the field polarization
out of the sagittal plane, as suggested for 3D topological
insulators [54].
We now consider WSM surface magnetoplasmon
modes, which turn out to have an unusual magnetic field
dependence. As this effect is distinct from the zero-field
anomalous SPP discussed so far, we restrict our atten-
tion to the nonretarded limit and neglect corrections due
to the separation of the Weyl nodes. The dielectric ten-
sor takes the form ε1(ω) = ε∞ + 4πi
ω σ. The conductiv-
ity can be derived in a semiclassical framework, in which
the Berry curvature modifies the semiclassical equation of
motion [25, 40, 41, 55, 56]. In particular, the Berry curva-
ture induces a longitudinal magnetoconductivity [40, 41]
(cid:18)
(cid:19)
α
π
ω2
c
Ω2
p
σ(cid:107)(ω) =
iΩ2
p
4πω
1 +
,
(7)
pωc/4π(ω2 − ω2
whereas in an ordinary metal, there is only the first
Drude term and no dependence on the magnetic field.
Here, ωc = ev2B/µc denotes the cyclotron frequency and
Eq. (7) holds for frequencies ω (cid:29) τ−1, where τ is the
inelastic intranode scattering time [40]. Note that the
relative strength of the Drude and the anomalous term
depends on the magnetic field and the doping density.
The remaining components of the conductivity tensor
pω/4π(ω2 − ω2
take the standard Drude form σ⊥ = iΩ2
c )
and σxy = Ω2
c ), provided that Berry cur-
vature corrections to the density of states and the in-
trinsic orbital moment are neglected [25].
In the non-
retarded limit, the electric field is given by E = −∇φ,
where the electrostatic potential solves Poisson's equa-
tion ∇2φ = 4πρ and ρ is the induced charge density. On
the vacuum side, the potential solves ∇2φ = 0. Combin-
ing Poisson's equation with the current and the continu-
ity equation, we obtain (setting ε∞ = 1)
∇ · (σ∇φ) = 0.
∇2φ +
(8)
4πi
ω
4
FIG. 2. Surface plasmon dispersion of a Weyl semimetal with
broken inversion symmetry for ωb0 /Ωp = 0.5, and 1. The
notation is the same as in Fig. 1.
We make the ansatz φ = φieiqxx+iqyye−iωte−κjz and
impose the continuity of φ at the boundary. Integrating
Eq. (8) across the interface, we find the second boundary
condition
φ(cid:48)(0+) − φ(cid:48)(0−) +
[σ∇φ]z(0+) = 0,
4πi
ω
(9)
where the prime denotes a derivative with respect to z.
These conditions are of course equivalent to demanding
the continuity of the parallel components of E and the
perpendicular component of D.
For a magnetic field parallel to the surface, we obtain
the surface plasmon condition
(cid:16)
(cid:17)
ω2 − (Ω2
p − ω2 + ω2
α
c )
1 +
π
p − ω2) sin2 θ = 0,
−2ωωc sin θ − (Ω2
cos2 θ
ω2
c
Ω2
p
(10)
where θ is the relative angle between the magnetic field
and the parallel momentum q. The term proportional to
sin θ implies that the surface magnetoplasmon is non-
reciprocal,
i.e., the frequency depends on the sign of
qy. For a dispersion perpendicular to the magnetic field
(cos θ = 0), we find ω2 − ωωcsgn(q) − Ω2
p/2 = 0, which
is the same form as in ordinary metals [57]. There is a
modification of the plasmon dispersion for modes that
propagate along the magnetic field (sin θ = 0), for which
we find
(cid:0)Ω2
p + ω2
c
(cid:1).
ω2 =
1 + α
π
2 + α
π
ω2
c
Ω2
p
ω2
c
Ω2
p
(11)
The anomalous correction to σ(cid:107) changes the magnetic
field dependence of the plasmon frequency. Compar-
ing with the standard surface plasmon relation ω2 =
ε∞/(ε∞ +1)Ω2
p, we interpret the correction as an anoma-
lous contribution to the dielectric constant of the medium
p ∼ B2n−4/3, which acquires a magnetic
∆ε∞ = αω2
field dependence. If the anomalous term dominates (at
c /πΩ2
small α and doping), the plasmon mode is equal to the
bulk plasmon frequency, ω2 = Ω2
p. For a magnetic field
that is perpendicular to the surface, we obtain the same
anomalous surface plasmon dispersion as in Eq. (11).
In summary, we predict a rich (and experimentally
observable) structure of surface plasmon polaritons in
doped Weyl semimetals. In particular, we show the fol-
lowing: (a) There is a quantum surface plasmon mode
with unusual density dependence; (b) for broken inver-
sion symmetry, the retarded dispersion is strongly af-
fected by the chemical potential imbalance; (c) for bro-
ken time-reversal symmetry, the dispersion depends on
the Weyl node separation, which acts similar to an inter-
nal magnetic field; (d) a nonreciprocal dispersion arises
naturally even without external magnetic fields; and (e)
the magnetoplasmon mode acquires an additional longi-
tudinal magnetic field dependence. These effects are ex-
perimentally observable signatures of the chiral anomaly
in Weyl semimetals and could point the way to future
technological applications of these systems.
This work is supported by LPS-MPO-CMTC, Mi-
crosoft Q, and JQI-NSF-PFC (J.H. and S.D.S.), and
Gonville and Caius College (J.H.).
[1] R. H. Ritchie, Phys. Rev. 106, 874 (1957).
[2] E. A. Stern and R. A. Ferrell, Phys. Rev. 120, 130 (1960).
[3] C. J. Powell and J. B. Swan, Phys. Rev. 115, 869 (1959).
[4] F. J. Garc´ıa de Abajo, Rev. Mod. Phys. 82, 209 (2010).
[5] R. H. Ritchie, E. T. Arakawa, J. J. Cowan, and R. N.
Hamm, Phys. Rev. Lett. 21, 1530 (1968).
[6] N. Marschall, B. Fischer, and H. J. Queisser, Phys. Rev.
Lett. 27, 95 (1971).
[7] B. Rothenhausler and W. Knoll, Nature 332, 615 (1988).
[8] M. Malmqvist, Nature 361, 186 (1993).
[9] W. Srituravanich, N. Fang, C. Sun, Q. Luo, and X. Zhang,
Nano Letters 4, 1085 (2004).
[10] W. L. Barnes, A. Dereux, and T. W. Ebbesen, Nature
424, 824 (2003).
[11] O. Vafek and A. Vishwanath, Annual Review of Con-
densed Matter Physics 5, 83 (2014).
[12] T. Wehling, A. Black-Schaffer,
and A. Balatsky, Ad-
vances in Physics 63, 1 (2014).
[13] H. Nielsen and M. Ninomiya, Nuclear Physics B 185, 20
(1981).
[14] H. Nielsen and M. Ninomiya, Nuclear Physics B 193, 173
(1981).
5
[22] P. Hosur and X.-L. Qi, Phys. Rev. B 91, 081106 (2015).
[23] M. Kargarian, M. Randeria, and N. Trivedi, Scientific
Reports 5, 12683 EP (2015).
[24] A. A. Zyuzin and V. A. Zyuzin, Phys. Rev. B 92, 115310
(2015).
[25] F. M. D. Pellegrino, M. I. Katsnelson, and M. Polini,
Phys. Rev. B 92, 201407 (2015).
[26] P. Goswami, G. Sharma, and S. Tewari, Phys. Rev. B
92, 161110 (2015).
[27] S.-Y. Xu,
I. Belopolski, N. Alidoust, M. Neupane,
G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-
C. Lee, S.-M. Huang, H. Zheng, J. Ma, D. S. Sanchez,
B. Wang, A. Bansil, F. Chou, P. P. Shibayev, H. Lin,
S. Jia, and M. Z. Hasan, Science 349, 613 (2015).
[28] B. Q. Lv, H. M. Weng, B. B. Fu, X. P. Wang, H. Miao,
J. Ma, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen,
Z. Fang, X. Dai, T. Qian, and H. Ding, Phys. Rev. X 5,
031013 (2015).
[29] S.-Y. Xu, N. Alidoust, I. Belopolski, Z. Yuan, G. Bian,
T.-R. Chang, H. Zheng, V. N. Strocov, D. S. Sanchez,
G. Chang, C. Zhang, D. Mou, Y. Wu, L. Huang, C.-C. Lee,
S.-M. Huang, B. Wang, A. Bansil, H.-T. Jeng, T. Neupert,
A. Kaminski, H. Lin, S. Jia, and M. Zahid Hasan, Nat
Phys 11, 748 (2015).
[30] S. Borisenko, D. Evtushinsky, Q. Gibson, A. Yaresko,
T. Kim, M. N. Ali, B. Buechner, M. Hoesch, and R. J.
Cava, arXiv:1507.04847 (2015).
[31] A. B. Sushkov, J. B. Hofmann, G. S. Jenkins, J. Ishikawa,
S. Nakatsuji, S. Das Sarma, and H. D. Drew, Phys. Rev.
B 92, 241108 (2015).
[32] S. Borisenko, Q. Gibson, D. Evtushinsky, V. Zabolotnyy,
B. Buchner, and R. J. Cava, Phys. Rev. Lett. 113, 027603
(2014).
[33] T. Liang, Q. Gibson, M. N. Ali, M. Liu, R. J. Cava, and
N. P. Ong, Nat Mater 14, 280 (2015).
[34] M. Neupane, S.-Y. Xu, R. Sankar, N. Alidoust, G. Bian,
C. Liu, I. Belopolski, T.-R. Chang, H.-T. Jeng, H. Lin,
A. Bansil, F. Chou, and M. Z. Hasan, Nat Commun 5,
3786 (2014).
[35] Q. Li, D. E. Kharzeev, C. Zhang, Y. Huang, I. Pletikosic,
A. V. Fedorov, R. D. Zhong, J. A. Schneeloch, G. D. Gu,
and T. Valla, arXiv:1412.6543 (2014).
[36] Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng,
D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai,
Z. Hussain, and Y. L. Chen, Science 343, 864 (2014).
[37] H.-J. Kim, K.-S. Kim, J.-F. Wang, M. Sasaki, N. Satoh,
A. Ohnishi, M. Kitaura, M. Yang, and L. Li, Phys. Rev.
Lett. 111, 246603 (2013).
[38] J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan,
W. Wang, R. J. Cava, and N. P. Ong, arXiv:1503.08179
(2015).
[39] P. Goswami, J. H. Pixley, and S. Das Sarma, Phys. Rev.
B 92, 075205 (2015).
[15] H. Nielsen and M. Ninomiya, Physics Letters B 130, 389
[40] D. T. Son and B. Z. Spivak, Phys. Rev. B 88, 104412
(1983).
(2013).
[16] K. Fujikawa, Phys. Rev. Lett. 42, 1195 (1979).
[17] A. A. Zyuzin and A. A. Burkov, Phys. Rev. B 86, 115133
[41] B. Z. Spivak and A. V. Andreev, arXiv:1510.01817
(2015).
(2012).
[42] S. Das Sarma, E. H. Hwang, and H. Min, Phys. Rev. B
[18] P. Goswami and S. Tewari, Phys. Rev. B 88, 245107
91, 035201 (2015).
(2013).
[19] P. Hosur and X. Qi, Comptes Rendus Physique 14, 857
(2013).
[20] F. Wilczek, Phys. Rev. Lett. 58, 1799 (1987).
[21] A. G. Grushin, Phys. Rev. D 86, 045001 (2012).
[43] S. Li and A. V. Andreev, Phys. Rev. B 92, 201107 (2015).
[44] M. Lv and S.-C. Zhang, International Journal of Modern
Physics B 27, 1350177 (2013).
[45] J. Hofmann and S. Das Sarma, Phys. Rev. B 91, 241108
(2015).
[46] J. Zhou, H.-R. Chang, and D. Xiao, Phys. Rev. B 91,
tions 14, 1223 (1974).
035114 (2015).
[47] J. M. Pitarke, V. M. Silkin, E. V. Chulkov, and P. M.
Echenique, Reports on Progress in Physics 70, 1 (2007).
[48] S. Das Sarma and E. H. Hwang, Phys. Rev. Lett. 102,
[52] K. Chiu and J. Quinn, Il Nuovo Cimento B 10, 1 (1972).
[53] M. S. Kushwaha, Surface Science Reports 41, 1 (2001).
[54] A. Karch, Phys. Rev. B 83, 245432 (2011).
[55] D. Xiao, M.-C. Chang, and Q. Niu, Rev. Mod. Phys. 82,
206412 (2009).
1959 (2010).
[49] R. F. Wallis, J. J. Brion, E. Burstein, and A. Hartstein,
[56] R. Lundgren, P. Laurell, and G. A. Fiete, Phys. Rev. B
Phys. Rev. B 9, 3424 (1974).
90, 165115 (2014).
[50] R. Camley, Surface Science Reports 7, 103 (1987).
[51] A. Hartstein and E. Burstein, Solid State Communica-
[57] A. L. Fetter, Phys. Rev. B 32, 7676 (1985).
6
Supplemental Material: Surface plasmon polaritons in topological Weyl semimetals
1Condensed Matter Theory Center and Joint Quantum Institute, Department of Physics, University of Maryland,
2T.C.M. Group, Cavendish Laboratory, University of Cambridge, Cambridge CB3 0HE, United Kingdom
College Park, Maryland 20742-4111 USA
Johannes Hofmann1,2 and Sankar Das Sarma1
1
In this supplemental material, we provide details on the calculation of the retarded surface plasmon polariton
dispersion. The interface between the vacuum and the Weyl semimetal (WSM) is in the x-y plane at z = 0, with the
vacuum for z < 0 and the WSM for z > 0. The evanescent ansatz (3) for the electric field at the boundary solves the
wave equation (4), where the displacement field D is related to the electric field E through the dielectric tensor ε(ω),
Eq. (2). Hence, the electric field solves a homogeneous equation M E = 0, where
q2
M =
y − κ2
−qxqy
∓iqxκj ∓iqyκj q2
j −qxqy ∓iqxκj
j ∓iqyκj
x + q2
y
x − κ2
q2
− ω2
c2 ε(ω),
(S1)
M =
−κ2
with the positive sign on the vacuum side and negative on the WSM side. The zeros of the determinant of M
determine the decay constant κj. On the vacuum side (j = 0), we find κ2
c2 for any orientation of q. On the
WSM side (j = 1), we have for the various geometries shown in Fig. 1:
0 = q2 − ω2
Case (a): b = (0, 0, b), q = (0, q, 0)
q2 − κ2
1 − ω2
iω2
c2 ε2(ω)
c2 ε1
0
−κ2
− iω2
c2 ε2
1 − ω2
−iqκ1
0
c2 ε1 −iqκ1
q2 − ω2
c2 ε1
,
1 = q2 − ω2
κ2
c2 ε1 ± 1
ε1
(cid:114)
− ω2
c2 ε2
2
(cid:16)
q2 − ω2
c2 ε1
(cid:17)
(S2)
Case (b): b = (b, 0, 0), q = (q, 0, 0)
M =
c2 ε1
1 − ω2
0
−iqκ1
0
q2 − κ2
1 − ω2
iω2
c2 ε2
−iqκ1
c2 ε1 − iω2
c2 ε2
q2 − ω2
c2 ε1
,
κ2
1 = q2 +
(cid:19)
− ε1
± 1
2ε1
(cid:114)
4q2 ω2
c2 ε1ε2
2 +
ω4
c4 ε4
2
(S3)
Case (c): b = (b, 0, 0), q = (0, q, 0)
1 − ω2
0
0
q2 − κ2
M =
c2 ε1
c2 ε1 −iqκ1 − iω2
q2 − ω2
c2 ε2
c2 ε1
where we define as in the main text ε1(ω) = ε∞(1− Ω2
ω2 ) and ε2(ω) = ε∞ωb/ω with ωb = 2e2b/πε∞. In Eqs. (S2)-(S4),
we also note the results for the decay constant κ1 on the WSM side. For b0 (cid:54)= 0 and q = (q, 0, 0) as shown in Fig. 2,
we have
κ2
1 = q2 +
c2 ε2
(S4)
2
ε1
,
p
− ε1
ω2
c2
0
−κ2
1 − ω2
−iqκ1 + iω2
0
(cid:18) ε2
(cid:19)
ω2
c2
2
2ε1
(cid:18) ε2
,
−κ2
M =
c2 ε1(ω)
1 − ω2
− cκ1
ε2
− icq
ε2
Ωp
Ωp
q2 − κ2
cκ1
Ωp
ε2
1 − ω2
c2 ε1(ω)
iω2
c2 ε2(ω)
−iqκ1
ε2
q2 − ω2
icq
Ωp
c2 ε1(ω)
1 = q2 − ω2
κ2
c2 ε1 − ω4
2c4 ε2
2 +
1
2
,
(cid:114)
ω8
c8 ε4
2 + 4
ω6
c6 ε1ε2
2,
(S5)
/ω2 with ω2
b0
= 2e2b0Ωp/πcε∞. There are, in general, two linearly independent
where we define ε2(ω) = ε∞ω2
b0
solutions for E on the vacuum and on the WSM side which correspond to modes that are localized at the interface. We
impose as a boundary condition the continuity of the parallel component of the electric field E and the perpendicular
component of the displacement field Dz as well as the continuity of the magnetic field B = c
set of four linearly independent constraints. Setting the determinant of this constraint matrix equal to zero, we find
the SPP condition for the various cases:
iω∇ × E. This gives a
(cid:18)
1a+κ1aκ1b+κ2
(cid:3)− ω2
1b)1+q2(cid:2)κ0(1−1)+(κ1a+κ1b)1
1b)1
c2 1
(κ1a+κ1b)(κ0+κ1a+κ1b+q2(1−1)
(cid:18)
κ1a+κ1b+κ0(1−1)
(cid:19)
(cid:3)− ω2
c2 1
= 0. (S7)
2
(cid:19)
= 0.
(S6)
Case (a):
1κ0κ1aκ1b(κ1a+κ1b)+q2(cid:2)κ1aκ1b+κ0(κ1a+κ1b)+(κ2
Case (b):
κ1aκ1b(κ0+κ1a+κ1b)+κ0(κ2
1a+κ1aκ1b+κ2
Case (c):
ε1κ1 + κ0(ε2
(S8)
Here, we denote by κ1a and κ1b the two solutions corresponding to ± in Eqs. (S2)-(S4). For b0 (cid:54)= 0, the condition
is stated in Eq. (6) of the main text. In the case b (cid:54)= 0, the result is similar as for the as for a metal in a constant
external magnetic field [S1, S2], and the SPP conditions agree when correcting for the difference in the dielectric
tensor. The explicit form of the dielectric components in a WSM is, of course, different from the Drude form in an
external magnetic field.
1 − ε2
2) − qε2 = 0.
[S1] K. Chiu and J. Quinn, Il Nuovo Cimento B 10, 1 (1972).
[S2] R. F. Wallis, J. J. Brion, E. Burstein, and A. Hartstein, Phys. Rev. B 9, 3424 (1974).
|
1208.0548 | 1 | 1208 | 2012-08-02T17:17:47 | Probing a Single Nuclear Spin in a Silicon Single Electron Transistor | [
"cond-mat.mes-hall"
] | We study single electron transport across a single Bi dopant in a Silicon Nanotransistor to assess how the strong hyperfine coupling with the Bi nuclear spin $I=9/2$ affects the transport characteristics of the device. In the sequential tunneling regime we find that at, temperatures in the range of $100 mK$, $dI/dV$ curves reflect the zero field hyperfine splitting as well as its evolution under an applied magnetic field. Our non-equilibrium quantum simulations show that nuclear spins can be partially polarized parallel or antiparallel to the electronic spin just tuning the applied bias. | cond-mat.mes-hall | cond-mat |
Probing a Single Nuclear Spin in a Silicon Single Electron Transistor
F. Delgado(1), R. Aguado(2), and J. Fern´andez-Rossier(1,3)1
(1) International Iberian Nanotechnology Laboratory (INL), Av. Mestre Jos´e Veiga, 4715-330 Braga,
Portugal
(2) Instituto de Ciencia de Materiales de Madrid (ICMM-CSIC), Cantoblanco, 28049 Madrid,
Spain
(3) Departamento de F´ısica Aplicada, Universidad de Alicante, 03690 San Vicente del Raspeig,
Spain
(Dated: 25 June 2018)
We study single electron transport across a single Bi dopant in a Silicon Nanotransistor to assess how the
strong hyperfine coupling with the Bi nuclear spin I = 9/2 affects the transport characteristics of the device.
In the sequential tunneling regime we find that at, temperatures in the range of 100mK, dI/dV curves reflect
the zero field hyperfine splitting as well as its evolution under an applied magnetic field. Our non-equilibrium
quantum simulations show that nuclear spins can be partially polarized parallel or antiparallel to the electronic
spin just tuning the applied bias.
PACS: 73.23.Hk, 31.30.Gs, 74.55.+v, 75.75.-c
The amazing progress both in the silicon process-
ing technologies and in the miniaturization of silicon
based transistors has reached the point where single-
dopant transistors have been demonstrated.1 -- 7 Whereas
this progress has been fueled by the development of clas-
sical computing architectures, it might also be used for
quantum computing. In this regard, the electronic and
nuclear spins of single donors in silicon are very promis-
ing building blocks for quantum computing.8 -- 10 Progress
along this direction makes it necessary to implement sin-
gle spin readout schemes both for electronic and nuclear
spins. Single electronic spin readout has been demon-
strated, both in GaAs quantum dots as well as in P doped
Silicon Nanotransistors.11,12
The readout of the quantum state of a single nuclear
spin, much more challenging, has been demonstrated for
NV centers in diamond taking advantage of single spin
optically detected magnetic resonance afforded by the ex-
traordinary properties of that system.13 Single nuclear
spin readout with either optical14 or a combined electro-
optical techniques15 has been proposed, but remains to
be implemented. Here we explore the electrical readout
of a single nuclear spin, more suitable for an indirect
band-gap host like Si. A preliminary step is to construct
a circuit whose transport is affected by the quantum state
of the nuclear spin. There is ample experimental evidence
of the mutual influence of many nuclear spins and trans-
port electrons in III-V semiconductor quantum dots in
the single electron transport regime.16 -- 19 In particular,
Kobayashi et al. have reported hysteresis in the dI/dV
upon application of magnetic fields, reflecting the real-
ization of different ensemble of nuclear states coupled to
the electronic spin via hyperfine coupling.19
Here we propose a device where a single nuclear spin
is probed in single electron transport. We model the
single electron transport in a silicon nanotransistor such
that, in the active region, transport takes place through
a single Bi dopant, see Fig. 1. We show that, at suffi-
ciently low temperatures, the dI/dV curves of this de-
vice probe the hyperfine structure of the dopant level. In
FIG. 1. (color online) a) Scheme of the Si:Bi FinFET nan-
otransistor. b) Trapping Coulomb potential of the Bi dopant
an single energy level participating in the transport.
turn, the occupations of the nuclear spin states are af-
fected by the transport electrons. Whereas single dopant
transistors have been demonstrated for single P, As and
in Si,3,4,6,12 we choose Bi because it has a much
B,
larger hyperfine splitting,20 -- 22 due to both a larger nu-
clear spin I = 9/2 and a larger hyperfine coupling con-
stant (A ≈ 6.1µeV). The zero-field splitting of the Bi
donor level is given by 5A and has been observed by
electron spin resonance20 -- 22 and in photoluminescence
experiments with many dopants.23
We consider the sequential transport regime, where the
occupation of the donor level fluctuates between q = 0
and q = 1. In the q = 0 state, the nuclear spin interacts
only with the external field.
In the q = 1 state, the
electron and the nuclear spin are hyperfine coupled. The
Hamiltonian that describes both states reads20 -- 22,24
H = q (cid:16)ǫd + eVG + A~S · ~I + ωeSz(cid:17) + ωN Iz,
(1)
where ǫd is the donor energy level with respect to the
Fermi energy, which we take as EF = 0, and VG denotes
an external gate voltage. We assume that valley degen-
eracies of the donor level are split-off and neglect the
valley degree of freedom. The third term is the hyper-
fine coupling, and the last two, where ωe = geµBBz
and ωN = gnµN Bz, correspond to the electron and
nuclear Zeeman terms, with ge (gn) the electron (nu-
2
q = 0 and q = 1 manifolds. The corresponding transition
rates are be calculated using the Fermi golden rule with
Htun as the perturbation:26
0 X
hM Iz(m), σi2 ,
(3)
Γη
m,M = Γη
σ
where Iz, σi ≡ Izi ⊗ σi. In the following we take the
applied bias convention µS − µD = eV , with µS = eV /2
and µD = −eV /2. For a given temperature, bias and
gate voltages and Hamiltonian parameters, we obtain the
steady state solution of the master equation, ignoring the
effect of the fast-decaying coherences. This yields the
steady state occupations Pm(V ) and PM (V ).
We consider the sequential tunneling regime, in which
the energy level broadening induced by coupling to the
electrodes Γ0 is small, Γ0 ≪ kBT . This also justifies the
markovian approximation implicit in the Bloch-Redfield
master equation. In this regime, current flows when the
bias enables charge fluctuations of the dopant level. The
steady state current corresponding to electrons flowing
from the source electrode to the dopant level is given by
I = e X
m,M
nPm(V )fS(∆M,m)ΓS
m,M
−PM (V ) [1 − fS(∆M,m)] ΓS
m,Mo,
(4)
where ∆M,m = ǫM − ǫm and fS(ǫ) = f (ǫ − µS) is the
Fermi function relative to the chemical potential of the S
electrode. The first term in the right hand-side of Eq.(4)
represents the electrons flowing from the S electrode to
the empty Bi, while the second one corresponds to elec-
trons flowing from the q = 1 Bi to the S electrode. In
steady state, the continuity equation ensures that cur-
rent between the dopant and the drain is the same than
the source-dopant current.
Figure 3a) shows the differential conductance dI/dV
map for zero-applied magnetic field, with I = I/(eΓ0)
and Γη
0 = Γ0/2. At zero bias, the conductance is zero
except at the special value of VG for which the addition
energy vanishes. Far from this point, the zero-bias charge
of the dopant state, hereafter denoted with q0, is either
q0 = 0 or q0 = 1. The finite bias conductance has a
peak whenever the bias energy, eV /2, matches the energy
difference between two states with different charge, m for
q = 0 and M for q = 1, that are permitted by the spin
selection rule implicit in Eq.(3). The height of the peak is
proportional to both the non-equilibrium occupations Pm
and PM and to the quantum mechanical matrix element
Γη
m,M . This determines the very different spectra when
the zero bias charge in the dopant is q = 0 or q = 1. The
width of the dI/dV peaks is proportional to kBT , so that
the dI/dV spectra can resolve the hyperfine structure
provided that kBT is smaller than the splitting of the
levels. The energy differences inside the F = 4 and F = 5
manifolds, see Fig. 2), are roughly proportional to A.
Thus, while the zero-field splitting can be resolved at
T = 0.3 K, temperature must be significantly below 50
mK to resolve the finite field structure, see Fig. 3c).
FIG. 2. (color online) Scheme of the current-induced allowed
transition for the a) q = 1 charged system and b) q = 0
uncharged system. It has been assumed that ωN ≪ kBT ≪
ωe . ∆0.
clear) g-factors and µB (µN ) the Bohr (nuclear) mag-
neton. In equilibrium, i.e., at zero bias, the occupation
of the dopant level depends on the value of the addi-
tion energy, which ignoring the Zeeman terms and the
tiny correction due to the hyperfine coupling, is given by
ε0(VG) ≡ ǫd + eVG.
We denote the q = 0 eigenstates as mi. Their energies
read as ǫm ≡ ωN Iz. The eigenenergies and eigenvectors
of q = 1 are denoted by ǫM and M i. The q = 1 zero-
field Hamiltonian A~I · ~S can be diagonalized in terms of
the total angular operator F , resulting in two multiplets
(F=4, F=5) with energies EF =4 = −11A/4 and EF =5 =
9A/4, and a zero-field splitting ∆0 = 5A ≈ 30µeV. At
finite magnetic field, the exact eigenvalues of H can also
be calculated analytically.22 The corresponding energy
levels are shown in Fig. 2.
The tunneling Hamiltonian between the single Bi
dopant level and the source and drain electrodes reads
as
Htun = X
λσ
Vλ (cid:0)d†
σcλσ + h.c(cid:1) ,
(2)
where operator cλ,σ annihilates an electron with spin σ
and orbital quantum number λ ≡ η, ~k, with wave vec-
tor ~k and electrode index η = S, D, while operator dσ
annihilates a spin σ electron in the dopant level. The
scattering rate for the tunneling process, ignoring the
hyperfine coupling, is given by Γη
Vη2ρη, where ρη
is the density of states of the electrode. Our model is
very similar to the one used to describe single electron
transport through a quantum dot exchanged coupled to
a single Mn atom.25,26
0 = 2π
The dissipative dynamics of the electro-nuclear spin
system under the influence of the coupling to the elec-
trodes is described by a Bloch-Redfield (BR) master
equation.26,27 The coupling to the reservoir, given by the
tunneling Hamiltonian, involves transitions between the
3
FIG. 3. (color online) a) and b) Contour plot of the dI/dV
vs. applied bias V and on-site energy ε0 at zero magnetic field
(left) and B = 0.6T (right). c) and d) Conduction spectrum
dI/dV as a function of applied bias for different magnetic
fields at ε0 = −0.4µeV and ε0 = 0.8meV respectively. White
horizontal lines in panel a) and b) marks the values of ε0 in the
2D plots c) and d). In all cases, T = 10mK and Γ0 = 0.1µeV.
Let us consider first the q0 = 1 case (left panel in
Fig. 3). At 10 mK only the ground state(s) is (are) oc-
cupied. Thus, a single transition is seen, from the q = 1
to the q = 0 states. As the magnetic field is ramped,
the energy of the transition increases, reflecting the elec-
tronic Zeeman shift. In contrast, in the q0 = 0 case (right
panel in Fig. 3), all the Zeeman split nuclear levels are
equally populated, even down to mK temperatures. Spin
conservation selection rule implicit in Eq. (2) connects
these 10 quasi-degenerate states of the q = 0 manifold to
the hyperfine spin-split levels of the q = 1 manifold with
different energies. As a result, the dI/dV curve reveals
2 peaks at zero field, reflecting the splitting between the
F = 4 and F = 5 states. At higher fields, the two zero-
field peaks split in up to 10 peaks, that can be resolved
at low enough temperature [see Figs. 3c) and 4b)].
Interestingly, the application of a bias to the q0 = 0
state, for which the nuclear spin states are randomized,
can result in a finite average nuclear magnetic moment.
We show this in Fig. 4a) for finite B. At zero bias, the
charge of the dopant level is q0 = 0, and the nuclear spins
are randomized. When the bias hits the addition energy
a selective depopulation of a given Iz level of the q = 0
manifold starts, in favor of a q = 1 state that mixes the
Iz and Iz ±1 components, resulting in a net accumulation
of nuclear spin. When all the transitions to the F = 4
manifold are allowed, the nuclear spin vanishes again.
Then, when the bias permits the transitions to the F =
5 manifold, the nuclear spin accumulation starts in the
opposite direction. Thus, when eV /2 matches the center
of the F = 4 multiplet, see Fig. 4a), the nuclear spins
tend to align antiparallel to the electronic spin. Then,
when eV /2 reaches the center of the F = 5 multiplet,
FIG. 4.
(color online) a) Average electronic occupation of
the Bi, hQi/e (black line) and nuclear and electronic spins,
hIzi (red line) and hSzi (blue line), respectively. b) dI/dV
vs. bias for different temperatures. Same parameters as 3c)
with B = 0.6T.
the nuclear spins prefer aligning parallel to the electronic
spin.
Whereas all our results discussed so far refer to steady
state conditions, it is worth pointing out that there are
two very different time scales in the dynamics of the sys-
tem. Whereas the charge equilibrates in the dopant level
in a time scale set by 1/Γ0, the nuclear spin relaxation,
dominated by many events of hyperfine exchange with
the electronic spin and subsequent recharging of the Bi,28
takes place at a much longer time scale, hundreds of time
larger than 1/Γ0, but still much shorter than the intrinsic
T1 time of the nuclear spin. Thus, charge fluctuations in
the Bi induce nuclear spin relaxation.28
We finally discuss the experimental feasibility of our
proposal with state of the art techniques. First, accord-
ing to our simulations, see Fig. 4b), the finite field hyper-
fine splitting is resolved at 10 mK but not a 20 mK. At 40
mK the 2 humps associated to the F = 4 and F = 5 man-
ifolds are clearly resolved. Keeping the transport in the
sequential tunneling regime requires that Γ0 ≪ kBT ,
which at 10mK, translates into I ≪ 200pA. This is within
reach of experimental setups.12,16,19,29,30
In conclusion, we have studied the single electron
transport spectroscopy of the hyperfine structure of a Bi
dopant in a silicon nanotransistor. We have shown that,
at sufficiently low temperatures, and when the dopant
is ionized with a gate, the dI/dV corresponding to se-
quential transport can resolve the hyperfine spectrum of
the electron in the donor level.
In addition, the non-
equilibrium transport at finite field results in a hyper
polarization of the nuclear spin state, or nuclear spin ac-
cumulation. These results are different from our previous
work, where we considered the same system in a different
transport regime, cotunneling, and we showed that in-
elastic cotunneling of the dopant in the q = 1 state could
also resolve the hyperfine spectrum and drive the nuclear
spin states out of equilibrium.24 Future work should de-
termine how, in the cotunneling regime, the appearance
of the Kondo effect7 competes with the reported effect.
This work has been financially supported by MEC-
Spain (Grant Nos.
FIS2010-21883-C02-01, FIS2009-
08744, and CONSOLIDER CSD2007-0010) as well as
Generalitat Valenciana, grant Prometeo 2012-11.
1H. Sellier, G. P. Lansbergen, J. Caro, S. Rogge, N. Collaert,
I. Ferain, M. Jurczak, and S. Biesemans, Phys. Rev. Lett. 97,
206805 (2006).
2M. Pierre, R. Wacquez, X. Jehl, M. Sanquer, M. Vinet, and
O. Cueto, Nature nanotechnology 5, 133 (2009).
3G. P. Lansbergen, G. C. Tettamanzi, J. Verduijn, N. Collaert,
S. Biesemans, M. Blaauboer, and S. Rogge, Nano Letters 10,
455 (2010).
4K. Y. Tan, K. W. Chan, M. Mottonen, A. Morello, C. Yang, J. v.
Donkelaar, A. Alves, J.-M. Pirkkalainen, D. N. Jamieson, R. G.
Clark, et al., Nano Letters 10, 11 (2010).
5V. N. Golovach, X. Jehl, M. Houzet, M. Pierre, B. Roche, M. San-
quer, and L. I. Glazman, Phys. Rev. B 83, 075401 (2011).
6M. Fuechsle, J. A. Miwa, S. Mahapatra, H. Ryu, S. Lee,
O. Warschkow, L. C. L. Hollenberg, G. Klimeck, and M. Y. Sim-
mons, Nature Nanotechnology 7, 242 (2012).
7G. C. Tettamanzi, J. Verduijn, G. P. Lansbergen, M. Blaauboer,
M. J. Calder´on, R. Aguado, and S. Rogge, Phys. Rev. Lett. 108,
046803 (2012).
8B. Kane, Nature 393, 133 (1998).
9D. P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard, and K. B.
Whaley, Nature 408, 339 (2000).
10T. D. Ladd, F. Jelezko, R. Laflamme, Y. N. Y, C. Monroe, and
J. L. O'Brien, Nature 464, 45 (2010).
11J. M. Elzerman, R. Hanson, L. H. W. van Beveren, B. Witkamp,
L. M. K. Vandersypen, and L. P. Kouwenhoven, Nature 430, 431
(2004).
12A. Morello, J. J. Pla, F. A. Zwanenburg, K. W. Chan, K. Y.
Tan, H. Huebl, M. Mottonen, C. D. Nugroho, C. Yang, J. A. v.
Donkelaar, et al., Nature 467, 687 (2010).
13P. Neumann, J. Beck, M. Steiner, F. Rempp, H. Fedder, P. R.
Hemmer, J. Wrachtrup, and F. Jelezko, Science 329, 542 (2010).
4
14K.-M. C. Fu, T. D. Ladd, C. Santori, and Y. Yamamoto, Phys.
Rev. B 69, 125306 (2004).
15D. Sleiter, N. Y. Kim, K. Nozawa, T. D. Ladd, M. L. W. Thewalt,
and Y. Yamamoto, New Journal of Physics 12, 093028 (2010).
16J. R. Petta, J. M. Taylor, A. C. Johnson, A. Yacoby, M. D.
Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard, Phys.
Rev. Lett. 100, 067601 (2008).
17D. J. Reilly, J. M. Taylor, J. R. Petta, C. M. Marcus, M. P.
Hanson, and A. C. Gossard, Science 321, 817 (2008).
18S. Foletti, H. Bluhm, D. Mahalu, V. Umansky, and A. Yacoby,
Nature Physics 5, 903 (2009).
19T. Kobayashi, K. Hitachi, S. Sasaki, and K. Muraki, Phys. Rev.
Lett. 107, 216802 (2011).
20R. E. George, W. Witzel, H. Riemann, N. V. Abrosimov,
N. Notzel, M. L. W. Thewalt, and J. J. L. Morton, Phys. Rev.
Lett. 105, 067601 (2010).
21G. W. Morley, M. Warner, A. M. Stoneham, P. T. Greenland,
J. van Tol, C. W. M. Kay, and G. Aeppli, Nature Materials 9,
725 (2010).
22M. H. Mohammady, G. W. Morley, and T. S. Monteiro, Phys.
Rev. Lett. 105, 067602 (2010).
23T. Sekiguchi, M. Steger, K. Saeedi, M. L. W. Thewalt, H. Rie-
mann, N. V. Abrosimov, and N. Notzel, Phys. Rev. Lett. 104,
137402 (2010).
24F. Delgado and J. Fern´andez-Rossier, Phys. Rev. Lett. 107,
076804 (2011).
25A. L. Efros, E. I. Rashba, and M. Rosen, Phys. Rev. Lett. 87,
206601 (2001).
26J. Fern´andez-Rossier and R. Aguado, Phys. Rev. Lett. 98, 106805
(2007).
27C. Cohen-Tannoudji, G. Grynberg, and J. Dupont-Roc, Atom-
Photon Interactions (Wiley and Sons, INC., New York, 1998).
28L. Besombes, Y. Leger, J. Bernos, H. Boukari, H. Mariette, J. P.
Poizat, T. Clement, J. Fern´andez-Rossier, and R. Aguado, Phys.
Rev. B 78, 125324 (2008).
29J. Baugh, Y. Kitamura, K. Ono, and S. Tarucha, Phys. Rev.
Lett. 99, 096804 (2007).
30T. S. Jespersen, K. Grove-Rasmussen, J. Paaske, K. Muraki,
T. F. an J. Nygard, and K. Flensberg, Nature Physics 7, 348
(2011).
|
1108.5256 | 2 | 1108 | 2011-09-05T15:29:49 | Second-harmonic generation from coupled plasmon modes in a single dimer of gold nanospheres | [
"cond-mat.mes-hall"
] | We show that a dimer made of two gold nanospheres exhibits a remarkable efficiency for second-harmonic generation under femtosecond optical excitation. The detectable nonlinear emission for the given particle size and excitation wavelength arises when the two nanoparticles are as close as possible to contact, as in situ controlled and measured using the tip of an atomic force microscope. The excitation wavelength dependence of the second-harmonic signal supports a coupled plasmon resonance origin with radiation from the dimer gap. This nanometer-size light source might be used for high-resolution near-field optical microscopy. | cond-mat.mes-hall | cond-mat |
Second-harmonic generation from coupled plasmon modes
in a single dimer of gold nanospheres
A. Slablab1, L. Le Xuan1,∗ M. Zielinski1, Y. de Wilde2, V. Jacques1, D. Chauvat1, and J.-F. Roch1
1Laboratoire de Photonique Quantique et Mol´eculaire,
Ecole Normale Sup´erieure de Cachan and CNRS, UMR 8537, F-94235 Cachan, France and
2Institut Langevin, ESPCI ParisTech and CNRS, UMR 7587, F-75231 Paris, France
(Dated: June 5, 2018)
We show that a dimer made of two gold nanospheres exhibits a remarkable efficiency for second-
harmonic generation under femtosecond optical excitation. The detectable nonlinear emission for
the given particle size and excitation wavelength arises when the two nanoparticles are as close as
possible to contact, as in situ controlled and measured using the tip of an atomic force microscope.
The excitation wavelength dependence of the second-harmonic signal supports a coupled plasmon
resonance origin with radiation from the dimer gap. This nanometer-size light source might be used
for high-resolution near-field optical microscopy.
INTRODUCTION
During the last decade, nonlinear nanoparticles have
been extensively studied as new light sources at the
nanoscale. In particular, nanoparticles consisting of non-
centrosymmetric material exhibit second-harmonic gen-
eration (SHG) [1–4] which can be used for nonlinear opti-
cal microscopy [5–9]. For nanoparticles made of pure no-
ble metals, high electron polarisability could lead to much
stronger nonlinear effects. However, inversion symmetry
of the metallic crystalline structure forbids bulk second-
order electric dipole response. For a metallic nanosphere,
second-order polarization associated to induced surface
dipole moments and volumic quadrupole moments pro-
duces a weak SHG signal [10–12], with a strong depen-
dence to the nanoparticle shape and size [13, 14].
Efficient SHG at the nanoscale can be obtained by plas-
mon enhancement at the surface of a metallic tip [15],
or by coupled plasmon modes in engineered metallic
nanostructures with specific geometry, like T-shaped gold
nano-dimers [16], bowtie-shaped nano-antenna [17], and
gold nanowires [18].
In that context, a simple dimer
structure consisting of two metallic nanospheres appears
as a testbed for studying the plasmon coupling influ-
ence on the nonlinear optical response. Indeed, the lin-
ear scattering of this composite nanostructure exhibits
a wealth of specific properties depending on its geomet-
rical parameters [19–24]. The coupled plasmon modes
have resonance frequencies which can be tuned over the
whole visible spectrum by changing the dimer geome-
try [21]. Moreover, the electromagnetic field density is
greatly enhanced at the dimer gap and highly efficient
four-wave mixing has been observed from such metallic
dimers [25, 26].
In this Letter, we explore the second-order nonlin-
ear properties of a single dimer consisting of two gold
nanospheres (GNs) with controlable distance. We show
that this dimer nanostructure leads to highly efficient
SHG under femtosecond optical illumination in spite of
its apparent centrosymmetry considered as a whole. The
relative position of the two GNs is adjusted with nanome-
ter accuracy using the tip of an atomic force microscope
(AFM) [23]. We show that the SHG signal under the ex-
perimental excitation wavelength-particle sizes condition
is strongly enhanced when the two GNs are very close to
contact, with a strong dependence on their mutual size
ratio. The variation of the SHG intensity with the exci-
tation wavelength supports the role of an near-infrared
(IR) resonance resulting from the coupling of the plas-
mon oscillations in the two gold nanospheres [21, 25].
EXPERIMENT AND RESULTS
The principle of the experiment is shown in Fig. 1(a).
The nonlinear optical response of coupled GNs is inves-
tigated using a nano-optomechanical setup consisting of
an AFM (Asylum Research, MFP-3D BIO) on top of a
self-made inverted optical microscope. The GNs (100-
nm diameter) were purchased from the British Biocell
International (BBI) corporation. The colloidal solution,
which is stabilized in water by citric acid, is deposited by
spin coating on a standard 150-µm thick glass coverslip.
Before spin coating, the glass coverslip was silanized in
order to functionalize its surface with NH+
3 groups. Since
the GNs have negative charges on their surface, an elec-
trostatic interaction allows to efficiently catch them on
the glass surface during the spin coating, thus leading to
well-dispersed GNs on the substrate. The AFM is used
both to record the surface topography of the sample and
to perform mechanical manipulation of the GNs. The
formation of a single gold dimer is gradually achieved by
pushing a GN toward another with the AFM tip, thus
controlling the interparticle distance from large separa-
tion to contact between the two spheres. A titanium-
doped sapphire (Ti:Sa) laser emitting 100 fs pulses in
the 800-1000 nm wavelength range at 80 MHz repetition
rate is tightly focused onto the sample through a high nu-
merical aperture microscope objective (NA = 1.4, corre-
sponding to a maximum collection half angle of 68◦). The
2
from the second-order nonlinear process [27]. We at-
tribute this emission line to SHG from the dimer struc-
ture. Furthermore, the emission spectra clearly exhibit a
maximum of SHG when the excitation laser wavelength
is within the 950 to 1000 nm range (Fig. 2(b)), giving ev-
idence for an infrared resonance associated to the dimer
structure.
This spectral behavior of the second-harmonic re-
sponse is well explained by the analysis of Ref. [21] which
theoretically describes the linear optical response of a
gold dimer in the nearly touching regime. When the
two GNs are getting close to contact, the two single
nanosphere plasmon modes are coupled, ending in a new
plasmon mode which resonance is rapidly shifted towards
the infrared. We confirm this prediction with a finite
difference time domain (FDTD) simulation of the linear
optical properties of a gold dimer, performed with the
Lumerical software [28]. In agreement with Ref. [21], a
resonant behavior of the scattering cross-section is ob-
served as well as a rapid shift of this resonance to the
infrared as the gap is reduced. Choosing a 0.08 nm ef-
fective gap distance for touching GNs of 100 nm size, we
compute the linear scattering cross-section of the dimer,
displayed in Fig. 2(c). The dimer scattering cross-section
variation closely matches the measured SHG excitation
FIG. 2: Second-harmonic radiation of individual gold dimers.
(a)-Emission spectrum from a gold dimer recorded for an ex-
citation laser wavelength at λex = 980 nm, showing a strong
SHG spectral peak and a much weaker and broader two-
photon excited luminescence (TPEL). (b)-Emission spectra
recorded while tuning the excitation laser wavelength λex from
850 to 1010 nm. (c)-FDTD simulation of the scattering cross-
section of a dimer consisting of two 100 nm GNs separated by
a 0.08 nm effective gap distance.
(a)-Experimental setup. AFM tip: Olympus,
FIG. 1:
AC160TS used as purchased; MO: oil immersion microscope
objective (×100, NA = 1.4); DM: dichroic mirror; FM:
switchable mirror directing the collected light either to a spec-
trograph or to a silicon avalanche photodiode (APD). To-
pography measurements are done in the AFM tapping mode
while the contact mode is used to perform mechanical manip-
ulation of the GN. Optical images are recorded by scanning
the sample in x and y directions while recording the emitted
photons with the APD. The MO is mounted on a piezoelec-
tric transducer in order to adjust the laser beam focus on the
z-axis. (b)-When two GNs are in contact (see AFM image), a
strong nonlinear emission is observed (see optical image). No
emission is observed for isolated single GNs. (c)-When two
isolated GNs are brought in contact using the AFM tip, an
associated bright emission spot appears in the optical image.
light emitted by the GNs is collected with the same ob-
jective, spectrally filtered from the remaining excitation
light using a dichroic beamsplitter, and finally directed
either to a spectrograph or to a silicon avalanche pho-
todiode (APD) working in the photon counting regime.
The AFM tip and the excitation laser beam are carefully
aligned on the same axis. Simultaneous record of the to-
pography and the optical response then allows to monitor
the onset of second-order nonlinear effects as the dimer
is formed.
A typical realization of the experiment is depicted in
Figs. 1(b) and (c). For the laser input mean power range
used in the experiment (∼ 500 µW), the nonlinear optical
emission is not observed for isolated GNs (see Fig. 1(b)),
whereas a strong nonlinear optical signal clearly appears
when two isolated GNs are brought to contact using the
AFM-based nanopositioning technique (Fig. 1(c)).
Emission spectra recorded from this gold dimer while
tuning the excitation laser wavelength from 850 to
1000 nm are shown in Fig. 2(b). For each excitation
wavelength, a strong narrow emission peak at half the
excitation wavelength is observed alongside with a weak
and broad two-photon excited luminescence (TPEL)
background (Fig. 2(a)). In addition, the peak intensity
scales quadratically with the laser power, as expected
DMAFM tipSpectrographAPDFM200 nm300 nmAFM imageOptical imagefs laser(a)(b)(c)GNsMO@ 990 nmxyz(b)250020001500100050005405205004804604404201000980960940920900880Emission wavelength (nm)λem=λex2Intensity (a.u.)Excitation wavelength (nm)Intensity (a.u.)(a)300025002000150010005000560520480440Emission wavelength (nm)Emission wavelength (nm)SHGTPELExcitation wavelength (nm)Scattering cross section (a.u.) 120011001000900(c)3
allows to infer when the two GNs are nearly in contact.
As one GN is approached toward the other with nano-
metric steps, one could expect a gradual rise of the
SHG signal. However, no SHG is detected until the two
GNs are very close to contact (see inset in Fig. 4-middle
panel), in which case a high count rate of SHG photons is
observed (Fig. 5-left panel) at IR resonance wavelength.
By pushing further one GN towards the other, the ob-
served SHG signal remains unchanged, allowing to con-
clude that the SHG signal only appears when the two
GNs are very close to contact. This behavior is related
to the sudden increase of the excitation field at the gap re-
gion for very small interparticle distance. Indeed, the in-
duced accumulation of opposite charges discussed above
leads to a strong enhancement of the excitation electro-
magnetic field at the dimer gap. Using the measured
particles sizes (particle diameters of 100 nm and 80 nm)
from Fig. 4 (middle panel) as parameters and consid-
ering the GNs completely spherical, FDTD simulations
depicted in Fig. 4 (right panel) shows that the excitation
field gains about 6 orders of magnitude from an inter-
FIG. 4: (Left panel) (a) to (d)-Topography images showing
the assembly of a dimer by using the AFM tip to move one
GN towards the other. (Central panel) A cross-section (red
dashed line in (a)) is used to estimate the distance d between
the two GNs. The SHG signal is observed only when the two
GNs are in contact as shown in the insets of the cross-section
graphs. (Right panel) FDTD simulation of the electromag-
netic field intensity building up at the dimer gap as this gap
is reduced. The gap values are those inferred from the fit of
the cross-sections. In the case (d) of contact between the two
spheres, a 0.08 nm value is taken in order to agree with the
spectral behavior shown in Fig. 2. The simulation has been
done with spherical particles of 100 nm and 80 nm in diame-
ter, as measures topographically. Note the log-scale showing
a six-order-of-magnitude increase between the two extreme
values of the gap.
FIG. 3: Polar diagrams showing the SHG efficiency as a func-
tion of the angle of the linearly-polarized excitation laser for
two different dimers (see topography image).
Intense lobes
along the dimer axis are observed.
spectra.
In particular, it exhibits a minimum close to
920 nm and a maximum close to 990 nm, as observed in
Fig. 2(a). Such a correlation supports the resonant plas-
monic origin of the SHG emission.
To further characterize the nature of the coupled plas-
mon mode, a polarization analysis is performed while ex-
citing the dimer in resonance at λex = 990 nm and rotat-
ing the linear excitation polarization using a half-wave
plate placed before the dichroic beamsplitter (Fig. 1(a)).
The polar diagrams measured for two different dimers
are shown in Fig. 3. The SHG intensity vanishes when
the excitation polarization is perpendicular to the dimer
axis, and increases to a maximum value when the polar-
ization is set along the dimer axis, as revealed by the cor-
responding AFM topographic images. This dipolar like
response can be explained with a simple model based on
the coupling between the plasmon oscillations in the two
GNs. At the dimer gap, a strong accumulation of oppo-
site charges close to each other is induced. Any charge
oscillation driven by external light within the gap is ac-
companied by a charge redistribution over each particle
in order to maintain intraparticle charge neutrality. This
redistribution finally leads to a large dipole strength of
the gold dimer considered as a whole. While the interac-
tion of light with charges is weak for an exciting electric
field perpendicular to the dimer axis, the plasmon mode
coupling becomes efficient when the field is applied along
this axis, as experimentally observed (see Fig. 3).
Now we study the onset of the nonlinear optical re-
sponse while varying the interparticle distance from large
separation to contact. For that purpose, a gold dimer
is gradually formed with nanometric steps by pushing a
GN toward another with the AFM tip. After each dis-
placement, an AFM topography image and a raster scan
SHG image are jointly recorded. Four intermediate sit-
uations are shown in Fig. 4(a)-(d). Despite a relatively
small AFM tip radius (≈ 7 nm), a direct measurement
of the gap distance d between the two nanoparticles is
obviously not possible. However, this parameter can be
inferred with a few nanometers accuracy from the dip ob-
served in a cross section of the topographic image of the
dimer (see central panel of Fig. 4), once the size of the
GNs and the tip radius are known. The corresponding fit
0°45°90°135°180°225°270°315°0°45°90°135°180°225°270°315°200 nmSHGDistance (nm)0200400(b)SHGSHGd∼20nmd∼4nmd∼0nm300 nmSHG100(c)100100SHGd∼38nm100 nm(a)100-202468logE2(d)particle distance decreasing from 4.0 to 0.08 nm. SHG
originates from this huge field which is confined within
a nanometric volume located at the dimer gap. This
localized excitation is compatible with the SHG raster-
scan images (see Fig.1(c)) which show bright spots with a
diffraction-limited size (≈ 320 nm FWHM), as expected
for a point-like emission.
DISCUSSION
Second-order nonlinear optical processes require an
overall noncentrosymmetry which can originate from a
combination of different effects, such as intrinsic geomet-
rical features of the dimer, excitation field dissymetry,
or our specific detection geometry. Since a simple size
difference between the two GNs would break the dimer
centrosymmetry [25], we measured the SHG efficiency as
a function of the size difference h between the two GNs.
The experiment is performed for 19 single dimers in close
contact geometry while keeping the excitation at 990 nm
wavelength. As shown in Fig. 5(a), the SHG efficiency
decreases by almost two orders of magnitude when h is
changed from 0 to 35 nm. This result is a priori in con-
tradiction with the above argument where SHG efficiency
might increase with the size difference between the two
GNs of the dimer.
To understand this behavior, we compute a FDTD sim-
ulation of the optical scattering cross section of the dimer
as a function of the h parameter. As the asymmetry be-
tween the two GNs is increased, the resonance resulting
4
from the coupling of the plasmon modes is blue-shifted
and its amplitude decreases (Fig. 5(b)). From these sim-
ulations, we can then compute the fourth power of the
electric field amplitude at the gap while exciting at the
fixed 990 nm wavelength. This quantity, which is rele-
vant to the locally enhanced field around the gap, is found
to decrease by roughly two orders of magnitude when h
varies from 0 to 35 nm (see inset in Fig. 5(b)), the same
factor as measured. It confirms that the decrease in SHG
intensity matches the wavelength shift in resonance, and
that the SHG originates from the neighborhood region
around the dimer gap, rather than from an overall bro-
ken symmetry corresponding to the association of two
GNs with different size.
In our specific case, the SHG might originate from the
specific experimental geometry. The tightly focused ex-
citation beam creates a field dissymmetry from one side
to the other of the gap plan in the light propagation
direction. This dissymmetry induces nonlocally excited
electric-dipole second-order processes inside the spheres
and surrounding the gap, and locally excited quadrupolar
processes from the metallic surfaces [11, 12, 14]. Any re-
sulting off-axis radiation can then be efficiently collected
by the high numerical aperture objective.
In addition,
the SHG could be enhanced either by specific geomet-
rical features at the gap associated to facets of the two
GNs [30] or by organic layers on the GNs surface [31].
CONCLUSION AND PROSPECTS
In conclusion, we have experimentally demonstrated
that highly efficient SHG is obtained from two GNs when
very close to contact, for the given geometry-size rela-
tion and optical excitation wavelength. Spectral analysis
of the SHG response and polarization analysis support a
coupled plasmon resonance origin of the nonlinear emis-
sion. Since SHG is emitted from a nanometric volume at
the dimer gap, it might be used as a nanosource for the
development of high-resolution near-field imaging. Using
the AFM-tip pushing technique, more complex nanos-
tructures can be assembled, such as a trimer nanoparticle
which can lead to a tunable nano-half wave plate [32], or
a succession of metallic beads with decreasing sizes which
forms an efficient plasmonic nano-lens with high field con-
centration [33]. Such structures are likely to exhibit sur-
prising nonlinear properties in the optical domain.
FIG. 5: (a) SHG efficiency plotted on a log-scale as a function
of the size difference h between the two GNs of an asymmetric
dimer (see inset). The excitation wavelength is at λ = 990
nm and laser input mean power is (∼ 500 µW). (b) Simula-
tion of the scattering cross-section for different values of h.
Inset: Log-scale plot of the fourth power of the electric field
amplitude at the dimer gap Egap4 as a function of h, for a
constant 990 nm excitation wavelength.
Acknowledgements
We dedicate this work to our esteemed colleague Do-
minique Chauvat who passed away during the prepara-
tion of the manuscript. We are grateful to S. Perruchas
and T. Gacoin for providing us with the GNs sample. We
h=10nmh=20nmh=30nm1300120011001000900wavelength (nm)Scattering cross-section (a.u.)h(nm)h100090011001200Wavelength (nm)1300(a)(b)h=0nmEgap4101424101524101623020100456710023456710002403020100103210h(nm)201030400SHG emission (x10 cts.s )-1 3thank J.-J. Greffet and F. Marquier for priceless discus-
sions. This work is supported by C'Nano Ile-de-France.
∗ Electronic address: xuan-loc.le@m4x.org
[1] J. C. Johnson, H. Yan, R. D. Schaller, P. B. Petersen,
P. Yang and R. J. Saykally, "Near-Field Imaging of Non-
linear Optical Mixing in Single Zinc Oxide Nanowires,"
Nano Lett. 2, 279-283 (2002).
[2] A. B. Djurisi´c and Y. H. Leung, "Optical properties of
ZnO nanostructures," Small 2, 944-961 (2006).
[3] M. Zielinski, D. Oron, D. Chauvat, J. Zyss, " Second-
harmonic generation from a single core/shell quantum
dot," Small 5, 2835-2840 (2009).
[4] M. Zielinski, S. Winter, R. Kolkowski, C. Nogues, D.
Oron, J. Zyss, D. Chauvat, "Nanoengineering the sec-
ond order susceptibility in semiconductor quantum dot
heterostructures," Opt. Express 19, 6657-6670 (2011).
[5] L. Bonacina, Y. Mugnier, F. Courvoisier, R. Le Dantec,
J. Extermann, Y. Lambert, V. Boutou, C. Galez, J.-P.
Wolf, " Polar Fe(IO3)3 nanocrystals as local probes for
nonlinear microscopy," Applied Physics B 87, 399-403,
(2007).
[6] L. Le Xuan, C. Zhou, A. Slablab, D. Chauvat, C.
Tard, S. Perruchas, T. Gacoin, P. Villeval, J.-F. Roch,
"Photostable second-harmonic generation from a single
KTiOPO4 nanocrystal for nonlinear microscopy," Small
4, 1332-1336 (2008).
[7] C.-L. Hsieh, R. Grange, Y. Pu, D. Psaltis, " Three-
dimensional harmonic holographic microcopy using
nanoparticles as probes for cell imaging," Opt. Express
17, 2880-2891(2009).
[8] A. V. Kachynski, A. N. Kuzmin, M. Nyk, I. Roy, P. N.
Prasad, "Zinc Oxide Nanocrystals for Non-resonant Non-
linear Optical Microscopy in Biology and Medicine," J.
Phys. Chem. C 112, 10721-10724 (2008).
[9] Y. Nakayama, P. J. Pauzauskie, A. Radenovic, R. M.
Onorato, R. J. Saykally, J. Liphardt, P. Yang, "Tunable
nanowire nonlinear optical probe," Nature 447, 1098-
1101 (2007).
[10] V. L. Brudny, B. S. Mendoza, W. L. Mochan, "Second-
harmonic generation from spherical particles," Phys.
Rev. B 62, 11152-11162 (2000).
[11] J. Dadap, J. Shan, T. F. Heinz, "Theory of optical
second-harmonic generation from a sphere of centrosym-
metric material: small-particle limit," J. Opt. Soc. Am.
B 21, 1328-1347 (2004).
[12] J. Butet, J. Duboisset, G. Bachelier, I. Russier-Antoine,
E. Benichou, C. Jonin, P.-F. Brevet, "Optical Second
Harmonic Generation of Single Metallic Nanoparticles
Embedded in a Homogeneous Medium," Nano Lett. 10,
1717-1721 (2010).
[13] J. Nappa, G. Revillod, I. Russier-Antoine, E. Benichou,
C. Jonin, P.-F. Brevet, "Electric dipole origin of the
second harmonic generation of small metallic particles,"
Phys. Rev. B 71, 165407-165410 (2005).
[14] J. Shan, J. I. Dadap, I. Stiopkin, G. A. Reider, T. F.
Heinz, "Experimental study of optical second-harmonic
scattering from spherical nanoparticles," Phys. Rev. A
73, 023819-023822 (2006).
[15] A. Bouhelier, M. R. Beversluis, A. Hartschuh, L.
5
Novotny, "Near-field second-harmonic generation in-
duced by local field enhancement," Phys. Rev. Lett. 90,
013903-013906 (2003).
[16] H. Husu, B. K. Canfield, J. Laukkanen, B. Bai, M. Kuit-
tinen, J. Turunen, M. Kauranen, "Local-field effects in
the nonlinear optical response of metamaterials," Meta-
materials 2, 155-168 (2008).
[17] T. Hanke, G. Krauss, D. Trautlein, B. Wild, R. Brats-
chitsch, A. Leitenstorfer, "Efficient Nonlinear Light
Emission of Single Gold Optical Antennas Driven by
Few-Cycle Near-Infrared Pulses," Phys. Rev. Lett. 103,
257404-257407 (2009).
[18] A. Benedetti, M. Centini, C. Sibilia, M. Bertolotti, "En-
gineering the second harmonic generation pattern from
coupled gold nanowires," J. Opt. Soc. Am. B 27, 408-
416 (2010).
[19] W. Rechberger, A. Hohenau, A. Leitner, J. R. Krenn,
B. Lamprecht, F. R. Aussenegg, "Optical properties of
two interacting gold nanoparticles," Opt. Commun. 220,
137-141 (2003).
[20] T. Atay, J.-H. Song, A. V. Nurmikko, "Strongly inter-
acting plasmon nanoparticle pairs: From dipole-dipole
interaction to conductively coupled regime," Nano Lett.
4, 1627-1631 (2004).
[21] I. Romero, J. Aizpurua, G. W. Bryant, F. D. Garc´ıa de
Abajo, "Plasmons in nearly touching metallic nanoparti-
cles: singular response in the limit of touching dimers,"
Opt. Express 14, 9988-9999 (2006).
[22] A. L. Lereu, G. Sanchez-Mosteiro, P. Ghenuche, R.
Quidant, N. F. Van Hulst, "Individual gold dimers inves-
tigated by far- and near-field imaging," Journal of Mi-
croscopy 229, 254-258 (2008).
[23] S. Schietinger, M. Barth, T. Aichele, O. Benson,
"Plasmon-Enhanced Single Photon Emission from a
Nanoassembled Metal Diamond Hybrid Structure at
Room Temperature," Nano Lett. 9, 1694-1698 (2009).
[24] P. K. Jain and M. A. El-Sayed, "Plasmonic coupling
in noble metal nanostructures," Chem. Phys. Lett. 487,
153-164, (2010).
[25] M. Danckwerts and L. Novotny, "Optical frequency mix-
ing at coupled gold nanoparticles," Phys. Rev. Lett. 98,
026104-026107 (2007).
[26] S. Palomba and L. Novotny, "Near-Field Imaging with a
Localized Nonlinear Light Source," Nano Lett. 9, 3801-
3804 (2009).
[27] R. W. Boyd, Nonlinear Optics (Academic Press, New
York, 1992).
[28] We use the gold dielectric constants reported in Ref. [29]
for a 500-1400 nm wavelength range and we do not take
into account the influence of the substrate. The two GNs
are supposed to be in air and the polarization of the
excitation field is linear oriented along the dimer axis.
[29] P. B. Johnson and R. W. Christy, "Optical constants of
the noble metals," Phys. Rev. B 6, 4370-4379 (1972).
[30] S.-C. Yang, H. Kobori, C.-L. He, M.-H. Lin, H.-Y. Chen,
C. Li, M. Kanehara, T. Teranishi, S. Gwo, "Plasmon Hy-
bridization in Individual Gold Nanocrystal Dimers: Di-
rect Observation of Bright and Dark Modes," Nano Lett.
10, 632-637 (2010).
[31] P. Rooney, A. Rezaee, S. Xu, T. Manifar, A. Hassan-
zadeh, G. Podoprygorina, V. Bohmer, C. Rangan, S.
Mittler, "Control of surface plasmon resonances in dielec-
trically coated proximate gold nanoparticles immobilized
on a substrate," Phys. Rev. B 77, 235446-235452 (2008).
[32] Z. Li, T. Shegai, G. Haran, H. Xu, "Multiple-particle
nanoantennas for enormous enhancement and polariza-
tion control of light emission," ACS Nano 3, 637-642
(2009).
[33] K. Li, M. I. Stockman, D. J. Bergman, "Self-Similar
Chain of Metal Nanospheres as an Efficient Nanolens,"
Phys. Rev. Lett. 91, 227402-227405 (2003).
6
|
1303.3529 | 2 | 1303 | 2013-09-19T19:22:06 | Hall Drag and Magnetodrag in Graphene | [
"cond-mat.mes-hall"
] | Massless Dirac fermions in graphene at charge neutrality form a strongly interacting system in which both charged and neutral (energy) modes play an important role. These modes are essentially decoupled in the absence of a magnetic field, but become strongly coupled when a field is applied. We show that these ideas explain the recently observed giant magnetodrag, arising in classically weak fields when electron density is tuned near charge neutrality. We predict strong Hall drag in this regime, which is in stark departure from the weak coupling regime, where theory predicts the absence of Hall drag. Energy-driven magnetodrag and Hall drag arise in a wide temperature range and at weak magnetic fields, and feature an unusually strong dependence on field and carrier density. | cond-mat.mes-hall | cond-mat |
Hall Drag and Magnetodrag in Graphene
1 Department of Physics, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA and
2 School of Engineering and Applied Sciences, Harvard University, Cambridge, Massachusetts 02138, USA
Justin C. W. Song1,2 and Leonid S. Levitov1
Massless Dirac fermions in graphene at charge neutrality form a strongly interacting system in
which both charged and neutral (energy) modes play an important role. These modes are essentially
decoupled in the absence of a magnetic field, but become strongly coupled when a field is applied.
We show that these ideas explain the recently observed giant magnetodrag, arising in classically
weak fields when electron density is tuned near charge neutrality. We predict strong Hall drag in
this regime, which is in stark departure from the weak coupling regime, where theory predicts the
absence of Hall drag. Energy-driven magnetodrag and Hall drag arise in a wide temperature range
and at weak magnetic fields, and feature an unusually strong dependence on field and carrier density.
Graphene near charge neutrality (CN) hosts an intrigu-
ing electron-hole system with unique properties[1 -- 10].
Our understanding of the behavior at CN would greatly
benefit from introducing ways to couple the novel neu-
tral modes predicted at CN to charge modes which can
be easily probed in transport measurements. There is a
long history of employing magnetic field for such a pur-
pose, since transport in charge-neutral plasmas is ultra-
sensitive to the presence of external magnetic fields[11].
A new interesting system in which magnetotransport
at CN can be probed are atomically thin graphene double
layer G/hBN/G structures[12, 13]. Strong Coulomb cou-
pling between adjacent layers in these systems results in
strong Coulomb drag, arising when current applied in one
(active) layer induces a voltage in the adjacent (passive)
layer[13 -- 22]. Recent measurements[13] revealed drag re-
sistance that peaks near CN and has dramatic magnetic
field dependence, with the peak value increasing by more
than an order of magnitude (and changing sign) upon
application of a relatively weak B field. Strong mag-
netic field dependence of drag has been observed pre-
viously in other double layer two-dimensional electron
gas (2DEG) heterostructures[23 -- 25], however these ex-
periments were carried out in the quantum Hall regime,
whereas the anomalous magnetodrag found in Ref.[13]
occurs at classically weak fields B <∼ 1 T.
Here we explain this puzzling behavior in terms of an
energy-driven drag mechanism which involves coupled
energy and charge transport[20, 22] (see Fig.1). En-
ergy transport plays a key role because of fast verti-
cal energy transfer due to interlayer Coulomb coupling
in G/hBN/G systems[20] and relatively slow electron-
lattice cooling[28, 29]. As a result, current applied in one
layer can create a spatial temperature gradient for elec-
trons in both layers, giving rise to thermoelectric drag
voltage. The effect peaks at CN, since thermoelectric re-
sponse is large close to CN[8 -- 10] and diminishes as 1/EF
upon doping away from CN[30, 31]. Drag arising from
this mechanism depends on thermoelectric response and,
unlike the conventional momentum drag mechanism, it
is insensitive to the electon-electron interaction strength.
Another interesting effect that can be probed in these
FIG. 1: Energy-driven magnetodrag in a double layer
graphene heterostructure close to CN. (a) Schematic of charge
current, temperature gradients, and electric field in the two
layers that give rise to a negative ρdrag(cid:107)
. (b,c) Magnetodrag
resistivity, ρdrag(cid:107)
, obtained from Eqs.(11),(13). Parameter val-
ues: B = 0.6 T, n0 = 1011 cm−2, T = 150 K, and ρ0 = h
3e2 .
The B = 0 dependence taken from the model of drag at zero
B field [20, 21].(d) Experimentally measured magnetodrag re-
sistivity in G/hBN/G heterostructures, reproduced from Ref.
[13] for the same B values as in (c). Application of magnetic
field leads to a giant negative drag at CN. Note the simi-
larity between data and theoretically predicted drag density
dependence, B dependence, and sign.
systems is that of Hall drag. It has long been argued that,
at weak coupling, no Hall voltage can arise in the passive
layer in the presence of current in the active layer [26, 27].
This is because transferred momentum is parallel to ve-
locity, allowing only a longitudinal "back-current" to de-
velop in the passive layer. As we shall see, a very differ-
ent behavior arises at strong coupling, owing to the long-
range energy currents leading to electron-lattice tempera-
ture imbalance. Close to CN, the magnitude of the cross-
couplings between charge and energy currents becomes
large, producing a finite Hall drag, VH = Rdrag
H I(cid:107).
H and Rdrag(cid:107)
As we will see, energy currents result in Hall and mag-
netodrag resistances, Rdrag
, that are large and
peak near CN, see Fig.1 and Fig.2. These large values
arise even for classically weak fields B ∼ 0.1 T, exceed-
ing by two orders of magnitude the values found previ-
ously in GaAs/AlGaAs 2DEG heterostructures [23 -- 25] at
similar fields. The mechanism based on coupled energy
and charge transport predicts large and negative drag at
CN that matches recent experiments (see Fig.1c,d). Our
mechanism naturally leads to Hall drag because verti-
cal energy transfer between layers does not discriminate
between longitudinal and transverse heat currents since
temperature profile is a scalar field. This stands in con-
trast to conventional momentum driven drag, where mo-
mentum transfer is parallel to the applied current [26, 27].
Heat current and an electric field, induced by charge
current and temperature gradients, are coupled via the
thermoelectric effect altered by the B field,
∇T
T
.
jq = Qj, E = Q
(1)
Here Q is a 2 × 2 matrix, of which diagonal compo-
nents describe the Peltier and Thompson effects, and off-
diagonal components describe the Nernst-Ettingshausen
effect. Onsager reciprocity requires that Q in both the
expressions for jq and E are the same [see analysis fol-
lowing Eq.(9)]. As an example of how our mechanism
produces drag consider the Hall bar geometry, see Fig.1.
When a longitudinal charge current is applied in the ac-
tive layer (for B (cid:54)= 0) a transverse (Ettingshausen) heat
current develops in both layers through efficient verti-
cal energy transfer. Nernst voltage in the passive layer
results in a longitudinal magnetodrag of a negative sign.
To obtain the electric field in layer 2 induced by cur-
rent applied in layer 1, we first need to understand the
coupling of temperature profiles T1,2(r) in the two layers.
Energy transport in the system can be described by
−∇κ1∇δT1 + a(δT1 − δT2) + λδT1 = −∇ ·(cid:0)Q(1)j(cid:1)
−∇κ2∇δT2 + a(δT2 − δT1) + λδT2 = 0
with a the energy transfer rate between the two layers
[20], λ the electron-lattice cooling rate, and δTi = Ti− T0
(here T0 is the lattice temperature, equal for both layers).
Here we focus on a Hall-bar geometry of two parallel
rectangular layers of dimensions L×W , L (cid:29) W , pictured
in Fig.1a. For the sake of simplicity, we treat the electric
and heat currents as independent of the x coordinate
along the bar. In layer 1, current is injected at x = −L/2
and drained at x = L/2. In layer 2, the Hall drag voltage
arising across the device, VH and the longitudinal drag
voltage, V(cid:107), are evaluated as
VH =
E(2)
y dy, V(cid:107) =
L
W
E(2)
x dy
(3)
(cid:90) W/2
−W/2
(cid:90) W/2
−W/2
2
The electric and thermal variables may depend on the
transverse coordinate y, see below.
Boundary conditions for a Hall bar require electric cur-
rent being tangential to the side boundaries, y = ±W/2,
and zero temperature imbalance at the ends, x = ±L/2,
reflecting that the current and voltage contacts act as
ideal heat sinks. The electric current parallel to the
boundaries y = ±W/2 gives rise to the Ettingshausen
heat current that may have a component transverse to
the Hall bar. The divergence of this heat current, ap-
pearing on the right hand side of Eq.(2), acts as an effec-
tive boundary delta function source in the heat transport
equations. Boundary conditions can profoundly influence
the symmetry of the resultant drag resistivity, see below.
We consider the case of a spatially uniform Q in both
layers. The ideal heat sinks at x = ±L/2 mean that no
temperature imbalance develops in the x-direction (ex-
cept for some "fringing" heat currents near the contacts
which give a contribution small in W/L (cid:28) 1, which we
will ignore in the following discussion). Since no temper-
ature gradients are sustained in the x-direction for from
the ends, we can reduce our problem Eq.(2) to a quasi-1D
problem with temperature profiles that only depend on
the y-coordinate. As a result, the only heat source arises
from the Ettingshausen effect Q(1)j = (Q(1)
yx j)(cid:98)y.
To describe transport in the presence of such a source,
we will expand temperature variables in both layers in
a suitable orthonormal set of functions. Here it will be
convenient to use eigenstates of the operator ∂2
y with zero
Neumann boundary conditions at y = ±W/2, given by
(cid:19)
(cid:18) 2πn
W
(cid:18) 2π(n + 1
(cid:19)
2 )
y
,
W
un(y) = A cos
y
,
vn(y) = A sin
A = (2/W )1/2, n = 0, 1, 2... From the symmetry of the
source in Eq.(2) we expect δT1,2(y) to be odd in y. Thus
only the functions vn(y) are relevant, giving
δT1,2(y) =
δ T1,2(qn)A sin qny,
qn =
2π(n + 1
2 )
W
.
(cid:88)
qn
nκ1δ T1 + a(δ T1 − δ T2) + λδ T1 = Fn
q2
nκ2δ T2 + a(δ T2 − δ T1) + λδ T2 = 0
q2
and Fn = 2A(−1)nQ(1)
where κ1,2 = κ(1,2)
Eq.(4), we find the temperature profile in layer 2:
xx
yx j. Solving
(4)
δT2(y) =
aFn
L1(qn)L2(qn) − a2 vn(y),
(5)
(cid:88)
n≥0
where Li(qn) = κiq2
n + a + λ (i = 1, 2). Since electron-
lattice cooling is very slow [28, 29], with the correspond-
ing cooling length values in excess of few microns, we
will suppress λ in what follows. Because the boundaries
(2)
For each n we obtain a pair of algebraic equations
3
e2L11/T , and thermal conductivity is κ = L22/T 2. Com-
paring to the heat current due to an applied charge cur-
rent, Eq.(1), we identify L21 = −eQL11.
Onsager reciprocity demands that the cross-couplings
21(−B) where B is the applied magnetic
obey L12(B) = LT
field (note the transposed matrix). In an isotropic sys-
tem the off-diagonal components of L obey L(xy)(B) =
L(yx)(−B). As a result, Onsager reciprocity reduces to
L12(B) = L21(B)
(9)
in an isotropic system. Applying Eq. 9 to Eq. 8 in an
11 QL11∇T.
open circuit, we find E = −e−1∇µ = T −1L−1
For an isotropic system Q = Qxx1 + iQxyσ2, L = Lxx1 +
iLxyσ2, so that [Q, L] = 0, which gives Eq.(1).
Several different regimes arise depending on the rela-
tion between the interlayer cooling length ξ and the bar
width W . Using Eq.(3) and Eq.(1) we obtain
(cid:32)
(cid:33)
(cid:32)
V(cid:107)
VH
=
Rdrag(cid:107)
Rdrag
H
−Rdrag
Rdrag(cid:107)
H
,
(10)
(cid:33)(cid:32)
(cid:33)
I(cid:107)
0
−LG(ξ)
W T κ
giving the magnetodrag and Hall drag resistance values
−G(ξ)
T κ
Q(1)
Q(1)
Rdrag
H =
xy Q(2)
xx , Rdrag(cid:107) =
xy Q(2)
xy ,
(11)
where we used Qxx = Qyy and Qxy = −Qyx for an
isotropic system. For a narrow bar (or, slow cooling),
we have ξ/W (cid:29) 1 and G → 0, giving vanishingly small
H,(cid:107) . For a wide bar (or, fast cooling) we have G → 1
Rdrag
so that Rdrag
H,(cid:107) saturates to a universal value independent
of the interlayer cooling rate. For typical device param-
eters, we estimate ξ ≈ 40 nm at T = 300 K [20]. Since
L, W are a few mircons for typical graphene devices, we
expect them to be firmly in the G = 1 regime, with the
Hall drag and magnetodrag attaining universal values in-
dependent of the electron-electron interaction strength.
To describe the density and B field dependence, we use
a simple model for Q. Measurements indicate[8, 9] that
thermopower and the Nernst effect in graphene are well
described by the Mott formula [33], giving
(cid:32)
(cid:33)
Q =
π2
3e
k2
BT 2ρ
∂[ρ−1]
∂µ
,
ρ =
ρ(cid:107)
ρH
−ρH ρ(cid:107)
,
(12)
with ρ the resistivity, e < 0 the electron charge, and µ the
chemical potential. We use a simple phenomenological
model [34] relevant for classically weak B fields:
ρ0(cid:112)1 + n2/n2
0
ρ(cid:107) =
,
ρH =
−Bn
e(n2 + n2
0)
,
(13)
where ρ0 is the resistivity peak value at the Dirac point,
n is the carrier density, and parameter n0 accounts for
broadening of the Dirac point due to disorder. We ac-
count for disorder broadening of the density of states,
dn/dµ = (n2 + n2
0)1/4(cid:0)2/(π¯h2v2
F )(cid:1)1/2
.
FIG. 2:
(a) Schematic of charge current, temperature gra-
dients, and electric field in the two layers of a Hall bar that
produces Hall drag. (b,c) Density dependence of Hall drag
resistance, predicted from Eqs.(11),(13) for the same param-
eter values as in Fig.1. (d) Density dependence of Qxx, Qxy,
see text.
in the transverse (y-direction) are free, a finite temper-
ature imbalance between the edges can arise, given by
∆T = δT2(y = W/2) − δT2(y = −W/2). We find
∆T = 4A2(cid:88)
ξ = (cid:112)κ1κ2/aκ. We evaluate the sum using the iden-
tity(cid:80)∞
where we defined κ = κ1 + κ2 and a length scale
aQ(1)
yx j
L1L2 − a2 =
1 − tanhπc
n(1 + ξ2q2
q2
n)
(cid:88)
to obtain
Q(1)
yx j
(cid:16)
, (6)
n≥0
n≥0
W κ
8
1
n=0
(n+ 1
2 )4+c2(n+ 1
2 )2 = π2
2c2
πc
∆T =
W Q(1)
yx j
κ
G(ξ), G(ξ) = 1 − 2ξ
W
tanh
. (7)
(cid:17)
(cid:18) W
2ξ
(cid:19)
Connecting ∆T with the drag voltage, and in particu-
lar determining its sign, requires taking full account of
Onsager reciprocity. This analysis is presented below.
In the same way that the applied charge current, j,
in layer 1 causes a heat current (Peltier/Ettingshausen),
a temperature imbalance in layer 2, ∆T , can sustain
voltage drops across the sample (Thermopower/Nernst).
These two effects are related by Onsager reciprocity con-
straints. The cross couplings in the coupled energy and
charge transport equations [32] arise from
(cid:18) −j
(cid:19)
jq
=
(cid:18) eL11/T eL12
L21/T L22
(cid:19)(cid:32) ∇µ
(cid:33)
∇ 1
T
(8)
where L are 2 × 2 matrices and e is the carrier charge.
In this notation, the electrical conductivity equals σ =
H
= (W/L)Rdrag(cid:107)
From Eqs.(11),(12),(13) and the Wiedemann-Franz re-
lation for κ, we obtain ρdrag(cid:107)
and Rdrag
(see Fig.1b,c and Fig.2b,c, respectively).
In that, we
used the parameter values n0 = 1011 cm−2, ρ0 = h
3e2 ,
and a representative temperature, T = 150 K. These
values match device characteristics (disorder broadening,
n0, and peak resistivity, ρ0) described in Ref.[13]. As a
sanity check, we plot the components of Q (in Fig.2d)
which show the behavior near CN matching thermopower
and Nernst effects measured in graphene[8, 9].
Analyzing magnetodrag, we find that ρdrag(cid:107)
peaks at
dual CN, taking on large and negative values (Fig.1b,c).
Magnetodrag peak exhibits a steep B dependence,
ρdrag(cid:107),peak ∝ −B2, bearing a striking resemblance to mea-
surements reproduced in Fig.1d. In particular, our model
explains the negative sign of the measured magnetodrag.
Hall drag is large and sign-changing (see Fig.2b,c), tak-
ing on values consistent with measurements[35]. Interest-
ingly, the map in Fig.2b indicates that the sign of Rdrag
is controlled solely by carrier density in layer 2, breaking
the n1 ↔ n2 symmetry between layers. This behavior
does not contradict Onsager reciprocity, it arises as a
consequence of the asymmetric boundary conditions for
the Hall bar: free boundary at y = ±W/2 and ideal heat
sinks at the ends, δT (x = ±L/2) = 0. This allows for
finite temperature gradients to be sustained across the
bar but not along the bar, see Fig.2a.
H
For other geometries, the temperature gradient can
be obtained by balancing the heat flux due to thermal
conductivity against the net heat flux in the two lay-
ers, (κ1 + κ2)∇δT = D Q(1)j1 (see Eq.(24) of Ref.[22]).
The quantity D can in principle be obtained by solving
heat transport equations. Adopting the same approach
as above, we find a magneto and Hall-drag resistivity
ρdrag =
1
T κ
Q(2)D Q(1), E2 = ρdragj1.
(14)
clarify,
We wish to
in connection to
For isotropic heat flow, D = 1. In this case, since Q(1)
and Q(2) commute, the resulting drag is layer-symmetric,
n1 ↔ n2[22]. In particular, Hall drag for D = 1 vanishes
on the diagonal n1 = −n2. In contrast, for anisotropic
heat flow, such as that discussed above, we expect a
generic tensor D (cid:54)= 1 and thus no layer symmetry.
recent
measurements,[35] that layer symmetry n1 ↔ n2 implies
a swap of current and voltage contacts. Layer symme-
try, which implies D = 1 in Eq.(14), will therefore only
hold for Hall bars equipped with wide voltage contacts,
for which the contact and the bar widths are compa-
rable. This is indeed the case for the cross-shaped de-
vices used in Ref.[13]. However it is not the case for a
Hall bar with noninvasive voltage probes which are much
narrower than the bar width, as assumed in our analysis
above. Noninvasive probes, which have little effect on
4
temperature distribution in the electron system, trans-
late into D (cid:54)= 1 and no layer symmetry.
In summary, magnetic field has dramatic effect on drag
at CN because it induces strong coupling between neu-
tral and charge modes, which are completely decoupled in
the absence of magnetic field in a uniform system. Field-
induced mode coupling leads to giant drag that dwarfs
the conventional momentum drag contribution as well as
a remnant drag due to spatial inhomogeneity[20]. Our es-
timates indicate that these two contributions are orders
of magnitude smaller than the predicted magnetodrag,
which also has an opposite sign. The giant magnetodrag
and Hall drag values attained at classically weak mag-
netic fields, along with the unique density dependence
and sign, make these effects easy to identify in experi-
ment. The predicted magnetodrag is in good agreement
with findings in Ref.[13]. Magnetic field, coupled with
drag measurements at CN, provides a unique tool for
probing the neutral modes in graphene.
We acknowledge useful discussions with A. K. Geim,
P. Jarillo-Herrero, L. A. Ponomarenko, and financial sup-
port from the NSS program, Singapore (JS).
[1] J. Gonzales, F. Guinea, and M.A.H. Vozmediano, Nucl.
Phys. B424, 595 (1994); Phys. Rev. B 59, R2474 (1999).
[2] D. E. Sheehy and J. Schmalian, Phys. Rev. Lett. 99,
226803 (2007).
[3] D. T. Son, Phys. Rev. B 75, 235423 (2007).
[4] O. Vafek, Phys. Rev. Lett. 98, 216401 (2007).
[5] A. B. Kashuba, Phys. Rev. B 78, 085415 (2008).
[6] L. Fritz, J. Schmalian, M. Mueller, S. Sachdev, Phys.
Rev. B 78, 085416 (2008).
[7] M. Mueller, L. Fritz, S. Sachdev, Phys. Rev. B 78, 115406
(2008).
[8] Y. M. Zuev, W. Chang, and P. Kim, Phys. Rev. Lett.
102, 096807 (2009).
[9] P. Wei, W. Bao, Y. Pu, C. N. Lau, and J. Shi, Phys. Rev.
Lett. 102, 166808 (2009).
[10] J. G. Checkelsky and N. P. Ong, Phys. Rev. B 80,
081413(R) (2009).
[11] L. P. Pitaevskii, E.M. Lifshitz, Physical Kinetics (Perga-
mon, Oxford, 1981)
[12] L. Britnell, et al. Science 335, 947 (2012).
[13] R. V. Gorbachev, et al. Nature Physics 8, 896 (2012).
[14] S. Kim, I. Jo, J. Nah, Z. Yao, S. K. Banerjee, and E.
Tutuc, Phys. Rev. B 83, 161401 (2011).
[15] W.-K. Tse and S. Das Sarma, Phys. Rev. B 75, 045333
(2007).
[16] B. N. Narozhny, Phys. Rev. B 76, 153409 (2007).
[17] E. H. Hwang, R. Sensarma, and S. Das Sarma , Phys.
Rev. B 84, 245441 (2011).
[18] N. M. R. Peres, J. M. B. Lopes dos Santos, and A. H.
Castro Neto, Europhys. Lett. 95, 18001 (2011).
[19] M. I. Katsnelson, Phys. Rev. B 84, 041407 (2011).
[20] J. C. W. Song and L. S. Levitov, Phys. Rev. Lett. 109,
236602 (2012).
[21] B. N. Narozhny, M. Titov, I. V. Gornyi, and P. M. Os-
trovsky, Phys. Rev. B 85, 195421 (2012).
[22] J. C. W. Song, D. A. Abanin, L. S. Levitov, Nano Lett.,
5
10.1021/nl401475u (2013).
McEuen, Nature Physics, 9 103 (2013)
[23] N. K. Patel, E. H. Linfield, K. M. Brown, M. Pepper, D.
A. Ritchie, and G. A. C. Jones, Semicond. Sci. Technol.,
12, 309 (1997).
[29] A. C. Betz, S. H. Jhang, E. Pallecchi, R. Ferreira, G.
F´eve, J-M. Berroir, and B. Placais, Nature Physics, 9
109 (2013).
[24] H. Rubel, A. Fischer, W. Dietsche, K. von Klitzing, and
[30] J. M. Ziman, Principles of the Theory of Solids, Cam-
K. Eberl, Phys. Rev. Lett. 78 1763 (1997).
bridge University Press (1979).
[25] M. P. Lilly, J. P. Eisenstein, L. N. Pfeiffer, and K. W.
[31] E. H. Hwang, E. Rossi, and S. Das Sarma Phys. Rev. B
West, Phys. Rev. Lett., 80 1714 (1998).
80, 235415 (2009).
[26] A. Kamenev and Y. Oreg, Phys. Rev. B, 52 7516 (1995).
[27] M. C. Bonsager, K. Flensberg, B. Y.-K. Hu, and A.-P.
Jauho, Phys. Rev. Lett., 77 1366 (1996). The authors
note that Hall drag vanished for systems with inversion
symmetry and when the carriers in the active layer can
be described by a drifted Fermi-Dirac distribution.
[28] M. W. Graham, S-F. Shi, D. C. Ralph, J. W. Park, P. L.
[32] H. B. Callen, Phys. Rev., 73 1349 (1948).
[33] M. Jonson and S. M. Girvin, Phys. Rev. B, 29 1939
(1984).
[34] D. A. Abanin, et. al., Science, 332 328 (2011) and ac-
companying online supplement.
[35] A. K. Geim, private communication
|
1005.4780 | 1 | 1005 | 2010-05-26T10:22:08 | Power laws in surface state LDOS oscillations near a step edge | [
"cond-mat.mes-hall"
] | In this paper we indicate a general method to calculate the power law that governs how electronic LDOS oscillations decay far away from a surface step edge (or any local linear barrier), in the energy range when only 2D surface states are relevant. We identify the critical aspects of the 2D surface state band structure that contribute to these decaying oscillations and illustrate our derived formula with actual examples. | cond-mat.mes-hall | cond-mat |
Power laws in surface state LDOS oscillations near a step edge
Rudro R. Biswas1,3∗ and Alexander V. Balatsky2,3
1Department of Physics, Harvard University, Cambridge, MA 02138
2Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545
3Center for Integrated Nanotechnologies, Los Alamos National Laboratory, Los Alamos, NM 87545
(Dated: May 25, 2018)
In this paper we indicate a general method to calculate the power law that governs how electronic
LDOS oscillations decay far away from a surface step edge (or any local linear barrier), in the energy
range when only 2D surface states are relevant. We identify the critical aspects of the 2D surface
state band structure that contribute to these decaying oscillations and illustrate our derived formula
with actual examples.
PACS numbers: 03.65.Nk, 07.79.Cz, 73.20.-r, 73.20.At
I.
INTRODUCTION
For over a couple of decades now Scanning-Tunneling
Microscopy (STM) experiments have been used to ob-
serve the effects of perturbations to electronic surface
states, in the form of atomic defects, corrals and step
edges1–4. The quantum electronic response to defects
can give us basic information about the band structure
of the scattered quasiparticles5; it also encodes informa-
tion about their nature and this is useful when probing
correlated phases2,6–8. In this paper, we shall consider
the case of a step edge on a 2D surface and calculate the
spatial decay of standing waves created in the surface
LDOS, far away from the step edge. We shall show, by
a simple process of power counting, that the geometry of
the constant energy cut of the quasiparticle band struc-
ture and a qualitative knowledge of the character of the
quasiparticle wavefunctions provide enough information
to pin down the power with which the LDOS oscillations
decay far away from the step edge.
II. THEORY
We begin by considering a general band structure for
the 2D surface states, whose cross-section at the energy
of observation Eobs is shown in Figure 1. Also shown in
the figure is the orientation of the surface step edge -
parallel to the y-direction. Because of the conservation of
momentum parallel to the edge during a scattering pro-
cess, the incoming and outgoing states must be connected
by straight lines perpendicular to the step edge. Some
such processes are also marked in Figure 1. The arrows
joining the initial and final states denote the wave-vector
of LDOS oscillations that particular scattering process
would give rise to. Obtaining the total LDOS involves
summing up these oscillations. The most coherent con-
tributions to this sum come from regions where the scat-
tering wave-vectors change the slowest as we move par-
allel to the step edge, i.e, changing only the ky of the
scattering states. We denote the 'identifying' scattering
wave vector in each such region as the 'characteristic'
wave vector of that region.
As a very common example, for a circular constant
energy cut as in a 2DEG (Figure 2), the most coherent
contributions come from the scatterings around the di-
ameter - the characteristic wave vector in this case is
thus the diameter, ∆0.
The new electronic LDOS far away from the step edge
is now provided by (below, 'new' refers to the new energy
eigenstates while 'init' and 'fin' refer to the initial and
FIG. 1: (Color online) A constant energy cut of a generic
2D electronic surface state (SS) band structure, taken at the
energy Eobs of an STM probe; on the right is shown the ori-
entation of the surface step edge (or any linear barrier) in
question. Quasiparticles are scattered 'horizontally', preserv-
ing ky. The regions where the scattering wave-vectors vary
the slowest (locally) with ky are shaded – blue for incident
and pink for reflected states. The 'characteristic' scattering
wave vectors are also indicated by arrows and numbers.
final scattering states, respectively):
ψnew2
E
(cid:88)
(cid:88)
(cid:122)
(cid:88)
E
E
ρ(x, E) =
=
=
ψinit + rψfin2 + transmitted from other side
x−independent part
(cid:16)ψinit2 + rψfin2 + transmitted
(cid:105)
(cid:125)(cid:124)
(cid:88)
(cid:104)
+ 2
Re
r ψ
†
init · ψfin
(cid:123)
(cid:17)
E
(1)
Writing the energy-momentum eigenstates as ψk(x) =
χeik·x, where χ denotes an 'internal' part involving the
spin and other internal components, the x-dependent
part of the LDOS can be summarized as
δρ(x, E) ∝(cid:88)(cid:90)
(cid:104)
(cid:16)
†
f · χi
χ
(cid:17)
ei∆kxx(cid:105)
(2)
dky ρ0(ky)Re
r(ky)
0
The sum is over the various regions of coherent scatter-
ing, each one corresponding to a characteristic scattering
vector. The outer limits of these integrals are not impor-
tant as the oscillations there de-cohere rapidly. ρ0(ky)
is a DOS factor (it multiplicatively converts the measure
dky to a product of the length of the band curve enclosed
between ky and ky + dky and the DOS in that region).
The x-dependence of a characteristic oscillation far
away from the step edge may be found from the
above expression by writing down the lowest order ky-
dependencies of the relevant quantities near each charac-
teristic wave-vector (δky is the ky-displacement from the
associated characteristic wave vector):
ρ0(ky) ∼ ρ0δkα
r(ky) ∼ r0δkβ
†
f · χi ∼ δkγ
χ
∆kx ∼ ∆0 + ∆1δkη
(3)
y
y
y
y
(cid:104)
r0 ei(∆0x+∆1µ)(cid:105)
α+β+γ+1
η
Re
Changing the integration variable δky to the variable µ =
δkη
y x in (2) and using (3), we obtain our central result
δρ(x, E) ∝(cid:88) ρ0
(cid:90) dµ
∼(cid:88)ρ0r0 sin(∆0x + φ)
α+β+γ+1
µ
µ
x
η
x(α+β+γ+1)/η
(4)
This asymptotic behavior is correct for x (cid:29) (∆k)−1,
where ∆k is the characteristic size of the region in mo-
mentum space where the scaling laws (3) hold. The
power of decay of the oscillations coming from each char-
acteristic region is thus given by (α + β + γ + 1)/η, which
may be evaluated using our knowledge of the band struc-
ture in that region. Note, however, that we haven't been
able to evaluate the total strength of the scattering pro-
cess which requires a much more detailed calculation in-
cluding evaluating the reflection amplitudes themselves.
This exercise provides us with the combination of possi-
ble power laws we can try to fit actual experimental data
to, if an idea of the band structure exists.
2
III. EXAMPLES
A. 2DEG1
The scattering wave-vector varies slowest around the
equator (see Figure 2). Thus, ∆0 = diameter of circle
= 2kE, α = β = γ = 0 (assuming, quite reasonably, that
the reflection amplitude is nonzero and smooth across
normal scattering). Also, from the geometry of the band,
we get η = 2. Using (4), this tells us that:
δρ(x, E) ∼ sin(∆0x + φ)
x1/2
(∆0 = 2kE)
(5)
FIG. 2: (Color online) The characteristic wave-vector ∆0 in
the case of a circular band (for 2DEGs with rotational invari-
ance)
B. Strong Topological Insulator (circular band cut)
1. Generic barrier
This case is illustrated in Figure 3 and is realized for
the gapless surface states in Strong Topological Insula-
tors like Bi2Se3 and Bi2Te3 (at energies near the Dirac
point). The scattering wave-vector varies slowest around
the equator (as in the 2DEG case), where ∆0=diameter
of circle, α = 0, β = 1 since the reflection coefficient
changes sign9 as one crosses the diameter/case of normal
reflection (can be any odd power; should be linear gener-
ically), γ = 1 because the spins are exactly antiparallel
for scattering states at the diameter and thus the lowest
order overlap is linear in δky, and η = 2 as in the 2DEG
case. This gives rise to:
δρ(x, E) ∼ sin(∆0x + φ)
x3/2
(∆0 = 2kE)
(6)
This result agrees with numerical calculations for partic-
ular cases of the model describing the step edge10.
FIG. 3: (Color online) Constant energy cut of a circular sur-
face band on a STI surface. The spins, indicated as block
arrows, are antiparallel for normal scattering (the character-
istic scattering process for the STI surface state band that is
circular) – the spin overlap magnitude is thus generically a
linear function of the angle of incidence. The same may be
said for the reflection amplitude magnitude which is a certain
gauge is antisymmetric in the angle of incidence.
2.
'Perfect' reflection
The scattering wave-vector varies slowest around the
equator (as in the 2DEG case), where ∆0=diameter of
circle, α = β = 09 (since the reflection amplitude is con-
stant in magnitude near normal incidence), γ = 1 as
argued for the previous case, and η = 2. This gives rise
to:
δρ(x, E) ∼ sin(∆0x + φ)
x
(∆0 = 2kE)
(7)
In the actual case, there will always be a region near
normal incidence where the reflection amplitude will be-
come linear (because it is antisymmetric). This means
that 'very' far away ∼ the inverse of the ky-span of the
region where r is linear, the previous result (6) for the
generic barrier should hold.
C. Bi2Te3 (with hexagonal warping)3
If we are far away from the Dirac point, the surface
band of Bi2Te3 exhibits hexagonal warping11. The fol-
lowing results hold when the STM bias maintains our
observation energy in that regime.
1. Step edge ⊥ ΓM direction
The scattering wave-vector varies slowest around the
equator (as in the 2DEG case), where ∆0=diameter of
3
circle, α = 0, β = 1, γ = 1 and η = 2 exactly as argued
before for the circular STI band. However, the extent of
this region is very small and the scattering is found to be
dominated by processes connecting the hexagonal 'cor-
ners' (marked by bold arrow in Figure 4), with a charac-
3. For the latter case, since
teristic scattering vector knest
the spin states have a finite overlap with each other at
the hexagon corners11, we have γ = 0. Also, assuming
that the reflection coefficient is smooth for the relevant
scattering processes, β = 0. Finally, α = 0 (DOS is finite
and smooth) and the overwhelmingly 'linear' nature of
the bands yield η = 1. Putting these together, we obtain
the observed variation3
δρ(x, E) ∼ sin(knestx + φ)
x
(8)
FIG. 4: (Color online) Scattering processes from a step edge
on Bi2Te3 oriented perpendicular to the ΓM direction, in the
energy range where the band exhibits hexagonal warping. The
scattering vector varies linearly near knest, as indicated by the
angle made by the dotted lines, leading to η = 1. The weaker
characteristic scattering process is denoted by the dotted ar-
row. Spins are indicated as block arrows.
2. Step edge ⊥ ΓK direction
The wave-vectors vary slowest around the equator (as
in the 2DEG case) and from the considerations of the
cicular STI surface band above, we can conclude that
there should be characteristic oscillations at 2kΓK decay-
ing as x−3/2 (or 1/x for a 'perfect' reflector).
In this
case, because of the larger extent of the characteristic
scattering region around the diameter, these oscillations
may be strong and observable. Because of the presence
of the linear band shape with larger spectral presence
and reflection strengths near the corners, we can also
observe LDOS oscillations from the corner→corner scat-
tering processes, decaying as 1/x. Of course, from our
simple calculation we cannot reliably predict which of
the two processes discussed above have the stronger sig-
nature.
4
over the long-distance behavior) for features at the 'char-
acteristic' scattering vectors. The FT near those points
can then be fitted to the abovementioned power laws (or
a logarithm, for the case of a 1/x decay) to recover the
spatial decay power laws.
V. CONCLUSION
We have outlined a method to calculate the possible
oscillatory power laws governing the decay of LDOS per-
turbations next to a step edge or some such linear barrier
on a surface, in the energy range when only electronic
surface states are relevant and the surface band struc-
ture is qualitatively known. To find these laws we need
to identify the characteristic scattering regions (Figure
1), compute the scaling powers of the relevant quantities
(3) and from that obtain the possible oscillatory powers
(4).
Acknowledgments
This work was supported by the US DOE thorough
BES and LDRD and by the University of California
UCOP program T027-09. We would also like to acknowl-
edge illuminating discussions with M. Crommie, D. Hal-
dane, H. Manoharan and members of M. Z. Hasan and
A. Yazdani's research groups.
FIG. 5: (Color online) Scattering processes from a step edge
on Bi2Te3 oriented perpendicular to the ΓK direction, in the
energy range where the band exhibits hexagonal warping.
Spins are indicated as block arrows.
IV.
ISOLATING CONTRIBUTIONS USING THE
1-D FOURIER TRANSFORM
The Fourier Transform of the LDOS data may be used
to observe signatures from more than one set of scattering
processes. For oscillations decaying as sin(Kx + φ)/xn,
scaling analysis tells us that the Fourier transform looks
like F (k) ∼ k ∓ Kn−1, when k ∼ ±K. Thus, one way
to look for contributions to these oscillations would be
to scan the 1-D Fourier transform of the LDOS (taken
∗ Electronic address: rrbiswas@physics.harvard.edu
1 M. F. Crommie, C. P. Lutz, and D. M. Eigler, Nature 363,
524 (1993).
2 A. Yazdani, B. A. Jones, C. P. Lutz, M. F. Crommie, and
D. M. Eigler, Science 275, 1767 (1997).
3 Z. Alpichshev, J. G. Analytis, J.-H. Chu, I. R. Fisher, Y. L.
Chen, Z. X. Shen, A. Fang, and A. Kapitulnik, Phys. Rev.
Lett. 104, 016401 (2010).
4 A. Richardella, P. Roushan, S. Mack, B. Zhou, D. A.
Huse, D. D. Awschalom, and A. Yazdani, Science 327,
665 (2010).
5 L. Petersen, P. T. Sprunger, P. Hofmann, E. Laegsgaard,
B. G. Briner, M. Doering, H.-P. Rust, A. M. Bradshaw,
F. Besenbacher, and E. W. Plummer, Phys. Rev. B 57,
R6858 (1998).
6 J. E. Hoffman, E. W. Hudson, K. M. Lang, V. Madhavan,
H. Eisaki, S. Uchida, and J. C. Davis, Science 295, 466
(2002).
7 J. E. Hoffman, K. McElroy, D. H. Lee, K. M. Lang,
H. Eisaki, S. Uchida, and J. C. Davis, Science 297, 1148
(2002).
8 S. H. Pan, E. W. Hudson, K. M. Lang, H. Eisaki, S. Uchida,
and J. C. Davis, Nature 403, 746 (2000).
9 R. R. Biswas and A. V. Balatsky, arXiv:0912.4477 (2009).
10 X. Zhou, C. Fang, W.-F. Tsai, and J. Hu, Phys. Rev. B
80, 245317 (2009).
11 L. Fu, Phys. Rev. Lett. 103, 266801 (2009).
|
1104.3104 | 1 | 1104 | 2011-04-15T16:28:18 | Observation of Intra- and Inter-band Transitions in the Optical Response of Graphene | [
"cond-mat.mes-hall",
"cond-mat.mtrl-sci"
] | The optical conductivity of freely suspended graphene was examined under non-equilibrium conditions using femtosecond pump-probe spectroscopy. We observed a conductivity transient that varied strongly with the electronic temperature, exhibiting a crossover from enhanced to decreased absorbance with increasing pump fluence. The response arises from a combination of bleaching of the inter-band transitions by Pauli blocking and induced absorption from the intra-band transitions of the carriers. The latter dominates at low electronic temperature, but, despite an increase in Drude scattering rate, is overwhelmed by the former at high electronic temperature. The time-evolution of the optical conductivity in all regimes can described in terms of a time-varying electronic temperature. | cond-mat.mes-hall | cond-mat | Observation of Intra- and Inter-band Transitions in the Optical Response of Graphene
Leandro M. Malard1, Kin Fai Mak1, A. H. Castro Neto2,3, N. M. R. Peres4, and
Tony F. Heinz1
1Departments of Physics and Electrical Engineering, Columbia University, 538 West 120th Street, New York, New York
10027, USA
2Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, Massachusetts 02215, USA
3Graphene Research Centre, National University of Singapore, 2 Science Drive 3, 117542, Singapore
4Departamento de Física e Centro de Física, Universidade do Minho, P-4710-057 Braga, Portugal
(April 15, 2011)
The optical conductivity of freely suspended graphene was examined under non-equilibrium conditions
using femtosecond pump-probe spectroscopy. We observed a conductivity transient that varied strongly
with the electronic temperature, exhibiting a crossover from enhanced to decreased absorbance with
increasing pump fluence. The response arises from a combination of bleaching of the inter-band transitions
by Pauli blocking and induced absorption from the intra-band transitions of the carriers. The latter
dominates at low electronic temperature, but, despite an increase in Drude scattering rate, is overwhelmed
by the former at high electronic temperature. The time-evolution of the optical conductivity in all regimes
can described in terms of a time-varying electronic temperature.
PACS: 78.67.Wj, 78.47.D-, 78.47.jb
The optical properties of graphene have been the
subject of much attention [1-3]. Interest in this topic is
generated both by the insight that the optical response of
graphene provides into the nature of its excited states and
their interactions and by the importance of understanding
graphene’s optical
response
for emerging photonic
applications [2, 4]. The response of graphene to excitation
by femtosecond laser pulses has attracted particular interest
[5-11]. Probing dynamics on the femtosecond time scale
provides
information about electron-electron, electron-
phonon, and phonon-phonon interactions and also has
important
implications
for
the behavior of
recently
developed ultrafast photonic devices [2, 4].
To date all ultrafast pump-probe measurements of
graphene have observed an induced bleaching of the
absorption under optical excitation [7-9]. This behavior
arises from Pauli blocking of the strong inter-band optical
transitions [7-9] and is expected for electronic temperatures
sufficiently high to induced filling of the states. The optical
response of graphene arises, however,
from
two
fundamental processes: inter-band and intra-band optical
transitions [3, 12-14]. Here we report observation of a
dominant intra-band contribution to the transient optical
response, with a corresponding enhanced absorption. We
this effect
identify
through probing freely suspended
graphene samples at relatively low pump fluence. With
increasing pump fluence, the transient optical response
exhibits a crossover from enhanced absorption to bleaching
as Pauli blocking of inter-band transitions becomes more
prominent.
We explain the observed temporal evolution of the
optical absorption for all pump fluences within
the
framework of a thermalized electronic energy distribution
that is in equilibrium with a set of strongly coupled optical
phonons. Cooling of this subsystem, and the decrease in
electronic temperature, is controlled by energy loss of the
strongly coupled optical phonons to lower energy phonons
through anharmonic decay. A fluence-independent time
constant of 1.4 ps is deduced, which is consistent with
recent time-resolved Raman scattering measurements of the
cooling of zone-center optical phonons [15]. In addition to
the importance of these measurements for understanding
ultrafast carrier dynamics and applications in ultrafast
photonics,
response at high carrier
intra-band
the
temperature provides insight into carrier dynamics in a
regime relevant to high-field charge transport in graphene
[16-19]. In particular, our experiments reveal a sharp
increase in the Drude scattering rate with increasing
temperature of the electrons and optical phonons, as also
suggested by high-field electrical transport measurements in
carbon nanotubes [20, 21] and graphene [16-19]. Our
results shown to be compatible with the behavior expected if
electron-optical phonon scattering plays a dominant role in
determining
the carrier scattering rate
in
this high
temperature regime.
In our experimental study we made use of freely
suspended graphene samples. Such samples allow us to
eliminate both accidental doping effects and energy transfer
associated with the substrate. We prepared exfoliated
graphene samples that were freely suspended over trenches
patterned in transparent SiO2 substrates [22]. The substrates,
with trenches of width 4 – 5 µm and depth of ~3 µm (as
characterized by atomic force microscopy), were carefully
cleaned by chemical etching (nanostrip) before we deposited
graphene by mechanical exfoliation from kish graphite.
Areas of graphene of single-layer thickness were identified
by optical microscopy and further characterized by Raman
spectroscopy [22]. The low doping levels of the samples
were reflected in the large (~15 cm-1) width of the G-mode
and characteristic asymmetry of the 2D-mode, while the low
level of defects was indicated by the absence of the
disorder-induced D-mode.
The optical pump-probe measurements were
performed using a modelocked Ti-sapphire laser producing
pulses of ~300-fs duration at an 80-MHz repetition rate. The
800-nm wavelength output of the laser provided the probe
pulses, while radiation at 400 nm, obtained by frequency
doubling in a β-barium borate (BBO) crystal, served as the
pump excitation. Both the pump and probe beams were
focused onto the samples with a single 40× objective to
yield Gaussian spots of comparable widths of ~1.5 µm
(FWHM). The absorbed pump fluence was determined
using the absorbance of graphene at 400 nm (1.8πα = 4.14
% [23]). To account for the spatial variation of the pump
and probe beams, the effective fluence was determined by
weighting the absorbed probe fluence using the spatial
1
profile of the probe beam. Effective absorbed fluences F
4.0 µJ/cm2 were
between
0.5
and
investigated
experimentally. We note that over this range of pump
fluences, saturation of absorbance in graphene at the pump
wavelength
pump-probe
[4]. The
negligible
is
measurements were performed by modulating the pump
laser at 20 kHz and detecting the synchronous change in
probe transmission. The induced modulation of the probe
beam for the lowest pump fluence was ~10-6. For a
suspended thin film of material like our graphene sample,
the fractional change in transmittance,
, is given by
the change in the sample absorbance
. The later is
proportional to the change in the real part of the optical
, according
sheet conductivity of graphene,
to
[14]. We can
therefore directly
convert our experimentally observed change in transmission
into a fundamental material property of the graphene,
namely,
.
Figure 1(a) shows the measured transient response
of graphene for a comparatively low absorbed pump fluence
of F = 0.5 µJ/cm2. An increase in the optical conductivity
(enhanced absorption) is seen. The rise time of this
response is comparable to our experimental time resolution,
while the relaxation can be fit to a single exponential decay
with a time constant of τexp = 3.1 ps.
FIG. 1 (color online): Measured dynamics (closed circles) of
the transient optical conductivity of graphene for excitation
with different absorbed pump fluences F: (a) 0.50 µJ/cm2,
(b) 2.03 µJ/cm2, (c) 3.04 µJ/cm2, and (d) 4.06 µJ/cm2. The
red curves are fits based on the model described in the text,
which include both intra- and inter-band contributions to the
optical response.
To analyze our data, we consider models for the
optical conductivity of graphene with the electronic system
described by a Fermi-Dirac distribution at (electronic)
temperature T. The assumption of a thermalized energy
distribution for the electronic excitations is justified by the
time resolution of >100 fs in our measurements. On this
time scale, the excited electrons and holes are equilibrated
with one another, as well with the strongly coupled optical
phonons (near the Γ- and K-points) [5, 24].
The predicted change in the optical conductivity of
graphene under excitation arises from both intra- and inter-
band contributions,
. For the
case of thermalized electronic energy distributions relevant
in our case, the optical response of graphene has been
examined in several theoretical investigations [3, 12, 13].
Within a picture of non-interacting electrons, the induced
change in the real part of the optical sheet conductivity
follows directly from the corresponding changes in the intra-
and inter-band terms:
(1)
Here
denotes the photon energy of the probe beam, and
Γ is the carrier scattering rate. For our suspended graphene
samples, we have negligible doping and take the chemical
potential to be at the Dirac point. We also neglect any
transient separation of the chemical potentials for the
electrons and holes. While a dynamical separation of the
chemical potentials is common in semiconductors under
ultrafast excitation, on the relevant time scale the effect is
expected to be insignificant in graphene because of the
presence of rapid Auger processes for this zero-gap material
[4, 11].
The photo-induced
the optical
in
change
conductivity in Eqn. (1) from intra-band optical transitions,
, is positive in sign. The enhanced conductivity
arises from the presence of additional free carriers after
optical excitation and scales linearly with the electronic
temperature T. The inter-band term
yields a
negative contribution to the conductivity, a bleaching of the
absorption, from Pauli blocking of the strong inter-band
transitions at the photon energy
. In our regime of
>> kBT , the fractional change in the inter-
band conductivity is slight and varies nearly exponentially
in T. Accordingly, the intra-band contribution dominates at
comparatively low temperatures (low fluence), but, as
shown in Fig. 2, is overwhelmed by the inter-band terms as
the temperature increases. Thus the transient increase in
absorption seen in Fig. 1(a) for low pump fluence reflects
the change in intra-band optical response.
In order to apply the analysis of graphene optical
response presented in Eqn. (1) to our experiment, we
introduce a simple model to describe the temporal evolution
of the electron temperature T(t) induced by the pump laser.
As we shall see, this treatment predicts transient optical
conductivities compatible both with the low-fluence data
already presented and the high-fluence response discussed
below. To describe T(t), we first note that on the time scale
of our measurement, the deposited energy from the pump
pulses rapidly equilibrates among the electronic excitations
and the strongly coupled optical phonons [5, 24]. Energy
then leaves this coupled subsystem through anharmonic
decay of the phonons on the picosecond time scale, which
2
we describe phenomenologically by an effective phonon
lifetime τph [25]. To determine the resulting electronic
temperature T(t) we need to know the heat capacity of the
sub-system consisting of the electrons and the strongly
coupled phonons. For this purpose, we use an Einstein
model consisting of two phonon modes, the Γ- and K-point
phonons. Since only a small part of the phonon branch gets
populated, we considered only a fraction of the graphene
Brillouin zone to obtain the phonon heat capacity. The
fraction is determined by comparing results from [24] and
the initial condition is fixed by the absorbed fluence F. (For
details of this analysis, please refer to the Auxiliary
Materials [26].) The electronic contribution to the heat
capacity is minor and is neglected. Using this heat capacity,
we the find T(t) for any given excitation condition and,
using Eqn. (1), then obtain the temporal evolution of the
optical conductivity. In fitting our results, we treating the
phonon lifetime τph and the Drude scattering rate Γ as
adjustable parameters. As we discuss below, we use an
effective scattering rate Γ that depends on the pump fluence,
but is independent of time, to account for the variation of the
scattering rate with the temperature of the electrons and
optical phonons.
Fig. 2 (color online): Change in the optical conductivity of
graphene as a function of electronic temperature calculated
from Eq. (1). Black and red curves show the expected
contributions from intra- and inter-band terms, respectively.
The blue curve shows the total response.
The measured
the optical
time evolution of
conductivity in Fig. 1(a) is reproduced well with this model.
We use an effective scattering rate of Γ = 27 meV and a
phonon lifetime of τph =1.4 ps. This lifetime differs
significantly from time constant (τexp = 3.1 ps) of the
conductivity transient, a difference that reflects the strong
temperature dependence of the phonon heat capacity: The
electronic temperature, which tracks that of the strongly
coupled optical phonons and is described by τexp, falls more
slowly than the energy content of the optical phonons,
which is characterized by τph. The inferred phonon lifetime
of τph =1.4 ps, we note, lies between that obtained by time-
resolved Raman scattering measurements for graphite (2.2
ps) [24] and for graphene (1.2 ps) [15]. The slightly longer
phonon lifetime compared to that measured for graphene
[15] is attributed to the fact that our graphene samples are
suspended, thus eliminating
possible decay channels
involving the generation with surface polar phonons in the
substrate [16, 17, 19].
At higher pump fluences (and, correspondingly,
higher electronic temperatures), the optical conductivity is
expected to decrease as state filling of the inter-band
transitions begins to dominate the transient response (Fig.
2). The measured conductivity transients at higher pump
fluence are shown in Fig. 1 (b-d). With increasing pump
fluence, the initial response is indeed negative. At later
times when the electronic temperature drops, the intra-band
contribution becomes dominant and a positive conductivity
change is observed.
The experimental data of Fig. 1(b-d) can be fit (red
curves) with the same model of the electronic temperature
described above, but now with the inclusion of the inter-
band optical response of graphene. Thus model can then
accurately describe
the dynamics of
the conductivity
transients at all fluences [Fig. 1(a-d)]. In this fitting process,
we use a single, fluence-independent parameter of τph =1.4
ps for the optical phonon lifetime. This parameter reflects
the anharmonic decay rate of these optical phonons. It
should not increase with pump fluence unless the excitation
of the resulting phonon decay modes also increased
significantly [27], which is not expected this regime. On the
other hand, the fitting procedure implies that the effective
carrier scattering rate Γ increases with pump fluence. The
inferred variation of Γ is shown in Fig. 3 as a function of
absorbed fluence (bottom scale) and of the peak electronic
temperature (top scale) in our model. At high fluence,
corresponding to a peak temperature of 1,700 K, we find Γ
= 52 meV. Both this rate and that inferred for lower fluences
are significantly higher than the scattering rates implied by
conventional transport measurements [28] obtained for a
more modest temperature range.
To understand the origin of the enhanced scattering
rate, recall that the electron-phonon coupling in graphene is
especially strong for the optical phonon modes located at the
center and edge of the Brillouin zone [5, 25]. This coupling
is not, however, important for conventional low-bias charge
transport measurements at room
temperature. In
this
conventional regime, the optical phonons are not thermally
excited and cannot be absorbed in a scattering event, while
the electron energy is too low to scatter through the
emission of optical phonons. This situation changes when
the temperature of the electrons and optical phonons
becomes comparable to that characterizing the optical
phonon,
[20, 21]. Electron scattering with optical
phonons is then allowed and the contribution of this
scattering process to the temperature dependence of the
Drude is significant [29].
To obtain specific predictions for the increased
Drude scattering rate Γ from electron – optical phonon
interactions, we consider the zone-center and zone-edge
phonons to be dispersionless branches with energies of 200
meV and 150 meV, respectively. We then can then
calculate (see Auxiliary Materials [26]), the expected
temperature dependence of Γ from electron- optical phonon
scattering. As shown in Fig. 3, this contribution increases
nearly linearly with T for the relevant temperatures. By
using published values for the electron-phonon coupling for
both the zone-center [30] and zone-edge [31] optical
phonons, we obtain semi-quantitative agreement with the
experimental results for Γ as a function of pump fluence.
Better agreement is not anticipated, since the treatment
involves several significant simplifications. We first note
that the deduced effective Drude scattering rate Γ actually
3
corresponds to fitting the response over a range of electronic
temperatures. Here we assume that the result is dominated
by the behavior near the peak electronic temperature. More
elaborate modeling could take this effect into account. In
terms of
the optical
the underlying description of
conductivity at high temperature, our treatment could be
improved by a more detailed analysis of the different
phonon modes, considered here as two dispersionless
the scattering process. More
in
branches,
involved
examine possible
should
fundamentally, one
also
contributions
to Γ
from electron-electron
scattering
processes [32], which are known to influence the dc
conductivity [33]. Finally, at high electronic temperatures
(T > 1600K), the experimental values for Γ level off with
increasing T. This effect is absent in the model and suggests
that screening of the electron interactions [34] may play a
significant role in this regime of high carrier densities.
Fig. 3 (color online): Drude scattering rate
(closed
circles) as a function of absorbed pump fluence (bottom
scale) and the inferred peak electronic temperature (top
scale). The points are inferred from modeling of the
experimental transient optical response of the graphene
sample. The black line corresponds to the predicted
temperature dependence of
arising from electron- optical
phonon interactions, as described in the text and in the
Auxiliary Materials [26].
An interesting point of comparison for our results
is with those obtained in high-field dc transport studies. We
find that the values for Γ deduced here for high electronic
temperatures are also broadly consistent with those obtained
in high-field transport measurements in metallic carbon
nanotubes at current saturation [20, 21]. In this regime,
electron- optical phonon scattering is also considered to be
the dominant contribution to Γ. For high-bias transport
measurements in graphene [16, 17, 19], the carrier scattering
rate is also significantly enhanced compared to the low-field
behavior. This effect has, however, been attributed to
interaction with of the carriers with polar phonons in the
substrate on which the graphene is deposited. Our results
imply that even for suspended graphene, which lacks these
substrate-mediated interactions, the scattering rate will be
strongly enhanced at elevated electronic temperatures. This
effect will places a fundamental limit on current flow in
graphene at electrical high bias.
In conclusion, we have examined ultrafast carrier
dynamics in freely suspended graphene samples by optical
pump-probe measurements. A crossover of the graphene
optical conductivity transients from enhanced to decreased
absorption is observed as the pump fluence is increased.
This behavior can be understood as the result of the
existence of both intra- and inter-band contributions to the
optical response of graphene. Analysis of the data implies an
optical phonon lifetime of 1.4 ps and enhanced carrier
scattering rates at high electronic temperatures. Our
experiment demonstrates the importance of free carrier
absorption in the visible spectral range for graphene under
non-equilibrium conditions and opens up new opportunities
for probing fundamental charge transport properties of this
2-dimensional system by means of optical measurements.
We thank Drs. Y. Wu, M. Koshino, D. Song and
H. Yan for fruitful discussions and C. H. Lui for sample
preparation. We acknowledge support from the Nanoscale
Science and Engineering Initiative of the National Science
Foundation and from the MURI program of the Air Force
Office of Scientific Research. AHCN acknowledges the
partial support of the U.S. DOE under grant DE-FG02-
08ER46512, and ONR grant MURI N00014-09-1-1063. LM
acknowledges support from the CNPq program in Brazil.
[1] A. H. Castro Neto et al., Rev. Mod. Phys. 81, 109 (2009).
[2] F. Bonaccorso et al., Nat. Photonics 4, 611 (2010).
[3] N. M. R. Peres, Rev. Mod. Phys. 82, 2673 (2010).
[4] Z. P. Sun et al., ACS Nano 4, 803 (2010).
[5] T. Kampfrath et al., Phys. Rev. Lett. 95, 187403 (2005).
[6] M. Breusing, C. Ropers, and T. Elsaesser, Phys. Rev. Lett. 102,
086809 (2009).
[7] P. A. George et al., Nano Lett. 8, 4248 (2008).
[8] D. Sun et al., Phys. Rev. Lett. 101, 157402 (2008).
[9] R. W. Newson et al., Opt. Express 17, 2326 (2009).
[10] C. H. Lui et al., Phys. Rev. Lett. 105, 127404 (2010).
[11] T. Winzer, A. Knorr, and E. Malic, Nano Lett. 10, 4839
(2010).
[12] N. M. R. Peres, F. Guinea, and A. H. Castro Neto, Phys. Rev.
B 73, 125411 (2006).
[13] L. A. Falkovsky, and A. A. Varlamov, Eur. Phys. J. B 56, 281
(2007).
[14] K. F. Mak et al., Phys. Rev. Lett. 101, 196405 (2008).
[15] K. Kang et al., Phys. Rev. B 81, 165405 (2010).
[16] I. Meric et al., Nat. Nano. 3, 654 (2008).
[17] M. Freitag et al., Nano Lett. 9, 1883 (2009).
[18] A. Barreiro et al., Phys. Rev. Lett. 103, 076601 (2009).
[19] A. M. DaSilva et al., Phys. Rev. Lett. 104, 236601 (2010).
[20] J. Y. Park et al., Nano Lett. 4, 517 (2004).
[21] E. Pop et al., Phys. Rev. Lett. 95, 155505 (2005).
[22] S. Berciaud et al., Nano Lett. 9, 346 (2009).
[23] K. F. Mak, J. Shan, and T. F. Heinz, Phys. Rev. Lett. 106,
046401 (2011).
[24] H. Yan et al., Phys. Rev. B 80, 121403 (2009).
[25] N. Bonini et al., Phys. Rev. Lett. 99, 176802 (2007).
[26] See EPAPS Document for more information.
[27] I. Chatzakis et al., submitted (2011).
[28] K. I. Bolotin et al., Phys. Rev. Lett. 101, 096802 (2008).
[29] T. Stauber, N. M. R. Peres, and A. H. Castro Neto, Phys. Rev.
B 78, 085418 (2008).
[30] A. H. C. Neto, and F. Guinea, Phys. Rev. B 75, 045404
(2007).
[31] M. Lazzeri et al., Phys. Rev. B 78, 081406 (2008).
[32] E. H. Hwang, B. Y. K. Hu, and S. Das Sarma, Phys. Rev. B
76, 115434 (2007).
[33] A. A. Kozikov et al., Phys. Rev. B 82, 075424 (2010).
[34] E. H. Hwang, and S. Das Sarma, Phys. Rev. B 79, 165404
(2009).
4
Auxiliary Material
Observation of Intra- and Inter-band Transitions in the
Optical Response of Graphene
Leandro M. Malard1, Kin Fai Mak1, A. H. Castro Neto2,3, N. M. R. Peres4, and
Tony F. Heinz1
1Departments of Physics and Electrical Engineering, Columbia University, 538 West 120th Street, New
York, New York 10027, USA
2Department of Physics, Boston University, 590 Commonwealth Avenue, Boston, Massachusetts 02215,
USA
3Graphene Research Centre, National University of Singapore, 2 Science Drive 3, 117542, Singapore
4Departamento de Física e Centro de Física, Universidade do Minho, P-4710-057 Braga, Portugal
1) Determination of the electronic temperature
Here we describe the model employed for the heat capacity of the strongly coupled
optical phonons. These excitations are assumed to be in thermal equilibrium with the
electronic excitations, but, because of the very low heat capacity of electronic
excitations in graphene, the heat capacity is dominated by the phonons. Knowledge
of this heat capacity then allows us to convert the experimental value of the absorbed
pump fluence into the electronic temperature of this subsystem. With the temperature
in hand, we predict the optical response from Eq. (1) of the main text to calculate for a
comparison with the experimental data.
Strong electronic coupling is present for the zone-center phonons (G band, energy of
200 meV) and the zone-edge phonons (D band, energy of 150 med) [A1, A2, A3].
Using the Einstein model for heat capacity including these two phonon modes, we
obtain for the energy density (per unit area) as a function of temperature [A4]:
5
. (E1)
Here A is the Brillouin zone area, f is the fraction of Brillouin zone filled by the G and
D optical phonons and
denotes the G or D phonon energy. We determined f in
Eq. (E1) by comparison with a previous experiment in which transient phonon
temperatures for femtosecond laser excitation of graphite were determined by time-
resolved Raman scattering [A5, A6]. Fitting these earlier measurements to a
polynomial yields the following relation between the absorbed fluence (or energy
density) per graphene layer and the temperature:
. (E2)
We find the fraction f by comparing Eq. (E1) and (E2). Eq. (E1) then yields the
energy density as a function of temperature, even for temperatures below the range
for which the experimental data from the earlier experiments are available.
To model the temporal evolution of the sub-system of the electronic excitations and
the strongly coupled phonons, we assume that the energy relaxes as a rate of 1/τph
through the anharmonic decay, with Eq. (E1) yielding the variation of the
temperature. This temperature dynamics is then insert in Eq. (1) from the main text of
the paper to fit our pump-probe dynamics experiment. Because of the non-linear
relationship between the energy density and temperature, the decay time of these two
quantities are not the same, as can see in Fig. S1.
6
Figure S1: (color online) Temporal evolution of the energy density (black) and
temperature (red) calculated for the case of an absorbed fluence of 0.5
. The
decay times for the energy density and temperature are, respectively, 1.4 and 3.1 ps.
2) Calculation of the electronic scattering rate in graphene at high temperatures
The intra-band contribution to the optical conductivity scales with the electron
scattering rate Γ. For the regime of temperature of the electrons and strongly coupled
optical phonons, we expect that scattering of the electrons by these phonons will
contribute significantly to the overall scattering rate Γ. Here we calculate this
contributions and find that it yields a rate similar to that inferred from the experiment.
As above, we model the strongly coupled phonons by two branches of dispersionless
optical phonons, the 200 meV zone-center and 150 meV zone-edge phonons.
7
The self-energy of the electrons in graphene due to optical phonons is given by [A7]:
1
1
9
2
N c
β
Q , iωn − iνm ) , [E3]
0 ( k −
D0 (iνm )
Gbb
∑
m
Σ( k , iωn ) / = −
∂t
⎛
⎜
∂a
⎝
2
⎞
⎟
⎠
M cω0
∑
Q→
where
is the electron-phonon coupling as defined in Ref. [A8],
is the
€
carbon-carbon distance,
is the hopping in the tight-binding model for graphene, MC
is the carbon mass,
is the phonon frequency. All the other parameters are equally
defined by Ref. [A7]. After performing the Matsubara summation, the imaginary part
of the self-energy is given by
1
9π
2
∂t
ℑΣ( k , iωn ) / = −
⎛
⎞
∑
⎜
⎟
4
N c
M cω0
∂a
⎝
⎠
Q
] δ(ε k −ε k −
[
Q + ω0 )
+ n B (ω0 ) + n F (ε k −
Q )
] δ(ε k −ε k −
[
+ n B (ω0 ) + 1 − n F (ε k −
Q )
Q − ω0 )
[
] δ(ε k +ε k −
Q )
+ n B (ω0 ) + n F (−ε k −
Q + ω0 )
[
] δ(ε k +ε k −
Q − ω0 ),
Q )
+ n B (ω0 ) + 1 − n F (−ε k −
where nB and nF are the Bose and Fermi distribution functions,
[E3]
and
€
. The two first terms are intra-band scattering processes and the last two
are inter-band ones.
The scattering rate is obtained from
. [E4]
To compute
, we need to solve the following integrals appearing in Eq. [E3]:
8
[E5]
where
. The
-functions in Eq. [E3] guarantee that we can replace
in the Fermi functions with
for the first and second integral, and with
in the fourth one. The third integral does not contribute to the scattering
rate, since its argument is always positive. Computing the three integrals, we find that
the scattering rate of an electron with momentum
has, as a consequence of phase-
space restrictions, different forms for different energy ranges:
1. For
:
2. For
:
where C is given by
[E6]
[E7]
[E8]
9
and
. The electron phonon couplings
used for our
calculations are given by
[A8] and
[A9] for the zone-center and zone-edge phonons, respectively.
Now we need to average the energy dependent scattering rate given by Eq. [E6-E7] to
find the average scattering rate [A10]:
[E9]
where
is the density of states of graphene.
Finally to compute the results for the temperature dependence of the scattering rate
shown in Fig. 3 of the main manuscript, we summed the contributions from the two
optical phonon branches:
[E10]
References:
[A1] – S. Piscanec et al., Phys. Rev. Lett. 93, (2004).
[A2] – M. Lazzeri and F. Mauri, Phys. Rev. B 73, 165419 (2006).
[A3] – N. Bonini et al., Phys. Rev. Lett. 99, 176802 (2007).
[A4] – N. W. Ashcroft and N. D. Mermin, Solid State Physics (Thomson, 1976).
[A5] – H. Yan et al., Phys. Rev. B 80, 121403R (2009).
[A6] – C. H. Lui et al., Phys. Rev. Lett. 105, 127404 (2010).
[A7] – T. Stauber, N. M. R. Peres and A. H. Castro Neto, Phys. Rev. B 78, 085418
(2008).
[A8] – A. H. Castro Neto and F. Guinea, Phys. Rev. B 75, 045404 (2007).
[A9] – M. Lazzeri et al., Phys Rev. B. 78, 081406R (2008).
[A10] – Paul Harrison, Quantum Wells, Wires and Dots: theoretical and
computational physics of semiconductor nanostructures (Wiley, 2005).
10
|
1508.04433 | 2 | 1508 | 2015-10-29T18:27:09 | Helical Quantum Edge Gears in 2D Topological Insulators | [
"cond-mat.mes-hall",
"cond-mat.str-el"
] | We show that two-terminal transport can measure the Luttinger liquid (LL) parameter $K$, in helical LLs at the edges of two dimensional topological insulators (TIs) with Rashba spin-orbit coupling. We consider a Coulomb drag geometry with two coplanar TIs and short-ranged spin-flip inter-edge scattering. Current injected into one edge loop induces circulation in the second, which floats without leads. In the low-temperature ($T \rightarrow 0$) perfect drag regime, the conductance is $(e^2/h)(2 K + 1)/(K + 1)$. At higher $T$ we predict a conductivity $\sim T^{-4K+3}$. The conductivity for a single edge is also computed. | cond-mat.mes-hall | cond-mat |
Helical Quantum Edge Gears in 2D Topological Insulators
Yang-Zhi Chou,1, ∗ Alex Levchenko,2 and Matthew S. Foster1, 3
1Department of Physics and Astronomy, Rice University, Houston, Texas 77005, USA
2Department of Physics, University of Wisconsin-Madison, Madison, Wisconsin 53706, USA
3Rice Center for Quantum Materials, Rice University, Houston, Texas 77005, USA
(Dated: September 11, 2018)
We show that two-terminal transport can measure the Luttinger liquid (LL) parameter K, in
helical LLs at the edges of two-dimensional topological insulators (TIs) with Rashba spin-orbit
coupling. We consider a Coulomb drag geometry with two coplanar TIs and short-ranged spin-flip
interedge scattering. Current injected into one edge loop induces circulation in the second, which
floats without leads.
In the low-temperature (T → 0) perfect drag regime, the conductance is
(e2/h)(2K + 1)/(K + 1). At higher T we predict a conductivity ∼ T −4K+3. The conductivity for a
single edge is also computed.
PACS numbers: 71.10.Pm, 72.15.Nj, 74.25.F-
The edge states that encircle two-dimensional (2D)
topological insulators (TIs) realize a novel electronic heli-
cal Luttinger liquid (HLL) phase [1 -- 3]. Distinct from an
ordinary one-dimensional (1D) quantum wire and from a
quantum Hall edge, a helical edge consists of two coun-
terpropagating modes forming a Kramers pair. The left-
and right-moving channels interact through Coulomb re-
pulsion, but time reversal symmetry protects the edge
from the opening of a gap and from Anderson localization
due to impurities. The combination of topological pro-
tection and electron correlations implies that a TI edge
is an ideal Luttinger liquid at low temperatures [4, 5].
Experimental evidence for helical edge states in HgTe
[6] and InAs/GaSb [7] includes a quantized conductance
G ≃ 2e2/h [7, 8].
In the absence of electrical contacts and magnetic
fields, a HLL forms a closed, unbreakable loop. This
topology of the edge has so far received little atten-
tion.
In this Letter, we propose a TI device geome-
try in which edge loops rotate as interlocking "gears"
through Coulomb drag [9 -- 13]. Our main result is that
the strength of electron correlations encoded in the Lut-
tinger parameter can be directly obtained in such a device
using a two-terminal dc conductance measurement.
Correlations are generically strong in 1D electron fluids
because two particles cannot exchange positions without
scattering or tunneling. These correlations are encoded
in the Luttinger parameter K [14]. Measuring K in a
nontopological 1D electronic system (or "wire") is pos-
sible but delicate. For instance, the zero-temperature
(T = 0) dc transport through a perfectly clean wire gives
a quantized conductance independent of K [15 -- 17]. In a
long wire, disorder tends to induce Anderson insulating
behavior. At temperatures T & vkF /kB, inelastic scat-
tering due to irrelevant umklapp interactions gives a con-
ductivity that depends on T through a power law [18];
here v and kF , respectively, denote the charge velocity
and Fermi wave vector. The disorder-induced scattering
may lead to a qualitatively similar effect [19]. The tem-
perature exponent in conductance can reveal the Lut-
tinger parameter K, but a large temperature range is
needed to fit the data. The tunneling zero bias anomaly
is also predicted to encode K, but measurements often
contain contributions from other mechanisms [20].
In the simplest version of HLL physics that realizes the
quantum spin Hall effect [4, 5, 21, 22], the z component
of spin is assumed to be conserved in a TI. As a result,
the edge electrons carry well-defined Sz currents. When
0
I'
1
I
1
I
2
V
FIG. 1. Using helical quantum edge gears to measure the
Luttinger parameter. We consider Z2 TI edge states in two
adjacent topological regions. The blue and red arrows indi-
cate the propagation directions of edge electrons with oppo-
site helicities. The left TI is connected to external leads; I1
and I ′
1 denote the currents of the edges connected to these.
The right TI edge floats as an electrically isolated closed loop.
Rashba spin-orbit coupling [1] enables Coulomb drag due to
short-ranged spin-flip scattering [23] between the adjacent
edges. This induces a current I2 that circulates in the right
edge.
In the case of identical TIs with an interacting edge
region of size L → ∞, at zero temperature strong backscat-
tering "locks" the currents I1 = I2, associated to perfect drag
[10, 12]. We then predict that the zero-temperature conduc-
1)/V = (e2/h)(2K + 1)/(K + 1), where K
tance is G = (I1 + I ′
is the Luttinger parameter. In a real system of finite length
L ≫ ξ and at temperatures T satisfying v/L . kBT ≪ ∆
[11] with ξ = v/∆ and ∆ the Mott gap of the antisymmetric
mode, the result for G holds up to terms exponentially small
in L/ξ and ∆/kBT [10 -- 12]. Here v is the charge velocity.
Rashba spin-orbit coupling (SOC) is present [1] (generi-
cally expected in the absence of inversion symmetry), Sz
symmetry is sabotaged. New spin-flip interactions [23 --
25] are then allowed on TI edges.
We show that the Luttinger parameter enters the con-
ductance in a Coulomb drag geometry consisting of two
coplanar TI regions with Rashba SOC. Over a segment of
length L, proximate HLL edge states are separated by a
gap narrow enough to allow short-ranged Coulomb scat-
tering but wide enough to prevent tunneling. In Fig. 1,
we consider two identical helical edges. Current I1 is in-
jected by external leads. Short-ranged spin-flip scatter-
ing [23] between edges induces a current I2 in the right TI
edge loop, which floats without leads. At zero tempera-
ture, the two proximate edge segments develop a locking
state of perfect drag (I1 = I2) [10, 12] for an infinitely
long interacting region L → ∞. An additional current
I′1 flows in parallel between the contacts. The zero-
temperature two-terminal conductance G = (I1 + I′1)/V
is
G =
e2
h h1 + (1 + 1/K)−1i =
e2
h (cid:18) 2K + 1
K + 1 (cid:19) ,
(1)
where (1 + 1/K) is the dimensionless resistance of the
locked edges, as explained below. For a finite locking
length L ≫ ξ and at temperatures T satisfying v/L .
kBT ≪ ∆ [11], Eq. (1) holds up to exponentially small
corrections in L/ξ and ∆/kBT [10 -- 12]. Here ξ ≡ v/∆
is the length scale associated to the gapped "antilocking"
mode with I1 = −I2; ∆ is the energy gap.
We also discuss dissipative finite-temperature trans-
port in this geometry. In contrast to the usual setup for
Coulomb drag [9, 13], the system is naturally character-
ized in terms of conductances or conductivities:
I2 (cid:21)=(cid:20) G11 G12
(cid:20) I1
G21 G22 (cid:21)(cid:20) V1
V2 (cid:21) , Gij = σij /L,
where the labels 1 and 2 indicate the active and passive
systems respectively. For our TI edges, the passive sys-
tem is a closed HLL loop with V2 = 0, I1 = σ11V1/L, and
I2 = σ21V1/L. We compute the intraedge and transcon-
ductivities using the Kubo formula and bosonization, em-
ploying the effective potential formalism [26, 27]. Both
σ11 and σ21 give T −4K+3 (T −4K+2) behavior in the ab-
sence (presence) of disorder, above the locking transition.
Finally, we compute the conductivity of a single
edge due to the least irrelevant symmetry-allowed (one-
particle umklapp) interaction term. We find asymptotic
T −2K−1 (T −2K−2) behavior in the high- (low-) T limits,
in the presence of disorder, consistent with [28], and we
also obtain the full result for the clean limit. Power-law
scaling of conductance as a function of temperature and
bias voltage that may be attributable to Luttinger liquid
physics was recently observed in InAs/GaSb quantum
spin Hall devices [29].
Model. -- The edge states of a 2D TI can be expressed
in terms of right (R) and left (L) mover fermion fields.
The kinetic term is
2
(2)
H0 = −ivF Z dx(cid:2)R†(x)∂xR(x) − L†(x)∂xL(x)(cid:3) ,
where vF is the Fermi velocity of the edge band. Time-
reversal symmetry is encoded by R(x) → L(x), L(x) →
−R(x), and i → −i. Left and right movers interact via
intraedge Coulomb repulsion.
We focus on the coplanar geometry in Fig. 1 and
consider the backscattering components of the interedge
Coulomb interaction. An additional interedge Luttinger
interaction does not modify our results for the locking
regime if the distal portion of edge loop 2 is much longer
than the interacting segment of length L; otherwise the
parameter K in Eq. (1) encodes a combination of inter-
and intraedge correlations.
In the presence of Rashba
SOC, the following interedge backscattering terms are
allowed by symmetry [23]:
H− = U−Z dxhei2(kF 1−kF 2)xL†1R1R†2L2 + H.c.i ,
H+ = U+Z dxhei2(kF 1+kF 2)xL†1R1L†2R2 + H.c.i ,
(3)
(4)
where kF 1 (kF 2) indicates the Fermi momentum in the
first (second) edge. These are defined relative to an edge
Dirac point, which is a commensurate (time-reversal in-
variant) momentum [2]. The U− interaction describes
normal backscattering, while U+ is a two-particle umk-
lapp interaction. Additional one-particle umklapp inter-
action terms are also allowed,
HU = Xa=1,2
UaZ dxhe−i2kF a xR†aLaR†¯aR¯a
− ei2kF a xL†aRaL†¯aL¯a + H.c.i,
(5)
where a is the index of the edge, ¯1 = 2, and ¯2 = 1.
It is worth mentioning that all of these interactions are
disallowed in the presence of Sz conservation (in each
edge) [23, 30].
For simplicity, we assume the two HLLs are identi-
cal, so that kF 1 = kF 2 ≡ kF and U1 = U2 ≡ U . The
dominant interedge interaction at T = 0 is the nonumk-
lapp backscattering H−; the others are irrelevant at long
In order to include Lut-
wavelengths for kF 6= 0 [14].
tinger liquid effects, we use bosonization [14, 31]. The
individual edge loop HLLs are described by
Hb,0 =
v
2 Xa=1,2Z dx(cid:20)K (∂xφa)2 +
1
K
(∂xθa)2(cid:21) ,
(6)
where K is the Luttinger parameter and v is the charge
velocity. K = 1 and v = vF corresponds to the free
fermion limit. The density (n) and current (I) can be
expressed in terms of the axial fields as na = ∂xθa/√π
and Ia = −∂tθa/√π, respectively. The interedge interac-
tion H− is bosonized to
Hb,− =
U−
2π2α2 Z dx cosh√4π (θ1 − θ2)i ,
(7)
where α is an ultraviolet length scale.
Perfect current drag and dc conductance. -- At zero
temperature, two infinite HLLs form an interedge lock-
ing state [10, 12] due to the two-particle backscatter-
ing term in Eq. (7). The locking state is characterized
by θ1(t, x) = θ2(t, x) + cm, where cm = (m + 1/2)√π
is a constant and m ∈ Z. This state exhibits per-
fect current drag [10], I1 = I2 in Fig. 1. The conduc-
tance of the locked edges (both carrying current I1) is
I1/V = (e2/h)[K/(K + 1)]. This can be understood as
the series resistor combination of a spinless LL connected
to leads with resistance h/e2 [15 -- 17] and one with peri-
odic boundary conditions and resistance h/Ke2 [32, 33].
An explicit Green's function calculation confirms this re-
sult [34], which is also independent of disorder. Equation
(1) is obtained by adding the parallel I′1 edge channel.
For a finite interacting region of length L and nonzero
temperature T , we require that L ≫ ξ and kBT ≪ ∆.
Occasional phase slips between the drive and slave cir-
cuits give rise to corrections that are exponentially small
in L/ξ and ∆/kBT [10 -- 12]. For L = 1 µm in InAs/GaSb
with v ∼ vF = 3 × 104 m/s, this gives a lower bound
for ∆ of the order of v/L = 0.02 meV. We assume
that kBT is larger than the latter to avoid coherent in-
stanton effects [11]. By comparison, the bulk minigap
is of the order of 4 meV [7]. The Mott gap takes the
form [14] ∆ ∼ pK U−v /α. Using α = 1 nm gives
∆ ∼ 20 meVpK (U−/v). The interaction strength U−
is obtained from the inter-edge Coulomb potential, medi-
ated by matrix elements determined by the Rashba SOC
in each TI (since it vanishes in its absence). The result
will depend on microscopic details that we do not analyze
here.
Finite temperature corrections. -- Above a crossover
temperature T ∗ ∼ ∆/kB[12], inelastic electron-electron
collisions due to the interedge interactions in Eqs. (3) -- (5)
can be treated perturbatively. In addition, we consider
intraedge collisions due to electron-electron interactions
[Eq. (14), below] and forward-scattering potential disor-
der. Ordinary backscattering (random mass) disorder is
forbidden by time-reversal symmetry. We ignore irrel-
evant backscattering disorder terms with extra deriva-
tives that are not expected to impact the conductivity
in isolation [28] and which give subleading corrections in
combination with interactions. Forward-scattering disor-
der is encoded in Himp =Pa=1,2R dx ηa(x) na(x), where
ηa(x) is a random potential obeying ηa(x) = 0 and
ηa(x)ηb(x′) = gη δa,bδ(x − x′). The ··· denotes disorder
averaging, while gη characterizes the disorder strength.
3
To compute the conductivity, we evaluate interaction
corrections to the inverse boson propagator via the effec-
tive potential method [26, 27]. We use replicas to average
over disorder. The retarded boson correlation function is
ab
, (8)
=h G(R)(ω, k)i−1
ab −h Π(R)(ω, k)iab
h G(R)(ω, k)i−1
where a, b ∈ {1, 2} indicate the edges. The noninter-
acting propagator is G(R)(ω, k), while Π(R) denotes the
self-energy describing the interaction corrections. Equa-
tion (8) is a matrix Dyson equation. At second order in
the coupling constants, Π(R) contains an imaginary part
that determines the scattering rates; the real part does
not contribute to dc conductivity.
In the limit ω → 0
with k = 0,
ImhΠ(R)
ab i = −2ωΞab + O(cid:0)ω2(cid:1) ,
(9)
2 Ξ− + ΞU + ΞW and Ξ12 = Ξ21 = 1
where Ξab is the "rate" (inverse mean free path) as-
sociated to Πab. The components are Ξ11 = Ξ22 =
2 Ξ+ + 1
1
2 Ξ−.
ΞU is due to the one-particle backscattering term HU .
Ξ+ and Ξ− correspond to the two-particle backscatter-
ing interactions H+ and H−, respectively. ΞW is due
to intraedge inelastic electron-electron collisions, HW in
Eq. (14). ΞU and ΞW affect only the diagonal elements
of the self-energy, while Ξ+ and Ξ− contribute to all of
the components.
2 Ξ+ − 1
Temperature dependences of all the scattering rates
are obtained analytically. For kBT ≫ max(gη/v, vkF ),
ΞU ∝ T 2K−1, ΞW ∝ T 2K+1, and Ξ± ∝ T 4K−3. In the
presence of disorder and for kBT ≪ gη/v, ΞU ∝ T 2K,
ΞW ∝ T 2K+2, and Ξ± ∝ T 4K−2. The additional power
of T comes from disorder smearing of kF [35]. Full
crossovers with and without disorder are determined by
the explicit forms of ΞU,W,± provided in Ref. [34].
The dc conductivity can be obtained through Kubo
formula [36]. The intraedge dc conductivity is
e2
1
σ11 = −
π
e2
2h(cid:20)
=
lim
ω→0
Imhω G(R)
11 (ω, k)i
1
Ξ+ + ΞU + ΞW
+
1
Ξ− + ΞU + ΞW(cid:21) .
(10)
This expression is very different from the conductivity
of an isolated edge, discussed below. Both intraedge
and interedge interactions contribute to Eq. (10), but
the intraedge contribution ΞW is subleading comparing
to the interedge rates. The finite-temperature behavior
for σ11 is summarized as follows. For clean edges and
kBT ≫ vkF ,
σ11 ∼(T −4K+3,
T −2K+1,
for K ≤ 1,
for K > 1.
(11)
With smooth disorder and kBT ≪ gη/v,
for K ≤ 1,
for K > 1.
σ11 ∼(T −4K+2,
T −2K,
K ≤ 1 (repulsive interactions) is the physical situation.
The transconductivity is
σ21 =
e2
2h(cid:20)
1
Ξ+ + ΞU + ΞW −
1
Ξ− + ΞU + ΞW(cid:21) .
(13)
This can be measured by shorting the distal part of pas-
sive edge loop with an ideal (zero input impedance) cur-
rent meter. The leading temperature dependence of the
drag conductivity is the same as the intraedge conduc-
tivity. In the usual case, one measures instead the drag
resistivity [9, 13]. Here this evaluates to ρD = −ρ12 =
(h/2e2) [Ξ− − Ξ+] , independent of the interedge U and
intraedge W interactions.
In the case of clean identi-
cal edges, a positive drag resistivity with leading T 4K−3
behavior is obtained. This is the same result found previ-
ously for spinless Luttinger liquids [11, 12] and TI edges
with small magnetic fields [30].
Single edge. -- Finally, we consider dc conductivity of a
single edge in isolation, in the presence of Rashba SOC.
The least irrelevant intra-edge electron-electron interac-
tion term allowed by time-reversal symmetry that can
give a finite transport lifetime is
HW = W Z dx :(cid:8)ei2kF xL†(x)R(x)R†(x)[−i∂xR(x)]
+e−i2kF xR†(x)L(x)L†(x)[−i∂xL(x)] + H.c.(cid:9) :, (14)
where :O: denotes the normal ordering of O. This is a
one-particle spin-flip umklapp term. Similar one-particle
backscattering interactions appear in Refs. [24, 25, 28],
but the full temperature dependence of the dc conductiv-
ity was not determined. The interaction correction due
to Eq. (14) can be described by a self-energy with imag-
inary part ImhΠ(R)
k = 0 and ω → 0. We find that
W (ω, k)i = −2ωΞW + O(cid:0)ω2(cid:1) , when
22Kπ2K+3KΓ [−K − 3]
W 2 α2K
ΞW =
(v)2 l2K+1
T
Γ [K + 2]
+ γ2
γ/π
−∞
dy"
×Z ∞
2π (cid:1)2
(cid:0)y − kF lT
cosh (2πy) − cos (πK)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
×
sin (πK)
+ (kF → −kF )#
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Γ(cid:2) 4+K
2 + iy(cid:3)
Γ(cid:2) 2−K
2 + iy(cid:3)
2
,
(15)
where W = W/(π3/2α) and lT ≡ v/kBT denotes the
thermal de Broglie wavelength. The disorder is en-
coded in γ ≡ lT (K/v)2gη/2π. The dc conductivity is
σdc = (e2/h)(1/ΞW ). At zero temperature where the
HLL exhibits ballistic transport, σdc diverges.
For a clean noninteracting edge (K = 1 and v = vF ),
the conductivity reduces to
4
(12)
σdc =
e2
h
(v)2 l3
T
W 2π3
6 [cosh (kF lT ) + 1]
2π )4 + 5
2 ( kF lT
2π )2 + 9
.
(16)
(cid:2)( kF lT
16(cid:3)
At high temperatures kF lT ≫ 1, this is proportional
to T −3;
in the opposite limit the umklapp scattering
is thermally activated, giving σdc ∼ T exp (kF lT ). Lut-
tinger interactions modify the high-temperature behav-
ior to T −2K−1, while disorder leads to σdc ∼ T −2K−2
for lT ≫ (v)2g−1
η , again due to smearing of the Fermi
momentum [35]. The disordered result is consistent with
earlier predictions [24, 25, 28]. The T −2K−1 behavior
is the most important temperature dependence in the
clean edge due to the intraedge interactions, in the pres-
ence of Rashba SOC. The responsible interaction term in
Eq. (14) will be generated by renormalization irrespective
of whether it arises in a particular microscopic model.
Summary and discussion - In this work, we have shown
that low-temperature edge state transport measurements
for two proximate HLLs can quantify the value of the
Luttinger parameter in the presence of spin-flip interedge
electron-electron scattering. The latter is enabled by
Rashba SOC within each TI, as can arise in InAs/GaSb.
In contrast to the usual setup for Coulomb drag, the
passive circuit floats without leads and provides a much
stronger source of scattering for the active circuit edge
than intraedge interactions, which are negligible at low
temperature. Because of the topological protection, this
result is immune to disorder but receives exponentially
small corrections for a long, but finite interacting region.
In the same device geometry, both the intraedge con-
ductivity and the transconductivity show the same lead-
ing temperature dependence for T above the crossover
scale to the low-temperature locking regime. Thus, two-
terminal conductivity gives an alternative route to detect
Coulomb drag physics. We have also computed the con-
ductivity correction due to the least irrelevant symmetry-
allowed interaction in a given edge. This gives T −2K−1
temperature dependence for a clean edge.
We close with some observations and avenues for fu-
ture work.
In general, negative drag is possible when
kF 1 − kF 2 ≫ kF 1 + kF 2. Eq. (4) instead of Equation
(3) dominates the interedge interactions at low temper-
ature in this case. For two almost identical edges but
kF 1 = −kF 2, a perfect antiparallel current locking can oc-
cur; the two-terminal conductance is still given by Eq. (1)
in the T → 0 limit. The finite-temperature behavior will
be qualitative the same as the parallel drag situation.
The generic kinetic theory of Coulomb drag between he-
lical edge states, that also includes the forward-scattering
long-ranged component of the Coulomb interaction, is an
important topic for future work [37]. Understanding how
a HLL edge state thermalizes via the various scattering
mechanisms has crucial implications for nonequilibrium
spectroscopy [38, 39]. It will also be interesting to study
the noise [40] for the two-helical-edge setup described
here.
Y.-Z.C. and M.S.F. thank R.-R. Du, L. Du, D. Na-
telson, and A. Nevidomskyy for useful discussions. A.L.
thanks N. Kainaris, I. Gornyi and D. Polyakov for multi-
ple important discussions and ongoing collaboration on a
related problem. A.L. and M.S.F. acknowledge the hos-
pitality of the Spin Phenomena Interdisciplinary Center
(SPICE), where this work was completed. Y.-Z.C. and
M.S.F. acknowledge funding from the Welch Foundation
under Grant No. C-1809 and from an Alfred P. Sloan
Research Fellowship (No. BR2014-035). Y.-Z.C. also ac-
knowledges hospitality of the Michigan State University.
A.L. acknowledges funding from NSF Grants No. DMR-
1401908 and No. ECCS-1407875.
∗ yc26@rice.edu
5
[23] Y. Tanaka and N. Nagaosa, Phys. Rev. Lett. 103, 166403
(2009).
[24] T. L. Schmidt, S. Rachel, F. von Oppen, and L. I. Glaz-
man, Phy. Rev. Lett. 108, 156402 (2012).
[25] N. Lezmy, Y. Oreg, and M. Berkooz, Phys. Rev. B 85,
235304 (2012).
[26] M. E. Peskin and D. V. Schroeder, An Introduction to
Quantum Field Theory (Westview, Boulder, 1995).
[27] Z. Ristivojevic, P. Le Doussal, and K. J. Wiese, Phys.
Rev. B 86, 054201 (2012).
[28] N. Kainaris, I. V. Gornyi, S. T. Carr, and A. D. Mirlin,
Phys. Rev. B 90, 075118 (2014).
[29] T. Li, P. Wang, H. Fu, L. Du, K. A. Schreiber, X. Mu, X.
Liu, G. Sullivan, G. A. Csathy, X. Lin, R.-R. Du, Phys.
Rev. Lett. 115, 136804 (2015).
[30] V. A. Zyuzin and G. A. Fiete, Phys. Rev. B 82, 113305
(2010).
[31] R. Shankar, Acta Phys. Pol. B 26, 1835 (1995).
[32] W. Apel and T. M. Rice, Phys. Rev. B 26, 7063(R)
(1982).
[33] C. L. Kane and M. P. A. Fisher, Phys. Rev. B 46, 15233
(1992).
[34] See supplemental material.
[35] G. A. Fiete, K. Le Hur, and L. Balents, Phys. Rev. B 73,
[1] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 146802
165104 (2006).
(2005).
[36] J. Sirker, R. G. Pereira, and I. Affleck, Phys. Rev. B 83,
[2] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045
035115 (2011).
(2010).
[37] N. Kainaris, A. Levchenko, I. Gornyi, and D. Polyakov,
[3] X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057
in preparation.
[38] C. Altimiras, H. Le Sueur, U. Gennser, A. Cavanna, D.
Mailly, and F. Pierre, Nature Physics 6, 34 (2010).
[39] S. S. Apostolov and A. Levchenko, Phys. Rev. B 89,
201303(R) (2014).
[40] Y. M. Blanter and M. Buttiker, Physics Reports 336, 1
(2000).
(2011).
[4] C. Xu and J. E. Moore, Phys. Rev. B 73, 045322 (2006).
[5] C. Wu, B. A. Bernevig, and S.-C. Zhang, Phys. Rev. Lett.
96, 106401 (2006).
[6] M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buh-
mann, L. W. Molenkamp, X.-L. Qi, and S.-C. Zhang,
Science 318, 766 (2007).
[7] I. Knez, R.-R. Du, and G. Sullivan, Phys. Rev. Lett. 107,
136603 (2011).
[8] A. Roth, C. Brune, H. Buhmann, L. W. Molenkamp, J.
Maciejko, X.-L. Qi, and S.-C. Zhang, Science 325, 294
(2009).
[9] A. G. Rojo, Journal of Physics: Condensed Matter 11,
R31 (1999).
[10] Y. V. Nazarov and D. V. Averin, Phys. Rev. Lett. 81,
653 (1998).
[11] V. V. Ponomarenko and D. V. Averin, Phys. Rev. Lett.
85, 4928 (2000).
[12] R. Klesse and A. Stern, Phys. Rev. B 62, 16912 (2000).
[13] B.
preprint
Levchenko,
Narozhny
and
A.
arXiv:1505.07468 (2015).
[14] T. Giamarchi, Quantum Physics in One Dimension (Ox-
ford University Press, Oxford, 2004).
[15] D. L. Maslov and M. Stone, Phys. Rev. B 52, R5539
(1995).
[16] V. V. Ponomarenko, Phys. Rev. B 52, R8666 (1995).
[17] I. Safi and H. J. Schulz, Phys. Rev. B 52, R17040 (1995).
[18] T. Giamarchi, Phys. Rev. B 44, 2905 (1991).
[19] D. L. Maslov, Phys. Rev. B 52, R14368 (1995).
[20] V. V. Deshpande, M. Bockrath, L. I. Glazman, and A.
Yacoby, Nature 464, 209 (2010).
[21] C. L. Kane and E. J. Mele, Phy. Rev. Lett. 95, 226801
(2005).
[22] B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. 96,
106802 (2006).
Helical Quantum Edge Gears in 2D Topological Insulators:
SUPPLEMENTAL MATERIAL
We set = kB = 1 except as noted.
BOSONIZATION CONVENTIONS
We adopt the standard field theoretic bosonization method. The fermionic fields can be described by chiral bosons
via
6
R(x) =
1
√2πα
ei√π[φ(x)+θ(x)], L(x) =
1
√2πα
ei√π[φ(x)−θ(x)],
where α is some ultraviolet length scale.
The time reversal operations in the bosonic language read: φ → −φ +
√π
2 , θ → θ −
√π
2 , and i → −i.
In the imaginary time formalism, the Luttinger action for the ath edge reads
BOSONIC ACTIONS
i (∂xθa) (∂τ φa) +
v
2 Zτ,x (cid:20)K (∂xφa)2 +
xφa(cid:1) cos(cid:16)√4π θa + 2kF,ax(cid:17) ,
S0,a =Zτ,x
SW,a = WaZτ,x (cid:0)∂2
Simp,a =Zτ,x
ηa(x)
1
√π
∂xθa,
1
K
(∂xθa)2(cid:21) ,
where Rτ,x
is the short hand notation for R dτ dx.
The inter-edge interactions correspond to
SU = Xa=1,2
S+ = −U+
2π2α2 Zτ,x
U−
2π2α2 Zτ,x
S− =
Ua
π3/2α Zτ,x
(∂xφ¯a) sinh√4πθa + 2kF axi ,
cosh√4π (θ1 + θ2) + 2(kF 1 + kF 2)xi ,
cosh√4π (θ1 − θ2) + 2(kF 1 − kF 2)xi ,
where a labels the edge, with ¯1 = 2 and ¯2 = 1.
We can integrate over ∂xφ1 and ∂xφ2 exactly. The θ-only action is
a (∂xθa)2i
1
Sθ = Xa=1,2
− Xa=1,2
− Xa=1,2
i Wa
2vaKa Zτ,x h(∂τ θa)2 + v2
vaKa Zτ,x
vaKaπ3/2α Zτ,x
iU¯a
(∂τ ∂xθa) cos(cid:16)√4π θa + 2kF ax(cid:17)
(∂τ θa) sin(cid:16)√4πθ¯a + 2kF ¯ax(cid:17) + S+ + S−,
(17)
(18)
(19)
(20)
(21)
(22)
(23)
(24)
(25)
(26)
where we have dropped terms not contributing to the boson self-energy at the second homogeneous order of the
coupling constants.
7
Forward-scattering potential disorder is averaged using the replica trick. The quadratic part of the action becomes
Sθ,0 =
1
2vK Xa=1,2
R
Xn=1Z dτ dxh(∂τ θa,n)2 + v2 (∂xθa,n)2i −
gη
2π Xa=1,2
R
Xn,m=1Z dτ dτ′dx [∂xθa,n(τ )] [∂xθa,m(τ′)] ,
(27)
where n, m are the replica indexes, R is the number of replica, and gη is the variance of the disorder potential [η(x),
introduced in the main text]. All interaction terms are diagonal in the replica space.
DC CONDUCTANCE
We derive the zero temperature dc conductance result [Eq. (1)] in this section. We consider the device geometry
in Fig. 1. The edge carrying current I1 has Luttinger parameter K and charge velocity v. It is connected to external
leads, which we treat as non-interacting free fermion reservoirs with K = 1 and v = vF . The electric field is applied
between the leads. The passive circuit edge carrying current I2 forms a closed loop with uniform Luttinger parameter
K and velocity v. For simplicity, we assume that this loop is infinitely long (but see below). The two edges interact
via Eq. (7) in the region −L/2 ≤ x ≤ L/2. At zero temperature, the locking condition holds across this span of length
L. We can replace the sine-Gordon term in Eq. (7) by a mass term,
Hb,M =
M 2
2 Z L/2
−L/2
dx (cid:20) θ1(x) − θ2(x) − c0
√2
(cid:21)2
.
(28)
The constant c0 can be absorbed by shifting θ2 → θ2 − c0; the locking condition becomes θ1 = θ2.
The current I1 can be expressed in terms of retarded Green's functions,
e2
πZ L/2
hI1(x)i = i
12 (ω; x, x′)i , (29)
where E is the external electric field in the region −L/2 ≤ x ≤ L/2, and η1,2(x) are the random forward scattering
potentials. The retarded Green's functions in the above formula are determined by
[∂x′η2(x′)]hω G(R)
11 (ω; x, x′)i − i
dx′ [E − ∂x′η1(x′)] lim
ω→0hω G(R)
dx′ lim
ω→0
−L/2
−∞
e2
πZ ∞
ω2
v1(x)K1(x) + ∂xh v1(x)
K1(x) ∂xi − M 2(x)
M 2(x)
2
2
where
M 2(x)
2
ω2
vK + v
K ∂2
x − M 2(x)
2
G(R)
11 (ω; x, x′) G(R)
21 (ω; x, x′) G(R)
12 (ω; x, x′)
22 (ω; x, x′)
G(R)
= δ(x − x′)1,
1,
K1(x) =(K, −L/2 ≤ x ≤ L/2
x > L/2
v1(x) =(v,
−L/2 ≤ x ≤ L/2
x > L/2
M (x) =(M, −L/2 ≤ x ≤ L/2
vF ,
0,
x > L/2
(30)
(31)
(32)
(33)
The retarded Green function can be solved by imposing the following boundary conditions [15, 19]. (i) G(R)
ab
continuous everywhere in x. (ii) v1(x)
x ∼ x′,
K1(x) ∂x G(R)
11 , v1(x)
K1(x) ∂x G(R)
12 , ∂x G(R)
21 , and ∂x G(R)
is
22 are continuous for x 6= x′. (iii) For
K1(x)
K1(x)
(cid:20) v1(x)
(cid:20) v1(x)
h v
h v
K
K
∂x G(R)
∂x G(R)
11 (ω; x, x′)(cid:21)x=x′+0+ −(cid:20) v1(x)
12 (ω; x, x′)(cid:21)x=x′+0+ −(cid:20) v1(x)
K1(x)
∂x G(R)
∂x G(R)
11 (ω; x, x′)(cid:21)x=x′−0+
12 (ω; x, x′)(cid:21)x=x′−0+
∂x G(R)
∂x G(R)
21 (ω; x, x′)ix=x′+0+ −h v
22 (ω; x, x′)ix=x′+0+ −h v
K
K
K1(x)
∂x G(R)
∂x G(R)
21 (ω; x, x′)ix=x′−0+
22 (ω; x, x′)ix=x′−0+
= 0
= 1.
= 1,
= 0,
(34)
(35)
(36)
(37)
(iv) The retarded Green functions obey the outgoing wave conditions.
8
Expanding G(R)
ab
ditions, we find that
in each region in terms of propagating and/or evanescent waves and imposing all boundary con-
lim
ω→0hω G(R)
11 (ω; x, x′)i = lim
ω→0hω G(R)
12 (ω; x, x′)i = −i
K
2(1 + K)
.
The result is independent of x. The current expression in Eq. (29) becomes
hI1(x)i =
e2
π
K
2(1 + K)Z L/2
−L/2
dx′ [E − ∂x′ η1(x′)] −
e2
π
K
=
(1 + K)
e2
h
[EL − η1(L/2) + η1(−L/2)] −
(1 + K)
K
2(1 + K)Z ∞
K
dx′ [∂x′ η2(x′)]
−∞
[η2(∞) − η2(−∞)] ,
e2
h
(38)
(39)
where η1(L/2) = η1(−L/2) = 0 because the first edge is connected to free fermion leads, and η2(∞) = η2(−∞) due
to the periodic boundary condition. The conductance of the locked edges is then
G = hI1i
EL
=
K
1 + K
e2
h
.
(40)
Adding the contribution of the parallel edge carrying I′1 gives Eq. (1).
In the presence of the inter-edge Luttinger interactions, the Luttinger parameter in the region −L/2 ≤ x ≤ L/2 is
modified. The two-terminal conductance is unchanged if we assume that the distal part of the passive edge is much
longer than the interacting part, as above. Then this passive edge is effectively connected to the Luttinger liquid
leads with Luttinger parameter K, and this gives resistance (1/K)h/e2. For a finite total length l > L of the passive
edge, the value of the measured Luttinger parameter is between the intra-edge value K and the Luttinger parameter
for the symmetric mode, and it also depends on the ratio of the interacting region length L to the total length l.
SCATTERING RATES
At second order in the interaction coupling strengths, there are four distinct self-energies, ΠW , ΠU , Π+, and Π−.
Coupling constant mixing will occur at higher orders, but is prevented here by vertex operator charge neutrality
conditions. We are interested in the imaginary part of the retarded self-energies. In the long-wavelength and low-
energy limits, Im(cid:2)Π(R)(ω, k)(cid:3) ≈ −2ωΞ.
In this section we provide the explicit results for the scattering rates that enter the intra-edge and transconductivities
in Eqs. (10) and (13). The scattering rate due to Eq. (14) already appears in Eq. (15). Here γ = K 2βgη/2πv is a
parameter indicating the ratio of the effective disorder strength to the temperature.
In the clean limit (γ → 0), the Lorentzian distributions in Eq. (15) becomes delta functions. The scattering rate
reduces to
sin (πK)
Γ [−K − 3]
Γ [K + 2]
ΞW =(cid:16) W αK(cid:17)2 22K+1π2K+3
v2K+3β2K+1 K
where β is the inverse temperature. When T ≫ vkF , the clean scattering rate is proportional to T 2K+1.
2π i
2π i
In the presence of disorder and at low temperatures such that γ ≫ 1, the scattering rate becomes
v2K+3β2K+2
ΞW =(cid:16) W αK(cid:17)2 22Kπ2K+1
cosh (vβkF ) − cos (πK)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
√π Γ(cid:2)2 + K
2(cid:3) Γ(cid:2) K
Γ(cid:2) 5
2 + K(cid:3)
2 + i vβkF
2 + i vβkF
Γh 4+K
Γh 2−K
2π (cid:17)2
(cid:16) vβkF
2(cid:3) Γ(cid:2) 5
2(cid:3) Γ(cid:2) 1
K sin (πK)
2 + K
βγ/π
+ γ2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2 + K
2(cid:3) Γ [−K − 3]
.
(42)
2
,
(41)
The term in square brackets is independent of temperature, so that ΞW ∼ T 2K+2.
The inter-edge single particle backscattering Hamiltonian in Eq. (5) contains two terms, U1 and U2. U1 and U2
correct the dc conductivity in edge 1 and 2, respectively. The scattering rates are
9
ΞU,a =(cid:16) UaαK(cid:17)2 22K−3π2K+1
v2K+1β2K−1
1
K
Γ [1 − K]
Γ [2 + K]
2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
γ/π
+
γ/π
dy
×Z ∞
−∞
2π (cid:17)2
(cid:16)y − vβkF,a
In the high-temperature limit γ → 0 and vβkF,a → 0, ΞU is proportional to T 2K−1.
where Ua = Ua/π3/2α.
In
the low-temperature limit γ ≫ 1, ΞU is proportional to T 2K. In the main text, we assume two identical edges and
U1 = U2, so that ΞU,1 = ΞU,2 = ΞU .
2π (cid:17)2
(cid:16)y + vβkF,a
+ γ2
+ γ2
sin (πK)
cosh (2πy) − cos (πK)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Γ(cid:2) K
2 + iy(cid:3)
Γ(cid:2)− K
2 + iy(cid:3)
,
(43)
The scattering rates due to Eqs. (3) and (4) are given by
Γ [1 − 2K]
Γ[2K]
v4K−1β4K−3
Ξ± =(cid:16) U±α2K(cid:17)2 24K−3π4K
4π (cid:17)2
(cid:16)y − vβk±
×Z ∞
−∞
F
¯γ/π
2
where ¯γ = K 2βgη/πv, U± = ∓U±/2π2α2, and k±F = 2(kF 1 ± kF 2).
vβk±F → 0, Ξ± is proportional to T 4K−3. In the low-temperature limit ¯γ ≫ 1, it is proportional to T 4K−2.
cosh (2πy) − cos(2πK)(cid:12)(cid:12)(cid:12)(cid:12)
Γ [1 − K + iy](cid:12)(cid:12)(cid:12)(cid:12)
4π (cid:17)2
(cid:16)y + vβk±
dy
Γ [K + iy]
sin(2πK)
+ ¯γ2
+ ¯γ2
¯γ/π
+
F
In the high-temperature limit ¯γ → 0 and
,
(44)
|
1905.04034 | 1 | 1905 | 2019-05-10T09:42:03 | Stochastic dynamics and pattern formation of geometrically confined skyrmions | [
"cond-mat.mes-hall"
] | Ensembles of magnetic skyrmions in confined geometries are shown to exhibit thermally driven motion on two different time scales. The intrinsic fluctuating dynamics ($t\sim 1~$ps) is governed by short-range symmetric and antisymmetric exchange interactions, whereas the long-time limit ($t\gtrsim10\,$ns) is determined by the coaction of skyrmion-skyrmion-repulsion and the system's geometry. Micromagnetic simulations for realistic island shapes and sizes are performed and analyzed, indicating the special importance of skyrmion dynamics at finite temperatures. We demonstrate how the competition between skyrmion mobility and observation time directly affects the addressability of skyrmionic bits, which is a key challenge on the path of developing skyrmion-based room-temperature applications. The presented quasiparticle Monte Carlo approach offers a computationally efficient description of the diffusive motion of skyrmion ensembles in confined geometries, like racetrack memory setups. | cond-mat.mes-hall | cond-mat | a
Stochastic dynamics and pattern formation of geometrically confined skyrmions
Alexander F. Schaffer,1, ∗ Levente R´ozsa,2 Jamal Berakdar,1 Elena Y. Vedmedenko,2 and Roland Wiesendanger2
1Institut fur Physik, Martin-Luther-Universitat Halle-Wittenberg, D-06099 Halle (Saale), Germany
2Institut fur Nanostruktur- und Festkorperphysik, Universitat Hamburg, D-20355 Hamburg, Germany
(Dated: May 13, 2019)
Ensembles of magnetic skyrmions in confined geometries are shown to exhibit thermally driven motion on
two different time scales. The intrinsic fluctuating dynamics (t ∼ 1 ps) is governed by short-range symmetric
and antisymmetric exchange interactions, whereas the long-time limit (t (cid:38) 10 ns) is determined by the coac-
tion of skyrmion-skyrmion-repulsion and the system's geometry. Micromagnetic simulations for realistic island
shapes and sizes are performed and analyzed, indicating the special importance of skyrmion dynamics at finite
temperatures. We demonstrate how the competition between skyrmion mobility and observation time directly
affects the addressability of skyrmionic bits, which is a key challenge on the path of developing skyrmion-based
room-temperature applications. The presented quasiparticle Monte Carlo approach offers a computationally ef-
ficient description of the diffusive motion of skyrmion ensembles in confined geometries, like racetrack memory
setups.
Magnetic skyrmions1 -- 3 are quasiparticles which are con-
sidered as possible carriers of information for future storage
devices. Their specific chirality is determined by the antisym-
metric Dzyaloshinskii-Moriya exchange interaction (DMI)4,5.
The DMI can be induced by a broken inversion symmetry in
a crystal itself (e.g. MnSi)2 or by interfacing a heavy metal
layer (e.g. Pt, W, Ir) with a ferromagnetic material (e.g. Fe,
Co)6,7.
Conceptually the utilization of these topologically non-
trivial8 quasiparticles in so-called racetrack setups is of great
interest 9 -- 11: Skyrmions can be manipulated (written and
deleted)12,13 and addressed individually14 -- 16 on a magnetic
stripe, allowing a memory device extension from a pure
surface density into the third dimension by pushing the
quasiparticle-hole-train back and forth, e.g. by applying elec-
trical currents17,18. The low threshold driving current19 along
with the small size and high stability of the skyrmions are key
features of this concept.
To connect experimental and theoretical model systems
with technological applications, investigating the influence of
finite temperatures is of crucial importance. The bits on a
racetrack-based memory device need to fulfill two main fea-
tures, stability against external perturbations and addressabil-
ity. Both are affected by thermal fluctuations, as shown below.
The stability of skyrmions was examined in several publi-
cations over the last years20 -- 24. R´ozsa et al.21 investigated the-
oretically periodic two-dimensional Pd/Fe double-layers on
Ir(111) and determined the phase diagram as a function of ex-
ternal field and temperature, which includes field-polarized,
skyrmion lattice, spin spiral, fluctuation-disordered and para-
magnetic regions.
Skyrmion lifetimes in the fluctuation-
disordered regime were calculated. The lifetimes of iso-
lated skyrmions in racetrack geometries for Pd/Fe/Ir(111) and
Co/Pt(111) systems were investigated in Refs.22,23, and differ-
ent mechanisms were revealed for the collapse of a skyrmion
inside the track and at the boundary.
The diffusive motion of skyrmions at finite temperature
has also attracted significant research attention lately25 -- 27.
Previous studies concentrated on diffusion in infinite or ex-
tended geometries, but a clarification of the role of the sam-
ple shape still seems to be missing in the case where the sys-
tem size becomes comparable to that of the skyrmions. Ef-
fects of this kind directly impede the addressability, which
is indispensable when using skyrmions for storage technol-
ogy, e.g.
in racetrack memory devices. Since the number
of skyrmions during the diffusive motion remains constant,
quasiparticle models have been developed for their description
in this limit25,26,28,29, which are primarily based on the Thiele
equation30. The advantage of such a collective-coordinate de-
scription over micromagnetic or atomistic spin dynamics sim-
ulations is its significantly lower computation cost.
Due to the finite temporal resolution, imaging techniques
on the atomic length scale like spin-polarized scanning tun-
neling microscopy (SP-STM)31 or magnetic force microscopy
(MFM)32 cannot access temporally the diffusive motion of
the skyrmions.
Instead, the time-averaged skyrmion proba-
bility distribution is imaged which may well be different from
the zero-temperature configuration or a snapshot of a simula-
tion. Here we will discuss the formation, stability and address-
ability of diffusive skyrmion configurations. Using micro-
magnetic simulations, it is shown that the complex interplay
of the repulsive interaction between the skyrmions33,34 along
with the confinement effect of the nanoisland and the ther-
mally induced skyrmion diffusion leads to a pattern formation
of the skyrmion probability distribution on the nanosecond
time scale. A computationally efficient quasiparticle Monte
Carlo method is introduced, which is found to yield compa-
rable results to time-averaged micromagnetic simulations re-
garding the skyrmion probability distribution. The simple im-
plementation and high speed of such a method may make it
advantageous over micromagnetic simulations when the ac-
tual number of skyrmions has to be determined based on a
time-averaged experimental image.
RESULTS
Skyrmion stability. We study metastable skyrmions and
their motion and mobility at finite temperatures. At first the
phase space has to be explored with respect to the exter-
nal magnetic field and temperature. A detailed description
of the phase diagram for extended systems of ultrathin bi-
2
fields lead to shrunken skyrmions, which in turn enables the
magnetic island to host a larger number of quasiparticles. In
this regime the topological charge is always close to an in-
teger, with the small deviation caused by the tilting of the
spins at the edge of the sample. Reaching field values above
B ≈ 3.5 T, the system becomes completely field-polarized
and skyrmions are not stabilized anymore starting from a ran-
dom configuration, even at zero temperature. Previous calcu-
lations37 -- 39 indicated that skyrmions on the lattice collapse at
around B = 4.5 T in the system.
In order to prevent the appearance of spin spiral states and
to be able to consider the skyrmions as well-defined quasi-
particles, we choose B = 1.5 T for the following calcula-
tions, unless mentioned otherwise. Figure 1 shows that sev-
eral metastable configurations associated with different topo-
logical charges are accessible starting from randomly gener-
ated initial states. However, in this approach not all stable
structures are covered, as the included ones are obtained from
relaxing random initial configurations. Different configura-
tions may also be generated in a controlled way by adding
skyrmions to the field-polarized state one-by-one, and relax-
ing the state at zero temperature. This is shown in Fig. 2 for
the disk-shaped and a triangular geometry for B = 1.5 T.
On top of the transitions due to a variation of the magnetic
field, the impact of finite temperature is also of crucial im-
portance.
In Fig. 3a temperature effects on the topological
charge are displayed. Starting from a relaxed configuration
at T = 0 K, the temperature is increased every 500 ps by
∆T = 2 K and the topological charge is averaged over time
at each temperature. The results show a first discontinuity at
T1 ≈ 40 K. Up to this temperature the thermal fluctuations are
relatively weak, and the number of well-defined skyrmions in-
side the sample remains constant. The small standard devia-
tion of the topological number in this regime can be attributed
to the thermal motion of the spins at the edge. However,
the combination of thermal fluctuations and the repulsion be-
tween the skyrmions triggers the escape of a single skyrmion
out of the sample at T1 ≈ 40 K. Hence, the total topological
number is changed from NSk ≈ −5 to NSk ≈ −4, see Sup-
plementary Video 1. This change is also shown in the plot of
the topological number susceptibility in Fig. 3b, defined as
(cid:0)(cid:104)N 2
Sk(cid:105) − (cid:104)NSk(cid:105)2(cid:1) ,
(1)
χN =
1
T
(cid:82) t1
t0
with (cid:104)NSk(cid:105) = 1
NSk(t)dt. Only a minor deviation
t1−t0
from the otherwise smooth function is visible. This means that
the thermal fluctuations are still not strong enough to perturb
the quasiparticles drastically.
A first characteristic change in the slope of the topological
susceptibility can be seen at T2 ≈ 50K. Comparing Figs. 3a
and b it is obvious that the fluctuations of topological num-
ber are increasing with temperature, and around this point the
lifetime of skyrmions is reduced below the averaging window.
The average topological charge continuously approaches zero
above this temperature. In this disordered regime the lifetime
of skyrmions is shorter than the observation time, as the fluc-
tuations allow a collapse inside the sample. Hence, this tem-
perature range is not favorable for information storage appli-
FIG. 1. Field dependence of the total topological charge. Red re-
gion: external fields below the strip-out instability leading to spin
spiral segments; dark green:
transition regime; bright green: en-
semble of skyrmions; blue: completely field-polarized island, where
skyrmions collapse due to the strong external field. Solid red line
indicates the magnetic field B = 1.5 T used for subsequent calcula-
tions. For each step in the vertically applied field 20 randomly mag-
netized configurations of a 21 nm diameter nanodisk were relaxed
towards the nearest metastable state on the energy hypersurface at
T = 0 K.
atomic layers of Pd/Fe on an Ir(111) substrate hosting mag-
netic skyrmions can be found, e.g., in Refs.21,35. Here, simula-
tions are performed for the same system in the micromagnetic
framework, by solving the Landau -- Lifshitz -- Gilbert equation
(LLG)36, as described in the Methods section.
In Fig. 1 possible equilibrium spin configurations are in-
vestigated in a 21-nm-diameter nanodisk of 0.4 nm thickness
at zero temperature. For each fixed external magnetic field
(B = Bz), 20 different randomly generated initial states
are considered and relaxed towards the closest local mini-
mum of the total energy hypersurface. Subsequently, the total
skyrmion number NSk of the relaxed states is calculated and
plotted against the external magnetic field in Fig. 1 to gain an
overview of the possible metastable states. Below the strip-
out instability field37, skyrmions as localized cylindrical spin
structures are not stable and the configuration consists of spin
spiral segments. For these the topological number is not a
well-defined quantity, as spin spiral structures can extend over
the whole island, and therefore the boundaries have a signif-
icant effect. This is reflected by the continuous distribution
of topological charge values found below 0.75 T in the red
regime in Fig. 1, whereas discrete skyrmion numbers corre-
spond to discrete steps in the total topological charge. The
upper boundary of this region is reasonably close to the strip-
out field of B = 0.65 T37,38 determined for extended systems.
With increasing the external field, individual skyrmions
may be stabilized in the system, where they will coexist with
spin spiral segments. This corresponds to the darker green
area in Fig. 1 around B ≈ 1 T, where besides the continuous
distribution also discrete steps can be observed in the topolog-
ical charge.
In the following bright green region stronger magnetic
01234-8-6-4-2024B(T)NSkB03
FIG. 3. Temperature dependence of the topological number.
Disk-shaped island (21 nm diameter) for B = 1.5 T. a Total topo-
logical number (cid:104)NSk(cid:105) averaged over 0.5 ns with error bars indicat-
ing the standard deviation. b Topological number susceptibility χN
(Eq. (1)), with the dashed black lines indicating three characteristic
temperatures discussed in the main text. Solid red line marks the
temperature used for successive simulations.
Skyrmion diffusion. With these first results, giving in-
formation on the segment of the parameter space we deal
with, we will focus on the influence of thermal fluctuations
on the dynamics of the skyrmions. In Fig. 4a and 5 results
for different initial numbers of skyrmions presented in Fig. 2
are shown for the circular and triangular geometries, respec-
tively. The first row shows the magnetic configuration after
20 ps of simulation time and the second row the final state,
which means the magnetization state after our total simula-
tion time of 20 ns. Only the out-of-plane z component is
shown in grayscale. The magnetic configuration after only
20 ps is qualitatively very similar to the final one, displaying
slightly deformed skyrmion configurations due to the thermal
fluctuations compared to the zero-temperature initial states in
Fig. 2. This short time scale on which the short-range Heisen-
berg and Dzyaloshinskii-Moriya interaction dominate the dy-
namics and cause shape deformations, is examined more ex-
tensively below.
The third row corresponds to the time-averaged z compo-
nent, calculated over the simulation time divided into 1000
snapshots.
In our understanding this result comes as close
as possible to real-space scanning-probe measurements, due
to the limited time resolution and finite measurement dura-
tion in such techniques. According to Ref.37, typical limits
FIG. 2. Relaxed states for different initial skyrmion numbers.
Brightness corresponds to the out-of-plane component of the mag-
netization characterized by the polar angle θ, whereas the color de-
notes the in-plane orientation described by the azimuthal angle ϕ,
as shown in the color map. B = 1.5 T, T = 0 K. System size: a
21 × 21 × 0.4 nm3; b 30 × 30 × 0.4 nm3. The saturation magneti-
zation is locally set to zero in the dark gray areas.
cations.
Finally, at about T3 ≈ 100K the paramagnetic regime is en-
tered, which is characterized by a completely disordered time-
averaged magnetic configuration. Here the average topologi-
cal charge is very close to zero. This behavior is also indicated
by a change of sign in the first derivative dχN (T )/dTT =T 3.
In the following, the external parameters are fixed to T =
T0 = 15 K and B = B0 = 1.5 T, shown by the solid red
lines in Figs. 1 and 3. This ensures that the thermal fluctu-
ations do not lead to a collapse or escape for the skyrmions,
only to a diffusive motion, and that the number of particle-like
skyrmions will be a constant during the simulation time.
of the time resolution in SP-STM are tres > 5 ms. With the
present simulation parameters we found that increasing the
simulation time further does not lead to a significant change in
the time-averaged images, indicating that similar observations
may be expected on the much longer experimental time scales
as well. The observed two characteristic time scales can be
attributed to different interaction types of clearly distinguish-
able energy scales: local deformations on the picosecond time
scale of the magnetic texture are dominated by thermal excita-
tions competing with the short-range symmetric and antisym-
metric exchange interactions, while the translational motion
leading to the complex pattern formation over tens of nanosec-
onds is caused by the repulsive interactions between pairs of
skyrmions and between skyrmions and the boundaries.
In case of the disk, it is extremely challenging to make a
clear statement about the number of skyrmions in the system
based on the time-averaged images only. As the symmetry
of the system is a cylindrical one, also the resulting picture is
cylindrically symmetric, consisting of concentric bright and
dark circles and rings. The small contrast differences along
the angular direction within a single diffusive ring-like area
(cf.
third row in Fig. 4a) arise because of the finite simula-
tion time, but also as an artifact of the finite grid, leading to a
deviation from the ideal radial symmetry. For all initial con-
figurations the total skyrmion number is conserved. Since the
skyrmions repulse each other, the radius of the resulting gray
ring increases for higher skyrmion densities, before one of
them becomes localized quite strongly in the center. This be-
havior is analyzed in Fig. 4b, where the out-of-plane magneti-
zation component depending on the distance from the center
of the disk is calculated by integration over the angular coor-
dinate. For comparison, the static profile of a single skyrmion
in the center of the disk is plotted as well (black curve). The
positions of the local minima of the integrated functions are
shown in Fig. 4c. Without performing the averaging over time,
the profile of the skyrmion rotates from mz = −1 in the cen-
ter towards mz = 1 in the field-polarized background. Due to
the diffusive motion of the skyrmions, the time-averaged im-
ages show a smaller difference compared to the background
magnetization at the approximate positions of the skyrmions,
providing a measure for the localized nature of the quasiparti-
cles in the sample. For an increasing number of skyrmions not
only the radius of the resulting ring-shaped pattern increases,
but so does the localization of the skyrmions.
For the triangular system in Fig. 5, the lower symmetry
of the geometry is reflected in the resulting time-averaged
pictures. The symmetry of the spin configuration coincides
with that of the sample for one, three and six skyrmions
without thermal fluctuations, as shown in Fig. 2b.
In the
time-averaged images in the third row of Fig. 5, this re-
sults in well-localized skyrmions with a strong dark contrast
for these numbers of quasiparticles in the system. Interest-
ingly, also the adjacent configurations show almost the same
time-averaged pattern as the highly symmetric configuration,
e.g.
two or four skyrmions compared to the case of three
skyrmions. For the case where one skyrmion is missing from
the highly ordered configuration, the remaining skyrmions im-
itate the absent skyrmion and effectively show up as an addi-
4
FIG. 4. Diffusive skyrmion motion on a circular island. a z com-
ponent of the magnetization (white +z, black −z) for an ascending
total skyrmion number. First row shows the configuration after 20 ps,
the second row after 20 ns of elapsed time, and the third row repre-
sents the time average for the total simulation time over 1000 snap-
shots. b Time-averaged signal as a function of the distance from the
center of the disk for different skyrmion numbers, averaged over the
angular coordinate. Black line shows the static profile of a central
single skyrmion. c Radii rmin corresponding to the local minima of
the time-averaged signal in panel b, in dependence on the skyrmion
number. Disk system (21 × 21 × 0.4 nm3), T = 15 K, B = 1.5 T.
The saturation magnetization is locally set to zero in the dark gray
areas.
tional "phantom skyrmion". In contrast, a surplus skyrmion
smears out in the time-averaged picture, so that mostly the
high-symmetry points remain observable, even though they
can also be slightly broadened -- cf. the 6th and 7th columns
in Fig. 5. For an even larger numbers of skyrmions, the repul-
sive interaction between them in combination with tempera-
ture fluctuations is strong enough for the skyrmions to escape
at the boundary during the simulation time, as shown in the
decreased number of quasiparticles in the second row com-
pared to the first row in the 8th and 9th columns in Fig. 5.
1Skprofilemin20ps20nsav.02468-1.0-0.50.00.51.05
FIG. 5. Diffusive skyrmion motion on a triangular island. z
component of the magnetization (white +z, black −z). First row
shows the configuration after 20 ps, the second row after 20 ns of
elapsed time, and the third row represents the time average for
the total simulation time over 1000 snapshots. Triangular system
(30 × 30 × 0.4 nm3), T = 15 K, B = 1.5 T. The saturation magne-
tization is locally set to zero in the dark gray areas.
To identify the different time regimes more clearly, the
convergence behavior of the time-averaged pictures is inves-
tigated. The time-integrated pictures are calculated up to
each snapshot time and compared to the averaged image after
20 nsvia a matching parameter, where a complete agreement
is denoted by a value of 1. The mathematical definition of
the matching parameter is given in the Methods. The results
are plotted in Fig. 6. Panels a and b show the convergence
behavior for different skyrmion occupations for the disk and
for the triangular geometry. For the circle there are no ma-
jor differences in terms of the convergence speed for different
skyrmion numbers. Only the ensembles of three and seven
skyrmions exhibit a slightly faster convergence, which is ex-
plicable by the resemblance of those states to the close-packed
arrangement. As expected, the close-packed array possesses
a lower mobility.
In the case of the triangle, for three and
six skyrmions the value 1 is reached much faster than for the
other configurations. This effect can again be explained by the
strong localization of the skyrmions in these cases, preventing
an exchange of their positions. A different trend is visible for
the eight-skyrmion case (purple line in Fig. 6b), which shows
a much slower convergence. This delayed behavior arises due
to the non-conserved skyrmion number, which can be seen
in Fig. 5. Consequently, after the escape of the skyrmions it
takes more time until the averaged picture has adapted to this
change.
For exploring the stochastic dynamics on the short time
scale, which we suppose to be governed by the skyrmion
deformations in contrast to the slow skyrmion displacement,
multiple simulations are performed for the same initial con-
figuration, two skyrmions inside the disk relaxed at zero tem-
perature, but with different pseudorandom number seeds. The
simulations cover a time of 100 ps each with 50 different ini-
tializations. After every 0.5 ps a snapshot is taken and the
positions of the two skyrmions extracted. On the scale of the
simulation time no noteworthy displacement of each of the
skyrmions takes place, however a deformation of the mag-
netic textures is visible. The mean skyrmion radius relative
to the zero temperature radius calculated as described in the
Methods, is averaged over the different simulations, shown in
Fig. 7. Here, two features are remarkable. Firstly, the radius
FIG. 6. Convergence behavior of the time-integrated magneti-
zation for different skyrmion numbers. The match between the
final time-averaged pattern after 20 ns and the time average taken
until time t is defined such that a value of 1 corresponds to perfect
agreement, see the Methods for details. a and b show the conver-
gence of the matching parameter for the disk and triangular setup,
respectively, for different skyrmion numbers.
increases, approaching an enhancement of about 10 % after
100 ps. Secondly, the radius is oscillating with a high fre-
quency of approximately 60 GHz, estimated from the average
time between the peaks. This oscillation can be identified with
internal skyrmion modes, like the so-called breathing modes
(cf. Ref.38: at T = 0 K, B = 1.5 T and the triangular atomic
lattice the breathing mode frequency is f = 70.88 GHz,
which is expected to decrease at higher temperature). These
findings support the result that in confined geometries, differ-
ent time scales are important for the skyrmion dynamics. On
the short time scale the internal modes and deformations of
the skyrmions are dominant, whereas on the longer time scale
the interaction leads to a slowly converging pattern formation.
All of the discussed examples showcase the strong influ-
ence of the interaction between skyrmions as well as of the ge-
ometrical aspects on their mobility, that can vary between de-
localization and a strong localization, e.g., in the center of the
disk in Fig. 4a. This feature of variable mobility, depending
on the geometry and packing density of skyrmions, is of gen-
eral nature. In real experimental systems further contributions
may affect the mobility, and ultimately the measured position
of the skyrmion as well. As an example of such a feature,
defects in the grown thin film nanoislands will be discussed.
In Fig. 8, a nanoisland geometry inspired by the experimental
results in Refs.7,12 is considered. Additionally to the previous
simulation parameters, a magnetic defect is inserted in the top-
left corner of the island, treated as a simulation cell with the
12345678920ps20nsav.6
FIG. 8. Diffusive skyrmion motion on an asymmetric nanoisland
including a defect. z component of the magnetization (white +z,
black −z) is imaged. In the top-left corner a defect is modeled by
freezing the magnetization direction in a single simulation cell along
the +x direction. a Snapshots of the magnetization at different times,
as well as the resulting time-averaged picture (av.). b Time-averaged
signal along the dashed arrow shown in the lower right part of panel
a. Simulation box: 30 × 30 × 0.4 nm3, T = 25 K, B = 1.5 T. The
saturation magnetization is locally set to zero in the dark gray areas.
skyrmion and the boundary needs to be quantified. Due to
the Neumann boundary condition40 at the edge of nanois-
lands, the DMI leads to a noncollinear spin texture, similar
to a skyrmion41. Therefore, the interaction mechanism is the
same between the skyrmions and between a skyrmion and the
boundary, as has been studied in previous publications33. For
details on the calculation and the explicit functions of the po-
tentials see the Methods section.
potentials
quasiparticle
Starting from the modeled skyrmion-skyrmion and
skyrmion-boundary
simulations
are executed, in order to compare the calculated probability
densities with the time-averaged configurations obtained from
full-fledged micromagnetic simulations. The main assump-
tion for this endeavor is that neither the thermal fluctuations
lead to the collapse, creation or escape of skyrmions, nor does
a high skyrmion density lead to a merging of the skyrmions
on the island as discussed in Ref.39. A triangular setup
containing two quasiparticles serves as the model system.
The side of the triangle is chosen to be 30 nm, the same as in
FIG. 7. Fast deformation dynamics of two skyrmions on a disk-
shaped island. Two skyrmions on a disk (21 × 21 × 0.4) nm3 are
initialized from a relaxed zero-temperature configuration. Afterward,
50 different seeds for the pseudorandom number generator are used
to simulate the skyrmion dynamics over 100 ps, saved every 0.5 ps.
The results are analyzed by calculating the centers of the skyrmions,
followed by the computation of the average skyrmion radius (see
Methods) for each of them as shown in a and b. Averaging over the
different simulations smears out most of the pure stochastic dynam-
ics, yielding an oscillating pattern superposed with a general trend
of an increased skyrmion radius compared to the zero-temperature
configuration. Bright blue points show data averaged over different
simulations, the solid blue line corresponds to data smoothened by a
Gaussian filter in order to calculate the peak positions (red dots).
magnetization frozen along the in-plane +x direction. The
initial configuration including 5 skyrmions is evolved in time
over a span of 20 ns, as shown in the snapshots in panel a.
During the time evolution one skyrmion is pinned at the de-
fect, whereas the others are moving rather freely. The result-
ing time average of the images reflects this behavior, as only
a single spot is visible with a well-defined position in the up-
per left corner, and the rest of the quasiparticle features form
a blurred trace mostly following the geometry of the island.
Additionally, the profile of the out-of-plane spin component
is shown in Fig. 8b along the dashed arrow in the lower right
panel of Fig. 8a.
In this representation the discrepancy be-
tween the measured profiles of the different skyrmions is even
more pronounced. Where the localized skyrmion appears as a
sharp dip almost reaching mz = −1, the superposition of the
other four skyrmions leads to broader and shallower features,
and therefore no clear indication of their positions.
Quasiparticle model. The localization of the quasiparti-
cles discussed above may also be interpreted as the probabil-
ity density of finding a skyrmion at a selected position over
the complete simulation time. Such a probabilistic interpreta-
tion is capable of describing the different contrast levels ob-
served for nominally equivalent skyrmions, and it proves to be
valuable in the interpretation of scanning-probe experimental
results where the characteristic measurement times are signif-
icantly longer than the time scale of the diffusive motion. In
this section, a simplified quasiparticle model is introduced,
motivated by the time-averaged data of the micromagnetic
simulations, offering a time-efficient opportunity in predict-
ing the complex pattern formation of the skyrmion probability
distribution discussed above.
For the purpose of using a quasiparticle approach, the re-
pulsive potential between pairs of skyrmions and between the
5ns10ns20nsav.10nmab051015202530-1.0-0.50.00.51.0y(nm)〈mz〉the case of the micromagnetic calculations shown in Figs. 2b
and 5. The motion of the skyrmions inside the potential is
calculated by modeling them as interacting random walkers,
using an implementation of a Markov-chain Monte Carlo
algorithm with Metropolis transition probabilities to study
their diffusion, described in detail in the Methods section.
This method offers a fast way of obtaining the cumulative
probability density function of the positions of the two
particles, which in turn can be compared to the time-averaged
4 and 5 obtained from micromagnetic
images in Figs.
simulations.
In the following, a quantitative comparison
between full micromagnetic simulations and Monte Carlo
calculations will be presented to review the validity of the
simplified quasiparticle approach. For this endeavor, we
tracked the positions of two skyrmions in the triangular
system from the micromagnetic simulation snapshots after
every 20 ps and calculated the probability density distribution
of the center coordinates. The resulting distribution along
with the Monte Carlo result after 105 Monte Carlo steps is
shown in Fig. 9. The largest difference between the two
results arises because of the substantially less data for the
micromagnetic calculation. As the histogram is calculated
with only 103 snapshots compared to 105 Monte Carlo steps,
this behavior is expected. Nevertheless, the characteristic
features are similar. In both results the three peak densities
are not radially symmetric, but are resembling the triangular
boundary shape. Hence this deformation is not an artifact
of the Monte Carlo simulation. Additionally, the average
distance between the density maxima may serve as a good
quantity for comparing both methods. As Fig. 9b lacks the
resolution for calculating the peak positions reliably, we start
instead from the time-integrated magnetization pictures in
Fig. 5. Based on the averaged result for two skyrmions, we
extract again the peak positions and calculate subsequently
the mean distance. The positions for the Monte Carlo calcu-
lations are at (11.0, 6.5) nm, (15.0, 13.5) nm, (19.0, 6.5) nm,
the micromagnetic
(11.1, 6.6) nm,
(15.0, 13.5) nm, (19.0, 6.5) nm. This leads to an average
distance of 8.0 nm for both the Monte Carlo calculations
and the micromagnetic simulations, supporting the validity
of the simplified quasiparticle approach in the investigated
parameter space.
simulations
deliver
The model is capable of predicting the skyrmion distribu-
tion in the long time limit and thereby serves as a method for
computationally efficient calculations for larger or more com-
plex systems. As an example, the probability density func-
tions of the Monte Carlo simulations are shown in Fig. 10 af-
ter 105 Monte Carlo steps for different temperatures and sam-
ple sizes. Due to the repulsion between the two skyrmions,
equilibrium positions are found close to the vertices of the
triangle, with energy barriers between these local energy min-
ima. At low temperature (T = 1 K) and a small island sizes
(30 nm edge length) in the upper left panel in Fig. 10, no tran-
sitions between the minima occur during the duration of the
simulation, and two skyrmions may be observed in the ob-
tained probability densities. As the temperature is increased
to T = 15 K or T = 30 K in the first row of Fig. 10, the
transition rate between the preferential positions is enhanced,
7
FIG. 9. Comparison of the Monte Carlo and micromagnetic sim-
ulations. The probability density function for two skyrmions inside
a triangular island as obtained from a: micromagnetic simulations,
and b: from calculations following the Monte Carlo algorithm. For
the micromagnetic distribution, the coordinates of the skyrmions are
extracted from the snapshots taken every 20 ps over 20 ns of simu-
lation time at T = 15 K. Panel b is generated over 105 Monte Carlo
steps.
Probability density function for two skyrmions in-
FIG. 10.
side a triangular model potential calculated following a Monte
Carlo algorithm. The temperatures are T = 1 K, T = 15 K and
T = 30 K, the edge lengths of the triangle are 30 nm, 45 nm and
60 nm. Probability distributions for the centers of the two skyrmions
are calculated over 105 Monte Carlo steps.
and an additional "phantom skyrmion" appears in the proba-
bility density function, in remarkable agreement with the mi-
cromagnetic results as analyzed before. By enlarging the size
of the island (second and third rows in Fig. 10 for 45 nm and
60 nm edge length), the energy surface becomes flatter, which
in turn leads to higher transition probabilities between the en-
ergy minima at the same temperature. Therefore, raising the
temperature or increasing the size of the system have similar
effects on the resulting probability density functions for the
skyrmions. Consequently, large island sizes or high tempera-
tures result in completely delocalized skyrmions as indicated
in the bottom right panel in Fig. 10.
solve the LLG equation,
mi(t) = − γ
1 + α2
mi × Beff
(cid:104)
i + αmi ×(cid:16)
(cid:17)(cid:105)
8
,
(2)
mi × Beff
i
for every simulation cell mi of the discretized magnetization
vector field. Here, γ0 = 1.76 × 1011(T−1s−1) is the gyro-
magnetic ratio of an electron and α is the Gilbert damping
parameter. The time- and space-dependent effective magnetic
field,
Beff
i (t) = Bext
i + Bexch
i + Bd
i + Ba
i + Bdmi
i + Bth
i (t) ,
(3)
i
i
;
the external field Bext
exchange in-
is composed of
teraction field Bexch
= 2Aexch/M sat∆mi, with Aexch
the saturation magne-
the exchange stiffness and M sat
i = M sat Kij ∗ mj,
tization; demagnetizing field Bd
where details on the calculation of
the demagnetiz-
ing kernel K can be found in Ref.40; uniaxial mag-
netocrystalline anisotropy field Ba
i = 2K u/M satmzez,
with K u the anisotropy constant; and the field gener-
ated by the Dzyaloshinskii-Moriya interaction Bdmi
=
2D/M sat(∂mz/∂x, ∂mz/∂y,−∂mx/∂x− ∂my/∂y)T , with
D the strength of the interfacial Dzyaloshinskii-Moriya in-
teraction. The simulation parameters were determined ex-
perimentally in Ref.7 based on the field dependence of the
skyrmion profile: M sat = 1.1 MA/m, D = 3.9 mJ/m2,
Aexch = 2 pJ/m, K u = 2.5 MJ/m3 and α = 0.05.
Additionally, an effective thermal field is included in Eq. (3)
i
as
(cid:115)
Bth
i = η
2αkBT
M satγ0∆V ∆t
,
(4)
where η is a random vector generated according to a standard
normal distribution independently for each simulation cell and
time step. kB is Boltzmann's constant, T is the temperature,
∆V is the size of the simulation cell and ∆t is the time step
of simulation.
During the simulations, the cell size was set to ∆V =
0.3 × 0.3 × 0.4 nm3. The linear size of the cell is comparable
to the lattice constant a = 0.271 nm of the Pd/Fe bilayer on
Ir(111). This means that the micromagnetic simulations per-
formed here should closely resemble the results of atomistic
simulations where the skyrmions collapse when their size be-
comes comparable to the lattice constant37,39, and where the
use of temperature-independent model parameters is justified.
The total skyrmion number N Sk in the simulations was cal-
culated via
(cid:90)
(cid:18) ∂m
∂x
m ·
× ∂m
∂y
(cid:19)
dxdy .
(5)
DISCUSSION
In this paper we studied the temperature-driven diffusive
motion of ensembles of magnetic skyrmions in finite mag-
netic islands. Based on experimentally determined system
parameters for ultrahin Pd/Fe bilayers on Ir(111), we per-
formed full-fledged micromagnetic simulations for various
nanoisland geometries. For moderate temperatures and exter-
nal magnetic fields, magnetic skyrmions with a long lifetime
are present in the system, so that the topological charge is con-
served. Our study of this model system showed two different
time scales on which the stochastic dynamics of skyrmions
takes place. On the short time scale (t ∼ 1 ps), the fluctu-
ating dynamics governed by the symmetric and antisymmet-
ric exchange interaction is most prominent, which essentially
leads to local deformations, exciting internal modes of the
skyrmions. In the long-time limit (t (cid:38) 10 ns), interactions
between pairs of skyrmions, as well as between the quasi-
particles and the boundaries, are dominant. These thermally
activated mechanisms lead to complicated time-averaged pic-
tures of the skyrmion distribution, where the number of quasi-
particles in the system is not immediately apparent. These
results can be compared directly to images obtained from
time-integrating measurement techniques such as SP-STM or
MFM, the time-resolution of which (∼ 5 ms37) is typically
much longer than the timescale of the inherent dynamics of
skyrmions. Our findings open an alternative way compared to
conventional interpretations of experimental results obtained
with scanning probe methods on mobile arrays of skyrmions,
by differentiating between the various time scales of magneti-
zation dynamics.
Furthermore, we proposed and developed a Monte Carlo
quasiparticle model, which can be of great use to efficiently
calculate the stochastic motion of larger systems of skyrmions
in arbitrarily shaped geometries. According to the results, the
time-averaged images of the micromagnetic simulations can
be qualitatively well reproduced by the probability density
distributions obtained from the simple quasiparticle model,
as long as it is possible to treat the skyrmions as stable ob-
jects with lifetimes significantly longer than the simulation
time at the given temperature. This method can be used to
describe larger systems containing more skyrmions at vari-
ous temperatures in a computationally efficient way, including
model applications in storage technology like racetrack mem-
ory devices. Here, not only the stability of the information bit
is important, but also its addressability, which is immediately
affected by the diffusive motion of the skyrmions.
METHODS
NSk =
1
4π
Micromagnetic
simulations. Finite-temperature mi-
cromagnetic calculations, using the open-source, GPU-
accelerated software package mumax340, were performed to
Processing of the micromagnetic results. For the results
shown in Fig. 6, the z component of the magnetization aver-
aged up to time t is stored in grayscale pictures as a matrix,
Aij(t) = (mz(xi, yj, t) + 1)/2, where xi and yj denote the
position of the micromagnetic simulation cell. The distance
induced by the Frobenius norm between the image averaged
up to time t, Aij(t), and the image averaged over the whole
simulation length τ = 20 ns, Aij(τ ) is divided by the square
root of the number of matrix elements and subtracted from 1,
yielding the matching parameter
M (t) = 1 −
(Aij(t) − Aij(τ ))2
NxNy
.
(6)
(cid:118)(cid:117)(cid:117)(cid:116) Nx(cid:88)
Ny(cid:88)
i=1
j=1
This procedure gives a matching parameter of 1 for a perfect
agreement of the compared pictures.
In order to determine the skyrmion radius as a function of
time in Fig. 7, firstly the contour lines where the z component
of the magnetization is zero are calculated for the skyrmion at
each time step. The contour lines are discretized on N points
in space, Li(t) = (xi(t), yi(t)), i ∈ (1, ..., N ). Subsequently,
the center of the skyrmion is calculated via
Li(t) ,
(7)
N(cid:88)
i=1
c(t) =
1
N
N(cid:88)
i=1
which in turn is used to obtain the average skyrmion radius
¯r(t) =
1
N
Li(t) − c(t) .
(8)
Finally, the radius is normalized with respect to the zero tem-
perature radius r0 obtained from the relaxed initial configura-
tion of choice.
Calculation of the skyrmion-skyrmion and skyrmion-
boundary interaction potentials. For the calculation of the
potential, the total energy of a stripe-shaped model system
(75 × 30 × 0.4 nm3) was investigated with micromagnetic
simulations for two different cases.
In the first case, Neu-
mann boundary conditions were used along the long side, and
a homogeneous magnetization was relaxed leading to a canted
rim. Afterwards a skyrmion was added to the system at a cer-
tain distance from the boundary. Because of the repulsive na-
ture of the interaction, a standard relaxation of the skyrmion
position was not suitable for determining the distance depen-
dence. Instead, we fixed the central magnetic moments in the
skyrmion (in a circular region of 0.9 nm diameter) at a specific
position and relaxed the magnetic texture in the other cells.
Analogously, in the second case the skyrmion-skyrmion inter-
action was calculated by fixing one skyrmion and changing
the position of the other skyrmion, using periodic boundary
conditions along both directions in the plane. From the ob-
tained total energies we subtracted a reference spectrum for a
homogeneously magnetized periodic surrounding around the
skyrmion fixed at different positions in the mesh. Small devi-
ations arising due to the finite mesh are mostly canceled out
by this procedure.
The results for the interaction strengths are shown in
Fig. 11. A smooth energy function rapidly decaying with
9
increasing distance between skyrmion and boundary is ob-
tained. For small distances between the boundary and a
skyrmion or between two skyrmions, a deformation of the
quasiparticles is visible in the micromagnetic simulations,
also leading to a bending of the potential in Fig. 11. Dur-
ing the construction of the quasiparticle model it is assumed
that the thermal energy kBT is lower than the energy value
where the potential function starts to bend, which is around
5 meV (60 K) for the skyrmion-skyrmion interaction potential
in Fig. 11.
In this regime it can be assumed that for suffi-
ciently short simulation times the strong deformation of the
skyrmions due to the interactions becomes exceedingly rare,
and the basic spin structure is maintained. This temperature
regime is in reasonable qualitative agreement with the range
where the escape, collapse or creation of skyrmions may also
be excluded, discussed in detail in Fig. 3.
It was shown
in Ref.33 that the interaction potentials in this regime may
be approximated by the exponential model function E(r) =
a exp(−r/b). This is confirmed by the simulation results in
Fig. 11, with the fitted model parameters a = 0.211 meV,
b = 1.343 nm for skyrmion-boundary (Sk-Bnd) repulsion and
a = 1.246 meV, b = 1.176 nm for skyrmion-skyrmion (Sk-
Sk) interaction. The similar characteristic length scales b be-
tween the two cases support that the same physical mecha-
nism is responsible for the interactions, namely the formation
of chiral noncollinear spin structures due to the DMI as dis-
cussed in the main text. The different values of the parameter
a can be reformulated as a shift of the exponential function
along the r axis. While this distance is measured between two
magnetization cells pointing oppositely to the homogeneous
background in the case of two skyrmions, in the case of the
skyrmion-boundary repulsion it is measured between the cen-
ter of the skyrmion and the magnetization at the edge which is
only slightly tilted from the homogeneous background, mean-
ing that the same interaction strength is reached at a smaller
distance in the latter scenario.
Quasiparticle simulations. For the calculation of the
stochastic motion of skyrmions following the quasiparticle ap-
proach, the Metropolis algorithm42 was utilized. With this
method,
the probability density function converges to the
Boltzmann distribution determined by the energy functional
of the system. For the triangular geometry, the potential sur-
face was computed by taking the superposition of the poten-
tials shown in Fig. 11 from the three boundaries for every grid
point of the chosen finite, rectangular mesh. The cell size
was ∆V = 0.5 × 0.5 × 0.4 nm. The initial positions of the
skyrmions were randomly chosen inside the confined struc-
ture, excluding the case in which both quasiparticles start from
the same simulation cell. At each time step, a possible adja-
cent position is selected for each skyrmion simultaneously via
pseudo-random numbers. If the energy of the attempted new
state, including interaction between the skyrmions, is lower
than the energy of the initial one, the skyrmion will move
there. If not, the transition into this state happens with a prob-
ability of p(∆E) = exp(−∆E/kBT ) following the Boltz-
mann distribution. Here ∆T = E2 − E1 is the energy dif-
ference between the final and the initial states, kB is Boltz-
mann's constant and T is the temperature. Subsequently, new
attempted positions are generated and accepted or refused as
before until a fixed number of simulation steps is reached.
10
FIG. 11. Potential energy depending on the skyrmion-skyrmion
and skyrmion-boundary distance. Micromagnetic interaction en-
ergies between skyrmion and skyrmion (blue squares), and be-
tween skyrmion and boundary (red circles) on a finite nanoisland.
Solid lines display exponential fits based on the model function
E(r) = ae−r/b; the fitting parameters are skyrmion-boundary (Sk-
Bnd): a = 0.211 meV, b = 1.343 nm; skyrmion-skyrmion (Sk-Sk):
a = 1.246 meV, b = 1.176 nm.
∗ Corresponding author. alexander.schaeffer@physik.uni-halle.de
1 Bogdanov, A. & Yablonskii, D. Thermodynamically stable vor-
tices in magnetically ordered crystals. The mixed state of magnets.
Zh. Eksp. Teor. Fiz 95, 178 -- 182 (1989).
2 Muhlbauer, S. et al. Skyrmion lattice in a chiral magnet. Science
323, 915 -- 919 (2009).
3 Nagaosa, N. & Tokura, Y. Topological properties and dynamics
of magnetic skyrmions. Nat. Nanotechnol. 8, 899 -- 911 (2013).
4 Dzyaloshinsky, I. A thermodynamic theory of weak ferromag-
netism of antiferromagnetics. J. Phys. Chem. Sol. 4, 241 -- 255
(1958).
5 Moriya, T. Anisotropic superexchange interaction and weak fer-
romagnetism. Phys. Rev. 120, 91 -- 98 (1960).
6 Heinze, S. et al. Spontaneous atomic-scale magnetic skyrmion
lattice in two dimensions. Nat. Phys. 7, 713 -- 718 (2011).
7 Romming, N., Kubetzka, A., Hanneken, C., von Bergmann, K.
& Wiesendanger, R. Field-dependent size and shape of single
magnetic skyrmions. Phys. Rev. Lett. 114, 177203 (2015).
8 Bogdanov, A. & Hubert, A. The stability of vortex-like structures
in uniaxial ferromagnets. J. Magn. Magn. Mater. 195, 182 -- 192
(1999).
9 Parkin, S. S., Hayashi, M. & Thomas, L. Magnetic domain-wall
racetrack memory. Science 320, 190 -- 194 (2008).
10 Sampaio, J., Cros, V., Rohart, S., Thiaville, A. & Fert, A. Nu-
cleation, stability and current-induced motion of isolated mag-
netic skyrmions in nanostructures. Nat. Nanotechnol. 8, 839 -- 844
(2013).
11 Tomasello, R. et al. A strategy for the design of skyrmion race-
track memories. Sci. Rep. 4, 6784 (2014).
12 Romming, N. et al. Writing and deleting single magnetic
skyrmions. Science 341, 636 -- 639 (2013).
13 Schaffer, A. F., Durr, H. A. & Berakdar, J. Ultrafast imprinting
of topologically protected magnetic textures via pulsed electrons.
Appl. Phys. Lett. 111, 032403 (2017).
14 Hanneken, C. et al. Electrical detection of magnetic skyrmions
by non-collinear magnetoresistance. Nat. Nanotechnol. 10, 1039 --
1042 (2015).
15 Gobel, B., Mook, A., Henk, J. & Mertig, I. Magnetoelectric ef-
fect and orbital magnetization in skyrmion crystals: Detection and
characterization of skyrmions. Phys. Rev. B 99, 060406 (2019).
16 Maccariello, D. et al. Electrical detection of single magnetic
skyrmions in metallic multilayers at room temperature. Nat. Nan-
otechnol. 13, 233 -- 237 (2018).
17 Slonczewski, J. C. Current-driven excitation of magnetic multi-
layers. J. Magn. Magn. Mater. 159, L1 -- L7 (1996).
18 Iwasaki, J., Mochizuki, M. & Nagaosa, N. Universal current-
velocity relation of skyrmion motion in chiral magnets. Nat. Com-
mun. 4, 1463 (2013).
19 Jonietz, F. et al. Spin transfer torques in MnSi at ultralow current
densities. Science 330, 1648 -- 1651 (2010).
20 Hagemeister, J., Romming, N., von Bergmann, K., Vedmedenko,
E. & Wiesendanger, R. Stability of single skyrmionic bits. Nat.
Commun. 6, 8455 (2015).
21 R´ozsa, L., Simon, E., Palot´as, K., Udvardi, L. & Szunyogh, L.
Complex magnetic phase diagram and skyrmion lifetime in an ul-
trathin film from atomistic simulations. Phys. Rev. B 93, 024417
(2016).
22 Lobanov, I. S., J´onsson, H. & Uzdin, V. M. Mechanism and activa-
tion energy of magnetic skyrmion annihilation obtained from min-
imum energy path calculations. Phys. Rev. B 94, 174418 (2016).
23 Bessarab, P. F. et al. Lifetime of racetrack skyrmions. Sci. Rep. 8,
3433 (2018).
24 von Malottki, S., Dup´e, B., Bessarab, P., Delin, A. & Heinze,
S. Enhanced skyrmion stability due to exchange frustration. Sci.
Rep. 7, 12299 (2017).
25 Schutte, C., Iwasaki, J., Rosch, A. & Nagaosa, N. Inertia, diffu-
sion, and dynamics of a driven skyrmion. Phys. Rev. B 90, 174434
(2014).
26 Miltat, J., Rohart, S. & Thiaville, A. Brownian motion of mag-
netic domain walls and skyrmions, and their diffusion constants.
Phys. Rev. B 97, 214426 (2018).
27 Z´azvorka, J. et al. Thermal skyrmion diffusion used in a reshuffler
device. Nat. Nanotechnol. 1 (2019).
28 Lin, S.-Z., Reichhardt, C., Batista, C. D. & Saxena, A. Particle
model for skyrmions in metallic chiral magnets: Dynamics, pin-
ning, and creep. Phys. Rev. B 87, 214419 (2013).
29 Reichhardt, C., Ray, D. & Reichhardt, C. J. O. Collective transport
properties of driven skyrmions with random disorder. Phys. Rev.
Lett. 114, 217202 (2015).
30 Thiele, A. A. Steady-state motion of magnetic domains. Phys.
Rev. Lett. 30, 230 -- 233 (1973).
31 Wiesendanger, R. Spin mapping at the nanoscale and atomic
scale. Rev. Mod. Phys. 81, 1495 -- 1550 (2009).
32 Martin, Y. & Wickramasinghe, H. K. Magnetic imaging by force
microscopywith 1000 A resolution. Appl. Phys. Lett. 50, 1455 --
1457 (1987).
33 Lin, S.-Z., Reichhardt, C., Batista, C. D. & Saxena, A. Particle
model for skyrmions in metallic chiral magnets: Dynamics, pin-
ning, and creep. Phys. Rev. B 87, 214419 (2013).
34 R´ozsa, L. et al. Skyrmions with attractive interactions in an ultra-
thin magnetic film. Phys. Rev. Lett. 117, 157205 (2016).
35 Bottcher, M., Heinze, S., Egorov, S., Sinova, J. & Dup´e, B. B -- T
phase diagram of Pd/Fe/Ir (111) computed with parallel tempering
Monte Carlo. New J. Phys. 20, 103014 (2018).
36 Gilbert, T. L. A phenomenological theory of damping in ferro-
magnetic materials. IEEE Trans. Magn. 40, 3443 -- 3449 (2004).
37 Leonov, A. et al. The properties of isolated chiral skyrmions in
thin magnetic films. New J. Phys. 18, 065003 (2016).
38 R´ozsa, L., Hagemeister, J., Vedmedenko, E. Y. & Wiesendanger,
R. Localized spin waves in isolated k π skyrmions. Phys. Rev. B
98, 224426 (2018).
39 Siemens, A., Zhang, Y., Hagemeister, J., Vedmedenko, E. &
Wiesendanger, R. Minimal radius of magnetic skyrmions: stat-
ics and dynamics. New J. Phys. 18, 045021 (2016).
40 Vansteenkiste, A. et al. The design and verification of MuMax3.
AIP Advances 4, 107133 (2014).
11
41 Rohart, S. & Thiaville, A. Skyrmion confinement in ultrathin film
nanostructures in the presence of Dzyaloshinskii-Moriya interac-
tion. Phys. Rev. B 88, 184422 (2013).
42 Metropolis, N. et al. Equation of state calculations by fast com-
puting machines. J. Chem. Phys. 21, 1087 -- 1092 (1953).
ACKNOWLEDGEMENTS
Financial
support was provided by the Deutsche
Forschungsgemeinschaft
(DFG) via CRC/TRR 227 and
SFB 762, by the European Union via the Horizon 2020
research and innovation program under Grant Agreement
No. 665095 (MAGicSky), by the Alexander von Humboldt
Foundation, and by the National Research, Development and
Innovation Office of Hungary under Project No. K115575.
AUTHOR CONTRIBUTIONS
A.F.S. conducted and analyzed the micromagnetic and
quasiparticle Monte Carlo simulations. E.Y.V. proposed and
developed a general concept of this investigation. All authors
discussed the results and contributed to the manuscript.
COMPETING INTERESTS STATEMENT
The authors declare no competing financial interests.
DATA AVAILABILITY
The code and the data that support this work's findings are
available from the corresponding author on request.
|
1608.03879 | 1 | 1608 | 2016-08-12T19:21:14 | Magnetotransport signatures of the proximity exchange and spin-orbit couplings in graphene | [
"cond-mat.mes-hall"
] | Graphene on an insulating ferromagnetic substrate---ferromagnetic insulator or ferromagnetic metal with a tunnel barrier---is expected to exhibit giant proximity exchange and spin-orbit couplings. We use a realistic transport model of charge-disorder scattering and solve the linearized Boltzmann equation numerically exactly for the anisotropic Fermi contours of modified Dirac electrons to find magnetotransport signatures of these proximity effects: proximity anisotropic magnetoresistance, inverse spin-galvanic effect, and the planar Hall resistivity. We establish the corresponding anisotropies due to the exchange and spin-orbit couplings, with respect to the magnetization orientation. We also present parameter maps guiding towards optimal regimes for observing transport magnetoanisotropies in proximity graphene. | cond-mat.mes-hall | cond-mat | a
Magnetotransport signatures of the proximity exchange and spin-orbit couplings in
graphene
Jeongsu Lee∗ and Jaroslav Fabian
Institute for Theoretical Physics, University of Regensburg, 93040 Regensburg, Germany
Graphene on an insulating ferromagnetic substrate -- ferromagnetic insulator or ferromagnetic
metal with a tunnel barrier -- is expected to exhibit giant proximity exchange and spin-orbit cou-
plings. We use a realistic transport model of charge-disorder scattering and solve the linearized Boltz-
mann equation numerically exactly for the anisotropic Fermi contours of modified Dirac electrons
to find magnetotransport signatures of these proximity effects: proximity anisotropic magnetoresis-
tance, inverse spin-galvanic effect, and the planar Hall resistivity. We establish the corresponding
anisotropies due to the exchange and spin-orbit coupling, with respect to the magnetization orien-
tation. We also present parameter maps guiding towards optimal regimes for observing transport
magnetoanisotropies in proximity graphene.
PACS numbers: 72.20.Dp 72.80.Vp 73.22.Pr 73.63.-b
Dirac electrons in pristine graphene have wesizableak
spin-orbit coupling [1] and no magnetic moments, lim-
iting prospects for spintronics [2]. This can be partly
remedied by functionalizing graphene with adatoms and
admolecules, which can induce sizable local magnetic mo-
ments and spin-orbit coupling, leading to marked spin
transport fingerprints [3 -- 8]. A more systematic and, im-
portant, spatially uniform way to induce spin properties
in graphene is by proximity effects. Being essentially a
surface (or two), graphene can "borrow" properties from
its substrates. Placing graphene on a slab of a ferromag-
netic insulator, or a ferromagnetic metal with a tunnel
barrier, is expected to induce giant proximity exchange
as well as spin-orbit coupling in the Dirac electron band
structure. This is supported by first principles calcula-
tions [9 -- 13] as well as by recent experiments on graphene
on yttrium iron garnet [14, 15] and on graphene on EuS
[16]. In effect, proximity graphene on ferromagnetic sub-
strates should be an ultimately thin ferromagnetic layer,
with giant spin-orbit coupling, forming a perfect play-
ground for both spintronics experiment and theory [17].
An important question is: what transport ramifica-
tions can we expect in such a magnetic graphene with
strong spin-orbit coupling? On one hand, in ferromag-
netic metals the exchange coupling is typically much
greater than spin-orbit coupling. On the other hand,
in semiconductor heterostructures, which are the best
case studies for structure-induced spin-orbit coupling in
its transport signatures [18, 19], there is no ferromag-
netic exchange and spin splitting can be due to the Zee-
man interaction which is, for realistic values of magnetic
field, much weaker than spin-orbit coupling. Proxim-
ity graphene should be intermediate between those two
extremes: the proximity exchange and spin-orbit cou-
plings are expected to be similar, on the order of 1 - 10
meV [9, 10, 20]. Perhaps the main effect of the interplay
of exchange and spin-orbit couplings -- magnetotransport
anisotropies -- should be well pronounced and make for
useful, experimentally testable signatures of the spin
proximity effects.
In this paper we solve numerically exactly a realistic
Boltzmann transport model, with long-range charge scat-
terers, for Dirac electrons in the presence of proximity
exchange and spin-orbit couplings. We start with the
anisotropic band structure, as in Fig. 1, and explore its
ramifications in transport. Specifically, we introduce and
calculate the proximity anisotropic magnetoresistance, as
an analog of the anisotropic lateral magnetoresistance
in ferromagnetic metal/insulator slabs [21], characteriz-
ing interfacial spin-orbit fields. We also present magne-
toanisotropies of the planar Hall effect and inverse spin-
galvanic effect. Finally, we give parameter maps indicat-
FIG. 1. (Color online). Scheme of magnetoanisotropic trans-
port experiment in proximity graphene. Polar (θ) and az-
imuthal (φ) angles define the magnetization orientation with
respect to the applied electric field. (a) Linear energy dis-
persion of pristine graphene can be modified by (b) (intrin-
sic and Bychkov-Rashba) spin-orbit coupling or (c) exchange
field, both leading to spin splitting. (d) The interplay of the
two interactions makes the bands anisotropic with respect to
the magnetization orientation, here shown as out-of-plane and
in-plane.
ESOC (a) intrinsic(b) proximity SOC(c) proximity exchange(d) SOC and exchangeEEESOC + exchange exchange Ein-planeout-of-planeMagnetic Insulator mI ✓EmzkEkcontours shifted relative to each other.
2
To investigate electrical transport, we solve the lin-
earized Boltzmann equation for the above model, assum-
ing spatial homogeneity. In the presence of a longitudinal
electric field E, the non-equilibrium distribution function
is f = f0 + δf , where f0 is the equilibrium Fermi-Dirac
function. We use the ansatz δf = −e(−∂f0/∂E)u·E and
consider long-range Coulomb scattering, which is the es-
tablished model for resistivity in graphene [23]. The un-
known vector u is found by solving the integral equation
(obtained from the Boltzmann equation in linear order
in E),
v(k) = 2πni(vF rs)2
×(cid:72)
(cid:48)
dk
∇k(cid:48) Ek(cid:48) F (k,k
q2ε(q)2 [u(k) − u(k(cid:48))] ,
(cid:48)
)
EF
(2)
where Ek is the energy corresponding to wave vector k,
v(k) is the group velocity, ni is the concentration of scat-
terers, the effective fine structure constant rs ≈ 0.8 [24],
F (k, k(cid:48)) = Ψ(k)†Ψ(k(cid:48))2 is the overlap between the in-
cident (k) and scattered (k(cid:48)) states Ψ. For example, for
pristine graphene F (k, k(cid:48)) = (1 + cos θkk(cid:48))/2. For sim-
plicity, spin and pseudospin indices are implicit in the
momentum labels k. The integral is over the Fermi con-
tour of Fermi energy EF , and the transferred momentum
is q = k−k(cid:48). The dielectric function ε is calculated from
the random phase approximation [24 -- 26].
ε(q) =(cid:26) 1+ qs
q
1+ πrs
2 + qs
q − qs
√
q2−4k2
2q2
F
−rs sin
−1( 2kF
if q≤2kF
q ) if q>2kF
,(3)
where qs = 4kF rs. The Fermi wave vector kF is taken
from the pristine graphene case corresponding to a given
electron density. The integral equation, Eq. (2), is solved
numerically exactly [27], taking the energy spectrum and
eigenstates of the effective hamiltonian, Eq. (1).
Knowing vectors u for the Fermi contour momenta k,
the conductivity tensor is obtained from,
viujδ(Ek − EF ).
(4)
(cid:90) dk
2π
σij =
e2
h
We plot the calculated longitudinal conductivity of
graphene as a function of carrier density n, with and
without proximity effects, in Fig. 3(a). We use ni = 80×
1010 cm−2 as a representative density of long-range scat-
terers. The carrier density, unlike in pristine graphene,
depends not only on the Fermi level but also on the
strength of the proximity interactions, λI and λex,
n(EF ) = 2 × 1
2π
F + 2EF λI + λ2
ex
(5)
(cid:3) /(vF ).
(cid:2)E2
The factor 2 takes into account the valley degeneracy.
The carrier density is independent of λBR and the di-
rection of the magnetization. In all the plots we fix the
(Color online). Fermi contours and spin texture
FIG. 2.
(left), and the band structure along kx (middle) and ky (right)
for different directions of the exchange field. The direc-
tion of the exchange field is indicated by the large arrows;
small arrows give the spin projections. (a) The out-of-plane
(OOP) exchange field separates two spin subbands, while the
Bychkov-Rashba field leads to a distinctive spin texture de-
pending on the z-projection of the real spin, which interacts
with exchange field.
(b) The in-plane (IP) exchange field
splits the bands, but also deforms the Fermi circles. We
used EF = 100 meV, λI = 0 meV, λBR = 10 meV, and
λex = 25 meV.
ing regions of large transport magnetoanisotropies.
Dirac electrons in graphene in the presence of proxim-
ity exchange and spin-orbit couplings are described by
the minimal Hamiltonian [17],
H = H0 + HI + HBR + Hex.
(1)
pristine
Here,
graphene Hamiltonian is H0 =
vF (τzσxkx + σyky) with pseudospin (sublattice) Pauli
matrices σσσ and τz = ±1 for K and K(cid:48) points. The Fermi
velocity is vF = (3/2)ta0 ≈ 8.6 × 107 cm/s for t = 2.7
eV and the inter-atomic distance of carbons in graphene
a0 = 1.42 A[22]. The proximity intrinsic-like spin-orbit
coupling is given by the Hamiltonian HI = λI τzσzsz,
with parameter λI and (true) spin Pauli matrices s. The
intrinsic coupling opens a gap of 2λI . The Bychkov-
Rashba Hamiltonian, HBR = λBR(τzσxsy − σysx), with
parameter λBR describes the proximity spin-splitting due
to spin-orbit coupling and lack of space inversion sym-
metry. Finally, the spin-dependent hybridization with
the ferromagnet leads to a proximity exchange, Hex =
λexm·s with parameter λex and magnetization orienta-
tion m.
There are two important magnetic configurations to
consider: out-of-plane and in-plane magnetizations, de-
picted in Fig. 2 (see also Fig. 1).
In the out-of-plane
case the spin up and spin down bands are spin split, but
the band structure (and thus Fermi contour) remains
isotropic. On the other hand, in the in-plane case the
band structure is markedly anisotropic, with the Fermi
kxkxkykykxkxkyky3
(Color online).
FIG. 4.
(a) Calculated proximity induced
anisotropic magnetoresistance (PAMR) and (b) planar Hall
resistivity are shown when the exchange field is in-plane (θ =
π/2) for λBR = 20 meV. PAMR quantifies the longitudinal
magnetoresistance as a function of magnetization orientation
φ. The interplay of spin-orbit coupling and exchange field
leads to a net anisotropic resistivity with C2v symmetry while
the off diagonal elements of the resistivity tensor are non-zero.
Other parameters are n = 1012 cm−2, λI = 0 meV, λex =
10 meV. The insets are polar plot representations.
Rashba field, a spin density transverse to the current
appears as a demonstration of the inverse spin-galvanic
effect [28 -- 30] . The shift of the spin subbands due to the
electric field, combined with the spin texture due to the
Bychkov-Rashba spin-orbit interaction, leads to a spin
polarization. This is also expected to happen in graphene
[31, 32], along with the spin-galvanic effect [33, 34]. The
non-equilibrium spin density caused by the electric field
can be calculated as,
δ(cid:104)S(cid:105) =
(cid:90)
dk
(2π)2 δf (k)s(k),
(6)
where s(k) is the spin (represented by Pauli matrices s)
expectation value of the state k. For our proximity model
in the presence of long-range Coulomb scatterers, the cal-
culated inverse spin-galvanic effect is shown in Fig. 3(b)
as a function of exchange coupling. With increasing mag-
netization the induced transverse spin is reduced, as the
exchange coupling aligns the spins and deforms the ro-
tational spin texture of the Bychkov-Rashba field. The
spin densities can be giant. In fields of 1 V/µm, which
FIG. 3.
(a) (Color online). Calculated longitudinal conduc-
tivity as a function of carrier density, for pristine and proxim-
ity graphene. For proximity graphene we show the conductiv-
ity in the presence of Bychkov-Rashba and exchange coupling
only, and in the presence of intrinsic spin-orbit coupling only.
(b) Inverse spin-galvanic effect (scheme in the left inset) in
proximity graphene. Spin density induced (and normalized)
by electric field (which is along x-axis) with respect to the
exchange interaction, when the magnetization is out-of-plane
(OOP) and in-plane (IP). In-plane magnetization can be ei-
ther parallel (along x-ais) or perpendicular (along y-axis) to
the electric field E. In the right inset, the angle dependance of
Sy(φ) is shown in the polar plot with ∆Sy = Sy−Smin
, for the
electric field E = 1 V/cm, and for λex = 10 meV in percent-
y = Sy(EM). The carrier density in
age with respect to Smax
this plot is n = 1012 cm−2 and λI = 0 meV, λBR = 20 meV.
y
carrier density, instead of the Fermi level. The conduc-
tivity for three different combinations of parameters is
shown in Fig. 3(a). The linear dependence on n is well
reproduced. While λBR and λex bring about relatively in-
significant changes ((cid:46) 2.5 %), the presence of λI =10 meV
lowers the conductivity by about ∼10 % at a fixed car-
rier density. Thus, in terms of modifying the magni-
tude of the conductivity, the proximity effects (unless not
inducing additional scattering, which would need to be
investigated case-by-case) are rather weak, being more
pronounced with the inclusion of the intrinsic spin-orbit
coupling, than with the Bychkov-Rashba and exchange
effects. However, as we will see shortly, the anisotropic
effects are quite pronounced.
When current flows in the presence of a Bychkov-
������������������n/niσxx(e2/h)λI=λBR=λex=0λI=0,λBR=λex=10λI=10,λBR=λex=0ni=80×1010cm-2(a) ������������������������������λex(meV)<Sy>(107cm-2)OOPIP(EM)IP(E⟂M)E=1V/cm(b) Sjinv. spin-galvanic Sy/Smaxy(%)10y x �π�π-���-���-���-������ϕPAMR(%)θ=π/���(a) �π�π-�-����ϕρxy(Ω)θ=π/���(b) 4
FIG. 5. (Color online). Parameter maps of PAMR(θ = π/2, φ = π/2). (a) PAMR as a function of λBR and λI , for a fixed
λex = 10 meV. (b) PAMR as a function of λBR and λex, for λI = 0. Note that PAMR changes sign around the λBR = λex line.
In both maps the carrier density is 1012 cm−2.
are still achievable in graphene, the spin density could
reach 1011 cm−2, corresponding to about 10% of spin
polarization. The largest induced spin accumulation is
in the out-of-plane configuration for large exchange. The
magnetoanisotropy of the inverse spin-galvanic effect can
be very large, as seen in Fig. 3(b). The presence of the
current-induced spin accumulation, as well as its mag-
netoanisotropy, could be detected in the same proximity
structure, by measuring the transverse voltage, as in non-
local spin injection [2].
To quantify the transport magnetoanisotropy, we in-
troduce proximity induced anisotropic magnetoresistance
(PAMR), as a ratio of the resistivities R (or conductiv-
ities σ) for a given magnetization orientation (θ, φ) (see
Fig. 1),
PAMR[E] =
=
R(θ, φ) − R(θ, 0)
σxx(θ, 0) − σxx(θ, φ)
R(θ, 0)
σxx(θ, φ)
,
(7)
analogously to the tunneling anisotropic magnetoresis-
tance effect [35 -- 37]. PAMR refers to the changes in the
longitudinal magnetoresistance as the magnetization di-
rection varies with respect to the direction of the external
electric field that drives the current. When the magne-
tization is out-of-plane (θ = 0), the broken time reversal
symmetry and strong spin-orbit coupling can lead to the
novel quantum anomalous Hall effect [10], or crystalline
magnetoanisotropy [21]. However, here we focus on the
regime in which PAMR is most pronounced (θ = π/2).
As shown in Fig. 4(a), PAMR exhibits a C2v symme-
try due to the interplay between the Bychkov-Rashba
and exchange interactions. The expected magnitudes of
PAMR are about 1%, similar to what is observed in ferro-
magnetic metals [21]. The anisotropic resistivity tensor
has non-zero off-diagonal elements due to the presence
of exchange and spin-orbit couplings. This leads to the
planar Hall effect, shown in Fig. 4(b). The magnitude of
the planar Hall effect could reach up to 4 Ω which greater
than the typical values studied in metallic ferromagnetic
systems [38].
What values can PAMR reach for a reasonable range
of proximity parameters? Figure 5 shows two parameter
maps, one with the Bychkov-Rashba and intrinsic, the
other with the Bychkov-Rashba and exchange couplings.
We see two distinct features.
(i) First, in Fig. 5(a) a
horizontal line around λBR ∼ λex ≈ 10 meV separates
two regions. For λBR (cid:46) 10 meV, increasing λI increases
PAMR. For λBR (cid:38) 10 meV, PAMR initially increases
with increasing λI , reaching a maximum of about 1 %
around λI ∼ 7 meV, beyond which PAMR decreases. (ii)
Second, in Fig. 5(b) the line λBR = λex marks a sharp
crossover between weak and strong PAMR. However, this
crossover is not uniform. PAMR is largest for large values
of both λex and λBR slightly greater than λex. The reason
why this region gives the largest PAMR (more than 1%)
is that in this parameter range there is a band crossing
between the strongly spin-orbit coupled subbands.
In conclusion, we used a realistic transport model to
predict magnetotransport anisotropies in graphene with
proximity exchange and spin-orbit couplings. We predict
marked anisotropies in the magnetoresistance, with sim-
ilar values as reached in ferromagnetic metal junctions
and slabs. The calculated PAMR depends strongly on
(a) (b) the spin-orbit coupling and exchange parameters. We
also calculated the magnetoanisotropies of the planar
Hall and inverse spin-galvanic effects. All these magne-
toanisotropies should be a sensitive tool to probe prox-
imity effects in graphene.
We thank Denis Kochan, Jonathan Eroms, and Cosimo
Gorini for useful discussions. This work was supported
by the DFG SFB 689 and the European Union Sev-
enth Framework Programme under Grant Agreement No.
604391 Graphene Flagship.
∗ jeongsu.lee@physik.uni-regensburg.de
[1] M. Gmitra, S. Konschuh, C. Ertler, C. Ambrosch-Draxl,
and J. Fabian, Phys. Rev. B 80, 235431 (2009).
[2] I. Zuti´c, J. Fabian, and S. Das Sarma, Rev. Mod. Phys.
76, 323 (2004).
[3] D. Kochan, M. Gmitra, and J. Fabian, Phys. Rev. Lett.
112, 116602 (2014).
[4] J. Bundesmann, D. Kochan, F. Tkatschenko, J. Fabian,
and K. Richter, Phys. Rev. B 92, 081403 (2015).
[5] D. Van Tuan and S. Roche, Phys. Rev. Lett. 116, 106601
(2016).
[6] A. Ferreira, T. G. Rappoport, M. A. Cazalilla,
and
A. H. Castro Neto, Phys. Rev. Lett. 112, 066601 (2014).
[7] J. Wilhelm, M. Walz, and F. Evers, Phys. Rev. B 92,
014405 (2015).
[8] M. R. Thomsen, M. M. Ervasti, A. Harju, and T. G.
Pedersen, Phys. Rev. B 92, 195408 (2015).
[9] H. X. Yang, A. Hallal, D. Terrade, X. Waintal, S. Roche,
and M. Chshiev, Phys. Rev. Lett. 110, 046603 (2013).
[10] Z. Qiao, W. Ren, H. Chen, L. Bellaiche, Z. Zhang, A. H.
MacDonald, and Q. Niu, Phys. Rev. Lett. 112, 116404
(2014).
[11] P. Lazi´c, K. D. Belashchenko, and I. Zuti´c, Phys. Rev.
B 93, 241401 (2016).
[12] C. B. Crook, C. Constantin, T. Ahmed, J.-X. Zhu, A. V.
Balatsky, and J. T. Haraldsen, Sci. Rep. 5, 12322 (2015).
and J. Fabian,
[13] K. Zollner, M. Gmitra, T. Frank,
arXiv:1607.08008 (2016).
[14] Z. Wang, C. Tang, R. Sachs, Y. Barlas, and J. Shi, Phys.
Rev. Lett. 114, 016603 (2015).
[15] J. C. Leutenantsmeyer, A. A. Kaverzin, M. Wojtaszek,
and B. J. van Wees, arXiv:1601.00995 (2016).
[16] P. Wei, S. Lee, F. Lemaitre, L. Pinel, D. Cutaia, W. Cha,
F. Katmis, Y. Zhu, D. Heiman, J. Hone, J. S. Moodera,
5
and C.-T. Chen, Nat. Mater. advance online publication,
DOI: 10.1038 (2016).
[17] W. Han, R. K. Kawakami, M. Gmitra, and J. Fabian,
Nat. Nanotechnol. 9, 794 (2014).
[18] M. Trushin and J. Schliemann, Phys. Rev. B 75, 155323
(2007).
[19] O. Chalaev and D. Loss, Phys. Rev. B 77, 115352 (2008).
[20] A. Avsar, J. Y. Tan, T. Taychatanapat, J. Balakrishnan,
G. K. W. Koon, Y. Yeo, J. Lahiri, A. Carvalho, A. S.
Rodin, E. C. T. O'Farrell, G. Eda, A. H. Castro Neto,
and B. Ozyilmaz, Nat. Commun. 5, 4875 (2014).
[21] T. Hupfauer, A. Matos-Abiague, M. Gmitra, F. Schiller,
and
J. Loher, D. Bougeard, C. H. Back, J. Fabian,
D. Weiss, Nat. Commun. 6, 7374 (2015).
[22] N. M. R. Peres, Rev. Mod. Phys. 82, 2673 (2010).
[23] S. Adam, E. H. Hwang, V. M. Galitski,
and
S. Das Sarma, Proc. Natl. Acad. Sci. U.S.A. 104, 18392
(2007).
[24] E. H. Hwang and S. Das Sarma, Phys. Rev. B 75, 205418
(2007).
[25] E. H. Hwang, S. Adam, and S. Das Sarma, Phys. Rev.
Lett. 98, 186806 (2007).
[26] E. H. Hwang and S. Das Sarma, Phys. Rev. B 77, 195412
(2008).
[27] We discretize the Fermi contour (100 to 500 points usu-
ally suffice) and solve the resulting sets of linear equations
algebraically.
[28] S. D. Ganichev, E. L. Ivchenko, V. V. Bel'kov, S. A.
Tarasenko, M. Sollinger, D. Weiss, W. Wegscheider, and
W. Prettl, Nature (London) 417, 153 (2002).
[29] S. Ganichev and W. Prettl, Intense terahertz excitation
of semiconductors (Oxford University, New York, 2006).
[30] S. D. Ganichev and L. E. Golub, Phys. Status Solidi B
251, 1801 (2014).
[31] A. Dyrda(cid:32)l, M. Inglot, V. K. Dugaev, and J. Barna´s,
Phys. Rev. B 87, 245309 (2013).
[32] A. Dyrda(cid:32)l, J. Barna´s, and V. K. Dugaev, Phys. Rev. B
89, 075422 (2014).
[33] Ka Shen, G. Vignale, and R. Raimondi, Phys. Rev. Lett.
112, 096601 (2014).
[34] J. S´anchez, L. Vila, G. Desfonds, S. Gambarelli, J. P. At-
tan´e, J. M. De Teresa, C. Mag´en, and A. Fert, Nat.
Commun. 4, 2944 (2013).
[35] A. Matos-Abiague, M. Gmitra,
and J. Fabian, Phys.
Rev. B 80, 045312 (2009).
[36] A. Matos-Abiague and J. Fabian, Phys. Rev. B 79,
155303 (2009).
[37] L. Brey, C. Tejedor, and J. Fern´andez-Rossier, Appl.
Phys. Lett. 85, 1996 (2004).
[38] H. X. Tang, R. K. Kawakami, D. D. Awschalom, and
M. L. Roukes, Phys. Rev. Lett. 90, 107201 (2003).
|
1311.5678 | 1 | 1311 | 2013-11-22T08:53:53 | Electronic and Transport Properties of Molecular Junctions under a Finite Bias: A Dual Mean Field Approach | [
"cond-mat.mes-hall"
] | We show that when a molecular junction is under an external bias, its properties can not be uniquely determined by the total electron density in the same manner as the density functional theory (DFT) for ground state (GS) properties. In order to correctly incorporate bias-induced nonequilibrium effects, we present a dual mean field (DMF) approach. The key idea is that the total electron density together with the density of current-carrying electrons are sufficient to determine the properties of the system. Two mean fields, one for current-carrying electrons and the other one for equilibrium electrons can then be derived.By generalizing the Thomas-Fermi-Dirac (TFD) model to non-equilibrium cases, we analytically derived the DMF exchange energy density functional. We implemented the DMF approach into the computational package SIESTA to study non-equilibrium electron transport through molecular junctions. Calculations for a graphene nanoribbon (GNR) junction show that compared with the commonly used \textit{ab initio} transport theory, the DMF approach could significantly reduce the electric current at low biases due to the non-equilibrium corrections to the mean field potential in the scattering region. | cond-mat.mes-hall | cond-mat |
Electronic and Transport Properties of Molecular Junctions under a Finite Bias: A
Dual Mean Field Approach
Shuanglong Liu1, Yuan Ping Feng1, and Chun Zhang1,2∗
1Department of Physics and Graphene research centre,
National University of Singapore, 2 Science Drive 3, Singapore 117542
2Department of Chemistry, National University of Singapore, 3 Science Drive 3, Singapore 117543
We show that when a molecular junction is under an external bias, its properties can not be
uniquely determined by the total electron density in the same manner as the density functional
theory (DFT) for ground state (GS) properties.
In order to correctly incorporate bias-induced
nonequilibrium effects, we present a dual mean field (DMF) approach. The key idea is that the total
electron density together with the density of current-carrying electrons are sufficient to determine
the properties of the system. Two mean fields, one for current-carrying electrons and the other
one for equilibrium electrons can then be derived. Calculations for a graphene nanoribbon (GNR)
junction show that compared with the commonly used ab initio transport theory, the DMF approach
could significantly reduce the electric current at low biases due to the non-equilibrium corrections
to the mean field potential in the scattering region.
Since the pioneering work of Aviram and Ratner1,
molecular electronics has attracted a great deal of in-
terests due to its promise for future electronics technol-
ogy. The central topic in theoretical research of molecular
electronics is to understand the quantum electron trans-
port at the molecular level by relating the electric cur-
rent passing through the molecular junction to its intrin-
sic electronic properties. The commonly used ab initio
approach combines the quantum transport theory based
on non-equilibrium Green's function (NEGF) techniques
and the computational method based on density func-
tional theory (DFT).2,3 The approach has been applied to
describe the quantum electron transport through various
types of molecular junctions.4 -- 8, and great success has
been achieved in understanding quantum electron trans-
port and also inspiring novel applications in molecular
electronics.
Despite its great success, there are still two problems
in the current approach: 1) whether or not the DFT is
good enough for molecular junctions under a finite bias is
questionable, and 2) when quantitatively compared with
experiments, in most cases, theoretically calculated elec-
tric current is significantly higher. For some molecular
junctions, there might be orders of magnitude differences
between experiment and theory. 4,5,9 In this paper, we
examine in details the bias-induced noneqiulibrium ef-
fects in molecular junctions that are neglected in the
current approach and propose a new ab initio method
to incorporate these nonequilibrium effects.
Molecular junctions connected with two reservoirs can
be modeled as open systems with the general structure
shown in Fig. 1. The system is divided into three re-
gions: left, right reservoirs, and the molecular-scale scat-
tering region in the middle. Electrons in two reservoirs
are assumed to be in their local equilibrium and can be
described by single-electron or mean-field Hamiltonians
∗Electronic address: phyzc@nus.edu.sg
Scattering
Region
Left
Reservoir
Right
Reservoir
Molecular Center
FIG. 1: A molecular scale junction consists of an interact-
ing scattering region and two non-interacting reservoirs. At
t = −∞, an infinitely high single-electron barrier potential
(the gray line) is applied at the left contact so that no elec-
tric current is flowing through. The barrier potential slowly
decreases to zero from t = −∞ to t = 0. The system reaches
the desired non-equilibrium steady state at t = 0.
(HL and HR for left and right reservoirs respectively) so
that two constant chemical potentials µl and µr can be
defined in left and right reservoirs respectively. The bias
voltage across the system can then be defined by the dif-
ference between chemical potentials in two reservoirs as
Vb = (µl − µr)/e. Electrons in the scattering region are
described by an interacting Hamiltonian HS. The total
Hamiltonian for such an open system can be written as
H = HL + HR + HS + HT where HT is the tunneling
term. The goal of the transport theory is to relate the
electronic and transport properties of the steady state
of the system to the Hamiltonian and the voltage bias,
which is extremely difficult (if not impossible) with the
presence of the complicated interaction term HS in the
scattering region.
For simplicity, we set the temperature to be zero
throughout the paper. When µl equals µr, the whole sys-
tem is in equilibrium; the DFT is applicable, and then
the interaction term HS can be replaced by the mean-
field DFT Hamiltonian. Using NEGF techniques, the
mean-field Schrodinger equation together with the open-
system boundary conditions can be solved,10,11 and in
turn the electronic properties can be worked out.
In
practice, due to the problems of existing DFT function-
als and also the fact that the conductance calculations
rely on single-electron orbital, calculations may not be
accurate.12 -- 19 When µl is not the same as µr, the sit-
uation however is different. In this case, the scattering
region is driven out of equilibrium, and whether or not
its electronic structures in principle can be described by
DFT needs to be carefully examined.
Theoretically, the non-equilibrium steady state of the
junction can be achieved by the following adiabatic time-
dependent process. At beginning t = −∞, an infinitely
high single-electron barrier potential is applied at the left
contact as shown in Fig. 1 so that there is no electric cur-
rent flowing through the system. The two parts separated
by the barrier can be denoted as LB (the part on the left
of the barrier) and RB (the part on the right of the bar-
rier). The coupling between LB and RB is then gradually
turned on by slowly decreasing the barrier potential to
zero from t = −∞ to t = 0. The final state at t = 0 is the
non-equilibrium steady state we desire. For such a time-
dependent system, the Runge-Gross (RG) theorem, the
foundation of time-dependent DFT,20 claims that the ex-
ternal potential at time t, vext(r, t), can be uniquely de-
termined by the time-dependent electron density ρ(r, t)
together with the initial state ψ(t = −∞) up to a triv-
ial additive time-dependent function c(t). If initially the
system is in a stationary ground state, the initial state
ψ(t = −∞) itself is a functional of initial electron den-
sity ρ(r, t = −∞) according to the first Hohenberg-Kohn
(HK) theorem21, and then the initial-state dependence
of the external potential in RG theorem can be elimi-
nated. As a result, at t = 0, the external potential and
in turn the Hamiltonian can be uniquely determined by
ρ(r, t = 0), justifying the commonly used ab initio trans-
port theory that combines DFT and NEGF.
Unfortunately, when the external bias Vb is not zero,
initially (t = −∞), the system is in a non-equilibrium
state instead of a stationary ground state. The non-
equilibrium initial state consists of two separated parts,
LB and RB, each of which is in its own equilibrium.
From the first HK theorem, we know that the exter-
nal potential of each part at t = −∞ can be deter-
mined by its electron density up to an arbitrary con-
stant. Combining two parts together, the external po-
tential of the whole system at t = −∞ is determined by
ρ(r, t = −∞) in whole space up to two independent ar-
bitrary constants cl (from LB) and cr (from RB), which
can be expressed as vext(r, t = −∞) ≡ evext[ρ(r, t =
−∞)] + clr∈LB + crr∈RB. Two constants, cl and cr,
shift the Hamiltonian and chemical potentials in LB and
RB, respectively. Only knowing the electron density, cl
and cr are not determined, then the Hamiltonian of the
whole system and also the bias voltage Vb cannot be
determined, resulting in a fact that the initial Hamil-
tonian and in turn the initial state in general can not
be uniquely determined by the electron density alone.
With given evext[ρ(r, t = −∞)], cl and cr simply shift µl
and µr respectively. Without losing generality, the exter-
nal potential can also be written as vext(r, t = −∞) ≡
2
the definition of the bias voltage Vb, we have vext(r, t =
eevext[ρ(r, t = −∞)] + µlr∈LB + µrr∈RB. According to
−∞) ≡ eevext[ρ(r, t = −∞)] + eVbr∈LB + µrr∈(LB+RB).
We see immediately that the initial external potential
and in turn the initial state is determined by the ini-
tial electron density together with the bias voltage Vb
up to a trivial additive constant for the whole system
µr. Consequently, the initial-state dependence at t = 0
in RG theorem can be replaced by the voltage depen-
dence, leading to an important theorem that forms the
basis of the ab initio transport theory: When the system
reaches the steady state (t = 0), its external potential,
Vext(r, t = 0), is uniquely determined by the steady-state
electron density together with the bias voltage Vb up to
a trivial additive constant. As a direct consequence, the
energy of the steady state can be written as a voltage-
dependent density functional E[ρ(r), Vb]. When Vb = 0,
the functional goes to the commonly used DFT one.
Next, we show that the external parameter Vb can be
determined by intrinsic properties of the system. For
this purpose, as shown in Eq. 1, we divide the total
electron density of the steady state ρt into two parts and
name these two parts the equilibrium density ρe and non-
equilibrium density ρn, respectively.
ρt(r) = −iZ µl
ρe(r) = −iZ µr
ρn(r) = −iZ µl
µr
−∞
−∞
G<(r, ǫ)dǫ = ρe + ρn
G<(r, ǫ)dǫ
G<(r, ǫ)dǫ
(1)
Here we assume µr < µl. All physical quantities in
Eq. 1 are defined for the steady state at t = 0. The
term G< is the non-equilibrium lesser Green's function
that includes effects of both non-equilibrium distribution
and many-body interactions.2,3,10,11 The non-equilibrium
density ρn as defined has a physical meaning of the den-
sity of current-carrying electrons. Note that according
to the definition, ρn can also be computed in reservoirs
although reservoirs are assumed to be in equilibrium.
With the assumption of the mean-field reservoirs and
also the lesser Green's function in non-interacting sys-
tems2,3,22, it is straitforward to show that given ρe and
ρn, the total electron density ρt and the voltage bias Vb
are determined. Considering the fact that ρe and ρn can
also be determined by ρt and Vb, we therefore proved
the one-to-one correspondence between these two sets of
variables. Now we have one of the major results of the
paper, the foundation of the proposed ab initio approach:
For molecular junctions under a finite bias, the steady-
state properties of the system are uniquely determined
by the equilibrium and non-equilibrium electron densi-
ties, ρe and ρn, as defined in Eq. 1. The aforementioned
voltage dependent energy functional can then be written
as E[ρe, ρn].
We now try to generalize the existing DFT-based
ab initio transport theory2,3 to include bias-induced
nonequilibrium effects. For this purpose, we assume a
stationary principle for the non-equilibrium steady state:
The variation of the the steady-state energy functional
is zero, δEδρe ,δρn = 0. By taking the variations of the
energy functional with respect to ρe and ρn in the scatter-
ing region, we obtained two effective mean field equations
(Eq. 2),
(cid:18)−
(cid:18)−
1
2
1
2
∂exc
∂ρe (cid:19) φe
∂ρn(cid:19) φn
∂exc
∇2 + vext + VH +
j = λe
j φe
j ,
∇2 + vext + VH +
j = λn
j φn
j ,
(2)
where φe and φn are single-electron orbitals that con-
tribute to ρe and ρn respectively.
In the derivation of
above equations, the generalized local density approxi-
mation, Exc = R exc[ρe, ρn](r)dr, is used where exc is
the exchange-correlation energy density of the uniform
electron gas. The Hartree potential VH in the scattering
region can be obtained by solving the Poisson equation
with correct boundary conditions.10,11 Two coefficients,
λe and λn, are energies of corresponding orbitals. These
two equations have to be solved self-consistently together
with the correct open-system boundary conditions, and
NEGF techniques are powerful in matching the required
boundary conditions. Defining two mean-field exchange-
correlation potentials as V e
, Eq.
2 suggests that the non-equilibrium electrons or current-
carrying electrons experience a different mean-field ef-
fective potential from the equilibrium electrons do. We
therefore name the proposed method the dual-mean-field
(DMF) approach. The DMF equations (Eq. 2) are key
results of the paper, which provide the theoretical basis
for the investigation of electronic properties of molecular
junctions under a finite bias.
xc = ∂exc
∂ρn
xc = ∂exc
∂ρe
and V n
In general, the exchange-correlation energy density exc
can be written as the summation of the exchange part
and the correlation part, exc = ex + ec. The DMF ex-
change energy density can be worked out by general-
izing the Thomas-Fermi-Dirac (TFD) model23 to non-
equilibrium cases by applying an external bias voltage to
the uniform electron gas. By placing the uniform elec-
tron gas between two non-interacting reservoirs with dif-
ferent chemical potentials, the exchange energy density
(Eq. 3) can be analytically derived (The derivation can
be found in supporting information), and then two DMF
x = ∂ex
exchange potentials defined as V e
∂ρn
can be computed,
x = ∂ex
∂ρe
and V n
ex(ρt, η) =
1
4
(1 + η)
4/3(cid:20) −(cid:0)1 − η2(cid:1)2
+η4ln (η) + η4 + η3 −
η2 + η +
(ρt),
(3)
ln (1 + η)
3
2(cid:21)eT F D
x
1
2
where η(r) = ρn(r)/ρt(r) which is called the non-
3
TranSIESTA
DMF
18
16
14
12
10
)
A
μ
(
t
n
e
r
r
u
C
8
6
4
2
0
0
0.1
0.2
0.3
Voltage (Volts)
0.4
0.5
FIG. 2:
I-V curves for a GNR junction calculated from
both DMF approach and the commonly used ab initio
transport theory via the software TranSIESTA. Inset: The
atomic structure of the GNR junction and the iso-surface
of the difference between the exchange energy potentials of
current-carrying electrons calculated from DMF approach and
TranSIESTA,δV = V n
. The exchange poten-
tials were calculated under 0.2 V. The iso-surface value is 15
meV .
x − V T ranSIEST A
x
1+η(cid:17)1/3
equilibrium index in this paper, and η = (cid:16) 1−η
. Ac-
cording to the definition, η takes the value between 0
and 1, which measures the extent of the nonequilibrium.
When η = 0, the exchange density reduces to the equilib-
rium TFD one eT F D
. The DMF
correlation energy density is however challenging. In this
paper, we set the correlation functional to be the same
as the DFT one, which may not be a bad approximation
for weakly correlated systems under low biases.
4π3 (cid:0)3π2ρt(cid:1)4/3
(ρt) = − 1
x
We now are ready to apply the DMF approach to real
molecular scale junctions. As a case study, we choose
a junction made of two zigzag graphene nanoribbons
(Z-GNRs) with different widths as shown in the inset
of Fig. 2. The GNRs have been regarded as one of
the most promising building blocks for graphene-based
electronic devices.24 Since the long-range magnetic order
may not be stable under a finite temperature for such a
one-dimensional system, we follow previous studies to set
the total spin of the system zero.25,26 For Z-GNR based
junctions under low biases, It has been well known that
the current flows through edge states. We have imple-
mented the DMF approach into the SIESTA computa-
tional package.27 For comparison, we performed calcu-
lations with both the DMF approach and the commonly
used DFT based transport method via the function Tran-
SIESTA built in SIESTA11 (More computational details
can be found in supporting information).
In Fig. 2, I-V curves from DMF and TranSIESTA
calculations for the GNR junction are presented. When
bias is small (< 0.1 V), the DMF approach essentially
0.1
η=0
FIG. 3: Non-equilibrium index η(r) in the scattering region
of the GNR junction. The color map was plotted in the plane
2.2 A above the GNR. The system goes from local equilibrium
to non-equilibrium when the color changes from red to blue.
reproduces the TranSIESTA results. Starting from 0.1
V, significant deviations between two approaches occur,
and the current from DMF calculation is always lower
than that calculated from TranSIESTA. To understand
this, we plot in the inset of Fig. 2 the iso-surface of the
difference between the DMF non-equilibrium exchange
potential (V n
x ) and the DFT exchange potential calcu-
lated from TranSIESTA. The potentials were calculated
for the bias voltage 0.2 V. We see that the exchange po-
tential increases significantly at edges which are places
the electric current flows through. For other parts of the
system, the non-equilibrium correction to the potential is
not that important. The increase of the exchange poten-
tial leads to a higher scattering barrier in the scattering
region, and in turn, decreases the current as we see in
Fig. 2.
The non-equilibrium index in the DMF approach,
η(r), provides detailed spacial information for the non-
equilibrium steady state of the system. To demonstrate
this, in Fig. 3, we plot the color contour map of η(r) in
4
the scattering region of the GNR junction under 0.2 V.
From the figure, we can see that in the scattering region,
the extent of the non-equilibrium at different places are
quite different: Edges are far away from equilibrium while
electrons in the center of ribbons are still approximately
in local equilibrium.
In conclusion, we propose a DMF approach to de-
scribe electronic and transport properties of molecular-
scale junctions under a finite bias. We show that two elec-
tron densities, ρe and ρn, are needed to uniquely deter-
mine the properties of the steady state of non-equilibrium
molecular junctions. Subsequently, two mean fields, one
for current-carrying electrons and the other one for equi-
librium electrons, can be derived. The transport prop-
erties can then be calculated from the mean-field poten-
tial that the current-carrying electrons experience. The
case study for a GNR junction shows that the DMF ap-
proach could lead to significantly lower electric current
than the commonly used transport theory. For molec-
ular junctions that have localized molecular orbitals in
the scattering region, the non-equilibrium corrections to
the mean-field potential in the DMF approach will cause
significant variations of these localized orbitals and lead
to more profound changes in transport properties, which
will be discussed in our future studies.
We acknowledge the support from Ministry of Educa-
tion (Singapore) and NUS academic research grants (R-
144-000-325-112 and R144-000-298-112). Computations
were performed at the Graphene Research Centre and
Centre for Computational Science and Engineering at
NUS. See Supplementary Material Document No.xxxxx
for the generalized TFD model and computational de-
tails. For information on Supplementary Material, see
http://www.aip.org/pubservs/epaps.html. The compu-
tational codes for all calculations are available online or
from the author.
1 A. Aviram, and M. A. Ratner, Chem. Phys. Lett. 29, 277
11 M. Brandbyge, J. Mozos, P. Ordejon, J. Taylor, and K.
(1974)
Stokbro, Phys. Rev. B 65, 165401 (2002)
2 J. Taylor, H. Guo, and J. Wang, Phys. Rev. B 63, R121104
12 S. Y. Quek, L. Venkataraman, H. J. Choi, S. Louie, M. S.
(2001)
Hybertsen, and J. B. Neaton, Nano Lett. 7, 3477 (2007)
3 Y. Xue, S. Datta, and M. A. Ratner, Chem. Phys. 281,
13 S. H. Ke, H. U. Baranger, and W. Yang, J. Chem. Phys.
151 (2002)
126, 201102 (2007)
4 S. M. Lindsay, and M. A. Ratner, Adv. Mater. 19, 23
14 C. Toher, A. Filippetti, S. Sanvito, and K. Burke, Phys.
(2007)
Rev. Lett. 95, 146402 (2005)
5 J. Tomfohr, and O. F. Sankey, J. Chem. Phys. 120, 1542
15 P. Delaney, and J. C. Greer, Phys. Rev. Lett. 93, 036805
(2004)
(2004)
6 C. Zhang, M. H. Du, H. Cheng, X. Zhang, A. Roitberg,
16 M. Koentopp, K. Burke, and F. Evers, Phys. Rev. B 73,
and J. L. Krause, Phys. Rev. Lett. 92, 158301 (2004)
121403 (2006)
7 K. Stokbro, J. Taylor, and M. Brandbyge, J. Am. Chem.
17 K. S. Thygesen, and A. Rubio, Phys. Rev. B 77, 115333
Soc. 125, 3674 (2003)
8 S. Barraza-Lopez, M. Kindermann, and M. Y. Chou, Nano
Lett. 12, 3424 (2012)
(2008)
18 K. S. Thygesen, Phys. Rev. Lett. 100, 166804 (2008)
19 J. B. Neaton, M. S. Hybertsen, and S. G. Louie, Phys.
9 L. Shen, M. G. Zeng, S. W. Yang, C. Zhang, X. F. Wang,
Rev. Lett. 97, 216405 (2006)
and Y. P. Feng, J. Am. Chem. Soc. 132, 11481 (2010)
10 J. Taylor, H. Guo, and J. Wang, Phys. Rev. B 63, 245407
(2001)
20 E. Runge, and E. K. U. Gross, Phys. Rev. Lett. 52, 997
(1984)
21 P. Hohenberg, and W. Kohn, Phys. Rev. B 136, B864
(1964)
22 C. Zhang, X. Zhang, P. S. Krsti´c, H. P. Cheng, W. H. But-
ler, and J. M. Maclaren, Phys. Rev. B 69, 134406 (2004)
23 N. H. March, Adv. in Phys. 6, 1 (1957)
24 A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S.
Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109
(2009)
25 Q. Yan, B. Huang, J. Yu, F. Zheng, J. Zang, J. Wu, B.
5
Gu, F. Liu, and W. Duan, Nano Lett. 7, 1469 (2007)
26 Z. Li, H. Qian, J. Wu, B. Gu, and W. Duan, Phys. Rev.
Lett. 100, 206802 (2008)
27 J. Solar, E. Artacho, J. Gale, A. Garc´ıa, J. Junquera, P.
Ordej´on, and D. S´anchez-Portal, J. Phys.: Condens. Mat-
ter 14, 2745 (2002)
|
1907.00601 | 2 | 1907 | 2019-07-02T20:56:51 | Electrically driven spin torque and dynamical Dzyaloshinskii-Moriya interaction in magnetic bilayer systems | [
"cond-mat.mes-hall"
] | Efficient control of magnetism with electric means is a central issue of current spintronics research, which opens an opportunity to design integrated spintronic devices. However, recent well-studied methods are mostly based on electric-current injection, and they are inevitably accompanied by considerable energy losses through Joule heating. Here we theoretically propose a way to exert spin torques into magnetic bilayer systems by application of electric voltages through taking advantage of the Rashba spin-orbit interaction. The torques resemble the well-known electric-current-induced torques, providing similar controllability of magnetism but without Joule-heating energy losses. The torques also turn out to work as an interfacial Dzyaloshinskii-Moriya interaction which enables us to activate and create noncollinear magnetism like skyrmions by electric-voltage application. Our proposal offers an efficient technique to manipulate magnetizations in spintronics devices without Joule-heating energy losses. | cond-mat.mes-hall | cond-mat | Electrically driven spin torque and dynamical
Dzyaloshinskii-Moriya interaction in magnetic bilayer systems
Akihito Takeuchi,1 Shigeyasu Mizushima,1 and Masahito Mochizuki2, 3
1Department of Physics and Mathematics,
Aoyama Gakuin University, Sagamihara, Kanagawa 229-8558, Japan
2Department of Applied Physics, Waseda University,
Okubo, Shinjuku-ku, Tokyo 169-8555, Japan
3PRESTO, Japan Science and Technology Agency,
Kawaguchi, Saitama 332-0012, Japan
Abstract
Efficient control of magnetism with electric means is a central issue of current spintronics re-
search, which opens an opportunity to design integrated spintronic devices. However, recent well-
studied methods are mostly based on electric-current injection, and they are inevitably accompa-
nied by considerable energy losses through Joule heating. Here we theoretically propose a way to
exert spin torques into magnetic bilayer systems by application of electric voltages through taking
advantage of the Rashba spin-orbit interaction. The torques resemble the well-known electric-
current-induced torques, providing similar controllability of magnetism but without Joule-heating
energy losses. The torques also turn out to work as an interfacial Dzyaloshinskii-Moriya interaction
which enables us to activate and create noncollinear magnetism like skyrmions by electric-voltage
application. Our proposal offers an efficient technique to manipulate magnetizations in spintronics
devices without Joule-heating energy losses.
PACS numbers:
9
1
0
2
l
u
J
2
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
2
v
1
0
6
0
0
.
7
0
9
1
:
v
i
X
r
a
1
Introduction
Recent research in spintronics [1 -- 8] is seeking efficient ways to control magnetism in
materials and realize high-performance magnetic devices beyond conventional techniques
based on classical electromagnetism. To manipulate the magnetization in magnets, we need
to exert torques to drive them. Spin-transfer torque is one such torque and, mediated by the
exchange interaction, originates with the angular momentum transfer from the spins of the
conducting electrons to magnetizations in the magnetic structure [9 -- 11]. Another is the so-
called β-term torque originating from the spin relaxation due to the spin-orbit interactions
and/or magnetic impurity scatterings, which is perpendicular to the spin transfer torque and
thus behaves as a dissipative torque [11 -- 13]. These two torques enable the magnetization
to be driven and switched via the injection of electric currents [14 -- 17].
The Rashba-type spin-orbit interaction (RSOI) has recently attracted interest as a means
to manipulate conduction-electron spins [5, 18]. This interaction is of relativistic origin and
becomes active in systems with broken spatial inversion symmetry such as surfaces and
interfaces [5, 18 -- 21]. Because the RSOI mediates mutual coupling between spins and orbital
momenta of the conduction electrons, it works as an effective magnetic field acting on their
spins. The strength and direction of the effective magnetic field that each electron feels
are governed by the momentum of the electron. Therefore, the RSOI can be a source of
nontrivial torques acting on the magnetization through its control on the spin polarizations
of the conduction electrons [22 -- 27]. The strength of the RSOI can be tuned by applying
a gate voltage normal to the surface or interfacial plane [19], which modulates the extent
of the spatial inversion asymmetry. Thus, an alternate-current (AC) gate voltage produces
a time-varying RSOI, and thereby offers a potential technique to produce electrically an
oscillating magnetic texture via the nontrivial Rashba-mediated torques. This technique
must be useful for manipulation of magnetic skyrmions in a magnetic bilayer system, which
have recently attracted great interest from the viewpoint of possible application to high-
performance memory devices [28, 29].
Recent theoretical studies demonstrated that this time-varying RSOI induces an AC spin
current [30 -- 32]. Ho et al. derived an expression for the torque T from the AC spin cur-
rent [33], finding that this expression does not contain a term corresponding to the β-term
torque. It is known that the spin-transfer torque alone cannot drive magnetic domain walls,
2
but other types of torques such as the β-term torque and the spin-orbit torque [34 -- 36]
are required to drive them. However, their theory was based on a continuity equation for
the conduction electron spins where the contribution of spin relaxation torque is totally ne-
glected. The lack of a β-term torque may be a consequence of this crude approximation [37].
Therefore, we reexamine whether the β-term torque in the time-dependent Rashba electron
systems is present using a more elaborate theoretical method.
In this paper, we derive using the quantum field theory an explicit analytical formula for
the torque arising from the AC RSOI in the magnetic bilayer systems. We demonstrate that
the appropriate incorporation of the spin relaxation effect leads to a formula including both
the spin-transfer torque and the β-term torque given in terms of the spin current proportional
to the time derivative of the RSOI. We find that the derived expression may be regarded
as a time-dependent Dzyaloshinskii-Moriya interaction (DMI) at the interface and may be
exploited to manipulate noncollinear magnetic textures such as domain walls, magnetic
vortices, and skyrmions. For a demonstration, we performed a numerical simulation of
the electrical activation of a two-dimensional skyrmion lattice, where the effective AC DMI
induces a periodic expansion and contraction of the skyrmions. This collective excitation
of the skyrmions resembles the breathing mode [38, 39] observed in a skyrmion crystal
under out-of-plane microwave magnetic fields. Efficient techniques to control and drive
magnetism using electric fields are subject of intensive studies in the recent spintronics
research because electric fields in insulators are not accompanied by Joule-heating losses.
Note that electric currents and resulting Joule-heating losses can be significantly suppressed
by using a ferromagnet/insulator bilayers or a metallic bilayer fabricated on an insulating
substrate. Our finding provides a new efficient technique to control magnetism using AC
electric fields and means to electrically excite eigenmodes of noncollinear spin textures.
Results
Model
We consider a magnetic bilayer system, that is, a two-dimensional electron system on top
of the surface of a magnet (see Fig. 1). This system is fabricated on an insulating substrate
to enhance the effects of the electric gate voltage acting on the ferromagnet/heavy-metal
3
interface through preventing the electric-current flow. The total Hamiltonian of this electron
system has four contributing terms, H = HK + HR + Hex + Himp with
HK =
1
2me Z d2r(cid:12)(cid:12)pψ(r, t)(cid:12)(cid:12)
αR(t)
2
− εFZ d2r ψ†(r, t)ψ(r, t),
HR = −
Z d2r ψ†(r, t)(p × σ)zψ(r, t),
Hex = JexZ d2r m(r) · ψ†(r, t) σψ(r, t),
Himp = Z d2r vimp(r)ψ†(r, t)ψ(r, t),
(1)
(2)
(3)
(4)
where me and p denote the mass and momentum, respectively, of a conduction electron,
εF the Fermi energy, σ the Pauli matrices, and ψ† (ψ) the creation (annihilation) operator
of a conduction electron. The term HK represents the kinetic energies of the conduction
electrons with εF being the Fermi energy, while the term HR describes the time-varying
RSOI with αR(t) being the time-dependent coupling coefficient. The term Hex represents
the exchange interaction between the conduction-electron spins and the local magnetization
with Jex and m being the coupling constant and the normalized local magnetization vector,
respectively. The term Himp represents the scattering potentials from spatially distributed
nonmagnetic impurities and determines the relaxation time of electrons given by τ . More
conretely, we consider the impurity potential given by vimp(r) = uimpPi δ(r − Ri), where
uimp denotes the strength of impurity scattering, Ri positions of the impurities, and δ(r) the
Dirac delta function. When we take an average over the impurity positions as vimp(r) = 0
and vimp(r)vimp(r′) = nimpu2
impδ(r − r′), the relaxation time of electrons is given by τ =
/2πνenimpu2
imp in the first Born approximation. Here, nimp denotes the concentration of
impurities and νe = me/2π2 the density of state. In the present study, we focus on the spin
torques induced by the time-varying RSOI but neglect those induced by the magnetization
dynamics. Note that there should be a feedback effect that the spin states of the conduction
electrons are modulated by the magnetization dynamics, but we neglect this subsequent
effect on the spin torques.
Spin torque arising from time-dependent RSOI
The torque induced by the electron spins via the exchange interaction is defined as
m × s,
T =
Jexa2
4
(5)
(cid:85)(cid:74)(cid:78)(cid:70)(cid:14)(cid:69)(cid:70)(cid:81)(cid:15)(cid:1)(cid:51)(cid:66)(cid:84)(cid:73)(cid:67)(cid:66)(cid:1)(cid:87)(cid:70)(cid:68)(cid:85)(cid:80)(cid:83)
DR(t)
(cid:79)(cid:80)(cid:79)(cid:68)(cid:80)(cid:77)(cid:77)(cid:74)(cid:79)(cid:70)(cid:66)(cid:83)(cid:1)
(cid:78)(cid:66)(cid:72)(cid:79)(cid:70)(cid:85)(cid:74)(cid:91)(cid:66)(cid:85)(cid:74)(cid:80)(cid:79)(cid:84)
m(r)
(cid:71) (cid:70) (cid:83) (cid:83) (cid:80)(cid:78) (cid:66) (cid:72) (cid:79) (cid:70) (cid:85)
(cid:77) (cid:66) (cid:90) (cid:70) (cid:83)
(cid:1)
(cid:73) (cid:70) (cid:66) (cid:87) (cid:90) (cid:14)(cid:78) (cid:70) (cid:85) (cid:66) (cid:77)
(cid:74) (cid:79) (cid:84) (cid:86) (cid:77) (cid:66) (cid:85) (cid:74) (cid:79) (cid:72) (cid:1) (cid:84) (cid:86) (cid:67) (cid:84) (cid:85) (cid:83) (cid:66) (cid:85) (cid:70)
AC
(cid:34)(cid:36)(cid:1)(cid:72)(cid:66)(cid:85)(cid:70)(cid:1)
(cid:87)(cid:80)(cid:77)(cid:85)(cid:66)(cid:72)(cid:70)
FIG. 1: Schematic illustration of the time-dependent Rashba electron system interacting with local
magnetizations m(r). The Rashba parameter αR(t) is time-modulated by an external AC electric
voltage. The insulating substrate prevents electric-current flows and enhances the effects of the
electric voltage acting on the RSOI-hosting interface.
where a is the lattice constant, s = hψ† σψi is the conduction electron spin density, and
the brackets denote the quantum expectation value. We assume a metallic bilayer system
in which the condition Jex < εF usually holds and an adiabatic case with a slowly varying
magnetization texture of q ≪ kF where q is the wavenumber of local magnetization and kF
is the Fermi wavenumber. We also assume that a weak magnitude of αR (αRkF ≪ εF) and
low frequencies Ω (Ω ≪ εF) for the time-dependent RSOI. In this perturbation regime, we
obtain the analytical formula of the torque in the form
T = T 1+T 2+T 3 = −
a
D1(m×∇)zm+
a
D2(m×∇)zm−
a
βRD2m×h(m×∇)zmi, (6)
where each coefficient is defined in the condition with εF ≫ /τ , Jex ≫ /τ and εF − Jex ≫
/τ as (η = /2τ )
D1(t) =
νea
2πτ (cid:20) εF
εF − Jex(cid:19) − 2(cid:21)αR(t),
ln(cid:18) εF + Jex
Jex
ex − η2)
J 2
ex(J 2
ex + η2)2
(J 2
dαR(t)
dt
,
D2(t) = νeaεFτ
βR =
2Jexη
ex − η2 .
J 2
(7)
(8)
(9)
We note that D1 vanishes in the clean limit with τ → ∞. Note that D1 does not vanish in
the three-dimensional case or in the half-metallic case with Jex > εF [40] even in the clean
limit.
5
In Eq. (6), the first two contributions, T 1 + T 2, describe an effective DMI [41, 42], which
is given in the continuum form as
HDMI =
D1 − D2
a
ǫαβz Z d2r (m × ∇αm)β.
(10)
(Note that m × (a2/)δHDMI/δm leads to T 1 + T 2). The contribution D1 (∝ αR) appears
even in the steady Rashba system [40, 43, 44]. Recent experiments indeed demonstrated
voltage-induced variations of the interfacial DMI [45, 46], which were ascribed to this steady
Rashba contribution. Quite recently, an experimental observation of gigantic variation of the
DMI that reaches 130 % has been reported for the Ta/FeCoB/TaOx multilayer system [46].
In contrast, the contribution D2 (∝ ∂tαR) appears only in a driven Rashba system with a
time-dependent RSOI.
This interfacial DMI may be tuned by an electric gate voltage via the RSOI, and, more
interestingly, an oscillating DMI may be achieved by applying an AC gate voltage. The
Rashba coefficient αR(t) in the driven Rashba system is a sum of steady and time-dependent
components, αR(t) = α0 + αext(t) with αext(t) = αext sin (Ωt). For metallic (semiconducting)
bilayer systems, the strength of this Rashba-induced DMI is roughly estimated to be D1 ∼
0.1 meV (6 × 10−6 meV) and D2 ∼ 5 × 10−3 meV (2 × 10−6 meV). Here we assume typical
parameter values [19 -- 21], i.e., a = 5 A, εF = 4 eV (10 meV), kF = 1 A−1 (0.01 A−1),
Jex/εF = 0.25 (0.5), τ = 10−14 s (10−12 s), α0 = 2 eV·A (0.07 eV·A), αext/α0 = 0.1, and
Ω/2π = 1 GHz. The strength of the Rashba-mediated DMI is relatively strong (weak) in
metal (semiconductor) systems. We also note that the magnitude of D2 being proportional to
∂tαext(t) may be tuned by changing the amplitude and frequency of the AC gate voltage. The
relative strength of D2 or the ratio D2/D1 tends to be small. Note that the ratio D2/D1 is
approximately given by ΩεFτ 2/2π, which takes ∼ 10−4 (10−2) for metallic (semiconducting)
bilayer systems when a typical frequency of Ω=1 GHz is assumed. Namely, the ratio D2/D1
tends to be larger for the semiconducting system, whereas the absolute value of D2 tends
to be larger for the metallic system. We should choose an appropriate system depending on
the target phenomena or the experiments.
In Eq. (6), the last two terms proportional to D2 may be rewritten as
T 2 + T 3 ∝ (js · ∇)m − βRm × (js · ∇)m.
(11)
by adopting a definition of the spin current js ≡ (e/a)D2 z × m. We find that these terms
6
have equivalent forms with the spin-transfer torque and the nonadiabatic torque associ-
ated with the spin current js, respectively. With an AC-dependent αR(t), the spin current
js ∝ ∂tαR(t) z × m gives rise to AC torques.
In the clean limit with /Jexτ ≪ 1, the
coefficient βR is reduced to /Jexτ . Finally, the second term is exactly identical in form to
the conventional current-induced nonadiabatic torque although τ corresponds to a differ-
ent time scales, specifically, the relaxation time of the conduction electrons (spins) in the
present (current-induced) case [11]. Assuming the above-mentioned material parameters for
the metallic bilayer systems, we evaluate the values of js = (e/a)D2 and βR as ∼ 2 A/m and
∼ 0.07, respectively. These values are large enough to induce the magnetization dynamics.
Application of oscillating DMI
The AC torques in the driven Rashba system may be exploited to activate resonances of
the magnetic textures. For a demonstration, we numerically show that the breathing mode
of a skyrmion crystal can be excited by the RSOI-mediated AC torques. We start with the
continuum limit of the spin model for a two-dimensional magnet,
H = Z d2r
J
2
(∇m)2 −
µBµ0
a2 Z d2r m · H.
(12)
The model contains the ferromagnetic exchange interaction and the Zeeman interaction with
an external magnetic field H = H z. We adopt typical material parameters of a = 5 A,
J = 1 meV and µ0H = 34 mT. The dynamics of magnetizations obeys the Landau-Lifshitz-
Gilbert equation
m = −γmm ×(cid:18)−
a2
γm
δH
δm(cid:19) + αGm × m + T 1,
(13)
where γm is the gyromagnetic ratio, and αG (= 0.04) is the Gilbert-damping coefficient. We
incorporate T 1 as the Rashba-mediated torque whereas T 2 and T 3 are neglected because
they are much smaller than T 1 in the metallic bilayer systems.
that T 1 (T 2 and T 3) originates from D1 (D2), and the ratio D2/D1 is 10−12-10−14 as men-
(Equation (6) indicates
tioned above.) The coupling coefficient D1(t) is composed of a steady component D0 and
the time-dependent component Dext(t) as D1(t) = D0 + Dext(t). We solve this equation
numerically using the fourth-order Runge-Kutta method for a system of 140 × 162 nm2,
including N = 280 × 324 magnetizations, applying a periodic boundary condition. Without
the AC electric voltage, the Rashba-mediated torque has a steady component only with
7
D1 = D0(= 0.09 meV). In this situation, the torque stabilizes the skyrmion crystal with
hexagonally packed N´eel-type skyrmions (Fig. 2(a)) at low temperatures by effectively work-
ing as a static DMI, as exemplified by Eq. (10).
Next, we simulated the dynamics of this skyrmion crystal in the presence of AC gate volt-
ages by switching on the Rashba-mediated AC torque with Dext(t) 6= 0. The time evolution
of the numerical simulation is performed for every 0.01 ps. We first calculate the dynami-
cal magnetoelectric susceptibility χ(Ω) to identify the resonance frequency of this skyrmion
crystal. We trace a time profile of the net magnetization M (t) = (1/Na2)R d2r m(r, t)
and ∆M (t) = M (t) − M (0) after applying a short electric-field pulse by switching on the
time-dependent component of the torque Dext(t) = 0.05D0 for an interval of 0 ≤ t ≤ 1. We
obtain its spectrum (Fig. 2(b)) by calculating the Fourier transform of ∆M (t). From this
spectrum, we find that the skyrmion crystal has a resonance at Ω/2π = 1.72 GHz. Note
that the resonance frequency (the time period) of the skyrmion eigenmode is determined by
the spin-wave gap which is proportional to D2
0 (D−2
0 ).
We then apply a microwave electric field to this skyrmion-hosting magnetic bilayer sys-
tem by switching on the AC component of the torque Dext(t) = 0.05D0 sin (Ωt). We ex-
amined both resonance with Ω/2π = 1.72 GHz and off-resonance with Ω/2π = 1 GHz.
For the former, we observe a breathing-type mode where all the skyrmions constituting the
skyrmion crystal uniformly contracted and expanded periodically (Fig. 2(c)). For the latter,
the induced breathing oscillation of the skyrmions is small. We find that the periodically
modulated skyrmion diameter is nearly proportional to the time-dependent DMI coefficient.
The present simulation demonstrated the electrical activation of a skyrmion resonance. The
induced dynamical DMI is also expected to realize the electrical creation of skyrmions [48 --
53]. The electrical writing of skyrmions with the RSOI-mediated DMI in the field polarized
ferromagnetic state and even the helical state should be established in the future study.
Discussion
In summary, we have theoretically derived a precise formula of the spin torque in a
time-varying Rashba electron system driven by the AC gate voltage. The obtained formula
contains not only the spin-transfer torque but also the β-term torque associated with the AC
spin current that is proportional to the time derivative of the RSOI. These AC torques can
8
FIG. 2: Eigenmodes of N´eel-type skyrmion crystal activated by an AC DMI. (a) Skyrmion crys-
tal with hexagonally packed N´eel-type skyrmions. In-plane and out-of-plane components of the
magnetizations are shown by arrows and color map, respectively. (b) Imaginary part of the dy-
namical magnetoelectric susceptibility. (c) Snapshots of the electrically activated breathing motion
at t = 0, π/2Ω, π/Ω and 3π/2Ω for the resonant condition with Ω/2π = 1.72 GHz, and (d) those
for an off-resonant condition with Ω/2π = 1 GHz.
excite resonance of magnetic textures through acting as an interfacial AC DMI. Indeed, we
have numerically demonstrated that the effective AC DMI can activate the breathing mode
of magnetic skyrmions. We have confirmed that not only the crystallized magnetic skyrmions
(skyrmion crystal) but also isolated skyrmions in ferromagnets can be excited resonantly by
application of a microwave field, which offers a better experimental feasibility because a
lot of ferromagnet/heavy-metal bilayer systems turned out to host magnetic skyrmions as
topological defects [28, 54 -- 56]. Recent theoretical studies revealed that activation of the
skyrmion breathing mode under application of a magnetic field inclined from the vertical
direction induces translational motion of the skyrmions [57 -- 59], which provides a means to
9
drive magnetic skyrmions electrically with a low energy consumption. Our finding provides
a promising technique to manipulate noncollinear magnetic textures with a great efficiency
that have potential applications in memory, logic, and microwave devices.
There are several types of devices to realize the proposed effects. The devices must
have two important features, i.e., (1) an interface that hosts the SOI due to the broken
inversion symmetry and (2) insulating nature to prevent the electric-current flow. One
possible type of device is a ferromagnet/insulator bilayer system [45, 46]. On the contrary,
we proposed another type of device with ferromagnet/heavy-metal bilayer fabricated on an
insulating substrate.
In the latter system, we expect much stronger SOI because of the
heavy-metal layer. In this case, the required insulating nature is taken up by the insulating
substrate. The issue which system is appropriate is left for future study. The RSOI is
originally very strong in the latter system, but its electric tunability may be low because
the applied electric field mainly acts on the heavy-metal/insulator layer but not on the
ferromagnet/heavy-metal interface. In addition, due to the short screening length in metal,
the electric field decays quickly and may hardly reach the interface that hosts the RSOI. On
the other hand, the ferromagnet/insulator bilayer system can originally have a weak RSOI
only, but its electric tunability can be large because the applied electric field directly acts
on the ferromagnet/insulator interface that hosts the RSOI.
It should be also mentioned that several types of bilayer systems with interfacial DMI have
been intensively studied recently, where driven spin torques and generations of skyrmion-type
noncollinear magnetic textures using the interfacial DMI have been experimentally demon-
strated not only for ferromagnet/heavy-metal systems but also for ferromagnet/transition-
metal-dichalcogenide systems [60] and ferromagnet/topological-insulator systems [61].
In
the latter two cases, the SOI is much more complicated than the simple Rashba model for
the ferromagnet/heavy-metal system considered in the present study [62], and it is unclear
which of many terms emerged at the interfaces of these systems can be modulated by AC
gate voltage in reality [63]. These problems require further investigations, which are left for
future researches. In experiments, we need to take care of magnetic anisotropies in the fer-
romagnetic layer because the stability and the resonance modes of skyrmions are sensitively
affected by them [64]. Our results will provide a firm basis and a good starting point for
future experimental and theoretical studies.
10
Methods
Diagonalization of exchange interaction
We need to calculate the electron spin density vector s to obtain the spin torque. However,
as the exchange coupling Jex is generally large compared with other energy scales such as
the kinetic energy and the RSOI, we cannot regard it as a perturbation. Instead, we need
to perform a diagonalization of the exchange interaction term [11, 33]. For this purpose, we
employ a 2 × 2 unitary matrix U ≡ n · σ with n = (sin (θ/2) cos φ, sin (θ/2) sin φ, cos (θ/2)).
Here, θ and φ are the polar and azimuthal angles of the local magnetization vector m =
(sin θ cos φ, sin θ sin φ, cos θ). As relation m· U † σ U = σz holds, the exchange interaction term
may be diagonalized using a new operator of the conduction electron Ψ expressed in the local
coordinates rotating along with the spatially modulated magnetization vectors. Hereafter,
this coordinate system is referred to as the rotated spin frame. The new operator Ψ is related
to the original operator ψ defined in the global coordinates via a unitary transformation as
ψ = U Ψ. Finally, the total Hamiltonian is rewritten with the Ψ operators as
H =
1
2me Z d2r(cid:12)(cid:12)(cid:2)p + eAα(r, t)σα(cid:3)Ψ(r, t)(cid:12)(cid:12)
+JexZ d2r Ψ†(r, t)σzΨ(r, t)
2 − εFZ d2r Ψ†(r, t)Ψ(r, t)
meα2
R(t)
2
−
Z d2r Ψ†(r, t)Ψ(r, t) +Z d2r vimp(r)Ψ†(r, t)Ψ(r, t),
where −e (< 0) is the electron charge. The non-Abelian gauge potential A appears as a
by-product of the diagonalization of the exchange interaction. This gauge potential contains
two contributions [33], denoted A = A(ex) + A(so). The first term A(ex) is a gauge potential
originating from the spatial variation of the magnetization structure whereas the second
term A(so) comes from the RSOI. Each gauge potential is defined as
A(ex)α
µ
σα ≡ −
A(so)α
µ
σα ≡ −
i
e
meαR
e
ǫµβz U † σβ U = −
meαR
e
ǫµβzRαβ σα,
U †∇µ U =
(n × ∇µn)ασα,
e
with Rαβ σα ≡ U † σβ U = (2nαnβ − δαβ)σα. Here Rαβ is an element of a 3 × 3 orthogonal
matrix. The symbols δαβ and ǫαβγ denote the Kronecker delta and the Levi-Civita antisym-
metric tensor, respectively. In the rotated spin frame, the Ψ-electron spin density is given by
11
Sα = hΨ†σαΨi. The spin density S in the rotated spin frame is related to the spin density s
in the original frame via R as sα = RαβSβ. Therefore, the spin torque T is rewritten using
S± = Sx ± iSy,
T α =
Jexa2
2
ǫαβγmβ Xσ=±
(Rγx − iσRγy)Sσ.
In the calculation, we consider the impurity potential given by
vimp(r) = uimpXi
δ(r − Ri),
where uimp denotes the strength of impurity scattering, Ri the positions of the impurities,
and δ(r) the Dirac delta function. When we take an average over the impurity positions as
vimp(r) = 0,
vimp(r)vimp(r′) = nimpu2
impδ(r − r′)
the relaxation time of electrons is given by τ = /2πνenimpu2
imp in the first Born approxi-
mation. Here, nimp denotes the concentration of impurities and νe = me/2π2 the density
of state.
Calculation of spin torque arising from AC RSOI
The Ψ-electron spin density S± is written in terms of the path-ordered Green function
S±(r, t) = −iTrhσ± G<(r, t; r, t)i,
where σ± = σx ± iσy and Tr signifies the trace over the spin indices. The lesser
component of the path-ordered Green function [47] is represented by G<(r, t; r′, t′) =
(i/)hΨ†(r′, t′)Ψ(r, t)i. In the present system, the Dyson equation is given as
G(r, t; r′, t′) = g(r − r′, t − t′) +ZC
where C denotes the Keldysh contour and V is defined as
dt′′ Z d2r′′ g(r − r′′, t − t′′) V (r′′, t′′) G(r′′, t′′; r′, t′),
V (r′′, t′′) = −
ie
2me(cid:20) ∂
∂r′′
µ
Aα
µ(r′′, t′′) + Aα
µ(r′′, t′′)
+
e2
2me
Aα
µ(r′′, t′′)Aα
µ(r′′, t′′)I −
12
∂
∂r′′
µ(cid:21)σα
R(t′′)I + vimp(r′′)I.
me
2 α2
Here g denotes the noninteracting Green function given in the Fourier space as gk,ω =
G<
is the identity matrix.
(1/2)Pσ=±(I + σσz)gk,ω,σ where I
bol < represents the relation [RC dt′′ G1(t, t′′) G2(t′′, t′)]< = R ∞
1 (t, t′′) Ga
2 (t′′, t′) +
2(t′′, t′)] [47]. The retarded, advanced, and lesser Green functions (gr, ga and
k,ω,σ) where fω is the Fermi distribu-
k,ω,σ)∗ =
1/(ω − εk + εFσ + iη) where εk = 2k2/2me and εFσ = εF − σJex. The spin density S±
tion function. The retarded (advanced) Green function is defined as gr
g<) are mutually related by g<
k,ω,σ = fω(ga
k,ω,σ − gr
−∞ dt′′ [ Gr
1(t, t′′) G<
The superscript sym-
k,ω,σ = (ga
can be obtained by iteration of this equation. The dominant contributions are given by
the first-order perturbation expansions in A(so) and up to first order in A(ex). After some
algebra, this equation is reduced to (see Supplementary Materials for details)
S± = −
2e
JexahD1 − (1 ∓ iβR)D2i(cid:16)m × A(ex)±(cid:17)z
,
where A(ex)± = A(ex)x ±iA(ex)y. Substituting this result into the definition of the spin torque
and using the relations Pσ=±(Rαx − iσRαy)A(ex)σ
µ
iσRαy)iσA(ex)σ
µ
= −(/e)(m × ∇µm)α and Pσ=±(Rαx −
= (/e)∇µmα, we thus have obtained the result given in Eqs. (6-9).
Note on the Numerical Simulations
Our numerical simulation with the LLG equation corresponds to the micromagnetic sim-
ulation based on the continuum spin model with the exchange stiffness A = 4.0 × 10−14
[Jm−1], the continuum DM parameter B = 1.44 × 10−5 [Jm−2], the magnetic field µ0Hz=34
mT, and the saturation magnetization Ms = 3.5 × 104 [Am−1] for a system size of 140 nm
× 162 nm × 2 nm. This simulation can be performed with commercial or free softwares
such as OOMMF and mumax. In the present study, the continuum spin model is mapped
to the lattice spin model by dividing the continuum space into identical rectangular cells.
More concretely, dividing the continuum space into identical rectangular cells of 0.5 nm×0.5
nm×2 nm, we obtain the lattice spin model with the normalized magnetization vectors mi
where the exchange coupling J=1 meV, the DM parameter D0/J=0.09, and the magnetic
field µBµ0Hz/J=0.004.
[1] Ohno, H. Making nonmagnetic semiconductors ferromagnetic. Science 281, 951 (1998).
13
[2] Wolf, S. A., Awschalom, D. D., Buhrman, R. A., Daughton, J. M., von Moln´ar, S., Roukes,
M. L., Chtchelkanova, A. Y. & Treger, D. M. Spintronics: a spin-based electronics vision for
the future. Science 294, 1488 (2001).
[3] Zuti´c, I., Fabian, J. & Das Sarma, S. Spintronics: fundamentals and applications. Rev. Mod.
Phys. 76, 323 (2004).
[4] Chappert, C., Fert, A. & Van Dau, F. N. The emergence of spin electronics in data storage.
Nat. Mater. 6, 813 (2007).
[5] Manchon, A., Koo, H. C., Nitta, J., Frolov, S. M., & Duine, R. A. New perspectives for
Rashba spin-orbit coupling. Nat. Mat. 14, 871 (2015).
[6] Endoh, T., & Honjo, H. A recent progress of spintronics devices for integrated circuit appli-
cations. J. Low Power Electron. Appl. 8, 44 (2018).
[7] Jungwirth, T., Sinova, J., Manchon, A., Marti, X., Wunderlich, J., & Felser, C. The multiple
directions of antiferromagnetic spintronics. Nat. Phys. 14, 200 (2018).
[8] Baltz, V., Manchon, A., Tsoi, M., Moriyama, T., Ono, T., & Tserkovnyak, Y. Antiferromag-
netic spintronics. Rev. Mod. Phys. 90, 015005 (2018).
[9] Slonczewski, J. C. Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1 (1996).
[10] Berger, L. Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev.
B 54, 9353 (1996).
[11] Tatara, G., Kohno, H. & Shibata, J. Microscopic approach to current-driven domain wall
dynamics. Phys. Rep. 468, 213 (2008).
[12] Zhang, S. & Li, Z. Roles of nonequilibrium conduction electrons on the magnetization dynam-
ics of ferromagnets. Phys. Rev. Lett. 93, 127204 (2004)
[13] Thiaville, A., Nakatani, Y., Miltat, J. & Suzuki, Y. Micromagnetic understanding of current-
driven domain wall motion in patterned nanowires. Eyrophys. Lett. 69, 990 (2005).
[14] Klaui, M., Vaz, C. A. F., Bland, J. A. C., Wernsdofer, W., Faini, G., Cambril, E. & Hey-
derman, L. J. Domain wall motion induced by spin polarized currents in ferromagnetic ring
structures. Appl. Phys. Lett. 83, 105 (2003).
[15] Tsoi, M., Fontana, R. E. & Parkin, S. S. P. Magnetic domain wall motion triggered by an
electric current. Appl. Phys. Lett. 83, 2617 (2003).
[16] Yamaguchi, A., Ono, T., Nasu, S., Miyake, K., Mibu, K. & Shinjo, T. Real-space observation of
14
current-driven domain wall motion in submicron magnetic wires. Phys. Rev. Lett. 92, 077205
(2004).
[17] Yamanouchi, M., Chiba, D., Matsukura, F. & Ohno, H. Current-induced domain-wall switch-
ing in a ferromagnetic semiconductor structure. Nature 428, 539 (2004).
[18] Rashba, E. I. Properties of semiconductors with an extremum loop. 1. Cyclotron and combi-
national resonance in a magnetic field perpendicular to the plane of the loop. Sov. Phys. Solid
State 2, 1109 (1960).
[19] Nitta, J., Akazaki, T., Takayanagi, H. & Enoki, T. Gate control of spin-orbit interaction in
an inverted In0.53Ga0.47As/In0.52Al0.48As heterostructure. Phys. Rev. Lett. 78, 1335 (1997).
[20] Ast, C. R., Henk, J., Ernst, A., Moreschini, L., Falub, M. C., Pacil'e, D., Bruno, P. Kern, K.
& Grioni, M. Giant spin splitting through surface alloying. Phys. Rev. Lett. 98, 186807 (2007)
[21] Nakagawa, T., Ohgami, O., Saito, Y., Okuyama, H., Nishijima, M. & Aruga, T. Transition be-
tween tetramer and monomer phases driven by vacancy configuration entropy on Bi/Ag(001).
Phys. Rev. B 75, 155409 (2007).
[22] Obata, K. & Tatara, G. Current-induced domain wall motion in Rashba spin-orbit system.
Phys. Rev. B 77, 214429 (2008).
[23] Manchon, A. & Zhang, S. Theory of nonequilibrium intrinsic spin torque in a single nanomag-
net. Phys. Rev. B 78, 212405 (2008).
[24] Manchon, A. & Zhang, S. Theory of spin torque due to spin-orbit coupling. Phys. Rev. B 79,
094422 (2009).
[25] Kim, K.-W., Seo, S.-M., Ryu, J., Lee, K.-J. & Lee, H.-W. Magnetization dynamics induced
by in-plane currents in ultrathin magnetic nanostructures with Rashba spin-orbit coupling.
Phys. Rev. B 85, 180404(R) (2012).
[26] Miron, I. M., Gaudin, G., Auffret, S., Rodmacq, B., Schuhl, A., Pizzini, S., Vogel, J. &
Gambardella, P. Current-driven spin torque induced by the Rashba effect in a ferromagnetic
metal layer. Nat. Mater. 9, 230 (2010).
[27] Miron, I. M., Moore, T., Szambolics, H., Buda-Prejbeanu, L. D., Auffret, S., Rodmacq, B.,
Pizzini, A., Vogal, J., Bonfim, M., Schuhl, A. & Gaudin, G. Fast current-induced domain-wall
motion controlled by the Rashba effect. Nat. Mater. 10, 419 (2011).
[28] Fert, A., Cros, V. & Sampaio, J. Skyrmions on the track. Nat. Nanotech. 8, 152 (2013).
[29] Soumyanarayanan, A., Reyren, N., Fert, A. & Panagopoulos, C. Emergent phenomena induced
15
by spin-orbit coupling at surfaces and interfaces. Nature 539, 509 (2016).
[30] Mal'shukov, A. G., Tang, C. S., Chu, C. S. & Chao, K. A. Spin-current generation and
detection in the presence of an ac gate. Phys. Rev. B 68, 233307 (2003).
[31] Zhang, S.-f. & Zhu, W. The limit spin current in a time-dependent Rashba spin-orbit coupling
system. J. Phys.: Condens. Matter 25, 075302 (2013).
[32] Ho, C. S., Jalil, M. B. A. & Tan, S. G. Spin force and the generation of sustained spin current
in time-dependent Rashba and Dresselhaus systems. J. Appl. Phys. 115, 183705 (2014).
[33] Ho, C. S., Jalil, M. B. A. & Tan, S. G. Gate-control of spin-motive force and spin-torque in
Rashba SOC system. New J. Phys. 17, 123005 (2015).
[34] Ando, K., Takahashi, S., Harii, K., Sasage, K., Ieda, J., Maekawa, S., & Saitoh, E. Electric
manipulation of spin relaxation using the spin Hall effect. Phys. Rev. Lett. 101, 036601 (2008).
[35] Emori, S., Bauer, U., Ahn, S.-M., Martinez, E., & Beach, G. S. D. Current-driven dynamics
of chiral ferromagnetic domain walls. Nat. Mat. 12, 611 (2013).
[36] Ryu, K.-S., Thomas, L., Yang, S.-H., & Parkin, S. Chiral spin torque at magnetic domain
walls. Nat. Nanotech. 8, 527 (2013).
[37] Tatara, G. & Entel, P. Calculation of current-induced torque from spin continuity equation.
Phys. Rev. B 78, 064429 (2008).
[38] Mochizuki, M. Spin-wave modes and their intense excitation effects in skyrmion crystals. Phys.
Rev. Lett. 108, 017601 (2012).
[39] Onose, Y., Okamura, Y., Seki, S., Ishiwata, S. & Tokura, Y. Observation of magnetic excita-
tions of skyrmion crystal in a helimagnetic insulator Cu2OSeO3. Phys. Rev. Lett. 109, 037603
(2012).
[40] Ado, I. A., Qaiumzadeh, A., Duine, R. A., Brataas, A. & Titov, M. Asymmetric and symmetric
exchange in a generalized 2D Rashba ferromagnet. Phys. Rev. Lett. 121, 086802 (2018).
[41] Dzyaloshinskii, I. E. Thermodynamics theory of "weak" ferromagnetism in antiferromagnetic
substances. Sov. Phys. JETP 5, 1259 (1957).
[42] Moriya, T. Anisotropic superexchange interaction and weak ferromegnetism. Phys. Rev. 120,
91 (1960).
[43] Kim, K.-W., Lee, H.-W., Lee, K.-J. & Stiles, M. D. Chirality from interfacial spin-orbit
coupling effects in magnetic bilayers. Phys. Rev. Lett. 111, 216601 (2013).
[44] Kikuchi, T., Koretsune, T., Arita, R. & Tatara, G. Dzyaloshinskii-Moriya interaction as a
16
consequence of a Doppler shift due to spin-orbit induced intrinsic spin current. Phys. Rev.
Lett. 116, 247201 (2016).
[45] Nawaoka, K., Miwa, S., Shiota, Y., Mizuochi, N. & Suzuki, Y. Voltage induction of interfacial
Dzyaloshinskii-Moriya interaction in Au/Fe/MgO artificial multilayer. Appl. Phys. Express 8,
063004 (2015).
[46] Srivastava, T. et al. Large-voltage tuning of Dzyaloshinskii-Moriya interactinos: a route toward
dynamic control of skyrmion chirality. Nano Lett. 18, 4871 (2018).
[47] Haug, H. & Jauho, A. P. Quantum Kinetics in Transport and Optics of Semiconductors
(Springer, 2007).
[48] Mochizuki, M., & Watanabe, Y. Writing a skyrmion on multiferroic materials. Appl. Phys.
Lett. 107, 082409 (2015).
[49] Mochizuki, M., Creation of skyrmions by electric field on chiral lattice magnetic insulators.
Adv. Electron. Mater. 2, 1500180 (2016).
[50] Schott, M. et al. The skyrmion switch: Turning magnetic skyrmion bubbles on and off with
an electric field. Nano Lett. 17, 3006 (2017).
[51] Huang, P. et al. In situ electric field skyrmion creation in magnetoelectric Cu2OSeO3. Nano
Lett. 18, 5167 (2018).
[52] Kruchkov, A. J. et al. Direct E field control of the skyrmion phase in a magnetoelectric
insulator. Sci. Rep. 8, 10466 (2018).
[53] Wang, L. et al. Ferroelectrically tunable magnetic skyrmions in ultrathin oxide heterostruc-
tures. Nat. Mat. 17, 1087 (2018).
[54] Woo, S. et al. Observation of room-temperature magnetic skyrmions and their current-driven
dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501 (2016).
[55] Yu, G. et al. Room-Temperature Skyrmion Shift Device for Memory Application. Nano Lett.
17, 261 (2017).
[56] Jiang, W. et al. Blowing magnetic skyrmion bubbles. Science 349, 283 (2015).
[57] Wang, W., Beg, M., Zhang, B., Kuch, W. & Fangohr, H. Driving magnetic skyrmions with
microwave fields. Phys. Rev. B 92, 020403(R) (2015).
[58] Takeuchi, A. & Mochizuki, M. Selective activation of an isolated magnetic skyrmion in a
ferromagnet with microwave electric fields. Appl. Phys. Lett. 113, 072404 (2018).
[59] Ikka, M., Takeuchi, A. & Mochizuki, M. Resonance modes and microwave-driven translational
17
motion of a skyrmion crystal under an inclined magnetic field. Phys. Rev. B 98, 184428 (2018).
[60] Liu, R. H., Lim, W. L. & Urazhdin, S. Control of current-induced spin-orbit effects in a
ferromagnetic heterostructure by electric field. Phys. Rev. B 89, 220409(R) (2014).
[61] Lv, W. et al. Electric-Field Control of Spin-Orbit Torques in WS2/Permalloy Bilayers. ACS
Appl. Mater. Interfaces 10, 2843 (2018).
[62] Belashchenko, K. D., Kovalev, A. A. & van Schilfgaarde, M. First-principles calculation of the
spin-orbit torques in a Co/Pt bilayer. arXiv:1810.11003.
[63] Gmitra, M., Kochan, D., Hogl, P. & Fabian, J. Trivial and inverted Dirac bands and the
emergence of quantum spin Hall states in graphene on transition-metal dichalcogenides. Phys.
Rev. B 93, 155104 (2016).
[64] Nakamura, M. et al. Emergence of Topological Hall Effect in Half-Metallic Manganite Thin
Films by Tuning Perpendicular Magnetic Anisotropy. J. Phys. Soc. Jpn. 87, 074704 (2018).
Acknowledgements
This work was supported by JSPS KAKENHI (Grant No. 17H02924 and No. 16H06345),
Waseda University Grant for Special Research Projects (Project Nos. 2017S-101, 2018K-
257), and JST PRESTO (Grant No. JPMJPR132A).
18
Supplementary Material for "Electrically driven
spin torque and dynamical Dzyaloshinskii-Moriya
interaction in magnetic bilayer systems"
Calculation of Ψ-electron spin density S±
We provide details of the derivation of the Ψ-electron spin density S±.
In Fig. 3, we
present the Feynman diagrams associated with the Ψ-electron spin density induced by the
time-dependent RSOI. Here the vertex corrections due to the nonmagnetic impurity scat-
terings are not considered because they are negligible in the present case.
A(so)
S
A(ex)
FIG. 3:
Diagrammatic representations of the dominant contributions of the Ψ-electron spin
density S. The solid, wavy, and dotted lines represent the Green function, the RSOI A(so), and
the exchange interaction A(ex) in the rotated spin frame, respectively.
Expanding the lesser component with respect to q and Ω, the spin density S±(q, Ω) is
19
written
S±(q, Ω) =
(q′, Ω)A(ex)z
ν
(q − q′) + A(so)z
µ
(q′, Ω)A(ex)±
ν
(q − q′)i(cid:27)
2e
µ
Xq′ hA(so)±
k,ω,σ)2 − gr
k,ω,−σ(ga
µ
ie4
m2
e (cid:26) ± qνA(so)±
σXk Xω
×Xσ=±
Ω(cid:20) ± qνA(so)±
ie4
2m2
e
+
µ
(q, Ω) −
fωkµkνhga
k,ω,−σ(gr
k,ω,σ)2i
A(so)±
µ
(q′, Ω)A(ex)z
ν
(q − q′)(cid:21)
k,ω,σ(ga
k,ω,−σ)2 − (gr
k,ω,σ)2ga
k,ω,−σ
(q, Ω) −
2e
Xq′
dω(cid:20)gr
kµkν(cid:26) dfω
k,ω,−σ)ga
k,ω,−σ +
1
2
gr
k,ω,σ(gr
k,ω,σ − gr
k,ω,−σ)gr
k,ω,−σ(cid:21)
×Xσ=±
(σ ∓ 1)Xk Xω
k,ω,σ − ga
ga
k,ω,σ(ga
+
1
2
−fωhga
k,ω,−σgr
−gr
µ
ie23
m2
e
+
A(so)z
×Xσ=±
ΩXq′
(σ ∓ 1)Xk Xω
k,ω,σ(ga
k,ω,σ − ga
ga
−fωh(ga
k,ω,−σ)2(ga
1
2
−
k,ω,−σga
k,ω,σ(ga
k,ω,σ − ga
k,ω,−σ)2 + (ga
k,ω,−σ)2(ga
k,ω,σ)2
k,ω,σ(gr
k,ω,σ − gr
k,ω,−σ)2 − (gr
k,ω,−σ)2(gr
k,ω,σ)2i(cid:27)
(q′, Ω)A(ex)±
(q − q′)
ν
kµkν(cid:26) dfω
dω(cid:20)gr
k,ω,σga
k,ω,−σga
k,ω,σ − gr
k,ω,−σgr
k,ω,σga
k,ω,−σ
k,ω,−σ)ga
k,ω,−σ −
1
2
gr
k,ω,σ(gr
k,ω,σ − gr
k,ω,−σ)gr
k,ω,−σ(cid:21)
k,ω,σ)2 − (gr
k,ω,−σ)2(gr
k,ω,σ)2i(cid:27).
(14)
,
(15)
Using the following relation,
± i∇νA(so)±
µ
+
2e
the spin density S±(r, t) is rewritten as
A(so)±
µ A(ex)z
ν
=
2e
A(so)z
µ A(ex)±
ν
S±(r, t) =
8e23
m2
e
A(so)z
µ
(r, t)A(ex)±
ν
+
4e23Jex
m2
e
(r, t)
µ
∂A(so)z
∂t
A(ex)±
ν
±iIm(cid:20)σ
dfω
dω
gr
k,ω,σ(ga
k,ω,−σ)2ga
(r)ImXσ=±
σXk Xω
(r)Xσ=±Xk Xω
k,ω,σ + 2Jexfω(ga
fωkµkνga
k,ω,−σ(ga
k,ω,σ)2
kµkν(cid:26)dfω
dω
k,ω,−σ)3(ga
k,ω,σ)3(cid:21)(cid:27).
Re(ga
k,ω,σ − gr
k,ω,σ)(ga
k,ω,−σ)2ga
k,ω,σ
(16)
Here we neglect contributions from the products of ga to calculate terms proportional to the
first order in Ω as they only give small corrections compared with those from the products
20
of gr and ga. Therefore, the spin density reduces to
S±(r, t) =
8e23
m2
e
−
Jex
2
∂A(so)z
µ
A(ex)±
ν
(r)(cid:20)A(so)z
µ
(r, t)ImXσ=±
σXk Xω
(Re ∓ iσIm)Xk Xω
2 hC (2)
(r, t) −
Jex
µ
(r, t)
∂t Xσ=±
(r)(cid:26)C (1)
µν A(so)z
=
8e2
me
A(ex)±
ν
fωkµkνga
k,ω,−σ(ga
k,ω,σ)2
dfω
dω
kµkνgr
k,ω,σ(ga
k,ω,−σ)2ga
k,ω,σ(cid:21)
µν ∓ iC (3)
µ
µν i ∂A(so)z
∂t
(r, t)
(cid:27).
(17)
The summations over k and ω are performed in the following manner,
(ω − ε + εF,−σ − iη)(ω − ε + εF,σ − iη)2
ε
2
me
νe
2π
C (1)
µν =
=
=
=
4πJ 2
ex
4πJ 2
ex
fωkµkνga
k,ω,−σ(ga
k,ω,σ)2
νe
νe
0
σXk Xω
ImXσ=±
δµνImXσ=±Z ∞
δµνImXσ=±
δµνImXσ=±
δµνImZ Jex
δµνImZ Jex
δµν(cid:26) εF
ex − η2
Jex
νe
νe
−Jex
−Jex
4πJ 2
ex
4πJ 2
ex
νeη
16πτ Jex
F − J 2
ε2
Jexη
ε
ε − σJex
−∞
−∞
−∞
−∞
−∞
dω
dω
dω
dε Z 0
dε Z 0
dε Z 0
σZ 0
σZ −σJex
dε Z 0
dε εh ln (ε + εF − iη) + iπi
ln(cid:20) (εF + Jex)2 + η2
(εF − Jex)2 + η2(cid:21) − 2
εF + Jex(cid:19) − tan−1(cid:18)
ω + ε + εF − iη
dω
−∞
η
ε
(cid:20) tan−1(cid:18)
= −
= −
= −
+
ω + ε + εF,σ − iη
ω + ε + εF − iη
η
εF − Jex(cid:19)(cid:21)(cid:27),
(18)
21
C (2)
µν =
=
=
2
me
νe
2π
ReXσ=±Xk Xω
δµνReXσ=±Z 0
δµνReXσ=±
−∞
νe
8πJex
dfω
dω
kµkνgr
k,ω,σ(ga
k,ω,−σ)2ga
k,ω,σ
dε
(ε + εF,σ + iη)(ε + εF,−σ − iη)2(ε + εF,σ − iη)
ε
σZ 0
−∞
dε(cid:20)
1
σεFJex
i2η(σJex − iη)2(cid:18)
ε + εF − σJex − iη(cid:19) −
1
−
−
1
ε + εF + σJex − iη(cid:19)
(ε + εF − σJex + iη)2(cid:21)
1
εF − σJex + iη
σJex − iη
ε + εF − σJex + iη
−
1
i2η(cid:18)
1
ε + εF − σJex + iη
=
νe
8πJex
δµνReXσ=±
σ(cid:26)
σεFJex
i2η(σJex − iη)2(cid:20) ln(cid:18)εF − σJex + iη
εF + σJex − iη(cid:19) − i2π(cid:21)
σJex − iη(cid:27)
1
−
= −
−
−
1
εF − σJex − iη(cid:19) − i2π(cid:21) +
i2η(cid:20) ln(cid:18) εF − σJex + iη
ex − η2)
ex + η2)
ex + η2)2 δµν(cid:26)1 −
ex − η2)
η
νeτ εF(J 2
2(J 2
1
η(J 2
πεF(J 2
2π(cid:20) tan−1(cid:18)
ex + η2)2
(J 2
εF + Jex(cid:19) + tan−1(cid:18)
ex − η2)(cid:20) tan−1(cid:18)
2πεFJex(J 2
η
εF − Jex(cid:19)(cid:21)
+
ηJex
ex − η2)
2π(J 2
ln(cid:20) (εF + Jex)2 + η2
(εF − Jex)2 + η2(cid:21)
η
εF + Jex(cid:19) − tan−1(cid:18)
η
εF − Jex(cid:19)(cid:21)(cid:27),
(19)
C (3)
µν =
2
me
ImXσ=±
σXk Xω
δµνImXσ=±Z 0
−∞
νe
8πJex
=
dfω
dω
kµkνgr
k,ω,σ(ga
k,ω,−σ)2ga
k,ω,σ
dε(cid:20) εF(σ2Jex − iη)
σ2Jex(σJex − iη)2(cid:18)
1
ε + εF − σJex + iη
−
1
ε + εF + σJex − iη(cid:19)
εF − σJex + iη
σJex − iη
1
(ε + εF − σJex + iη)2(cid:21)
−
1
σ2Jex(cid:18)
=
νe
8πJex
δµνImXσ=±
ε + εF − σJex + iη
1
1
−
ε + εF − σJex − iη(cid:19) −
σ2Jex(σJex − iη)2(cid:20) ln(cid:18)εF − σJex + iη
(cid:26) εF(σ2Jex − iη)
εF − σJex − iη(cid:19) − i2π(cid:21) +
1
1
σ2Jex(cid:20) ln(cid:18) εF − σJex + iη
εF + σJex − iη(cid:19) − i2π(cid:21)
σJex − iη(cid:27)
ex + η2)
8πJ 3
ex
ln(cid:20) (εF + Jex)2 + η2
(εF − Jex)2 + η2(cid:21)
νeεFJex
ex + η2)2 δµν(cid:26)1 −
η(J 2
ex + η2)
2πεFJ 2
ex
+
η(3J 2
η
εF + Jex(cid:19) + tan−1(cid:18)
η
εF − Jex(cid:19)(cid:21)
2(J 2
1
2π(cid:20) tan−1(cid:18)
ex + η2)2
4πεFJ 3
(J 2
ex (cid:20) tan−1(cid:18)
−
= −
−
−
η
εF + Jex(cid:19) − tan−1(cid:18)
η
εF − Jex(cid:19)(cid:21)(cid:27).
(20)
22
|
1808.05960 | 2 | 1808 | 2018-10-22T14:27:36 | Dynamical density response and optical conductivity in topological metals | [
"cond-mat.mes-hall"
] | Topological metals continue to attract attention as novel gapless states of matter. While there by now exists an exhaustive classification of possible topologically nontrivial metallic states, their observable properties, that follow from the electronic structure topology, are less well understood. Here we present a study of the electromagnetic response of three-dimensional topological metals with Weyl or Dirac nodes in the spectrum, which systematizes and extends earlier pioneering studies. In particular, we argue that a smoking-gun feature of the chiral anomaly in topological metals is the existence of propagating chiral density modes even in the regime of weak magnetic fields. We also demonstrate that the optical conductivity of such metals exhibits an extra peak, which exists on top of the standard metallic Drude peak. The spectral weight of this peak is transferred from high frequencies and its width is proportional to the chiral charge relaxation rate. | cond-mat.mes-hall | cond-mat | Dynamical density response and optical conductivity in topological metals
Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada
(Dated: October 23, 2018)
A.A. Burkov
Topological metals continue to attract attention as novel gapless states of matter. While there
by now exists an exhaustive classification of possible topologically nontrivial metallic states, their
observable properties, that follow from the electronic structure topology, are less well understood.
Here we present a study of the electromagnetic response of three-dimensional topological metals with
Weyl or Dirac nodes in the spectrum, which systematizes and extends earlier pioneering studies. In
particular, we argue that a smoking-gun feature of the chiral anomaly in topological metals is the
existence of propagating chiral density modes even in the regime of weak magnetic fields. We also
demonstrate that the optical conductivity of such metals exhibits an extra peak, which exists on
top of the standard metallic Drude peak. The spectral weight of this peak is transferred from high
frequencies and its width is proportional to the chiral charge relaxation rate.
I.
INTRODUCTION
Topological metal (TM) is a recently discovered new
phase of matter.1 -- 18 It is characterized by topological in-
variants, defined on the Fermi surface,19 -- 22 rather than in
the whole Brillouin zone (BZ), as in topological insulators
(TI). Such Fermi surface topological invariants arise as
a consequence of monopole-like singularities in the elec-
tronic structure, Weyl nodes, whose significance was em-
phasized early on by Volovik and by Murakami.19,22
Perhaps the most interesting feature of TM is that
their electronic structure topology leads not only to spec-
troscopic manifestations in the form of edge states,5 a
feature they share with TI, but also to nontrivial re-
sponse. This novel response is usually described as being
a consequence of the chiral anomaly,23 which may be un-
derstood in the following way. While the appearance of
gapless Weyl nodes in the spectrum has a topological ori-
gin, it also leads to an emergent symmetry, or an emer-
gent conservation law, namely conservation of the chiral
charge. This conservation law becomes increasingly more
precise as the Fermi energy of the TM approaches the
Weyl nodes. However, this apparent low-energy conser-
vation law is violated when the system is coupled to an
electromagnetic field. The origin of this violation lies in
the fact that the chiral symmetry can never be an exact
symmetry of a (3+1)-dimensional Dirac fermion on a lat-
tice, as first pointed out by Nielsen and Ninomiya,24 as
a single (or, more generally, an odd number) Dirac point
in the BZ is topologically incompatible with the chiral
symmetry. Thus, while the chiral symmetry appears to
be present when one focuses only on states at the small
Fermi surface, enclosing the Weyl points, the global lack
of chiral symmetry manifests in the electromagnetic re-
sponse of the system. This property is of great interest
both because it has a topological origin and because it
is contrary to one of the fundamental postulates of the
standard theory of metals, which states that anything of
observable consequence in a metal involves only states on
the Fermi surface.
The chiral anomaly in TM has numerous predicted ob-
servable consequences, which include negative longitudi-
nal magnetoresistance (LMR),25,26 giant planar Hall ef-
fect (PHE),27,28 and anomalous Hall effect.29 While most
of these have already been observed experimentally in
various TM materials,18,30 -- 34 none of these phenomena
by themselves may be regarded as smoking-gun manifes-
tations of the chiral anomaly, in the sense that all of them
may in principle arise from unrelated sources, and these
sources all have to be ruled out before the chiral anomaly
origin may be claimed. An excellent discussion of these
issues in the case of the negative LMR may be found in
Ref. 35.
As first discussed by Altland and Bagrets,36 a truly
unique feature of the chiral anomaly is the highly un-
usual dependence of the transport properties, such as the
sample conductance, on the relevant length (and time or
frequency, as will be shown in this paper) scales. In an
ordinary three-dimensional (3D) metal the conductance
scales linearly with the sample size L
G(L) = σL,
(1)
where the Drude conductivity σ is related to the density
of states at the Fermi energy g and the diffusion constant
D by the Einstein relation
σ = e2gD.
(2)
Corrections to Eq. (1) are small in good metals, the small
parameter being 1/kF (cid:96), where kF is the Fermi momen-
tum (¯h = c = kB = 1 units are used henceforth) and (cid:96)
is the mean free path; the corrections arise only at very
low temperatures as a result of quantum interference phe-
nomena. The scaling of Eq. (1) is partly a consequence of
the fact that, in an ordinary metal in the diffusive trans-
port regime, i.e. at length scales, longer than the mean
free path (cid:96) and time scales longer than the momentum
relaxation time τ , no intrinsic hydrodynamic (i.e. long)
length scales remain, besides the sample size L.
However, as discussed in Ref. 36, in a TM two addi-
tional hydrodynamic length scales emerge. These are the
chiral charge diffusion length
(cid:112)
Lc =
Dτc,
(3)
8
1
0
2
t
c
O
2
2
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
2
v
0
6
9
5
0
.
8
0
8
1
:
v
i
X
r
a
G(L) =
e2Nφ
2π
,
(5)
where
where τc (cid:29) τ is the chirality relaxation time, and
La = D/Γ,
(4)
where Γ = eB/2π2g and B is the applied magnetic field.
La is a new purely quantum mechanical magnetic-field-
√
related length scale, which is distinct from the magnetic
eB and which arises from the chiral
length (cid:96)B = 1/
anomaly. It is related to the magnetic length as La ∼
(cid:96)(kF (cid:96)B)2 and is thus much longer than the mean free
path in the weak-field (quasiclassical) regime kF (cid:96)B (cid:29) 1,
which we will be interested in here. Transport properties
of TM may then be shown to depend strongly on the
interplay of the three length scales: L, Lc, and La.27,36
In particular, the strength of the negative LMR and the
PHE depends on the parameter Lc/La, getting stronger
as this ratio increases.
Particularly striking phenomena arise when La < L <
c/La,36 which is an extended and accessible range when
L2
Lc/La (cid:29) 1.
In this regime the sample conductance is
given by
where Nφ = L2/2π(cid:96)2
B is the number of magnetic flux
quanta, piercing the sample with cross-section area L2.
This means that in the regime La < L < L2
c/La the
sample transports electric current as Nφ one-dimensional
(1D) conduction channels and the conduction is ballistic
and dissipationless [of course Eq. (5) only represents the
dominant part of the conductance and ordinary dissipa-
tive Ohmic conduction is also present]. This is striking
because it arises in a 3D metal with a Fermi surface and
in the weak magnetic field regime kF (cid:96)B (cid:29) 1. The ex-
istence of such ballistic quasi-1D transport regime is a
smoking-gun manifestation of the chiral anomaly in 3D
TM.
In this paper we further elaborate on this striking prop-
erty of TM and consider their related dynamical proper-
ties.
In particular, we demonstrate that the quasi-1D
transport regime manifests in dynamics as chiral propa-
gating density modes, which exist in a range of wavevec-
tor values given by
La/L2
c < q < 1/La.
(6)
This "one-dimensionalization" of the electron dynamics
is a unique property of TM, related to the chiral anomaly.
We also demonstrate that related phenomena exist in
frequency-dependent properties of TM. In particular we
demonstrate that the frequency dependence of the op-
tical conductivity of TM has a non-Drude form, where
an extra narrow peak exists at low frequencies, whose
width scales as 1/τc while height is a function of the
ratio Lc/La. The spectral weight of this extra peak is
transferred from high frequencies.
The rest of the paper is organized as follows. In Sec-
tion II we calculate the full density response function of
a simple model of a TM in an external magnetic field.
2
We analyze the eigenmode structure of the density re-
sponse function and demonstrate the presence of chiral
propagating density modes when La/L2
c < q < 1/La. In
Section III we relate the existence of these propagating
chiral modes to observable transport properties of TM.
We also demonstrate that similar phenomena exist in the
frequency domain: we analyze the frequency dependence
of the optical conductivity and point out its non-Drude
nature. We conclude in Section IV with a brief discussion
of the main results.
II. DENSITY RESPONSE FUNCTION OF A
TOPOLOGICAL METAL
We start from the simplest model of a TM, which con-
tains the necessary ingredients to capture the physics we
want to describe. The simplest such model is the follow-
ing model of a lattice Dirac fermion
H = tγ0γµ sin kµ + ∆(k)γ0,
∆(k) = t(3 − cos kx − cos ky − cos kz),
(7)
(8)
and γµ are Dirac gamma matrices in, for example, the
Weyl representation
γ0 = τ x, γi = −iτ yσi, i = 1, 2, 3.
(9)
This model describes two Weyl nodes of opposite chi-
rality at the Γ-point in the BZ (the effects we will be
discussing do not depend on the momentum-space sep-
aration between the Weyl nodes). Since a single Dirac
point in the BZ is incompatible with the chiral symme-
try, Eq. (7) also has an essential property, shared by all
real Weyl and Dirac semimetals, that the chiral symme-
try (chiral charge conservation) is only an approximate
low-energy symmetry of Eq. (7), which emerges when H
is expanded to linear order in k near the Γ-point. In this
case we have
H = tγ0γµkµ,
(10)
and the chirality operator γ5 = iγ0γ1γ2γ3 = τ z com-
mutes with H, which is no longer true once nonlinear
terms are included. This gives a finite (but small) chiral
charge relaxation rate, which is an essential property of
a Weyl or Dirac semimetal.
We add a uniform magnetic field in the z-direction
B = B z, and choose the Landau gauge for the vector
potential A = xB y. We will ignore the Zeeman effect for
simplicity. To find the eigenstates of H in the presence
of the magnetic field, we expand to first order in kx,y,
while keeping the full kz dependence. This is a good ap-
proximation in the regime of weak magnetic fields when
kF (cid:96)B (cid:29) 1, which we will be interested in. For computa-
tional convenience we also make the following canonical
transformation in the original Weyl representation of the
gamma-matrices:
τ x,y → σzτ x,y, σx,y → τ zσx,y.
(11)
This brings the Hamiltonian to the form
H = t(σxπx + σyπy) + m(kz)σz,
(12)
where π = −i∇ + eA is the canonical momentum and
m(kz) = tτ z sin kz + ∆(0, 0, kz)τ x.
(13)
Diagonalizing Eq. (12), we find the eigenstate wavefunc-
tions
n, s, p, ky, kz(cid:105) = zsp
+ zsp
n↑τ (kz)n − 1, ky, kz,↑, τ(cid:105)
n↓τ (kz)n, ky, kz,↓, τ(cid:105),
(14)
where n is an integer Landau level index, ↑,↓ label the
two eigenvalues of σz, τ = ± are the two eigenval-
ues of τ z, and s, p = ±. Here and throughout sums
over repeated indices will be implicit. The amplitudes
zsp
nστ (kz) may be regarded as components of an eigenvec-
tor zsp
(cid:115)
(cid:32)(cid:115)
n (kz)(cid:105) ⊗ up(kz)(cid:105), where
n (kz)(cid:105) = vsp
(cid:115)
(cid:32)(cid:115)
t sin kz
m(kz)
1√
2
1 + p
1 − p
, p
up
n(kz)(cid:105) =
(cid:33)
(cid:33)
,
t sin kz
m(kz)
vsp
n (kz)(cid:105) =
1√
2
1 + sp
m(kz)
n(kz)
, s
1 − sp
m(kz)
n(kz)
.
(15)
The corresponding energy eigenvalues are given by
(cid:113)
2ω2
Bn + m2(kz),
nsp(kz) = sn(kz) = s
(16)
where m(kz) = 2t sin kz, and ωB = t/(cid:96)B, for all n ≥ 1.
The lowest Landau level (LLL), corresponding to n = 0,
is special: it does not have the s label and its eigenenergy
and the corresponding eigenvector are given by
and
0p(kz) = −pm(kz),
0(kz)(cid:105) = (0, 1).
vp
(17)
(18)
We add to the Hamiltonian Eq. (12) random impurity
potential V (r), which we take to be of the Gaussian white
noise form with (cid:104)V (r)(cid:105) = 0 and
(cid:104)V (r)V (r(cid:48))(cid:105) = γ2δ(r − r(cid:48)).
(19)
We take the impurity potential to be independent of
the spin and orbital pseudospin indices. Physically this
means that the impurities are taken to be nonmagnetic
and the potential is smooth enough that its spatial vari-
ation on the scale of the unit cell of the crystal is neg-
ligible. The last assumption is not essential, but does
simplify the subsequent calculations.
3
We will evaluate the density response for the above
model of a TM using the self-consistent Born approxi-
mation (SCBA) and the ladder approximation to perform
the impurity averaging. This is a conserving approxima-
tion, meaning it preserves exact conservation laws and
sum rules, and amounts physically to neglecting quantum
interference effects. This is justified in the quasiclassical
transport regime, which we will confine ourselves to: we
assume that we are interested in the density response
at length scales much longer than the inverse Fermi mo-
mentum and time scales much longer than the inverse
Fermi energy; the impurity scattering is taken to be weak
enough, so that kF (cid:96) (cid:29) 1 and, as already mentioned,
magnetic field is also assumed to be weak, which means
kF (cid:96)B (cid:29) 1. Finally, we will assume that the Fermi energy
is close to the Dirac point F (cid:28) t (but F τ (cid:29) 1), which
defines the regime of a TM. The last condition ensures
the near conservation of the chiral charge, as will be seen
explicitly below.
The calculation of the SCBA impurity self-energy in
a similar model has already been discussed in detail in
Ref. 26. We will thus omit the details of this calculation
here and simply quote the result. One obtains that in the
quasiclassical transport regime the impurity scattering
rate is independent of both the Landau level index n and
the longitudinal momentum component kz and is given
by the standard SCBA expression
1
τ
=
πγ2g
2
,
(20)
where the density of states at the Fermi energy is given
by
g =
F
πt2
Θ[F − m(kz)],
dkz
2π
(21)
(cid:90) π
−π
Θ(x) being the Heaviside step function.
We evaluate the density response function by summing
the impurity ladder diagrams. We start from the most
general retarded density matrix response function, de-
fined as
χα1α2,α3α4(r, tr(cid:48)
(cid:48))
, t
(cid:48))(cid:104)[α1α2
†
(r, t),
α3α4
= −iΘ(t − t
(r(cid:48)
(cid:48))](cid:105),
, t
(22)
where the density matrix is defined as
α1α2(r, t) = Ψ†
α2
(r, t)Ψα1
(r, t),
(23)
and α = (στ ) is a composite index, which encodes both
the spin and orbital pseudospin labels.
The standard procedure to find the real-time response
function Eq. (22) is to start from the corresponding
imaginary-time response function
χα1α2,α3α4(r, τr(cid:48)
(cid:48))
, τ
, τ − τ
(cid:48))Gα4α2(r(cid:48)
= −Gα1α3 (r, r(cid:48)
(24)
where Gαα(cid:48)(r, r(cid:48), τ − τ(cid:48)) is the exact imaginary-time
Green's function, which depends on both r and r(cid:48) sepa-
rately due to both the lack of translational symmetry in
, r, τ
(cid:48) − τ ),
4
FIG. 1. Diagrammatic representation of (a) SCBA Green's
function. Thin line represents the bare Green's function,
thick line is the SCBA impurity-averaged Green's function
and the dashed line represents the disorder potential correla-
tor (cid:104)V (r)V (r(cid:48))(cid:105) = γ2δ(r − r(cid:48)). (b) Density response function
χ. (c) Bethe-Salpeter equation for the diffusion vertex D.
The physical meaning of the two contributions to the
density response function, χI and χII , is that χI arises
from states on the Fermi surface, while all filled states
contribute to χII . χII thus represents equilibrium part of
the response and is easily shown to be a diagonal matrix,
with the nonzero matrix elements equal to −g. On the
other hand, χI represents the dynamical nonequilibrium
part of the density response and is given by
χI (q, Ω) = − iΩ
2πγ2I(q, Ω)D(q, Ω),
(33)
d3r d3r
(cid:48)
e
−iq·(r−r(cid:48))
Iα1α2,α3α4 (q, Ω) =
, Ω)GA
α1α3
γ2
L3
(r(cid:48)
α4α2
, r, 0),
× GR
(r, r(cid:48)
GR,A being the
retarded and advanced real-time
impurity-averaged SCBA Green's functions. They are
explicitly given by
(cid:88)
αα(cid:48) (r, r(cid:48)
GR,A
(cid:104)r, αn, s, p, ky, kz(cid:105)(cid:104)n, s, p, ky, kzr(cid:48), α(cid:48)(cid:105)
(34)
, ω)
ω − ξnsp(kz) ± i/2τ
,
=
nspkykz
(35)
where ξnsp(kz) ≡ nsp(kz) − F .
For a general direction of the wavevector q, the eval-
uation of I(q, Ω) is severely complicated by the fact
that contributions of different Landau levels are mixed
in Eq. (34). This is not the case only when q = qz,
when translational symmetry in the xy-plane leads to
decoupling of the individual Landau level contributions.
D = 1 + ID,
where I ≡ γ2P 0. The solution of this equation is
(29)
where
(cid:90)
(cid:88)
the presence of a random impurity potential, and the lack
of gauge invariance in the presence of an external mag-
netic field. One then performs impurity averaging, which
restores translational invariance in the density response
function and gives
χα1α2,α3α4 (q, iΩ) =
1
β
where
Pα1α2,α3α4(q, iω, iω + iΩ),
iω
(25)
Pα1α2,α3α4(r − r(cid:48)
= −(cid:104)Gα1α3 (r, r(cid:48)
, iω, iω + iΩ)
, iω + iΩ)Gα4α2 (r(cid:48)
, r, iω)(cid:105),
(26)
is the impurity-averaged generalized polarization bubble,
and β = 1/T is the inverse temperature. In the quasiclas-
sical regime we are interested in, P may be evaluated by
summing all the SCBA diagrams for the impurity self-
energy and the ladder vertex corrections, as shown in
Fig. 1. The result of this diagram summation may be
written in a shorthand matrix notation as
P = P 0D,
(27)
where P 0 is the bare polarization bubble, in which only
the self-energy corrections are included
P 0
α1α2,α3α4
= −Gα1α3 (r, r(cid:48)
is
(r − r(cid:48)
, iω, iω + iΩ)
, iω + iΩ)Gα4α2 (r(cid:48)
Gαα(cid:48)(r, r(cid:48), iω) here
the disorder-averaged SCBA
Green's function, which still depends on r and r(cid:48) sepa-
rately since it is a gauge-dependent quantity in the pres-
ence of an external magnetic field. The vertex part D,
which is also known as the diffusion propagator, or dif-
fuson, satisfies the following Bethe-Salpeter equation
, r, iω).
(28)
D = (1 − I)−1.
(30)
(cid:90) ∞
To obtain the real-time retarded response function we
analytically continue to real frequency iΩ → Ω+iη, which
gives
−∞
d
2πi
χ(q, Ω) =
nF () [P(q, + iη, + Ω + iη)
− P(q, − iη, + Ω + iη) + P(q, − Ω − iη, + iη)
− P(q, − Ω − iη, − iη)] .
(31)
In the low-frequency limit, when Ω (cid:28) F , this simplifies
to
χ(q, Ω) = − iΩ
2π
(cid:90) ∞
P(q,−iη, Ω + iη)
d nF ()ImP(q, + iη, + Ω + iη)
− 1
π
≡ χI (q, Ω) + χII (q, Ω).
−∞
(32)
=+=+ <latexit sha1_base64="TwKQu5I6W5bc9HOXn7rGty6B3x4=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGtCYD+sNr+nNgVeJX5IGlGgP61+DUUIzyZSlghjT973UBjnRllPBZrVBZlhK6ISMWd9RRSQzQT6/dYbPnDLCUaJdKYvn6u+JnEhjpjJ0nZLY2Cx7hfif189sdB3kXKWZZYouFkWZwDbBxeN4xDWjVkwdIVRzdyumMdGEWhdPzYXgL7+8SroXTd9r+veXjdZNGUcVTuAUzsGHK2jBHbShAxRieIZXeEMSvaB39LForaBy5hj+AH3+AP+rjjI=</latexit><latexit sha1_base64="TwKQu5I6W5bc9HOXn7rGty6B3x4=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGtCYD+sNr+nNgVeJX5IGlGgP61+DUUIzyZSlghjT973UBjnRllPBZrVBZlhK6ISMWd9RRSQzQT6/dYbPnDLCUaJdKYvn6u+JnEhjpjJ0nZLY2Cx7hfif189sdB3kXKWZZYouFkWZwDbBxeN4xDWjVkwdIVRzdyumMdGEWhdPzYXgL7+8SroXTd9r+veXjdZNGUcVTuAUzsGHK2jBHbShAxRieIZXeEMSvaB39LForaBy5hj+AH3+AP+rjjI=</latexit><latexit sha1_base64="TwKQu5I6W5bc9HOXn7rGty6B3x4=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGtCYD+sNr+nNgVeJX5IGlGgP61+DUUIzyZSlghjT973UBjnRllPBZrVBZlhK6ISMWd9RRSQzQT6/dYbPnDLCUaJdKYvn6u+JnEhjpjJ0nZLY2Cx7hfif189sdB3kXKWZZYouFkWZwDbBxeN4xDWjVkwdIVRzdyumMdGEWhdPzYXgL7+8SroXTd9r+veXjdZNGUcVTuAUzsGHK2jBHbShAxRieIZXeEMSvaB39LForaBy5hj+AH3+AP+rjjI=</latexit><latexit sha1_base64="TwKQu5I6W5bc9HOXn7rGty6B3x4=">AAAB63icbVBNSwMxEJ3Ur1q/qh69BIvgqeyKoMeiF48VbC20S8mm2W5okl2SrFCW/gUvHhTx6h/y5r8x2+5BWx8MPN6bYWZemApurOd9o8ra+sbmVnW7trO7t39QPzzqmiTTlHVoIhLdC4lhgivWsdwK1ks1IzIU7DGc3Bb+4xPThifqwU5TFkgyVjzilNhCGtCYD+sNr+nNgVeJX5IGlGgP61+DUUIzyZSlghjT973UBjnRllPBZrVBZlhK6ISMWd9RRSQzQT6/dYbPnDLCUaJdKYvn6u+JnEhjpjJ0nZLY2Cx7hfif189sdB3kXKWZZYouFkWZwDbBxeN4xDWjVkwdIVRzdyumMdGEWhdPzYXgL7+8SroXTd9r+veXjdZNGUcVTuAUzsGHK2jBHbShAxRieIZXeEMSvaB39LForaBy5hj+AH3+AP+rjjI=</latexit>=(a)(b)(c)5
near conservation is a defining property of a TM, as dis-
cussed above. Thus we may project the original 16 × 16
matrix onto the 2 × 2 subspace, describing the coupled
transport of the electric and the chiral charge, which is
accomplished as
Iab(q, Ω) =
Iα1α2,α3α4(q, Ω)Γb
(36)
Γa
α3α4 ,
α2α1
1
4
where a, b = 0, 5, corresponding to the electric (0) or chi-
ral (5) charges, and Γa,b are the corresponding operators,
i.e.
Γ0 = τ 0σ0 = 1, Γ5 = τ zσ0 = τ z.
(37)
After a tedious, but straightforward, calculation, we ob-
tain
Fortunately, this is in fact the case of primary interest to
us, since the chiral anomaly leads to unusual transport
phenomena in the direction of the magnetic field. Thus
we will take q = qz henceforth.
In this case the evaluation of I(q, Ω) is relatively
straightforward, particularly in the weak magnetic field
regime kF (cid:96)B (cid:29) 1 that we are interested in. An addi-
tional simplification arises from the fact that we are not
interested in the whole 16 × 16 matrix I, which contains
a lot of unnecessary information. We are interested only
in the response of conserved, or nearly conserved, quan-
tities, which will always dominate everything else at long
times and long distances.
In a generic TM, we expect
only two such quantities to exist: the electric charge,
which is strictly conserved, and the chiral charge, whose
(cid:18) 1 − iΩτ − iq(cid:96)
(cid:19)
− i(F /t)2(1 − iΩτ )4
1 − iΩτ + iq(cid:96)
,
ln
8(q(cid:96))5
i
2q(cid:96)
(cid:20) i
I00(q, Ω) =
I55(q, Ω) =
I05(q, Ω) = I50(q, Ω) =
2q(cid:96)
(cid:21)
ln
(cid:18) 1 − iΩτ − iq(cid:96)
1 − iΩτ + iq(cid:96)
(cid:19)
− (F /t)2(1 − iΩτ )
12(q(cid:96))2
+
(F /t)2(1 − iΩτ )3
4(q(cid:96))4
,
i
2(kF (cid:96)B)2
q(cid:96)
(1 − iΩτ )2 + (q(cid:96))2 .
(38)
Substituting this into Eq. (33), we obtain the dynami-
cal nonequilibrium contribution to the density response
χI (q, Ω), while the equilibrium contribution is a diagonal
matrix given by
55(q, Ω) = −g,
(39)
χII
00(q, Ω) = χII
as already mentioned above.
A comment is in order here. As can be seen from
Eq. (38), only the off-diagonal matrix element I05 de-
pends on the magnetic field. This is true in the quasi-
classical limit kF (cid:96)B (cid:29) 1 only, and is a consequence of
the fact that in this limit we may ignore the effect of
the magnetic field on the density of states. Summation
over the Landau level index n, which arises when evaluat-
ing Eq. (38), may in this case be replaced by integration
and the magnetic field dependence disappears to lead-
ing order in 1/kF (cid:96)B. In contrast, the off-diagonal matrix
element I05 arises entirely from the contribution of the
n = 0 Landau level. This contribution is proportional to
1/(kF (cid:96)B)2, but leads to large effects at long length scales
and long times, as will be seen below, provided τc/τ (cid:29) 1.
Eqs. (32), (33), (38) and (39) give a general expression
for the density response function of a TM in the quasi-
classical regime
χ(q, Ω) = −g[iΩτI(q, Ω)D(q, Ω) + 1].
(40)
This expression is valid in either diffusive Ωτ, q(cid:96) (cid:28) 1 or
ballistic Ωτ, q(cid:96) (cid:29) 1 limits and may be used, in particular,
to study the ballistic-diffusive crossover regime. We will
start by analyzing the two limits.
A. Ballistic regime
In this regime all components of the matrix I are small
and thus D ≈ 1. Physically this means that we are look-
ing at short length and time scales at which the impurity
scattering may be ignored. While the response function
χ(q, Ω) is a 2 × 2 matrix, only its χ00(q, Ω) component
describes observable density response. Taking the limit
Ωτ, q(cid:96) → ∞ in Eqs. (38), (40) we obtain
(cid:18) Ω − qt + iη
(cid:19)(cid:21)
Ω + qt + iη
.
(41)
χ00(q, Ω) = −g
1 +
Ω
2qt
ln
(cid:20)
This is just the familiar Lindhard function (in the limit
q (cid:28) kF and Ω (cid:28) F ), describing the density response
of a clean Fermi liquid with the Fermi velocity t. The
imaginary part of χ00(q, Ω), which is determined by the
branch cuts of the Lindhard function
Imχ00(q, Ω) = g
Θ(qt − Ω),
πΩ
2qt
(42)
describes the excitation spectrum of the Fermi liquid,
which forms a particle-hole continuum. Thus in the bal-
listic regime and in the weak magnetic field limit chi-
ral anomaly has no effect on the density response of a
TM [its effects appear only at order 1/(kF (cid:96)B)2, which is
negligible compared to Eq. (41)]. This of course will no
longer be true if we tune the Fermi energy to zero (i.e.
to the ideal Weyl or Dirac semimetal limit), but this is
a fine-tuned, non-generic situation, and is of somewhat
less interest for this reason.
B. Diffusive regime
The situation is much more interesting in the diffusive
limit Ωτ, q(cid:96) (cid:28) 1. In this case I ≈ 1, and multiple impu-
rity scattering needs to be taken into account. Expanding
in Taylor series in Ωτ and q(cid:96), we obtain
I00(q, Ω) ≈ 1 + iΩτ − Dq2τ,
I05(q, Ω) = I50(q, Ω) ≈ iΓqτ,
I55(q, Ω) ≈ 1 + iΩτ − τ /τc − Dq2τ.
(43)
Here D = t2τ /3 = t(cid:96)/3 is the diffusion constant,
Γ =
eB
2π2g
=
t
2(kF (cid:96)B)2 ,
(44)
is a new transport coefficient, which describes the chiral-
anomaly-induced coupling between the electric and the
chiral charge densities, and
1
τc
=
2
F
20 t2τ
,
(45)
is the chiral charge relaxation rate. Note that the fact the
chiral charge relaxation rate vanishes in the limit F → 0
is a consequence of our assumption that the impurity
potential is diagonal in the spin and orbital indices and
thus commutes with the chiral charge operator γ5 = τ z.
In general this is not the case and we can expect some
residual chiral charge relaxation even in the F → 0 limit.
In the diffusive regime the dynamics of the density re-
sponse is determined by the poles of the diffusion propa-
gator D, instead of the branch cuts of the response func-
tion, as in the ballistic limit. From Eq. (43), the inverse
diffusion propagator is given by
D−1(q, Ω) =
(cid:18) −iΩτ + Dq2τ
−iΓqτ
(cid:19)
−iΩτ + τ /τc + Dq2τ
−iΓqτ
,
6
which gives the following explicit expression for the 00
component of the matrix response function, which corre-
sponds to the observable electric charge density response
(cid:20) Ω(Ω + i/τc + iDq2)
(Ω − Ω+)(Ω − Ω−)
(cid:21)
− 1
.
(51)
χ00(q, Ω) = g
We now note that the frequency Ω0 is purely imaginary
at the smallest momenta when
where we have introduced two new length scales
q <
1
2Γτc
=
La
2L2
c
≡ 1
L∗
,
(cid:112)
Lc =
Dτc,
(52)
(53)
which has the meaning of the chiral charge diffusion
length and
La =
D
Γ
=
2
3
(cid:96)(kF (cid:96)B)2.
(54)
La is a magnetic-field-related length scale, distinct
from the magnetic length, which arises from the chiral
anomaly. It is a long hydrodynamic length scale in the
weak magnetic field regime, in the sense that La (cid:29) (cid:96), but
it may still be much smaller that either the chiral charge
diffusion length Lc or the sample size L. In fact, the ra-
tio Lc/La quantifies the strength of the chiral-anomaly-
related density response phenomena, as will be seen be-
low.
Thus when q < 1/L∗ the eigenfrequencies of the diffu-
sion propagator are purely imaginary, which corresponds
to ordinary diffusion (nonpropagating) modes. However,
when q > 1/L∗ (which may be a very small momentum
when the ratio Lc/La is large), Ω0 is real, which signals
the emergence of a pair of propagating modes in this
regime. The modes are only weakly damped as long as
(46)
Ω0 ≈ Γq > Dq2,
(55)
The zeros of the determinant of this matrix determine
the eigenmode frequencies
where
Ω± = ±Ω0 − i(Dq2 + 1/2τc),
Ω0 =(cid:112)Γ2q2 − 1/4τ 2
c .
(47)
(48)
which defines the upper limit on the wavevector q =
1/La, above which the propagating modes disappear.
The propagating modes thus exist in the interval
1/L∗ < q < 1/La.
(56)
This interval is significant when Lc/La (cid:29) 1.
Within this interval of q the density response function
The diffusion propagator itself may then be written as
takes the following approximate form
D(q, Ω) =
1/τ
(Ω − Ω+)(Ω − Ω−)
(cid:18) iΩ − Dq2 − 1/τc
×
−iΓq
(49)
Taking into account that in the diffusive regime I ≈ 1,
iΩ − Dq2
−iΓq
.
we obtain from Eq. (40)
χ(q, Ω) ≈ −g[iΩτD(q, Ω) + 1],
(50)
(cid:19)
χ00(q, Ω) = g
Ω2
0
(Ω + iDq2)2 − Ω2
0
,
(57)
where Ω0 = Γq. This is the density response func-
tion of an effective 1D system with the Fermi velocity
Γ = t/2(kF (cid:96)B)2 (cid:28) t. Note that this is very different
from the 1D response one would obtain in a TM in the
quantum limit kF (cid:96)B < 1, when only the lowest n = 0
Landau level contributes to the density response. In this
case one gets Nφ 1D modes, which correspond to Nφ or-
bital states within the LLL. The Fermi velocity of these
1D modes is equal to the microscopic Fermi velocity t. In
our case, while the ultimate origin of the 1D dynamics is
still the LLL, its emergence is only possible in the diffu-
sive regime and thus requires multiple impurity scatter-
ing. The corresponding Fermi velocity Γ is proportional
to the applied magnetic field and is much smaller than t
in the quasiclassical weak-field regime. Such effectively
1D density response, with propagating rather than diffu-
sive density dynamics, which exists in a 3D metal with a
large Fermi surface (kF (cid:96) (cid:29) 1) in a weak magnetic field
(kF (cid:96)B (cid:29) 1) is a truly unique feature of TM and should
be regarded as their true smoking-gun characteristic.
On the other hand, when La > Lc, propagating modes
do not exist for any q and one obtains a pair of standard
diffusion modes
Ω+ = −iDq2, Ω− = −iDq2 − i/τc,
(58)
which correspond to independent diffusion of the electric
and the chiral charge densities.
III. TRANSPORT IN TOPOLOGICAL METALS
It is very useful to also look at the transport properties,
which follow from the density response, described in Sec-
tion II. In addition to providing further insight into the
physical meaning of the results, discussed in the previous
section, this will also allow us to calculate experimentally
measurable physical quantities, such as the frequency-
and scale-dependent conductivity.
A. Scale-dependent conductance
It is easy to show that Eqs. (49) and (50) for the dif-
fusion propagator and the generalized density response
function are equivalent to the following transport equa-
tion in real space and time, that the electric n0 and chiral
n5 charge densities must satisfy
∂n0
∂t
∂n5
∂t
= D∇2(n0 + gV0) + Γ · ∇(n5 + gV5),
= D∇2(n5 + gV5) − n5 + gV5
τc
+ Γ · ∇(n0 + gV0),
(59)
where V0 and V5 are external electric and chiral poten-
tials correspondingly and we have generalized to an ar-
bitrary magnetic field direction, which is why the coeffi-
cient Γ ∝ B has become a vector. The chiral potential
V5 may arise, for example, in a situation when the inver-
sion symmetry is broken, in which case the Weyl nodes
of different chirality will generally be located at differ-
ent energies, V5 being precisely this energy difference.
Otherwise this should simply be regarded as a fictitious
potential, which couples linearly to the chiral charge nc.
Indeed, Fourier transforming Eq. (59) we obtain
7
(cid:19)
.
(60)
(cid:19)
(cid:18) n0
n5
= −g(cid:2)iΩτ + D−1(q, Ω)(cid:3)(cid:18) V0
(cid:19)
(cid:18) V0
V5
= −g[iΩτD(q, Ω) + 1]
,
(61)
V5
D−1(q, Ω)
This gives(cid:18) n0
(cid:19)
n5
which is equivalent to Eq. (50).
Solving Eq. (59) in the steady state, assuming a uni-
form sample of linear size L, attached to normal metal
leads (in which the chiral electrochemical potential n5 +
gV5 = 0) in the z-direction (i.e. the current flows along
the magnetic field), one obtains the following expression
for the scale-dependent sample conductance27,36
G(L) =
e2Nφ
2π
F (L/La, L/Lc),
(62)
where the scaling function F (x, y) is given by
(cid:112)1 + y2/x2 + tanh
(1 + y2/x2)3/2
(cid:112)1 + y2/x2
(cid:16) x
2
(cid:17) . (63)
F (x, y) =
y2
2x
This scaling function exhibits crossover behaviors which
exactly match the corresponding crossovers
in the
wavevector dependence of the diffusion modes, described
in Section II.
Indeed, when x (cid:28) y, which means La (cid:29) Lc, we have
F (x, y) ≈ 2/x, which gives
G(L) ≈ e2gDL = σL,
(64)
which is simply the standard Ohmic conductance, with a
small magnetic-field dependent correction, which goes as
(Lc/La)2, and which we have ignored here for the sake
of brevity.27 This corresponds to the regime, in which we
have two independent diffusion modes, given by Eq. (58),
corresponding to independent diffusion of the electric and
the chiral charges.
On the other hand, when La (cid:28) Lc, or x (cid:29) y, we
obtain
F (x, y) ≈
1
y2/2x + tanh(x/2)
.
(65)
This exhibits a regime of quasiballistic conductance with
G(L) ≈ e2Nφ
2π
,
which is realized when
La < L < L∗.
(66)
(67)
This corresponds precisely to the range of the wavevec-
tors q in Eq. (56), for which propagating modes exist
when La (cid:28) Lc. Thus, one of the observable manifesta-
tions of the existence of quasi-1D propagating modes in a
TM is the quasiballistic conductance, given by Eq. (66).
It is instructive to see what the quasiballistic con-
ductance regime corresponds to directly in terms of the
transport equations Eq. (59). In this regime both the sec-
ond derivative D∇2n0,5 and the relaxation n5/τc terms
may be ignored and we obtain
∂n0
∂t
∂n5
∂t
= Γ
= Γ
∂n5
∂z
∂n0
∂z
,
.
(68)
Introducing the left- and right-handed charges as nR,L =
(n0 ± n5)/2 we obtain
∂nR
∂t
∂nL
∂t
,
∂nR
= Γ
∂z
= −Γ
∂nL
∂z
.
(69)
Eq. (69) describes two chiral bosonic density modes,
which propagate along and opposite to the direction of
the applied magnetic field. Such "bosonization" of the
electron dynamics, which occurs in a 3D metal in a weak
quasiclassical magnetic field, is a characteristic smoking-
gun feature of a TM.
Eq. (69) means, in particular, that a density distur-
bance, created in a TM in magnetic field, with split into
two chiral modes, which will propagate ballistically in
opposite directions, spatially separating electrons of dif-
ferent chirality. It might be possible to detect this effect
optically.37
B. Optical conductivity
Optical conductivity of TM has been studied before,
with a focus mostly on the interband transition ef-
fects.38 -- 42 Here we will demonstrate that low-frequency
intraband optical conductivity is qualitatively affected by
the chiral anomaly, which has not been noticed before.
From the general expression for the density response
function Eq. (40) we may easily obtain the frequency-
dependent conductivity. Indeed, electric charge conser-
vation requires that
σzz(Ω) = −e2 lim
q→0
iΩ
q2 χ00(q, Ω).
(70)
A straightforward calculation then gives
σzz(Ω) =
σ
1 − iΩτ
1 − iΩτc + (Lc/La)2
1 − iΩτc
,
(71)
where σ = e2gD is the zero-field DC conductivity. Eval-
uating the real part, one obtains
(cid:34)
(cid:18) Lc
(cid:19)2 1 − Ω2τ τc
La
1 + Ω2τ 2
c
(cid:35)
Re σzz(Ω) =
σ
1 + Ω2τ 2
1 +
. (72)
In this paper we have studied density response in TM
and the corresponding experimentally observable phe-
nomena. We have argued that one of the truly unique
8
FIG. 2. (Color online) Frequency-dependent conductivity for
Lc/La = 1 (solid line) and Lc/La = 0 (dashed line), and
τ /τc = 0.04.
(cid:90) ∞
0
Eq. (72) is one of the main new results of this paper.
The prefactor in Eq. (72) is the standard Drude expres-
sion for the optical conductivity of a metal. The part
in the square brackets is a correction that arises in a
TM as a consequence of the chiral anomaly. This correc-
tion represents transfer of the spectral weight from high
frequencies into a new low-frequency peak, whose width
scales with the chiral charge relaxation rate 1/τc, while
height is proportional to the ratio (Lc/La)2. Importantly,
Eq. (72) satisfies the exact f -sum rule
dΩ Re σzz(Ω) =
πσ
2τ
,
(73)
which means that the appearance of the new low-
frequency peak indeed represents spectral weight trans-
fer, as it should, see Fig. 2.
It is instructive to examine the high-frequency limit of
τ τc. In this limit
√
Eq. (72), namely when Ω > 1/τc, 1/
we obtain
Re σzz(Ω) ≈
σ
1 + Ω2τ 2
1 − 1
3
.
(74)
(cid:34)
(cid:19)2(cid:35)
(cid:18) (cid:96)
La
The negative second term in the square brackets ex-
presses the reduction of the spectral weight at high fre-
quencies, induced by the chiral anomaly. We note that
while formally the whole expression may become negative
for La (cid:28) (cid:96), this would be outside of the regime of validity
of our theory, which assumes weak magnetic field regime
kF (cid:96)B (cid:29) 1 and thus La (cid:29) (cid:96). Within this regime, the real
part of the optical conductivity is always positive, as it
should be.
IV. DISCUSSION AND CONCLUSIONS
0.51.01.52.0Ωτ0.51.01.52.0Reσzz(Ω)/σ9
features of TM is the existence of propagating density
modes, which are induced by the combined effect of the
chiral anomaly and impurity scattering. The modes exist
only in the diffusive limit and disappear in the ballistic
regime. We have demonstrated that one of the observ-
able manifestations of the existence of such propagating
modes is the highly nontrivial scaling of the conductance
of a TM with the sample size, first pointed out by Altland
and Bagrets.36 We have also demonstrated an entirely
new phenomenon, namely a nontrivial frequency depen-
dence of the optical conductivity, which exhibits trans-
fer of the spectral weight from high frequencies, greater
than 1/
τ τc, into a new non-Drude low-frequency peak
of width 1/τc. The existence of this new narrow peak in
the optical conductivity is a smoking-gun consequence of
the chiral anomaly in TM.
√
One issue we have not touched upon in this pa-
per is the effect of the electron-electron, in particular
long-range Coulomb,
interactions. One might worry
that the Coulomb interactions could push the linearly-
dispersing sound-like mode Eq. (55) to the plasma fre-
quency, as happens in the case of the ordinary elec-
tronic zero sound mode, if short-range interactions are
replaced by Coulomb interactions. This does not happen
in our case, however, since the existence of the sound-
like mode has nothing to do with the electron-electron
interactions. Its physical origin lies in the effective "one-
dimensionalization" of the electron dynamics in a dirty
TM in the presence of even a weak magnetic field. What
this means is that the LLL dominates the density re-
sponse at long times and long distances even when many
higher Landau levels are occupied since the dynamics is
ballistic in the LLL while it is diffusive in the higher
Landau levels. This picture has nothing to do with
the electron-electron interactions and will not be signifi-
cantly modified by them, just as the ordinary low-energy
particle-hole continuum in a clean Fermi liquid is not sig-
nificantly affected by the interactions. The frequency of
e2t2g is not significantly af-
fected by a weak applied magnetic field43,44 and is much
larger than the frequency of the low-energy chiral den-
sity mode Ω0 = Γq, which arises within the low-energy
particle-hole continuum of the clean metal. This means
that the two modes do not interact with each other in
any significant way. However, the issue of collective plas-
mon modes in a dirty TM is interesting in its own right
and will be addressed in a future publication.
the plasmon modes ΩP ∼(cid:112)
ACKNOWLEDGMENTS
We thank Xi Dai for a useful discussion. Financial sup-
port was provided by Natural Sciences and Engineering
Research Council (NSERC) of Canada.
1 N. P. Armitage, E. J. Mele,
and A. Vishwanath, Rev.
(2014).
Mod. Phys. 90, 015001 (2018).
2 M. Z. Hasan, S.-Y. Xu, I. Belopolski, and S.-M. Huang,
Annual Review of Condensed Matter Physics 8 (2017).
3 B. Yan and C. Felser, Annual Review of Condensed Matter
Physics 8 (2017).
4 A. A. Burkov, Annual Review of Condensed Matter
Physics 9, 359 (2018).
5 X. Wan, A. M. Turner, A. Vishwanath,
Savrasov, Phys. Rev. B 83, 205101 (2011).
and S. Y.
6 A. A. Burkov and L. Balents, Phys. Rev. Lett. 107, 127205
(2011).
7 A. A. Burkov, M. D. Hook, and L. Balents, Phys. Rev. B
84, 235126 (2011).
8 G. Xu, H. Weng, Z. Wang, X. Dai, and Z. Fang, Phys.
Rev. Lett. 107, 186806 (2011).
9 S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J.
Mele, and A. M. Rappe, Phys. Rev. Lett. 108, 140405
(2012).
10 Z. Wang, Y. Sun, X.-Q. Chen, C. Franchini, G. Xu,
H. Weng, X. Dai, and Z. Fang, Phys. Rev. B 85, 195320
(2012).
11 Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Phys.
Rev. B 88, 125427 (2013).
12 Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng,
D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai,
Z. Hussain, and Y. L. Chen, Science 343, 864 (2014).
13 M. Neupane, S.-Y. Xu, R. Sankar, N. Alidoust, G. Bian,
C. Liu, I. Belopolski, T.-R. Chang, H.-T. Jeng, H. Lin,
A. Bansil, F. Chou, and M. Z. Hasan, Nat. Commun. 5
14 S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian,
C. Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, S.-
M. Huang, H. Zheng, J. Ma, D. S. Sanchez, B. Wang,
A. Bansil, F. Chou, P. P. Shibayev, H. Lin, S. Jia, and
M. Z. Hasan, Science 349, 613 (2015).
15 B. Q. Lv, N. Xu, H. M. Weng, J. Z. Ma, P. Richard, X. C.
Huang, L. X. Zhao, G. F. Chen, C. E. Matt, F. Bisti, V. N.
Strocov, J. Mesot, Z. Fang, X. Dai, T. Qian, M. Shi, and
H. Ding, Nat Phys 11, 724 (2015).
16 B. Q. Lv, H. M. Weng, B. B. Fu, X. P. Wang, H. Miao,
J. Ma, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen,
Z. Fang, X. Dai, T. Qian, and H. Ding, Phys. Rev. X 5,
031013 (2015).
17 L. Lu, Z. Wang, D. Ye, L. Ran, L. Fu, J. D. Joannopoulos,
and M. Soljaci´c, Science 349, 622 (2015).
18 E. Liu, Y. Sun, L. Muechler, A. Sun, L. Jiao, J. Kroder,
V. Suss, H. Borrmann, W. Wang, W. Schnelle, S. Wirth,
S. T. B. Goennenwein,
and C. Felser, ArXiv e-prints
(2017), arXiv:1712.06722 [cond-mat.mtrl-sci].
19 G. Volovik, The Universe in a Helium Droplet (Oxford:
Clarendon, 2003).
20 F. D. M. Haldane, Phys. Rev. Lett. 93, 206602 (2004).
21 G. E. Volovik, in Quantum Analogues: From Phase Tran-
sitions to Black Holes and Cosmology, Lecture Notes in
Physics, Vol. 718, edited by W. Unruh and R. Schtzhold
(Springer Berlin Heidelberg, 2007).
22 S. Murakami, New Journal of Physics 9, 356 (2007).
23 A. A. Zyuzin and A. A. Burkov, Phys. Rev. B 86, 115133
(2012).
24 H. Nielsen and M. Ninomiya, Physics Letters B 130, 389
(1983).
and N. P. Ong, Phys. Rev. X 8, 031002 (2018).
36 A. Altland and D. Bagrets, Phys. Rev. B 93, 075113
25 D. T. Son and B. Z. Spivak, Phys. Rev. B 88, 104412
(2016).
10
(2013).
26 A. A. Burkov, Phys. Rev. B 91, 245157 (2015).
27 A. A. Burkov, Phys. Rev. B 96, 041110 (2017).
28 S. Nandy, G. Sharma, A. Taraphder, and S. Tewari, Phys.
Rev. Lett. 119, 176804 (2017).
29 A. A. Burkov, Phys. Rev. Lett. 113, 187202 (2014).
30 J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan,
M. Hirschberger, W. Wang, R. J. Cava, and N. P. Ong,
Science 350, 413 (2015).
31 Q. Li, D. E. Kharzeev, C. Zhang, Y. Huang, I. Pletikosic,
A. V. Fedorov, R. D. Zhong, J. A. Schneeloch, G. D. Gu,
and T. Valla, Nat Phys 12, 550 (2016).
32 H. Li, H. Wang, H. He, J. Wang, and S.-Q. Shen, ArXiv
e-prints (2017), arXiv:1711.03671 [cond-mat.mes-hall].
33 N. Kumar, C. Felser,
and C. Shekhar, ArXiv e-prints
(2017), arXiv:1711.04133 [cond-mat.mes-hall].
34 Y. J. Wang, J. X. Gong, D. D. Liang, M. Ge, J. R. Wang,
(2018),
and C. J. Zhang, ArXiv e-prints
W. K. Zhu,
arXiv:1801.05929 [cond-mat.mtrl-sci].
35 S. Liang, J. Lin, S. Kushwaha, J. Xing, N. Ni, R. J. Cava,
37 Q. Ma, S.-Y. Xu, C.-K. Chan, C.-L. Zhang, G. Chang,
Y. Lin, W. Xie, T. Palacios, H. Lin, S. Jia, P. A. Lee,
P. Jarillo-Herrero, and N. Gedik, Nature Physics 13, 842
EP (2017).
38 C. J. Tabert and J. P. Carbotte, Phys. Rev. B 93, 085442
(2016).
39 S. Ahn, E. J. Mele, and H. Min, Phys. Rev. B 95, 161112
(2017).
40 B. Roy, V. Jurici´c, and S. Das Sarma, Scientific Reports
6, 32446 EP (2016).
41 Y. Sun and A.-M. Wang, Phys. Rev. B 96, 085147 (2017).
42 D. Neubauer, A. Yaresko, W. Li, A. Lohle, R. Hubner,
M. B. Schilling, C. Shekhar, C. Felser, M. Dressel, and
A. V. Pronin, ArXiv e-prints
(2018), arXiv:1803.09708
[cond-mat.mes-hall].
43 I. Panfilov, A. A. Burkov, and D. A. Pesin, Phys. Rev. B
89, 245103 (2014).
44 J. Zhou, H.-R. Chang, and D. Xiao, Phys. Rev. B 91,
035114 (2015).
|
1105.5209 | 2 | 1105 | 2011-10-17T06:07:41 | Gapless interface states between topological insulators with opposite Dirac velocities | [
"cond-mat.mes-hall"
] | The Dirac cone on a surface of a topological insulator shows linear dispersion analogous to optics and its velocity depends on materials. We consider a junction of two topological insulators with different velocities, and calculate the reflectance and transmittance. We find that they reflect the backscattering-free nature of the helical surface states. When the two velocities have opposite signs, both transmission and reflection are prohibited for normal incidence, when a mirror symmetry normal to the junction is preserved. In this case we show that there necessarily exist gapless states at the interface between the two topological insulators. Their existence is protected by mirror symmetry, and they have characteristic dispersions depending on the symmetry of the system. | cond-mat.mes-hall | cond-mat |
Gapless interface states between topological insulators with opposite Dirac velocities
Ryuji Takahashi1 and Shuichi Murakami1, 2
1Department of Physics, Tokyo Institute of Technology,
2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan
2PRESTO, Japan Science and Technology Agency (JST), Kawaguchi, Saitama 332-0012, Japan
(Dated: November 14, 2018)
The Dirac cone on a surface of a topological insulator shows linear dispersion analogous to optics
and its velocity depends on materials. We consider a junction of two topological insulators with
different velocities, and calculate the reflectance and transmittance. We find that they reflect the
backscattering-free nature of the helical surface states. When the two velocities have opposite signs,
both transmission and reflection are prohibited for normal incidence, when a mirror symmetry
normal to the junction is preserved. In this case we show that there necessarily exist gapless states
at the interface between the two topological insulators. Their existence is protected by mirror
symmetry, and they have characteristic dispersions depending on the symmetry of the system.
PACS numbers: 73.20.-r, 73.40.-c,73.43.-f,75.70.Tj
Recently physical phenomena originating from the
Dirac cones of electrons have been studied, in the context
of graphene sheet [1] or the topological insulator (TI) [2 --
5]. In a graphene sheet, novel transport phenomena are
predicted theoretically in p-n junction systems: for ex-
ample the Klein paradox [6], and the negative refraction
[7]. The TI in three dimensions (3D) [4, 5], such as Bi2Se3
[8, 9] and Bi2Te3 [10], has a single Dirac cone in its sur-
face states, as observed by angle-resolved photoemission
spectroscopy. Unlike graphene, the states on the Dirac
cone on the surface of the TI are spin filtered; they have
fixed spin directions for each wave number k. Because
the state at k and that at −k have the opposite spins,
the perfect backscattering from k to −k is forbidden.
Such linear dispersion is similar to photons. The ve-
locity of the Dirac cone on the surface of 3D TI depends
on materials. For example, the velocity for Bi2Te3 is
about 4 × 105m/s [10] depending on the direction of the
wave vector, and that for Bi2Se3 is approximately 5 × 105
m/s [8]. Therefore, when two different TIs are attached
together, the refraction phenomenon similar to optics is
expected at the junction.
In this Letter, we theoreti-
cally study the refraction of electrons at the junction be-
tween the surfaces of two TIs [Fig. 1(a)]. The resulting
transmittance and reflectance are different from optics,
reflecting prohibited perfect backscattering. In addition,
we show that when the velocities of the two TIs have op-
posite signs, neither refraction nor reflection is allowed for
the incident electron normal to the junction. In this case,
we can show that there necessarily exist gapless interface
states between the two TIs and the incident surface elec-
trons totally go into the interface states. As long as the
mirror symmetry with respect to the yz plane Myz is
preserved, the interface gapless states exist. These gap-
less states are formed at the interface between the same
Z2 nontrivial materials. As a result, these interface states
do not come from the Z2 topological number, but come
from the mirror Chern number[11], and are protected by
the mirror symmetry Myz.
The effective Dirac Hamiltonian of the surface states
on the xz plane is represented as
H = −iv[σx∂z − σy∂x],
(1)
where σx, σy are the Pauli matrices, and v is the Fermi
velocity. From the Hamiltonian one can obtain the
linear energy E = svk where k = k, and s = +1(−1)
corresponds to the upper (lower) cone, provided v > 0.
We consider a refraction problem between the two TIs,
which we call TI1 and TI2, with the incidence angle θ,
the transmission angle θ′, and the reflection angle θR
[Fig. 1(a)]. As in optics, the momentum conservation
requires θR = θ, and the wave functions are written as
ψI (x, z) = 1√2
eik(x sin θ+z cos θ)(1, e−iθ)t, ψT (x, z) =
1√2
=
1√2
are
the wave numbers on TI1 and TI2, respectively, and we
consider the Fermi energy EF > 0 (i.e above the Dirac
point), giving s = +1 for both of the TIs. Let v1 and v2
denote the velocities of the two TIs.
eik′(x sin θ ′+z cos θ ′)(1, e−iθ ′
eik(x sin θ−z cos θ)(1, −eiθ)t, where
and k′
ψR(x, z)
)t,
k
2
1
sin θ = v−1
We first assume v1 and v2 to be positive. The con-
servation of the momentum and the energy yields Snell's
law: k′ sin θ′ = k sin θ, v−1
sin θ′. Let r and
t denote the amplitude of the reflected and transmitted
wave, compared with the incident wave. The current
conservation in the z direction is written as R + T = 1,
where R ≡ r2 and T ≡ v2 cos θ ′
v1 cos θ t2 are the reflectance
and the transmittance, respectively. We note that the
wavefunction should eventually be discontinuous at the
junction when the velocities are different, as has been
studied in the context of graphene [12, 13]. The reason
is the following. Therefore, the current conservation at
the interface requires v1zψ12 = v2zψ22, where v1,2 is
a velocity, and the subscripts 1 and 2 represent TI1 and
TI2, respectively. Because in our case v1z 6= v2z, we
have ψ12 6= ψ22 at the junction, and the continuity
2
FIG. 2:
(Color online) Transport at the junction between
the surfaces of two TIs, whose velocities have different signs.
(a) Linear dispersion at kx = 0. The incident wave (I) is
perpendicular to the junction. Both the transmission (T)
and reflection (R) are prohibited due to spin conservation.
(b) Normal incidence. TI1 (red) and TI2 (blue) have the
velocities of opposite signs. The purple region represents the
interface.
two TIs. These interface states arise from hybridization
between the two surface states from the two TIs. To show
the existence of gapless interface states, we first write
down the effective Hamiltonian at the interface from the
two Dirac cones with hybridization:
H = (cid:18)H1 V
V † H2(cid:19) .
(3)
FIG. 1: (Color online) (a) Schematic of the refraction of the
surface states at the junction between the two TIs, TI1 and
TI2. (b)(c): Reflectance (red) and transmittance (blue) for
the ratios of the velocities of the two TIs: (b)v2/v1 = 0.6
and (c)v2/v1 = 1.4. The solid curves are the results for the
junction between two TIs, while the dotted curves show the
results for optics with p and s polarizations.
of the wavefunction is violated. The proper way is to
set the Hamiltonian to be Hermitian also at the bound-
Here H1(2) is the effective surface Hamiltonian for the
surface of TI1 (TI2) at the interface:
ary, i.e. H = −i(cid:2) 1
where v(z) is the velocity dependent on z. The resulting
coefficients are
2 [v(z)σx∂z + σx∂zv(z)] − v(z)σy∂x(cid:3) ,
r = i
sin θ ′−θ
cos θ+θ ′
2
2
e−iθ,
t = r v1
v2
cos θ
cos θ+θ ′
2
ei θ′
−θ
2
.
(2)
They satisfy the current conservation. The results are
plotted as the solid curves in Figs. 1(b)(c). The dotted
curves represent corresponding results for optics. Unlike
optics, for normal incidence (θ = 0), the perfect transmis-
sion (T = 1, R = 0) occurs, which reflects the prohibited
backscattering on the surface of the TI. This is similar
to graphene [6, 12, 13] but the transmittance in our case
monotonically decreases with the incidence angle.
Next, we consider the case where the velocities of the
two TIs have opposite signs, where we can no longer use
the above approach. One might think that it is similar to
the negative refraction in optics [14, 15], but it is not true
because the Fermi energy is above the Dirac point for the
two TIs. Furthermore, both reflection and transmission
are prohibited for normal incidence, because the incident
wave has no way to conserve its momentum kx along the
interface and spin simultaneously (see Fig. 2). Thus it is
a paradox what happens for normal incidence.
Our answer to this question is that gapless states exist
at the interface between the two TIs (the purple region
in Fig. 2(b)). The normally incident wave goes along the
surface of one TI, then into the interface between the
H1 = v1(σ × k)z, H2 = −v2(σ × k)z
(4)
and V is the hybridization at the interface. For simplic-
ity, we retain only the lowest order in k. In the expression
of H2, there is an extra minus sign; on the surface of TI2
in Fig. 2, the mode going in the +z direction evolves
from that going in the −y direction, whereby the extra
sign necessarily appears.
We explain the reason for justifying our model
in
Eqs. (3), (4). For simplicity we assumed that the surface
states on the xz surface for TI1 and TI2 are described
by the Dirac cone. Generic surface states with non-Dirac
types are covered in the later discussion using the mirror
Chern number[11]. We used here the fact that the Dirac
velocities for each TI have the same signs for the xy and
xz surfaces. It is because the signs of the Dirac veloci-
ties are determined by the mirror Chern number which
is the bulk quantity [11]. We also set the Dirac cones to
be isotropic for simplicity; the following results turn out
to be unaltered by anisotropy in the Dirac cones. We
henceforth impose the mirror symmetry with respect to
the yz plane Myz, because this symmetry preserved by
H1 and H2 sets the spins parallel to the x axis for the nor-
mally (kz) incident wave. By imposing this mirror sym-
metry Myz and time-reversal symmetry, V is expressed
as V = (cid:18) g ih
ih g(cid:19) , where g and h are real constants repre-
senting the hybridization between the two surface states.
3
rotational symmetry around the z axis, and is lifted when
it is broken by adding higher order terms in k, e.g. the
warping term in Bi2Se3 [16]. As seen in Fig. 3(b), the
dispersion becomes a collection of Dirac cones. There-
fore, for this example Hamiltonian, we could show that
there are gapless states at the interface when the system
has the mirror symmetry Myz.
We note that this method is generic, because the anal-
ysis is based only on the symmetry. The only assumption
is that the gapless point is near k = 0, and we can expand
the Hamiltonian in terms of k. To complement this argu-
ment, we show the existence of gapless interface states on
generic grounds. Because these gapless states are gener-
ated between two TIs with the same Z2 topological num-
bers, they are not protected in the same sense as the
surface states of three-dimensional TIs. In the following
we show that these gapless interface states are protected
by the mirror symmetry and the time-reversal symme-
try. Each TI with mirror symmetry is characterized by
the mirror Chern number [11]. When the system has the
mirror symmetry Myz, the surface modes are labeled
with the mirror eigenvalues M = ±i at kx = 0, corre-
sponding to the spin along −x and +x, respectively. The
mirror Chern number is obtained as nM = (n+i − n−i)/2
where n±i are the Chern numbers [17, 18] for the sub-
space of states with mirror eigenvalues M = ±i. We
have n+i = −n−i by the time-reversal symmetry. In our
case where the two surface Dirac cones have opposite ve-
locities, the mirror Chern numbers for the two TIs are
different. TI1 has n(1)
−i = +1,
M
and TI2 has n(2)
−i = −1 at kx = 0
M
plane. For the M = +i (Sx < 0) subspace, this corre-
sponds to the junction of two systems with Chern num-
bers n(1)
+i = −2,
it gives rise to two left-going chiral modes in the y di-
rection. On the other hand, for M = −i it also gives
two right-going chiral modes in the y direction. These
modes are schematically shown in Fig. 3(d). Therefore
it is natural to generate the two Dirac cones in the junc-
tion. Thus these gapless states are protected by the mir-
ror symmetry. If the mirror symmetry is not preserved,
the gapless states do not exist in general. This discussion
is generic, and is complementary to our discussion by the
surface Dirac Hamiltonian. Therefore, we conclude that
the gapless interface states exist for the generic cases with
mirror symmetry, even with e.g. lattice mismatch at the
interface. In real materials, mirror symmetry may be lost
by disorder in principle; nevertheless, if the the sample is
relatively clean, the gapless interface states are expected
to survive and can be measured experimentally.
+i = +1; because n(1)
+i = −1 and n(2)
+i = −1, n(1)
+i = +1, n(2)
= −1, i.e. n(1)
= 1, i.e. n(2)
+i − n(2)
The distance between the Dirac cones of the gapless
interface states are proportional to the magnitude of the
hybridization between the two TIs at the interface. When
the hybridization becomes as strong as the bandwidth,
the spacing between the interface Dirac cones is of the
FIG. 3: (Color online) (a)(b) Dispersion on the interface be-
tween the two TIs in Eq. (5) with velocities v1 = 1, v2 = −2.
In (a), the hybridization is g = 2, h = 1. There are two
Dirac points (kx, ky) = (0, ±p5/2) where the gap closes. In
(b), the hybridization is g = 2, h = 0 and the warping term
λ(k3
−)σz with λ = 0.4 added to H1 and H2. There ap-
pear six Dirac cones. (c) Illustration of the interface mode.
The surface current goes into the interface. (d) Schematic of
the dispersion of interface states on kx = 0.
+ + k3
From the Hamiltonian [Eq. (3)], the eigenvalues are cal-
culated as
E = ±r∆k ±q∆2
k − η, ∆k = g2 + h2 +
1 + v2
v2
2
2
k2,(5)
(6)
η = v2
1v2
2k4 + ∆2
0 − 2v1v2k2(h2 cos 2α − g2),
where we set kx+iky = keiα, and α is real. The condition
for existence of gapless interface states is
v2
1v2
2k4 − 2v1v2k2(h2 cos 2α − g2) + (g2 + h2)2 = 0. (7)
To solve this equation, we note that g is nonzero, whereas
h can become zero when additional symmetries such as
rotational symmetry with respect to the z axis are im-
posed. Then we can see that for v1v2 > 0 (the two veloc-
ities with the same signs), the interface states are gapped
by the hybridization.
Only when two velocities have opposite signs (v1v2 <
0), are there gapless states on the interface. Disper-
sion of the gapless states depends on whether h 6= 0
or h = 0. When h 6= 0, the solutions are (kx, ky) =
(cid:16)0, ±p(g2 + h2)/v1v2(cid:17) and there are gapless states on
the interface. The interface states have two Dirac cones
(Fig. 3 (a)). On the other hand, when h = 0 due to
rotational symmetry with respect to the z axis, the gap
closing points form a circle k2
. This de-
generacy on the circle in k space is due to the continuous
y = g2
v1v2
x + k2
order of inverse of the lattice spacing. In that case the
transport properties will be like the graphene, having
two Dirac cones at K and K' points. We note that in
graphene there are spin-degenerate Dirac cones, whereas
in the present case the interface Dirac cones are not spin
degenerate. From Fig. 3(d), when the wave number k
goes around one of the Dirac point, the spin direction
also rotates around the z axis (normal to the interface).
In the similar way as in graphene, one can consider the
valley degree of freedom as a pseudospin, and develop val-
leytronics [19, 20] similar to graphene. These interface
states can be measured via transport; for this purpose
one should suppress the surface transport by attaching
ferromagnets on the surface.
From the spin-resolved angle-resolved photoemission
spectra, all the TIs observed so far, such as Bi1−xSbx [21,
22], Bi2Se3[9], and Bi2Te3[9], have nM = −1. To realize
the protected interface states in experiments discussed in
this Letter, one needs to find a TI with nM = +1, i.e., the
surface Dirac cone with negative velocity, and the spins
on the upper cone is in the counterclockwise direction in
the k space. It is an interesting issue to search for such
TIs. The Dirac velocity v is nothing but the coefficient
λ in the Rashba spin-splitting term λ(σ × k)z in the
Hamiltonian. The Rashba coefficient λ originates from
an integral of a sharply peaked function near the nuclei,
which rapidly varies between positive and negative values
[23, 24]. Therefore, we expect that it can change sign in
principle. The sign of the mirror Chern number nM is
also related with the mirror chirality of the bulk Dirac
Hamiltonian describing the bands near the bulk gap [11].
Because the mirror chirality governs the sign of the g-
factor which can be negative or positive as a result of the
spin-orbit coupling, one may well expect that in some
materials nM can become +1.
In conclusion, we study refraction phenomena on the
junction between the two TI surfaces with different veloc-
ities. The resulting reflectance and transmittance reflect
the backscattering-free nature of the surface states of TIs.
When the velocities of the TI surface states for the two
TIs have different signs, we show that the gapless states
appear on the interface. The existence of the gapless
states is shown by using the mirror Chern number, and
thus is topologically protected by the mirror symmetry.
4
The authors are grateful to T. Oguchi for discus-
sions. This research is supported in part by Grant-in-
Aid for Scientific Research (No. 21000004 and 22540327)
from the MEXT, Japan and by Kurata Grant from Ku-
rata Memorial Hitachi Science and Technology Founda-
tion. R.T also acknowledges the financial support from
the Global Center of Excellence Program by MEXT,
Japan through the "Nanoscience and Quantum Physics"
Project of Tokyo Institute of Technology.
[1] K. S. Novoselov et al., Science 306, 666 (2004).
[2] C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801
(2005); ibid. 95, 146802 (2005).
[3] B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. 96
(2006) 106802.
[4] L. Fu, C. L. Kane and E. J. Mele, Phys. Rev. Lett.
98,106803(2007).
[5] J. E. Moore and L. Balents, Phys. Rev. B 75, 121306(R)
(2007).
[6] M. I. Katsnelson et al., Nature Physics 2, 620 - 625
(2006).
[7] V. V. Cheianov et al., Science 315, 1252 (2007).
[8] Y. Xia et al., Nature Phys. 5, 398 (2009).
[9] D. Hsieh et al., Nature (London) 460, 1101 (2009).
[10] Y. L. Chen et al., Science 325, 178 (2009).
[11] J. C. Y. Teo, L. Fu, C. L. Kane, Phys. Rev. B 78, 045426
(2008).
[12] A. Raoux, M. Polini, R. Asgari, A. R. Hamilton, R. Fazio,
and A. H. MacDonald, Phys. Rev. B81, 073407 (2010)
[13] A. Concha and Z. Tesanovi´c Phys. Rev. B82, 033413
(2010)
[14] V. G. Veselago, Sov. Phys. Usp. 10, 509 (1968).
[15] J. B. Pendry, Phys. Rev. Lett. 85, 3966 (2000).
[16] L. Fu, Phys. Rev. Lett. 103, 266801 (2009)
[17] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M.
den Nijs, Phys. Rev. Lett. 49, 405 (1982).
[18] M. Kohmoto, Ann. Phys. N.Y. 160, 343 (1985).
[19] O. Gunawan et al., Phys. Rev. Lett. 97, 186404 (2006).
[20] D. Xiao, W. Yao, and Q. Niu, Phys. Rev. Lett. 99, 236809
(2007).
[21] D. Hsieh et al., Science 323, 919 (2009).
[22] A. Nishide et al., Phys. Rev. B 81, 041309(R) (2010).
[23] M. Nagano, A. Kodama, T. Shishidou and T. Oguchi, J.
Phys.: Conden. Matter 21, 064239 (2009).
[24] G. Bihlmayer et al., Surf. Sci. 600, 3888 (2006).
|
1608.06436 | 1 | 1608 | 2016-08-23T09:33:38 | Dynamic dependence to domain wall propagation through artificial spin ice | [
"cond-mat.mes-hall"
] | Domain wall propagation dynamics have been studied in nanostructured artificial kagome spin ice structures. A stripline circuit has been used to provide localised pulsed magnetic fields within the artificial spin ice structure. This provides control of the system through electrically assisted domain wall nucleation events. Synchronisation of the pulsed fields with additional global magnetic fields and the use of a focussed magneto-optical Kerr effect magnetometer allows our experiments to probe the domain wall transit through an extended ASI structure. We find that the propagation distance depends on the driving field revealing field driven properties of domain walls below their intrinsic nucleation field. | cond-mat.mes-hall | cond-mat | a
Dynamic dependence to domain wall propagation through artificial spin ice
Department of Physics, Imperial College London, London SW7 2BZ, United Kingdom
D.M. Burn, M. Chadha, and W.R. Branford
(Dated: November 10, 2018)
Domain wall propagation dynamics have been studied in nanostructured artificial kagome spin
ice structures. A stripline circuit has been used to provide localised pulsed magnetic fields within
the artificial spin ice structure. This provides control of the system through electrically assisted
domain wall nucleation events. Synchronisation of the pulsed fields with additional global magnetic
fields and the use of a focussed magneto-optical Kerr effect magnetometer allows our experiments
to probe the domain wall transit through an extended ASI structure. We find that the propagation
distance depends on the driving field revealing field driven properties of domain walls below their
intrinsic nucleation field.
Magnetic meta-materials such as artificial spin ice
show behaviour arising from complex geometrical struc-
turing in addition to the original material properties.1,2
In these systems it is the combination of magnetic charge
interactions and topological constraints determine the
magnetisation behaviour of the system. Artificial spin
ice structures consisting of arrays of magnetic nano-
bars provide a 2D analogue to explore frustrated mag-
netic phenomena.3,4 These systems are of fundamen-
tal scientific interest5 -- 12 and have even been identified
as potentials for novel neural network or processing
technologies.3,13,14
Magnetisation reversal in artificial spin ice structures
composed of interconnected magnetic bars can be de-
scribed by an ensemble of magnetic domain wall (DW)
processes. The creation, annihilation and propagation
of these DWs throughout the system leads to magneti-
sation reversal within the bars as well as the transport
of both magnetic and topological charges throughout the
system. The conservation of both magnetic and topo-
logical charge provides constraints on the creation and
annihilation of DWs in the system. This reveals the
physical significance of the finer details of the micromag-
netic DW structure such as its chirality or topological
makeup when the DW interacts with a complex magnetic
structure.8,9,15 -- 19
The majority of our understanding of the magnetisa-
tion behaviour in artificial spin ice systems is based on ex-
periments combining thermal and quasi-static magnetic
fields applied to the entire system.10,12,20 -- 22 The role of
DWs have been typically investigated based on their nat-
ural occurrence,23 when an applied field exceeds the nu-
cleation field which is typically lower at the edges of the
structures. This approach is therefore limited in that we
can only investigate the internal behaviour of the system
once a process related to the edge of the system takes
place.
In this investigation the localised injection of DWs
along the length of a lithographically patterned mi-
crostrip is employed.24,25 Here the pulsed field DW in-
jection technique allows control over the DW nucleation
location within the system leading to significant experi-
mental advantages as the DW nucleation process can be
separated from a global applied field. Firstly, this allows
the behaviour of DWs in the system to be investigated
in a wider field range, even at lower fields than their nu-
cleation field. Secondly, this allows the magnetisation
dynamics in the system to be explored. This is of great
interest in the artificial spin ice system and understand-
ing of the behaviour in a dynamic context is necessary
for any future technological applications.
Our understanding of the propagation path of a DW
through a series of vertex structures can be explained
through topological considerations. Figure 1 shows a
two-vertex section of an artificial spin ice structure with
arrows representing the magnetisation orientation in each
bar and with topological defects pinned to the edges of
the structure associated with both the DWs and the ver-
tices.
Figure 1 shows the evolution of a down-chirality DW
incident upon a vertex with initial magnetisation sat-
urated to the left. During the interaction the −1/2
topological defect initially belonging to the DW becomes
pinned on the upper edge of the vertex. The +1/2 from
the DW follows the edge of the structuring and pairs with
the −1/2 initially associated with the vertex. This new
defect pair corresponds to a DW which is able propagate
along the lower branch at the vertex. The similar +ve
magnetic charges of the initial DW and the vertex pro-
vide a repulsive force which means there is an energy bar-
rier associated with this process which can be overcome
through the application of an applied magnetic field.
In figure 1(b) the -ve charge of the second vertex now
provides an attractive force on the positively charged
DW. Here the topological charges on the lower side of
the nanobar are opposite and therefore unwind when they
FIG. 1. Simple model of DW progation and annihilation
based on topological constraints in multiple vertices.
meet. The DW annihilates resulting in the two-in one-out
state illustrated in figure 1(c) where just a −1/2 topo-
logical defect from the incident DW now remains pinned
at the vertex. In this model, the DW no longer exists in
the system.
Figure 1(d) shows how a new DW can be injected into
the lower horizontal nanobar based on the state shown
in figure 1(c). The lower edge of the vertex contains zero
topological charge which is separated to form two defects
of +1/2 and −1/2 respectively. The +1/2 forms a pair
with the pre-existing −1/2 defect on the upper edge of
the vertex and represents a DW which can propagate
along the nanobar whilst the remaining −1/2 defect re-
mains at the lower edge of the vertex. This process in-
volves the separation of two opposite magnetic charges
which gives rise to an energy barrier which needs to be
overcome to complete this process.
The series of interactions illustrated in figure 1 shows
how an initial DW with down-chirality can interact with
two verticies in an artificial spin ice structure resulting
in a down-chirality DW in a subsequent nanobar with
similar geometry. The repeat of this process is consis-
tent with reversal events following from one another. A
similar series of interactions would also take place for
up-chirality DW which would propagate along the upper
branch at the first vertex. This would be followed by an
equivalent annihilation event at the second vertex and
the availability to nucleate a further up-chirality DW in
the final horizontal nanobar. In all cases, a sizable mag-
netic field is required to supply the energy to overcome
the energy barriers associated with moving like charges
towards one another, and separating zero charge into a
positive and negative charge pair.
This current understanding is based on the system
which maintains a minimum energy micromagnetic spin
configuration.
In this quasi-static regime energy must
be supplied to overcome energy barriers associated with
the transitions which nucleate the DW in new nanobars
throughout the system.
In this work we consider the
higher energy processes involved with dynamic propagat-
ing DWs and show deviations from our understanding of
these processes in the quasi-static regime.
By varying the bias field that drives DW motion we
investigate the importance of DW dynamics during prop-
agation through an artificial spin ice structure with fields
applied along the armchair geometry. Combining lo-
calised DW injection with a MOKE measurements with a
localised magnetisation probe we also infer on the length-
scales of propagation through the system at these fields.
Critically, our experiments probe the DW propagation
behaviour in fields below the intrinsic DW nucleation
field for these structures.
Artificial spin ice structures consisting of intercon-
nected NiFe nanowires were fabricated in a kagome geom-
etry using electron-beam lithography and thermal evapo-
ration. The bars were 700 nm x 150 nm in dimension and
were 10 nm thick. Further details about the patterning
of the magnetic structures can be found elsewhere26. A
2
2 µm wide Cr(5 nm)/Au(50 nm) microstrip was added
in a second lithographic process and is shown in figure 2
along with the simulated field profile expected from the
microstrip.27
The magnetisation reversal in the system was investi-
gated using magneto-optical Kerr effect (MOKE) magne-
tometry in the longitudinal geometry. Here, a focussed
laser spot with a ∼ 10 µm elongated footprint provided
a localised probe of the magnetisation reversal as illus-
trated in figure 2. The combination of quasi-static and
pulsed magnetic fields, supplied from external coils and
from the microstrip respectively, were used to drive the
magnetisation reversal in the sample. The Kerr signal
was averaged over 50 field cycles in each measurement.
By introducing a 5◦ angular offset between ASI struc-
turing and the applied field direction the magnetisation
reversal associated with DW nucleation events and DW
propagation through the system can be distinguished by
the reversal field.26 Additionally, by varying the contri-
bution from the quasi-static and pulsed fields, behaviour
from quasi-static energy dependent magnetisation re-
versal processes and time-dependent magnetisation pro-
cesses have been investigated.24,28
Initially the magnetisation behaviour was investigated
with a 1 Hz sinusoidal quasi-static applied magnetic field
and is shown in figure 3(a). Two transitions in the mag-
netisation occur at two distinct and relatively sharp re-
versal fields despite averaging over 50 field cycles and over
multiple nanowires within the illuminated laser footprint.
The two steps indicate the combination of two reversal
processes occurring during the magnetisation reversal as-
sociated with the nucleation field of a DW from a vertex
and the field required for a pre-existing DW to propagate
through a vertex.26
By introducing additional pulsed magnetic fields, fig-
ures 3(b) and (c) show modified hysteresis loops where
the arrows indicate the triggering of the pulsed field
FIG. 2. Image of a interconnected array of nanowires with
a Au stripline for applying localised pulsed magnetic fields.
700 nm x 150 nm
within the quasi-static field cycle resulting in an increase
in the magnetisation. At this point, the combination of
pulsed and quasi-static fields locally overcome the rever-
sal field leading to the injection of magnetic DWs at the
stripline. The additional pulsed field induced reversal
results in a reduced magnetisation reversal at the lower
quasi-static field and the higher field quasi-static reversal
remains unchanged.
The magnetisation reversal combines multiple rever-
sal processes which can be modelled as a summation of
tanh functions.29 The lines in figure 3 show a model fit
to the data where each transition is parameterised by a
reversal field, the relative change in magnetisation and a
parameter representing the shape of that transition. The
quasi-static reversal fields are symmetric with increasing
and decreasing field and share fitting parameters.
The shape of the hysteresis loop representing the in-
crease in magnetisation at the lower reversal field also
shows some broadening when a pulsed field is present.
This could represent a modified reversal field due to
a partially-reversed magnetisation state following the
pulse. This will be discussed in more detail later.
The pulsed-field induced magnetisation reversal de-
pends on both the pulsed field voltage and the trigger-
ing point within the quasi-static field cycle. Figure 4(a)
shows the minimum pulse voltage required to result in
the additional pulsed-field induced magnetisation rever-
sal steps in figure 3 (b and c). This is plotted as a
function of the quasi-static field at the point of pulsed
field triggering which can be considered as a static bias
field on the timescales of the pulsed field. The linear de-
crease represents the contributions to the total field at
the stripline where an increase in pulsed field amplitude
from a greater pulsed voltage allows magnetisation rever-
sal to take place at a lower quasi-static field. A linear fit
3
to this data provides a calibration of the stripline which
produces 10.8 ± 0.3 Oe/V. For quasi-static fields greater
than 122 Oe the magnetisation reversal is driven purely
by the quasi-static field. Therefore the effect of the pulsed
field in this field regime cannot be distinguished.
Figure 4(a) also compares the difference in behaviour
when the laser spot is positioned 0, 10 and 20 µm away
from the stripline. All the points fall on the same line
indicating the reversal process at the stripline is not af-
fected by the measurement position. However, when
measurements are performed with greater separation,
magnetisation reversal is only observed when the pulsed
field triggering occurs at larger quasi-static fields. This
feature allows us to probe the motion of the DWs through
an extended region of the ASI system.
For all measurement positions the combination of
pulsed and quasi-static fields still result in magnetisation
reversal at the stripline. This reversal is due to the lo-
caslied injection of magnetic DWs at the stripline which
propagate along the nanobars reversing the magnetisa-
tion near the stripline. At greater distances from the
stripline only the quasi-static field drives the DW propa-
gation and the magnetisation reversal represents the dy-
namic behaviour of DWs in fields below their nucleation
field.
Pulse lengths of 150 ns and 20 ns had little influence
on the DW injection process so the quasi-static bias field
dependence on DW propagation has been further investi-
gated with 20 ns long, 7.5 V pulses which are sufficient to
inject DWs when biased with a field greater than 108 Oe.
Figure 4(b) shows the pulsed-field-induced magnetisation
reversal contribution as a function of the quasi-static bias
field for various measurement positions. This is found
from the ratio in magnetisation change from the pulsed
and quasi-static fields in the hysteresis loops.
n
o
i
t
a
s
i
t
e
n
g
a
m
d
e
s
i
l
a
m
r
o
N
(a)
(b)
(c)
1
0
-1
1
0
-1
1
0
-1
-250
-200
-150
-100
0
-50
50
Applied field (Oe)
100
150
200
250
)
V
(
d
e
l
i
f
d
e
s
u
P
l
9
8
7
6
d
e
c
u
d
n
i
l
d
e
i
f
d
e
s
u
P
l
)
%
(
l
a
s
r
e
v
e
r
n
o
i
t
a
s
i
t
e
n
g
a
m
1.0
0.5
0.0
100
Distance to stripline
0 m
5 m
10 m
20 m
Distance to stripline
0 m
5 m
10 m
20 m
105
110
115
120
Quasi-static field (Oe)
FIG. 3. MOKE hysteresis loops showing the magnetisation
reversal driven by (a) a quasi-static magnetic field and (b) and
(c) with the addition of a pulsed magnetic field with triggering
indicated by the arrows. The loops show the behaviour 9V
20ns pulses with the laser at the stripline. The fitted lines
show a model fit to the data.
FIG. 4. a) Minimum pulsed voltage required to lead to mag-
netisation reversal when pulses are triggered at different QS
Bias fields. b)Pulsed field induced magnetisation reversal as
a function of quasi-static bias field measured at various sepa-
rations from the DW injection stripline. Pulses are 20 ns and
7.5 V
i
(
s
a
b
c
i
t
)
e
O
a
t
s
-
i
s
a
u
Q
120
115
110
105
100
-20
Pulsed field induced
magnetisation reversal
(%)
0
10
20
30
40
50
60
70
80
-10
0
10
20
Laser position ( m)
FIG. 5. Pulsed field induced magnetisation reversal measured
as a function of quasi-static bias field and measurement posi-
tion from the stripline. 20 ns Pulses at 7.5 V
Measurements at the stripline location show a large
pulsed-field induced magnetisation change which is most
significant when a large quasi-static field is used to drive
the DW propagation. Here, the result represents DWs
which reverse the magnetisation in nanobars near where
they are nucleated.
When the quasi-static bias field is reduced below
115 Oe the magnetisation change decreases. This indi-
cates that the proportion of pulsed field induced reversal
events within the region probed by the laser spot is re-
duced. This can be explained by DWs which are not able
to propagate so far through the structure in the lower
fields.
Measurements at greater distances from the stripline
show magnetisation reversal driven purely by the quasi-
static field (see field profiles in figure 2). Here a lower
magnetisation change is found as injected DWs must
propagate through a greater number of nanobars and ver-
tices before reaching the probed region. This means that
DW pinning is more likely, but at high quasi-static bias
fields, DW propagation up to 15 µm is still observed.
The DW propagation distance through the structure is
more clearly shown in in figure 5 where the pulsed-field-
induced magnetisation reversal is plotted as a function of
the measurement position and the quasi-static field. Here
a strong reversal is centered around the position of the
stripline at 0 µm which becomes more significant with
greater quasi-static fields. Again, as the fields approach
120 Oe the pulsed-field-induced reversal becomes indis-
tinguishable from the quasi-static field driven reversal.
With the large quasi-static fields, the distance over
4
which the reversal can be detected is much greater than
for the smaller quasi-static fields resulting in a triangular
shape in figure 5. This shows how the propagation of
a DW through the artificial spin ice structure depends
on the driving field applied to the DW. With low fields
the propagation distance is limited as multiple nanobars
and interconnecting verticies are encountered and pro-
vide pinning sites. However, with greater fields the pin-
ning from these become less significant allowing the DW
propagation over a greater number of nanobars and ver-
tices.
The simplified model of the DW path illustrated in
figure 1 relies on the external driving field exceeding the
nucleation field for a DW in these structures. However,
our results reveal that DWs are able to propagate below
this field with a driving field dependence to their propa-
gation length through the system.
The propagation at a reduced field can be explained by
considering the energetics associated with a propagating
DW. When considering the annihilation process between
the DW and vertex in figure 1(b) and (c) the energy
associated with the DW can be used to assist with the
nucleation of the DW in figure 1(d). This additional
energy would mean less is required externally from the
driving field.
The propagation length dependence can also be ex-
plained in terms of the energetics of the dynamically
propagating DWs. With a greater quasi-static driving
field the DWs travel with a greater energy. This means
that they can encounter, and overcome a greater number
energy barriers associated with the verticies before be-
coming pinned. The lower energy DWs become pinned
after fewer verticies and therefore travel a reduced dis-
tance through the artificial spin ice structuring.
In conclusion, we have probed the magnetisation rever-
sal in artificial spin ice systems through focussed MOKE
magnetometry. The combination of pulsed and quasi-
static magnetic fields allow for the injection of DWs and
the study of their dynamic interactions with the geo-
metrical structuring for a range of DW driving fields.
Our results demonstrate control over the location of in-
jected DWs within artificial spin ice structures and how
DW propagation distance depends on the external driv-
ing field.
Existing quasi-static models based on the manipula-
tion of magnetic and topological charges throughout the
system do not predict a length dependence to the propa-
gation. Our results, probing DWs in the dynamic regime
with a range of driving fields suggest changes for DWs
arriving at vertices in a higher energy state.
1 C. Nisoli, R. Moessner and P. Schiffer. Reviews of Modern
Condensed Matter 25, 363201 (2013).
Physics 85, 1473 -- 1490 (2013).
2 L.J. Heyderman and R.L. Stamps. Journal of Physics:
3 R.F. Wang, C. Nisoli, R.S. Freitas, J. Li, W. McConville,
B.J. Cooley, M.S. Lund, N. Samarth, C. Leighton, V.H.
Crespi and P. Schiffer. Nature 439, 303 (2006).
4 E. Mengotti, L.J. Heyderman, A.F. Rodr´ıguez, F. Nolting,
R.V. Hugli and H.B. Braun. Nature Physics 7, 68 (2010).
5 S. Ladak, D.E. Read, G.K. Perkins, L.F. Cohen and W.R.
Branford. Nature Physics 6, 359 (2010).
6 R.V. Hugli, G. Duff, B. O'Conchuir, E. Mengotti, A.F.
Rodr´ıguez, F. Nolting, L.J. Heyderman and H.B. Braun.
Philosophical transactions A 370, 5767 (2012).
7 A. Farhan, P.M. Derlet, A. Kleibert, A. Balan, R.V.
Chopdekar, M. Wyss, J. Perron, A. Scholl, F. Nolting
and L.J. Heyderman. Physical Review Letters 111, 057204
(2013).
8 W.R. Branford, S. Ladak, D.E. Read, K. Zeissler and L.F.
Cohen. Science 335, 1597 -- 600 (2012).
9 N. Rougemaille, F. Montaigne, B. Canals, M. Hehn, H. Ri-
ahi, D. Lacour and J.C. Toussaint. New Journal of Physics
15, 10 (2013).
10 Yi Qi, T. Brintlinger and J. Cumings. Physical Review B
77, 094418 (2008).
11 S. Zhang, I. Gilbert, C. Nisoli, G-W. Chern, M.J. Erickson,
L. O'Brien, C. Leighton, P.E. Lammert, V.H. Crespi and
P. Schiffer. Nature 500, 553 (2013).
12 U.B. Arnalds, A. Farhan, R.V. Chopdekar, V. Kapaklis,
A. Balan, E.T. Papaioannou, M. Ahlberg, F. Nolting, L.J.
Heyderman and B. Hjorvarsson. Applied Physics Letters
101, 112404 (2012).
13 S. Ladak, D.E. Read, T. Tyliszczak, W.R. Branford and
L.F. Cohen. New Journal of Physics 13, 023023 (2011).
14 P.E. Lammert, Xianglin Ke, Jie Li, C. Nisoli, David M.
Garand, Vincent H. Crespi and P. Schiffer. Nature Physics
6, 786 -- 789 (2010).
15 S.A. Daunheimer, O. Petrova, O. Tchernyshyov and
J. Cumings. Physical Review Letters 107, 167201 (2011).
16 O. Tchernyshyov and G-W. Chern. Physical Review Letters
5
95, 197204 (2005).
17 A. Pushp, T. Phung, C. Rettner, B. P. Hughes, S-H. Yang,
L. Thomas and S.S.P. Parkin. Nature Physics 9, 505
(2013).
18 S.K. Walton, K. Zeissler, D.M. Burn, S. Ladak, D.E. Read,
T. Tyliszczak, L.F. Cohen and W.R. Branford. New Jour-
nal of Physics 17, 013054 (2015).
19 P. Mellado, O. Petrova, Y. Shen and O. Tchernyshyov.
Physical Review Letters 105, 187206 (2010).
20 F Montaigne, D. Lacour, I.A. Chioar, N. Rougemaille,
D Louis, S Mc Murtry, H. Riahi, B.S. Burgos, T.O. Mente,
A. Locatelli, B. Canals and M. Hehn. Scientific reports 4,
5702 (2014).
21 Z. Budrikis, J.P. Morgan, J. Akerman, A. Stein, P. Politi,
S. Langridge, C.H. Marrows and R.L. Stamps. Physical
Review Letters 109, 037203 (2012).
22 C. Nisoli, Jie Li, Xianglin Ke, D. Garand, P. Schiffer and
Vincent H. Crespi. Physical Review Letters 105, 047205
(2010).
23 S. Ladak, S.K. Walton, K. Zeissler, T. Tyliszczak, D.E.
Read, W.R. Branford and L.F. Cohen. New Journal of
Physics 14, 045010 (2012).
24 M. Hayashi, L. Thomas, Y. B. Bazaliy, C. Rettner,
R. Moriya, X. Jiang and S.S.P. Parkin. Physical Review
Letters 96, 197207 (2006).
25 A. Himeno, T. Ono, S. Nasu, K. Shigeto, K. Mibu and
T. Shinjo. Journal of Applied Physics 93, 8430 (2003).
26 D.M. Burn, M. Chadha and W.R. Branford. Physical Re-
view B 92, 214425 (2015).
27 T J Silva, C S Lee, T M Crawford and C T Rogers. Journal
of Applied Physics 85, 7849 -- 7862 (1999).
28 D.M. Burn, E. Arac and D. Atkinson. Physical Review B
88, 104422 (2013).
29 B.K. Middleton, M.M. Aziz and J.J. Miles. IEEE Trans-
actions on Magnetics 36, 404 (2000).
|
1708.02628 | 1 | 1708 | 2017-08-08T19:53:28 | What is the best planar cavity for maximizing coherent exciton-photon coupling | [
"cond-mat.mes-hall"
] | We compare alternative planar cavity structures for strong exciton$-$photon coupling, where the conventional distributed Bragg reflector (DBR) and three unconventional types of cavity mirrors$-$ air/GaAs DBR, Tamm $-$ plasmon mirror and sub$-$wavelength grating mirror. We design and optimize the planar cavities built with each type of mirror at one side or both sides for maximum vacuum field strength. We discuss the trade$-$off between performance and fabrication difficulty for each cavity structure. We show that cavities with sub$-$wavelength grating mirrors allow simultaneously strongest field and high cavity quality. The optimization principles and techniques developed in this work will guide the cavity design for research and applications of matter$-$light coupled semiconductors, especially new material systems that require greater flexibility in the choice of cavity materials and cavity fabrication procedures. | cond-mat.mes-hall | cond-mat | What is the best planar cavity for maximizing coherent exciton-photon coupling
Zhaorong Wang,1 Rahul Gogna,2 and Hui Deng3, a)
1)EECS Department, University of Michigan, Ann Arbor, MI 48109,
USA
2)Applied Physics Department, University of Michigan, Ann Arbor, MI 48109,
USA
3)Physics Department, University of Michigan, Ann Arbor, MI 48109,
USA
We compare alternative planar cavity structures for strong exciton-photon coupling,
where the conventional distributed Bragg reflector (DBR) and three unconventional
types of cavity mirrors – air/GaAs DBR, Tamm-plasmon mirror and sub-wavelength
grating mirror. We design and optimize the planar cavities built with each type of
mirror at one side or both sides for maximum vacuum field strength. We discuss
the trade-off between performance and fabrication difficulty for each cavity struc-
ture. We show that cavities with sub-wavelength grating mirrors allow simultaneously
strongest field and high cavity quality. The optimization principles and techniques
developed in this work will guide the cavity design for research and applications of
matter-light coupled semiconductors, especially new material systems that require
greater flexibility in the choice of cavity materials and cavity fabrication procedures.
7
1
0
2
g
u
A
8
]
l
l
a
h
-
s
e
m
.
t
a
m
-
d
n
o
c
[
1
v
8
2
6
2
0
.
8
0
7
1
:
v
i
X
r
a
a)Electronic mail: dengh@umich.edu
1
Strong coupling between semiconductor quantum well (QW) excitons and cavity photons
leads to new quasi-particles – microcavity polaritons. Since their discovery1, planar micro-
cavity polaritons have become a fruitful ground for research on fundamental cavity quantum
electrodynamics, macroscopic quantum coherence, and novel device applications2–4. Crucial
for polariton research is a cavity with a strong vacuum field strength E at the QW plane
and a high quality factor Q. A strong vacuum field leads to stronger exciton-photon cou-
pling and thus a larger vacuum Rabi splitting between the polariton modes. A high quality
factor leads to long lifetime and coherence time of the cavity photon and correspondingly
the polariton. They together enable robust polariton modes, thermodynamic formation
of quantum phases, and polariton lasers with lower density threshold at higher operating
temperatures3.
Given the importance of a strong cavity field, there has been a few decades of effort to
improve the cavity field strength. Most commonly used in polariton research are planar
Fabry-Perot (FP) cavities formed by two distributed Bragg reflectors (DBRs), each DBR
consisting of either epitaxially-grown, closely lattice-matched alloys or amorphous dielectric
layers. This type of cavity can reach Q of tens of thousands in III-As based systems5,6, but
only about a thousand for III-N7,8 and even lower for other material systems9,10. The field
penetrates many wavelengths into the DBR, hence the field strength is not optimal11,12.
To achieve a stronger field, different cavity structures have been developed with reduced
effective cavity length, including using a metal mirror13 and using Al-oxide14–16 or air/GaAs
DBRs13,17 with larger index contrast. These structures showed greater polariton splitting
but unfortunately worse Q compared to AlGaAs-based DBRs; polariton lasing have not been
possible in these structures. Recently, cavities using a high index-contrast sub-wavelength
grating (SWG)18 were demonstrated for polariton lasers19. Not only does SWG allow greater
design flexibility20,21 and compatibility with unconventional materials, it is only a fraction in
thickness compared to typical DBRs, making it promising for shorter effective cavity length.
Yet the demonstrated SWG-based polariton cavity was not optimized for strong cavity field.
Here we demonstrate how to optimize these different types of cavity for the strongest
field, and compare their polariton splitting, quality factors and practicality for fabrication.
In particular, we find SWG cavities may allow simultaneously stronger field and high Q.
The cavity field relevant to the exciton-photon coupling is the vacuum fluctuation field.
2
Its strength E is normalized to the zero-point photon energy by
Z ǫ(r)E(r)2dV =
1
2
ωc,
(1)
where ǫ is the permittivity, and ωc is the cavity resonance energy. The integral is evaluated
in the entire space with a spatial-dependent ǫ. For a planar cavity, we consider the field
being confined in the z-direction but unbounded in the x and y directions. The integration
is thus evaluated in a quantization volume (Lx, Ly, Lz), where Lx and Ly are set to be much
larger than the wavelength, and Lz is set at some cutoff point where the field has decayed
significantly (∼1%).
The maximum field strength Emax serves as our main figure of merit for cavities designed
for strong coupling. Specifically, for a QW placed at the field maximum, the vacuum Rabi
splitting Ω is directly proportional to d · Emax, for d the QW exciton dipole moment. We
note that the Emax is closely related to the mode volume commonly used to characterize
photonic crystals cavity22,23 as well as the effective cavity length Lc for planar cavities11.
We choose Emax over Lc because the vacuum Rabi splitting is determined by not only Lc
but also ǫ as Ω ∝ (ǫLc)−1/2,12 while ǫ can vary by an order of magnitude in different
cavities. Therefore the maximum vacuum field strength is the most unambiguous quantity
to characterize the cavity for strong coupling. Also, we focus on high-Q cavities but do
not optimize for Q. This is because for cavities designed to have sufficiently high Q, the
experimentally achievable Q depends mainly on practical constrains such as chemical purity
and structural integrity of the fabricated structure.
In the following, we optimize the maximum vacuum field strength Emax for different
types of planar cavities. Based on Eq. 1, the Emax is enhanced in cavities with (i) a tightly
confined field profile E(r) and (ii) materials of low refractive indices. A tightly confined
field profile effectively reduces the range of integration, enhancing the field maximum. For
a planar cavity, this means using the shortest possible cavity length (half wavelength) and
using mirrors with shorter field penetration length. Equally important, lower refractive
indices lead to higher E(r), especially in the cavity layer where most of the field resides.
Vacuum or air has the lowest refractive index. Although not applicable to crystalline QWs
that needs mechanical support and surface protection, air-cavities can greatly enhance field
strength for QWs formed by two-dimensional (2D) van der Waals materials.
We use transfer matrix method for calculations of DBR-based cavities, and rigorous
3
coupled-wave analysis (RCWA) for SWG-based cavities. We obtain the cavity resonance
from the reflection spectrum of the cavity, using the real-valued bulk permittivity ǫ∞ for the
QW layer. Then we compute complex field distribution at the resonance and normalize the
field using Eq. 1. To include the exciton-photon coupling, we use a linear dispersion for the
QW layer24 modeled by a Lorentz oscillator25:
ǫ(e) = n2(e) = ǫ∞ +
f q22
mǫ0Lz
1
e2
0 − e2 − iγe
,
(2)
where f is the oscillator strength per unit area, q and m are the charge and mass of the
electron respectively, Lz is the QW thickness, e0 is the exciton energy, and γ is the exciton
linewidth. For a fair comparison of the different types of cavities, we focus on 2D half-
wavelength cavities with a single QW placed at the field maximum.26 We use III-As systems
as an example, because polariton cavities of highest qualities are all based on III-As systems.
The cavity is made of AlxGaAs alloys where the different Al content gives different refraction
index, with AlAs the lowest (n = 3.02), GaAs the highest (n = 3.68), and AlxGaAs a linear
interpolation (n = 3.68 − 0.66x). For the QW, we consider either a 12 nm GaAs/AlAs QW
or a 7 nm InGaAs/GaAs QW. For simplicity and ease of comparison, we assume both to
have the same exciton energy of 1.550 eV (800 nm)19, oscillator strength of f = 6 × 10−4A−2
27 and γ = 0.8 meV.
As a benchmark, we first describe the performance of the most widely used, monolithic
DBR-DBR cavities. DBRs are typically made of multiple pairs of high- and low-index layers,
all with an optical path length (OPL) of λ/4, for λ the cavity resonance wavelength. High
reflectance is achieved by maximal constructive interference among multi-reflections from
the layer interfaces. Light decays in a distance inversely proportional to the refraction index
contrast of the DBR pair. So a high index contrast is preferred for the DBR pair. For a
monolithic III-As cavity with a GaAs QW, we use AlAs (n = 3.02) for the low-index layer
and Al0.15Ga0.85As (n = 3.58) for the high index layer, where 15% AlAs alloy is included
instead of pure GaAs to avoid absorption at the QW exciton resonance. We consider a
bottom DBR with 20 pairs on GaAs substrate, and a top DBR with 16.5 pairs, matching
the reflectance of the bottom DBR. An experimental cavity Q of a few thousands can be
readily achieved with this structure. More DBR pairs will lead to a higher Q, but will have
negligible effect on the field confinement or the vacuum Rabi splitting.
To show the enhancement of the cavity vacuum field strength by reducing cavity length,
4
8
7
6
5
4
3
V
e
m
SWG-SWG cavity
Tamm-DBR hybrid
Air DBR
both-side
optimized
SWG-DBR hybrid
Air DBR both-side
from Ref[17]
Air DBR one-side
/2-cavity
Conventional DBR Cavities
with different cavity lengths
8 -cavity
2
1
1.5
2
2.5
E V
LxLy
3
10-6
FIG. 1. Calculated vacuum Rabi splitting Ω and maximum vacuum field strength maxeE for
various optimized cavities. All cavities are based on III-As material system. Single QW is embedded
at the field maximum. The vacuum field strength is calculated at the cavity resonance energy with
QW dispersion turned off, while the Rabi splitting is measured from the two reflection dips when
the QW dispersion is turned on. The dash-dotted line is a linear fit through all points.
we vary the cavity layer OPL from λ/2 to 8λ. As shown in Fig. 1, Emax increases with
decreasing cavity length, and the polariton splitting scales linearly with Emax. This confirms
the Emax as an appropriate figure of merit for optimizing exciton-photon coupling. The
field distribution of the best-performing λ/2 cavity is shown in Fig. 2(a). The maximum
field strength is Emax = 1.89 × 10−6V/pLxLy at the QW, but the field extends many
wavelengths into the DBRs due to the relatively small index contrast (< 1.2 : 1) of the
DBR layers. The reflectance spectrum shows a polariton splitting of 4.54 meV (Fig. 2(b)).
This conventional cavity can be fabricated by mature epitaxial growth technology as a
monolithic crystalline structure with few impurities or defects. However, the requirement
of lattice-matched materials, with resultantly small index contrast, leads to limited field
confinement.
For the second type of cavity, we increase the index contrast of the DBR layer by replacing
the AlAs layer by air or vacuum (n = 1), which we call an air-DBR. It can be created via
selective wet etching13,17. It represents the the highest possible refraction index contrast for
a DBR. The GaAs QW still needs to be embedded inside an AlAs cavity (as opposed to an
air cavity) to avoid detrimental surface effects. Each material layer – both the cavity layer
and the Al0.15GaAs layers in the DBRs, need to be sufficiently thick for mechanical stability.
5
(c)
10-6
(a)
10-6
1.5
1
0.5
0
1600
1800
2000
2200
2400
2600
z (nm)
(b)
1
e
c
n
a
t
c
e
l
f
e
R
0.8
0.6
0.4
0.2
0
1.544
(e)
10-6
2.5
2
1.5
1
0.5
0
QW dispersion
turned on
QW dispersion
turned off
1.547
1.55
1.553
1.556
Energy (eV)
0
1000
2000
3000
z (nm)
2
1.5
1
0.5
0
1
(f)
e
c
n
a
t
c
e
l
f
e
R
0.8
0.6
0.4
0.2
0
1.544
(d)
1
e
c
n
a
t
c
e
l
f
e
R
0.8
0.6
0.4
0.2
QW dispersion
turned on
QW dispersion
turned off
0
500
1000
1500
2000
z (nm)
0
1.544
1.547
1.55
1.553
1.556
Energy (eV)
QW dispersion
turned on
QW dispersion
turned off
1.547
1.55
1.553
1.556
Energy (eV)
FIG. 2. Vacuum field profile and Rabi splitting for conventional DBR cavity (a-b) and air DBR
cavity (c-f). The color stripes represent different materials– red is Al0.15Ga0.85As, lighter red is the
lower index AlAs, white region is air, and the thin stripe at the field maximum is a GaAs QW.
All three structures sit on GaAs substrate, which is not shown in (a) and (c). The vacuum field
strength has a unit of Volt/(normalization length in meters).
The experimentally realized structure used a minimum OPL of 3λ/4 for each suspended
layer17, which we also adopt. Based on this, Fig. 2(c) shows the optimal structure we
obtained. The top mirror consists of 3 pairs of Al0.15Ga0.85As/air layers. The air layer has
λ/4 OPL, the Al0.15Ga0.85As layer has 3λ/4 OPL (∼ 168 nm). The bottom DBR is the
same as in the DBR-DBR cavity (Fig. 2(a)). The maximum field strength is increased to
Emax = 2.29 × 10−6V/pLxLy. The corresponding vacuum Rabi splitting is Ω = 5.51 meV,
21% larger than the conventional DBR cavity.
Further increase of the field confinement can be obtained by replacing the bottom DBR
also by an air-DBR, as shown in Fig. 2(e). An experimentally fabricated air-DBR cavity17
used a 7 nm In0.13GaAs QW embedded in a GaAs high-index λ-cavity. For this structure,
we calculated Emax = 2.47 × 10−6V/pLxLy and Ω = 6.32 meV (Fig. 1), which is 39%
improvement over the conventional DBR cavity.
The structure can be further optimized by using a cavity layer of low refraction index.
However, an AlAs-cavity is incompatible with the selective wet-etching required to create
the bottom air-DBR; instead a Ga-rich layer is needed. Hence, for the cavity, we use a
Al0.4Ga0.6As (nr = 3.33) layer of λ/2 OPL sandwiched between two Al0.15Ga0.85As (nr =
3.58) layers of λ/4 OPL, for a total OPL of λ. Similarly the 3/4λ-OPL high-index DBR
6
layer can also be replaced by such sandwich structure to improve the index contrast. The
wet etching selectivity of AlAs over Al0.4Ga0.6As has been shown to be 107.28 A 12 nm GaAs
QW is placed at the center of the cavity. We obtain Emax = 2.91 × 10−6V/pLxLy and
Ω = 7.02 meV, 55% larger than conventional DBR cavity. This structure represents the
best Rabi splitting in realistic DBR-based III-As cavities.
To obtain even tighter field confinement, we consider another two types of cavities where
the DBR is replaced by mirrors with shorter penetration depth – a metal mirror and a
sub-wavelength grating mirror. Metal mirror, with a typical penetration depth of less than
100 nm, can be used to form Tamm-plasmon polaritons29. The optimal metal-DBR cavity
we find consists of 45 nm thick gold layer as the top mirror (Fig. 3(a)), the same bottom
DBR as in the conventional DBR-DBR cavity is used, and a λ/2 AlAs cavity layer slightly
adjusted to compensate for the change of the reflection phase from π.
In this structure, the field decays rapidly in the gold layer, resulting in a much shorter
effective cavity length (Fig. 3(a)), therefore a much higher vacuum field strength. We obtain
30 and Ω = 6.51 meV, 43% larger than conventional DBR
Emax = 2.47 × 10−6V/pLxLy
cavity. This is an impressive improvement considering only one side of the cavity is replaced
by a metal mirror. The structure was used to create the first organic polariton31,32.
In
III-As cavities, Grossmann et al13 additionally replaced the bottom DBR with an air DBR
to further improve the vacuum Rabi splitting to about double that of the conventional DBR
cavity.33 However, the drawback of this structure is the low cavity Q due to metal loss. The
cavity linewidth is comparable to the Rabi splitting; the system remains in the intermediate
coupling regime rather than the strong coupling regime (Fig. 3(b))34. The large loss in metal
also makes it difficult to achieve quantum degeneracy or low threshold laser.
To maintain both a high cavity Q and short cavity length, we use a III-As subwavelength
grating (SWG) for the fourth type of cavity18,19. The SWG is a thin layer of high-index
dielectric grating suspended in air, acting as the high-reflective mirror (Fig. 4(a)). To avoid
non-zero order evanescent diffraction waves entering the cavity layer, we include a 3/4λ
air-gap between the SWG and the bottom structure. The λ/2 cavity layer and bottom
DBR are identical to the conventional DBR cavity as in Fig. 2(a). We obtain Emax =
2.23 × 10−6V/pLxLy and Ω = 5.37 meV, both 19% larger compared to the conventional
DBR cavity and comparable to the single-sided air-DBR cavity. The advantage of the SWG-
DBR cavity is its high cavity quality and simplicity in fabrication. The structure is first
7
(a)
10-6
2
1.5
1
0.5
(b)
1
e
c
n
a
t
c
e
l
f
e
R
0.8
0.6
0.4
0.2
0
1000
1200
1400
z (nm)
1600
1800
0
1.52
QW dispersion
turned off
QW dispersion
turned on
1.53
1.54
1.55
1.56
1.57
1.58
Energy (eV)
FIG. 3. Vacuum field profile (a) and Rabi splitting (b) for Tamm-plasmon type cavity. The
DBR structure is the same with previous DBR cavities. The excitation wave is launched from the
substrate to top (left to right). The gold colored region represents a 45nm gold layer. Its refractive
index is taken from Olmon, et al35 for evaporated gold at 800nm. The cavity Q is around 60 inferred
from the linewidth, mainly due to the loss in the metal. The polariton splitting is 6.51meV, 43%
larger than conventional DBR cavity.
fabricated as a high-quality, monolithic crystal by epitaxial growth, as with the conventional
DBR cavities. The SWG is then created by lithography and etching. Because the grating is
the only air-suspended layer created, SWG-DBR hybrid cavity is easier to make and more
robust mechanically than an air-DBR. Polariton lasing and excellent coherence properties
have been demonstrated in SWG-DBR cavities19,36,37, but not yet in cavities using air-DBR
or metal mirrors.
Tighter cavity field confinement can be achieved when replacing the bottom DBR also by
an SWG. As shown in Fig. 4 (b), a λ/2 AlAs cavity layer sandwiched by two Al0.15Ga0.85As
λ/4 layers is air-suspended between two identical SWGs. Because of the evanescent diffrac-
tion waves from the SWGs, the field at the QW is slightly non-uniform. We record an
average vacuum strength at the QW to be 2.92 × 10−6V/pLxLy. The vacuum Rabi split-
ting is Ω = 7.58 meV, 67% larger than that of the conventional DBR cavity.
The fabrication of such a double-SWG cavity is more challenging than the SWG-DBR
hybrid cavity, but within reach of established fabrication techniques. The top-SWG and the
rest of the structure can be fabricated separately at first, then they can be bonded together
by cold welding38 or stacked together using the micromanipulation technique applied to
making 3D photonic crystals39. Alternatively, one could make the SWGs from Si/SiO2
wafer, then do a wafer bonding with the III-V active layer40. Additional practical challenges
include heat dissipation and proper design of current injection path for electrical-pumped
polariton devices. Finally we note that the SWG-based cavity may be particularly useful
8
(a)
-200
0
200
400
600
800
)
m
n
(
z
1000
1200
0
10-6
(b)
2
1.5
1
0.5
-200
0
200
400
)
m
n
(
z
600
800
200
400
x (nm)
0
200
400
x (nm)
10-6
3.5
3
2.5
2
1.5
1
0.5
(c)
1
e
c
n
a
t
c
e
l
f
e
R
0.98
0.96
0.94
0.92
QW dispersion
turned on
QW dispersion
turned off
1.545
1.55
1.555
(d)
1
e
c
n
a
t
c
e
l
f
e
R
0.98
0.96
0.94
0.92
QW dispersion
turned on
QW dispersion
turned off
1.545
1.55
1.555
Energy (eV)
Energy (eV)
FIG. 4. Vacuum field profile and Rabi splitting for SWG-DBR hybrid cavity (a-b) and double
SWG cavity (c-d). The design parameters for the SWG are, thickness of grating 95.6nm, period
503nm, filling factor 0.3, using Al0.15GaAs(n = 3.58), with air gap 511nm (left) and 144nm (right)
between SWG and the cavity layer. Bottom DBR in the hybrid cavity is the same with conventional
DBR cavity used above. The middle resonance in the reflection spectrum (d) is due to the exciton
absorption of the first-order grating diffraction mode (evanescent in air but propagating in the
cavity layer).
for coupling to 2D materials, where similar procedures of integration are already necessary.
Additionally, it is possible to hold 2D material with lower index materials, further enhancing
the vacuum field.
In summary, we have investigated the performance of strong coupling for several types
of planar cavities. We use maximum vacuum field strength as the main figure of merit,
in place of the effective cavity length or mode volume. The Tamm-DBR cavity can pro-
vide ∼ 43% improvement in Rabi splitting with only one mirror replaced by a gold thin
film. Researchers using active media with a large oscillator strength may find this struc-
ture easy to fabricate and yet still sufficient for reaching strong-coupling. However, its low
quality factor due to metal loss limits the coherence of the exciton-photon coupling and
the resulting polariton modes. The air-DBR type cavities are shown to provide 21% (55%)
improvement over conventional DBR cavities if one mirror (both mirror) is replaced with
an air-DBR. SWG based cavities provide 19% (67%) improvement when one (two) SWG
is used, comparable to (better than) that of air-DBR cavities. In practice, among all al-
ternative structures to conventional DBR-DBR cavities, SWG-DBR cavities are so far the
only ones where high cavity quality factor and polariton lasing have been demonstrated
experimentally. Our findings here may guide the design of cavity systems for the research
9
and applications of semiconductor exciton-polaritons, especially for increasing their operat-
ing temperature, lowering the density threshold for quantum phase transition and polariton
lasing, and incorporating newer materials.
ZW, HD acknowledge the support by the National Science Foundation (NSF) under
Awards DMR 1150593 and the Air Force Office of Scientific Research under Awards FA9550-
15-1-0240. We thank Pavel Kwiecien for his open-source RCWA code used for calculations
in this work.
REFERENCES
1C. Weisbuch, M. Nishioka, a. Ishikawa, and Y. Arakawa, Physical Review Letters 69,
3314 (1992).
2G. Khitrova and H. M. Gibbs, Reviews of Modern Physics 71 (1999).
3H. Deng and Y. Yamamoto, Reviews of Modern Physics 82, 14891537 (2010).
4D. Sanvitto and S. Kna-Cohen, Nature Materials 15, 1061 (2016).
5S. Reitzenstein, C. Hofmann, A. Gorbunov, M. Strau, S. H. Kwon, C. Schneider, A. Lf-
fler, S. Hfling, M. Kamp, and A. Forchel, Applied Physics Letters 90, 251109 (2007),
http://dx.doi.org/10.1063/1.2749862.
6B. Nelsen, G. Liu, M. Steger, D. W. Snoke, R. Balili, K. West, and L. Pfeiffer, Phys. Rev.
X 3, 041015 (2013).
7G. Christmann, R. Butte, E. Feltin, J.-F. Carlin, and N. Grandjean, Applied Physics
Letters 93, 051102 (2008), 5.
8A. Das, J. Heo, M. Jankowski, W. Guo, L. Zhang, H. Deng, and P. Bhattacharya, Phys.
Rev. Lett. 107, 066405 (2011).
9S. Kena-Cohen and S. R. Forrest, Nat Photon 4, 371 (2010).
10X. Liu, T. Galfsky, Z. Sun, F. Xia,
and E.-c. Lin, Nature Photonics 9, 30 (2015),
arXiv:1406.4826.
11Y. Yamamoto, S. MacHida, and G. Bjork, Physical Review A 44, 657 (1991).
12T. Norris, J. Rhee, C. Sung, Y. Arakawa, M. Nishioka, and C. Weisbuch, Physical review.
B, Condensed matter 50, 14663 (1994).
13C. Grossmann, C. Coulson, G. Christmann, I. Farrer, H. E. Beere, D. a. Ritchie, and J. J.
Baumberg, Applied Physics Letters 98, 231105 (2011).
10
14T. R. Nelson, J. P. Prineas, G. Khitrova, H. M. Gibbs, J. D. Berger, E. K. Lindmark, J.-H.
Shin, H.-E. Shin, Y.-H. Lee, P. Tayebati, and L. Javniskis, Applied Physics Letters 69,
3031 (1996).
15L. A. Graham, Q. Deng, D. G. Deppe, and D. L. Huffaker, Applied Physics Letters 70,
814 (1997).
16a. R. Pratt, T. Takamori, and T. Kamijoh, Applied Physics Letters 74, 1869 (1999).
17J. Gessler, T. Steinl, A. Mika, J. Fischer, G. S¸ek, J. Misiewicz, S. Hofling, C. Schneider,
and M. Kamp, Applied Physics Letters 104, 081113 (2014).
18M. C. Huang, Y. Zhou, and C. J. Chang-Hasnain, Nature Photonics 1, 119122 (2007).
19B. Zhang, Z. Wang, S. Brodbeck, C. Schneider, M. Kamp, S. Hofling, and H. Deng, Light:
Science & Applications 3, e135 (2014).
20Z. Wang, B. Zhang, and H. Deng, Physical Review Letters 114, 1 (2015).
21B. Zhang, S. Brodbeck, Z. Wang, M. Kamp, C. Schneider, S. Hofling, and H. Deng,
Applied Physics Letters 106 (2015), 10.1063/1.4907606.
22J. S. Foresi, P. R. Villeneuve, J. Ferrera, E. R. Thoen, G. Steinmeyer, S. Fan, J. D.
Joannopoulos, L. C. Kimerling, H. I. Smith, and E. P. Ippen, Nature 390, 143 (1997).
23P. T. Kristensen, C. Van Vlack, and S. Hughes, Optics Letters 37, 1649 (2012).
24Y. Zhu, D. J. Gauthier, S. E. Morin, Q. Wu, H. J. Carmichael, and T. W. Mossberg,
Physical Review Letters 64, 2499 (1990).
25R. Houdr´e, R. P. Stanley, U. Oesterle, M. Ilegems, and C. Weisbuch, Physical Review B
49, 16761 (1994).
26In all cavities, multiple QWs can be used to enhance the polariton splitting.
27L. C. Andreani and A. Pasquarello, Physical Review B 42, 8928 (1990).
28E. Yablonovitch, T. Gmitter, J. P. Harbison, and R. Bhat, Applied Physics Letters 51,
2222 (1987).
29M. Kaliteevski, I. Iorsh, S. Brand, R. A. Abram, J. M. Chamberlain, A. V. Kavokin, and
I. A. Shelykh, Physical Review B - Condensed Matter and Materials Physics 76, 1 (2007).
30For strong dispersive media like metal, the energy density in field normalization equation 1
is modified by Eq.6.126 in Jackson, Classical Electrodynamics, 3rd edition, 2001.
31D. G. Lidzey, D. D. C. Bradley, M. S. Skolnick, T. Virgili, S. Walker, and D. M. Whittaker,
Nature 395, 53 (1998).
32D. G. Lidzey, D. D. C. Bradley, T. Virgili, A. Armitage, M. S. Skolnick, and S. Walker,
11
Phys. Rev. Lett. 82, 3316 (1999).
33We have not been able to reproduce the field profile or reflection spectrum shown in Ref.13
in our simulation and instead found the 45 nm thick gold layer as the optimal for a Tamm-
plasmon mirror.
34V. Savona, L. C. Andreani, P. Schwendimann, and A. Quattropani, Solid State Commu-
nications 93, 733 (1995).
35R. L. Olmon, B. Slovick, T. W. Johnson, D. Shelton, S.-h. Oh, G. D. Boreman, and M. B.
Raschke, Phys. Rev. B 86, 1 (2012).
36J. Fischer, S. Brodbeck, B. Zhang, Z. Wang, L. Worschech, H. Deng, M. Kamp, C. Schnei-
der, and S. Hfling, Applied Physics Letters 104 (2014), 10.1063/1.4866776.
37S. Kim, B. Zhang, Z. Wang, J. Fischer, S. Brodbeck, M. Kamp, C. Schneider, S. Hfling,
and H. Deng, Physical Review X 6, 011026 (2016).
38S. K´ena-Cohen, M. Davan¸co, and S. Forrest, Physical Review Letters 101, 1 (2008).
39K. Aoki, D. Guimard, M. Nishioka, M. Nomura, S. Iwamoto, and Y. Arakawa, Nature
Photonics 2, 688 (2008).
40C. Sciancalepore, B. Ben Bakir, C. Seassal, X. Letartre, J. Harduin, N. Olivier, J. Fedeli,
and P. Viktorovitch, IEEE Photonics Journal 4, 399 (2012).
12
|
1808.09157 | 1 | 1808 | 2018-08-28T07:56:22 | Theory of charge-spin conversion at oxide interfaces: The inverse spin-galvanic effect | [
"cond-mat.mes-hall"
] | We evaluate the non-equilibrium spin polarization induced by an applied electric field for a tight-binding model of electron states at oxides interfaces in LAO/STO heterostructures. By a combination of analytic and numerical approaches we investigate how the spin texture of the electron eigenstates due to the interplay of spin-orbit coupling and inversion asymmetry determines the sign of the induced spin polarization as a function of the chemical potential or band filling, both in the absence and presence of local disorder. With the latter, we find that the induced spin polarization evolves from a non monotonous behavior at zero temperature to a monotonous one at higher temperature. Our results may provide a sound framework for the interpretation of recent experiments. | cond-mat.mes-hall | cond-mat | a
Theory of charge-spin conversion at oxide interfaces:
The inverse spin-galvanic effect
Gotz Seibolda, Sergio Caprarab, and Roberto Raimondic
aInstitut fur Theoretische Physik, BTU Cottbus-Senftenberg, PBox 101344, 03013 Cottbus,
bDipartimento di Fisica Universit`a di Roma Sapienza, piazzale Aldo Moro 5, I-00185 Roma,
Germany
cDipartimento di Matematica e Fisica, Universit`a Roma Tre, Via della Vasca Navale 84, 00146
Italy
Rome, Italy
ABSTRACT
We evaluate the non-equilibrium spin polarization induced by an applied electric field for a tight-binding model of
electron states at oxides interfaces in LAO/STO heterostructures. By a combination of analytic and numerical
approaches we investigate how the spin texture of the electron eigenstates due to the interplay of spin-orbit
coupling and inversion asymmetry determines the sign of the induced spin polarization as a function of the
chemical potential or band filling, both in the absence and presence of local disorder. With the latter, we
find that the induced spin polarization evolves from a non monotonous behavior at zero temperature to a
monotonous one at higher temperature. Our results may provide a sound framework for the interpretation of
recent experiments.
Keywords: Spin-orbit coupling, spin-charge conversion, oxides interfaces
1. INTRODUCTION
It is well known that the breaking of the inversion symmetry leads to the so-called Rashba spin-orbit coupling
(SOC),1 -- 3 where polar and axial vectors transform similarly.4 Basically this allows for two major possibilities of
charge to spin conversion: The spin Hall (SH)5 and the inverse spin galvanic (ISG) effect,6, 7 as well as for their
Onsager reciprocal effects. While the SH effect converts an electrical current into a spin imbalance at the sample
edges via an induced perpendicular spin current, the ISG effect creates a bulk non-equilibrium spin polarization
by a flowing electrical current.6, 8 -- 11 The inverse SG effect corresponds then to the production of electrical
current via the pumping of spin polarization.12, 13 Both the SG12, 13 and the ISG14 -- 20 effects have been observed
in semiconductors. In the first case an electrical current is measured after pumping spin polarized light (SG)
whereas in the second case Faraday and Kerr spectroscopies measure the spin polarization induced by the applied
current. The SG effect has also been very effectively measured by spin pumping from an adjacent ferromagnet
into a metallic interface,21 into a topological insulator surface23, 24 and more recently into the two-dimensional
electron gas (2DEG) in oxide LAO/STO heterostructures.25 -- 28 These latter materials have emerged29 -- 33 as
very promising materials for the SG and ISG effect, due to the large values of the Rashba SOC parameter α
as experimentally observed34 -- 37 and also theoretically calculated,38 -- 40 even though it is likely that, due to their
complex band structure, the available theory41 of the SG/ISG effect developed for the 2DEG in semiconductors
may not be able to capture a number of specific features. A first step in this direction has been made recently
by a combination of analytical diagrammatic and numerical approaches.42, 43
The layout of the paper is the following. In the next section we introduce a model for the electron states
relevant for describing transport at oxide LAO/STO interfaces. In section 3 we provide the necessary formalism
of linear response theory for the SG and ISG effects. In section 4 we introduce an approximate effective model
for electron states close to band minima. In section 5 we evaluate analytically the SG response for the effective
Further author information: (Send correspondence to Roberto Raimondi.)
Roberto Raimondi: E-mail: roberto.raimondi@uniroma3.it
H0 =
εxy
k
0
0
0
εxz
k
0
0
0
εyz
k
(1)
model, whereas in section 6 we introduce disorder and the necessary formalism to handle it. Finally in section
7 we present a fully numerical approach which includes both cases without and with disorder. We conclude in
section 8. A number of technical details are provided in the appendices.
2. THE MODEL
The electronic structure of the 2DEG at LAO/STO interfaces, perpendicular to the (001) crystal direction, is
usually described39, 44, 45 within a tight-binding Hamiltonian H0 for the Ti t2g orbitals, dxy, dxz,dyz, supplemented
by local atomic spin-orbit interactions with Hamiltonian Haso and an interorbital hopping with Hamiltonian HI
which is induced by the interface asymmetry.
The hopping between the d orbitals of two neighbouring cubic cells is mediated via intermediate jumps to p
orbitals. For instance, the hopping between two dxy orbitals along the x axis occurs via two successive hopping
dxy → py and py → dxy. In the first hop, the overlap, which is of order ∼ tpd, yields a positive sign, whereas
the sign is negative ∼ −tpd in the second one. Hence the effective dxy − dxy hopping goes like −t2
pd/∆E, with
∆E being the energy difference between d and p orbitals. As a result, in the basis xy(cid:105),xz(cid:105),yz(cid:105) the hopping
between similar orbitals reads as
with, setting to unity the lattice spacing,
εxy
k = −2t1[cos(kx) + cos(ky) − 2] − 4t3[cos(kx) cos(ky) − 1]
εxz
k = −2(t1 + t3)[cos(kx) − 1] − 2t2[cos(ky) − 1] + ∆
εyz
k = −2(t1 + t3)[cos(ky) − 1] − 2t2[cos(kx) − 1] + ∆
where the energy difference ∆ between the xy(cid:105) and xz(cid:105),yz(cid:105) states is due to the confinement of the 2DEG in
the xy-plane.46
The atomic SOC is given by
with τ i denoting the Pauli matrices.
Haso = ∆aso
0
iτ x
−iτ y
−iτ x
0
iτ z
iτ y
−iτ z
0
(2)
(cid:48)
Figure 1. At the interface an orbital polarization and (or) orbital displacement results in hopping processes ∼ t
pd as
between px- and zx-orbitals along the y-direction. The asymmetry is visualized by a small shift of px-orbitals along the
z-direction.
Hopping between different d orbitals may occur if inversion symmetry is broken, see Fig. 1. Consider, for
instance, the two hops along the y direction, dxy → px and px → dzx. While the first hop ∼ tpd is as the first
hop of the effective hopping between two dxy orbitals discussed above, the second hop ∼ t(cid:48)
pd will be forbidden in
the presence of inversion symmetry. To see this consider that
tpd = (cid:104)px, R +
a
2
yHdxy, R(cid:105),
where R is the lattice site of the dxy orbital. In the same way
t(cid:48)
pd = (cid:104)dzx, R + ayHpx, R +
a
2
y(cid:105).
In both cases, H is the full Hamiltonian. If H is invariant with respect to the inversion z → −z, then necessarily
t(cid:48)
pd = 0, because px is even, while dzx is odd. Clearly if H has terms which are not invariant for z → −z, then
t(cid:48)
pd (cid:54)= 0. As a result the interface asymmetry hopping reads44
(3)
HI = γ
0
2i sin(ky)
2i sin(kx)
−2i sin(ky) −2i sin(kx)
0
0
0
0
.
In the following we use the parameters, t1 = 0.277 eV, t2 = 0.031 eV, t3 = 0.076 eV, ∆ = 0.4 eV, ∆aso = 0.010 eV,
γ = 0.02 eV, which have been derived in Ref.39 from projecting DFT on the t2g Wannier states. Note that for
the splitting ∆ we take a value intermediate between the theoretical (∆ = 0.19 eV) and the experimental one
(∆ = 0.6 eV). The left panel of Fig. 2 shows the band dispersions along the x axis for these values of the
parameters. The bands come naturally in three pairs, which are split by the combined effect of the spin-orbit
coupling and the inversion symmetry breaking. For our analysis we have selected three different values of the
chemical potential for corresponding filling regimes. For µ = 0.3 eV, only the lowest pair of bands (1,2) is
occupied. The chemical potential µ = 0.425 eV is close to the Lifshitz point, where the spin-orbit splitting is
large and the pairs of bands (3,4) and (5,6) start to be filled. Finally, the chemical potential µ = 0.7 eV is in the
regime, where all pairs of bands (1,2), (3,4) and (5,6) are occupied.
We now analyze the chirality for each eigenstate band p = 1, . . . , 6 by computing the spin at each momentum
point of the Fermi surface (FS) according to
Sα(p, kF ) = (cid:88)n,σ,σ(cid:48)
Φ∗n,σ(p, kF )τ α
σ,σ(cid:48)Φn,σ(cid:48)(p, kF )
where Φn,σ(cid:48)(p, kF ) are the eigenfunctions of the system at momentum kF . The indices n = xy, xz, yz and σ label
the orbital and its spin. Then the chirality of the p-th band can be obtained from
kFS(p, kF (cid:19) .
α(p) = arcsin(cid:18) kF × S(p, kF ) · ez
Fig. 2 shows the chiralities for each pair of bands at selected chemical potentials and the corresponding FSs. For
the lowest pair of bands (1,2) the momentum dependent spin pattern displays a vortex-type structure with the
core centered at Γ = (0, 0). Thus, even when the FS changes from electron- to hole-like between µ = 0.5 and
µ = 0.6, the corresponding chiralities are always confined to α(p) ≈ ±π/2 without any sign change in α. For
the middle pair of bands (3,4) the spin structure is composed of two vortex patterns (with the same vorticity)
centered at (π, 0) and (0, π). As a consequence, the spin texture vanishes along the diagonals and a Rashba-type
description along this direction fails. In section 4 we will come back to this point. However, for small µ and
all other momenta the chirality also starts at α ≈ ±π/2 but then on average becomes smaller with increasing
chemical potential and eventually changes sign for µ ≈ 1.5. An analogous situation occurs for the uppermost
pair of bands where the 'spin-vortex core' is centered at (π, π). In this case the chiralities also change sign upon
increasing the chemical potential while at small µ one again recovers α ≈ ±π/2.
3. LINEAR RESPONSE THEORY
In this paper we aim at evaluating the spin polarization induced by an externally applied electric field. To be
definite we take the electric field along the x axis and the spin polarization along the y axis. To linear order in
the applied field we write the spin polarization as
sy(ω) = σISG(ω) Ex(ω),
(4)
Figure 2. Left panel: Structure of the t2g interface bands. The inset enlarges the region around the 'Lifshitz' point where
the spin-orbit splitting is large. The horizontal dashed lines in the main panel refer to three values of chemical potential:
µ = 0.3 eV (blue line), µ = 0.425 eV (red line) and µ = 0.7 eV (green line). The right panel displays the spin texture
for the three pairs of bands together with their Fermi surfaces. For the lower pair of bands (1,2) the spins point in the
opposite direction.
where σISG, the "conductivity" for the ISG effect, can be obtained by the zero-momentum limit of the Fourier
transform Ryx(ω) of the response function (henceforth the symbols in capital letters indicate the operators for
spin density and charge current) defined as
where the brackets stand for the quantum-statistical average and θ (t) is the Heaviside step function. The
frequency-dependent ISG conductivity reads
Ryx(t, r) = −ıθ(t)(cid:104)[Sy(t, r), Jx](cid:105) ,
(5)
σISG(ω) = lim
η→0+ (cid:60)(cid:20) Ryx(ω)
ı(ω + ıη)(cid:21) = −πδ(ω)R(cid:48)
yx(0) + P
R(cid:48)(cid:48)
yx(ω)
ω
≡ DISGδ (ω) + P
R(cid:48)(cid:48)
yx(ω)
ω
,
(6)
where the first term will be referred to as the Drude singular term and the second as the regular term, in analogy
with the terminology used in the case of the optical conductivity. Because under time reversal both the charge
current and the spin polarization are odd, according to the Onsager relation, the SG and ISG conductivities are
equal.41 For this reason we will use the term SG conductivity (SGC) for both direct and inverse effects. The
calligraphic symbol P stands for the principal part. The real R(cid:48)
yx(ω) parts of the
yx(ω) and the imaginary R(cid:48)(cid:48)
response function are related by the Kramers-Kronig relation (KKR)
∞
−∞
d ω(cid:48) R(cid:48)(cid:48)
yx(ω(cid:48))
ω(cid:48) − ω
.
∞
d ω σISG(ω) = 0
(7)
(8)
By integration over the frequency, thanks to the KKR, the SGC satisfies the following sum rule
R(cid:48)
yx(ω) =
1
π
due to the fact that for the SGC there is no 'diamagnetic' contribution as opposed to the optical conductivity.
In the following we are going to apply the above formulae to the model introduced in section 2. To this end,
it is instructive to consider first the case of the Rashba SOC for a 2DEG with quadratic dispersion relation in
−∞
-π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx kxkydkx~ dSyEx-π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx -π-π/20π/2π-π-π/20π/2πky kx a)b)c)Band 1Band 2Band 3Band 4Band 5Band 65FIG.5:Leftpanel:Structureofthet2ginterfacebands.Theinsetenlargestheregionaroundthe'Lifshitz'pointwherethespin-orbitsplittingislarge.TheverticallinesinthemainpanelrefertothechemicalpotentialsforwhichtheFermisurfaceisshowninFig.1.TherightpanelshowdisplaysthechiralspinstructurefortheupperofthethreepairofbandstogetherwiththeFS.Forthelowerbandsthespinspointintheoppositedirection.PERTURBATIVECALCULATIONAroundthe pointthenon-interactingpartofthehamiltonianreads"xyk=(t1+2t3)k2(12)"xzk=(t1+t3)k2x+t2k2y+ (13)"yzk=(t1+t3)k2y+t2k2x+ .(14)Theatomicspin-orbitcoupling(⇠⌧z,cf.Eq.(2))liftsthedegeneracybetweenxzandyzbutleavesthespindegeneracy,cf.Fig.7.OneobtainstheneweigenvaluesE± = +12(t1+t2+t3)k2±12q(t1+t3 t2)2(k2x k2y)2+4 2asoandeigenfunctionxz,"i=ia+,"i+ib ,"i(15)yz,"i= b+,"i+a ,"i(16)xz,#i=ia+,#i+ib ,#i(17)yz,#i=b+,#i a ,#i(18)witha=1p2r1+"xz "yz("xz "yz)2+4 2aso(19)⇡1p21+14 aso(t1+t3 t2)(k2x k2y) b=1p2r1 "xz "yz("xz "yz)2+4 2aso(20)⇡1p21 14 aso(t1+t3 t2)(k2x k2y) .InthisnewbasistheasymmetryhoppingisgivenbyHI=p2 (ky+ikx)xy,"ih+,"+h.c.+p2 (ky ikx)xy,#ih+,#+h.c.+p2 (ky ikx)xy,"ih ,"+h.c.+p2 (ky+ikx)xy,#ih ,#+h.c.andtheresidualcouplingofHasowiththexy-levelreadsHaso=p2 asoxy,"ih+,#+h.c.+p2 asoxy,#ih+,"+h.c. 12p2(t1+t3 t2)(k2x k2y)xy,"ih ,#+h.c. 12p2(t1+t3 t2)(k2x k2y)xy,#ih ,"+h.c..Inthefollowingwerestrictontheregionclosetothe -pointwhere aso>tk2andneglectthereforethetwoEk[eV](kx,0)the effective mass approximation. The insight gained in this simpler case will guide us also in the analysis of the
model with a complex band structure. We consider then the Rashba-Bychkov Hamiltonian3
H =
p2
2m
+ α(τ xpy − τ ypx),
(9)
where m is the effective mass and α the SOC. The 2DEG is confined to the xy plane and px and py are the
momentum operators along the two coordinate axes. Clearly there are two eigenvalues E±(p) = p2/2m± αp with
the corresponding eigenstates of (9) being plane waves whose spin quantization axis is fixed by the momentum
direction
p, s(cid:105) =
1
√2(cid:18) s ıe−ıθ
1
(cid:19) , s = ±1
where tan(θ) = py/px. The ISG response function at finite frequency and momentum reads
Ryx(ω, q) = (cid:88)p,s1,s2
(cid:104)p, s1Syp, s2(cid:105)(cid:104)p, s2Jxp, s1(cid:105)
f (Es1 (p) − µ) − f (Es2(p + q) − µ)
ω + ıη + Es1(p) − Es2 (p + q)
,
(10)
(11)
where f (E) is the Fermi distribution function at temperature T . Depending on the values of the spin indices,
one has intraband (s1 = s2 = ±1) and interband (s1 = −s2 = ±1) contributions. In the dynamic limit, when
the momentum goes to zero at finite frequency, the intraband contribution vanishes. For the model of Eq. (9)
the interband matrix elements for spin density Sy = τ y/2 and charge current Jx = (−e)(px/m − ατ y) read
(cid:104)p, sSyp,−s(cid:105) =
(cid:104)p,−sJxp, s(cid:105) = (−e)α(ıs) sin(θ),
(−ıs) sin(θ),
1
2
and the zero-momentum response function becomes
Ryx(ω) =
1
2
(−e)α(cid:88)ps
sin2(θ)
f (Es(p) − µ) − f (E−s(p) − µ)
ω + ıη + Es(p) − E−s(p)
.
(12)
At zero temperature, there are two FSs corresponding to the two spin helicity bands with Fermi momenta
p± = (cid:112)2mµ + (mα)2 ∓ mα. The evaluation of the imaginary part of the zero-momentum response function
leads to (η → 0+)
R(cid:48)(cid:48)
yx (ω) =
e
32α
ω [θ (ω − 2αp−) − θ (ω − 2αp−)] ,
(13)
showing an antisymmetric behavior with respect to the frequency ω. The spectral weight, at positive frequency,
is confined in the range 2αp+ < ω < 2αp−. The two frequencies delimiting the interval are nothing but the spin-
orbit splitting at the two Fermi surfaces. We note, and this will turn out useful when discussing the numerical
calculations, that at finite η, the imaginary part remains finite and acquires a linear-in-frequency behavior around
the origin, whose slope vanishes as η. The Drude weight, according to Eq. (6) can be easily obtained by the KKR
relation (7) to read
DISG = −
π
2
eN0α,
(14)
where N0 = m/(2π) is the single-particle density of states of the 2DEG. For the sake of simplicity we have
chosen units such = 1. There are two features worth noticing. The first is that the Drude weight is controlled
by the sign of the SOC. The second is that the Drude weight arises from the interband transitions between the
spin-orbit split bands. This must be compared with the case of optical conductivity for the electron gas, where
the Drude weight arises from the diamagnetic contribution to the current. In the present case, due to the sum
rule (8), the Drude low-frequency peak yields information about the spectral weight of interband transitions at
finite frequency. To the best of our knowledge this feature has not been noticed before.
In the following of the paper we will consider the effect of disorder, but it is instructive to make here an
heuristic discussion. In the presence of spin-independent disorder, due to the form (10) of the eigenstates, the
electron spin acquires a finite relaxation rate τ−1
. This mechanism, which is known as the Dyakonov-Perel
relaxation, arises because, at each scattering event, the change in momentum also affects the spin eigenstate. As
a result, in the diffusive approximation, ωτ (cid:28) 1, the spin density obeys a Bloch equation51
s
dsy
dt
= −
1
τs
(sy − s0),
(15)
where s0 = −eαN0τ E represents the steady-state nonequilibrium spin polarization8 induced by an applied
electric field E along the x axis and τ is the momentum relaxation scattering time (not to be confused with the
Pauli matrices τ i). According to Ref.51 the Dyakonov-Perel relaxation rate reads
By Fourier transforming (15) to frequency ω, one obtains the SGC in the form
1
τs
=
1
2τ
4α2p2
F τ 2
F τ 2 .
1 + 4α2p2
σISG(ω) = −eαN0
τ
τs
τ−1
s
ω2 + τ−2
s
,
(16)
(17)
which has a Lorentzian lineshape and evolves to a singular contribution in the weak scattering limit τ → ∞.
More precisely by integrating over frequency one obtains
∞
−∞
dωσISG(ω) = −eαN0π
τ
τs
π
2
= −
eαN0,
(18)
which reproduces the Drude weight of Eq. (14). Notice that in the last step we made use of the fact that the
spin relaxation time becomes twice the momentum relaxation time in the weak scattering limit according to
Eq. (16). Eq. (18) seems to violate the sum rule (8), but this is not the case. The form (17) for the SGC has
been derived in the diffusive approximation, which is valid for frequencies ω (cid:28) τ−1 well below the region of the
interband spectral weight. Hence, the form (17) captures only the low frequency spectral weight, which evolves
in the singular Drude weight in the limit of vanishing disorder. The effect of disorder is then to eliminate the
Drude singular contribution and to yield a finite SGC at zero frequency, which is the result of a finite slope of
the imaginary part of the response function. The microscopic approach in the presence of disorder is discussed
in section 6 and details about the frequency dependence are developed in the appendix B.
Around the Γ point the non-interacting part of the Hamiltonian (1) reads
4. EFFECTIVE MODELS
εxy
k = (t1 + 2t3)k2
k = (t1 + t3)k2
εxz
εyz
k = (t1 + t3)k2
x + t2k2
y + t2k2
y + ∆
x + ∆ ,
(19)
(20)
(21)
x + k2
y. The atomic SOC [∼ τ z, cf. Eq. (2)] lifts the degeneracy between xz and yz but leaves
where k2 = k2
the spin degeneracy, cf. Fig. 3. One obtains the new −(+), corresponding to the pairs of bands (3,4) and (5,6)
respectively,
E±σ = ∆ +
(t1 + t2 + t3)k2
1
2
±
1
2(cid:113)(t1 + t3 − t2)2(k2
x − k2
y)2 + 4∆2
aso
and eigenfunctions
with
1
√2(cid:114)1 +
a ≈
xz,↑(cid:105) = ıa+,↑(cid:105) + ıb−,↑(cid:105)
yz,↑(cid:105) = −b+,↑(cid:105) + a−,↑(cid:105)
xz,↓(cid:105) = ıa+,↓(cid:105) + ıb−,↓(cid:105)
yz,↓(cid:105) = b+,↓(cid:105) − a−,↓(cid:105)
(22)
(23)
(24)
(25)
εxz − εyz
(εxz − εyz)2 + 4∆2
1
√2(cid:18)1 +
1
4∆aso
aso ≈
(t1 + t3 − t2)(k2
x − k2
y)(cid:19) , b =(cid:112)1 − a2
In the basis +,↑(cid:105),+,↓(cid:105),−,↑(cid:105),−,↓(cid:105) the asymmetry hopping is given by
HI = √2γ(ky + ıkx)xy,↑(cid:105)(cid:104)+,↑ + h.c.
+ √2γ(ky − ıkx)xy,↓(cid:105)(cid:104)+,↓ + h.c.
+ √2γ(ky − ıkx)xy,↑(cid:105)(cid:104)−,↑ + h.c.
+ √2γ(ky + ıkx)xy,↓(cid:105)(cid:104)−,↓ + h.c.
and the residual coupling of Haso with the xy-level reads
Haso = √2∆asoxy,↑(cid:105)(cid:104)+,↓ + h.c.
+ √2∆asoxy,↓(cid:105)(cid:104)+,↑ + h.c.
x − k2
−
x − k2
(t1 + t3 − t2)(k2
(t1 + t3 − t2)(k2
1
2√2
1
2√2
−
y)xy,↑(cid:105)(cid:104)−,↓ + h.c.
y)xy,↓(cid:105)(cid:104)−,↑ + h.c. .
In the following we restrict to the region close to the Γ point, where ∆aso > tk2, and neglect therefore the two
latter terms in Haso resulting in the effective coupling structure depicted in panel (b) of Fig. 3. We can now
calculate the effective interactions between levels α,β in 2nd order perturbation theory
(cid:104)αH (2)β(cid:105) = −
1
2(cid:88)n (cid:18)
1
En − Eα
+
1
En − Eβ(cid:19) Hα,nHn,β
(26)
and α,β either corresponds to the xy or to the ± levels. For the xy states one finds
(cid:104)xy,↑ H (2)xy,↓(cid:105) ≈ −(cid:104)xy,↑ HI+,↑(cid:105)(cid:104)+,↑ Hasoxy,↓(cid:105)
− (cid:104)xy,↑ Haso+,↓(cid:105)(cid:104)+,↓ HIxy,↓(cid:105)
∆
∆
and similarly for (cid:104)xy,↓ H 2xy,↑(cid:105). Inserting the matrix elements yields an effective Rashba SOC ∼ kyτ x − kxτ y
H SOC
xy = −4
γ∆aso
∆ (cid:18)
0
ky + ıkx
ky − ıkx
0
(cid:19) = −αxy(τ xky − τ ykx)
(27)
with a negative coupling constant with αxy = 4γ∆aso/∆.
From Fig. 3 one can see that the same matrix elements also mediate the 2nd order interaction between the
+, σ and +,−σ states. Since in this case the denominator in Eq. (26) is negative we obtain a positive coupling
α+ = 4γ∆aso/∆ for the E+ states
H SOC
E+ = 4
γ∆aso
∆ (cid:18)
0
ky + ıkx
ky − ıkx
0
(cid:19) = α+(τ xky + τ ykx) .
(28)
Figure 3. Level structure and interactions of the three-band hamiltonian. Interactions of Haso ∼ ∆aso Eq. (2) are shown
in red and those of the asymmetry hopping HI Eq. (3) are indicated in blue. The red dashed lines correspond to a atomic
SO interaction which around the Γ-point (i.e. ∼ tk2 < ∆aso) is much smaller than the matrix elements represented by
the solid lines. Panel (a) corresponds to the original Hamiltonian whereas in (b) the atomic SOC between xz and yz has
been diagonalized.
Moreover the off-diagonal matrix elements in Eq. (28) are c.c. to those of Eq. (27) which means that the +, σ
and +,−σ states are interacting via a Dresselhaus coupling ∼ kyτ x + kxτ y.
The effective interactions between −, σ and −,−σ can be again obtained from 2nd order perturbation theory
in the limit tk2 < ∆aso which now involves the matrix elements represented by the dashed lines in Fig. 3. The
resulting effective coupling reads
γ(t1 + t3 − t2)(k2
x − k2
y)
0
ky + ıkx
−
H SOC
x − k2
and therefore corresponds to a linear Rashba SOC but with a coupling constant ∼ (k2
ky − ıkx
= −
y)(τ xky − τ ykx)
x − k2
y).
∆
0
(cid:19) = −β(k2
(cid:18)
5. THE CLEAN LIMIT
In this section we evaluate the Drude weight for the effective models discussed in section 4.
5.1 xy bands
In this case the eigenvalues and eigenvectors corresponding to the Hamiltonian (27) read
(29)
(30)
Exy
± =
k2
2m ± αxyk, ±(cid:105) =
1
√2(cid:18) ∓ıe−ıθ
1
(cid:19)
with px = cos(θ) and py = sin(θ). As for the Rashba model (9), the spin operator is simply the Pauli matrix
Sy = τ y/2 and the charge current is similar to the 2DEG case Jx = (−e)(kx/mτ 0 + αxyτ y). The interband
matrix elements read
(cid:104)k, sSyk,−s(cid:105) =
1
2
is sin(θ), (cid:104)k,−sJxk, s(cid:105) = −is sin(θ)(−e)αxy
and the response function (in the zero-temperature limit) gives
R(cid:48)
yx(ω → 0) =
1
2
(−e)αxy(cid:88)ps
sin2(θ)
θ(Exy
s (p) − µ) − θ(Exy
s (p) − Exy
Exy
−s(p)
−s(p) − µ)
=
1
2
(−e)αxyN0,
which leads to the Drude weight
with an opposite sign as compared to the 2DEG case of Eq. (14).
DISG
xy =
π
2
eαxyN0
(31)
a)xyxyxyxyxzxzyzyz+−−+b)5.2 E− bands
In this case the eigenvalues and eigenvectors corresponding to the Hamiltonian (29) read
E−
± =
k2
2m ± βζ k3, ±(cid:105) =
1
√2(cid:18) ∓ı ζ
e−ıθ
ζ
1
(cid:19)
(32)
where ζ = k2
x −
y)τ y/2, while the charge current has a more complicated structure as compared to the 2DEG case Jx =
k2
In this case the spin operator reads Sy = −γ(k2
y with kx = cos(θ) and ky = sin(θ).
x − k2
(−e)(cid:0) kx
m τ 0 − 2βkxkyτ x + β(3k2
x − k2
sin(θ)
y)τ y(cid:1). The interband matrix elements read
ζ
(cid:104)k, sτ yk,−s(cid:105) = ıs
ζ
ζ
(cid:104)k,−sτ yk, s(cid:105) = −ıs
ζ
ζ
(cid:104)k, sτ xk,−s(cid:105) = ıs
cos(θ)
ζ
ζ
(cid:104)k,−sτ xk, s(cid:105) = −ıs
ζ
cos(θ).
sin(θ)
In the response function
R(cid:48)
xy(0) = −(−e)(cid:88)k,,s
= (−e)
γ
2
0
2π
2π
1
2
(−γk2)(k2
dθ
2π
(k2
x − k2
y)2
γ
= (−e)
2
= (−e)γk2
0
F βk2
dθ
2π
x − k2
(k2
y)2
F N0(cid:18)−
4(cid:19)
1
y)k2(cid:104)−2βk2
x
k2
k+
dk
2π
k−
+ − k3
k3
−
6π
x − k2
k2
y
ζ
k2
y
ζ
k2
y + β(3k2
x
k2
y − k4
y)(cid:105) θ(µ − E−s (k)) − θ(µ − E−
2sβζk3
−s(k))
the factors ζ disappear and the Drude weight reads
DISG
− = (−e)(cid:16) π
4
γp2
F βp2
F N0(cid:17) .
5.3 E+ bands
In this case the eigenvalues and eigenvectors corresponding to the Hamiltonian (28) read
(33)
(34)
with kx = cos(θ) and ky = sin(θ). In this case the spin operator reads Sy = γ(k2
x − k2
y)τ y/2, while the charge
current is Jx = (−e)(cid:0) px
1
Exy
± =
k2
2m ± α+k, ±(cid:105) =
1 (cid:19)
√2(cid:18) ∓ıeıθ
m τ 0 + α+τ y(cid:1). The interband matrix elements are
(cid:104)k, sτ yk,−s(cid:105) = −ıs sin(θ),
(cid:104)k,−sτ yk, s(cid:105) = ıs sin(θ).
The response function is
R(cid:48)
xy(0) = −(−e)γ(cid:88)k,s
k2(k2
x − k2
y)k2
y
= −(−e)
= (−e)γk2
2π
dθ
2π
γ
2
x − k2
(k2
4(cid:19)
1
0
F α+N0(cid:18)−
s (k)) − θ(µ − E+
−s(k))
2sα+k
α+
2
θ(µ − E+
k+
k2dk
2π
y)k2
y
k−
and the Drude weight is
DISG
+ = (−e)(cid:16) π
4
γp2
F α+N0(cid:17) .
(35)
6. THE DISORDERED LIMIT
It is well known that in the presence of disorder, the Drude weight in the formula for the optical conductivity
is suppressed and the spectral weight goes into the regular part.
In the Drude model, the regular part, as
function of the frequency, has a Lorentzian shape whose width is controlled by the scattering rate τ−1 (not to
be confused with the Pauli matrices). Such a transfer of spectral weight from the singular to the regular part
occurs also in the case of the SGC. To this end we need to introduce disorder in our model. This will be done in
the numerical computation of the next section, whereas in this section we introduce disorder within the effective
models derived in section 4 by using the standard diagrammatic impurity technique. This technique has been
applied to the Rashba model for the evaluation of the ISG effect,8 anisotropy magnetoresistance47, 50 and spin
Hall effect.48 We review here the basic aspects by focusing on the case of the xy-bands, which is equivalent to the
Bychkov-Rashba model in the 2DEG. By following the standard procedure, disorder is introduced as a random
potential V (r), with zero average (cid:104)V (r)(cid:105) = 0 and white-noise correlations (cid:104)V (r)V (r)(cid:105) = niu2δ(r − r(cid:48)), with ni
being the impurity concentration. By Fermi golden rule, one associates a scattering rate τ−1 = 2πniu2N0, where
N0 is the single-particle density of state previously introduced in Eq. (14). We will consider the weak-disorder
limit which is controlled by the small parameter (EF τ )−1, with EF the Fermi energy.
In the diagrammatic
impurity technique, the first step is the introduction of the irreducible self-energy in the self-consistent Born
approximation for the electron Green function.
6.1 The case of the Exy bands
The Green function, due to the SOC of the lowest pair of bands H SOC
matrices as G = G0τ 0 + G1τ x + G2τ y and explictly reads
xy
of Eq. (27), can be expanded in Pauli
G(, k) =
G+(, k) + G−(, k)
2
τ 0 − (τ xky − τ y kx)
G+(, k) − G−(, k)
2
where
and the self-energy has the form
G±(, k) =(cid:2) − Exy
± (k) − Σ()(cid:3)−1
,
Σ() = ∓
ı
2τ
τ 0,
(36)
(37)
(38)
the minus and plus signs applying to the retarded (R) and advanced (A) sectors, respectively. The scattering
time τ entering Eq. (38) is exactly the one required by the Fermi golden rule.
It is worth noticing that the
self-energy is proportional to the identity matrix in the spin space.50 Once the Green function is known, we may
compute the SGC by means of the Kubo formula
Figure 4. Ladder diagrams for the determination of the dressed vertex. The gray-filled triangle represents the infinite
sum of diagrams, which results from repeated scattering. The solid lines with arrows are Green function propagators for
electrons, whereas the dashed lines represent the operation of impurity average.
σISG =
1
2π(cid:10)Tr(cid:2)SyGRJxGA(cid:3)(cid:11)dis av
(39)
which can be obtained from the expression (5), after averaging over the disorder configurations, represented as
(cid:104). . .(cid:105)dis av. In the above the Tr . . . symbol involves all degrees of freedom, i.e. spin and space coordinates. The
disorder average in Eq. (39) enters in two ways. The first is to use the disorder-averaged Green function given in
Eq. (36). The second is the introduction, to lowest order in the expansion parameter (EF τ )−1, of the so-called
ladder diagrams, which lead to vertex corrections. The vertex corrections procedure can be performed either for
the spin or charge vertex of Eq. (39). Here we consider the vertex correction for the charge current vertex. The
dressed vertex Jx obeys the Bethe -- Salpeter equation
Jx = Jx + niu2(cid:88)k
GR(, k)JxGA(, k),
(40)
which results from the infinite summation of ladder diagrams, as shown in Fig. 4. In terms of the dressed vertex
the SGC reads
σISG =
1
2π(cid:88)k
tr(cid:104)SyGR(, k)JxGA(, k)(cid:105) ,
(41)
where now the lower case trace symbol involves the spin degrees of freedom only. The problem is then reduced
to the solution of the Bethe -- Salpeter equation (40) and to the evaluation of the bubble (41). In general the
Bethe -- Salpeter equation is an integral equation. However, in the present case of white-noise disorder, the Bethe --
Salpeter equation becomes an algebraic one, even though still having a spin structure. In the appendix A we
provide the details of the solution of Eq. (40), which leads to
which shows that the vertex corrections exactly cancel the interband matrix elements of the charge current
vertex. As a result, the evaluation of Eq. (41) leads to
Jx = (−e)
kx
m
,
(42)
σISG
xy = eN0αxyτ,
(43)
which must be compared with the Drude weight evaluated in Eq. (31).
6.2 The case of the E− bands
According to the analysis of appendix A, the dressed charge current vertex reads
Jx = (−e)(cid:20) kx
m
The evaluation then of Eq. (41) leads to
τ 0 − 2βkxkyτ x + β(3k2
x − k2
y)τ y +
1
4
βp2
F τ y(cid:21) .
Γ(4)≡L=(c)=+(a)=+++...(b)=+++...p+q,αp,γp′+q,δp′,βϵn+ωνϵnσISG
− = −
5
8
eN0γβp4
F τ,
(44)
which has a sign opposite to that of the Exy bands. In Eq. (44) the combination βp2
SOC, whereas γp2
F is the spin dressing factor accounting for the interactions in the original model.
F plays the role of an effective
6.3 The case of the E+ bands
According to the analysis of appendix A, the dressed charge current vertex reads
The evaluation then of Eq. (41) leads to
Jx = (−e)
kx
m
τ 0.
which shows again a change of sign with respect to that of the E− bands. Also here the combination γp2
spin dressing factor accounting for the interactions in the original model.
F is the
σISG
+ = (γp2
F )eN0α+τ,
(45)
7. THE NUMERICAL APPROACH
In this section we present our numerical results. The starting point is the response function defined in Eq. (5),
which may be expressed as follows
Ryx(ω) =
1
N (cid:88)k,p
(fp − fk) (cid:104)pSyk(cid:105)(cid:104)kJxp(cid:105)
ω + iη + Ep − Ek
,
(46)
where k and p are quantum numbers labelling the eigenstates of the Hamiltonian. For instance, in the absence
of disorder, the index k includes the crystal momentum, the orbital and spin degrees of freedom. The symbol fk
stands for the Fermi function evaluated at the energy of the eigenstate k. In Eq. (46) N is the number of lattice
sites. The numerical evaluation is performed on a finite system and then it is convenient to separate from the
outset the Drude singular weight from the regular part as follows
DISG = −
π
N (cid:88)k,p
σISG
reg = lim
ω→0P
R(cid:48)(cid:48)
yx (ω)
ω
= −
fp − fk
Ep − Ek (cid:60)(cid:104)pSyk(cid:105)(cid:104)kJxp(cid:105),
N (cid:88)k,p
Ep − Ek (cid:61)(cid:104)pSyk(cid:105)(cid:104)kJxp(cid:105)
fp − fk
iη + Ep − Ek
1
(47)
(48)
,
where (cid:60) and (cid:61) indicate the real and imaginary parts.
Fig. 5 shows the behavior of the spin-orbit split gap at the Fermi surface. Inspection of Fig. 5 reveals that for
µ = 0.3 eV the gap has extrema at energies ∆ ≈ 0.005 and 0.009 eV which are expected to dominate the response
due to the "saddle-point" character of the corresponding states as discussed below. For µ = 0.7 , as shown in
Fig. 2, all bands are occupied and all the gaps will appear in the response function. In particular the pair of
bands (1,2) contributes to the gap at energies ∆ ≈ 0.015 eV, the pair of bands (3,4) at energies ∆ ≈ 0.005−0.015
eV, and the pair of bands (5,6) at energies ∆ ≈ 0.
Fig. 6 shows the frequency dependence of the real and imaginary parts of the response function Ryx(ω) for
the three chemical potentials µ = 0.3 eV, µ = 0.425 eV and µ = 0.7 eV. The underlying ground state is for a
homogeneous system but we investigate the influence of the particle-hole lifetime parameter η. As compared with
the approach discussed in Sec. 6 this mimics the inclusion of momentum relaxation without considering vertex
corrections. For µ = 0.3 eV, when only the lowest pair of bands is occupied, one may interpret the observed
behavior in terms of the Rashba model of Eq. (9). The low energy structure is determined by transitions across
Figure 5. Size of the spin-orbit split gap around the Fermi surface for µ = 0.3 eV (a) and µ = 0.7 eV (b,c). This is obtained
by determining the cut kF of each band with µ and then calculating the energy difference to the 'other' SO split band at
the same kF . If the SO interaction has some significant momentum dependence around kF the gap determined for each
of the two bands from a pair may slightly differ as in panel (a). The angle Θ is defined with respect to the kx-axis.
the spin-orbit split gap of the same t2g band. In fact, it is exactly in this energy range that the imaginary part of
Ryx(ω) develops a peak structure as discussed in Eq.(13). In the clean system, i.e. lifetime parameter η → 0, the
imaginary part vanishes for energies below the minimum gap excitations and therefore the slope of R(cid:48)(cid:48)
yx(ω), which
determines the regular ISG response σISG
yx,reg, is zero, whereas the limiting value of the real part fixes the Drude
weight of the singular contribution. As shown in Fig. 6 (panel (a)) a finite η (or, similarly, a finite temperature)
broadens the excitations and therefore induces a finite slope of R(cid:48)(cid:48)
yx(ω) at ω = 0 leading thus to a finite ISG
response. This is evidenced by making η larger in Fig. 6: the red curve is for η = 10−5, while the blue one for
η = 10−3. The fact, that the ISG response vanishes for a clean system is consistent with the analysis in Ref.41
and with the discussion at the end of section 3. According to the KKR (7), the sign of the imaginary part of
the ISG response function is determined by the sign of the effective Rashba SOC. The negative sign shown by
the numerical evaluation of Fig. 6 agrees with the sign found for the coupling −αxy in the effective model for the
lowest pair of xy bands in Eq. (31). We remind that the Drude weight is determined by the zero-frequency value
of the real part, which is obtained from the imaginary part via the KKR (7).
For the chemical potential µ = 0.425 eV, close to the Lifshitz point, all the bands are occupied and the
imaginary part of the response function gets contributions from the interband transitions across the spin-orbit
split gaps of all the pairs of bands as well as from the interband transitions involving different pairs of bands
simultaneously. In panel (b) of Fig. 6 this is evidenced by showing, together with the full imaginary part (red line)
also the contribution of the individual pairs of bands: (1,2) (green line), (3,4) (blue dashed line) and (5,6) (yellow
line). From this we conclude that the large spectral weight at energy ω = 0.02 is due to interband transition
between different pairs of bands. The green curve shows that the contribution due exclusively to the lowest pair
of bands (1,2) is still around the same energy as in panel (a) and hence the behavior of this pair of bands is still
2bandscuttheFermilevelforwhichthegapsareshowninFig.2b,c.Thepairofbandscorrespondingtothelowesteigenvalues(1,2)hasanopenFermisurfaceatµ=0.7(cf.Fig.1)andtheSOgapvanishesattheantinodalpoints.ThemiddleanduppermostpairofbandshaveafiniteSOgapallaroundtheFermisurface.Fortheuppermostpairitismaximalalongthediagonaldirec-tionwhereasforthemiddlepairitisminimalalongthediagonalandattheantinodalpoints.-π/20π/2-π/20π/2-π/20π/2-π/20π/2µ=0.7 eVµ=0.425 eVµ=0.3 eVFIG.1:FermisurfacesforthechemicalpotentialsindicatedbygreylinesinFig.??.WeproceedbycalculatingtheinverseEdelsteine↵ect(IEE)whichisdefinedviatheresponsefunction[8] IEE(!)=hhJx;Syii!i(!+i⌘)(4)00,0050,01∆ [eV]band 1band 200,0050,010,0150,02∆ [eV]band 3band 3-1-0,500,51Θ/π00,0050,010,0150,02∆ [eV]a) ∆(µ=0.3 eV)b) ∆(µ=0.7 eV)c) ∆(µ=0.7 eV)bands 1,2bands 3,4bands 5,6band 1band 2FIG.2:Sizeofthespin-orbitsplitgaparoundtheFermisur-faceforµ=0.3eV(a)andµ=0.7eV(b,c).ThisisobtainedbydeterminingthecutkFofeachbandwithµandthencal-culatingtheenergydi↵erencetothe'other'SOsplitbandatthesamekF.IftheSOinteractionhassomesignificantmo-mentumdependencearoundkFthegapdeterminedforeachofthetwobandsfromapairmayslightlydi↵erasinpanel(a).Theangle⇥isdefinedwithrespecttothekx-axis.withhhJx;Syii!=1NXk,p(fp fk)hpSykihkJxpi!+i⌘+Ep Ek(5)whereJxisthetotalchargecurrent(obtainedfromtheusualPeierlssubstitution)andSydenotesthetotaly-polarizedspin.OnefindsJx= Xk [2t1sin(kx)+4t3cos(ky)sin(kx)]c†k ,xyck ,xy Xk 2(t1+t3)sin(kx)c†k ,xzck ,xz Xk 2t2sin(kx)c†k ,yzck ,yz+2i Xk cos(kx)[c†k ,xyck ,yz c†k ,yzck ,xy]andforthespinSy=12Xk, , 0↵=xy,xz,yzc†k ,↵⌧y 0ck 0,↵.(6)TherealpartiscomposedofaIEE'Drude'andareg-ularpart< |