• Keine Ergebnisse gefunden

Spin-Orbit Interaction in Symmetric Wells with Two Subbands

N/A
N/A
Protected

Academic year: 2022

Aktie "Spin-Orbit Interaction in Symmetric Wells with Two Subbands"

Copied!
4
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Spin-Orbit Interaction in Symmetric Wells with Two Subbands

Esmerindo Bernardes,1John Schliemann,2,3Minchul Lee,3J. Carlos Egues,1,3,4and Daniel Loss3,4

1Instituto de Fı´sica de Sa˜o Carlos, Universidade de Sa˜o Paulo, 13560-970 Sa˜o Carlos, Sa˜o Paulo, Brazil

2Institute for Theoretical Physics, University of Regensburg, D-93040 Regensburg, Germany

3Department of Physics and Astronomy, University of Basel, CH-4056 Basel, Switzerland

4Kavli Institute for Theoretical Physics, University of California, Santa Barbara, California 93106, USA (Received 11 July 2006; published 17 August 2007)

We investigate the spin-orbit (SO) interaction in two-dimensional electron gases in quantum wells with two subbands. From the 88 Kane model, we derive a new intersubband-induced SO term which resembles the functional form of the Rashba SO but is nonzero even insymmetricstructures. This follows from the distinct parity of the confined states (even or odd) which obliterates the need for asymmetric potentials. We self-consistently calculate the new SO coupling strength for realistic wells and find it comparable to the usual Rashba constant. Our new SO term gives rise to a nonzero ballistic spin-Hall conductivity, which changes sign as a function of the Fermi energy ("F) and can induce an unusual Zitterbewegungwith cycloidal trajectorieswithoutmagnetic fields.

DOI:10.1103/PhysRevLett.99.076603 PACS numbers: 72.25.Dc, 71.70.Ej, 73.21.Fg, 85.75.d

The rapidly developing field of spintronics has generated a great deal of interest in spin-orbit (SO) coupling in semiconductor nanostructures [1]. For an n-doped zinc- blende semiconductor quantum well with only the lowest subband occupied, i.e., in a strictly 2D situation, there are two main contributions to the interaction of the spin and orbital degrees of freedom of electrons. One contribution is the Dresselhaus term, which results from the lack of in- version symmetry of the underlying zinc-blende lattice [2]

and is to lowest order linear in the crystal momentum [3].

This linearity is shared by the other contribution known as the Rashba term [4], which is due to structural inversion asymmetry and can be tuned by an electric gate across the well [5]. These two contributions can lead to an interesting interplay in spintronic systems [6].

In this Letter we consider yet another type of electronic SO coupling which, as we show, occurs in III-V (or II-VI) zinc-blende semiconductor quantum wells with more than one subband. We derive a new intersubband-induced SO interaction which resembles that of the ordinary Rashba model; however, in contrast to the latter, ours is nonzero even insymmetricstructures (Fig.1). We self-consistently determine the strength of this new SO coupling for realistic single and double wells and find it comparable to the Rashba constant [Figs. 2(a) and 2(b)]. We have investi- gated the spin-Hall effect and the dynamics of spin- polarized electrons due to this new SO term. We find (i) a nonzero ballistic spin-Hall conductivity which changes sign as a function of "F and (ii) an unusual Zitterbewegung [7] with cycloidal trajectories without magnetic fields (Fig.3). As derived below, for a symmetric well with two subbands our44electron Hamiltonian is

H p~2

2m

11z1

@x pxypyx; (1)

wherem is the effective mass, "o "e=2,"eand

"o are quantized energies of the lowest (even) and first excited (odd) subbands (corresponding to eigenstates jei and joi), respectively, measured from the bottom of the quantum well, x;y;z denote the Pauli matrices describing the subband (or pseudospin) degree of freedom, andx;y;z are Pauli matrices referring to the electron spin. The new intersubband-inducedSO couplingis

1

E2g 1 Eg2

P2

3 hej@zVzjoi

v

E2g Eg2

P2

3 hej@zhzjoi; (2) where Eg and are the fundamental and split-off band gaps in the well region [8],P is the Kane matrix element [9]. The parametersvanddenote valence-band offsets between the well and the barrier regions [10],Vzis the Hartree-type contribution to the electron potential, and hzis the structural quantum-well profile [11]. Note that

) (

: z

o ϕo

0 ) ( ) ( ) ( )

(

ϕeaϕoa ϕe aϕo a η

o e i a

a i

i

iϕ()2ϕ()2=0, =, α

δc

εe

εo ) (

: z

e ϕe

a a z

FIG. 1 (color online). Square well with its ground-stateez and first excited-state oz wave functions. The new intersubband-induced SO coupling in Eq. (4) is nonzero even in symmetric wells due to the distinct parities of ez (even) and oz (odd), which yield a nonvanishing matrix element for the derivative of the symmetric potential.

PRL99,076603 (2007) P H Y S I C A L R E V I E W L E T T E R S week ending 17 AUGUST 2007

0031-9007=07=99(7)=076603(4) 076603-1 © 2007 The American Physical Society

(2)

can be varied via external gates (Fig.2). Next we outline the derivation ofH in Eq. (1).

Kane Hamiltonian. —We start from the usual 88 Kane Hamiltonian describing the s-type conduction and thep-type valence bands around thepoint [12],

H88 Hc Hcv Hvc Hv

; (3)

where Hc is a 22 diagonal matrix with elements

p2=2m0Vc~r,m0is the bare electron mass,Hvis a6 6diagonal matrix with elementsp2=2m0Vv~r Egfor the heavy- and light-hole bands, and p2=2m0V~r Egfor the split-off band, Vi~r (ic; v;) denote arbitrary potentials (see below), andHcv Hvcyis

Hcv

p2

2 3

q z p6 0 p3z p3 0 p6

2 3

qz p2 p3 pz3 0

B@

1 CA; (4)

where~ P ~k,k~p=~ @is the electron wave vector,k kx iky, and P i@hSjpxjXi=m0 parametrizes the conduction-to-valence-band coupling, jSiandjXiare the usual periodic Bloch functions at thepoint.

Effective electron Hamiltonian: Folding down.—The Kane Hamiltonian (3) acts on an eight-component spinor y c vy in which the last six components v represent valence-band states. By eliminating the hole components from the Schro¨dinger equation H88

", where"is the eigenenergy, we can fold down this8 8 equation into a 22 effective equation for the conduction-band states only: H"~

c HcHcv"

Hv1Hvc~

c, ~

c is a renormalized conduction-electron spinor.

SO in symmetric wells. —Applying the above procedure to a quantum well, defined by the confining potentials [11]

Vi~r !Viz Vz ihz,ic; v;, we find H" HQW P2

3@2p11 12 ; pz; (5) HQW p 1

2mz; "ppz 1

2mz; "pzVcz; (6) where 1=mz; " 2P2=3@22=11=2 1=m0, 1 " p2=2m0Vz vhz Eg and 2

" p2=2m0Vz hz Eg. Equation (5) describes an electron in a quantum well (HQW term) with spin-orbit interaction (last term) [13]. The kinetic-energy operators above are complicated due to the position- and energy-dependent effective mass mz; ". Since Eg and Egare the largest energy scales in our system, we can simplify (5) and (6) by expanding1=1 and1=2 in the form 1=1 E1g f1 "p2=2m0Vz vhz=

Eg g and 1=2 Eg1f1 "p2=2m0 Vz hz=Eg g. To zeroth order 1 Eg, 2Eg, and HQWp2k=2mp2z=2m Vcz with (a constant effective mass) 1=m 2P2=3@22=Eg1=Eg 1=m0 [14]. Since the SO operator 11 12 ; pz !@z1=1 @z1=2, we need to keep the first-order terms in the expansions of 11 and12 which yield the leading nonzero contribution to the SO term in (5). We find11 12 ; pz 1=E2g 1=Eg2@zVz v=E2g =Eg 2@zhz.

Finally, we project this SO operator into the twolowest (spin-degenerate) eigenstates jiiz jk~kiijzi, h~rjk~kii expi ~kk~rkiz,ie; o, andz";#, of thesymmetric

-6 -4 -2 0 2 4

-0.2 0.0 0.2 0.4 0.6 0.8

η,αii(meVnm)

Vb(eV)

(a) GaInAs single well αη

αeo βe βo Vb=0

-20 -10 0 10 20 30

-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6

η,αii(meVnm)

Vb(eV) (b) InSb double well

αηe αo βi

Vb=0 12 13

-0.02 0.02

ε±(k)(meV)

k(nm−1) Vb=0

FIG. 2 (color online). Calculated SO coupling strengths as a function of the external gate Vb for realistic wells. (a) For the single GaInAs [16,20] well studied, the intersubband coupling is larger than the Dresselhausi and the Rashbai constants (ie; o). Note that jej joj and both change sign across Vb0(in contrast toi and). (b) For the InSb double well considered,shows a ‘‘resonant behavior’’ aboutVb0[sym- metric configuration, lower-left inset in (b)]. This occurs because the subband splitting"o"ereaches a minimum atVb0and the double-well wave functions are very similar (though of distinct parities) forVb0. This also makese oaround Vb0. Upper-right inset in (b): Energy dispersions " k~ [Eq. (10)] of the symmetric double well.

FIG. 3 (color online). Zitterbewegungdue to the SO coupling for distinct ratios =2"SO. Note the peculiar trajectories with the forward injected electrons moving backward (I) and even in aclosed path (II). This follows from the SO induced change in the curvature of the bands which renormalizes the effective masses. Here we useSOk0y =10,1SOm=@.

PRL99,076603 (2007) P H Y S I C A L R E V I E W L E T T E R S week ending 17 AUGUST 2007

076603-2

(3)

well (HQW) (Fig.1). This directly leads to the H in (1) with the SO coupling (2) [15]. Note that this new SO interaction is nonzero even in symmetric wells asarises from the coupling between the ground state (even) and the first excited state (odd) [Eq. (2)]. We can generalizeH to include the Rashbaand the linearized DresselhausSO couplings. Next we determine the magnitude of(and, ) for realistic quantum wells with two subbands.

Self-consistent calculation of the SO couplings. —We consider modulation-doped quantum wells similar to those experimentally investigated in Ref. [16]. Our wells, how- ever, havetwooccupied subbands. Similarly to Ref. [16], we study cases with constant chemical potentials [17]. By self-consistently solving Poisson and Schro¨dinger’s equa- tions we determine the energy levels "e; "o and the con- fined wave functionsiz,ie; oof the wells. We then calculate (i)via Eq. (2), (ii)ifrom equations similar to Eq. (2) for each subband, and (iii)ichijk2zjii,cis the bulk Dresselhaus SO parameter [18]. The structural sym- metry of the wells and their charge densities can be changed via a gate potentialVb.

Our calculated SO couplings,i, andi,ie; o, for anInAlAs=InGaAs=InAlAs single quantum- well (‘‘sam- ple 3’’ in [16]) are all comparable in magnitude [Fig.2(a)].

Note that our e=e ratio is consistent with the experi- mental one in Ref. [19]. In addition,e vsVb here agrees well with the experimental data in Fig. 3 (‘‘triangle up’’) of Ref. [16] (see also Fig. 4 in [20]). Ouri[21] andiare also consistent with those of Ref. [22]. Note that for the single well studied heredoes not vary appreciably with the gateVb, similarly toiand as opposed toi.

For adouble-well structure, on the other hand, we find that has a ‘‘resonant behavior,’’ changing by about an order of magnitude asVbis swept acrossVb0[Fig.2(b)]

(this may have a dramatic effect on Shubnikov– de Haas measurements).Vb0corresponds to a fully symmetric double well. In contrast to the single-well case,eando have opposite signs and undergo abrupt changes in magni- tudes nearVb 0[Fig.2(b), dashed lines]. Similarly to the single-well case,e and o are also essentially constant for a double well [Fig. 2(b), dotted lines]. A detailed account of our results will be presented elsewhere.

Having established that the new SO couplingis sizable, in what follows we focus on a fully symmetric well to investigate physical effects arising solely from.

Fully symmetric case: Eigensolutions. —Let us consider a two-subband well (single or double) described by the Hamiltonian H in (1) (we assume a negligible Dresselhaus term [19]). In the basis fjei";joi#;joi";jei#g H becomes

H~

"2k2

2m"e ik 0 0

ik "2m2k2"o 0 0

0 0 "2m2k2"o ik

0 0 ik "2k2

2m "e 0

BB BB

@

1 CC CC A:

(7)

Both the upper-left (U) and lower-right (L) blocks of H~ have eigenvalues

" k ~ k @; (8) with k @2k2=2m, @2 2k22, and ei- genvectors

j 1iUsin=2jei"cos=2eijoi#; (9) j 2iLcos=2joi"sin=2eijei#; (10) j 3iUcos=2jei"sin=2eijoi#; (11) j 4iLsin=2joi"cos=2eijei#: (12)

Here, ei kyikx=k, cos 1=

1 k=2 p , and k~ksin;cos (here we drop the ‘‘k’’ in k~k).

For k2 we can expand " k~ in (8) and define effective massesm m=1 2"SO=, where"SO 2m=2@2 is the energy scale of the new SO coupling. For the double well of Fig. 2(b), m is reduced by 5%

compared to the bulk value m. This could be measured via, e.g., cyclotron-resonance experiments [23].

Novel Zitterbewegung. — The dynamics of electron wave packets in wells with SO interaction exhibit an oscillatory motion [7]— theZitterbewegung. For our new SO interaction, a wave packet ji moves according to hj~rHtji where ~rHt Uy~rUis the position operator in the Heisenberg picture [UexpiHt="] with com- ponents

xHt 11x0 11px mt

@txy

2@2

yy

@py1z

cos2t 1 2@3

2xy

@z1px

@ 2

pyx pxypyx

sin2t 2t; (13) and yHt, obtained from Eq. (13) via the replacements px; x哫py; y, py; y哫px;x (i.e., a =2 rotation about the z axis). Similar expressions can be derived for the spin components iHt,ix; y; z[24].

For simplicity, we evaluate the expectation value of

~rHt for planes waves (‘‘wide wave packets’’). For a spin-up electron injected into the lowest subband along theyaxis with (group) velocityv~g @k0y=my, we find^

hxHti 2k0y

2@21cos2t; (14)

hyHti @k0y

m t2k0y

2@3 sin2t 2t; (15) assuming x0 y0 0. Equations (14) and (15) show that cycloidal motion is possible in our system. This differs PRL99,076603 (2007) P H Y S I C A L R E V I E W L E T T E R S week ending

17 AUGUST 2007

076603-3

(4)

qualitatively from the Rashba SOZitterbewegungwhich is always perpendicular to the initialv~g.

Figure 3 shows trajectories for three distinct

~

vg @k0y=my—all with^ k0y>0. We find motion op- posite toand along theyaxis (orbits I and III, respectively) and even a closed path (II). To understand this behavior we note that for k0y"o"e the linear-in-t terms in hyHti can be recast into@k0yt=m) the injected wave moves with the renormalized velocity vg@k0y=m. Hence, for orbit I, <1)m<0[25] andvg<0, for orbit II, 1)m! 1andvg0, and for orbit III, >1)m>0 and vg>0. Though remarkable, we stress that the orbits I and II occur forunusualparameters (e.g., "F< "SO=10). However, these orbits do show that our SO Hamiltonian has a physical mechanism allowing for cyclotronic motion without magnetic fields.

Spin-Hall conductivity zxy.—The spin-Hall effect is a convenient probe for SO effects in wells [26]. We have calculatedzxy(‘‘clean limit’’) in the presence of an exter- nal magnetic fieldBby following the approach of Rashba [27], which allows us to properly account for both the intrabranch and interbranch contributions in the Kubo formula [28]. Here we focus on the B!0 limit where we findzxy0for"F> "o(two subbands occupied) and

zxy e 8

1 1

1 32

21=2 233

(16) for "e< "F< "o (upper subband empty), where 1 2"SO=, 2 "F=2, and 3

21=4121=4

q . Note thatzxy is nonzero and non- universal in this range, shows a discontinuity at"F "o, and changes sign as a function of "F. Details of our calculation of zxyand a thorough discussion will be pre- sented elsewhere [29]. Here we just note that measure- ments ofzxy(versus"F) in symmetric two-subband wells offer a possibility to probe our new SO interaction [26].

Note that the Rashba and the linearized Dresselhaus spin- Hall conductivities are identically zero in the dc limit [27,30].

We have introduced an intersubband-induced SO inter- action in quantum wells with two subbands. The corre- sponding SO coupling (whose magnitude is similar to Rashba coupling) is nonzero even in symmetric wells. This new SO interaction gives rise to a nonzero spin-Hall con- ductivity, renormalizes the bulk mass by5%(measurable via cyclotron resonance [23]) in double wells, and can induce a cycloidalZitterbewegung. Weak antilocalization [20,31] should offer another possibility to measure.

We thank S. Erlingsson, D. S. Saraga, D. Bulaev, J. Lehmann, M. Duckheim, L. Viveiros, G. J. Ferreira, R.

Calsaverini, and E. Rashba for useful discussions. This work was supported by the Swiss NSF, the NCCR Nanoscience, DARPA, ONR, CNPq, FAPESP, DFG via SFB 689, and U.S.A. NSF PHY99-07949.

[1] Semiconductor Spintronics and Quantum Computation, edited by D. D. Awschalom, D. Loss, and N. Samarth (Springer, Berlin, 2002); I. Zutic et al., Rev. Mod. Phys.

76, 323 (2004).

[2] G. Dresselhaus, Phys. Rev.100, 580 (1955).

[3] M. I. Dyakonov and V. Y. Kachorovskii, Sov. Phys.

Semicond. 20, 110 (1986); G. Bastard and R. Ferreira, Surf. Sci.267, 335 (1992).

[4] Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984).

[5] G. Engelset al., Phys. Rev. B55, R1958 (1997); J. Nitta et al., Phys. Rev. Lett.78, 1335 (1997).

[6] J. Schliemannet al., Phys. Rev. Lett.90, 146801 (2003);

M. Trushin and J. Schliemann, Phys. Rev. B 75, 155323 (2007).

[7] J. Schliemann et al., Phys. Rev. Lett. 94, 206801 (2005).

[8] I. Vurgaftmanet al., J. Appl. Phys.89, 5815 (2001).

[9] E. O. Kane, J. Phys. Chem. Solids1, 249 (1957).

[10] Note that vEbgEgc and vb, whereEbg andb are, respectively, the fundamental and split-off band gaps in the barriers andcthe conduction- band potential offset [see, e.g., P. Pfeffer and W.

Zawadzki, Phys. Rev. B68, 035315 (2003)].

[11] R. Lassnig, Phys. Rev. B31, 8076 (1985).

[12] T. Darnhofer and U. Ro¨ssler, Phys. Rev. B 47, 16020 (1993).

[13] In (5) we neglect an effective Darwin term as it arises in relativistic quantum mechanics. This term leads to a rigid intrabandshift and can be absorbed into"e; "o.

[14] We treat m as a parameter in our model (the Kane effective mass neglects corrections from higher-lying bands).

[15] In Eq. (1) we assume thathej@zVjoiand (in turn) are real. If jjei, we have to replace xbycosx siny in (1). The freedom in fixinghas no physical consequences since a change incan be compensated by applying the phase factor expiz to the basis states:

cosxsinyexpiz x. Thus,H enjoys a U(1) gauge symmetry corresponding to a rotation about thezdirection in the~space.

[16] T. Kogaet al., Phys. Rev. Lett.89, 046801 (2002).

[17] Wells with constant areal densities yield similar results in the parameter range studied.

[18] J.-M. Jancuet al., Phys. Rev. B72, 193201 (2005).

[19] S. Giglbergeret al., Phys. Rev. B75, 035327 (2007).

[20] T. Kogaet al., Phys. Rev. B74, 041302(R) (2006).

[21] C. M. Huet al.[Phys. Rev. B60, 7736 (1999)] find a larger Rashba constant in the second subband.

[22] E. Shafiret al., Phys. Rev. B70, 241302(R) (2004).

[23] H. K. Nget al., Appl. Phys. Lett.75, 3662 (1999).

[24] E. Bernardeset al., Phys. Status Solidi C3, 4330 (2006).

[25] L. Senaet al., Phys. Rev. B72, 235309 (2005).

[26] V. Sihet al., Nature Phys.1, 31 (2005).

[27] E. I. Rashba, Phys. Rev. B70, 201309(R) (2004).

[28] J. Sinovaet al., Phys. Rev. Lett.92, 126603 (2004).

[29] M. Leeet al.(to be published).

[30] J. Sinova et al., Solid State Commun. 138, 214 (2006).

[31] V. A. Guzenko et al., Phys. Status Solidi C 12, 4227 (2007).

PRL99,076603 (2007) P H Y S I C A L R E V I E W L E T T E R S week ending 17 AUGUST 2007

076603-4

Referenzen

ÄHNLICHE DOKUMENTE

Therefore dynamical coupling produces a nonvanishing out-of-plane spin polarization with an oscillating time pattern resulting from the mixing of the static eigenstates such that

We calculate the electrically induced spin accumulation in diffusive systems due to both Rashba 共 with strength ␣兲 and Dresselhaus 共 with strength ␤兲 spin-orbit

We show however that the singularity is broadened and that the suppression of spin accumulation becomes physically relevant (i) in finite- sized systems of size L, (ii) in the

The interplay of the linear Bychkov-Rashba and Dresselhaus spin-orbit interactions drastically affects the plasmon spectrum: the dynamical structure factor exhibits variations

In conclusion, we studied the combined effect of long- range and magnetic disorders on voltage induced spin polar- izations and the related spin Hall currents in a Rashba 2DEG..

We considered the semiclassical description of ballistic transport through chaotic mesoscopic cavities in the presence of spin-orbit interactions.. Our focus was the calculation

Institut für Theoretische Physik, Universität Regensburg, D-93040 Regensburg, Germany 共 Received 16 May 2007; revised manuscript received 17 July 2007; published 2 November 2007 兲

This effect amounts in spin accumulation as a response to an applied in-plane electric field and is therefore a possible key ingredient towards all-electrical spin control