Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Adiabatic approximation for a uniform DC electric field

Zhuo Bin Siu and Mansoor B. A. Jalil


Computational Nanoelectronics and Nanodevices Laboratory,
Electrical and Computer Engineering Department, National University of Singapore, Singapore

Seng Ghee Tan


Data Storage Institute, Agency for Science, Technology and Research (A*STAR), Singapore

In this work, we show that the disorder-free Kubo formula for the non-equilibrium value of an ob-
servable due to a DC electric field, represented by Ex x in the Hamiltonian, can be interpreted as the
standard time-independent theory response of the observable due to a time- and position-independent
perturbation HM F . We derive the explicit expression for HM F and show that it originates from
the adiabatic approximation to k|Ex x in which transitions between the dierent eigenspinor states
of a system are forbidden. The expression for HM F is generalized beyond the real spin degree of
freedom to include other spin-like discrete degrees of freedom (e.g. valley and pseudospin). By
direct comparison between Kubo formula and the time-independent perturbation theory, as well as
the Sundaram-Niu wavepacket formalism, we show that HM F reproduces the eect of the E-field,
i.e. Ex x
, up to the first order. This replacement suggests the emergence of a new spin current term
that is not captured by the standard Kubo formula spin current calculation. We illustrate this via
the exemplary spin current for the heavy hole spin 3/2 Luttinger system. Finally, we apply the
formalism and derive an analogous HM F for the eects of a weakly position-dependent coupling
to the spin-like internal degrees of freedom. This gives rise to an anomalous velocity as well as
spin accumulation terms in spinpseudospin space in addition to those contained explicitly in the
unperturbed Hamiltonian.
2

I. INTRODUCTION

In the Spin Hall Eect (SHE) [15], the passage of an in-plane electric field in a two-dimensional electron gas
(2DEG) with spin orbit coupling (SOC) leads to the emergence of an out-of-plane spin accumulation. Murakami [6]
and Fujita [711], and their respective coauthors, had independently studied the SHE. They showed that the out-of-
plane spin accumulation can be understood as the response of the charge carriers as their spins align adiabatically
with the momentum-dependent SOC field. The direction of the SOC field changes in time due to the change in the
momentum of the charge carriers as they accelerate under the electric field. Mathematically, the electric field gives
rise to an eective out-of-plane magnetization term perpendicular to the SOI field in the Hamiltonian.
The usual derivation [711] of this eective magnetization involves a unitary transformation from the laboratory
frame, where the spin quantization axis is conventionally taken to be an arbitrary fixed z axis, to the eigenbasis frame
where the spin quantization axis now points along the SOC field. For concreteness, consider the spin 1/2 Hamiltonian

H(k) = B0 (k)I + B(k) + Ex x


(1)

where B0 (k)I are the spin-independent terms, B(k) is a momentum dependent SOC field that may also possibly
have momentum independent components (for example, a Rashba SOC with a uniform x magnetization Mx would
k) = (ky +Mx , kx )), and Ex x
give B( represents an electric field in the x direction. We denote the mth eigenspinor
k) as |m (k).
of B(
In the unitary transformation formalism, we consider a unitary transformation U which diagonalizes B( k) so

that U B(k) U = |B|z . Now since U is momentum dependent due to the momentum dependence of B(k), we have

U H(k)U = B0 (k)I + |B( x + iU kx U )


k)|z + Ex ( (2)

where there is now an additional iU kx U term due to the non-commutativity between the momentum and position
operators. iU kx U has diagonal spin components, which point along the same spin direction as B, as well as o-
diagonal components which point in a spin space direction perpendicular to B. The latter is usually identified as
the source of the out-of-plane spin accumulation in the SHE when B is the in-plane SOC field. This, however, raises
the issue of how we can reconcile the o-diagonal elements of iU kx U with employing an adiabatic approximation.
The o-diagonal elements of iU kx U correspond to transitions between eigenstates of B whereas the adiabatic
approximation is usually associated with forbidding transitions between eigenstates. One of the two main motivations
of this work, then, is to clarify how this unitary transformation formalism is actually consistent with taking an
adiabatic approximation.
Besides the emergence of spin accumulation, another hallmark of the SHE is the appearance of currents of out-
of-plane spin flowing in the in-plane direction perpendicular to the applied electric field. One common method to
obtain these spin currents is by using the Kubo formula [12] which, in the absence of impurity scattering, gives the
non-equilibrium value of an observable O due to a uniform DC electric field in the ith direction as
Imm (k)|O|m (k)m (k)|ji |m (k)
O/Ei = (nm ,k nm,k ) . (3)
(em,k em ,k )2
k,m =m

where nm,k and em,k are the Fermi-Dirac occupancy and energy of the |m (k) state respectively. Not surprisingly,
putting O = z gives consistent results for the out-of-plane spin accumulation given by the unitary transformation
formalism described earlier. Moreover, some of us ( S. G. T. and M. B. A. J.) have pointed out in our earlier works
[8, 10, 11] that the Berry curvature anomalous velocity in the ith direction (ki bkj b) bEj , b B/|
B| that is more
commonly derived by putting O = vi into the Kubo formula Eq. 3, or by applying the Sundaram-Niu wavepacket
formalism [13] can also be obtained from the Heisenberg equation of motion for x i = i[ by including a term
xi , H]
proportional to the o-diagonal terms of iU ki U into H.
The second motivation of this paper is therefore to clarify
the link between the unitary transformation formalism and the Kubo formula.
We claim that to first order in perturbation theory, spin accumulation and velocities can be predicted by replacing

Ex x in Eq. 1 by HM F Ex m,m =m |m (k)ikx m (k)|m (k)m (k)|. (The subscript M F stands for Murkami-
Fujita.) After introducing the notation we shall be using in this paper, we first motivate the introduction of HM F by
showing that the Kubo formula Eq. 3 can be interpreted as the first-order expectation value of the observable under
the perturbation of HM F . We next show that HM F originates from taking the adiabatic approximation of retaining
only the diagonal elements in the eigenspinor basis representation of the position operator U x U in Eq. 2. This
discarding of the o-diagonal terms physically corresponds to preventing transitions between dierent eigenspinor
3

states. We then show that the results of replacing Ex x with HM F in the Hamiltonian and applying the Heisenberg
equation of motion reproduces the same results as the Sundaram-Niu wavepacket formalism [13] for the spin evolution
and charge current. The wavapacket formalism provides a physical justification for discarding Ex x in our eective
Hamiltonian.
The addition of HM F into the eective Hamiltonian suggests the consideration of a new contribution for the spin
current flowing in the jth direction of the |m (k) state 21 m (k)|{ , i[xj , HM F ]}|m (k) that is not captured by
the usual practice of putting O = 12 { , i[xj , B0 (k)I + B(
k) ]} into the Kubo formula Eq. 3. We evaluate this
spin current for the Luttinger spin 3/2 system.
Using the same formalism we used to show the origin of HM F due to an electric field Ex x , we next derive the
analogous HM F ;mag for the perturbation due to a linear variation of the magnetization (x M ) x . We show that
the resulting HM F ;mag also results in a spin-dependent anomalous velocity and spin accumulations.

II. SYSTEM DEFINITION

Here we study translation-invariant systems with momentum-dependent spin-orbit interactions and possibly other
discrete internal degrees of freedom perturbed by a uniform DC electric field. For example, many emerging material
systems of interest in spintronics like silicene [1417] and MoS2 [1820], possess discrete degrees of freedom such as
the pseudospin and / or valley degrees of freedom, in addition to their real spins. Another example of a system with a
discrete internal degree of freedom is a topological insulator (TI) thin film which, unlike a semi-infinite thick TI slab,
possesses both a top as well as a bottom surface where the surface states localized at dierent surfaces can couple to
one another due to the finite thickness of the film [2123]. The low energy eective Hamiltonian for the surface states
of a TI thin film can thus be written as

H = v(k ) zz + x + M
+ Ez z (4)

where the i s and i s are Pauli matrices. Besides the real spin of the charge carriers, denoted as , there is another
discrete degree of freedom associated with whether the charge carriers are localized nearer the upper ( | + z +z |
) or lower (| z z |) surface of the film. The x term then represents the coupling between the two surfaces of
the film due to the finite thickness, and the Ez z term the spin-independent potential energy dierence between the
top and bottom surfaces. The M term represents the exchange coupling to either ferromagnetic dopants or an
adjoining ferromagnetic film.
For simplicity, we collectively refer to all of the discrete internal degrees of freedom other than the real spin as
pseduospin, and denote the corresponding operators as . We enumerate all the possible combinations of the i j
operators (including I I I ) as k . For instance, for the TI thin film Hamiltonian Eq. 4 we may define
i I i for i = 1, 2, 3, 4 I I , i x i4 for i = 5, 6, 7, etc.
We thus consider systems of the form

H0 = Bi (p)
i
H = H0 + E x .

where we have placed hats on top of p and x to emphasise that these are operators. In the absence of the electric
field, the translation invariance of the system leads to the momentum being a good quantum number. For notational
simplicity we temporarily restrict ourselves to working in one dimension and drop the vector arrow on top of k. (The
extension to multiple spatial dimensions is trivial)
Since momentum is a good quantum number, it is common to consider

(k| I )H0 (
p) = Bi (k)(k| i ) (5)

and call H0 (k) Bi (k)i the Hamiltonian instead, for example as in Eq. 1. Notice that the operator p has been
demoted to the numerical eigenvalue k, and the k| on the right hand side is usually not written out but implied
implicitly. The formal mathematical definition of H0 (k) will however turn out to be important when we consider the
expansion of the Ex x perturbation in the eigenspinor basis later in Sect. IV.
For a given momentum k, we denote the eigenstates of H0 as |k, m |k |m (k) where |k is the k momentum
state ket, and |m (k) is the mth eigenspinor of H0 at momentum k. The corresponding eigenenergy ek,m satisfies
p)|k, m = |k, mek,m .
H0 (
4

III. KUBO FORMULA

We briefly run through the textbook derivation of the Kubo formula Eq. 3. The purpose of this overview is two-
fold. We first wish to bring across the fact that the derivation is somewhat mathematically involved and not very
physically transparent. We will show in this, and the next section, that Eq. 3 can be interpreted more transparently
as giving the expectation value of O due to the perturbation by the adiabatic approximation of Ex x
. We secondly also
want to highlight one aspect of the derivation that is not very commonly discussed but which justifies our eventual
replacement of Ex x with HM F .
The textbook derivation of the Kubo formula starts from linear response theory. Linear response theory tells us
that the non-equilibrium value of the expectation value of an observable O due to a uniform DC electric field of
amplitude Eb in the bth direction is given by
i
O = lim Eb COj b ()
0
(i 1 )

= Eb lim Tr(OG(i(n + q)j b G(in ))
0 q+i
n

where COj b is the retarded correlation function between the observable O and the charge current j in the bth direction,
the Gs are complex frequency Matsubara Greens functions, and the sum n occurs over Matsubara frequencies. The
limit of taking 0 corresponds to first assuming that the electric field is actually AC with frequency and then
taking the DC limit of the electric field.
Using the standard summation formula 1 in n(in ) exp(in ) = j Res[f (zj )]n(zj ) exp(zj ), where zj are
the complex poles of the Fermi-Dirac distribution function n, to evaluate the complex frequency summation lands us
at
( )
i n n b
O = lim O, j, Eb . (6)
0 e e + + i
,

where and are collective labels for the quantum numbers (k, m) of the eigenstates of H0 (which does not include
), |k, m, and O, |O|. The 0 limit can be evaluated by exploiting the fact that
the Ex x
1 1 1 1 1
= ( )
+x x +x
to split the summand into a divergent part which we discard, and a non-divergent part which we retain. We will have
a bit more to say about the discarding of the divergent part later on. The non-divergent part in Eq. 6 is
n n
i b
O, j, Eb (7)
(e e )( + i + e e )
where it is now safe to put 0 explicitly. This gives the standard Kubo formula
n n ( )
b
O = Im O , j Eb . (8)
(e e )2 ,
=

In order to relate this to our results later and considering just the one-dimensional case for notational simplicity,
we use the fact m |j|m = k ek,m mm + k m |m (em em ) to obtain
n(ek,m ) n(ek,m )
O = Im(m (k)|O|m (k)k m (k)|m (k))E.

ek,m ek,m
k m=m

To isolate the contribution of a specific |k, n state to O, Ok,n we can set the occupancy factor for this state to 1
and all other states to 0. Note that there are actually two sets of terms in the sums above which contribute one
where m = n, m = n and m = n, m = n. The contributions of these two sets are equal. We thus have
2
Ok,n = Im(n (k)|O|m (k)k m (k)|n (k)E

ek,m ek,n
m =n
|m (k)ik m (k)|n (k)
=2 Ren (k)|O| E. (9)
ek,n ek,m
m =n
5

The derivation above is not very intuitive and is mathematically complicated with the use of complex frequencies
and first assuming an AC electric field and then taking the limit of the electric field frquency going to 0. However
Eq. 9 admits a more physically intuitive interpretation. In standard time-independent perturbation theory the first
order correction to the state |m due to a perturbation V , which we denote as |1m , is given by
|m m |V |m
|1m = . (10)

e m e m
m =m

Denoting the exact eigenspinor of H0 (k) + V as |


m , the expectation value of O is

m |O|
m m |O|m + 2Re(m |O|1m )

|m m |V |m
= m |O|m + 2Re m |O| (11)

ek,m ek ,m
m =m

to first order in V . Comparing Eq. 11 to Eq. 9, it is evident that the corresponding V in Eq. 9 would be

V = HM F Ex |n (k)ik n (k)|n (k)n (k)| (12)
n,n =n

This suggests that the eects of the Ex x


term in the Hamiltonian on those observables which can be obtained by
the Kubo formula can be reproduced, to first order in Ex , by replacing Ex x
with HM F . This invites the question of
where HM F comes from. We address this in the next section before showing that replacing Ex x
with HM F reproduces
the current and spin dynamics predicted by the Sundaram-Niu formalism.

IV. ORIGIN OF HM F

We pointed out at the end of Sect. II that Hamiltonian, H0 (k), actually corresponds to k| I H0 (
p). The addition
of the perturbation E x presents the question of what form the corresponding k| I x will take.
To answer this question, we note that the spinpseudospin identity I can be resolved in any set of orthogonal
basis states. For concreteness, let us take the example that the I s correspond to spin 1/2 operators. One set of basis
states that can be chosen is the spin up and down states, the up and down defined with respect to a constant spin
quantization axis. The basis states are hence momentum independent. We shall, for brevity, refer to this set of basis
states as the laboratory frame basis and where we need to refer to these states explicitly, always write the indices as
i ( for l aboratory) with a subscript i to dierentiate between dierent basis states. Another obvious choice of basis
states to resolve I into is in terms of the eigenspinors of H0 (k), |m (k), which unlike the laboratory frame basis,
are momentum dependent. We show in the appendix that

k|x= |m (k)m (k)|ik m k, m | + |m (k)ik k, m| (13)
m,m m

We note that if we had expanded the spin identity operator in the momentum-independent laboratory basis |i s
instead, Eq. 13 reduces to the more familiar result

k|
x = I (ik k|) (14)

There is a bit of technical subtlety involved in relating Eq. 13 to the corresponding (x + iU kx U ) term in Eq. 2
where the ket-bra pairs in the operators are not written out explicitly. There, one needs to be mindful of whether the
resulting expressions are in the laboratory frame or in the eigenbasis frame. In Eq. 2, we made use of

U = x
Ux + iU k U . (15)

in which the unitary transformation on the left hand side of the equal sign brings the x operator in the laboratory
frame to the eigenspinor frame on the right hand side. Since this is a just a basis transformation U is actually
just an identity matrix and we read o that U = m,i |i i |m (k) m (k)|. The tilde that appears on the

k in Eq. 15 reflects the subtlety that the k within U k U in Eq. 15 is understood to act only on the matrix
elements i |m (k) and not on the outermost ket-bra pair |i m (k)| sandwiching the matrix elements. That is,

ik U m,i |i k (i |m (k))m (k)| the rightmost m (k)| bra is not dierentiated. This comes from the
6
( )
|+ |
convention that in practical calculations the U are written as numerical matrices, i.e. U = where
|+ |
are the eigenspinor states and / are the lab spin up / down states, so that the only numbers that appear explicitly
to be dierentiated are the matrix elements.
The correspondence of Eq. 14 with Eq. 13 is that the second term on the right of Eq. 13 can be further expanded
(by adding resolutions of identity) to yield
( ) (
|m (k)ik k, m| = i (k k|) I k| |m m |k m m | (16)
m m,m

Adding this to the remaining m i|m (k)m (k)|k m k, m | = i|k m k, m | term in Eq. 13 leads to the recovery
of Eq. 14. As a consequence Eq. 15 should really be understood as
x e (k) + iU k U
l (k) = x

|kik k| I = |k, mik k, m| + i |k, mk| m |k m m |
m m,m

= |k, mik k, m| i|k, mk| k m |
m

where we defined x l (k) and x e (k) in the second line (with the subscripts l and e for l ab and eigenspinor frames
respectively, and the k argument stressing that this is actually the |kk| projection of the position operator ). Notice
that position operator x in Eq. 15 actually has dierent definitions in the laboratory ( xl (k)) and eigenbasis (
xe (k))
frames. Compared to x l (k), x
e (k) has an extra ik|k m (k)| term coming from the ik dierentiating the momentum
dependence of the basis ket. The iU k U that appears on the right of Eq. 15 can now be understood as canceling o
the contribution of this extra term.
Eq. 13 now allows us to tackle the question of where HM F comes from. In the spirit of the adiabatic approximation
where we forbid transitions between dierent eigenspinor states, we retain only the diagonal eigenspinor terms of the
iU k U piece in Eq. 13 in the adiabatic approximation to x -
( )
k|xad |m (k)im (k)|k m k, m| + |m (k)ik k, m|
m


compared to Eq. 13 we dont sum over the m = m terms in the first piece. ) Substituting in the expansion
(Note that
of the m |m (k)ik k, m| from Eq. 16 gives

k|
xad = ik k| I + k| |m ik m |m m |. (17)
m,m =m

The first piece on the right ik k| I is just k| x while the latter piece is exactly the HM F (without the factor
of E and with an extra k|) in Eq. 12. This shows that HM F originates from taking an adiabatic approximation to
k|
x.
The above results can also be obtained from the common usage notation of Eq. 15 by splitting up the eigenspinor
frame iU k U = d + o into a d iagonal part d and o-diagonal part o with

d = |m m |ik m m |, o = |m m|ik n n |
m m,n=m

We have, from Eq. 15,

U (
xe + iU k U )U = x
l U ( =x
xe + d)U l U oU, (18)

so that

U (
xe + diag(iU k U ))U = x
l i |m m |ik n n |. (19)
m,n=m

The left hand side of Eq. 19 has the physical meaning of transforming the adiabatic approximation of k|
x from the
eigenspinorframe back to the laboratory frame. With reference to Eq. 12, the second term on the right hand side of
Eq. 19 i m,n=m |m m |ik n n | is HM F /Ex as expected.
7

We therefore conclude this section by summarizing how the results in this and the preceding sections have addressed
our motivations in the introduction.
(i) We raised the issue of how taking the o -diagonal elements of iU k U in Eq. 2 is consistent with taking an
adiabatic approximation. We now see from Eq. 18 that an eective magnetization corresponding to the opposite
direction of the o-diagonal terms of iU k U comes from retaining only the diagonal elements of iU k U in the
eigenbasis frame. The key point to note in establishing this equivalence is that we need to transform the eigen-basis
frame x e + diag(iU k U ) back to the laboratory frame and not just diag(iU k U ) alone.
(ii) With the above formalism, we can establish the link between the Kubo formula and the unitary transformation
formalism. We found that the Kubo formula can be interpreted as the first order perturbation response to the
adiabatic approximation of the position dependence of the electric potential, which we termed as HM F .

V. SUNDARAM-NIU WAVEPACKET FORMALISM

We next demonstrate that the Heisenberg equations of motion where Ex x is replaced by HM F reproduces the
dynamics of the position and spin evolution predicted by the Sundaram-Niu wavepacket (SN) formalism [13] and
the extension of the latter by Cheng and Niu [24] to include spin. Physically, the SN formalism revolves around
a wavepacket localized about position x0 and momentum k0 . This provides an avenue for us to study quantum
states which have, in a sense, defined momentum and position values despite the Heisenberg uncertainty principle.
Operationally, the position localization oers an elegant solution for studying the eects of perturbations which
have the form of Ex x , which corresponds to, for example, the potential induced by a uniform DC electric field. In
applying the standard first-order time-independent perturbation theory in a SOI system one needs to evaluate matrix
elements like k, m|
x|k, m. One might naively conclude that such a term might diverge as x . The real space
localization of the wavepacket around x0 oers the reassurance that we need only take into account the eects of E x

at the immediate neighbourhood of x0 thus avoiding the divergence at |x| .
In Cheng and Nius extension of the SN formalism to include spin, the wavepacket state | is given by

| = |k, mm a(k).
m


Here the i s are the normalized weight of the ith eigenspinors such that i i i = 1. i is a dynamic variable
much like x and k while a(k) is an envelope function peaking around k = k0 where it has the form exp(i(k)). The
requirement that the wavepakcet is localized around x0 , |
x| = x0 , imposes some conditions on (k). Relegating
the details of the derivation to the appendix, the time evolution of the expectation values of the spin and position
operators when H0 does not have explicit position or time dependence is

t (i ij j ) = k a (l l |ka i ij j i ij j |ka l l ) (20)



x a = (ka ek,m )m m + ikb m (kb m |ka m ka m |kb m )m (21)

where ij i (k)||j (k).

VI. VALIDITY OF HM F .

We have seen in the previous section that the Kubo formula can be interpreted as yielding the expectation values
of observables under the perturbation of HM F . Based on Eq. 12, this motivated the consideration of

He (k) H0 (k) + Ex |m (k)ikx m (k)|m (k)m (k)| (22)
m,m =m

as the eective Hamiltonian for capturing the eects of the electric field at a specific value of k for the original
Hamiltonian H = H0 (p) + Ex x
. We have shown that the expectation values of observables calculated with He will,
by construction, match those calculated by the Kubo formula to first order in Ex . We now show that applying the
Heisenberg equations of motion on Eq. 22 will also reproduce the dynamics of the position and spin operators as
predicted by the SN wavepacket formalism in Eqs. 21 and 20 respectively.
8

Now returning to multiple spatial dimensions, the Heisenberg equation of motions are given by (for notational
simplicity we temporarily denote |m (k) simply as |m)


m|x i |m = ... + m| i[xi , |aa|ikj bb|] |mEj
a=b

= +i(ki m|kj m + kj m|ki m)Ej

Noting k j = Ej with a minus sign, this result agrees with the generalized Sundaram-Niu result of Eq. 21.
Similarly, if we compare the Heisenberg equation of motion for a spin operator = ab |aab b| we have


m||m
= ... + m| i[, |aa|ikj bb|] |mEj
a=b

= ... + (ma a|kj m m|kj aam )Ej

which agrees with the generalized SN equation for spin evolution, i.e. Eq. 20. The agreement of the dynamics of
spin and position of He with the SN wavepacket approach provides a physical reason for why it may be justifiable
to do away with the operator x in Eq. 22. The localization of the wavepacket around x0 (which we may set to
0), implies that the wavepacket only feels the influence of position-dependent terms (such as the electric field) at
and around the immediate neighbourhood of x0 . The irrelevance of the position dependent terms away from x0
makes it unnecessary to include the entire position-dependent potential inside the eective Hamiltonian (recall that
x
= dx|xxx| is actually the integral of position eigenvalues and position projectors over all sapce) except at x0 .
The fact that influence of Ex x away from x0 = 0 may be ignored, which would be divergent if we had considered
all points in space, is perhaps also reflected in the discarding of the divergent 1 , 0 term in the lead-up to Eq.
7. HM F can then be interpreted as accounting for the eects of the gradient of the position dependent potential at
x0 = 0.
In the remainder of this paper, we show that Eq. 22 oers three advantages over the direct use of the Kubo formula
Eq. 6 .
i. Eq. 22 suggests the possibility of new spin current terms that are not captured by the Kubo formula, as we shall
proceed to elaborate on in the next section.
ii. The usual Kubo formula Eq. 6 may break down when there are degeneracies because of the ea eb factor in
the denominator. Eq. 22 validates the use of the degenerate perturbation theory prescription of finding the linear
combinations of degenerate eigenstates |m s for which the m |V |n = 0 when m |H0 |m = n |H0 |n for
m = n.
iii. The reasoning behind the construction of Eq. 22 suggests a similar approach to constructing an eective
Hamiltonian for the eects of a spatially varying magnetization. We describe this in Sect. VIII.
We demonstrate points (i) and (ii) in the next section where we discuss spin currents in the Luttinger heavy hole
system, and point (iii) in the section which follows thereafter.

VII. MURAKAMI-FUJITA SPIN CURRENT

In the previous section, we have shown that m | i[O, HM F ]|m matches the results of the Sundaram-Niu
wavefunction formalism for the operators O = x
, . This suggests that in general, there is a contribution to the spin
current linear in E of the form
1
ji(M F)
k,m m (k)| { , i[xi , HM F ]}|m (k). (23)

2
This is in addition to the usual term [12] calculated from the Kubo formula

1
ji(Kubo) k,m = 2Rem (k)| { , i[xi , H0 ]}|1m (k). (24)

2

(As a recap |1m was defined in Eq. 10 as the first order correction to |m due to the perturbation of HM F . )
This additional spin current contribution has been derived previously for the case of Rashba SOC [11]. We now
evaluate this new spin current for the spin 3/2 Luttinger system which was studied previously by Murakami [6].
9

The Hamiltonian for an isotropic spin 3/2 Luttinger system is given by


( )
1 5
H= (1 + 2 )k I 22 (k S)(k S)
2
2m 2

where 1 and 2 are material dependent parameters and S are the spin 3/2 operators.
The eigenstates consist of two pairs of degenerate energy states, the light hole states |L1 and |L2 with eigenenergy
k2 k2
2m (1 + 22 ) and heavy hole states |H1 and |H2 with eigenenergy 2m (1 22 ). The m indices in the notation
|m which we have been using in the previous sections hence run over m = {L1, L2, H1, H2}. Murakami used
the unitary transformation U = exp(iy ) exp(i z ) to transform x from the lab frame to the eigenspinor frame
l U x
x U = x
e + iU k U . Recalling that U = m,i |mm|i i |, we read o from the matrix exponential U that
3i ( ) 3i ( )
e 2 cos3 2 e 2 sin3 2
1 i ( ) 1 i ( )
3e 2 sin2 () csc 2 3e 2 sin 2 sin()
|H1 = 4 1 i () , |H2 = 2 1 i ( )
2 3e 2 sin 2 sin() 4 3e 2 sin2 () csc 2
3i ( ) 3i ( )
e 2 sin3 2 e 2 cos3 2
3i ( ) 3i ( )
14 3e 2 sin2 () csc 2 1
3e 2 sin 2 sin()
1 i ( ( ) ( )) 2
1 i ( ( ) ( ))
e 2 cos + 3 cos 3 2 )) e 2 sin 2 3 sin 3
|L1 = 4 1 i ( ( 2 ) ( 3 , |L2 = 41 i ( () ( 32 )) (25)
4 e 2 sin 2 3 sin 2 4 e 2 cos 2 + 3 cos 2
1
3i (
) 1
3i
2
(
)
2 3e
2 sin
2 sin() 4 3e
2 sin () csc
2

where (k, , ) are the k-space spherical coordinates.


Following the prescription of Eq. 12, HM F for an electric field in the z direction reads (in the usual spin 3/2
laboratory spin z basis)
i
0 i 3e 0 0
Ez 3e 0 2ei 0
HM F = i sin()


i .
0 i 3e
i
2k 0 2e
0 0 3e 0

It can be shown straightforwardly that choosing alternative linear combinations of the degenerate states (i.e. by
substituting |L1 = |L11 + |L21 , |L2 = |L12 + |L22 , expanding and enforcing orthonormality between |L1
and |L2 , and likewise for the heavy hole states) does not change the final expression for HM F .
Defining
kF
Ji(MF)

m dk dd m (k)|ji(M

F)
|m (k)
0

i(M F )
with m |j |m defined in Eq. 23 gives, after straightforward calculations for an electric field in the z direction,

Jx(M F)
x ,z
LH,HH = Jy(M F)
y ,z
LH,HH = 0,
Jz(M F)
x ,y ,z
LH,HH = 0,
3e
Jx(M F)
HH = Jy(M F)
HH = Ez kf ,
y x
~
e
Jx(M F)
LH = Jy(M F)
LH = Ez kf
y x
3~
where we now write out e and ~ (which have previously been set to 1) explicitly to show that the results are
dimensionally consistent. We will now show that the above spin current terms dier from those obtained via Kubo
i(Kubo)
formula J m . As a side-note, in implementing the Kubo formalism, the naive substitution of the eigenstates
Eq. 25 into the Kubo formula Eq. 9 is problematic because the degeneracy of the eigenstates leads to the presence of
terms in which the numerator, proportional to the HM F induced transition between degenerate eigenstates, is finite
but the denominator, equal to the energy dierence between the degenerate states, is zero. Now that we understand
that Eq. 9 is nothing but standard time-independent perturbation theory applied for the perturbation of HM F , we are
justified in applying degenerate perturbation theory. Specifically, for each pair of degenerate states say |L1 and |L2,
we diagonalize the projection of HM F in the basis of the degenerate states, in order to find the linear combinations
10

of the degenerate states |L1


= |L11 + |L21 , |L2 = |L12 + |L22 for which L1|H M F |L2
= 0 and similarly
for the heavy hole states. Using the |L
i states instead of the original linear combinations in Eq. 25 ensures that the
numerator in the Kubo formula is zero when the denominator is, so that these terms in the Kubo formula can be
ignored.
In implementing the above prescription, we find that the |H1 and |H2 states in Eq. 25 are already diagonal with
respect to HM F . The linear combination of the light hole states diagonalizing HM F are

1
12 i 32 e 2 i(3) sin()

1
i(1+3e2i )e 2 i(+)

=
|L1 4 2 ,
1 i(+)
e2 (2 cos()+i sin())

2 2
1 3 23i (
) ( (
) )
2 2e sin 2 sin() cot 2 + i

1 3 12 i(+3)
2i 2e sin()

1
e 2 i(+) ( sin()2i cos())

=
|L2 2 2 .
2i
(3+e )e 2
1
i(3)

4 2
3i ( ) ( ( ) )
1
2
3 2
2e sin 2 sin() cot 2 i

i(Kubo
We then find, after straightforward but tedious calculations, that the only non-zero j m s are

eEz kf
jx(Kubo) HH = jy(Kubo) LL = jy(Kubo) HH = jx(Kubo) LL = (1 + 22 ).
y x x y
2~2

VIII. POSITION-DEPENDENT DEGREE MAGNETIZATION

We next move on to study the eects of a weak position dependence of the coupling to the internal degrees of
freedom in the form M ( x)i i which we shall collectively refer to as the magnetization (in analogy with Mi i for
i being the real spin) for short. We assume that the position dependence is weak enough for a first order gradient
expansion of the magnetization to suce

H = B(k)i i + Mi (0)i + (x Mi )i x
.

The B(k)i i is the momentum-dependent SOC term and the Mi (0) are the magnetizations at the fixed x=0 and
are position-independent. We take H0 = (B(k)i + Mi (0))i , and treat (x Mi )i x as the perturbation. Note that
this is a dierent, but not necessarily mutually exclusive, limit from the more often studied case (Ref. 10 and
references therein) where the SOC is treated as a perturbation to the position dependent magnetization under the
strong magnetization limit.
Analogous to what we have done in the previous section, we now seek an expression for
( )


k|
xi = k, m|
xi dk |k , m k , m | (26)
m m

where the |k, ms are the eigenstates of H0 .


x|k , m , it can be shown that (for brevity, we
Employing a similar series of steps in our earlier derivation of k, m|
have dropped the subscript i in )
k, m|x|k , m
= i(k ((k k)m (k)||m (k )) (k k)ik m (k)||m (k ) (27)
= ik ((k k)m (k)||m (k )) + (k k)m (k)||ik m (k ) (28)
Using the expansion of Eq. 27 for k, m| x|k , m gives

dk |k, mk, m|
x|k , m k , m | = |k, m(m ||ik m k, m | + m ||m ik k, m |). (29)
11

Notice that each of the two terms on the right hand side are not individually Hermitian by themselves. The remedy
to this is to take half of the right hand side of Eq. 29, and add this to half of the left hand side of Eq. 29 evaluated
using the expansion of Eq. 28. The latter is (half of) the Hermitian conjugate of Eq. 29. The symmeterized version
of Eq. 29 thus reads

dk |k, mk, m|
x|k , m k , m |
1(
= (|k, m(m ||ik m ik m ||m )k, m | +
2 )
(|ik k, mm ||m k, m | + |k, mm ||m ik k, m |) (30)
We now take the adiabatic approximation of retaining only the diagonal terms in the eigenbasis. Diering from our
earlier experience with the expansion of dk |k, mk, m|x|k , m k , m | where the terms that come with the ik |
and |ik are already diagonal in the eigenspinor basis, the corresponding two terms here in the last line of Eq. 30
do contain o-diagonal elements. This raises the question of whether we should retain the m = m elements in these
terms.
If we retain m = m terms in the terms containing |ik k, m and its c.c. in the last line of Eq. 30 , we have

( )
|k, m m ||ik m k, m| + m ||m ik k, m |) + h.c.
m m

= (|kik k| |k, mm ||ik m k, m |) + h.c.
m=m

We treat the second term and its Hermitian conjugate as the perturbation

M F ;mag = 1 (x Mi )
H (|k, mm ||ik m k, m | + h.c.).
2 m=m

Comparing this against Eq. 12 for an electric field, one can interpret (xi M )i x
as a spin dependent electric field.
One could, however, obtain a closer agreement with the predictions of the Sundaram-Niu wavepacket formalism by
discarding the m = m terms in the last line of Eq. 30 in keeping with the spirit of the adiabatic approximation of
forbidding transitions between dierent eigenstates. We obtain

1 1 ( )
(k|(
xi )ad + h.c.) = |k, m m ||ik m k, m| + m ||m (ik m |k| + m |ik |) + h.c. . (31)
2 2 m

This is the analogue of Eq. 17, symmeterized to ensure that the term remains Hermitian. The 12 |m m |i |m m |
(|kik | |ik k|) term in Eq. 31 (the last term in Eq. 31 just to the left of the h.c. and its conjugate) ) is
analogous to ik k| I in Eq. 17. This analogy is in the sense that 21 |m m |i |m m | (|kik | |ik k|)
can be interpreted
into the eigenspinor basis of H0 , just like ik k| I is
as the projection of the spin part of x
actually ( m (|m m |I|m m |) ik k|. We treat the remaining terms in Eq. 31,
x M
HM F ;mag = |k, m(m ||ik m k, m| + m ||m ik m |k| + h.c.). (32)
2
|k,nn |HM F ;mag |m
as perturbations to |k, m and calculate the perturbed states |k, m1 mag = n=m en,k em,k . The first order
correction to the velocity due to HM F ;mag , 2Re(k, m| i[x, H0 ]|k, m1 mag ) then gives the SN velocity Eq. 21 with
k |k,m = (x Mi )k, m|i |k, m. This is consistent with treating (x Mi )i x
as a spin-dependent electric field.
Besides giving a velocity contribution, the two terms containing an outer ket-bra pair of the form |m k m |
and its h.c. in Eq. 32 may also give an eective magnetization pointing in a dierent direction from the magne-
tization direction of H0 . The Hamiltonian for this eective magnetization, HM is given by the o-diagonal terms
of HM F ;mag , which consist of the second term of Eq. 32, x2M |k, m(m ||m ik m |k| + h.c.. with their spin
diagonal components, i.e. the projection of |m k m | onto |m m | subtracted away. We obtain

1
HM = (x Mi ) m |i |m (|m ik m | |m ik m |m m | + h.c.)
2 m
1
= (x Mi ) (|m m |i |m ik m |n n | + h.c.) (33)
2
m,n=m
12

Only Mx (0) non-zero


Spin accumulation / Magnetization gradient x y x z y z
z X X X
z z X X X X
Only My (0) non-zero
Spin accumulation / Magnetization gradient x x z
z X X
z z X X
Only Mz (0) non-zero
Spin accumulation / Magnetization gradient x y z x z y z z z
x X X X
z X X
x z X X X X
z z X X

TABLE I. A checkmark in the cell indicates that a finite spin a b accumulation (rows) exists after integrating over the Fermi
magnetization is present.
surface for a (x Mc d )c d gradient (columns) when only the stated component of the constant M

For a two-level system where the i s correspond to the Pauli matrices, the eective magnetization evaluates to
zero. However, in systems with multiple coupled internal degrees of freedom, such as the TI thin film, described by
Eq. 4, some physically relevant choices of the i do result in finite eective magnetizations which in turn result in
spin accumulations.
We illustrate this by numerically evaluating the spin accumulations which result from a x direction Mi gradient for
the Mi s coupling to a and a z . The (x Ma )a s correspond to x gradients in a direction magnetization exchange
couplings to the top and bottom thin film surfaces where the couplings on both surfaces have the same sign and
magnitude. The (x Ma;z )a z correspond to exchange couplings of the same magnitude but opposite signs on the top
and bottom surfaces. An exchange coupling with dierent magnitudes on the top and bottom surfaces (for example,
due to diering distances from the ferromagnetic layer on top of the TI thin film) is then given by (Ma + Ma;z z )a
with denoting the degree of asymmetry between the two surfaces. We integrate the spin accumulations in the a b
(b = (I , z )) resulting from the magnetization gradient integrated over the Fermi surface, and denote the integrated
spin accumulation as a b .
Tables I summarizes which combinations of the (x Mc,d )c d gradients result in finite a b accumulations
when only the x, y and z component of the constant Mi i magnetization in Eq. 4 has non-zero components. These
correspond to the three scenarios where the magnetization gradient in the x direction is parallel to the constant
magnetization, perpendicular to the constant magnetization and in the in-plane direction, and in the out-of-plane
directions respectively. A detailed analysis of the spin accumulations resulting from the magnetization gradients will
be presented elsewhere.

IX. CONCLUSION

In this work, we showed that the non-equilibrium expectation value of an observable O due to a DC electric field
given by the Kubo formula Eq. 3 can be interpreted as the standard time-independent perturbation theory response
of the observable O to the perturbation HM F . HM F originates from taking the adiabatic approximation to k| x
in which transitions between dierent eigenspinor states are forbidden. We then argued that to first order in the
electric field, the eects of Ex x
on the spin accumulation and currents can be captured in an eective Hamiltonian
where E x is replaced by HM F . We supported this claim by showing that the Heisenberg equations of motion with
this eective Hamiltonian reproduces the dynamics for spin and current predicted by the Sundaram-Niu wavepacket
formalism. The eective Hamiltonian suggests the existence of a new spin current term which is not captured by the
standard Kubo formula, and which we then evaluated for the Luttinger heavy hole system. Finally, using the same
approach used to obtain HM F for an electric field, we derived a corresponding expression for HM F ;mag resulting from
a spatially varying magnetization. We showed that HM F ;mag , similar to HM F , leads to an anomalous velocity as well
as an eective magnetization.
13

X. ACKNOWLEDGMENTS

We thank the MOE Tier II grant MOE2013-T2-2-125 (NUS Grant No. R-263-000-B10-112), and the National
Research Foundation of Singapore under the CRP Program Next Generation Spin Torque Memories: From Fun-
damental Physics to Applications NRF-CRP9-2013-01 (R-263-000-B30-281) and Non-Volatile Magnetic Logic and
Memory Integrated Circuit Device NRF-CRP9-2011-01 (R-263-000-A73-592) for financial support.

[1] M. I. Dyakanov and V.I. Perel, Phys. Lett. A. 35, 459 (1971).
[2] A. A. Bakun et al., Pisma Zh. Eksp. Teor. Fiz. 40, 464 (1984).
[3] J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).
[4] Y. K. Kato et al., Science 306, 1910 (2004).
[5] T. Kimura et al., Phys. Rev. Lett. 98, 156601 (2007).
[6] S. Mukami, N. Nagaosa and S.C. Zhang, Science 301, 1348 (2003).
[7] T. Fujita, M. B. A. Jalil and S.G. Tan, J. Phys. Soc. Jpn. 78, 104714 (2009).
[8] T. Fujita, M. B. A. Jalil and S.G. Tan, New J. Phys. 12, 013016 (2010).
[9] S. G. Tan and M. B. A. Jalil, J. Phys. Soc. Jpn. 82, 094714 (2013).
[10] T. Fujita et al., J. Appl. Phys. 110, 121301 (2011).
[11] S. G. Tan et al., Sci. Rep. 5, 18409 (2015).
[12] J. Sinova et al., Phys. Rev. Lett. 92, 126603 (2004).
[13] G. Sundaram and Q. Niu, Phys. Rev. B 59, 14915 (1999).
[14] S. Cahangirov et al., Phys. Rev. Lett. 102, 236804 (2009).
[15] C.-C. Liu, W. Feng and Y. Yao, Phys. Rev. Lett. 107, 076802 (2011).
[16] P. Vogt, Phys. Rev. Lett. 108, 155501 (2012).
[17] L. Tao et al., Nature Nanotech. 10, 227 (2016).
[18] R. Ganatra and Q. Zhang, ACS Nano 8, 4074 (2014).
[19] B. Radisavljevic et al., Nature Nanotech. 6, 147 (2011).
[20] H. Wang et al., Nano Lett. 102, 12, 4674 (2012).
[21] J. Linder, T. Yokoyama and A. Sudb, Phys. Rev. B 80, 205401 (2009).
[22] C.-X. Liu et al., Phys. Rev. B 81, 041307 (2010).
[23] H.-Z. Lu et al., Phys. Rev. B 81, 115407 (2010).
[24] R. Cheng and Q. Niu, Phys. Rev. B 88, 024422 (2013).

XI. APPENDIX

A. Derivation of k|
x

In order to obtain the expression for k| I x


in terms of a momentum dependent spin basis we first derive the
x|k , m
expression for k, m|

x|k , m
k, m|

= dxk, m|xxx|k , m

1
= dx m (k)|m (k ) exp(i(k k)x)x
2
( )
= i m,m k + m (k)|k m (k k)

With this expression, we can then derive

k|
x

( )
= |m (k) dk i m,m k + m (k)|k m (k k)k , m |
m,m

=i |m (k)m (k)|k m k, m | + |m (k)ik k, m|
m,m m
14

B. SN formalism with spin

(We note that the following derivations is not our original work and follow that in Refs. [13] and [24] with a minor
extension. The derivations here are simpler as we do not consider explicit time and position dependence in H0 . We
have chosen to go into more detail in some portions which have not been described as explicitly in the two references.
)
Consider the wavepacket state | of Eq. V. In order to enforce | x| = xc we first evaluate

|x| = dk dk a (k)m
k, m|x|k , m m
( )

= im (ik )mm + m |k m m .
from which we conclude that

k (m m ) = xc im m |k m m , (34)
a fact which we take note of for now and save for further use later.
We now proceed to construct the Lagrangian given by
L = |idt (H0 + H )|.
We thus need

( ( )
idt | = dk |k , m t + x c xc )m + i m a(k ).

so that

|idt = im m + (t )m m + m x c m |xc m m
m |k m ) + x c m |x m ) m
= im m + dt ( kxc ) + k x c + i(k c m
Now for our purposes the perturbations we have in mind inside H are proportional to ( x xc ) (or position and
momentum independent, if we choose to include these as perturbations) . Since by definition | x| = xc then
x xc | = 0, the terms in H propotional to (x xc ) do not contribute to L. If we do choose to include terms
|
dependent only on (the parameter) xc into the perturbation V (xc ) then |H | = m
m |V (xc )|m m .
For the remainder of this section we consider |m to be r independent. The Lagrangian thus reads, after discarding
the total time derivative dt ( kxc ),
m |k m ) ek,m m .
L = im m + k x c + i(k m (35)
Restoring multiple spatial dimensions and dropping the c subscript from x, the Lagrangian reads
L = i m + ka x a + i(ka m |k m ) e m .
m a k,m m (36)
Taking functional derivatives with respect to ka , we have

( )
ka L = x a + im kb ka m |kb m m (ka ek,m )m m (37)

t ka = t (im m |ka m m )

= im (kb (kb m |ka m ))m (38)
so that we end up with the Euler-Lagrange equation

x a = (ka ek,m )m m + ikb m (kb m |ka m ka m |kb m )m

= (ka ek,m )m m + im (kb (kb m |ka m ka m |kb m )m (39)
Taking the functional derivative with respect to m gives
( )
= (k | + x
m a m |xa m ) + im |V (
r)|m + im ek,m )
m a m ka m

so that on taking CC we have


( )
m = ka m |ka m + x a m |xa m + im |V (
r)|m m iek,m (r)m

Together, these give (concentrating on the k a piece),


t i ij j = k a (l l |ka i )ij j i ij j |ka l l + ...
= k a (l l |ka i )ij j j ji i |ka l l + ... (Swap i j in second piece) (40)

You might also like