You are on page 1of 5

Spin imbalance in hybrid superconducting structures with spin-active interfaces

Oleksii Shevtsov

and Tomas Lofwander

Department of Microtechnology and Nanoscience MC2,


Chalmers University of Technology, SE-412 96 G oteborg, Sweden
(Dated: April 25, 2014)
We consider a heterostructure consisting of a normal metal and a superconductor separated by
a spin-active interface. At nite bias voltages, spin-ltering and spin-mixing eects at the interface
allow for an induced magnetization (spin imbalance) on the superconducting side of the junction,
which relaxes to zero in the bulk. Such interfaces are also known to host a pair of in-gap Andreev
bound states which were recently observed experimentally. We show that the presence of these states
inuences the induced spin imbalance. Motivated by recent experiments on spin-charge separation in
superconducting aluminum wires, we propose an alternative way to observe spin imbalance without
applying an external magnetic eld. We also suggest that the peculiar dependence of the spin
imbalance on the applied bias voltage permits an indirect bound state spectroscopy.
PACS numbers: 74.78.Na,73.63.-b,74.45.+c
Non-equilibrium phenomena in superconductors have
attracted much attention since the pioneering works on
charge imbalance by Clarke and co-workers [14]. They
found that an excess charge brought into the super-
conductor by tunneling electrons reduces the density of
Cooper pairs close to the interface which leads to a non-
vanishing resistance of this part of the superconductor.
Each electron tunneling into the superconductor also
brings along its spin moment. Therefore, if the number of
injected electrons is dierent for opposite spin projections
(e.g. by using a ferromagnet instead of a normal metal,
or by applying a magnetic eld) it is possible to induce
a non-equilibrium magnetization, or spin imbalance, to-
gether with the charge imbalance at the superconducting
side of the interface. In a normal metal, charge and spin
of an electron are bound together. On the other hand,
the nature of Bogoliubov quasiparticles in a superconduc-
tor is more complicated. Indeed, recent experiments [5
7] have demonstrated spin-charge separation [8], where
charge imbalance and spin imbalance relax away from
the interface with dierent characteristic length scales.
It is important to note that these experiments needed
the orbital pair-breaking eect of an external magnetic
eld to observe spin-charge separation.
Here, we propose an alternative way to observe spin
imbalance, which does not require a magnetic eld. Our
idea relies on the possibility of fabricating spin-active in-
terfaces [917]. A superconductor coated with a spin-
active layer hosts a pair of interface bound Andreev
states, whose properties are controlled by parameters of
the interface [11, 12]. They have been observed in re-
cent tunneling experiments on nanoscale superconductor-
ferromagnet junctions [18]. We will show in this Letter
that these states play an important role for the spin im-
balance eect near the interface.
Consider a junction between a normal metal (N) and
a superconductor (S) with a spin-active interface, as in
Fig. 1(a). Assuming a smooth specularly reecting in-
terface invariant in the transversal direction, we may
consider spatial dependence along the longitudinal z-axis
only. A simple model of such an interface can be quanti-
ed by the following scattering matrix connecting incom-
ing and reected electrons in the normal state [12],
o =

S
d
iS
nd
iS
nd
S
d

, S
d
= [R + ( )] e
i( )/2
,
S
nd
= [D + ( )] e
i( )/2
, (1)
N S e*
h
*
e*
h
*
(a) (b)
(c) (d)
E/k
B
T
C
E/k
B
T
C
E/k
B
T
C
z
/

0
N
/
N
F
z
/

0
z
/

0
N
/
N
F
N
/
N
F
^

z
FIG. 1. (color online). (a) A normal metal (N) - super-
conductor (S) junction with a spin-active interface is char-
acterized by an intrinsic magnetic moment directed along .
The closed trajectory indicates formation of Andreev surface
bound states. (b)-(d) Local density of states in S as func-
tion of distance from the interface for spin mixing angles
= 0, 0.49, and 0.83. The interface transparencies are
D

= D

= 0.06. Here, 0 is the superconducting coherence


length, Tc the critical temperature, and NF the density of
states at the Fermi energy in the normal state.
a
r
X
i
v
:
1
4
0
4
.
6
1
7
2
v
1


[
c
o
n
d
-
m
a
t
.
m
e
s
-
h
a
l
l
]


2
4

A
p
r

2
0
1
4
2
with reection coecients R = (

)/2, =
(

)/2, and transmission coecients D =


(

)/2, = (

)/2. They fulll


R
,
+D
,
= 1. For an impenetrable wall (R

= R

=1),
reections are accompanied by spin-dependent phase
shifts through the spin-mixing angle [1012]. This
leads to formation of surface bound states (sometimes
called Andreev states), trapped between the impenetra-
ble wall and the bulk of the superconductor by the super-
conducting gap in the spectrum. A Bohr-Sommerfeld
quantization rule can be set up by considering the closed
loop in Fig. 1(a). A spin-mixing phase /2 is picked
up during reection at the interface, where the signs
correspond to spin-up and spin-down states. An en-
ergy () dependent phase shift () is picked up
during Andreev reection, where the signs corresponds
to electron-hole and hole-electron conversion processes.
Here, () = arccos(/), and and are the magni-
tude and phase of the superconducting order parameter,
respectively [19]. The quantization condition becomes
2() = 2n, where n is an integer. The resulting
surface states appear at energies
ABS
= cos(/2)
[11, 12]. The wave functions of the surface states decay
into the bulk of the superconductor at a characteristic
length
ABS
() v
F
/ cos()

2
ABS
, where is
the angle between the quasiparticle trajectory and the
z-axis, and v
F
is the quasiparticle velocity at the Fermi
surface in the normal state. This length scale can be very
long if the bound state is close to the gap edge
ABS
.
However, after averaging over all angles, as in the local
density of states, Fig. 1(c)-(d), the bound state peak still
decays [20] at a short distance of the order of the super-
conducting coherence length
0
= v
F
/2k
B
T
C
, where
T
C
is the critical temperature. For a tunnel barrier
(D
,
> 0), the surface states broaden into resonances
of width D
,
. As conrmed experimentally [18],
the positive-energy and negative-energy resonance peaks
correspond to quasiparticle states with opposite spin pro-
jections, see Fig. 2(c)-(d).
We utilize the quasiclassical Greens function formal-
ism with appropriate boundary conditions [21, 22] involv-
ing the scattering matrix Eq. (1). We compute the re-
tarded, g
R
( p
F
, , z), advanced, g
A
( p
F
, , z), and Keldysh,
g
K
( p
F
, , z), Greens functions (hat refers to the 2 2
structure in Nambu space). Spin imbalance is calcu-
lated by integrating the Keldysh Greens function over
the Fermi surface and energy according to [23],
M(z) =
i
B
N
F
8

d
p
F
4

d Tr

g
K
( p
F
, , z)

, (2)
where = diag[,

] is a diagonal matrix in Nambu


space, is a vector of the three spin Pauli matrices,
B
is
the Bohr magneton, and N
F
is the density of states at the
Fermi level in the normal state. In order to separate out
the non-equilibrium part of spin imbalance, we split the
Keldysh Greens function into spectral and anomalous
[ ]
M



(
z
=
0
)
e
q
M



(
z
)
/
e
q
)
l
n
(
(a) (b)
k
B
T
C

B
N
F
M



(
z
=
0
)
/
e
q
z/
0
E/k
B
T
C
z
/

0
E/k
B
T
C
z
/

0
(c) (d)
N


/
N
F
N


/
N
F
FIG. 2. (color online). (a) Interface value Meq(0) as function
of spin-mixing angle. (b) Semi-log plot of Meq(z)/Meq(0) as
function of distance from the interface. (c)-(d) Spin-down
and spin-up local density of states N
,
(, z) as function of
distance for = 0.66. D

= D

= 0.06, and T = 0.01TC.


parts [21, 22],
g
K
=

g
R
g
A

tanh

2k
B
T
+ g
a
, (3)
where we omit arguments for compactness. Then the
spin imbalance Eq. (2) is expressed as M(z) = M
eq
(z) +
M
ne
(z). The rst term corresponds to the spectral part
of g
K
. It exists in equilibrium and is sometimes called
inverse proximity eect [2428]. The second term, re-
lated to the anomalous propagator g
a
, is a true non-
equilibrium contribution and it depends explicitly on ap-
plied bias voltage V .
It is known that there are several mechanisms respon-
sible for spin relaxation [29], among which are scatter-
ing against magnetic impurities or presence of spin-orbit
coupling in combination with momentum scattering by
e.g. scalar impurities. In this work we focus on the sim-
plest mechanism, namely scattering by magnetic impu-
rities characterized by a spin-ip length, l
sf
. It is well
known that the presence of a small fraction of magnetic
impurities can signicantly reduce the order parameter
and even drive the superconductor into a gapless regime
[30]. We consider the case l
sf

0
, for which the pair
breaking eect is small. Throughout the text we use
l
sf
314
0
and compute the bulk impurity self-energy
and the bulk order parameter self-consistently. In this
case we get 1.776k
B
T
C
. The results presented be-
low are for the experimentally relevant tunneling limit,
D
,
1. In this case, Andreev bound states have a well-
dened energy. In the tunneling limit, and for small spin-
mixing angles , the order parameter is only marginally
suppressed near the interface and self-consistency of self-
energies may be neglected when computing spin imbal-
3
(a)
k
B
T
C
eV/ k
B
T
C
eV/
(b)
z/
0
M



(
z
=
0
)
n
e
M



(
z
)
/
n
e
)
l
n
(
(c)
k
B
T
C

B
N
F
M



(
z
=
0
)
/
n
e
(
1
0


)
-
2
k
B
T
C

B
N
F
M



(
z
=
0
)
/
n
e
(
1
0


)
-
2
FIG. 3. (color online). Non-equilibrium part of spin imbalance. (a) Interface value as a function of bias voltage for = 0,
D

= 0.06, and D

= 0.02. (b) Interface value for = 0.32 (black circles), = 0.49 (blue rectangles), = 0.66 (red
diamonds), = 0.83 (green triangles), and D

= D

= 0.06. (c) Semi-log plot of Mne(z)/Mne(0) for eV = 3.2kBTC (solid


lines) and eV = 1.7kBTC (dashed lines). The grey rectangle in (a)-(b) depicts the subgap region. Temperature T = 0.01TC.
ance. Below we focus on such a non-self consistent cal-
culation (also for arbitrary ) of interface properties and
comment in the end on the eects of self-consistency. Fi-
nally, since there is a single spin quantization axis in the
problem given by the interface moment , the spin im-
balance is parallel to it, M(z) = M(z) .
We start by discussing the spectral part of spin imbal-
ance, see Fig. 2. By denition,
M
eq
(z) =

B
2

d [N

(, z) N

(, z)] tanh

2k
B
T
, (4)
where N
,
(, z) are spin-resolved local densities of states.
The magnitude of M
eq
(z) is determined by the weight of
the Andreev bound states in the total density of states,
see Fig. 2(c)-(d), which increases as a function of [12].
In Fig. 2(a) we plot the interface value as function of spin-
mixing angle. The decrease of M
eq
(0) for large values
of is due to overlap (in energy space) of the bound
state peaks. One can show that M
eq
(z) sin [28], and
vanishes at = 0 and . In Fig. 2(b) we show how the
equilibrium part of spin imbalance decays away from the
interface. It turns out that contributions to M
eq
(z) from
bound states and from continuum states quickly cancel
each other as we move into the bulk. Therefore, the
inverse proximity eect decays on the short coherence
length scale
0
independently of [2426, 28].
Let us now consider the non-equilibrium part of spin
imbalance, M
ne
(z) (see Fig. 3) [31]. There are two main
contributions: one coming from spin-ltering ( = 0,
D

,= D

) and the other from spin-mixing ( ,= 0,


D

= D

). In Fig. 3(a)-(b) we plot the interface val-


ues, M
ne
(0), as function of bias voltage for these two
components. The spin-ltering and spin-mixing compo-
nents have dierent symmetries under V V . The
spin-ltering component is an odd function since positive
and negative biases correspond to adding or withdrawing
majority spins. The spin-mixing component is an even
function because positive and negative biases correspond
either to populating an Andreev bound state at positive
energy with one spin projection or depopulating the cor-
responding negative energy state with the opposite spin
projection. Furthermore, their voltage dependences are
markedly dierent. The spin-ltering component is due
to injection into continuum states and depends on the
size of the bias window and grows linearly at large bias,
but is quenched in the sub gap region. On the other hand,
the spin-mixing component, Fig. 3(b), consists of a sharp
increase of spin imbalance for voltages corresponding to
the energy of the bound state, but saturates quickly when
the whole resonance lies in the bias window, since the dif-
ference between spin-resolved densities of states is small
in the continuum [eV [ > . For general parameters, the
two components are superimposed (not shown in the g-
ure), but we note that the spin-mixing component dom-
inates at the interface because the bound state is com-
pletely spin polarized and its occupation leads to a large
spin imbalance.
In Fig. 3(c) we show on a semi-log plot the spatial
dependences of the spin-ltering and spin-mixing contri-
butions. For the case of pure spin-ltering and eV > ,
we inject spin polarized quasiparticles into the contin-
uum states. This contribution relaxes through scatter-
ing against magnetic impurities and decays on the spin-
ip length scale, l
sf
[slowly decaying, magenta line, in
Fig. 3(c)]. For the case of pure spin-mixing, if eV <
we have three possibilities. First, when eV < ,
ABS
,
only evanescent states are populated and this contribu-
tion is small and decays on a coherence length scale (not
shown). Secondly, when
ABS
< eV < , we populate
the Andreev bound state and this contribution decays on
the corresponding length
ABS
(dashed lines) which can
be long when the bound state is close to the gap edge
(small spin mixing angle). Finally, when eV > ,
ABS
,
we populate both the Andreev bound state and some
states in the continuum. Then, close to the interface the
spatial dependence is determined by the Andreev bound
state, while for distances far enough that the bound state
has decayed, the dominant contribution comes from the
4
k
B
T
C
eV/
(b)
k
B
T
C
eV/
(c)
k
B
T
C
eV/
(a)
M



(
z
)
/
n
e

B
N
F

e
V

)
(
1
0
-
2
M



(
z
)
/
n
e

B
N
F

e
V

)
(
1
0
-
2
M



(
z
)
/
n
e

B
N
F

e
V

)
(
1
0
-
2
FIG. 4. Derivative of the anomalous part of spin imbalance with respect to bias voltage for: (a) = 0, D

= 0.06, and
D

= 0.02; (b) = 0.32 and D

= D

= 0.06 ; (c) = 0.66 and D

= D

= 0.06. Solid, dashed and dotted lines correspond


to z = 0, 0, and 20, respectively. The grey rectangle depicts the subgap region. Temperature T = 0.05TC.
continuum with the decay length l
sf
(solid lines).
Let us now discuss the implications of the behavior
of the non-equilibrium part of spin imbalance discussed
above for experiments. In Ref. [7] authors used the re-
sults of tunnel model calculations [32] in order to analyze
their data. They measured a non-local dierential con-
ductance in a NISIF structure (I stands for insulator
and F for ferromagnet) g
nl
dI
det
/dV
inj
in an exter-
nal magnetic eld. Here, I
det
is the current at the detec-
tor electrode in response to an injection voltage V
inj
. For
the detector current they used (see Eq. (6) in Ref. [7]),
I
det
=
G
det
eN
F

(Q

+ Q

) + P
det
(S

, (5)
where G
det
is the normal state detector conductance,
P
det
is the spin polarization of the detector, Q

,
are
the spin-up/down contributions to charge imbalance [6],
and S
,
are the spin-up/down densities induced by a
spin-polarized current [6]. In order to separate the spin
and charge imbalance parts they used symmetries of
these components with respect to the injection voltage.
The charge imbalance Q

= Q

+ Q

is anti-symmetric
with respect to V
inj
, since for negative values electrons
are injected into the system, while holes are injected
for positive values. At the same time, the spin imbal-
ance, created in their case by an external magnetic eld,
is symmetric with respect to V
inj
(see also Ref. [5]).
Therefore they obtained I
sym
det
(V
inj
) (S

) and
g
asym
nl
(V
inj
) d(S

)/dV
inj
[33]. Note that the dier-
ential conductance has opposite symmetry to the current.
In the experiment the presence of an external magnetic
eld was crucial as the spin imbalance eect was created
by spin polarization of the superconducting density of
states (Zeeman eect) [34, 35]. In our case the spin polar-
ization of the density of states comes from the interface-
induced Andreev bound states, see Fig. 2(c)-(d). We also
note that the spin-mixing contribution in our case is sym-
metric with respect to bias, see Fig. 3(b). In another ex-
periment, Ref. [5], they measured a non-local dierential
resistance in a FISIF structure in an external magnetic
eld. Again, the presence of magnetic eld was crucial
to observe the spin imbalance because otherwise the only
source of spin imbalance is the spin-ltering eect, which
has the same symmetry as the charge imbalance and is
much smaller than the latter [5], see Fig. 3(a). Finally,
the (orbital) pair-breaking eect of the external magnetic
eld [36] made the charge imbalance signal decay faster
[37] than the spin imbalance in both experiments, a sit-
uation that was called spin-charge separation [5, 8].
In Fig. 4(a)-(c) we plot the derivative of the non-
equilibrium part of spin imbalance with respect to the
bias voltage (the equilibrium part does not depend on
bias and is not relevant for the above mentioned exper-
iments). The spin-ltering part, Fig. 4(a), has the same
symmetry with respect to V as the non-local conductance
responsible for the charge imbalance (see Ref. [5, 7]),
which is why it cannot be observed in experiments [5].
On the other hand, the derivative of the spin-mixing
contribution, Fig. 4(b)-(c), looks very similar to the non-
local conductance due to spin imbalance in Ref. [7]. It
is important to note that peaks observed experimentally
occurred at voltages near the gap edges. In our case
the peak positions are determined by the Andreev bound
state energies
ABS
. Therefore this particular feature of
spin imbalance can be used for bound state spectroscopy.
Our analysis suggests that it is possible to observe the
spin imbalance signal by doing analogous non-local mea-
surements without applying an external magnetic eld.
We leave for future studies a quantication of the level
of spin-charge separation in our setup, since it is nec-
essary to compute the order parameter self-consistently
to properly describe charge imbalance. The charge im-
balance relaxation length is related to the conversion of
quasiparticle current ow near the interface to superow
in the bulk, which is neglected here. For spin imbalance,
self-consistency is not as crucial.
We would like to comment on the case of highly trans-
parent junctions. It is known that the width of Andreev
bound states is proportional to the barrier transparency
while its weight in the total density of states is propor-
tional to the reection coecient [12]. Thus for the case
5
of high transparency junctions, the resulting spin imbal-
ance signal will be reduced and it will be dicult to assign
a single decay length to the bound states as they will de-
velop into broad resonances. We therefore conclude that
it is desirable to work with spin-active tunnel junctions.
Finally, in case of a disordered sample, the nite mean
free path l reduces the superconducting coherence length,

0
=

l
0
/3 [38]. In our (non-self-consistent)
model this means that the results we presented above
still hold but the length scale is reduced to . In fully
self-consistent calculations the disorder might broaden
the bound states [11], however we believe our results to
be still valid. Consequently in order to test our predic-
tions it is recommended to work with clean samples to
obtain better spatial resolution of spin imbalance.
In summary, we propose a way to create a spin imbal-
ance signal by fabricating a hybrid structure with a spin-
active interface. It can in principle be observed with the
currently used non-local conductance measurement tech-
niques, as it possesses the same symmetry as the Zeeman-
induced spin imbalance signal already observed experi-
mentally, and is of opposite symmetry to the charge im-
balance signal. The advantage of our setup is that it
does not require an external magnetic eld and that the
characteristics of the spin imbalance are controlled by
parameters of the interface, which can in principle be en-
gineered [39]. This feature can be used for bound state
spectroscopy.

shevtsov@chalmers.se

tomas.lofwander@chalmers.se
[1] J. Clarke, Phys. Rev. Lett. 28, 1363 (1972).
[2] M. Tinkham and J. Clarke, Phys. Rev. Lett. 28, 1366
(1972).
[3] M. Tinkham, Phys. Rev. B 6, 1747 (1972).
[4] C. C. Chi and J. Clarke, Phys. Rev. B 19, 4495 (1979).
[5] C. H. L. Quay, D. Chevallier, C. Bena, and M. Aprili,
Nat. Phys. 9, 84 (2013).
[6] F. H ubler, M. J. Wolf, D. Beckmann, and
H. v. L ohneysen, Phys. Rev. Lett. 109, 207001 (2012).
[7] M. J. Wolf, F. H ubler, S. Kolenda, H. v. L ohneysen, and
D. Beckmann, Phys. Rev. B 87, 024517 (2013).
[8] S. A. Kivelson and D. S. Rokhsar, Phys. Rev. B 41, 11693
(1990).
[9] A. Millis, D. Rainer, and J. A. Sauls, Phys. Rev. B 38,
4504 (1988).
[10] T. Tokuyasu, J. A. Sauls, and D. Rainer, Phys. Rev. B
38, 8823 (1988).
[11] M. Fogelstr om, Phys. Rev. B 62, 11812 (2000).
[12] E. Zhao, T. L ofwander, and J. A. Sauls, Phys. Rev. B
70, 134510 (2004).
[13] A. Cottet and W. Belzig, Phys. Rev. B 72, 180503 (2005).
[14] A. Cottet and J. Linder, Phys. Rev. B 79, 054518 (2009).
[15] J. Linder, T. Yokoyama, A. Sudb, and M. Eschrig,
Phys. Rev. Lett. 102, 107008 (2009).
[16] J. Linder, M. Cuoco, and A. Sudb, Phys. Rev. B 81,
174526 (2010).
[17] Y. Asano, M. Ozaki, T. Habe, A. A. Golubov, and
Y. Tanaka, Phys. Rev. B 86, 024510 (2012).
[18] F. H ubler, M. J. Wolf, T. Scherer, D. Wang, D. Beck-
mann, and H. v. Lohneysen, Phys. Rev. Lett. 109,
087004 (2012).
[19] T. L ofwander, V. S. Shumeiko, and G. Wendin, Super-
cond. Sci. Technol. 14, R53 (2001).
[20] G. Metalidis, M. Eschrig, R. Grein, and G. Sch on, Phys.
Rev. B 82, 180503 (2010).
[21] M. Eschrig, Phys. Rev. B 61, 9061 (2000).
[22] M. Eschrig, Phys. Rev. B 80, 134511 (2009).
[23] J. W. J.W Serene and D. Rainer, Phys. Rep. 101, 221
(1983).
[24] F. S. Bergeret, A. F. Volkov, and E. K. B., Europhys.
Lett. 66, 111 (2004).
[25] F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys.
Rev. B 69, 174504 (2004).
[26] F. S. Bergeret, A. Levy Yeyati, and A. Martn-Rodero,
Phys. Rev. B 72, 064524 (2005).
[27] J. Xia, V. Shelukhin, M. Karpovski, A. Kapitulnik, and
A. Palevski, Phys. Rev. Lett. 102, 087004 (2009).
[28] R. Grein, T. L ofwander, and M. Eschrig, Phys. Rev. B
88, 054502 (2013).
[29] I.

Zutic, J. Fabian, and S. Das Sarma, Rev. Mod. Phys.
76, 323 (2004).
[30] J. C. Phillips, Phys. Rev. Lett. 10, 96 (1963).
[31] For the calculations of the non-equilibrium part Mne(z)
we use a tunneling cone D
,
() = D
,
(0)e
sin
2

[40],
which is an experimentally relevant model. We used =
80, which corresponds to a 14

wide tunneling cone. The


transmission probabilities D
,
in the text are the normal
incidence values D
,
(0). Note that it is not necessary
to use the tunneling cone model for Meq(z) since it is
proportional to reection coecients, which are close to
unity for all .
[32] H. L. Zhao and S. Hersheld, Phys. Rev. B 52, 3632
(1995).
[33] We note that the induced magnetization is M =
(|e|g/2m)(S

), where g is the electron g-factor, e is


the electron charge and m is the electron eective mass.
[34] P. M. Tedrow and R. Meservey, Phys. Rev. Lett. 26, 192
(1971).
[35] R. Meservey and P. Tedrow, Phys. Rep. 238, 173 (1994).
[36] K. Maki, Prog. Theor. Phys. 32, 29 (1964).
[37] F. H ubler, J. C. Lemyre, D. Beckmann, and
H. v. L ohneysen, Phys. Rev. B 81, 184524 (2010).
[38] N. Kopnin, Theory of Nonequilibrium Superconductivity
(Oxford University Press, New York, 2001).
[39] T. S. Khaire, M. A. Khasawneh, W. P. Pratt, and N. O.
Birge, Phys. Rev. Lett. 104, 137002 (2010).
[40] E. L. Wolf, Principles of Electron Tunneling Spec-
troscopy, 2nd ed. (Oxford University Press, New York,
2012).

You might also like