You are on page 1of 10

Orbital Kondo effect in double quantum dots

Piotr Trocha1, ∗
1
Department of Physics, Adam Mickiewicz University, 61-614 Poznań, Poland
(Dated: July 9, 2010)
Orbital Kondo effect in a system of two single-level quantum dots attached to external electron
reservoirs is considered theoretically. The dots are coupled via direct hoping term and Coulomb
interaction. The Kondo temperature is evaluated from the scaling approach and slave boson tech-
nique. The later method is also used to calculate linear conductance of the system. Nonlinear
conductance, in turn, is calculated in terms of the nonequilibrium Green function formalism.

PACS numbers: 72.15.Qm, 73.23.-b, 73.63.Kv

I. INTRODUCTION Ref. [4].


The orbital Kondo effect in double quantum dot
(DQD) systems was analyzed e.g. in Ref. [18]. How-
Kondo effect in electronic transport through quantum ever, our results are different, and the key difference con-
dots (QDs) strongly coupled to external leads is a many sists in a different symmetry of the couplings to exter-
body phenomenon which has been extensively studied in nal leads. Moreover, we use various techniques including
the last two decades1–11 . Spin fluctuations in the dot, scaling, slave boson, and nonequilibrium Green function
generated by coupling of the dot to external leads, give formalisms. Apart from this, we apply a different method
rise to a narrow peak in the dot’s density of states (DOS) to evaluate the lesser Green function. The paper is or-
at the Fermi level. This Rado-Suhl resonance results in ganized as follows. In section 2 we describe the model of
enhanced transmission through the dot, and leads to the a double quantum dot system. Renormalization of the
unitary limit of the linear conductance G at zero tem- dots’ levels and the Kondo temperature are discussed in
perature, G = 2e2 /h. The enhanced transmission is sup- section 3 in terms of the scaling approach. The slave bo-
pressed when a bias voltage is applied, and this leads to son technique is briefly described in section 4 and is used
the so-called zero-bias (Kondo) anomaly in differential there to estimate the Kondo temperature and calculate
conductance. The above described phenomenon arises the linear conductance. Basic formula for nonequilibrium
from the two-fold spin degeneracy, and is often referred Green functions and the corresponding numerical results
to as the spin Kondo effect. However, the Kondo phe- on the conductance and LDOS are presented and dis-
nomenon may also appear when the spin degree of free- cussed in section 5. Summary and final conclusions are
dom is replaced by any two-valued quantum number, e.g. given in section 6.
the one associated with an orbital degree of freedom (the
orbital Kondo effect)12 . A minimal realization of the or-
bital (spinless) Kondo phenomenon requires two orbital II. MODEL
discrete levels coupled to external leads13,14 . This can
be realized for instance in two single-level quantum dots
We consider two coupled single-level quantum dots
coupled to external electrodes15–24 . Coherent superpo-
connected to nonmagnetic electron reservoirs as shown
sition of virtual tunneling events, in which one electron
schematically in Fig.1. Each dot is attached to sepa-
tunnels from the dot QD1 (QD2) to one of the leads and
rate source and drain leads, so the tunneling paths via
then simultaneously another electron tunnels to the dot
the two orbitals can be analyzed separately as in recent
QD2 (QD1), leads to the Kondo resonance at low tem-
experiments15,16 . We consider the case when each dot
peratures.
is coupled symmetrically to the leads, while the corre-
In this paper we consider theoretically the Kondo phe- sponding coupling strengths for both dots may be dif-
nomenon in electronic transport through two QDs cou- ferent. Moreover, our considerations are limited to the
pled, in general, via both Coulomb interaction and hop- case of spinless electrons, which can be realized exper-
ping term. To evaluate the level renormalization and imentally for instance by applying a sufficiently strong
Kondo temperature of the system we use the scaling ap- external magnetic field lifting the spin degeneracy.
proach. The Kondo temperature is also evaluated from The system under consideration can be described by
the slave boson technique. Additionally, the latter tech- the extended Anderson Hamiltonian of the general form
nique is used to calculate the linear conductance. Then,
the nonequllibrium Green function formalism is used to Ĥ = Ĥleads + ĤDQD + Ĥtunnel . (1)
calculate the local density of states (LDOS) for both dots
and transport characteristics (differential conductance) The first term, Ĥleads , describes here the four leads in the
in the nonlinear response regimes. To calculate the rele- non-interacting quasi-particle approximation, Ĥleads =
vant Green’s functions from the corresponding equations ĤL1 + ĤL2 + ĤR1 + ĤR2 , with Ĥβi being the Hamiltonian
of motion we apply the decoupling scheme introduced in of the left (β = L) and right (β = R) lead attached to the
2

αΓ αΓ system. From the scaling equations we also estimate the


L1 QD1 R1 relevant Kondo temperature. The derived results will
be used subsequently for interpretation of the numerical
results on electronic transport and LDOS.
t
A. Renormalization of the QDs’ levels

L2 QD2 R2
Γ Γ Now, we apply the scaling technique to derive renor-
malized dots’ energy levels and begin with the limit of
FIG. 1: (color online) Schematic picture of the double quan- t = 0. In the scaling approach, the high-energy excited
tum dot system coupled to external leads. states (in the energy region of width δD at the band
edges) are removed, but their impact on the system is
taken into account via renormalized parameters of the
∑ † Hamiltonian. Here, we consider only second order pro-
ith dot (i = 1, 2), Ĥβi = k ϵkβi ckβi ckβi (for β = L, R cesses, where the leads’ electrons are scattered to the
and i = 1, 2). Here, c†kβi
(ckβi ) is the creation (annihi- band edges and back. To perform scaling we assume
lation) operator of an electron with the wave vector k ϵi + U ≫ D ≫ |ϵi |. After integrating out the band edge
in the lead βi, whereas ϵkβi denotes the corresponding states we arrive at the following renormalized parame-
single-particle energy. ters;
The second term of the Hamiltonian (1) describes the
double quantum dot system, Γī δD
ϵ̃i = ϵi − E0 + , (4)
∑ † 2π D
ĤDQD = ϵi di di + t(d†1 d2 + h.c.) + U n1 n2 , (2)
where E0 is the energy of empty DQD system (initially
i
E0 = 0) and the index ī = 1 for i = 2 and ī = 2 for i = 1.
where ni = d†i di is the particle number operator (i = Here, Γi is defined as Γi = ΓL R
i +Γi . This procedure leads
1, 2), ϵi is the discrete energy level of the i-th dot, t de- to the following scaling equation:
notes the inter-dot hopping parameter (assumed real), dϵ̃i Γ
and U is the inter-dot Coulomb integral. = − ī , (5)
The last term, HT , of Hamiltonian (1) describes elec- d ln D 2π
tron tunneling between the leads and dots, and takes the and to the level separation,
form ( )
∑∑ β † (Γ2 − Γ1 ) D
ĤT = (Vik ckβi di + h.c.), (3) ∆ϵ̃ = ϵ̃1 − ϵ̃2 = ϵ1 − ϵ2 + ln , (6)
2π e
D
k βi

β where D e is the band width at the end of scaling proce-


where Vik are the relevant tunneling matrix elements.
dure. When ϵ1 = ϵ2 = ϵ0 , the above equation shows that
Coupling of the dots to external leads can be param-
∑ β β∗ the initial degeneracy is generally lifted.
eterized in terms of Γβi (ϵ) = 2π k Vik Vik δ(ϵ − ϵkβi ). Scaling in the presence of direct tunneling between the
β
We assume that Γi is constant within the electron band, dots, t ̸= 0, is more complex in a general case. However,
Γβi (ϵ) = Γβi = const for ϵ ∈ ⟨−D, D⟩, and Γβi (ϵ) = 0 oth- we restrict our considerations to some limiting cases, i.e.
erwise. Here, 2D denotes the electron band width. We when the hoping term is weak, |t|/Γ ≪ 1, and when
assume the dots are symmetrically coupled to the leads, |t|/Γ ≫ 1. If the bare dots’ levels are degenerate, the di-
ΓL R L R
1 = Γ1 = αΓ, and Γ2 = Γ2 = Γ. The parameter α rect hopping term generally lifts the degeneracy. The two
takes into account difference in the coupling of the two eigenstates of the coupled quantum dots isolated from the
dots to external leads. Note, that for these parameters, leads correspond to the antibonding and bonding √ states,
each dot separately is coupled symmetrically to the two with the eigenenergies ϵ± = (ϵ1 + ϵ2 )/2 ± ∆ϵ2 + t2 ,
leads. where ∆ϵ = (ϵ1 − ϵ2 )/2. When the hoping term is small,
one can first perform scaling of the bare dots’ levels and
then incorporate nonzero t by substituting ϵ1(2) by ϵ̃1(2)
III. LEVEL RENORMALIZATION AND KONDO in the above expression for ϵ± . The situation changes
TEMPERATURE when tunneling coupling between the dots is larger than
the dot-lead coupling. To find the relevant energy lev-
Coupling of the dots to external leads gives rise to els involved in the Kondo effect, one has to diagonalize
renormalization of the energy levels of both dots. In the dot’s Hamiltonian first (transformation to the bond-
this section we use the scaling approach to derive some ing and antibonding states), and then perform scaling for
general formula for the renormalized levels in unbiased the energy levels ϵ+ and ϵ− . The corresponding scaling
3

equation has the form 1.0


slave boson
dϵ̃± Γ∓
=− , (7) poor man scaling
d ln D 2π

TK(α)/TK(α=1)
which is similar to Eq.(5). However, the effective cou-
pling of the new states to the leads acquires now the form 0.5
Γ± = Γ1 + Γ2 . Thus, the level separation in the limit of
strong hoping term is independent of the couplings Γ±
as they are both the same, Γ+ = Γ− .

B. Kondo temperature 0.0


0.00 0.25 0.50 0.75 1.00
Now we evaluate the Kondo temperature for t = 0 and α
ϵ1 = ϵ2 = ϵ0 using the ’poor man’ scaling approach26 .
To pursue this method one has to derive first the Kondo FIG. 2: Normalized Kondo temperature as a function of the
Hamiltonian by performing the Schrieffer-Wolf transfor- parameter α, obtained from the scaling method for U = 50Γ
mation. Then, the band width is reduced by eliminating and from the slave boson technique for U = ∞. The other
states with energy D − δD ≤ |ϵkβi | ≤ D and introducing parameters are ϵ0 = −3.5Γ and t = 0.
a new effective Kondo Hamiltonian, which has the same
form as the initial one, but with renormalized parameters
J˜+ , J˜− , J˜z1 and J˜z2 . All information on the high energy
where instead of p̃ we have spin polarization p of the
excitations is incorporated into these renormalized pa-
leads.
rameters.
To apply the renormalization group procedure we first Variation of the Kondo temperature with the parame-
reformulate the definitions of the coupling parameters Γi ter α is shown in Fig.2 (dashed line). TK reaches max-
in the following way: Γ1 = 2αΓ ≡ Γ̃(1 − p̃) and Γ2 = imum value for α = 1 (p̃ = 0) and vanishes for α → 0
(p̃ = 1). This behavior is similar to that for spin Kondo
2Γ ≡ Γ̃(1 + p̃), so that Γ̃ = (α + 1)Γ and p̃ = (1 − α)/(1 +
phenomenon in a QD coupled to ferromagnetic leads. We
α). After performing scaling procedure one arrives at the
also note that our results are in agreement with those
following scaling equations;
obtained in Ref.[28], where the authors mapped spinless
d(ρ J˜± ) DQD system onto a spinful generalized Anderson model.
= −ρ J˜± (ρJ˜z1 + ρJ˜z2 ), (8)
d ln D

d(ρJ˜zi ) IV. SLAVE BOSON APPROACH


= −2(ρJ˜± )2 , (9)
d ln D
∑ To estimate the Kondo temperature and calculate con-
where ρ = ρ1 = ρ2 , with ρi = β ρβi , and ρβi being the
ductance in the linear response regime, we apply now the
density of states in the lead βi (β = L, R and i = 1, 2). To
slave boson technique for U → ∞29 . This method relies
solve these equations we first find the scaling trajectories
on introducing auxiliary operators for the dots, and re-
(ρ J˜± )2 − (ρ J˜z1 )(ρ J˜z2 ) = 0 and ρ J˜z1 − ρ J˜z2 = const ≡
placing the electron creation and annihilation operators
ρJz10
− ρJz2
0
= p̃ρ(Jz1 0
+ Jz20
). This allows us to write only
one scaling equation instead of the two coupled equations by fi† b and b† fi , respectively. Here, b† creates an empty
(8) and (9), state, whereas fi† creates a singly occupied state with
an electron in the i-th dot. To eliminate non-physical
d(ρJ˜zi ) states, the following constraint has to be imposed on the
= −2(ρ J˜zi )[ρ J˜zi ∓ ρp̃(Jz1
0 0
+ Jz2 )], (10) new quasi-particles,
d ln D
∑ †
for i = 1, 2. One actually continues the scaling process Q= fi fi + b† b = 1. (12)
until D ≈ kB TK . Solving Eq. (10) one finds the Kondo i
temperature as the relevant scaling invariant,
{ } The above constraint prevents double occupancy of the
e exp − 1 arctanh(p̃) system (the DQD system is either empty or singly occu-
TK = D 0 + J0 ) , (11)
ρ(Jz1 z2 p̃ pied).
e
In the mean field approximation (MFA), the boson
0
with ρ(Jz1 + Jz20
) = 2πΓ |ϵ0 |(ϵU0 +U ) . The above formula field b is replaced by an independent of time real num-
resembles the corresponding one for the Kondo temper- ber, b(t) → ⟨b(t)⟩ ≡ b̄. This approximation, however, re-
ature in a single QD coupled to ferromagnetic leads27 , stricts considerations to the low bias regime (eV ≪ |ϵi |).
4

Introducing now the following renormalized parameters: 2.0


β β
t̄ = tb̄2 , V ik = Vik b̄, and ϵ̄i = ϵi + λ, where λ is the corre- α=1
sponding Lagrange multiplier, one can write the effective α=0.5
MF Hamiltonian as 1.5 t=0 α=0.1
∑∑ ∑

G ( e /h)
Ĥ M F = εkβi c†kβi ckβi + ϵ̄i fi† fi + (t̄f1† f2 + h.c.)

2
k βi i 1.0
∑∑ ( )
(V ik c†kβi fi + h.c.) + λ b̄2 − 1 .
β
+ (13)
k βi 0.5
The unknown parameters, b̄ and λ, have to be found self-
consistently from the following equations; 0.0
-4 -3 -2 -1 0 1 2
∑ ∫ dε ε0/Γ
b̄2 − i ⟨⟨fi |fi† ⟩⟩<
ε = 1, (14)
i

FIG. 3: Linear conductance vs. dots’ level position, ϵ1 =
ϵ2 = ϵ0 , for indicated values of α and t = 0, obtained from
∑ ∫ dε the slave boson method for U → ∞. The level splitting due
−i (ε − ϵ̄i )⟨⟨fi |fi† ⟩⟩< 2
ε + λb̄ = 0, (15) to renormalization is tuned out by external gates.
i

where ⟨⟨fi |fi† ⟩⟩<ε is the Fourier transform of the lesser


′ † ′ V. NON-EQUILIBRIUM GREEN FUNCTION
Green function defined as G< ii (t, t ) ≡ ⟨⟨fi (t)|fi (t )⟩⟩ =
<
† ′ APPROACH
i⟨fi (t )fi (t)⟩. These equations follow from the constraint
imposed on the slave boson field, Eq. (12), and from the
equation of motion for the slave boson operator. The Electric current flowing through a biased system is
determined by nonequillibrium retarded, advanced, and
lesser Green functions ⟨⟨fi |fi† ⟩⟩< ε as well as the retarded lesser Green functions of the dots, and can be calculated
Green functions ⟨⟨fi |fi† ⟩⟩r (the latter ones are required in from the formula derived by Meir et al25 . In turn, to cal-
the further calculations, too) have been determined from r(a)
culate the retarded (advanced) Green functions Gij (ϵ),
the corresponding equations of motion.
we have applied the equation of motion method. Within
The Kondo temperature can be introduced as11
this method one writes first the equation of motion for the
√ causal Green function Gij (ϵ), which generates new Green
2
TK ≡ ϵ̄20 + Γ , (16) functions. Then, one writes the equations of motion for
these new Green functions, which in turn contain new
with Γ = b̄2 (Γ1 + Γ2 ) and ϵ1 = ϵ2 = ϵ0 . Variation of the higher-order Green functions. The latter ones have to be
Kondo temperature (evaluated from the above equation) calculated approximately. To close the set of equations
with the parameter α is shown in Fig.2 (solid line). for the Green functions we have applied the decoupling
To study charge transport we assume the same elec- scheme introduced in Ref.[4]. Although such an approx-
trochemical potentials for the left leads and also equal imation does not describe properly the zero temperature
electrochemical potentials of the right leads. The linear limit, it is sufficient to describe the Kondo phenomenon
conductance is then calculated from the Landauer for- close to the Kondo temperature.
mula, in which the transmission matrix is taken at the The retarded/advanced Green functions contain occu-
Fermi level. The slave boson technique in the form pre- pation numbers, ni , and the interdot correlators, niī =
sented above, however, does not take into account the ⟨d†i dī ⟩, which can be calculated from the identities
level renormalization described in the preceding sections. ∫
Therefore, to include this renormalization we replace the dϵ <
ni = −i G (ϵ), (17)
bare dot levels by the renormalized ones (keeping the 2π ii
notation used for the bare dot levels). Alternatively one
may say that the renormalization is tuned out by external ∫
dϵ <
gate voltages. In Fig.3 the linear conductance is shown as niī = −i G (ϵ). (18)
a function of the dots’ energy level, ϵ1 = ϵ2 = ϵ0 , and for 2π īi
indicated values of the parameter α. The linear conduc-
tance reaches the unitary limit for ϵ0 ≪ −Γ. This limit is Thus, we still need the lesser Green functions G< ij (ϵ).
<
However, one can note that instead of Gij (ϵ), only
achieved owing to the tuning out the level splitting due ∫
to renormalization. From this figure also follows, that dϵ G<
ij (ϵ) is needed. This quantity can be found ex-
the Kondo temperature decreases with decreasing α, in actly (in contrast to the approach based on the Ng’s
agreement with the above discussion and Fig.2. approximation30 ) and to do this we apply the Heisenberg
5

1.0 α=0.05
calculated as
α=0.25
0.6 α=0.5 1
α=0.75
Di = − ℑ [Grii (ε)] , (21)

LDOS
0.5 α=1
π
where ℑ[A] denotes the imaginary part of A.
0.4
LDOS

The approximation scheme used to calculate the


0.0
-0.04 0.00 0.04 nonequillibrium Green functions does not take into ac-
ε/Γ count the level renormalization described in the preced-
0.2 QD1 ing sections. Therefore, to take this renormalization into
account we replace the bare dot levels by the renormal-
ized ones (keeping the notation used for the bare dot
levels), similarly as in the case of slave boson technique.
0.0 In the following numerical calculations we assume equal
0.2 dot energy levels, ϵi = ϵ0 (for i = 1, 2) (ϵ0 is mea-
0.16 α=0.05
α=0.25
sured from the Fermi level of the leads in equilibrium,
α=0.5 µLi = µRi = 0). Apart from this, we assume ϵ0 = −3.5Γ,
LDOS

α=0.75
0.08
α=1
the bandwidth 2D = 500Γ, and U = 50Γ.
LDOS

0.00
-0.04 0.00 0.04
0.1 ε/Γ A. Numerical results for t = 0

QD2
Let us start with the case when the dots are capac-
itively coupled only, t = 0. The LDOS for both dots
is plotted in Fig.4. The spectrum of each dot reveals
two resonances corresponding to the dot level and its
0.0 Coulomb counterpart (the latter not shown). Apart from
-6 -4 -2 0 2 this, a narrow peak emerges in the spectrum of each dot
ε/Γ at the Fermi level of the leads. The intensity and width
of this peak strongly depends on temperature, revealing
FIG. 4: Local density of states for the dots QD1 and QD2, all characteristic features typical of the Kondo resonance.
calculated as a function of energy for indicated values of the The resonance in LDOS originates from the many body
parameter α. The other parameters are ϵ0 = −3.5Γ, U = 50Γ processes which occur in the low temperature regime.
and t = 0.
Since the conditions ϵ0 < µβi and µβi < ϵ0 +U are obeyed
for the parameters assumed (Coulomb blockade regime),
only a single electron can occupy the DQD system and
equation of motion for the operators d†j (t)di (t). Then, sequential tunnelling processes are blocked. However,
one takes average from the obtained equation and makes higher-order tunnelling events are still allowed. Let us
use of the fact that ⟨(d/dt)d†j (t)di (t)⟩ = 0 in a steady assume that an electron initially occupies the dot QD1,
state. As a result one obtains the following equations; and the system is in the Coulomb blockade regime. Due
to the uncertainty principle, the electron from the dot
t(niī + nīi ) − iΓi ni QD1 can tunnel onto the Fermi level of one of the leads
∑ ∫ dϵ attached to QD1, while an electron from the Fermi level
= fβ (ϵ)Γβi (Grii − Gaii ), (19) of one of the leads attached to QD2 can tunnel to the

β dot QD2 in the time ~/|ϵ0 |. Interference of many such
events gives rise to the narrow peaks in LDOS at the
Fermi level.
i
t(ni + nī ) + (ϵī − ϵi )niī − (Γi niī + Γī niī ) For the fully symmetric model (α = 1), LDOS for the
2
∑ ∫ dϵ dot QD1 is the same as that for QD2. The situation is
= fβ (ϵ)[Γβi Griī + Γβī Gaīi ]. (20) different for α ̸= 1. As α decreases, the intensity of the
2π Kondo peak in LDOS of the dot QD2 also decreases and
β
disappears when α tends to zero. The opposite situation
These equations, together with the appropriate equations occurs in the LDOS of the dot QD1, where the Kondo
for the retarded/advanced Green functions, have to be peak becomes more and more pronounced with decreas-
solved numerically in a self-consistent way. ing α. This behavior is due to the fact that the intensity
The basic transport characteristics of the system, like of the Kondo peak in LDOS of the dot QD1 is mainly
conductance and differential conductance can be calcu- determined by the coupling strength between QD2 and
lated numerically using the formulas derived above. The the leads, while the Kondo peak for the dot QD2 is pre-
local density of states (LDOS) for the i-th dot can be dominantly determined by coupling of the dot QD1 to
6

0.8
QD1/QD2 QD1
total 0.6 QD2
0.6 α=1 α=0.75 total

0.10 α=0.8
G (e /h)

0.4
2

0.4

0.2

G (e /h)
0.2

2
0.0 0.0

QD1
0.4
QD1
0.08
QD2 QD2
0.4 total total
α=0.5 α=0.25
G (e /h)

0.2
2

0.2

0.06
-0.4 -0.2 0.0 0.2 0.4
eV/Γ
0.0 0.0
-0.4 -0.2 0.0 0.2 0.4 -0.4 -0.2 0.0 0.2 0.4
eV/Γ eV/Γ
FIG. 6: Differential conductance calculated for indicated
FIG. 5: Differential conductance for the dots QD1 and QD2, value of α in the case when the level slitting due to renormal-
and the total differential conductance calculated for indicated ization is not compensated by external gate voltages. The po-
values of α. The other parameters as in Fig.4. sitions of the dots’ levels have been estimated self-consistently
using Eq.(5). The other parameters as in Fig.4.

the leads. Accordingly, the Kondo peak in LDOS of the


dot QD1 (QD2) increases (decreases) with decreasing α, becomes smaller than 1, α < 1. Interestingly, the con-
while the Kondo peaks of both dots are equal for α = 1. ductance of the dot weakly coupled to the leads (in our
This behavior of the Kondo peaks in LDOS of both dots case the dot QD1) is larger than the conductance of the
is similar to that in the case of spin Kondo effect, where dot strongly coupled to the leads. For a sufficiently small
each spin channel is coupled differently to the leads (when value of α, the differential conductance of the dot QD2
the leads are ferromagnetic). appears as a broad background, whereas the differential
For α ≪ 1, the Kondo peak for the dot QD1 becomes conductance of the dot weakly coupled to the leads is
strongly asymmetric and the LDOS is totally suppressed then very narrow. This behavior follows the features of
for energies above the Fermi level, where the spectral LDOS of the dots QD1 and QD2 discussed above (Fig.4).
function is equal to zero. This situation is similar to Apart from this, we note that the total differential con-
that reported in Ref.[19]. Apart from this, position of ductance diminishes as α decreases, and its line width
the Kondo peak for QD2 slightly moves away from the also shrinks. This behavior indicates on the suppression
Fermi level with increasing α (towards positive energies), of the effective Kondo temperature as α decreases. Such
and becomes asymmetric for all values of α. a behavior stems from the fact that the rate of tunneling
It is also worth to note that position of the broad max- events leading to the Kondo resonance decreases since
imum (associated with the dot’s level) in LDOS of QD1 the dot QD1 becomes detached from the leads as α de-
(the dot whose coupling to the leads changes with α) is al- creases. This is also in agreement with our predictions
most unchanged with tunning α, whereas position of the on the α dependence of the Kondo temperature, derived
broad maximum in the LDOS of the dot QD2 (coupled to in Sections III and IV (see also Fig.2). Finally, when one
the leads with constant strength) varies with the parame- of the dot is totally disconnected from the leads (α = 0),
ter α. This becomes clear when considering the formulas the Kondo temperature vanishes, and no Kondo effect
for renormalized dots’ energy levels, Eq.(5). Apart from appears.
this, the intensity of the broad peak for the dot QD1 When ϵ1 ̸= ϵ2 (eg. when the level slitting due to renor-
decreases monotonically with increasing α, whereas the malization is not compensated by external gate voltages),
intensity of the broad peak in the LDOS of the dot QD2 the Kondo peak in differential conductance becomes split
depends on α in a more complex way. When α → 0, the and the two components are shifted from the Fermi level
broad peak in the LDOS of the dot QD1 is then most pro- and have rather low intensity, as shown in Fig.6. This
nounced, whereas its Coulomb counterpart (not shown) suppression of the Kondo anomaly resembles similar be-
is totally localized at ϵ0 + U . havior in the case of the spin Kondo effect.
As mentioned before, the resonances in LDOS lead to The presence of Kondo peaks also depends on the cou-
zero bias anomaly in the differential conductance of the pling strength of the dots to the leads. Above we assumed
DQD system. In Fig.(5) we show the differential conduc- relatively strong coupling for both dots, with some asym-
tance of both dots as a function of the bias voltage. For metry of this coupling described by the parameter α. We
a fully symmetric system, the differential conductance of have also examined the case when both dots are weakly
both dots is the same, but the situation changes when α coupled to the leads. There is no Kondo effect in such
7

0.12 0.03
α=0 α=0
α=0.05 α=0.05
α=0.25 α=0.25

QD2
0.08 QD1 0.02
LDOS

LDOS
0.04 0.01

0.00 0.00
0.12 0.12
α=0.5 α=0.5
α=0.75 α=0.75
α=1 α=1

QD1 QD2
0.08 0.08
LDOS

LDOS
0.04 0.04

0.00 0.00
-0.4 -0.2 0.0 0.2 0.4 -0.4 -0.2 0.0 0.2 0.4
ε/Γ ε/Γ

FIG. 7: Local density of states for the dot QD1, calculated for FIG. 8: Local density of states for the dot QD2, calculated for
indicated values of α and for t = −0.1Γ. The other parameters indicated values of α and for t = −0.1Γ. The other parameters
as in Fig.4. as in Fig.4.

a case, which is consistent with the recent experimental


observations15 .

the two peaks located at ±2t, one finds an additional


B. Numerical results for the case t ̸= 0 peak in the LDOS of the dot QD1, which is located at
the Fermi level. However, instead of the peak, a dip in
the LDOS of the dot QD2 appears at the Fermi level.
Now, we consider the situation when direct hopping
Possible explanation of this behavior relies on the tran-
between the dots is allowed. Figures 7 and 8 show LDOS
sitions/tunneling events which do not induce electron
for the dots QD1 and QD2. When both dots are equally
exchange between the molecular states ϵ+ and ϵ− , but
coupled to the leads (α = 1), a double peak structure
rather between original bare dot levels.
emerges in the LDOS of the dots QD1 and QD2, and
the LDOS is the same for both dots. The two peaks are
centered at ε = ±2t. This comes from the fact that when Differential conductance for several values of the asym-
t ̸= 0, the dots’ states hybridize into two molecular-like metry parameter α is shown in Fig.9. When the dots are
states with eigenenergies ϵ± = ϵ0 ± t. These new states connected in the T-shape geometry, α = 0, one finds two
are then involved in the Kondo phenomenon. If initially maxima centered at eV = ±2t, and one dip at eV = 0.
an electron occupies the level ϵ− , then it can tunnel into Suppression of the conductance at eV = 0 is a result of
the Fermi level of a given lead and simultaneously an- destructive quantum interference31 . When the coupling
other electron having energy +2t tunnels onto the level to the dot QD1 is turned on, then the dip structure dis-
ϵ+ . Coherent superposition of many such events results appears in the differential conductance. Instead of dip,
in the Kondo peak at the energy ε = 2t. In the same way one finds the third peak centered at eV = 0. For a fully
one may explain the presence of the Kondo peak at the symmetric system, α = 1, this peak vanishes and only
energy ε = −2t. the satellite maxima are present. It is also worth to note
However, the situation changes for α ̸= 1. Apart from that the conductance increases with increasing α.
8

0.06 electron reservoirs. Various techniques have been used to


a) α=0 describe basic features of the Kondo physics. First, we
α=0.05
used the scaling technique to evaluate the level renor-
0.05 malization and the corresponding Kondo temperature.
Then, we used the slave boson technique to calculate lo-
G (e /h)

cal density of states and linear conductance. To find non-


2

0.04 linear conductance we used the nonequillibrium Green


function method.
The numerical results show that transport character-
0.03 istics reveal typical Kondo phenomenon, similar to that
observed in a single quantum dot with spin degenerate
discrete level, coupled to external ferromagnetic leads.
-0.4 -0.2 0.0 0.2 0.4 In the case considered, the splitting due to level renor-
malization could be compensated by external gate volt-
0.50 α=0.25 ages, so one could reach the full Kondo anomaly. Such a
b) α=0.5 compensation, however, is not possible when the direct
α=0.75
tunneling between the dots is strong.
α=1
Acknowledgments
G (e /h)
2

0.25 The author is extremely gratefull to Prof. J. Barnaś for


fruitful discussions and would like to thank him for giv-
ing helpful advices during preparation of the manuscript.
This work, as part of the European Science Foundation
EUROCORES Programme SPINTRA, was supported by
0.00 funds from the Ministry of Science and Higher Educa-
-0.4 -0.2 0.0 0.2 0.4 tion as a research project in years 2006-2009 and the EC
eV/Γ Sixth Framework Programme, under Contract N. ERAS-
CT-2003-980409. The author also acknowledges support
FIG. 9: Differential conductance for the dots QD1 and QD2, by funds from Ministry of Science and Higher Education
and total differential conductance calculated for indicated val- as a research project N N202 169536 in years 2009-2011.
ues of α and for t = −0.1Γ. The other parameters as in Fig.4.

VI. SUMMARY AND CONCLUSIONS APPENDIX: GREEN’S FUNCTIONS

We have considered the orbital Kondo effect in a spin- Here we show explicit form of the derived dots’ Green
less system of two single-level quantum dots connected to functions Gij for i, j = 1, 2;

{[ ] }
1 U ( ) U ( )
Gii = 1+ n A − niī t̃ Ωīī + n t̃ − niī Ai Ωiī , (A.1)
M W ī ī W ī

{ [ ] }
1 U ( ) U ( )
Giī = ni t̃ − nīi Aī Ωīī + 1 + ni Ai − nīi t̃ Ωiī , (A.2)
M W W

where U ( )
Ωiī = t − γi Aī + αī t̃ ,
W
M = Ω11 Ω22 − Ω12 Ω21
W = A1 A2 − t̃2

(0) U ( )
Ωii = ϵ − ϵi − Σii + αi Aī + γī t̃ (0)
A1 = ϵ − ϵ1 − U − Σ11 − Σc11 − Σe22 − Σd22 ,
W
9

α 2 t
(0)
A2 = ϵ − ϵ2 − U − Σ22 − Σc22 − Σf11 − Σd11 , b(n)
Σii = |Vik | (ϵ − ϵkα − ∆ϵ)Fα(n) (ϵkα )
Λ

t̃ = t + Σa22 + Σb11 ,

α1 = ΣdI cI eI
22 + Σ11 + Σ22 ,
c(n)
∑ 2t2 (n)
Σii = |Vik |
α 2
F (ϵkα )
α2 = ΣdI fI cI Λ α
11 + Σ11 + Σ22 , kα

γ1 = ΣaI bI
22 + Σ11 ,

γ2 = ΣaI bI
22 + Σ11 . d(n)
∑ |Vik |
α 2
Σii = F (n) (ϵkα )
ϵ + ϵkα − ϵ1 − ϵ2 − U α
The self-energies are defined in the following way kα


a(n) α 2 t
Σii = |Vik | (ϵ − ϵkα + ∆ϵ)Fα(n) (ϵkα )
Λ

e(n)
∑ t(ϵ − ϵkα )(ϵ − ϵkα + ∆ϵ) − 2t2 (n)
Σii = |Vik |
α 2
Fα (ϵkα )
Λ

f (n)
∑ t(ϵ − ϵkα )(ϵ − ϵkα − ∆ϵ) − 2t2 (n)
Σii = |Vik |
α 2
Fα (ϵkα )
Λ

d(I)
∑ |V α |2 noninteracting system, i.e., U = 0 in the Hamiltonian
ik
Σii = fα (ϵkα )
ϵ − ϵkα (1).

with Λ = (ϵ−ϵkα )[(ϵ−ϵkα +∆ϵ)(ϵ−ϵkα −∆ϵ)−4t2 ], ∆ϵ = Assuming that ϵ1 = ϵ2 = ϵ0 the self-energies including
(n) (n)
ϵ1 − ϵ2 , Fα (ϵkα ) = fα (ϵkα ) for n = I and Fα (ϵkα ) = 1 Fermi distribution function can be calculated analytically
(0)
for n = 0. The self energy Σij is the self energy of the and expressed by means of digamma function.


Electronic address: ptrocha@amu.edu.pl Lett. 89, 136802 (2002); R. Aguado and D. C. Langreth,
1
S. M. Cronenwett et al., Science 281, 540 (1998); S. Sasaki, Phys. Rev. B 67, 245307 (2003).
8
S. De Franceschi, J. M. Elzerman, W. G. van der Wiel, M. T. Aono and M. Eto, Phys. Rev. B 63, 125327 (2001).
Eto, S. Tarucha, and L. P. Kouvenhoven, Nature (London) 9
R. Świrkowicz, J. Barnaś, and M. Wilczyński , Phys. Rev.
405, 764 (2000). B 68, 195318 (2003); R. Świrkowicz, M. Wilczyński, M.
2
J. Gores, D. Goldhaber-Gordon, S. Heemeyer, M. A. Kast- Wawrzyniak, and J. Barnaś, Phys. Rev. B 73, 193312
ner, H. Shtrikman, D. Mahalu, and U. Meirav, Phys. Rev. (2006); R. Świrkowicz, M. Wilczyński, and J. Barnaś, J.
B 62, 2188 (2000). Phys.: Condens. Matter 18, 2291 (2006).
3 10
L. I. Glazman and M. E. Raikh, JETP Lett. 47, 452 (1988); T. Kuzmenko, K. Kikoin, Y. Avishai, Phys. Rev. B 69,
T. K. Ng and P. A. Lee, Phys. Rev. Lett. 61, 1768 (1988). 195109 (2004); Phys. Rev. Lett. 96, 046601 (2006).
4 11
Y. Meir, N. S. Wingreen, and P. A. Lee, Phys. Lett. J. S. Lim, M.-S. Choi, R. López, and R. Aguado, Phys.
66,3048 (1991); Phys. Lett. 70, 2601 (1993). Rev. B 74, 205119 (2006).
5 12
K. Kang and B. I. Min, Phys. Rev. B 52, 10689 (1995). G. Grüner and A. Zawadowski, Rep. Prog. Phys. 37, 1497
6
P. Nordlander, M. Pustilnik, Y. Meir, N. S. Wingreen, and (1974); G. Zarád and A. Zawadowski, Phys. Rev. Lett. 72,
D. C. Langreth, Phys. Lett. 83, 808 (1999). 542 (1994).
7 13
R. Aguado and D. C. Langreth, Phys. Rev. Lett. 85, 1946 D. Boese, W. Hofstetter, and H. Schoeller, Phys. Rev. B
(2000); R. López, R. Aguado, and G. Platero, Phys. Rev. 64, 125309 (2001).
10

14 25
P. G. Silvestrov and Y. Imry, Phys. Rev. B 75, 115335 Y. Meir, N. S. Wingreen, Phys. Rev. Lett. 68, 2512 (1992).
26
(2007). A. C. Hewson, The Kondo Problem to Heavy Fermions
15
A.Hübel, K. Held, J. Weis, and K. v. Klitzing, Phys. Rev. (Cambridge University Press, Cambridge, U.K., 1993).
27
Lett. 101, 186804 (2008). J. Martinek, Y. Utsumi, H. Imamura, J. Barnaś, S.
16
U. Wilhelm et al., Physica (Amsterdam) 14E, 385 (2002). Maekawa, J. König, and G. Schön, Phys. Lett. 91,127203
17
Q.-F. Sun and H. Guo, Phys. Rev. B 66, 155308 (2002). (2003); D. Matsubayashi and M. Eto, Phys. Rev B 75,
18 165319 (2007).
D. Sztenkiel and R. Świrkowicz, J. Phys.: Condens. Matter
28
19, 256205 (2007). V. Kashcheyevs, A. Schiller, A. Aharony, and O. Entin-
19
D. Sztenkiel and R. Świrkowicz, J. Phys.: Condens. Matter Wohlman, Phys. Rev. B 75, 115313 (2007).
29
19, 386224 (2007). P. Coleman, Phys. Rev. B 29, 3036 (1984).
30
20
A. W. Holleitner, A. Chudnovskiy, D. Pfannkuche, K. T. K. Ng, Phys. Rev. Lett. , 3635 (1993).
31
Eberl, and R. H. Blick, Phys. Rev. B 70, 075204 (2004). P. Trocha, J. Barnaś, Phys. Rev. B 76, 165432 (2007).
32
21
S. Lipiński and D. Krychowski, Phys. Status Solidi b 243, C. Lacroix, J. Appl. Phys 53, 2131 (1982).
33
206 (2005). T. A. Costi, J. Phys. C: Solid State Phys. 19, 5665 (1986).
34
22
T. Pohjola, H. Schoeller, and G. Schön, Europhys. Lett., T. Lobo, M.S. Figueira, R. Franco, J. Silva-Valencia, and
54, 241 (2001). M.E. Foglio, Physica B 398, 446 (2007).
35
23
J. Wen, J. Peng, B. Wang, and D. Y. Xing, Phys. Rev. B H.-G. Luo, J.-J. Ying, and S.-J. Wang, Phys. Rev. B 59,
75, 155327 (2007). 9710 (1999).
36
24
T. Kubo, Y. Tokura, and S. Tarucha, Phys. Rev. B 77, T. Lobo, M. S. Figueira, and M. E. Foglio, Nanotechnology
041305(R) (2008). 17, 6016 (2006); ibib 21, 274007 (2010).

You might also like