8
Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker, Andreas Published in: Physica E-Low-Dimensional Systems & Nanostructures DOI: 10.1016/j.physe.2009.06.069 2010 Link to publication Citation for published version (APA): Pedersen, J., & Wacker, A. (2010). Modeling of cotunneling in quantum dot systems. In Physica E-Low- Dimensional Systems & Nanostructures (Vol. 42, pp. 595-599). Elsevier. https://doi.org/10.1016/j.physe.2009.06.069 Total number of authors: 2 General rights Unless other specific re-use rights are stated the following general rights apply: Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal Read more about Creative commons licenses: https://creativecommons.org/licenses/ Take down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim.

Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

LUND UNIVERSITY

PO Box 117221 00 Lund+46 46-222 00 00

Modeling of cotunneling in quantum dot systems

Pedersen, Jonas; Wacker, Andreas

Published in:Physica E-Low-Dimensional Systems & Nanostructures

DOI:10.1016/j.physe.2009.06.069

2010

Link to publication

Citation for published version (APA):Pedersen, J., & Wacker, A. (2010). Modeling of cotunneling in quantum dot systems. In Physica E-Low-Dimensional Systems & Nanostructures (Vol. 42, pp. 595-599). Elsevier.https://doi.org/10.1016/j.physe.2009.06.069

Total number of authors:2

General rightsUnless other specific re-use rights are stated the following general rights apply:Copyright and moral rights for the publications made accessible in the public portal are retained by the authorsand/or other copyright owners and it is a condition of accessing publications that users recognise and abide by thelegal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private studyor research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal

Read more about Creative commons licenses: https://creativecommons.org/licenses/Take down policyIf you believe that this document breaches copyright please contact us providing details, and we will removeaccess to the work immediately and investigate your claim.

Page 2: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

arX

iv:0

904.

3249

v2 [

cond

-mat

.mes

-hal

l] 2

5 M

ar 2

010

Proceedings of FQMT08, published in Physica E 42, 595 (2010), doi:10.1016/j.physe.2009.06.069

Modeling of cotunneling in quantum dot systems

Jonas Nyvold Pedersen and Andreas Wacker∗

Mathematical Physics, Lund University, Box 118, 22100 Lund, Sweden

Transport through nanosystems is treated within the second order von Neumann approach. Thisapproach bridges the gap between rate equations which neglect level broadening and cotunneling,and the transmission formalism, which is essentially based on the single-particle picture therebytreating many-particle interactions on an approximate level. Here we provide an alternative presen-tation of the method in order to clarify the underlying structure. Furthermore we apply it to theproblem of cotunneling. It is shown that both elastic and inelastic cotunneling can be describedquantitatively, while the transmission approach with a mean-field treatment of the interaction pro-vides an artificial bistability.

PACS numbers: 73.23.Hk,73.63.-b

I. INTRODUCTION

Nanostructure technology allows for the fabrication ofsmall structures, such as quantum dots [1], nanowires [2],carbon nanotubes [3], as well as molecular structures,whose electronic properties are dominated by a smallnumber of electrons. The insertion into electrical cir-cuits exhibits current-bias (IV ) relations, which stronglydiffer from conventional resistors. Typical features arethe suppression of current for low bias due to Coulombblockade and pronounced current steps, whose positioncan be easily controlled by a gate bias, thus suggestingtransistor action.Due to the small spatial dimensions, the level quantisa-

tion with a spacing ∆E is of fundamental relevance. Fur-ther relevant energy scales are: The Coulomb interactionbetween electrons (or similarly any other type of many-particle interaction [4]) of the order U = e2/C, whereC is the geometrical capacitance; the tunneling rate Γ/~for the transition of particles between the structure andthe contacts; as well as the thermal energy kBT . Fortypical nanostructured electronic systems studied exper-imentally, these energies are all in the range of 0.1–10meV, where Γ can be even smaller. Generic examples ofIV -characteristics are shown in Fig. 1.The task to describe the current through such struc-

tures quantitatively is an evolved topic and a large varietyof approaches has been used within the last two decades.The general starting point is to divide the Hamiltonianof the system as

H = HD +Hleads +HT . (1)

Here HD describes the central region (such as a singleor multiple quantum dot structure, a molecule, a car-bon nanotube, or a nanowire), which we refer to as thequantum dot in the following.

Hleads =∑

kσℓ

Ekσℓc†kσℓckσℓ, (2)

∗Electronic address: [email protected]

bias

curr

ent

different gate bias

currentcotunneling

broadening kBT+Γ

Level spacing ∆E

Coulomb repulsion U(+∆E)

blockade withCoulomb

FIG. 1: Schematic current-bias relation for a quantum dot,showing the qualitative influence of the key energy scales.

describes the leads, where σ =↑, ↓ denotes the spin, klabels the spatial wave functions of the contact states,and ℓ denotes the lead (typically ℓ = L/R for the leftand right lead, respectively). In most works a thermaloccupation of the leads

c†kσℓckσℓ

= fkσℓ =1

e(Ekσℓ−µℓ)/kBT + 1,

with chemical potential µℓ is assumed, where the biaseVbias = µL − µR drives the current in a two termi-nal system. Conventionally noninteracting leads areassumed.[37] Finally, HT describes the coupling betweenthe leads and the quantum dot.If HD does not contain many-particle interactions, it is

straightforward to derive a transmission formula (see [6]and references therein) yielding a particle current fromlead ℓ into the quantum dot

Jℓ =1

2π~

dE Tℓ→ℓ′(E) [fℓ(E) − fℓ′(E)] . (3)

Here the transmission function is determined by thesingle-particle spectrum of HD in connection with the

Page 3: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

2

coupling HT , which essentially broaden the levels. A for-mal tool is the nonequilibrium Green function formal-ism [7]. An inclusion of many-particle effects withinthe mean-field level is straightforward and exchange-correlation effects can be taken into account by moreinvolved mean-field potentials [8]. Also time-dependentsimulations are possible using time-dependent densityfunctional theory [9]. However, hardly any results ex-ist, which go beyond mean-field (a rather old example is[10]). Thus this approach works for arbitrary tempera-tures if correlation effects due to the Coulomb interactionin the system are small.Another approach is to start from the von-Neumann

equation

i~d

dtρ = −[ρ,H ] (4)

treating the time evolution of the density operator ρ con-taining both the system and the leads. As only the time-evolution of the system is of interest, one can trace-outthe lead degree of freedom in order to obtain the reduceddot density operator

ρdot(t) = Trleads {ρ(t)} . (5)

The key task is to find an approximate equation of motionfor ρdot(t) taking into account the coupling HT . Thisprovides (generalized) master equations [11].To lowest order in HT electron transitions between

the quantum dot and the leads are treated indepen-dently from each other (here called first order follow-ing the number of correlated transitions). Restrictingto diagonal elements of ρdot(t), one obtains to lowest or-der the Pauli master equations [12].[38] Furthermore, co-herences, represented by nondiagonal matrix elements,can be treated in different ways (frequently also denotedquantum master equations). The traditional approach isthe Wangness-Bloch Redfield kinetics [14, 15] which hasbeen applied to quantum dots in [16, 17]. While reason-able results are typically found, there is a fundamentalproblem due to the occurrence of negative occupationsas the equations are not of Lindblad type [18]. Lindblad-type kinetic equations were derived in [19, 20], but onlyfor the high-bias limit. All first-order approaches entirelyneglect broadening effects but can treat the interactionsin the system exactly by a diagonalization of the isolateddot Hamiltonian [21]

HD =∑

a

Ea|a〉〈a|, (6)

where the states |a〉 are many-particle states, which maybe highly correlated.In order to treat both broadening and interaction, we

introduced a second order approach based on the vonNeumann equation (4) which considers both the dynam-ics of ρdot(t) and higher-order tunneling transitions [22].This second order von Neumann approach (2vN) bridgesthe gap between the transmission formalism, which is

fully reproduced for systems without interactions at ar-bitrary temperatures, and the first order generalized mas-ter equation schemes, which are recovered in the limitsof high temperature or large bias.

In this paper we provide a slightly different presenta-tion of our 2vN method in Sec. II in order to highlightits structure. Its range of validity and previous resultsare summarized in Sec. III. In Sections IV,V we shownew results demonstrating its applicability to the casesof elastic and inelastic cotunneling, respectively. The fail-ure of mean-field models, providing a fictitious bistabilityis addressed in Sec. V as well.

II. THE SECOND ORDER VON NEUMANN

APPROACH

Using HD in its diagonal representation (6), the tun-neling between the states in the leads and the dot reads(see Appendix A of [22])

HT =∑

kσℓ,ab

Tba(kσℓ)|b〉〈a|ckσℓ + h.c. (7)

The matrix element Tba(kσℓ) is the scattering amplitudefor an electron in the state kσℓ tunneling from the leadonto the dot, thereby changing the dot state from thestate |a〉 to the state |b〉. Note that this amplitude van-ishes unless the number of electrons in state |b〉, Nb,equals Na + 1. Here we use the convention that the par-ticle number increases with the position in the alphabetof the denoting letter.

A general state vector for the entire system is writtenas |ag〉 = |a〉 ⊗ |g〉, with |g〉 = |{Nkℓσ}〉 denotingthe state of both leads, where Nkσℓ ∈ {0, 1}. Toensure the anti-commutator rules of the operators weuse the following notation |g − kσℓ〉 ≡ ckσℓ|g〉 and

|g + kσℓ〉 ≡ c†kσℓ|g〉. The order of indices is opposite tothe order of the operators. E.g. |g − k′σ′ℓ′ + kσℓ〉 =

c†kσℓck′σ′ℓ′ |g〉 = −ck′σ′ℓ′c†kσℓ|g〉 = −|g + kσℓ − k′σ′ℓ′〉,

taking into account the anti-commutation rules of theoperators. To simplify the notation, σℓ is only attachedto k the first time the index k appears in the equation,and in the following it is implicitly assumed to beconnected with k. Furthermore, we apply the conventionthat

kσ(ℓ) means summing over k and σ with a fixed ℓ

being connected to k in this sum.

The density matrix elements are defined as

ρ[n]ag;bg′ = 〈ag|ρ|bg′〉 (8)

where the label n provides the number of electron or holeexcitations needed to transform g into g′. Examples are

ρ[1]bg−k;ag and ρ

[2]cg−k;ag+k′ . We denote the elements as n-

ehx elements in the following.

Page 4: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

3

The von Neumann equation (4) gives equations of mo-tion of the type

i~d

dtρ[0]bg;b′g = (Eb − Eb′)ρ

[0]bg;b′g

+∑

a,kσℓ

Tba(k)ρ[1]ag+k;b′g +

c,kσℓ

T ∗cb(k)ρ

[1]cg−k;b′g

−∑

c,kσℓ

ρ[1]bg;cg−kTcb′(k)−

a,kσℓ

ρ[1]bg;ag+kT

∗b′a(k), (9)

and

i~d

dtρ[1]cg−kσℓ;bg = (Ec − Eb − Ek)ρ

[1]cg−k;bg

+∑

b′

Tcb′(k)δNk,1ρ[0]b′g;bg −

c′

ρ[0]cg−k;c′g−kTc′b(k)

+∑

k′σ′ℓ′

[

b′

Tcb′(k′)ρ

[2]b′g−k+k′ ;bg +

d

T ∗dc(k

′)ρ[2]dg−k−k′ ;bg

−∑

c′

ρ[2]cg−k;c′g−k′Tc′b(k

′)−∑

a

ρ[2]cg−k;ag+k′T

∗ba(k

′)]

.

(10)

Thereby we obtain a hierarchy of n-ehx density matrixelements, where the n-ehx elements show a phase rotationdue to the energy difference involved and are coupled todifferent (n − 1) and (n + 1)-ehx elements. In order tobreak the infinite hierarchy we neglect all n-ehx elementswith n ≥ 3 and obtain a closed set of equations.For all density matrix elements we perform a sum over

all possible lead configurations g and define

Φ[0]b′b =

g

ρ[0]b′g;bg, (11)

which are the elements of the reduced density matrixρdot, and

Φ[1]ba (kσℓ) =

g

ρ[1]bg−k;ag, (12)

which describe the transitions of electrons between theleads and the quantum dot. The particle current fromthe lead ℓ into the structure, Jℓ, equals the rate of changeof the occupation in the lead. This gives

Jℓ = −d

dt

kσ(ℓ)

c†kck

= −d

dt

g,,b,kσ(ℓ)

δNk,1ρ[0]bg,bg

= −2

~

kσ(ℓ),cb

{

g

T ∗cb(k)ρ

[1]cg−k;bg

}

= −2

~

kσ(ℓ),cb

ℑ{

T ∗cb(k)Φ

[1]cb (k)

}

.

(13)

Thus the Φ[1]ba (kσℓ) terms describe current ampli-

tudes. Note that all 1-ehx terms can be described by

Φ[1]ba (kσℓ) as ρ

[1]ag;b−kg =

(

ρ[1]bg−k;ag

)∗

and∑

g ρ[1]ag+k;bg =

g′ ρ[1]ag′;bg′−k.

Performing the sum∑

g to the time evolution (10)

some terms do not directly obtain the form Φ[n]ba . Here

we approximate

g

δNk,1ρb′g;bg ≈ fk∑

g

ρb′g;bg = fkΦ[0]b′b,

g

ρ[0]cg−k;c′g−k =

g′

δNk,0ρ[0]cg′;c′g′ ≈ (1 − fk)Φ

[0]cc′ .

Similar approximations are done for the ρ[1]bg−k;ag ele-

ments appearing in the equation for the 2-ehx terms.This approximation is a factorization of the lead occu-pations, which are assumed not to be affected by thetransition processes. The result is a closed set of dif-

ferential equations for the reduced density matrix Φ[0]b′b,

the current elements Φ[1]ba(kσℓ) and the similarly defined

2-ehx terms Φ[2]b′b(−kσℓ+k′σ′ℓ′; 0), Φ

[2]ca(−kσℓ−k′σ′ℓ′; 0).

Defining a discrete set of k-states, one can set up a col-umn vector consisting of all the elements of the density-matrix Φ =

(

Φ[0],Φ[1],Φ[2]

)

, where the sub-vectors con-tain all the elements of the density-matrix with a specificn-value, as well as the complex conjugates of the complexelements. The equation of motion for the vector Φ canbe cast on a matrix form

i~d

dtΦ =

E00 M01 0

M10 E11 M12

0 M21 E22

Φ = MΦ. (14)

The submatricesEnn are diagonal and contain the energydifferences between the states involved.Now we consider the Φ[2] terms in a stationary approx-

imation, yielding

Φ[2] =

(

−E22 + i0+)−1

M21Φ[1],

where the i0+ ensures causality, corresponding to theMarkov limit for the highest-order elements.Inserting the result into Eq. (14) leads to the matrix

equation

i~d

dt

(

Φ[0]

Φ[1]

)

=

(

E00 M01

M10 M11

)(

Φ[0]

Φ[1]

)

, (15)

where M11 = E11 +M12 (−E22 + i0+)−1

M21 is not di-agonal. Due to the k-dependence it is very laborious toexpress Φ[1] solely in terms of Φ[0] thereby reducing theproblem to a generalized master equation. The explicit,and completely general, expressions for the equation ofmotion for Φ[0] and Φ

[1] are given in Eqs. (10,11) of [22]and is the main result of that paper.[39]The sub-matrices M10 and M01 only contain elements

proportional to the tunneling amplitude Tba, while thematrix M11 involves terms proportional to T 2

ba. Thus the

Page 5: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

4

stationary solution of Eq. (15), together with the normal-

ization∑

b Φ[0]bb = 1, will contain all powers of Tba, and

so will the stationary occupations and coherences. Thatis, the approach does not provide a systematic expansionin powers of the tunneling coupling. For the current,which is proportional to TbaΦ

[1], all terms up to orderT 4ba as well as a class of higher-order terms are taken into

account.Summarizing, the 2vN approach considers the reduced

density matrix (11) and the current amplitudes (12) asvariables, which are determined by the closed set of dy-namical equations (15). In deriving these, three approxi-mations are applied; (i) only coherent processes involvingtransitions of at most two different k-states are consid-ered, i.e. all n-ehx terms with n ≥ 3 are assumed tovanish, (ii) the time-dependence of terms generating 2-ehx terms is neglected which corresponds to the Markovlimit, (iii) the level occupations in the leads, fkσℓ, is un-affected by the couplings to the dot, i.e. the densities inthe leads and on the dot can be factorized.

III. RANGE OF VALIDITY AND COMPARISON

WITH OTHER METHODS

For noninteracting systems, Eq. (3) gives the correctresult for arbitrary temperature and bias. In [22] wedemonstrated analytically that using the 2vN approach,Eq. (3) is fully recovered for a single-level system. Nu-merically, we also found full agreement for all ranges ofparameters, including double dots [23] and the ferromag-netic Anderson model with an applied magnetic field [24],where coherences are of central importance.[40] Thus wehave strong indications that the 2vN method is able totreat transport correctly for noninteracting systems overthe full temperature and bias ranges.For interacting systems and temperatures kBT ≫ Γ

or in the high bias limit, the 2vN reproduces the well-known results of the methods presented in [20, 25]. Thisdemonstrates the correct treatment of charging effects ininteracting systems.For the Anderson model with infinite Coulomb repul-

sion, the 2vN equations could be solved analytically, see[23], and the result agree with the diagrammatic real-time transport theory in the resonant tunneling approxi-mation [26], where the onset of Kondo physics is observed(see also Eq. (2) of [27]). However, in the Kondo limititself, the model misses the unitary limit and unphysicalresults are found, as the strong correlations between thedot and the leads are not properly reflected.The validity of the 2vN approach for time-dependent

problems has not been carefully investigated. As theMarkov approximation is invoked, it might not bevalid for strongly time-dependent systems, where non-Markovian effects are important due to memory effects,which are also relevant when evaluating higher-order mo-ments, as e.g., the noise [28, 29]. As an example, thecurrent through a single spinless level was presented in

[22] which does not exhibit the oscillations found from atime-dependent Green function approach [30].In a recent paper, Jin et al. also consider quantum

transport in the same spirit as in the 2vN approach bykeeping correlations between the leads and the dot andperforming an expansion in the tunneling Hamiltonian[31]. They report a proof that they obtain the 2vN ap-proach as a second-order expansion.

IV. ELASTIC COTUNNELING

As an example we apply the 2vN method in the elasticcotunneling regime for a two-level spinless system. Weshow that the 2vN results agree with a mean-field solu-tion and with a scattering result.The system is described by the Hamiltonian

H =∑

k,ℓ=L,R

Ekℓc†kℓckℓ +

kℓn

[

Vkℓnc†kℓdn + h.c

]

+∑

n

End†ndn + Ud†1d1d

†2d2,

(16)

with n = 1, 2 denoting the two dot states. The firstterm is the Hamiltonian of the leads, and the last twoterms are the two single-particle states of the dot andthe interaction between them. The second term is thetunnel Hamiltonian with the tunneling amplitudes Vkℓn.Below it is assumed that Vkℓn = xℓntk, i.e. thecouplings between both dot states and the lead stateshave a fixed phase factor xℓn, and tk is assumed to bea real number. The coupling constants are defined asΓℓn(E) = 2π

k |Vkℓn|2δ(E − Ekℓn) = |xℓn|

2Γ(E). Forthe 2vN calculations a constant value Γ for |E| ≤ 0.95Wis used, and it is assumed that Γ(E) = 0 for |E| > W .For 0.95W < |E| < W an elliptic interpolation is applied.Using the Hamiltonian above, the system can be in fourdifferent many-particle states, denoted |0〉, |1〉, |2〉, |d〉 =

d†2d†1|0〉, with energies 0, E1, E2, Ed, respectively.The transport is calculated in a setup with E1 ≪ µℓ

and E2 + U ≫ µℓ such that the state |1〉 is the groundstate. Sequential tunneling processes are blocked dueto the Coulomb interaction between the electrons, but aleakage current due to elastic cotunneling processes canoccur. Here, the 2vN results are compared with a mean-field solution embedded in a nonequilibrium Green func-tion framework, and a scattering formalism.In the mean-field solution, the interaction term in the

Hamiltonian is replaced with

Ud†1d1d†2d2 → U

{[

d†1d1

d†2d2

+ d†2d2

d†1d1

⟩]

−[

d†1d2

d†2d1

+ d†2d1

d†1d2

⟩]}

,

(17)

where the first [. . .]-bracket is the Hartree term andthe second the Fock term. The occupations are cal-culated self-consistently whereafter the current can beevaluated.[41]

Page 6: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

5

For the scattering result, the elastic second-order scat-tering rate, γRL

11 , is the sum over all processes where anelectron labelled k′L has been transferred from the leftcontact to the state kR in the right contact, leaving thedot state unchanged. This can be realized in two differ-ent ways, |1〉 → |0〉 → |1〉 and |1〉 → |d〉 → |1〉, where theamplitudes for these two processes are added coherently.The rate is calculated as follows [32]: Assume that

initially the dot is in state |1〉 and the leads are in a state|νLνR〉, i.e. the initial state is |i〉 = |νLνR1〉 with energyEi = EνLνR + E1. The probability for the leads to be inthe state νLνR is denotedWνLνR = WνLWνR , as the leadsare assumed uncorrelated. In the final state an electronk′L has been transferred from left to right ending up in

the state kR, i.e. the final state is |fkk′ 〉 = c†kRck′L|i〉with energy Ei−Ek′L+EkR. According to the T -matrixformalism the second-order scattering rates are [32]

γRL11 = 2π

kk′

νLνR

WνLνR

〈fkk′ |HT1

Ei −H0HT |i〉

2

× δ(Efkk′

− Ei),

(18)

and after some algebra we arrive at

γRL11 =

Γ2

dE∣

x∗L1xR1

E1 − E+

xR2x∗L2

E2 + U − E

2

× nF (E − µL)[1 − nF (E − µR)],

(19)

where nF (E) = [1 + eE/kBT ]−1 is the Fermi function,and energy-independent coupling constants are assumed(the Wide-Band Limit). The rate γLR

11 is found by inter-changing L ↔ R. Below we only calculate the integralfor kBT = 0, but note that for finite temperatures the in-tegral diverges and a regularization procedure is needed[33].For kBT = 0, µR = eVbias > 0 and µR = 0, we obtain

γRL11 =

Γ2

∫ eVbias

0

dE∣

x∗L1xR1

E1 − E+

xR2x∗L2

E2 + U − E

2

, (20)

and γLR11 = 0. Assuming the state |1〉 to be almost com-

pletely occupied (i.e. |E1|/Γ ≪ 1 and (E2 + U)/Γ ≫ 1),the second-order elastic cotunneling current is Iel.cotun =1~γRL11 .In Fig. 2 the current versus bias voltage is calculated in

the elastic cotunneling regime using the 2vN method, themean-field approximation in a Green function framework,and the second-order scattering method. For the latter,the calculation is for vanishing temperature, while theother calculations are for kBT = Γ/10. For E1 = −3Γalmost perfect agreement between all three methods isfound, while deviations between the scattering methodand the others occur for E1 = −2Γ, which is most likelydue to the fact that the state |1〉 is not fully occupied inthis case as assumed in the derivation of the second-orderscattering rate (19).

0 5 10 150

0.005

0.01

0.015

0.02

0.025

eVbias = µL (Γ)

Curr

ent

[

Γ h

]

MF, E1 = −3Γ

MF, E1 = −2Γ

2vN, E1 = −3Γ

2vN, E1 = −2Γ

Iel.cotun, E1 = −3Γ

Iel.cotun, E1 = −2Γ

FIG. 2: (Color online) The IV -characteristics for two differentvalues of E1 calculated using the 2vN approach, mean-field(MF) nonequilibrium Green functions, and, finally, the scat-tering cotunneling current Iel.cotun = 1

~(γRL

11 − γLR11 ) , where

the latter is calculated for kBT = 0. The chemical potentialsare µL = eVbias and µR = 0, and the phase factors are allequal xℓn = 1/

2, i.e. Γℓn = Γ/2. The other parameters are:E2 = 16Γ, U = 20Γ, kBT = Γ/10 and W = 50Γ.

In summary, from a comparison with both a mean-field solution and a scattering formalism, we have shownthat the 2vN method is able to quantitatively describeelastic cotunneling processes even for temperatures muchlower than the energy scale set by the coupling to leads,kBT ≪ Γ.

V. INELASTIC COTUNNELING

For higher bias voltages the agreement between the2vN approach and the mean-field solution is assumed tobe less perfect, as the latter can lead to bistable solutions[34], which are not present in a generalized master equa-tion approach. Fig. 3 shows a comparison between the2vN approach and the mean-field solution over the fullbias range for E1 = −2Γ, with the rest of the parametersas in Fig. 2.Considering first the 2vN result, the curve shows in-

creased current when the bias matches the energy differ-ence between the levels, eVbias = |E1−E2| = 18Γ. This isdue to the onset of inelastic cotunneling [32], which leadsto a population of the excited state, |2〉. After the inelas-tic cotunneling process, additional cotunneling-assistedsequential tunneling processes can occur [35, 36]. Finally,at eVbias = E2 + U = 36Γ sequential tunneling throughthe upper level becomes possible. The value for the cur-rent is consistent with the master equation result (3Γ/8~)plus additional cotunneling through the lower level.

Page 7: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

6

0 10 20 30 400

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.5

eVbias = µL (Γ)

Curr

ent

[

Γ h

]

2vNMean−field

FIG. 3: (Color online) The IV -characteristic over the full biasrange for E1 = −2Γ calculated within the mean-field approx-imation and using the 2vN approach. The other parametersare as in Fig. 2. The arrows indicate the direction of the biassweep in the mean-field calculation.

In contrast, the mean-field solution misses the onset ofinelastic cotunneling and instead a bistable behaviour isobserved, and it also overestimates the current after theonset of sequential tunneling. Both is due to an insuf-ficient description of the Coulomb interaction and em-phasizes the need for a method which can describe bothhigher-order tunneling processes (as elastic and inelasticcotunneling) and many-particle interactions.

VI. SUMMARY

The second order von Neumann approach providesa quantitative description of transport through nanos-tructures for arbitrary bias and temperatures above theKondo temperature. In the low-bias regime, the elasticcotunneling current is in very good agreement with boththe second-order scattering result and the mean-field so-lution, even for temperatures much lower than the energyscale set by the coupling to leads. Inelastic cotunneling isalso well captured by the 2vN approach, whereas a mean-field solution within a nonequilibrium Green functionsframework shows an artificial bistability in this regime.

[1] D. Bimberg, M. Grundmann, N. Ledentsov, QuantumDot Heterostructures, John Wiley & Sons Ltd., NewYork, 1999.

[2] C. Thelander, P. Agarwal, S. Brongersma, J. Eymery,L. Feiner, A. Forchel, M. Scheffler, W. Riess, B. Ohlsson,U. Gosele, L. Samuelson, Materials Today 9 (2006) 28.

[3] M. S. Dresselhaus, G. Dresselhaus, Carbon Nanotubes,Springer, Berlin, 2001.

[4] K. Capelle, M. Borgh, K. Karkkainen, S. M. Reimann,Phys. Rev. Lett. 99 (2007) 010402.

[5] I. Knezevic, Phys. Rev. B 77 (2008) 125301.[6] S. Datta, Electronic Transport in Mesoscopic Systems,

Cambridge University Press, Cambridge, 1995.[7] H. Haug, A.-P. Jauho, Quantum Kinetics in Transport

and Optics of Semiconductors, Springer, Berlin, 1996.[8] M. Brandbyge, J.-L. Mozos, P. Ordejon, J. Taylor,

K. Stokbro, Phys. Rev. B 65 (2002) 165401.[9] G. Stefanucci, C.-O. Almbladh, S. Kurth, E. K. U. Gross,

A. Rubio, R. van Leeuwen, N. E. Dahlen, U. von Barth,Lect. Notes Phys. 706 (2006) 479.

[10] K. Makoshi, T. Mii, Surface Science 357 (1996) 335.[11] H.-P. Breuer, F. Petruccione, Open Quantum Systems,

Oxford University Press, Oxford, 2006.[12] C. W. J. Beenakker, Phys. Rev. B 44 (1991) 1646.[13] H. Schoeller, G. Schon, Phys. Rev. B 50 (1994) 18436.[14] R. K. Wangsness, F. Bloch, Phys. Rev. 89 (1953) 728.[15] A. G. Redfield, IBM Journal of Research and Develop-

ment 1 (1957) 19.[16] U. Harbola, M. Esposito, S. Mukamel, Phys. Rev. B 74

(2006) 235309.[17] C. Timm, Phys. Rev. B 77 (2008) 195416.[18] G. Lindblad, Commun. math. Phys. 48 (1976) 119.

[19] T. H. Stoof, Y. V. Nazarov, Phys. Rev. B 53 (1996) 1050.[20] S. A. Gurvitz, Y. S. Prager, Phys. Rev. B 53 (1996)

15932.[21] J. M. Kinaret, Y. Meir, N. S. Wingreen, P. A. Lee, X.-G.

Wen, Phys. Rev. B 46 (1992) 4681.[22] J. N. Pedersen, A. Wacker, Phys. Rev. B 72 (2005)

195330.[23] J. N. Pedersen, B. Lassen, A. Wacker, M. H. Hettler,

Phys. Rev. B 75 (2007) 235314.[24] J. N. Pedersen, D. Bohr, A. Wacker, T. Novotny,

P. Schmitteckert, K. Flensberg, Phys. Rev. B 79 (2009)125403

[25] Y. V. Nazarov, Physica B 189 (1993) 57.[26] J. Konig, H. Schoeller, G. Schon, Phys. Rev. Lett. 76

(1996) 1715.[27] J. Konemann, B. Kubala, J. Konig, R. J. Haug,

Phys. Rev. B 73 (2006) 033313.[28] A. Braggio, J. Konig, R. Fazio, Phys. Rev. Lett. 96 (2006)

026805.[29] C. Flindt, T. Novotny, A. Braggio, M. Sassetti, A.-P.

Jauho, Phys. Rev. Lett. 100 (2008) 150601.[30] G. Stefanucci, C.-O. Almbladh, Phys. Rev. B 69 (2004)

195318.[31] J. Jin, X. Zheng, Y. Yan, Journal of Chemical Physics

128 (2008) 234703.[32] H. Bruus, K. Flensberg, Many-Body Quantum Theory in

Condensed Matter Physics, 1st Edition, Oxford Univer-sity Press, New York, 2004.

[33] M. Turek, K. A. Matveev, Phys. Rev. B 65 (2002) 115332.[34] B. Horvath, B. Lazarovits, O. Sauret, G. Zarand,

Phys. Rev. B 77 (2008) 113108.[35] V. N. Golovach, D. Loss, Phys. Rev. B 69 (2004) 245327.

Page 8: Modeling of cotunneling in quantum dot systems Pedersen, …lup.lub.lu.se/search/ws/files/3614385/1580502.pdf · Modeling of cotunneling in quantum dot systems Pedersen, Jonas; Wacker,

7

[36] R. Schleser, T. Ihn, E. Ruh, K. Ensslin, M. Tews,D. Pfannkuche, D. C. Driscoll, A. C. Gossard,Phys. Rev. Lett. 94 (2005) 206805.

[37] A finite lifetime of the lead states has recently been ad-dressed in [5].

[38] A higher-order expansion can be done within the dia-grammatic real-time transport theory [13].

[39] In [22], Φ[0]b′b

and Φ[1]ba(kσℓ) are denoted wb′b and φba(kσℓ),

respectively.[40] For practical reasons only two-level systems have been

studied so far.[41] The calculations are analogous to those presented in [24].