4
Thermoelectric effects in quantum dots M. Yoshida a, , L.N. Oliveira b a Unesp-Universidade Estadual Paulista, Departamento de F´ ısica, IGCE-Rio Claro, Brazil b Instituto de F´ ısica de S ~ ao Carlos, USP, Brazil article info PACS: 72.15.Qm 73.23.Hk 73.50.Lw Keywords: Kondo effect Fano interference Thermopower Numerical renormalization-group abstract We report a numerical renormalization-group study of the thermoelectric effect in the single-electron transistor (SET) and side-coupled geometries. As expected, the computed thermal conductance and thermopower curves show signatures of the Kondo effect and of Fano interference. The thermopower curves are also affected by particle–hole asymmetry. & 2009 Elsevier B.V. All rights reserved. 1. Introduction The transport properties of mesoscopic devices are markedly affected by electronic correlations. Gate potentials applied to such devices give experimental control over effects once accessible only in special arrangements. In particular, the Kondo effect and Fano anti-resonances have been unequivocally identified in the conductance of single-electron transistors (SET) [1–3]; of Ahar- onov–Bohm rings [4]; and of quantum wires with side-coupled quantum dots [5]. Another achievement was a recent study of the thermopower, a quantity sensitive to particle–hole asymmetry that monitors the flux of spin entropy [6]. This work presents a numerical renormalization-group study [10–12] of the thermo- electric properties of nanodevices. We consider a quantum dot coupled to conduction electrons in the two most widely studied geometries: the single-electron transistor (SET), in which the quantum dot bridges two-dimensional gases coupled to electro- des; and the T-shaped device, in which a quantum dot is side- coupled to a quantum wire. 2. Thermoelectric properties Thermoelectric properties are traditionally studied in two arrangements: the Seebeck (open circuit) and Peltier (closed circuit) setups [13]. In the former, the steady-state electric current vanishes. A temperature gradient drives electrons towards the coldest region, and induces an electric potential difference between the hot and the cold extremes. The expression S ¼DV =DT , where DV is the potential difference induced by the temperature difference DT , then determines the thermopower S. Since the electrons transport heat, the heat current Q can also be measured, and the thermal conductance k can be obtained from the relation Q ¼krT . In the Peltier setup, a current J is driven through a circuit kept at uniform temperature. The heat flux Q ¼ PJ is then measured and determines the Peltier coefficient P, which is proportional to the thermopower: P ¼ ST . We prefer the Seebeck setup. The transport coefficients are then computed from the integrals [7] I n ðT Þ¼ 2 h Z þD D e n @f ðeÞ @e Tðe; T Þ de ðn ¼ 0; 1; 2Þ; ð1Þ where Tðe; T Þ is the transmission probability at energy e and temperature T, f ðeÞ is the Fermi distribution and D is the half width of the conduction band. The electric conductance G, the thermo- power S, and the thermal conductance k are given by [7] G ¼ e 2 I 0 ðT Þ; ð2Þ S ¼ I 1 ðT Þ eTI 0 ðT Þ ; ð3Þ k ¼ 1 T I 2 ðT Þ I 2 1 ðT Þ I 0 ðT Þ ; ð4Þ ARTICLE IN PRESS Contents lists available at ScienceDirect journal homepage: www.elsevier.com/locate/physb Physica B 0921-4526/$ - see front matter & 2009 Elsevier B.V. All rights reserved. doi:10.1016/j.physb.2009.07.118 Corresponding author. E-mail address: [email protected] (M. Yoshida). Physica B 404 (2009) 3312–3315

Thermoelectric effects in quantum dots

Embed Size (px)

Citation preview

Page 1: Thermoelectric effects in quantum dots

ARTICLE IN PRESS

Physica B 404 (2009) 3312–3315

Contents lists available at ScienceDirect

Physica B

0921-45

doi:10.1

� Corr

E-m

journal homepage: www.elsevier.com/locate/physb

Thermoelectric effects in quantum dots

M. Yoshida a,�, L.N. Oliveira b

a Unesp-Universidade Estadual Paulista, Departamento de Fısica, IGCE-Rio Claro, Brazilb Instituto de Fısica de S ~ao Carlos, USP, Brazil

a r t i c l e i n f o

PACS:

72.15.Qm

73.23.Hk

73.50.Lw

Keywords:

Kondo effect

Fano interference

Thermopower

Numerical renormalization-group

26/$ - see front matter & 2009 Elsevier B.V. A

016/j.physb.2009.07.118

esponding author.

ail address: [email protected] (M. Yoshid

a b s t r a c t

We report a numerical renormalization-group study of the thermoelectric effect in the single-electron

transistor (SET) and side-coupled geometries. As expected, the computed thermal conductance and

thermopower curves show signatures of the Kondo effect and of Fano interference. The thermopower

curves are also affected by particle–hole asymmetry.

& 2009 Elsevier B.V. All rights reserved.

1. Introduction

The transport properties of mesoscopic devices are markedlyaffected by electronic correlations. Gate potentials applied to suchdevices give experimental control over effects once accessibleonly in special arrangements. In particular, the Kondo effect andFano anti-resonances have been unequivocally identified in theconductance of single-electron transistors (SET) [1–3]; of Ahar-onov–Bohm rings [4]; and of quantum wires with side-coupledquantum dots [5]. Another achievement was a recent study of thethermopower, a quantity sensitive to particle–hole asymmetrythat monitors the flux of spin entropy [6]. This work presents anumerical renormalization-group study [10–12] of the thermo-electric properties of nanodevices. We consider a quantum dotcoupled to conduction electrons in the two most widely studiedgeometries: the single-electron transistor (SET), in which thequantum dot bridges two-dimensional gases coupled to electro-des; and the T-shaped device, in which a quantum dot is side-coupled to a quantum wire.

2. Thermoelectric properties

Thermoelectric properties are traditionally studied in twoarrangements: the Seebeck (open circuit) and Peltier (closedcircuit) setups [13]. In the former, the steady-state electric current

ll rights reserved.

a).

vanishes. A temperature gradient drives electrons towards thecoldest region, and induces an electric potential differencebetween the hot and the cold extremes. The expressionS ¼ �DV=DT , where DV is the potential difference induced bythe temperature difference DT , then determines the thermopowerS. Since the electrons transport heat, the heat current Q can also bemeasured, and the thermal conductance k can be obtained fromthe relation Q ¼ �krT .

In the Peltier setup, a current J is driven through a circuit keptat uniform temperature. The heat flux Q ¼ PJ is then measuredand determines the Peltier coefficient P, which is proportional tothe thermopower: P ¼ ST .

We prefer the Seebeck setup. The transport coefficients arethen computed from the integrals [7]

InðTÞ ¼ �2

h

Z þD

�Den @f ðeÞ

@e Tðe; TÞde ðn ¼ 0;1;2Þ; ð1Þ

where Tðe; TÞ is the transmission probability at energy e andtemperature T, f ðeÞ is the Fermi distribution and D is the half widthof the conduction band. The electric conductance G, the thermo-power S, and the thermal conductance k are given by [7]

G ¼ e2I0ðTÞ; ð2Þ

S ¼ �I1ðTÞ

eTI0ðTÞ; ð3Þ

k ¼ 1

TI2ðTÞ �

I21ðTÞ

I0ðTÞ

� �; ð4Þ

Page 2: Thermoelectric effects in quantum dots

ARTICLE IN PRESS

T=10-9

T=10-53

4t = 0

M. Yoshida, L.N. Oliveira / Physica B 404 (2009) 3312–3315 3313

respectively. Our problem, therefore, is to compute Tðe; TÞ for acorrelated quantum dot coupled to a gas of non-interactingelectrons.

T=10-4

T=10-2

T=10-9

T=10-5

T=10-4

T=10-2

κ / T

[2e2 /

h]

T=10-9

T=10-5

T=10-4

T=10-2

εd

0

1

2

0

1

2

3

0

1

2

3

0

1

2

3

T=10-9

T=10-7

T=10-5

T=10-4

T=10-2

t = 0.16

t = -0.16

t = 0.32

-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1

Fig. 1. Thermal conductance k, normalized by the temperature T, as a function of

the dot energy for U ¼ 0:3D and four t s, at the indicated temperatures. The top

panel, with t ¼ 0, shows no sign of interference. The second (third) panel, with

t ¼ 0:16D (t ¼ �0:16D) displays a Fano antiresonance. In the bottom panel,

t ¼ 0:32D, the conductance vanishes in the Kondo valley as T-0.

3. Thermal conductance and thermopower of a SET

Recent experiments [3] have detected Fano anti-resonances incoexistence with the Kondo effect in SETs. The interferenceindicates that the electrons can flow through the dot or tunneldirectly from on electrode to the other. The transport properties ofthe SET can be studied by a modified Anderson model [9,14],which in standard notation is described by the Hamiltonian

H ¼Xk;a

ekcyk;ack;a þ tXk;k0

ðcyk;Lck;R þ H:c:Þ

þVXk;aðcyk;acd þ H:c:Þ þ Hd: ð5Þ

Here the quantum-dot Hamiltonian is Hd ¼ edcydcd þ Und;mnd;k,with a dot energy ed, controlled by a gate potential applied to thedot, that competes with the Coulomb repulsion U. The summationindex a on the right-hand side takes the values L and R, for the leftand right electrodes respectively. The tunneling amplitude t

allows transitions between the electrodes, while V couples theelectrodes to the quantum dot. The Hamiltonian (5) beinginvariant under inversion, it is convenient to substitute even ðþÞand odd ð�Þ operators ck7 ¼ ðckR7ckLÞ=

ffiffiffi2p

for the ckL and ckR.It results that only the ckþ are coupled to the quantum dot.For brevity, we define the shorthand g � prt, where r is thedensity of conduction states; and the dot-level width G � prV2.

In the absence of magnetic fields, the transmission probabilitythrough the SET is [8,9]

Tðe; TÞ ¼ T0 þ4G

ffiffiffiffiffiffiffiffiffiffiT0R0

p1þ g2

RfGd;dgðe; TÞ

þ2GðT0 � R0Þ

1þ g2IfGd;dgðe; TÞ; ð6Þ

where Gddðe; TÞ is the retarded Green’s function for the dot orbital,and we have defined T0 � 4g2=ð1þ g2Þ

2 and R0 � 1� T0.To compute T, we rely on the numerical-renormalization

group (NRG) diagonalization of the model Hamiltonian [10].Although the resulting eigenvectors and eigenvalues yield essen-tially exact results for IfGddgðe; TÞ, the direct computation ofRfGddg is unwieldy. We have found it more convenient to definethe Fermi operator

b �2G

1þ g2

� �1=2

cd þ4g2

1þ g2

� �1=21ffiffiffiffiffiffiffiprpX

k

ckþ; ð7Þ

because the imaginary part of its retarded Green’s functionGbbðe; TÞ is directly related to the transmission probability: aidedby the two equations of motion relating Gkk0 to Gdk, and Gkd to Gdd,straightforward manipulation of Eq. (6) show that Tðe; TÞ ¼�IfGb;bgðe; TÞ. In practice, we (i) diagonalize H iteratively [10];(ii) for each pair of resulting eigenstates ðjmS; jnS), compute thematrix elements /mjbsjnS; (iii) thermal average the results[15,16] to obtain IfGb;bgðe; TÞ; (iv) substitute the result forTðe; TÞ in Eq. (1); and (v) evaluate the integral for n ¼ 0;1;2 toobtain InðTÞ (n ¼ 0;1;2).

Fig. 1 shows numerical results for the thermal conductance as afunction of the gate energy ed. Well above or well below the Kondotemperature TK , we expect the thermal and electric conductancesto obey the Wiedemann–Franz law k=T ¼ p2G=3 and hence showthe thermal conductance normalized by the temperature T. Eachpanel represents a tunneling parameter t and displays the

thermal-conductance profile for the indicated temperatures. Allcurves were computed for G ¼ 10�2D, and U ¼ 0:3D.

The top panel shows the standard SET, with no direct tunnelingchannel. At the lowest temperature ðkBT ¼ 10�9DÞ, the modelHamiltonian close to the strong-coupling fixed point, the Kondoscreening makes the quantum dot transparent to electrons, so thatin the Kondo regime ½G5minðjedj;2ed þ UÞ� the Wiedemann–Franzlaw pushes the ratio 3k=p2T to the unitary limit 2e2=h. At highertemperatures, the Kondo cloud evaporates and the thermalconductance drops steeply. The maxima near ed ¼ 0 and ed ¼ �U

reflect the two resonances associated with the transitions c1d2c0

d

and c1d2c2

d .In the next two panels, the direct tunneling amplitude

substantially increased, the current through the dot tends tointerfere with the current bypassing the dot. To show that aparticle–hole transformation is equivalent to changing the sign ofthe amplitude t, we compare the curves with t ¼ 0:16D (secondpanel) with t ¼ �0:16D (third panel). At low temperatures, in theformer (latter) case, the interference between the c1

d2c0d and

c1d2c2

d transitions is constructive near ed ¼ 0 (ed ¼ �U) anddestructive near ed ¼ U (ed ¼ 0). At intermediate dot energies,ed � �U=2, the amplitudes for direct transition and for transitionthrough the dot have orthogonal phases and fail to interfere, so thatthe resulting current is the sum of the two individual currents.

In the bottom panel, the direct tunneling amplitude t isdominant. For gate potentials disfavoring the formation of a dot

Page 3: Thermoelectric effects in quantum dots

ARTICLE IN PRESS

M. Yoshida, L.N. Oliveira / Physica B 404 (2009) 3312–33153314

moment, heat flows from one electrode to the other. In the Kondoregime, however, at low temperatures, the Kondo cloud couplingthe dot to the electrode orbitals closest to it blocks transportbetween the electrodes. As the Kondo cloud evaporates, thethermal conductance in the ed � �U=2 rises with temperature, sothat the resulting profile is symmetric to the one in the top panel.The two resonances near ed ¼ 0 and �U, which are independent ofKondo screening, keep the thermal conductance low even atrelatively high temperatures.

Fig. 2 shows thermopower profiles for the same amplitudes t

discussed in Fig. 1. In contrast with the thermal conductance, thethermopower is sensitive to particle–hole asymmetry: heatcurrents due to holes (electrons) make it positive (negative). Forthe standard SET (t ¼ 0, top panel), the thermopower is negligibleat low temperatures and vanishes at the particle–hole symmetricparametrical point ed ¼ �U=2.

With jtj ¼ 0:16D, particle–hole symmetry is broken ated ¼ �U=2, and temperatures comparable to TK make thethermopower sizeable in the Kondo regime. For t ¼ 0:16D,the sensitivity to particle–hole asymmetry makes the interferencebetween electron (hole) currents constructive (destructive) forboth ed ¼ 0 and for ed ¼ U, while for t ¼ �0:16D it is destructive(constructive).

For t ¼ 0:32D, direct tunneling again dominant, in the Kondoregime ðed � �U=2Þ the thermopower becomes sensitive to theKondo effect, which is chiefly due to electrons (holes) above(below) the Fermi level. The thermopower therefore emerges as aprobe of direct-tunneling leaks in SETs, one that may help identifythe source of interference in this and other nanodevices.

T=10-9

T=10-3

T=10

T=10-9

T=10-5

T=10-4

T=10-2

S [k

B/ e

]

T=10-9

T=10-5

T=10-4

T=10-2

εd

-1.4

-0.7

0

0.7

1.4

-1

-0.5

0

0.5

-0.6

0

0.6

1.2

-1.2

-0.6

0

0.6

1.2 T=10-9

T=10-5

T=10-4

T=10-2

t = 0

t = 0.16

t = 0.32

t = - 0.16

-0.4 -0.3 -0.2 -0.1 0 0.1 0.2

Fig. 2. Thermopower as a function of ed for the four tunneling amplitudes t in Fig.

1. Again G ¼ 0:01D and U ¼ 0:3D. For each t, the profile of the thermal power is

presented at the indicated temperatures.

4. Side-coupled quantum dot

We have also studied the T-shaped device, in which the dot isside-coupled to the wire [5]. Again, we considered the Seebecksetup. The quantum wire now shunting the two electrodes,we drop the coupling proportional to t on the right-hand side ofEq. (5) and employ the standard Anderson Hamiltonian

Hs ¼X

k

ekcykck þ VX

k

ðcyk;scd þ H:c:Þ þ Hd; ð8Þ

where Hd ¼ edcydcd þ Und;mnd;k is the dot Hamiltonian. Thetransmission probability is now given by Tðe; TÞ ¼ 1þprV2

IfGd;dgðe; TÞ where Gd;dðe; TÞ. Following the procedure out-lined above, we have diagonalized the Hamiltonian Hs iterativelyand computed the electrical conductance, the thermal conduc-tance and the thermopower as functions of the gate potential ed.Fig. 3 displays results for U ¼ 0:3D, and G ¼ 0:01D.

Not surprisingly—the wire is equivalent to a large tunnelingamplitude, i.e., to t�D—the transport coefficients mimic those ofthe t ¼ 0:32D SET. At low temperatures ðT5TK Þ in the Kondoregime, for instance, the Kondo cloud blocks transport through thewire segment closest to the dot. The thermal and electricalconductances thus vanish for ed � �U=2. As the temperature rises,the evaporation of the Kondo cloud allows transport and bothconductances rise near the particle–hole symmetric point. At lowtemperatures, the sensitivity to particle–hole asymmetry en-hances the thermopower in the Kondo regime, a behavioranalogous to the bottom panel in Fig. 2.

κ / T

[2e2 /h

] T=10-9

T=10-7

T=10-5

T=10-3

T=10-2

S[k B

/e]

T=10-9

T=10-7

T=10-5

T=10-3

T=10-2

εd

0

1

2

3

-1.2

-0.6

0

0.6

1.2

0

0.5

1

G (2

e2 /h) T=10-9

T=10-7

T=10-5

T=10-3

T=10-2

-0.5 -0.4 -0.3 -0.2 -0.1 0 0.1 0.2

Fig. 3. The thermal conductance, thermopower and electric conductance are

presented as a function of ed . The temperature dependence of each quantity is also

presented.

Page 4: Thermoelectric effects in quantum dots

ARTICLE IN PRESS

M. Yoshida, L.N. Oliveira / Physica B 404 (2009) 3312–3315 3315

5. Conclusions

We have calculated the transport coefficients for the SET andthe side-coupled geometries. In both cases, the thermal depen-dence and the gate-voltage profiles show signatures of the Kondoeffect and of quantum interference. Our essentially exact NRGresults identify trends that can aid the interpretation of experi-mental results. In the side-coupled geometry, in particular, theKondo cloud has marked effects upon the thermopower.

Acknowledgments

This work has been funded by BZG. Financial support by theFAPESP and the CNPq is acknowledged.

References

[1] D. Goldhaber-Gordon, et al., Nature 391 (1998) 156.[2] D. Goldhaber-Gordon, et al., Phys. Rev. Lett. 81 (1998) 5225.[3] J. Gores, et al., Phys. Rev. B 62 (2000) 2188.[4] Y.R.D. Mahalu, H. Shtrikman, Phys. Rev. Lett. 88 (2002) 076601.[5] M. Sato, et al., Phys. Rev. Lett. 95 (2005) 066801.[6] R. Scheibner, et al., Phys. Rev. Lett. 95 (2005) 176602.[7] T. Kim, S. Hershfield, Phys. Rev. Lett. 88 (2002) 136601.[8] T. Kim, S. Hershfield, Phys. Rev. B 67 (2003) 165313.[9] W. Hofstetter, et al., Phys. Rev. Lett. 87 (2001) 156803.

[10] K.G. Wilson, Rev. Modern Phys. 47 (1975) 773;H.R. Krishna-murthy, J.W. Wilkins, K.G. Wilson, Phys. Rev. B 21 (1980) 1003.

[11] M. Yoshida, M.A. Whitaker, L.N. Oliveira, Phys. Rev. B 41 (1990) 9403.[12] V.L. Campo Jr., L.N. Oliveira, Phys. Rev. B 72 (2005) 104432.[13] J.M. Ziman, Principles of The Theory of Solids, Cambridge University Press,

Cambridge, 1972.[14] B.R. Bulka, P. Stefa �nski, Phys. Rev. Lett. 86 (2001) 5128.[15] W.C. Oliveira, L.N. Oliveira, Phys. Rev. B 49 (1994) 11986.[16] S.C. Costa, C.A. Paula, V.L. Lıbero, L.N. Oliveira, Phys. Rev. B 55 (1997) 30.