Effects of magnetic field on photon-induced quantum transport in a single dot-cavity system
Rauf Abdullah Nzar1, 2, †, , Fatah Aziz H1, Fatah Jabar M A1
Physics Department, College of Science, University of Sulaimani, Kurdistan Region, Iraq
Science Institute, University of Iceland, Dunhaga 3, IS-107 Reykjavik, Iceland

 

† Corresponding author. E-mail: nzar.r.abdullah@gmail.com

Abstract
Abstract

In this study, we show how a static magnetic field can control photon-induced electron transport through a quantum dot system coupled to a photon cavity. The quantum dot system is connected to two electron reservoirs and exposed to an external perpendicular static magnetic field. The propagation of electrons through the system is thus influenced by the static magnetic and the dynamic photon fields. It is observed that the photon cavity forms photon replica states controlling electron transport in the system. If the photon field has more energy than the cyclotron energy, then the photon field is dominant in the electron transport. Consequently, the electron transport is enhanced due to activation of photon replica states. By contrast, the electron transport is suppressed in the system when the photon energy is smaller than the cyclotron energy.

1. Introduction

Quantum dots (QDs) are a crucial electronic structure in technological devices[13] because of their unique properties, such as zero-dimensional confinement effect[4] and single electron charge effect (Coulomb blockade).[5] Several methods have been used to control electron motion in a QD system, one of which is to control the energy levels and the electron concentration in the QD systems using a plunger gate voltage.[6] On the other hand, applying a photon radiation that interacts with the electrons in the QD induces a fascinating physical phenomena called photon-assisted tunneling (PAT).[6,7] The PAT occurs in quantum systems when a photon radiation is applied to an electronic island connected to electron reservoirs.[8] The photon radiation forms extra channels in the electronic island, leading to a modification in the electron transport.[9] Therefore, the photon radiation can play an essential role in the transport process generating a photo-current that depends on the photon frequency.[10] Recently, the influences of photon field, in a vacuum state, on two-level electronic system,[11] and double quantum dots in the presence of a single mode micro-cavity system with both continuous wave and pulsed excitation are studied.[12] Based on the proposed schemes, a single photon generation can be obtained separately under both QD–cavity resonant and off-resonant conditions. The single photon source, in turn, becomes increasingly important in the very diverse range of technological applications.

In addition, external magnetic fields can be used to control the electron transport in nanodevices, which leads to several important effects including the change of the energy level spacing inside the QD[13] and hence the QD lowest energy state shrinks with increasing magnetic field. As a result, the Coulomb interaction between two spin degenerate electrons grows.[14] Furthermore, external magnetic field can form the edge state[15] and the localized state[16] in the electronic systems, and consequently the electron transport is reduced.

The combination of the aforementioned fields, namely the magnetic and photon fields, can result in a magneto-photon current in graphene[17] and superconductor.[18] In this work, we consider magneto-photo transport under the influence of a quantized single photon mode in a cavity and investigated its effect on electron transport through a QD system. In the presence of the photon cavity, extra channels are formed in the system, which open new windows for electron tunneling called photon-assisted tunneling process. In addition, we have also shown how an external static magnetic field can control photon-assisted tunneling in the QD system.

The rest of this paper is organized as follows. In Section 2, the description of the model and the theoretical formalism are shown. In Section 3, we demonstrate the results. In Section 4, conclusions will be presented.

2. Model and theory

The system under investigation is a two-dimensional (2D) electron gas exposed to an external static magnetic field and a quantized photon field at low temperature. We assume that the electronic system consists of a QD embedded in a quantum wire. The QD system is connected to two electron reservoirs with different chemical potentials. The electron–photon coupling system is described by the following Hamiltonian in the many-body (MB) basis:

where He is the Hamiltonian of the electronic system including electron–electron interaction

Herein, π = p + (e/c) A with p and being the canonical momentum and magnetic vector potential, respectively. The magnetic field is applied along the z axis, i.e., B = B, and and are the fermionic field operators with being the annihilation (creation) operators for an electron in the single electron state |i〉 corresponding to ψi. The QD potential can be described by

where U is the strength of the potential, and αx and αy are constants that determine the diameter of the QD. The plunger-gate voltage is described by Upg which is an electrostatic potential shifting the energy states of the QD system with respect to the chemical potential of the leads. The second term of Eq. (2) indicates the electron–electron interaction in the central system with UC being the Coulomb interaction potential.[19]

The second part of Eq. (1) can be written as Hγ = ħωγaa introducing the Hamiltonian of the free photon field with ħωγ being the photon energy, and a (a) the photon annihilation (creation) operators. The quantized vector potential of the cavity photon field, in the Coulomb gauge, is given by Âγ = A(â + â)e, where A is the amplitude of the photon field, related to the electron–photon coupling constant via gγ = eAawΩw/c, and e determines the photon polarization with either parallel e = ex or perpendicular e = ey to the electron motion. Note that aw is the effective magnetic length and Ωw is the effective confinement frequency of electrons of the QD system.

The last term on the right side of Eq. (1)

represents the full electron–photon interaction including both para- and dia-magnetic electron–photon interactions, respectively. The charge is ρ = −ψ and the charge current density is governed by

The electron–electron and the electron–photon interactions are treated by exact diagonalization in appropriately truncated Fock spaces.

Figure 1 shows the schematic diagram of the QD system (brown color) connected to two leads (black color) under the combined effects of the magnetic field B (red arrows) and the photon radiation (blue zigzag arrows). The chemical potential of the left lead μL is assumed to be higher than that of the right lead μR. Consequently, the transport is dominated by the left to right electron motions between the two leads through the central system as indicated by violet arrows.

Fig. 1. Schematic diagram of a QD system (brown color) connected to the left lead (black color) with chemical potential μL and and the right lead with chemical potential μR. The photon field is represented by the blue zigzag arrow. The external magnetic field B is labeled by the red arrow.

The Liouville–von Neumann equation is used to describe the time evolution of the many-body density operator of the closed system. However, in the case of open system when the central system is connected to the leads, we use a projection operator technique to derive a generalized master equation for the reduced density operator.[20,21] Since we are interested in the transient behavior of the system, we assume a non-Markovian approach valid to a weak coupling of the leads to the central system.[16]

Once we have the reduced density operator, one can calculate charge current and charge density in the system. The charge current is Ic(t) = IL(t) − IR(t), where IL(t) indicates the partial current from the left lead into the QD system, and −IR(t) refers to the partial current into the right lead from the QD system. The partial current can be introduced as where and are the time derivatives of the system’s reduced density matrix due to its coupling to the left and right leads, respectively,[19,22] and is the charge operator with the number operator

3. Results and discussions

We assume the QD system and the leads are made of GaAs semiconductor with effective electron mass m* = 0.067me and relative dielectric constant κ = 12.4. The parameters of the QD potential are U = −3.3 meV, and αx = αy = 0.03 nm−1. The cavity consists of a single photon mode with energy ħωγ = 0.3 meV, and the electron–photon coupling strength gγ = 0.1 meV. The chemical potential of the left and the right leads are μL = 1.2 meV and μR = 1.1 meV, respectively, implying the bias voltage Δμ = μLμR = 0.1 meV. The temperature of the leads before coupling to the QD system is T = 0.001 K. The confinement energy of electrons in the QD system is equal to that of the leads ħΩ0 = ħΩl = 2.0 meV. Finally, the photon field is linearly polarized and aligned with the x axis parallel to the direction of electron motion in the QD system.

In what follows, we explain the influences of the magnetic field on photon-induced electron transport through the QD system. Figure 2 shows the energy spectrum of the QD system versus the plunger-gate voltage including zero-electron states (0ES, golden diamonds), one-electron states (1ES, blue rectangles), and two-electron states (2ES, red circles). The chemical potential of the leads are indicated by two horizontal lines (black lines).

Fig. 2. The energy spectrum of the QD system without (a) and with (b) photon cavity versus the plunger-gate voltage Upg including zero-electron states (0ES, golden diamonds), one-electron states (1ES, blue rectangles), and two-electron states (2ES, red dots). The chemical potentials are μL = 1.2 meV and μR = 1.1 meV (black). The SE state in the bias window is almost doubly degenerate due to the small Zeeman energy.

In Fig. 2(a) the many-electron (ME) energy of the QD system, excluding the photon cavity, is demonstrated. For the selected range of the plunger-gate voltage, the first excited state lies between the two chemical potentials, inside the bias window, which in turn gets into resonance with the first sub-band energy of the leads located in the bias window. Therefore, an electron in the first sub-band of the left lead may perform electron tunneling into the first-excited state of the QD system. As a result, a peak in the charge current is formed at as shown in Fig. 3. In addition, it should be known that the ground state energy of the QD system is found below 0.8 meV (not shown).

Fig. 3. The charge current IQ is plotted as a function of plunger gate voltage Upg at time t = 220 ps for (a) the QD system without photon cavity for cyclotron energy ħωc ≃ 10−4 meV (blue solid), 0.34 meV (green dashed), and 0.86 meV (red dotted), and (b) the QD system with photon cavity for ħωγ > ħωc (blue solid), ħωγħωc (green dashed), and ħωγ < ħωc (red dotted), where the photon energy is ħωγ = 0.3 meV. The bias window is Δμ = 0.1 meV, gγ = 0.1 meV, and Nγ = 2.

In Fig. 3(a) the charge current versus the plunger-gate voltage Upg is plotted for three different values of the cyclotron energy ħωc ≃ 10−4 meV (blue solid), 0.34 meV (green dashed), and 0.86 meV (red dotted), corresponding to the magnetic field B = 0.0001 T, 0.2 T, and 0.5 T, respectively. We can clearly see that the peak current increases with the cyclotron energy. At ħωc ≃ 10−4 meV, the charge current is very weak. This can be attributed to the localization of charge density in the QD (see Fig. 4(a)). In contrast, for higher value of cyclotron energy (such as ħωc ≃ 0.86 meV) corresponding to the higher magnetic field, the charge is delocalized and slightly extended to the outside of the QD which can be understood as follows. The high magnetic field induces stronger Lorentz force that forms a circular motion of the electron charge density outside the dot (not shown), and consequently the charge current is enhanced (see the red dotted line in Fig. 3(a)).

Fig. 4. Charge density in the QD system at t = 220 ps without (a) and with (b) photon cavity in the main current peak at shown in Fig. 3 for the cyclotron energy ħωc = 10−4 meV. The effective magnetic length is aw = 33.72 nm, ħωγ = 0.3 meV, gγ = 0.1 meV, and Nγ = 2.

Now assuming the QD system is coupled to the cavity photon, the electron transport can be affected by both the static magnetic and the dynamic photon fields. For a photon energy ħωγ = 0.3 meV and the electron–photon coupling strength gγ = 0.1 meV, figure 2(b) demonstrates how the many-body (MB) energy varies as a function of the plunger-gate voltage. In the presence of the cavity, photon replica states are formed with different photon contents. The energy spacing between the photon replica states is appropriately equal to the photon energy at low electron–photon coupling strength. Therefore, the first excited state in the bias window is no longer active in the electron transport but instead the electrons from the left leads transfer to the photon replica of the first excited state in the QD system. Comparing to the energy spectrum of the QD system in Fig. 2(a) for which the photon field is neglected, in Fig. 2(b) two photon replica states at Upg = 0.1 and 0.7 mV (blue rectangles) are found in the bias window corresponding to and respectively.

Figure 3(b) shows the charge current as a function of the plunger-gate voltage in the presence of the photon cavity for three cases ħωγ > ħωc (blue solid), ħωγħωc (green dashed), and ħωγ < ħωc (red dotted), where the photon energy is ħωγ = 0.3 meV. The peak current (main peak) at is again found. In addition to the main peak, an extra side peak is observed at The existence of this side peak is due to the formation of the one photon replica of the first excited state.

In the case of ħωγ > ħωc, where ħωγ = 0.3 meV and ħωc ≃ 10−4 meV, the photon field is dominant. Comparing to the charge current in the absence of the cavity shown in Fig. 3(a) (blue solid), the current is increased in the main peak at which attributes to the fact that the charge density is stretched out of the QD, as shown in Fig. 4(b). This stretching effect is caused by the paramagnetic term of the electron–photon interaction. In addition, the contribution of the photon replica state with two photons can also enhance the charge current. It is worth mentioning that this happens because the energy of two photon replica state is higher than that of the first excited state in the energy spectrum. The higher states in the energy spectrum are less bound in the system and actively contribute to the electron transport.

We should also note that the current is almost unchanged when ħωγħωc (green dashed) with ħωc ≃ 0.34 meV at the main peak recapturing the same value of current as was found in the absence photon cavity.

In addition, when ħωγ < ħωc (red dotted line in Fig. 3(b)) with ħωc ≃ 0.86 meV, the magnetic field effect is dominant. At this high cyclotron energy, the energy spacing between photon replica states is increased and the photon replica states weakly contribute to the electron transport. As a result, the charge current is decreased in the main peak.

Another interesting aspect of this issue is the influence of the external magnetic field on the current in the side peak displayed in Fig. 3(b). The formation of side peak is totally due to the photon cavity. It can be clearly seen that the current is high at low cyclotron energy when ħωγ > ħωc (blue solid), indicating that the photon-induced current should be generated at low magnetic field. There are two reasons for the high current here: the photon replica states are more active in the electron transport at low cyclotron energy, and the stretching of charge density in the QD system due to the photon cavity. To explain the enhancement of current at the side peak, the charge density at Upg = 0.1 mV is shown in Fig. 5 for the low cyclotron energy (ħωc ≃ 10−4 meV), i.e., ħωγ > ħωc, and the high cyclotron energy (ħωc ≃ 0.86 meV), i.e., ħωγ < ħωc. In Fig. 5(a) the charge density is mostly distributed outside the QD and near the contact area to the leads. The charge accumulation in the contact area leads to the stronger charging to the QD system from the leads, and then the charge current at the side peak is increased. Therefore, we emphasize that the PAT process requires the following condition ħωγ > ħωc.

Fig. 5. Charge density in the QD system at t = 220 ps with photon cavity in the side current peak at Upg = 0.1 mV shown in Fig. 3(b) for the case of ħωγ > ħωc (a) and ħωγ < ħωc (b). The effective magnetic length is aw = 33.72 nm, ħωγ = 0.3 meV, gγ = 0.1 meV, and Nγ = 2.

By contrast, at high cyclotron energy (ħωc ≃ 0.86 meV) when ħωγ < ħωc, the magnetic field dominates the electron transport. The magnetic field causes the charge accumulation around the QD, as shown in Fig. 5(b), and then the current is reduced at the side peak.

4. Conclusions

We have investigated the influence of a static magnetic field on photon-induced transport through a QD coupled to a quantized photon cavity. It was found that the cavity forms photon replica states and their contribution to the electron transport can be affected by the external magnetic field. Therefore, two different regimes are studied, low and high magnetic fields. At the low magnetic field regime (low cyclotron energy), where the cyclotron energy is assumed to be lower than the photon energy, the photon replica states formed in the presence of the cavity actively contribute to the transport. Consequently, the charge density in the system are stretched and then the current is increased.

On the other hand, at the high magnetic field regime (high cyclotron energy), when the cyclotron energy is higher than the photon energy, the magnetic field is dominant and the photon-induced current is suppressed. As a result, we emphasis that the photon-induced transport or photon-assisted transport can be obtained when the photon energy is higher than the cyclotron energy in the system.

Reference
1ImamogluAYamamotoY 1994 Phys. Rev. Lett. 72 210
2LossDDiVincenzoD P 1998 Phys. Rev. 57 120
3DiVincenzoD P 2005 Science 309 2173
4PetroffP MSchmidtK HRibeiroG MLorkeAKotthausJ 1997 Jpn J. Appl. Phys. 36 4068
5KouwenhovenL PMcEuenP LSingle Electron Transport Through a Quantum Dot in NanotechnologyTimpGregoryNew YorkSpringer471535471–535
6FujisawaTvan der WielW GKouwenhovenL P 2000 Physica 7 413
7KouwenhovenL PJauharSMcCormickKDixonDMcEuenP LNazarov YuVvan der VaartN CFoxonC T1994Phys. Rev. B502019
8ShibataKUmenoAChaK MHirakawaK 2012 Phys. Rev. Lett. 109 077401
9IshibashiKAoyagiY2002Physica B314437
10KouwenhovenL PJauharSOrensteinJMcEuenP LNagamuneYMotohisaJSakakiH 1994 Phys. Rev. Lett. 73 3443
11GuoY JNieW J 2015 Chin. Phys. 24 094205
12YeHPengY WYuZ YZhangWLiuY M 2015 Chin. Phys. 24 114202
13MaksymP AChakrabortyT 1990 Phys. Rev. Lett. 65 108
14van der WielW GDe FranceschiSElzermanJ MFujisawaTTaruchaSKouwenhovenL P 2002 Rev. Mod. Phys. 75 1
15ThomasI2010Semiconductor NanostructuresNew YorkOxford University Press
16AbdullahN RTangC SGudmundssonV 2010 Phys. Rev. 82 195325
17HagenmüllerDCiutiCPhys. Rev. Lett.109267403
18MaissenCScalariGValmorraFBeckMFaistJCibellaSLeoniRReichlCCharpentierCWegscheiderW 2014 Phys. Rev. 90 205309
19AbdullahN R2015Cavity-photon Controlled Electron Transport through Quantum Dots and Waveguide Systems PhD ThesisUniversity of IcelandReykjavik, Iceland
20NakajimaS 1958 Prog. Theor. Phys. 20 948
21ZwanzigR 1960 J. Chem. Phys. 33 1338
22AbdullahN RTangC SManolescuAGudmundssonV 2016 ACS Photonics 3 249