License: CC BY 4.0
arXiv:2402.15499v1 [cond-mat.mes-hall] 23 Feb 2024
11institutetext: Zach D. Merino 22institutetext: University of Waterloo, ON, N2L 3G1, 22email: [email protected] 33institutetext: Bohdan Khromets 44institutetext: University of Waterloo, ON, N2L 3G1, 44email: [email protected] 55institutetext: Jonathan Baugh 66institutetext: Institute for Quantum Computing 77institutetext: University of Waterloo, ON, N2L 3G1 77email: [email protected]

Simulated Charge Stability in a MOSFET Linear Quantum Dot Array

Zach D. Merino    Bohdan Khromets    and Jonathan Baugh
Abstract

In this study, we address challenges in designing quantum information processors based on electron spin qubits in electrostatically-defined quantum dots (QDs). Numerical calculations of charge stability diagrams are presented for a realistic double QD device geometry. These methods generalize to linear QD arrays, and are based on determining the effective parameters of a Hubbard model Hamiltonian that is then diagonalized to find the many-electron ground state energy. These calculations enable the identification of gate voltage ranges that maintain desired charge states during qubit manipulation, and also account for electrical cross-talk between QDs. As a result, the methods presented here promise to be a valuable tool for develo** scalable spin qubit quantum processors.

1 Introduction

In recent years, certain Quantum Computing (QC) implementations have demonstrated a quantum advantage over classical processors Zhong2020 . While further demonstrations of quantum supremacy may be realized, quantum processors (QPs) consisting of thousands to millions of physical qubits will be required to perform large-scale fault-tolerant QC.

The scaling requirements for fault tolerant QC make electrostatically-defined quantum dot (QD) spin qubits a promising approach. In particular, spin qubits in silicon (Si) have attracted much recent interest due to their compatibility with industrial complementary metal-oxide semiconductor (CMOS) fabrication techniques, which could yield high quality, densely packed qubits Saraiva2021 ; Undseth2023 . A schematic Si double QD device is shown in figure 1. A 2-dimensional electron gas accumulates at the Si/SiO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT interface due to positive voltage applied to a gate electrode. Mesoscopic QDs are formed beneath the plunger gate fingers, localized by the presence of screening gates. The QDs are tuned into the few-electron regime via the plunger gates, where tunnel coupling to a reservoir (not shown) allows for charge transfer. The spin of an unpaired electron in a QD with an odd number of charges constitutes the qubit. Spin qubits in Si and SiGe devices have exhibited long coherence compared to quantum gate operation time scales Veldhorst2014 , high-fidelity gate operations required for fault-tolerant QC Mills2022 , qubit control via applied electric gate signals Undseth2023 , and qubit operation at relatively high temperatures (14141-41 - 4 K) Petit2020 . Furthermore, scalable spin qubit architectures suitable for implementing surface code quantum error correction have been proposed Veldhorst2017 ; Buonacorsi2019 .

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 1: A schematic MOSFET double QD device in Si: (0(a)) Side cross-sectional view showing a 2-dimensional electron gas (2DEG) formed by a plunger gate electrode. (0(b)) Top-view cross-section showing the double well confining potential formed by the plunger (V1,V2subscript𝑉1subscript𝑉2V_{1},V_{2}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) and tunnel (W1subscript𝑊1W_{1}italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) electrodes. The electrostatic potential energy at the SiO2/Si interface is calculated by a finite element Poisson solver based on a 3D device model. (0(c)) Contour plot of the potential energy landscape in (b).

Scaling QPs generally increases the complexity in initializing, controlling, and measuring quantum computational states. Each additional QD adds independent control parameters (plunger and tunnel gates) and is a source of cross-talk between neighboring QDs Heinz2021 ; Undseth2023 . A critical requirement is the ability to predict the charge state of a linear QD array for a given vector of gate voltages. For spin qubits in singly-occupied QDs, it is necessary to determine the control boundaries within which single charge occupations are maintained. Thus, accurately modeling the charge state of a realistic linear QD array as a function of gate voltages is a necessary first step, prior to the design of spin qubit control pulses.

Previous studies have employed Hubbard models as approximate descriptions of linear arrays of tunnel-coupled QDs Hensgens2017 ; Secchi2023 . Given certain assumptions about the underlying electrostatic potential, others have shown that effective parameters of the Hamiltonian can be calculated DasSarma2011 ; Yang2011 and a resulting charge configuration determined Jafari2008 . In this work, electrostatic potential landscapes for realistic QD device geometries found by a finite element Poisson solver Birner2007 are used as inputs for calculations that determine stable charge states as a function of gate electrode voltages.

2 Hubbard Model for Quantum Dots

2.1 Hamiltonian

A generalized Hubbard Hamiltonian developed by DasSarma2011 ; Yang2011 for electrostatically defined QD systems is composed of

H=Hμ+Ht+HU+HJ,𝐻subscript𝐻𝜇subscript𝐻𝑡subscript𝐻𝑈subscript𝐻𝐽H=H_{\mu}+H_{t}+H_{U}+H_{J},italic_H = italic_H start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT + italic_H start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT , (1)

where terms describing the chemical potential, Hμsubscript𝐻𝜇H_{\mu}italic_H start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, tunnel coupling, Htsubscript𝐻𝑡H_{t}italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, Coulomb repulsion, HUsubscript𝐻𝑈H_{U}italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT, and spin exchange, HJsubscript𝐻𝐽H_{J}italic_H start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, are included. The first term, Hμsubscript𝐻𝜇H_{\mu}italic_H start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, represents the occupation of spin states in an array of N𝑁Nitalic_N quantum dots:

Hμ=σ=,[i=1Nμin^i,σ]subscript𝐻𝜇subscript𝜎delimited-[]superscriptsubscript𝑖1𝑁subscript𝜇𝑖subscript^𝑛𝑖𝜎H_{\mu}=-\sum_{\sigma=\uparrow,\downarrow}\left[\sum_{i=1}^{N}\mu_{i}\hat{n}_{% i,\sigma}\right]italic_H start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_σ = ↑ , ↓ end_POSTSUBSCRIPT [ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT ] (2)

Here, i𝑖iitalic_i is the index of each QD in a linear array of N𝑁Nitalic_N QDs, μisubscript𝜇𝑖\mu_{i}italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are their chemical potentials, σ{,}𝜎\sigma\in\{\uparrow,\downarrow\}italic_σ ∈ { ↑ , ↓ } is the spin state, and n^i,σsubscript^𝑛𝑖𝜎\hat{n}_{i,\sigma}over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT is the electron number operator for the spin state σ𝜎\sigmaitalic_σ of the i𝑖iitalic_ith QD. The number operator is defined as n^i,σ=c^i,σc^i,σsubscript^𝑛𝑖𝜎superscriptsubscript^𝑐𝑖𝜎subscript^𝑐𝑖𝜎\hat{n}_{i,\sigma}=\hat{c}_{i,\sigma}^{\dagger}\hat{c}_{i,\sigma}over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT = over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT, where c^i,σsuperscriptsubscript^𝑐𝑖𝜎\hat{c}_{i,\sigma}^{\dagger}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT (c^i,σ)subscript^𝑐𝑖𝜎(\hat{c}_{i,\sigma})( over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_i , italic_σ end_POSTSUBSCRIPT ) are the usual creation (annihilation) operators.

The second term, Htsubscript𝐻𝑡H_{t}italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, represents the tunnel couplings between QDs where an electron with spin state σ𝜎\sigmaitalic_σ is allowed to tunnel to an adjacent QD.

Ht=σ=,[j=1N1tj,j+1(c^j,σc^j+1,σ+c^j+1,σc^j,σ)]subscript𝐻𝑡subscript𝜎delimited-[]superscriptsubscript𝑗1𝑁1subscript𝑡𝑗𝑗1superscriptsubscript^𝑐𝑗𝜎subscript^𝑐𝑗1𝜎superscriptsubscript^𝑐𝑗1𝜎subscript^𝑐𝑗𝜎H_{t}=-\sum_{\sigma=\uparrow,\downarrow}\left[\sum_{j=1}^{N-1}t_{j,j+1}\left(% \hat{c}_{j,\sigma}^{\dagger}\hat{c}_{j+1,\sigma}+\hat{c}_{j+1,\sigma}^{\dagger% }\hat{c}_{j,\sigma}\right)\right]italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - ∑ start_POSTSUBSCRIPT italic_σ = ↑ , ↓ end_POSTSUBSCRIPT [ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT ( over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j + 1 , italic_σ end_POSTSUBSCRIPT + over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j + 1 , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT ) ] (3)

Here, tj,j+1subscript𝑡𝑗𝑗1t_{j,j+1}italic_t start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT are the tunnel coupling, or inter-site hop** terms for adjacent QDs. The operators c^j,σc^j+1,σsuperscriptsubscript^𝑐𝑗𝜎subscript^𝑐𝑗1𝜎\hat{c}_{j,\sigma}^{\dagger}\hat{c}_{j+1,\sigma}over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j + 1 , italic_σ end_POSTSUBSCRIPT, (c^j+1,σc^j,σ)superscriptsubscript^𝑐𝑗1𝜎subscript^𝑐𝑗𝜎\left(\hat{c}_{j+1,\sigma}^{\dagger}\hat{c}_{j,\sigma}\right)( over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j + 1 , italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT over^ start_ARG italic_c end_ARG start_POSTSUBSCRIPT italic_j , italic_σ end_POSTSUBSCRIPT ), account for j(j+1)𝑗𝑗1j\rightarrow(j+1)italic_j → ( italic_j + 1 ), (j(j+1))𝑗𝑗1\left(j\leftarrow(j+1)\right)( italic_j ← ( italic_j + 1 ) ), tunneling events.

The third term, HUsubscript𝐻𝑈H_{U}italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT, is the Coulomb repulsion between all electron pairs.

HU=i=1NUin^i,n^i,+σ1,σ2{,}[j=1N1Uj,j+1n^j,σ1n^j+1,σ2]subscript𝐻𝑈superscriptsubscript𝑖1𝑁subscript𝑈𝑖subscript^𝑛𝑖subscript^𝑛𝑖subscriptsubscript𝜎1subscript𝜎2delimited-[]superscriptsubscript𝑗1𝑁1subscript𝑈𝑗𝑗1subscript^𝑛𝑗subscript𝜎1subscript^𝑛𝑗1subscript𝜎2H_{U}=\sum_{i=1}^{N}U_{i}\hat{n}_{i,\uparrow}\hat{n}_{i,\downarrow}+\sum_{% \sigma_{1},\sigma_{2}\in\{\uparrow,\downarrow\}}\left[\sum_{j=1}^{N-1}U_{j,j+1% }\hat{n}_{j,\sigma_{1}}\hat{n}_{j+1,\sigma_{2}}\right]italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i , ↑ end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i , ↓ end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∈ { ↑ , ↓ } end_POSTSUBSCRIPT [ ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j , italic_σ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT over^ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_j + 1 , italic_σ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ]

The intra-Coulomb, or onsite, energy, Uisubscript𝑈𝑖U_{i}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, accounts for the Coulomb repulsion energy of electrons on the same QD, while the inter-Coulomb energy, Uj,j+1subscript𝑈𝑗𝑗1U_{j,j+1}italic_U start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT, accounts for the Coulomb energy between pairs of electrons in adjacent QDs. In general, the inter-Coulomb energy accounts for all combinations of electron pairs in different QDs, however, in this work only nearest neighbor contributions are considered.

The last term, HJsubscript𝐻𝐽H_{J}italic_H start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT, accounts for energy contributions from spin-exchange, pair-hop**, and occupation-modulated hop**. However, this term is much smaller compared to the other three components and will be neglected for the purposes of this work.

2.2 Effective Parameters

The inter-Coulomb repulsion is calculated using the expression given in Yang2011 ,

Uij=𝑑𝒓𝟏𝑑𝒓𝟐|φi(𝒓𝟏)|2|φj(𝒓𝟐)|2v(𝒓𝟏,𝒓𝟐)subscript𝑈𝑖𝑗superscriptsubscriptdifferential-dsubscript𝒓1superscriptsubscriptdifferential-dsubscript𝒓2superscriptsubscript𝜑𝑖subscript𝒓12superscriptsubscript𝜑𝑗subscript𝒓22𝑣subscript𝒓1subscript𝒓2U_{ij}=\int_{-\infty}^{\infty}d\boldsymbol{r_{1}}\int_{-\infty}^{\infty}d% \boldsymbol{r_{2}}\left|\varphi_{i}(\boldsymbol{r_{1}})\right|^{2}\left|% \varphi_{j}(\boldsymbol{r_{2}})\right|^{2}v(\boldsymbol{r_{1}},\boldsymbol{r_{% 2}})italic_U start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT | italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_v ( bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT , bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) (4)

where φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the single-particle ground state wave function of the i𝑖iitalic_ith QD and,

v(𝒓𝟏,𝒓𝟐)e24πϵ0ϵr1|𝒓𝟏𝒓𝟐|𝑣subscript𝒓1subscript𝒓2superscript𝑒24𝜋subscriptitalic-ϵ0subscriptitalic-ϵ𝑟1subscript𝒓1subscript𝒓2v(\boldsymbol{r_{1}},\boldsymbol{r_{2}})\equiv\frac{e^{2}}{4\pi\epsilon_{0}% \epsilon_{r}}\frac{1}{\left|\boldsymbol{r_{1}}-\boldsymbol{r_{2}}\right|}italic_v ( bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT , bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) ≡ divide start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG | bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT - bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT | end_ARG

is the Coulomb repulsion term. The onsite repulsion term for the i𝑖iitalic_ith QD is defined likewise but contains only the single-particle wave function φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the QD in consideration:

Ui=Uiisubscript𝑈𝑖subscript𝑈𝑖𝑖U_{i}=U_{ii}italic_U start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_U start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT (5)

The wave functions φisubscript𝜑𝑖\varphi_{i}italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT are determined by solving the Schrödinger equation

[𝒑^i22m*+Vi(𝒓)]φi(𝒓)=εiφi(𝒓),delimited-[]superscriptsubscriptbold-^𝒑𝑖22superscript𝑚subscript𝑉𝑖𝒓subscript𝜑𝑖𝒓subscript𝜀𝑖subscript𝜑𝑖𝒓\displaystyle\left[\frac{\boldsymbol{\hat{p}}_{i}^{2}}{2m^{*}}+V_{i}\left(% \boldsymbol{r}\right)\right]\varphi_{i}\left(\boldsymbol{r}\right)=\varepsilon% _{i}\varphi_{i}\left(\boldsymbol{r}\right),[ divide start_ARG overbold_^ start_ARG bold_italic_p end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT end_ARG + italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) ] italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) = italic_ε start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) , (6)

where m*superscript𝑚m^{*}italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT is the electron effective mass, for all electrostatic potentials Vi(𝒓)subscript𝑉𝑖𝒓V_{i}\left(\boldsymbol{r}\right)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) representing localized individual QDs. An example of the procedure that allows us to determine Vi(𝒓)subscript𝑉𝑖𝒓V_{i}\left(\boldsymbol{r}\right)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) for each QD in a double-dot potential is shown in figure 2. Here, each of the regions near the local minima (describing individual QDs) is fitted to a 2D Gaussian function while the other ones are masked with a high potential value. The procedure preserves the effects of gate voltage cross-talk on QDs even after the isolation of their electrostatic potentials.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: Electrostatic potential, ΦΦ\Phiroman_Φ, for: (1(a)) left single QD and masked right QD (1(b)) right single QD and masked left QD. Regions are in circled where there is a 99%absentpercent99\approx 99\%≈ 99 %, 95%absentpercent95\approx 95\%≈ 95 %, and 65%absentpercent65\approx 65\%≈ 65 % probability of finding an electron.

While the equations (4) and (5) are well defined, the integral expressions appear to be singular due to the Coulomb potential term v(𝒓𝟏,𝒓𝟐)𝑣subscript𝒓1subscript𝒓2v(\boldsymbol{r_{1}},\boldsymbol{r_{2}})italic_v ( bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT , bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) when 𝒓𝟏=𝒓𝟐subscript𝒓1subscript𝒓2\boldsymbol{r_{1}}=\boldsymbol{r_{2}}bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT = bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT. However, the apparent singularity can be avoided by performing the change of variable

𝒓=𝒓𝟏𝒓𝟐𝒓subscript𝒓1subscript𝒓2\displaystyle\boldsymbol{r}=\boldsymbol{r_{1}}-\boldsymbol{r_{2}}bold_italic_r = bold_italic_r start_POSTSUBSCRIPT bold_1 end_POSTSUBSCRIPT - bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT

and integrating 𝒓𝒓\boldsymbol{r}bold_italic_r in polar coordinates. This yields the following integral that can be performed numerically:

Uij=e24πϵ0ϵr0d|𝒓|02π𝑑θ𝑑𝒓𝟐|φi(𝒓+𝒓𝟐)|2|φj(𝒓𝟐)|2subscript𝑈𝑖𝑗superscript𝑒24𝜋subscriptitalic-ϵ0subscriptitalic-ϵ𝑟superscriptsubscript0𝑑𝒓superscriptsubscript02𝜋differential-d𝜃superscriptsubscriptdifferential-dsubscript𝒓2superscriptsubscript𝜑𝑖𝒓subscript𝒓22superscriptsubscript𝜑𝑗subscript𝒓22\displaystyle U_{ij}=\frac{e^{2}}{4\pi\epsilon_{0}\epsilon_{r}}\int_{0}^{% \infty}d\left|\boldsymbol{r}\right|\int_{0}^{2\pi}d\theta\int_{-\infty}^{% \infty}d\boldsymbol{r_{2}}\left|\varphi_{i}(\boldsymbol{r}+\boldsymbol{r_{2}})% \right|^{2}\left|\varphi_{j}(\boldsymbol{r_{2}})\right|^{2}italic_U start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG italic_e start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 italic_π italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ϵ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d | bold_italic_r | ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT italic_d italic_θ ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_d bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT | italic_φ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r + bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | italic_φ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_italic_r start_POSTSUBSCRIPT bold_2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (7)

Next, to determine the chemical potential from equation (2), Vi(𝒓)subscript𝑉𝑖𝒓V_{i}\left(\boldsymbol{r}\right)italic_V start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_italic_r ) is defined by masking all but the i𝑖iitalic_ith QD’s electrostatic potential, and the single-electron ground state energy is found by solving equation (6).

μi=εi(g)subscript𝜇𝑖subscriptsuperscript𝜀𝑔𝑖\displaystyle\mu_{i}=\varepsilon^{(g)}_{i}italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_ε start_POSTSUPERSCRIPT ( italic_g ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT

Lastly, the tunnel coupling term from equation (3) is derived from a two-level Hamiltonian that includes interdot tunneling in a double QD, and by masking the electrostatic potential for all but the j𝑗jitalic_jth QD pair,

tj,j+1=12(E(e)j,j+1E(g)j,j+1)2+(ε(g)jε(g)j+1)2t_{j,j+1}=\frac{1}{2}\sqrt{\Bigr{(}E^{(e)}_{j,j+1}-E^{(g)}_{j,j+1}\Bigl{)}^{2}% +\Bigr{(}\varepsilon^{(g)}_{j}-\varepsilon^{(g)}_{j+1}\Bigl{)}^{2}}italic_t start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG square-root start_ARG ( italic_E start_POSTSUPERSCRIPT ( italic_e ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT - italic_E start_POSTSUPERSCRIPT ( italic_g ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_ε start_POSTSUPERSCRIPT ( italic_g ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_ε start_POSTSUPERSCRIPT ( italic_g ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (8)

where Ej,j+1e/gsubscriptsuperscript𝐸𝑒𝑔𝑗𝑗1E^{e/g}_{j,j+1}italic_E start_POSTSUPERSCRIPT italic_e / italic_g end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_j + 1 end_POSTSUBSCRIPT are the first excited/ground state energies for the double QD potential of the j𝑗jitalic_jth QD pair and εj/j+1(g)subscriptsuperscript𝜀𝑔𝑗𝑗1\varepsilon^{(g)}_{j/j+1}italic_ε start_POSTSUPERSCRIPT ( italic_g ) end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j / italic_j + 1 end_POSTSUBSCRIPT are the ground state energies corresponding to the single QD potentials.

2.3 Charge State via Exact Diagonalization

To determine the charge state of a linear QD array at a given voltage configuration, the components of the Hamiltonian defined by equations (2)-(2.1) are calculated for a chosen charge configuration basis. The Hubbard model (1) accounts for all charge and spin states with up to two electrons per QD following the Pauli exclusion principle. For example, a choice of basis for a double QD with total electron number 2 is the following:

|α1c1,c2,|0,0=|,|α2c2,c2,|0,0=|0,|α3c1,c2,|0,0=|,|α4c2,c1,|0,0=|,|α5c1,c1,|0,0=|,0|α6c1,c2,|0,0=|,\begin{aligned} \left|\alpha_{1}\right\rangle&\equiv c_{1,\uparrow}^{\dagger}c% _{2,\uparrow}^{\dagger}\left|0,0\right\rangle=\left|\uparrow,\uparrow\right% \rangle\\ \left|\alpha_{2}\right\rangle&\equiv c_{2,\downarrow}^{\dagger}c_{2,\uparrow}^% {\dagger}\left|0,0\right\rangle=\left|0,\uparrow\downarrow\right\rangle\\ \left|\alpha_{3}\right\rangle&\equiv c_{1,\downarrow}^{\dagger}c_{2,\uparrow}^% {\dagger}\left|0,0\right\rangle=\left|\downarrow,\uparrow\right\rangle\end{% aligned}\qquad\begin{aligned} \left|\alpha_{4}\right\rangle&\equiv c_{2,% \downarrow}^{\dagger}c_{1,\uparrow}^{\dagger}\left|0,0\right\rangle=\left|% \uparrow,\downarrow\right\rangle\\ \left|\alpha_{5}\right\rangle&\equiv c_{1,\downarrow}^{\dagger}c_{1,\uparrow}^% {\dagger}\left|0,0\right\rangle=\left|\uparrow\downarrow,0\right\rangle\\ \left|\alpha_{6}\right\rangle&\equiv c_{1,\downarrow}^{\dagger}c_{2,\downarrow% }^{\dagger}\left|0,0\right\rangle=\left|\downarrow,\downarrow\right\rangle\end% {aligned}start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 1 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 2 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | ↑ , ↑ ⟩ end_CELL end_ROW start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 2 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 2 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | 0 , ↑ ↓ ⟩ end_CELL end_ROW start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 1 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 2 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | ↓ , ↑ ⟩ end_CELL end_ROW start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 2 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 1 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | ↑ , ↓ ⟩ end_CELL end_ROW start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 1 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 1 , ↑ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | ↑ ↓ , 0 ⟩ end_CELL end_ROW start_ROW start_CELL | italic_α start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT ⟩ end_CELL start_CELL ≡ italic_c start_POSTSUBSCRIPT 1 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT 2 , ↓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT | 0 , 0 ⟩ = | ↓ , ↓ ⟩ end_CELL end_ROW

The ordering of the basis is arbitrary, however, the above order was chosen for consistency with Jafari2008 so that the Hamiltonian has only diagonal contributions from Hμsubscript𝐻𝜇H_{\mu}italic_H start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT and HUsubscript𝐻𝑈H_{U}italic_H start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT and non-zero off diagonal element contributions from Htsubscript𝐻𝑡H_{t}italic_H start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. In general, the elements of the Hamiltonian are given by Hnm=αn|H|αmsubscript𝐻𝑛𝑚quantum-operator-productsubscript𝛼𝑛𝐻subscript𝛼𝑚H_{nm}=\left\langle\alpha_{n}\right|H\left|\alpha_{m}\right\rangleitalic_H start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT = ⟨ italic_α start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | italic_H | italic_α start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ⟩. After determining the matrix elements Hnmsubscript𝐻𝑛𝑚H_{nm}italic_H start_POSTSUBSCRIPT italic_n italic_m end_POSTSUBSCRIPT for a given gate voltage configuration, the Hamiltonian is diagonalized to find the charge state associated with the ground state energy. This allows us to identify the charge state of a linear QD array as a function of the voltages applied to the gate electrodes.

3 Results and Discussion

The following subsections 1) compare analytic and simulated results for onsite and inter-Coulomb energy calculated for a theoretical double QD electrostatic potential, and 2) present a charge stability diagram (CSD) calculated from realistic double QD electrostatic potentials. An electrostatic potential of two QDs separated by d𝑑ditalic_d is approximated by a quartic function with a confinement strength parameter ω𝜔\omegaitalic_ω:

U(x,y)=m*ω22[(x2d24)2d2+y2]𝑈𝑥𝑦superscript𝑚superscript𝜔22delimited-[]superscriptsuperscript𝑥2superscript𝑑242superscript𝑑2superscript𝑦2U(x,y)=\frac{m^{*}\omega^{2}}{2}\left[\frac{(x^{2}-\frac{d^{2}}{4})^{2}}{d^{2}% }+y^{2}\right]italic_U ( italic_x , italic_y ) = divide start_ARG italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 end_ARG [ divide start_ARG ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 4 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] (9)

and 2D Gaussian functions are used to approximate the single-particle ground state wave functions of each QD as described in Yang2011 . For the CSD, the electrostatic potential as a function of plunger gate voltages, 𝑽=[V1,V2]𝑽subscript𝑉1subscript𝑉2\boldsymbol{V}=[V_{1},V_{2}]bold_italic_V = [ italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ], and tunnel gate voltages, 𝑾=[W1]𝑾delimited-[]subscript𝑊1\boldsymbol{W}=[W_{1}]bold_italic_W = [ italic_W start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ], is calculated via finite-element Poisson solver nextnano Birner2007 , and then used as input to calculate the Hubbard model parameters and determine the charge state for a given voltage configuration.

3.1 Numerical Accuracy and Efficiency

Figure 2(a) plots the onsite energy as a function of confinement strength (in angular frequency units), ω1subscript𝜔1\omega_{1}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and compares the analytic solution to numerically obtained results with a varying number of mesh grid points. Note that the confinement strength ω1subscript𝜔1\omega_{1}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is directly related to the QD radius, r0=/(m*ω1)subscript𝑟0Planck-constant-over-2-pisuperscript𝑚subscript𝜔1r_{0}=\sqrt{\hbar/(m^{*}\omega_{1})}italic_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG roman_ℏ / ( italic_m start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG. For weakly confined electrons (large QD), accurate results can be obtained using fewer mesh grid points, however, as the QD size decreases the number of mesh grid points must increase.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 3: Comparison of numerical integration results to analytic results based on a quartic electrostatic potential. (2(a)) Onsite Coulomb energy U1subscript𝑈1U_{1}italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT versus QD confinement energy, ω1subscript𝜔1\omega_{1}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. (2(b)) Interdot Coulomb energy U12subscript𝑈12U_{12}italic_U start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT versus dot separation, d𝑑ditalic_d. (2(c)) Comparison of two integration methods, showing the average CPU time versus number of mesh grid points evaluated to obtain equivalent U1subscript𝑈1U_{1}italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT results. The data points labelled ‘fast’ refer to an integration algorithm that evaluates mesh grid points only where the electron probability density is non-negligible.

Electrostatically-defined QDs typically have radii on the order of tens of nanometers, giving ω1011𝜔superscript1011\omega\approx 10^{11}italic_ω ≈ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT-1012superscript101210^{12}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT rad/sec, which reduces the required number of mesh grid points and therefore the computation time compared to smaller QDs. Similarly, figure 2(b) shows the inter-Coulomb energy as a function of QD separation parameter, d𝑑ditalic_d, for two QDs with confinement ω1=ω2=51012subscript𝜔1subscript𝜔25superscript1012\omega_{1}=\omega_{2}=5\cdot 10^{12}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 5 ⋅ 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT rad/sec. A greater number of mesh grid points is required for numerical accuracy as interdot separation d𝑑ditalic_d is increased, which is a feature of the mesh grid formalism used in the numerical polar integration.

The numerical integration in (7) involves a quadruple integral, with each variable being integrated over n𝑛nitalic_n points within the coordinate boundaries defining the double QD electrostatic potential. Therefore, the number of mesh grid points to be evaluated during the numerical integration grows as 4nsuperscript4𝑛4^{n}4 start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Figure 2(c) shows how the CPU time scales with the number of mesh grid points for the brute force numerical integration as well as an improved integration algorithm. The improved algorithm only evaluates mesh grid points in regions where the electron is likely to be located, which significantly reduces the overall mesh grid size and therefore the required CPU time.

3.2 Charge Stability Diagram

The effective parameter calculations in section 3.1 were performed for a quartic electrostatic potential that was a function of two control parameters, i.e., the confinement strength ω𝜔\omegaitalic_ω and dot separation d𝑑ditalic_d. In realistic devices, besides the chosen 3D geometry, the control parameters are the plunger, 𝑽𝑽\boldsymbol{V}bold_italic_V, and tunnel, 𝑾𝑾\boldsymbol{W}bold_italic_W, gate voltage vectors that govern the potential landscape.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: Charge stability diagrams (CSD), where [n,m]𝑛𝑚[n,m][ italic_n , italic_m ] labels the electron occupation numbers of the two dots. (3(a)) Theoretical CSD based on a capacitive model in which μi(V1,V2)subscript𝜇𝑖subscript𝑉1subscript𝑉2\mu_{i}(V_{1},V_{2})italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) accounts for cross talk between QDs, while t12subscript𝑡12t_{12}italic_t start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT, U1subscript𝑈1U_{1}italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, U2subscript𝑈2U_{2}italic_U start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and U12subscript𝑈12U_{12}italic_U start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT are held constant. (3(b)) CSD resulting from the numerical Hubbard model approach. The gate voltage-dependent Hubbard parameters are calculated based on realistic QD potentials from a finite element Poisson solver. Two pairs of triple points (in red) are indicated; the left pair are in a regime of stronger tunnel coupling than the pair on the right.

We determine the CSD for a realistic double QD device shown in 0(c) by first constructing a charge basis consisting of 16 charge states with total charge ranging from 0 to 4 electrons, with the Hubbard model restriction that no more than 2 electrons can occupy each QD. The Hubbard Hamiltonian is calculated and diagonalized to find the charge state with lowest energy. This calculation is repeated for all simulated voltage configurations.

In figure 3(a), a theoretical CSD is constructed assuming a capacitive model DasSarma2011 and quartic potential. In the model, μi(V1,V2)subscript𝜇𝑖subscript𝑉1subscript𝑉2{\mu}_{i}(V_{1},V_{2})italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) are functions of the plunger gate voltages for a double QD device, and t12subscript𝑡12t_{12}italic_t start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT, U1subscript𝑈1U_{1}italic_U start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and U12subscript𝑈12U_{12}italic_U start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT are assumed to be constant and voltage-independent. Figure 3(b) shows the CSD resulting from the numerical Hubbard model approach outlined in the present work applied to realistic double QD potentials, such as the one shown in figure 0(c). Although the voltage ranges of the plunger gates are different, the general features shown in figures 3(a) and 3(b) are comparable. Cross-capacitance effects are evident as the sloped charge transition lines. There are two primary advantages in using our Hubbard model approach with realistic QD potentials to determine the CSD in figure 3(b):

  1. 1.

    Subtle voltage dependencies in the onsite and inter-Coulomb energy are captured, which can be seen in figure 3(b) by the varying charge transition line slopes and triple point separations.

  2. 2.

    Strong voltage dependencies on the interdot tunnel coupling are captured, evident as variable smoothing of the charge transitions about the triple points for different regions of the CSD.

4 Conclusion

In this work, an algorithmic approach has been presented for calculating the effective parameters of a Hubbard Hamiltonian that models electrons in realistic linear QD arrays. This was used to identify regions of charge stability as a function of control voltages which captures features missed by other simplified models. Although the results presented are for a QD array of length two, the method generalizes to arbitrary-sized arrays. Further development is underway to perform larger QD array calculations for application to a node network surface code. These methods are a valuable tool within the larger toolbox of simulation methods applicable to the development of scalable spin qubit quantum processors.

References

  • (1) Birner, S., Zibold, T., Andlauer, T., Kubis, T., Sabathil, M., Trellakis, A., Vogl, P.: nextnano: General purpose 3-d simulations. IEEE Transactions on Electron Devices 54(9), 2137–2142 (2007). DOI 10.1109/ted.2007.902871
  • (2) Buonacorsi, B., Cai, Z., Ramirez, E.B., Willick, K.S., Walker, S.M., Li, J., Shaw, B.D., Xu, X., Benjamin, S.C., Baugh, J.: Network architecture for a topological quantum computer in silicon. Quantum Science and Technology 4(2), 025,003 (2019). DOI 10.1088/2058-9565/aaf3c4
  • (3) Das Sarma, S., Wang, X., Yang, S.: Hubbard model description of silicon spin qubits: Charge stability diagram and tunnel coupling in si double quantum dots. Physical Review B 83(23) (2011). DOI 10.1103/physrevb.83.235314
  • (4) Heinz, I., Burkard, G.: Crosstalk analysis for single-qubit and two-qubit gates in spin qubit arrays. Physical Review B 104(4), 045,420 (2021). DOI 10.1103/physrevb.104.045420
  • (5) Hensgens, T., Fujita, T., Janssen, L., Li, X., Van Diepen, C., Reichl, C., Wegscheider, W., Das Sarma, S., Vandersypen, L.: Quantum simulation of a fermi–hubbard model using a semiconductor quantum dot array. Nature 548(7665), 70–73 (2017). DOI 10.1038/nature23022
  • (6) Jafari, S.A.: Introduction to hubbard model and exact diagonalization. arXiv (2008). DOI 10.48550/ARXIV.0807.4878. URL https://arxiv.longhoe.net/abs/0807.4878
  • (7) Mills, A.R., Guinn, C.R., Gullans, M.J., Sigillito, A.J., Feldman, M.M., Nielsen, E., Petta, J.R.: Two-qubit silicon quantum processor with operation fidelity exceeding 99%. Science Advances 8(14) (2022). DOI 10.1126/sciadv.abn5130
  • (8) Petit, L., Eenink, H.G.J., Russ, M., Lawrie, W.I.L., Hendrickx, N.W., Philips, S.G.J., Clarke, J.S., Vandersypen, L.M.K., Veldhorst, M.: Universal quantum logic in hot silicon qubits. Nature 580(7803), 355–359 (2020). DOI 10.1038/s41586-020-2170-7
  • (9) Saraiva, A., Lim, W.H., Yang, C.H., Escott, C.C., Laucht, A., Dzurak, A.S.: Materials for silicon quantum dots and their impact on electron spin qubits. Advanced Functional Materials 32(3) (2021). DOI 10.1002/adfm.202105488
  • (10) Secchi, A., Troiani, F.: Theory of multidimensional quantum capacitance and its application to spin and charge discrimination in quantum dot arrays. Physical Review B 107(15) (2023). DOI 10.1103/physrevb.107.155411
  • (11) Undseth, B., Xue, X., Mehmandoost, M., Rimbach-Russ, M., Eendebak, P.T., Samkharadze, N., Sammak, A., Dobrovitski, V.V., Scappucci, G., Vandersypen, L.M.: Nonlinear response and crosstalk of electrically driven silicon spin qubits. Physical Review Applied 19(4), 044,078 (2023). DOI 10.1103/physrevapplied.19.044078
  • (12) Veldhorst, M., Eenink, H.G.J., Yang, C.H., Dzurak, A.S.: Silicon CMOS architecture for a spin-based quantum computer. Nature Communications 8(1) (2017). DOI 10.1038/s41467-017-01905-6
  • (13) Veldhorst, M., Hwang, J.C.C., Yang, C.H., Leenstra, A.W., de Ronde, B., Dehollain, J.P., Muhonen, J.T., Hudson, F.E., Itoh, K.M., Morello, A., Dzurak, A.S.: An addressable quantum dot qubit with fault-tolerant control-fidelity. Nature Nanotechnology 9(12), 981–985 (2014). DOI 10.1038/nnano.2014.216
  • (14) Yang, S., Wang, X., Das Sarma, S.: Generic hubbard model description of semiconductor quantum-dot spin qubits. Physical Review B 83(16) (2011). DOI 10.1103/physrevb.83.161301
  • (15) Zhong, H.S., Wang, H., Deng, Y.H., Chen, M.C., Peng, L.C., Luo, Y.H., Qin, J., Wu, D., Ding, X., Hu, Y., Hu, P., Yang, X.Y., Zhang, W.J., Li, H., Li, Y., Jiang, X., Gan, L., Yang, G., You, L., Wang, Z., Li, L., Liu, N.L., Lu, C.Y., Pan, J.W.: Quantum computational advantage using photons. Science 370(6523), 1460–1463 (2020). DOI 10.1126/science.abe8770