File failed to load: https://mathjax.xml-journal.net/platformTools/js/MathJax-master/config/TeX-AMS-MML_HTMLorMML.js
J. Semicond. > 2024, Volume 45 > Issue 10 > 101701

REVIEWS

The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing

Zheng Zhou, Yixin Li, Zhiyuan Wu, Xinping Ma, Shichang Fan and Shaoyun Huang

+ Author Affiliations

 Corresponding author: Shaoyun Huang, syhuang@pku.edu.cn

DOI: 10.1088/1674-4926/24050043CSTR: 32376.14.1674-4926.24050043

PDF

Turn off MathJax

Abstract: Electron spins confined in semiconductor quantum dots (QDs) are one of potential candidates for physical implementation of scalable quantum information processing technologies. Tunnel coupling based inter exchange interaction between QDs is crucial in achieving single-qubit manipulation, two-qubit gate, quantum communication and quantum simulation. This review first provides a theoretical perspective that surveys a general framework, including the Helter−London approach, the Hund−Mulliken approach, and the Hubbard model, to describe the inter exchange interactions between semiconductor quantum dots. An electrical method to control the inter exchange interaction in a realistic device is proposed as well. Then the significant achievements of inter exchange interaction in manipulating single qubits, achieving two-qubit gates, performing quantum communication and quantum simulation are reviewed. The last part is a summary of this review.

Key words: exchange interactionquantum dotstunnel couplingquantum computation

Various physical computing devices have been explored in quantum mechanical regime for more than two decades to implement the primitive idea of quantum computation[14], which can be seen as superior to classical computation in exponentially faster executing certain computational tasks and simulating complex quantum systems[57]. The electron spin degree of freedom confined in semiconductor quantum dots (QDs) is one of leading candidates to construct physical quantum bits (qubits) for quantum information processors[814]. The fabrication processes of semiconductor QDs can take advantages of modern semiconductor chip engineering[12, 15, 16]. Remarkably, the processes of large-scale and high-quality quantum dot arrays on silicon are fully compatible to an industry-standard CMOS platform[17, 18]. The quantum dot arrays provide convenient building-blocks to achieve large-scale integration of qubits for universal quantum computation[19]. The precise control of interaction between quantum dots is an essential step in single qubit manipulation, two-qubit gate execution, quantum communication and quantum simulation[15]. The underlying physical mechanism leading to the interaction between quantum dots is the exchange interaction ($J$) between the confined electron spins in each dot. The exchange interaction, in general, is in short range, strong between confined electrons on the identical dot (intra-dot exchange). The tunnel coupling, however, favors finite exchange interaction taking place in long range between the nearest neighbor dots, and even between remote dots via intermediate dot(s) (inter-dot exchange). The latter is termed as superexchange[20, 21]. We limit our discussion on the inter-dot exchange interaction between the nearest neighbor quantum dots and prefer to refer it simply to exchange interaction if not specifically stated. Coupling more than two quantum dots, in principle, can be achieved by the assistance of the coupling between the nearest neighboring quantum dots. The comprehensive study of the exchange interaction between quantum dots is indispensable to exemplify the electrical method to control the interaction desired in developing the exchange dependent universal quantum computations.

Quantum dots are man-made microstructures and behave in many ways as artificial atoms with reduced energy scales[22, 23]. Therefore, the interactions between quantum dots can be explored using the physical model in study of atom bonding. For two electrons residing in double quantum dots, the method used to study a hydrogen molecule, such as the Heitler−London approach[24, 25] and the Hund−Mulliken approach[26] can help us understand the exchange interactions between the quantum dots. Furthermore, combining Hubbard and constant interaction models, we can develop electrical methods to control $J$ in a realistic device.

The review concentrates on exploring the underlying mechanism of the exchange interaction between the nearest neighbor quantum dots. The physical mechanisms of exchange interaction and practical method to control the exchange interaction are discussed in Section 2. The role of exchange interaction in constructing qubits, realizing two-qubit gates, performing quantum communication and quantum simulation is intensively surveyed in Section 3. Section 4 is the summary of the review.

The exchange interaction had been discovered independently by Heisenberg[27] and Dirac[28] in 1920s. The Pauli exclusion principle states that no two fermions can have the same four electronic quantum numbers[29]. Thus, the principle exerts a strong influence on each other because a fermion occupying a state excludes the others from it[30]. This is equivalent to a strong repulsive force which is the source of exchange interaction between fermions[31]. The repulsion requires that the wave function of indistinguishable fermions is in antisymmetricity, i.e. the wave function changes sign when two fermions are exchanged. The physical effect of the antisymmetric requirement increases the curvature of wavefunctions to prevent the overlap of the states occupied by indistinguishable fermions. As a result, the exchange antisymmetricity increases the expectation value of the distance between two fermions when their wave functions overlap[32].

Electrons with spin half are fermions. The overall wave function of the system composed of electrons must be antisymmetric. This means if the orbital wave function is symmetric, the spin one must be antisymmetric, vice versa[33]. The exchange interaction $J$ is, therefore, the energy difference between the antisymmetric orbital wave function (symmetric spin wave function) and the symmetric orbital wave function (antisymmetric spin wave function) to reflect the Pauli repulsion principle[24], as shown in Fig. 1. In this section, we study the exchange interaction between quantum dots and propose electrical methods to control the exchange interaction.

Fig. 1.  (Color online) The exchange interaction $J$ is the energy difference between the antisymmetric orbital wave function and the symmetric orbital wave function of two electrons.

The Heitler−London (HL) approach was used to describe valence bonding between two atoms and succeeded in explaining the formation of hydrogen molecule[25]. Consequently, the interaction between double quantum dots with one electron in each dot can be approximated by employing the HL approach[34, 35].

The HL approach, applied to double quantum dots, associates the Hamiltonian as follows[34]:

$$ \hat H = \sum\limits_{k = 1}^2 {\hat h({{\boldsymbol{r}}_k})} + C({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_{2}}) . $$ (1)

Here, $ \hat h({{\boldsymbol{r}}_k}) = - \dfrac{{{\hbar ^2}}}{{2m}}\nabla _{{r_k}}^2 + V({{\boldsymbol{r}}_k}) $ is the one-body Hamiltonian of the k-th electron containing kinetic energy and confinement energy in an external potential, $ C({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_{ 2}}) = \dfrac{{{e^2}}}{{4\pi {\varepsilon _0}{\varepsilon _{\text{r}}}\left| {{{\boldsymbol{r}}_1} - {{\boldsymbol{r}}_2}} \right|}} $ the Coulomb interaction between two electrons located at $ {{\boldsymbol{r}}_{\text{1}}} $ and $ {{\boldsymbol{r}}_{\text{2}}} $ separately, $ e\text{ }=\text{ }1.602\text{ }\times\text{ }10^{-19}\text{ C} $ the elementary charge, $ m{\text{ }} = {\text{ }}9.1{\text{ }} \times {\text{ }}{10^{ - 31}}{\text{ kg}} $ the mass of free electron. In a crystal, we can replace $ m $ with the effective mass ${m^*}$ as an approximation[36].

For gate-defined double quantum dots in two dimensional electron gas (2DEG) with identical confinement profile, we can simulate the $ V({\boldsymbol{r}}) $ with a quartic potential, in which two harmonic wells possess identical frequency of ${\omega _0}$ centered at $( - a,0)$ and $( + a,0)$, as shown in Fig. 2(a)[37]. The $ V({\boldsymbol{r}}) $ reads:

Fig. 2.  (Color online) (a) Quartic potential as illustrated in Eq. (2) with $ \hbar\omega_0=3\text{ meV} $ and $ a=19\text{ nm} $. The potential is used to simulate the coupling of two electrons locating in two harmonic wells centered at $ \left( { - a,0} \right) $ and $(a,0)$. The effective Bohr radius of harmonic well is ${a_{\text{B}}} = \sqrt {\hbar /m{\omega _0}} $. (b) The exchange coupling strength $J$ between two spins as a function of the inter-dot spacing $d = a/{a_{\text{B}}}$ with $ \hbar\omega_0=3\text{ meV} $, $ a_{\text{B}}=19\text{ nm} $ and $c = 2.4$ [See Eq. (6)].
$$ V({\boldsymbol{r}}) = V(x,y) = \frac{{m\omega _0^2}}{2}\left[ {\frac{1}{{4{a^2}}}{{\left( {{x^2} - {a^2}} \right)}^2} + {y^2}} \right] . $$ (2)

The (anti)symmetric two-electron orbital state-vector ($\left| {{\psi _ - }} \right\rangle $) $\left| {{\psi _ + }} \right\rangle $ can be generated by combining single-dot ground-state orbital wave-functions as follows:

$$ \begin{split} & \left| {{\psi _ - }} \right\rangle = \frac{{\left| {a\bar a} \right\rangle - \left| {\bar aa} \right\rangle }}{{\sqrt {2\left( {1 - {\Omega ^2}} \right)} }}, \\ & \left| {{\psi _ + }} \right\rangle = \frac{{\left| {a\bar a} \right\rangle + \left| {\bar aa} \right\rangle }}{{\sqrt {2\left( {1 + {\Omega ^2}} \right)} }}. \end{split} $$ (3)

Here, $ \left| a \right\rangle $ and $\left| {\bar a} \right\rangle $ are the Dirac notations to indicate the electron ground state in the quantum dot centering at $\left( {a,0} \right)$ and $\left( { - a,0} \right)$, respectively. $\left| {a\bar a} \right\rangle = \left| a \right\rangle \left| {\bar a} \right\rangle $ and $\left| {\bar aa} \right\rangle = \left| {\bar a} \right\rangle \left| a \right\rangle $ are two-electron product states. $ \Omega = \int {{\varphi _{ - a}}({\boldsymbol{r}})\varphi _{ + a}^*({\boldsymbol{r}}){\rm{d}^2}r} = \left\langle {\left. a \right|\overline a } \right\rangle $ is the overlap integral of two orbital wave functions. ${\varphi _{ - a}}({\boldsymbol{r}}) = \left\langle {\left. {\boldsymbol{r}} \right|\overline a } \right\rangle $ and ${\varphi _{ + a}}({\boldsymbol{r}}) = \left\langle {\left. {\boldsymbol{r}} \right|a} \right\rangle $ denote the one-electron ground orbital functions centered at $\left( { - a,0} \right)$ and $\left( {a,0} \right)$, respectively. Analogy to harmonic oscillator eigenfunctions, they can be expressed as follows:

$$ \begin{split} & {\varphi _{ + a}}(x,y) = \sqrt {\frac{{m{\omega _0}}}{{\pi \hbar }}} {\rm{e}^{ - m{\omega _0}\left[ {{{(x - a)}^2} + {y^2}} \right]/2\hbar }}, \\ & {\varphi _{ - a}}(x,y) = \sqrt {\frac{{m{\omega _0}}}{{\pi \hbar }}} {\rm{e}^{ - m{\omega _0}\left[ {{{(x + a)}^2} + {y^2}} \right]/2\hbar }}. \end{split} $$ (4)

The energy difference between two states $\left| {{\psi _ - }} \right\rangle $ and $\left| {{\psi _ + }} \right\rangle $ in Eq. (3), which differ only in their symmetry, is referred as the exchange energy and is written as follows:

$$ J = {\varepsilon _{\text{t}}} - {\varepsilon _{\text{s}}} = \left\langle {{\psi _ - }\left| {\hat H} \right|{\psi _ - }} \right\rangle - \left\langle {{\psi _ + }\left| {\hat H} \right|{\psi _ + }} \right\rangle . $$ (5)

From Eqs. (1)−(5) of the Heitler–London approach and making use of the Darwin–Fock solution to the isolated dots, we find[38] :

$$ J = \frac{{\hbar {\omega _0}}}{{\sinh \left( {2{d^2}} \right)}}\left\{ {c\left[ {{\rm{e}^{ - {d^2}}}{I_0}({d^2}) - 1} \right] + \frac{3}{4}\left( {1 + {d^2}} \right)} \right\} . $$ (6)

Here, unitless $c = \sqrt {2\pi } \left[ {{e^2}/\left( {4\pi {\varepsilon _0}{\varepsilon _{\text{r}}}{a_{\text{B}}}} \right)} \right]/2\hbar {\omega _0}$ is the ratio of Coulomb energy to confinement energy, and ${I_0}$ the zeroth-order Bessel function. If we set the energy of harmonic wells $\hbar {\omega _0}$ to be $3{\text{ meV}}$, we have ${a_{\text{B}}}$ at 19 nm and $c$ at 2.4. Numerical calculation to Eq. (6) shows that $J$ decays exponentially with increasing inter-dot spacing $d \equiv a/{a_{\text{B}}}$, as shown in Fig. 2(b).

The HL approach, however, does not contain any information about the double occupied states[15]. We have to take into account the hybridization between single-occupied state (1, 1) and double-occupied states (2, 0) as well as (0, 2) by extending the Heitler−London approach to the Hund−Mulliken approach[26], where $({n_1},{n_2})$ means that the number of electrons in quantum dot centered at $( - a,0)$ and $( + a,0)$, respectively. Herein, we limit ourselves in $\left| {S(2,0)} \right\rangle $, $\left| {S(0,2)} \right\rangle $, $ \left| {S(1,1)} \right\rangle $, and $\left| {{T_0}(1,1)} \right\rangle $ four elementary states. $\left| S \right\rangle = \left( {\left| { \uparrow \downarrow } \right\rangle - \left| { \downarrow \uparrow } \right\rangle } \right)/ \sqrt 2 $ is the singlet state of two electron spins and ${T_0} = \left( {\left| { \uparrow \downarrow } \right\rangle + \left| { \downarrow \uparrow } \right\rangle } \right)/\sqrt 2 $ is the triplet state of two electron spins at ${S_{{\text{z12}}}} = 0$, where ${S_{{\text{z12}}}}$ is the spin projection of two electron spins in an arbitrary direction (e.g. the direction of external static magnetic field). The state $\left| {{T_0}(0,2)} \right\rangle $ is energetically inaccessible due to great intra exchange interaction with the state $\left| {S(0,2)} \right\rangle $[8]. When an external magnetic field is applied, the triplet states of $\left| {{T_ + }} \right\rangle = \left| { \uparrow \uparrow } \right\rangle $ and $\left| {{T_ - }} \right\rangle = \left| { \downarrow \downarrow } \right\rangle $ at ${S_{{\text{z12}}}} = \pm 1$ are lifted from the state $\left| {{T_0}(1,1)} \right\rangle $ and are no longer accessible, too.

Two single-occupied states $ \Psi _{{ + }}^{\text{s}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) $ and $ \Psi _ - ^{\text{s}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) $ in either symmetric (+) or antisymmetric (-) forms, and two double-occupied states $ \Psi _{ - a}^{\text{d}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) $ and $ \Psi _{ + a}^{\text{d}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) $ at $ - a$ and $ + a$ positions can be constructed in ground states as follows:

$$ \left\{ \begin{aligned} & \Psi _{ - a}^{\text{d}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) = {\Phi _{ - a}}({{\boldsymbol{r}}_1}){\Phi _{ - a}}({{\boldsymbol{r}}_2})&S(2,0) \; \\ & \Psi _{ + a}^{\text{d}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) = {\Phi _{ + a}}({{\boldsymbol{r}}_1}){\Phi _{ + a}}({{\boldsymbol{r}}_2})&S(0,2) \; \\ & \Psi _ + ^{\text{s}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) = [{\Phi _{ + a}}({{\boldsymbol{r}}_1}){\Phi _{ - a}}({{\boldsymbol{r}}_2}) + {\Phi _{ - a}}({{\boldsymbol{r}}_1}){\Phi _{ + a}}({{\boldsymbol{r}}_2})]/\sqrt 2 &S(1,1)\; \\ & \Psi _ - ^{\text{s}}({{\boldsymbol{r}}_1},{{\boldsymbol{r}}_2}) = [{\Phi _{ + a}}({{\boldsymbol{r}}_1}){\Phi _{ - a}}({{\boldsymbol{r}}_2}) - {\Phi _{ - a}}({{\boldsymbol{r}}_1}){\Phi _{ + a}}({{\boldsymbol{r}}_2})]/\sqrt 2 &{T_0}(1,1). \\ \end{aligned} \right. $$ (7)

Here, ${\Phi _{ + a}} = \left( {{\varphi _{ + a}} - g{\varphi _{ - a}}} \right)/\sqrt {1 - 2\Omega g + {g^2}} $ and ${\Phi _{ - a}} = \left( {{\varphi _{ - a}} - g{\varphi _{ + a}}} \right)/ \sqrt {1 - 2\Omega g + {g^2}} $ are orthogonalized single-electron wave functions centered at $( + a,0)$ and $( - a,0)$, respectively, and $g = \left( {1 - \sqrt {1 - {\Omega ^2}} } \right)/\Omega $ is a coefficient used to achieve the orthogonalization. Since the symmetricity between orbital and spin wave functions must be opposite, we can write the corresponding spin state respect to the orbital wave function as shown in Eq. (7).

Taking Eq. (7) as the bases, we can express the Hamiltonian of Eq. (1) in a matrix as follows:

$$ \hat H = 2\varepsilon + \left( {\begin{array}{*{20}{c}} U&{{V_{\text{x}}}}&{\sqrt 2 {t_{\text{H}}}}&0 \\ {{V_{\text{x}}}}&U&{\sqrt 2 {t_{\text{H}}}}&0 \\ {\sqrt 2 {t_{\text{H}}}}&{\sqrt 2 {t_{\text{H}}}}&{{V_{{\text{inter}}}} + {V_{\text{x}}}}&0 \\ 0&0&0&{{V_{{\text{inter}}}} - {V_{\text{x}}}} \end{array}} \right). $$ (8)

The elements of matrix in Eq. (8) are as follows:

$$ \begin{split} & \varepsilon = \left\langle {{\Phi _a}\left| {\hat h} \right|{\Phi _a}} \right\rangle = \left\langle {{\Phi _{ - a}}\left| {\hat h} \right|{\Phi _{ - a}}} \right\rangle , \\ & {t_{\text{H}}} = \left\langle {{\Phi _a}\left| {\hat h} \right|{\Phi _{ - a}}} \right\rangle + W = \left\langle {{\Phi _{ - a}}\left| {\hat h} \right|{\Phi _a}} \right\rangle + W, \\ & W = \left\langle {{\Phi _a}{\Phi _a}\left| C \right|{\Phi _a}{\Phi _{ - a}}} \right\rangle = \left\langle {{\Phi _{ - a}}{\Phi _{ - a}}\left| C \right|{\Phi _{ - a}}{\Phi _a}} \right\rangle , \\ & {V_{{\text{inter}}}} = \left\langle {{\Phi _a}{\Phi _{ - a}}\left| C \right|{\Phi _a}{\Phi _{ - a}}} \right\rangle = \left\langle {{\Phi _{ - a}}{\Phi _a}\left| C \right|{\Phi _{ - a}}{\Phi _a}} \right\rangle , \\ & {V_{\text{x}}} = \left\langle {{\Phi _{ - a}}{\Phi _a}\left| C \right|{\Phi _a}{\Phi _{ - a}}} \right\rangle = \left\langle {{\Phi _a}{\Phi _{ - a}}\left| C \right|{\Phi _{ - a}}{\Phi _a}} \right\rangle , \\ & U = \left\langle {{\Phi _a}{\Phi _a}\left| C \right|{\Phi _a}{\Phi _a}} \right\rangle = \left\langle {{\Phi _{ - a}}{\Phi _{ - a}}\left| C \right|{\Phi _{ - a}}{\Phi _{ - a}}} \right\rangle . \end{split} $$ (9)

Here, the Coulomb matrix element $U$ is intra-dot Coulomb interaction energy, ${V_{{\text{inter}}}}$ inter-dot Coulomb interaction, ${V_{\text{x}}}$ the exchange energy in the tunnel barrier, and $W$ remaining energy[39].

The exchange energy between two dots can be determined by the energy gap between the singlet and triplet ground states by solving the eigenvalues of the Hamiltonian in Eq. (8), and reaches at:

$$ J=\frac{1}{2}\left(\sqrt{U_{\text{H}}^{\text{2}}+16t_{\text{H}}^{\text{2}}}-U_{\text{H}}\right)-2V_{\text{x}}, $$ (10)

where $ {U_{\text{H}}} \equiv U - {V_{{\text{inter}}}} $ is on-site repulsion renormalized by long-range Coulomb interactions[37].

When the two dots are weakly coupled, ${V_{\text{x}}}$ and $W$ were negligible since they are very small[40]. Though either ${V_{\text{x}}}$ or $W$ is not small enough compared with $J$ for relatively strong coupled double quantum dots, they robustly stay constant in most of experiments[26]. Therefore, we set ${V_{\text{x}}}$ and $W$ as zero, and Eq. (10) can be simplified as follows[41]:

$$ J=\frac{1}{2}\left(\sqrt{U_{ }^2+16t_{\text{c}}^{\text{2}}}-U\right), $$ (11)

where

$$ {t_{\text{c}}} = \left\langle {{\Phi _a}\left| {\hat h} \right|{\Phi _{ - a}}} \right\rangle = \left\langle {{\Phi _{ - a}}\left| {\hat h} \right|{\Phi _a}} \right\rangle . $$ (12)

If the double quantum dots are at weak coupling limit, i.e. the Hubbard ratio $ {t}_{c}/U\ll 1 $, the Eq. (11) can be further simplified to:

$$ J \approx \frac{{4t_{\text{c}}^{\text{2}}}}{U}. $$ (13)

In order to control the exchange interaction in a realistic device, we correlate the device parameters with the exchange interaction $J$ using the Hubbard model. Gate-defined double quantum dots can be numerically simulated in a GaAs/AlGaAs heterostructure[42], as shown in Fig. 3(a). By doping the AlGaAs layer with Si atoms, we can introduce free electrons in the quantum well at the interface between AlGaAs and GaAs layers. The electrons accumulate at the GaAs/AlGaAs interface forming a 2DEG, a thin (~10 nm) sheet of electrons that can only move along the interface. A negative potential applied on barrier gates ${V_{\text{b}}}$, ${V_{{\text{b1}}}}$, and ${V_{{\text{b2}}}}$ locally depletes the underneath 2DEG to form double quantum dots. By adjusting ${V_{\text{b}}}$, we can manipulate the barrier height between neighboring quantum dots to control the coupling strength between the two dots. Gate voltages ${V_{{\text{g1}}}}$ and ${V_{{\text{g2}}}}$ are used to alter the number of electrons residing in left dot 1 and right dot 2, respectively. The total Hamiltonian of the double quantum dots can be expressed by the Hubbard model as follows[43, 44]:

Fig. 3.  (Color online) (a) Double quantum dots in GaAs/AlGaAs heterostructure. (b) Two dimensional stability diagram of double quantum dots. The purple dashed-line is the position where $\Delta \varepsilon $ is zero, the yellow dashed-arrow indicates the direction along which $\Delta \varepsilon $ increases.
$$ {\hat H_{{\text{tot}}}} = \hat U({\hat n_1},{\hat n_2}) + \sum\limits_{\sigma = \uparrow , \downarrow } {{t_{\text{c}}}\left( {\hat c_{1,\sigma }^\dagger {{\hat c}_{2,\sigma }} + h.c.} \right)} , $$ (14)

where the operator $\hat c_{1,\sigma }^\dagger $(${\hat c_{2,\sigma }}$) charges (discharges) an electron in QD 1(2) with spin $\sigma = \uparrow , \downarrow $. ${\hat n_i} = \sum\nolimits_{\sigma = \uparrow , \downarrow } {\hat c_{i,\sigma }^\dagger \hat c_{i,\sigma }^{}} = {\hat n_{i \uparrow }} + {\hat n_{i \downarrow }}$ is a number operator of QD $i$ ($i$ = 1, 2), $ {t_{\text{c}}} $ a hopping matrix element, known as "tunnel coupling" between two quantum dots[45, 46].

We limit our discussion on no more than two electrons accommodating in the double quantum dots. The total electrostatic energy of the double quantum dots, $ \hat U({\hat n_1},{\hat n_2}) $, can be expressed as follows[45] :

$$ \begin{split} & \hat U({{\hat n}_1},{{\hat n}_2}) = \sum\limits_{i = 1}^2 {\left[ {\frac{{{E_{{\text{C}}i}}}}{2}{{\hat n}_i}({{\hat n}_i} - 1) + {\varepsilon _i}{{\hat n}_i}} \right]} + {{\hat n}_1}{{\hat n}_2}{E_{{\text{Cm}}}} + f({V_{{\text{g1}}}},{V_{{\text{g2}}}}), \\ & {\varepsilon _1} \equiv - \left( {{C_{{\text{g1}}}}{V_{{\text{g1}}}}{E_{{\text{C1}}}} + {C_{{\text{g2}}}}{V_{{\text{g2}}}}{E_{{\text{Cm}}}}} \right)/e,\\ &{\varepsilon _2} \equiv - \left( {{C_{{\text{g2}}}}{V_{{\text{g2}}}}{E_{{\text{C2}}}} + {C_{{\text{g1}}}}{V_{{\text{g1}}}}{E_{{\text{Cm}}}}} \right)/e, \\ & f({V_{{\text{g1}}}},{V_{{\text{g2}}}}) \equiv\left[ \frac{1}{2}C_{{\text{g1}}}^{\text{2}}V_{{\text{g1}}}^{\text{2}}{E_{{\text{C1}}}} + \frac{1}{2}C_{{\text{g2}}}^{\text{2}}V_{{\text{g2}}}^{\text{2}}{E_{{\text{C2}}}}+ {C_{{\text{g1}}}}{V_{{\text{g1}}}}{C_{{\text{g2}}}}{V_{{\text{g2}}}}\right.\\ &\quad\quad\quad\quad\; \times{E_{{\text{Cm}}}} \bigg]/{e^2}.\\[-8pt] \end{split} $$ (15)

Here, ${E_{{\text{C}}i}}$ is the charging energy of individual single dot i, ${E_{{\text{Cm}}}}$the electrostatic coupling energy between the two dots.

Since ${t_{\text{c}}}$($\sim 10{\text{ }}\mu {\text{eV}}$) is significantly smaller than $U({n_1},{n_2})$ ($\sim 5{\text{ meV}}$)[47], the Froehlich−Nakajima transformation (also known as Schrieffer–Wolff transformation) can be used to obtain ${\hat H_{{\text{eff}}}}$ in (1, 1) charge state subspace[4850] from ${\hat H_{{\text{tot}}}}$ as follows:

$$ {\hat H_{{\text{eff}}}} = \hat U(1,1) + J\left( {{{{\boldsymbol{\hat S}}}_1} \cdot {{{\boldsymbol{\hat S}}}_2} - \frac{{{\hbar ^2}}}{4}} \right)/{\hbar ^2}, $$ (16)

where $ {{\boldsymbol{\hat S}}_i} = \dfrac{\hbar }{2}{\displaystyle\sum}_{\begin{subarray}{l} \tau = \uparrow , \downarrow \\ \tau ' = \uparrow , \downarrow \end{subarray}} {\hat c_{i\tau }^\dagger {\hat{{\boldsymbol{{σ}}}}_{\tau \tau '}}\hat c_{i\tau '}} $ is the spin operator of the electron in dot i[51], and ${\hat{{\boldsymbol{{σ}}}}} \equiv \left( {{\hat\sigma _x},{\hat\sigma _y},{\hat\sigma _z}} \right)$ is Pauli operator. $J$ is the exchange interaction between two electron spins and can deduced from Eq. (14):

$$ J = \frac{{4t_{\text{c}}^{\text{2}}}}{{\dfrac{{2\left[ {U(0,2) - U(1,1)} \right]\left[ {U(2,0) - U(1,1)} \right]}}{{\left[ {U(2,0) - U(1,1)} \right] + \left[ {U(0,2) - U(1,1)} \right]}}}}. $$ (17)

Based on Eq. (15), we can rewrite Eq. (17) as follows:

$$ J = \frac{{4t_{\text{c}}^{\text{2}}}}{{\dfrac{{2\left[ {{E_{{\text{C2}}}} - \Delta \varepsilon - {E_{{\text{Cm}}}}} \right]\left[ {{E_{{\text{C1}}}} + \Delta \varepsilon - {E_{{\text{Cm}}}}} \right]}}{{{E_{{\text{C1}}}} + {E_{{\text{C2}}}} - 2{E_{{\text{Cm}}}}}}}}, $$ (18)

where $\Delta \varepsilon = {\varepsilon _2} - {\varepsilon _1}$ is a detuning energy and can be manipulated by ${V_{{\text{g1}}}}$ and ${V_{{\text{g2}}}}$[8, 13] as shown in Eq. (15). The detuning direction is illustrated by a yellow-dashed arrow in the two-dimensional stability diagram of double quantum dots, as shown in Fig. 3(b).

The tunnel coupling ${t_{\text{c}}}$ in Eq. (18) is defined in Eq. (12). For the single-occupied two quantum dots, the tunneling coupling ${t_{\text{c}}}$ hybridizes charge states (1, 0) and (0, 1). Taking $\left| {{\Phi _{ + a}}} \right\rangle $ and $\left| {{\Phi _{ - a}}} \right\rangle $ as bases, we can express the Hamiltonian (Eq. (14)) as follows:

$$ \hat H = \left( {\begin{array}{*{20}{c}} {{\varepsilon _1}}&{{t_{\text{c}}}} \\ {{t_{\text{c}}}}&{{\varepsilon _2}} \end{array}} \right) . $$ (19)

${\varepsilon _1}$ and ${\varepsilon _2}$, the energy of charge states (1, 0) and (0, 1), can be controlled by gate voltages ${V_{{\text{g1}}}}$ and ${V_{{\text{g2}}}}$, respectively, as shown in Eq. (15). The eigenstates and corresponding eigenenergies[45] of Hamiltonian (Eq. (19)) are shown in Table 1.

Table 1.  The eigenstates and the corresponding eigenenergy of the Hamiltonian (Eq. (19)).
EigenstateEigenenergyDeviation
$\left| {{\psi _{\text{A}}}} \right\rangle $${\varepsilon _{\text{M}}} + \sqrt {{{(\Delta \varepsilon )}^2}/4 + t_{\text{c}}^{\text{2}}} $$\Delta = \sqrt {{{\left( {\Delta \varepsilon } \right)}^2} + 4t_{\text{c}}^{\text{2}}} $
$\left| {{\psi _{\text{B}}}} \right\rangle $${\varepsilon _{\text{M}}} - \sqrt {{{(\Delta \varepsilon )}^2}/4 + t_{\text{c}}^{\text{2}}} $
DownLoad: CSV  | Show Table

The terms in the Table 1 can be expressed as follows:

$$ \begin{split} & \left| {{\psi _{\text{A}}}} \right\rangle = \cos \frac{\theta }{2}\left| {{\Phi _{ - a}}} \right\rangle + \sin \frac{\theta }{2}\left| {{\Phi _{ + a}}} \right\rangle , \\ & \left| {{\psi _{\text{B}}}} \right\rangle = - \sin \frac{\theta }{2}\left| {{\Phi _{ - a}}} \right\rangle + \cos \frac{\theta }{2}\left| {{\Phi _{ + a}}} \right\rangle , \\ & \tan \theta = 2{t_{\text{c}}}/\Delta , \\ & {\varepsilon _{\text{M}}} = \left( {{\varepsilon _1} + {\varepsilon _2}} \right)/2. \end{split} $$ (20)

The inter-dot tunnel coupling strength ${t_{\text{c}}}$ hybridizes charge state (0, 1) with (1, 0) around the energy degeneracy point $\Delta \varepsilon = 0$, resulting in an avoiding cross with an energy splitting of 2${t_{\text{c}}}$, as shown in Fig. 4(a). We can adiabatically transfer an electron from the right quantum dot to the left quantum dot by pushing $\Delta \varepsilon $ towards the positive side, as shown in Fig. 3(b). The transition process does not contribute a tunnel current through the device but can be monitored by a proximal charge sensor[52]. When tunnel coupling ${t_{\text{c}}}$ is smaller than the single-electron level spacing of individual dot, the profile of sensor conductance in the transition from (0, 1) to (1, 0) state, vice versa, can be described by a thermal equilibrium two-level model as follows[53, 54]:

Fig. 4.  (Color online) (a) Energy of $\left| {{\psi _{\text{A}}}} \right\rangle $ and $\left| {{\psi _{\text{B}}}} \right\rangle $ versus the detuning energy $\Delta \varepsilon $. (b) The electrostatic potential $U$ of quantum dots in Fig. 3(a) along the X-axis at Y ~ $0.08{\text{ }}\mu {\text{m}}$. The barrier height ${E_{\text{B}}}$ between two quantum dots can be manipulated by the barrier gate voltage ${V_{\text{b}}}$. (c) The barrier height ${E_{\text{B}}}$ is negatively proportional to the barrier gate voltage ${V_{\text{b}}}$. The black straight line is a linear fit.
$$ {g_{\text{s}}} = {g_0} + \delta g\frac{{\Delta \varepsilon }}{\Delta }\tanh \left( {\frac{\Delta }{{2{k_{\text{B}}}{T_{\text{e}}}}}} \right), $$ (21)

where ${g_{\text{s}}}$ is the conductance of charge sensor, ${T_{\text{e}}}$ the electron temperature, $\Delta = \sqrt {{{\left( {\Delta \varepsilon } \right)}^2} + 4t_{\text{c}}^{\text{2}}} $ the half of the energy difference between $\left| {{\psi _{\text{A}}}} \right\rangle $ and $\left| {{\psi _{\text{B}}}} \right\rangle $, and ${k_{\text{B}}}$ the Boltzmann constant. Since $\Delta $ can be extracted experimentally according to Eq. (21), ${t_{\text{c}}}$ can be obtained subsequently.

Since the confined electrons of quantum dots locate at the interface between GaAs and AlGaAs (Fig. 3(a)) and the transport of electrons is primarily along the X-axis, the potential energy profile along the X-axis at Y ~ 0.08 $\mu {\text{m}}$ can be used to describe the two quantum dots, as shown in Fig. 4(b). The approximation can help us to simplify the calculation cost of the coupling strength. The barrier height ${E_{\text{B}}}$ between two quantum dots can be generated and manipulated by the barrier gate voltage ${V_{\text{b}}}$, as shown in Fig. 4(c). With the help of simplified one-dimensional model, as shown in Fig. 5(a), we can further explore the quantitative relationship between ${t_{\text{c}}}$ and barrier height ${E_{\text{B}}}$. The tunnel coupling ${t_{\text{c}}}$ mixes the wave functions of two quantum dots $\left| {{\Phi _{ + a}}} \right\rangle $ and $\left| {{\Phi _{ - a}}} \right\rangle $ to form a bonding state $\left| {{\psi _{\text{B}}}} \right\rangle $ and an antibonding state $\left| {{\psi _{\text{A}}}} \right\rangle $, as shown in Table 1 and Eq. (20). By solving single-electron stationary Schrodinger equation in appendix A, we can deduce the tunnel coupling strength ${t_{\text{c}}}$ from the energy difference between $\left| {{\psi _{\text{B}}}} \right\rangle $ and $\left| {{\psi _{\text{A}}}} \right\rangle $, as follows:

Fig. 5.  (Color online) (a) Infinitely deep double-well model. $a$ is the width of the well, $L$ the width of the barrier between two wells, and ${E_{\text{B}}}$ the barrier height between two wells. For simplicity, the potential out of the double wells is set infinitely high due to Coulomb blockade effect. (b) The tunnel coupling ${t_{\text{c}}}$ as a function of barrier height ${E_{\text{B}}}$. The data points are calculated based on the device shown in Fig. 3(a). The curve is a fit to Eq. (22).
$$ {t_{\text{c}}} = \frac{{4{\rm{e}^{ - {q_0}L}}}}{{a{q_0}}}E_{\text{n}}^{{\text{(0)}}}, $$ (22)

where $ E_{\text{n}}^{{\text{(0)}}} \equiv \dfrac{{{\pi ^2}{\hbar ^2}}}{{2m{a^2}}} $ and $ {q_0} \equiv \sqrt {\dfrac{{2m({E_{\text{B}}} - E_{\text{n}}^{{\text{(0)}}})}}{{{\hbar ^2}}}} $.

Eq. (18) provides two independent electrical approaches to manipulate the exchange interaction in experiments as described in the following.

1. The tunnel coupling strength ${t_{\text{c}}}$ via the barrier gate voltage applied between two neighboring quantum dots. The exponential relationship between tunnel coupling strength ${t_{\text{c}}}$ and barrier height ${E_{\text{B}}}$ can be inspired from Eq. (22). Moreover, the simulation shows that ${E_{\text{B}}}$ is linearly proportional to the barrier gate voltage ${V_{\text{b}}}$, as shown in Fig. 4(c). Combining with Eq. (18), the facts implies that the exchange interaction $J$ is primarily an exponential function of ${V_{\text{b}}}$, as shown in Fig. 6(a).

Fig. 6.  (a) The exchange energy $J$ as a function of the barrier gate voltage ${V_{\text{b}}}$ as described in Eqs. (18) and (22). (b) The exchange energy $J$ as a function of detuning energy $\Delta \varepsilon $ as described in Eq. (18).

2. The detuning energy $\Delta \varepsilon $ via the plunge gate voltage applied on each quantum dot. Two plunge gate voltages can be adjusted to control the detuning energy $\Delta \varepsilon $, hence to manipulate the exchange interaction $J$ according to Eq. (18). The exchange interaction $J$ is shown as a function of the detuning energy $\Delta \varepsilon $ in Fig. 6(b).

Manipulating the gate voltages ${V_{{\text{g1}}}}$ and ${V_{{\text{g2}}}}$ also slightly alternates the barrier height ${E_{\text{B}}}$ by crosstalk effects. Equivalently, ${V_{\text{b}}}$ makes the crosstalk to the potential of double quantum dots, so that the linear relationship between ${E_{\text{B}}}$ and ${V_{\text{b}}}$ is accordingly masked. The crosstalk is routinely and efficiently compensated using virtual gates, which are specific linear combinations of physical gate voltages according to mutual capacitance matrix[55, 56]. The virtual gate can also help to make a sole change of the exchange interaction between target pair quantum dots without affecting the exchange interaction of other pair quantum dots. Several other methods have been formulated to explore the exchange interaction between quantum dots such as Hartree−Fock[57] and full interaction methods[58]. Magnetic fields and spin−orbit coupling also influence the exchange interaction, and can be found in details elsewhere[5963].

Electrons in multiple quantum dots can be in any available quantum states out of ground states. We routinely choose two lowest energy quantum states as the computational bases to define $\left| 0 \right\rangle $ and $\left| 1 \right\rangle $ states. An effective Hamiltonian of the two-level system can be obtained from the exact Hamiltonian by Schrieffer−Wolff (SW) unitary transformation which decouples the computational basis and the high-energy subspaces[50]. In 1998, Loss and DiVincenzo proposed a scheme to construct a physical qubit using the single spin state of electron in one single semiconductor quantum dot, known as Loss−Divincenzo (LD) qubit[64]. A focused local oscillating magnetic fields[65], however, is required in nanometer-scale devices to manipulate spin states and the power dissipation due to the generation of magnetic field is not compatible with cryogenic environments. One can use a high frequency electric field to control spin states with the assistance of either spin−orbit coupling effect or static slanting field of micromagnet to mitigate the technique challenges[66, 67]. Alternatively, the spin qubit can be realized in a quantum decoherence-free subspace[15] of spin states in multiple quantum dots, such as S−T0 qubit[68] and exchange-only qubit[69]. Since multiple coupled quantum dots are involved in the schemes, the exchange interaction between neighboring dots is indispensable to achieve the construction and the manipulation of quantum states.

One of the straightforward extensions of LD qubit is the S−T0 qubit constructed from double quantum dots with charge state (1, 1). The apparent advantage to use S−T0 qubit is the immunity to the global magnetic fluctuations[15]. The construction and operation of S−T0 qubit requires the exchange interaction and a magnetic field gradient across the quantum dots, as shown in Fig. 7(a). The Hamiltonian of the two-spin system is as follows:

Fig. 7.  (Color online) (a) The construction of S−T0 qubit. The exchange interaction $J$ between quantum dots, and a magnetic field gradient $\Delta B$ is required to build a S−T0 qubit with double quantum dots. (b) States with two electron spins. The states with color mapping of orange and cyan color are the computational states of S−T0 qubit. The gray color mapped states are the leakage states. (c) Two states of $\left| S \right\rangle $ and $\left| {{T_0}} \right\rangle $ form a two-level subspace. (d) Exchange coupling $J$ and a longitudinal magnetic-field gradient $\Delta B$ provide two orthogonal manipulation axes for S−T0 qubit operation.
$$ {\hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}} = \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{1}} + \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} = \frac{1}{2}\mu {B_1}\sigma _{\text{z}}^{\text{1}} + \frac{1}{2}\mu {B_2}\sigma _{\text{z}}^{\text{2}} + \frac{1}{4}J\left( {{{\boldsymbol{\sigma }}_1} \cdot {{\boldsymbol{\sigma }}_2} - 1} \right), $$ (23)
$$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{1}} \equiv \frac{1}{2}\mu {B_1}\sigma _{\text{z}}^{\text{1}} + \frac{1}{2}\mu {B_2}\sigma _{\text{z}}^{\text{2}}, $$ (24)
$$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} \equiv \frac{1}{4}J\left( {{{\boldsymbol{\sigma }}_1} \cdot {{\boldsymbol{\sigma }}_2} - 1} \right). $$ (25)

Here, $\mu = {g^*}{\mu _\rm{B}}$ is the magnetic moment of the electron in quantum dot, ${g^*}$ the effective g-factor which is related to spin−orbit interaction[70, 71] and keeps positive as an example, ${B_1}$(${B_2}$) the magnetic field strength in the quantum dot 1(2), $J$ the exchange interaction between electron spins in dots 1 and 2, ${{\boldsymbol{\sigma }}_1}$(${{\boldsymbol{\sigma }}_2}$) the Pauli operator of quantum dot 1(2), and $\sigma _{\text{z}}^{\text{1}}$ ($\sigma _{\text{z}}^{\text{2}}$) the electron spin Pauli-Z operator of quantum dot 1(2). The eigenstates and corresponding eigenenergies of the Hamiltonians $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{1}} $ and $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} $ are shown in Table 2.

Table 2.  The eigenstates and corresponding eigenenergies of Hamiltonians $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^1 $ and $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^2 $ with $\Delta B \equiv {B_2} - {B_1}$.
HamiltonianEigenstateEigenenergy
$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{1}} $$\left| { \uparrow \downarrow } \right\rangle $$ - \mu \Delta B/2$
$\left| { \downarrow \uparrow } \right\rangle $$\mu \Delta B/2$
$\left| { \uparrow \uparrow } \right\rangle $$\mu ({B_1} + {B_2})/2$
$\left| { \downarrow \downarrow } \right\rangle $$ - \mu ({B_1} + {B_2})/2$
$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} $$\left| S \right\rangle $$ - J$
$\left| {{T_0}} \right\rangle $$0$
$\left| { \uparrow \uparrow } \right\rangle $$0$
$\left| { \downarrow \downarrow } \right\rangle $$0$
DownLoad: CSV  | Show Table

The state subspaces used to define S−T0 qubit and leakage states are shown in Fig. 7(b), where ${S_{{\text{z12}}}}$ is the spin projection of two electron spins in the direction of external static magnetic field and ${S_{12}}$ is the total spin quantum number. When there is no magnetic field gradient (the Hamiltonian reduced to $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} $), the energy-gap $J$ protected two states $\left| S \right\rangle $ and $\left| {{T_0}} \right\rangle $ form a two-level system and encode $\left| 0 \right\rangle $ and $\left| 1 \right\rangle $, respectively, known as S−T0 qubit, as shown in Fig. 7(c). The applied magnetic field ${B_1}$ and ${B_2}$ separate the leakage states $\left| { \uparrow \uparrow } \right\rangle $ and $\left| { \downarrow \downarrow } \right\rangle $ from the computational state $\left| {{T_0}} \right\rangle $ to avoid quantum information leakage, as shown in Figs 7(b) and 7(c). We can use energy-selective initialization[72] to prepare the two electrons in $\left| S \right\rangle $ state for further quantum computation.

The $ \left| S \right\rangle $ and $ \left| {{T_0}} \right\rangle $ states can be set as the north and the south poles of the Bloch sphere, respectively. The quantum states $\left| { \uparrow \downarrow } \right\rangle $ and $\left| { \downarrow \uparrow } \right\rangle $ are thus indicated on the equatorial plane of the Bloch sphere, as shown in Fig. 7(d). For an arbitrary initial state of the system $\left| \psi \right\rangle = a\left| { \uparrow \downarrow } \right\rangle + b\left| { \downarrow \uparrow } \right\rangle =c\left| S \right\rangle + d\left| {{T_0}} \right\rangle $ ($a$, $b$, $c$, and $d$ are complex numbers and constrained by ${\left| a \right|^2} + {\left| b \right|^2} = 1$ and ${\left| c \right|^2} + {\left| d \right|^2} = 1$), the final state reaches at $\left| \psi \right\rangle = \left( {a{\rm{e}^{i\mu \Delta Bt/\hbar }}\left| { \uparrow \downarrow } \right\rangle + b\left| { \downarrow \uparrow } \right\rangle } \right)/\sqrt 2 $ after time $t$ under the magnetic field gradient $\Delta B$ (the global phase ${\rm{e}^{ - i\mu \Delta Bt/2\hbar }}$ is ignored). The evolution equals to the state-vector rotation around the Y-axis with precession frequency $\mu \Delta B/h$ on the Bloch sphere. On the other hand, the final state reaches at $ \left| \psi \right\rangle = \left( {c{\rm{e}^{iJt/\hbar }}\left| S \right\rangle + d\left| {{T_0}} \right\rangle } \right)/\sqrt 2 $ after time $t$ under the exchange interaction $J$. The evolution equals to the state-vector rotation around the Z-axis with precession frequency $J/h$ on the Bloch sphere. Consequently, the exchange interaction and the magnetic field gradient provide two independent manipulation axes for the S−T0 qubit operation. Universal single qubit operations can be achieved by implementing arbitrary rotation of state vector around Y and Z axes[73] , as shown in Fig. 7(d).

As for a measurement, we can project the charge state (1, 1) to (0, 2). The singlet state $ S\left( {1,1} \right) $ can transit to $ S\left( {0,2} \right) $ accompany with one electron spin-conservation tunneling from left to right dot. The triplet state $ {T_0}\left( {1,1} \right) $ fails in the transition either to energetically accessible $ S\left( {0,2} \right) $ state because of spin conservation or to energetically inaccessible $ {T_0}\left( {0,2} \right) $ state so that the transfer of electron is forbidden[74, 75]. The event of electron transfer can be monitored by a proximal charge sensor[54]. Accordingly, the final state can be distinguished in either $\left| S \right\rangle $ or $\left| {{T_0}} \right\rangle $ state.

The LD qubit requires a statistic magnetic field and a high frequency alternating electro-magnetic field as independent manipulation axes. On the other hand, the S−T0 qubit requires a large magnetic field gradient, which cannot be turned on/off on demand. The implementation of micromagnet and focused alternating electro-magnetic field in quantum chips is also technically challenging[17]. Exchange-only qubit constructed by the spins of three electrons, alternatively, demonstrates feasibility to achieve universal quantum computation only using the full-electrical exchange interaction between neighboring dots to mitigate the challenges[69, 76, 77, 78].

Fig. 8(a) shows three-electron spin system, where ${J_{12}}$ (${J_{23}}$) is the exchange interaction between neighboring quantum dots 1(2) and 2(3). The Hamiltonian of the three-spin system is shown as follows:

Fig. 8.  (Color online) (a) The construction of exchange-only qubit. Only the exchange interaction ${J_{12}}$ and ${J_{23}}$ between quantum dots are required to build the exchange-only qubit with linearly coupled triple quantum dots. (b) States with three electron spins. The states with color mapping of orange and cyan color are the computational states of exchange-only qubit. The gray color mapped states are the leakage states. (c) Two states of $\left| D \right\rangle $ and $\left| {D'} \right\rangle $ form a two-level subspace. $\left| D \right\rangle $ is the mixture of $\left| {{D_{ + 1/2}}} \right\rangle $ and $\left| {{D_{ - 1/2}}} \right\rangle $, $\left| {D'} \right\rangle $ is the mixture of $\left| {D_{ + 1/2}'} \right\rangle $ and $\left| {D_{ - 1/2}'} \right\rangle $. (d) The exchange coupling parameters of ${J_{12}}$ and ${J_{23}}$ provide two independent manipulation axes of the exchange-only qubit manipulation.
$$ {\hat H_{{\text{E}} - {\text{O}}}} = \hat H_{{\text{E}} - {\text{O}}}^1 + \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} = \frac{1}{4}{J_{12}}\left( {{{\boldsymbol{\sigma }}_1} \cdot {{\boldsymbol{\sigma }}_2} - 1} \right) + \frac{1}{4}{J_{23}}\left( {{{\boldsymbol{\sigma }}_2} \cdot {{\boldsymbol{\sigma }}_3} - 1} \right) , $$ (26)
$$ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} \equiv \frac{1}{4}{J_{12}}\left( {{{\boldsymbol{\sigma }}_1} \cdot {{\boldsymbol{\sigma }}_2} - 1} \right) , $$ (27)
$$ \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} \equiv \frac{1}{4}{J_{23}}\left( {{{\boldsymbol{\sigma }}_2} \cdot {{\boldsymbol{\sigma }}_3} - 1} \right) . $$ (28)

The eigenstates and corresponding eigenenergies of the Hamiltonians $ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ and $ \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} $ are shown in Table 3.

Where

Table 3.  The eigenstates and corresponding eigenenergies of Hamiltonian $ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ and $ \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} $[41].
Hamiltonian Eigenstate Eigenenergy
$ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ $\left| {{Q_{ + 3/2}}} \right\rangle $ 0
$ \left| {{Q_{ + 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 3/2}}} \right\rangle $ 0
$\left| {{{\bar D}_{ + 1/2}}} \right\rangle $ 0
$\left| {{{\bar D}_{ - 1/2}}} \right\rangle $ 0
$ \left| {{{\bar D}'}_{ + 1/2}} \right\rangle $ $ - {J_{12}}$
$ \left| {{{\bar D}'}_{ - 1/2}} \right\rangle $ $ - {J_{12}}$
$\hat H_{{\text{E}} - {\text{O}}}^{\text{2}}$ $\left| {{Q_{ + 3/2}}} \right\rangle $ 0
$ \left| {{Q_{ + 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 3/2}}} \right\rangle $ 0
$\left| {{D_{ + 1/2}}} \right\rangle $ 0
$\left| {{D_{ - 1/2}}} \right\rangle $ 0
$\left| {D_{ + 1/2}'} \right\rangle $ $ - {J_{23}}$
$\left| {D_{ - 1/2}'} \right\rangle $ $ - {J_{23}}$
DownLoad: CSV  | Show Table
$$ \begin{split} & \left| {{D_{ + 1/2}}} \right\rangle = \left( {\left| \uparrow \right\rangle \left| {{T_0}} \right\rangle - \sqrt 2 \left| \downarrow \right\rangle \left| {{T_ + }} \right\rangle } \right)/\sqrt 3 = \left( {\left| { \uparrow \uparrow \downarrow } \right\rangle + \left| { \uparrow \downarrow \uparrow } \right\rangle - 2\left| { \downarrow \uparrow \uparrow } \right\rangle } \right)/\sqrt 6 , \\ & \left| {{D_{ - 1/2}}} \right\rangle = \left( {\left| \downarrow \right\rangle \left| {{T_0}} \right\rangle - \sqrt 2 \left| \uparrow \right\rangle \left| {{T_ - }} \right\rangle } \right)/\sqrt 3 = \left( {\left| { \downarrow \downarrow \uparrow } \right\rangle + \left| { \downarrow \uparrow \downarrow } \right\rangle - 2\left| { \uparrow \downarrow \downarrow } \right\rangle } \right)/\sqrt 6 ,\\ & \left| {D_{ + 1/2}'} \right\rangle = \left| \uparrow \right\rangle \left| S \right\rangle = \left( {\left| { \uparrow \uparrow \downarrow } \right\rangle - \left| { \uparrow \downarrow \uparrow } \right\rangle } \right)/\sqrt 2 , \\ & \left| {D_{ - 1/2}'} \right\rangle = \left| \downarrow \right\rangle \left| S \right\rangle = \left( {\left| { \downarrow \downarrow \uparrow } \right\rangle - \left| { \downarrow \uparrow \downarrow } \right\rangle } \right)/\sqrt 2 . \end{split} $$ (29)
$$ \begin{split} & \left| {{{\bar D}_{ + 1/2}}} \right\rangle = \left( {\left| {{T_0}} \right\rangle \left| \uparrow \right\rangle - \sqrt 2 \left| {{T_ + }} \right\rangle \left| \downarrow \right\rangle } \right)/\sqrt 3 = \left( {\left| { \downarrow \uparrow \uparrow } \right\rangle + \left| { \uparrow \downarrow \uparrow } \right\rangle - 2\left| { \uparrow \uparrow \downarrow } \right\rangle } \right)/\sqrt 6 , \\ & \left| {{{\bar D}_{ - 1/2}}} \right\rangle = \left( {\left| {{T_0}} \right\rangle \left| \downarrow \right\rangle - \sqrt 2 \left| {{T_ - }} \right\rangle \left| \uparrow \right\rangle } \right)/\sqrt 3 = \left( {\left| { \uparrow \downarrow \downarrow } \right\rangle + \left| { \downarrow \uparrow \downarrow } \right\rangle - 2\left| { \downarrow \downarrow \uparrow } \right\rangle } \right)/\sqrt 6 , \\ & \left| {\bar D_{ + 1/2}'} \right\rangle = \left| S \right\rangle \left| \uparrow \right\rangle = \left( {\left| { \uparrow \downarrow \uparrow } \right\rangle - \left| { \downarrow \uparrow \uparrow } \right\rangle } \right)/\sqrt 2 , \\ & \left| {\bar D_{ - 1/2}'} \right\rangle = \left| S \right\rangle \left| \downarrow \right\rangle = \left( {\left| { \downarrow \uparrow \downarrow } \right\rangle - \left| { \uparrow \downarrow \downarrow } \right\rangle } \right)/\sqrt 2 . \end{split} $$ (30)

And

$$ \begin{gathered} \left| {{Q_{ + 3/2}}} \right\rangle = \left| { \uparrow \uparrow \uparrow } \right\rangle , \\ \left| {{Q_{ + 1/2}}} \right\rangle = \left( {\left| { \uparrow \uparrow \downarrow } \right\rangle + \left| { \uparrow \downarrow \uparrow } \right\rangle + \left| { \downarrow \uparrow \uparrow } \right\rangle } \right)/\sqrt 3 , \\ \left| {{Q_{ - 1/2}}} \right\rangle = \left( {\left| { \downarrow \downarrow \uparrow } \right\rangle + \left| { \downarrow \uparrow \downarrow } \right\rangle + \left| { \uparrow \downarrow \downarrow } \right\rangle } \right)/\sqrt 3 , \\ \left| {{Q_{ - 3/2}}} \right\rangle = \left| { \downarrow \downarrow \downarrow } \right\rangle . \\ \end{gathered} $$ (31)

The state subspaces of the three electron spins used to define exchange-only qubit and the leakage states are shown in Fig. 8(b), where ${S_{{\text{z123}}}}$ is the spin projection of three electron spins in an arbitrary direction, ${S_{23}}$ the total spin of the electrons confined in the rightmost two quantum dots, ${S_{123}}$ the total spin quantum number. At the limit of ${J_{12}}/{J_{23}} \to 0$ (the Hamiltonian reduced to $ \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} $), the exchange interaction ${J_{23}}$ provides an energy spacing between $\left| D \right\rangle $ and $\left| {D'} \right\rangle $ as shown in Fig. 8(c) and Table 3, where $\left| D \right\rangle $ is the mixture of $\left| {{D_{ + 1/2}}} \right\rangle $ and $\left| {{D_{ - 1/2}}} \right\rangle $, and $\left| {D'} \right\rangle $ is the mixture of $ \left| {D_{ + 1/2}'} \right\rangle $ and $ \left| {D_{ - 1/2}'} \right\rangle $[69, 77]. The two states of $\left| D \right\rangle $ and $\left| {D'} \right\rangle $ form a two-level system can be defined as $\left| 0 \right\rangle $ and $\left| 1 \right\rangle $, respectively, known as exchange-only qubit. We can first use energy-selective initialization to prepare the last two electrons in a spin singlet state, which lets ${S_{23}} = 0$. As the spin state of the first electron remains random, the three-electron system is initialized in $\left| {D'} \right\rangle $ which gives ${S_{123}} = 1/2$. The exchange interactions ${J_{12}}$ and ${J_{23}}$ keep the quantum number ${S_{123}}$ unchanged and allow the computational states within the ${S_{123}} = 1/2$ subspace[77].

We choose $\left| {D'} \right\rangle $ and $\left| D \right\rangle $ as the north and the south poles of Bloch sphere. The eigenstates $\left| {\bar D} \right\rangle $ and $ \left| {\bar D'} \right\rangle $ of $ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ are also indicated on the Bloch sphere, where $\left| {\bar D} \right\rangle $ is the mixture of $ \left| {{{\bar D}_{ + 1/2}}} \right\rangle $ and $ \left| {{{\bar D}_{ - 1/2}}} \right\rangle $, and $\left| {\bar D'} \right\rangle $ is the mixture of $ \left| {\bar D_{ + 1/2}'} \right\rangle $ and $ \left| {\bar D_{ - 1/2}'} \right\rangle $, as shown in Fig. 8(c). According to Eqs. (29) and (30), the angle between the state vector of $\left| {D'} \right\rangle $ and the state vector of $ \left| {{\bar D}''} \right\rangle $ is 120°, as shown in Fig. 8(d).

Referring to the S−T0 qubit, we can rotate the state vector separately around the blue-color axis with the precession frequency of ${J_{12}}/h$, and around the red-color axis with the precession frequency of ${J_{23}}/h$, as shown in Fig. 8(d). This implies that we can achieve arbitrary single qubit operations by implementing the rotations of Bloch vector around the two axes. Therefore, the two independent exchange interactions between neighboring quantum dots are sufficient to achieve universal single qubit operations[13, 15].

Like that of S−T0 qubit, the readout of exchange-only qubit relies on Pauli spin blockade of two neighboring dots. The right most two spins form either the singlet state out of $ \left| {D'} \right\rangle $ state or the triplet state out of the $ \left| D \right\rangle $ state, as shown in Fig. 9. We can project the triple-quantum-dot state from charge state (1, 1, 1) to (1, 0, 2) after qubit manipulations. The detection of either singlet or triplet state from a proximal charge sensor reveals the final state of either $ \left| {D'} \right\rangle $ or $ \left| D \right\rangle $ state, respectively[41].

Fig. 9.  (Color online) The spin configuration of $ \left| {D'} \right\rangle $ and $ \left| D \right\rangle $. ${S_1}$ is the spin quantum number of electron in the first quantum dot. ${S_{23}}$ is the total spin quantum number of two electrons, which makes the second and the third quantum dot singly occupied.

The crucial challenge of exchange-only qubit is to keep the leakage states out of the computational states because the hyperfine interactions from residual nuclear spins fluctuate the total spin and leak the qubit information into ${S_{123}} = 3/2$ states[77]. An external magnetic field $ B $ can be applied additionally to circumvent the leakage to the quantum states $\left| {{Q_{ \pm 3/2}}} \right\rangle $[78, 79].

The most important application of exchange interaction is the implementation of quantum gate operations between quantum dot qubits[80, 81]. In this section, we take SWAP gate[72, 82] and CNOT gate[83, 84] for examples to explore the role of exchange interaction in achieving two-qubit logical gates.

The SWAP gate is subject to exchange the quantum states of two qubits[85] and illustrated in a quantum circuit, as shown in Fig. 10(a). $\left| a \right\rangle $ and $\left| b \right\rangle $ refer to arbitrary quantum states. For LD qubits, we can set spin up as $\left| 1 \right\rangle $ and spin down as $\left| 0 \right\rangle $. The SWAP gate on two LD qubits can be achieved by manipulating the exchange interaction between them[86]. Taking the schematic diagram of two quantum dots in Fig. 10(b) for an example, we can write the Hamiltonian of the system when the exchange interaction $J$ is on as follows:

Fig. 10.  (Color online) (a) The quantum circuit of SWAP gate. (b) Double quantum dots with one single electron in each dot. ${{\boldsymbol{S}}_{1}}$(${{\boldsymbol{S}}_{2}}$) is the spin of electron in left (right) quantum dot. The exchange interaction $J$ is required to obtain the SWAP gate in specific duration.
$$ {\hat H_{{\text{SWAP}}}} = J(t)\left( {{{{\boldsymbol{\hat S}}}_1} \cdot {{{\boldsymbol{\hat S}}}_2} - \frac{{{\hbar ^2}}}{4}} \right)/{\hbar ^2}. $$ (32)

The state subspace required for achieving the SWAP gate is shown in Table 4.

Table 4.  The eigenstates and corresponding eigenenergies of Hamiltonian $ {\hat H_{{\text{SWAP}}}}. $
HamiltonianEigenstateEigenenergy
$ {\hat H_{{\text{SWAP}}}} $$\left| S \right\rangle $$ - J$
$\left| {{T_0}} \right\rangle $$0$
$\left| { \uparrow \uparrow } \right\rangle $$0$
$\left| { \downarrow \downarrow } \right\rangle $$0$
DownLoad: CSV  | Show Table

The initial state of the system, setting at $\left| { \uparrow \downarrow } \right\rangle =( \left| S \right\rangle + \left| {{T_0}} \right\rangle )/\sqrt 2 $ state, as shown in the upper panel of Fig. 10(b) evolves into $\left( {{e^{i\int_0^t {J(t')\rm{d}t'} /\hbar }}\left| S \right\rangle + \left| {{T_0}} \right\rangle } \right)/\sqrt 2 $ after time $t$. If $t$ satisfies $ \int_0^t {J(t')\rm{d}t'} /\hbar = \pi ,3\pi ,5\pi ... $, the $\left| { \downarrow \uparrow } \right\rangle = \left( {\left| {{T_0}} \right\rangle - \left| S \right\rangle } \right)/\sqrt 2 $ state succeed after the spin SWAP gate, as shown in the bottom panel of Fig. 10(b). $\left| { \uparrow \uparrow } \right\rangle $ and $\left| { \downarrow \downarrow } \right\rangle $ are eigenstates of $ {\hat H_{{\text{SWAP}}}} $, so the swap of these states is trivial.

There is no requirement of magnetic field in the SWAP gate operation. The presence of magnetic field gradient (such as nuclear spin field gradient), on the contrary, compromises the fidelity of the SWAP gate. Therefore, Ⅳ-group material platforms could favor a higher-fidelity SWAP gate operation due to the naturally less nonzero nuclear spins[82].

The quantum circuit representation of CNOT gate is shown in Fig. 11(a), in which qubit 1 with state $\left| a \right\rangle $ acts as control bit and qubit 2 with state $\left| b \right\rangle $ acts as target bit. The target qubit sustains if the control qubit is set to $\left| 0 \right\rangle $, whereas the target bit is flipped. The gate action is, thus, expressed as $\left| {a,b} \right\rangle \to \left| {a,b \oplus a} \right\rangle $ in formal logic[85]. Similar to the definition in 3.2.1, we can set spin up as $\left| 1 \right\rangle $ and spin down as $\left| 0 \right\rangle $.

Fig. 11.  (Color online) (a) The quantum circuit of CNOT gate. (b) Double quantum dots with one single electron in each dot. ${{\boldsymbol{S}}_1}$(${{\boldsymbol{S}}_2}$) is the spin of electron in left (right) quantum dot. The exchange interaction $J$ between two quantum dots, and a magnetic field gradient $\Delta {\boldsymbol{B}}$ are required to achieve the CNOT gate operation. (c) Schematic energy level diagram adopted from Table 5. (d) The spin of electron in the left quantum dot is flipped with absorbing a photo from an alternating magnetic field at frequency of $f_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{L}}$ when the spin of electron in the right quantum dot is $\left| \uparrow \right\rangle $.

In addition to the exchange interaction, a magnetic field gradient is required to implement the CNOT gate operation in the double quantum dots, as shown in Fig. 11(b). The Hamiltonian of the double dots is close to $ {\hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}} $ and can be expressed as follows:

$$ {\hat H_{{\text{CNOT}}}} = \mu {B_1}S_{\text{z}}^{\text{1}}/\hbar + \mu {B_2}S_{\text{z}}^{\text{2}}/\hbar + J\left( {{{{\boldsymbol{\hat S}}}_1} \cdot {{{\boldsymbol{\hat S}}}_2} - \frac{{{\hbar ^2}}}{4}} \right)/{\hbar ^2}. $$ (33)

${{\boldsymbol{\hat S}}_1}$(${{\boldsymbol{\hat S}}_2}$) is the spin operator of electron spin in quantum dot 1(2). $ S_{\text{z}}^{\text{1}} $ ($ S_{\text{z}}^{\text{2}} $) is the Z-axis component of electron spin operator in quantum dot 1(2).

The state subspace required to achieve CNOT gate is shown in Table 5.

Table 5.  The eigenstates and corresponding eigenenergies of the Hamiltonian ${\hat H_{{\text{CNOT}}}}.$
HamiltonianEigenstateEigenenergy
$ {\hat H_{{\text{CNOT}}}} $$ \left| { \uparrow \uparrow } \right\rangle $$\mu ({B_1} + {B_2})/2$
$ \left| { \Downarrow \Uparrow } \right\rangle $$\left[ { - J + \sqrt {{J^2} + {{\left( {g\Delta B\mu } \right)}^2}} } \right]/2$
$ \left| { \Uparrow \Downarrow } \right\rangle $$\left[ { - J - \sqrt {{J^2} + {{\left( {g\Delta B\mu } \right)}^2}} } \right]/2$
$ \left| { \downarrow \downarrow } \right\rangle $$ - \mu ({B_1} + {B_2})/2$
DownLoad: CSV  | Show Table

The presence of exchange interaction makes a hybridization between $ \left| { \downarrow \uparrow } \right\rangle $ and $ \left| { \uparrow \downarrow } \right\rangle $ to generate $ \left| { \Downarrow \Uparrow } \right\rangle $ and $ \left| { \Uparrow \Downarrow } \right\rangle $ states as the eigenstates. The hybridization states $ \left| { \Downarrow \Uparrow } \right\rangle $ and $ \left| { \Uparrow \Downarrow } \right\rangle $ collapse to $ \left| { \downarrow \uparrow } \right\rangle $ and $ \left| { \uparrow \downarrow } \right\rangle $, respectively, at the $ \Delta B\gg J $ limit[83].

The energy diagram of CNOT gate subspace is shown in Fig. 11(c), where $ hf_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\;\text{L}} $is energy difference between spin states $\left| \uparrow \right\rangle $ and $\left| \downarrow \right\rangle $ in the left dot with the right spin at state $\left| \uparrow \right\rangle $. So as to $ hf_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \downarrow \right\rangle }^{\;\text{L}} $, $ hf_{\left| {{\psi _{\text{L}}}} \right\rangle = \left| \uparrow \right\rangle }^{\;\text{R}} $, and $ hf_{\left| {{\psi _{\text{L}}}} \right\rangle = \left| \downarrow \right\rangle }^{\;\text{R}} $. The shift of single electron resonance frequency relies on another state and precisely matches the exchange splitting as follows:

$$ hf_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{L}} - hf_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \downarrow \right\rangle }^{\text{L}} = hf_{\left| {{\psi _{\text{L}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{R}} - hf_{\left| {{\psi _{\text{L}}}} \right\rangle = \left| \downarrow \right\rangle }^{\text{R}} = J. $$ (34)

The spin of right quantum dot is set as the control bit and the spin of left quantum dot as the target bit. The electron spin of left quantum dot is flipped by an alternating magnetic field of frequency $f_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{L}}$ when the electron spin of right quantum dot is $\left| \uparrow \right\rangle $, as shown in Fig. 11(d). The electron spin of left quantum dot perseveres without absorbing a photo from the alternating magnetic field of frequency $f_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{L}}$ if the electron spin of right quantum dot is $\left| \downarrow \right\rangle $. The CNOT gate based on two LD qubits is, thus, implemented[83].

Quantum state transmission is a crucial technique for quantum computation and quantum communication[87]. The tunable tunnel barrier is capable of adjusting the exchange interaction between semiconductor quantum dots, allowing an adiabatic spin transport protocol known as spin-coherent transport through adiabatic passage (spin-CTAP)[88, 89], as shown in Fig. 12(a). Qubit 1 is initially isolated and prepared in an arbitrary quantum state, whereas qubits 2 and 3 are coupled and prepared in a singlet state (indicated in dotted square). The exchange interaction between qubits 1 and 2 (2 and 3) is then turned on (off) adiabatically. The quantum state of qubit 1 is transmitted to qubit 3 when the evolution is complete. Spin states can be transferred between distant dots by adiabatic modulation without motion of the electrons, known as adiabatic quantum teleportation, too[90].

Fig. 12.  (Color online) (a) Spin-coherent transport through adiabatic passage. (b) Long-range coupler of spins to transfer an arbitrary spin from the left most Alice to the right most Bob[91]. (c) A representation of triple quantum dots for simulation of interaction-driven Mott metal−insulator transition. (d) Energy spectrum of low spin state and ferromagnetic state as a function of tunnel coupling strength, where $S$ is the total spin number of the three electrons.

An effective long-range exchange coupling can be established between distant spins (also known as super-exchange). In 2003, Bose proposed a spin chain as a long-range coupler of spins, known as the spin bus[91], which was developed later by Friesen et al.[92]. This scheme involves intermediate quantum dot chains as the medium (multi-QD mediator)[21] to transfer electron spins from the sender "Alice" to the receiver "Bob", as shown in Fig. 12(b). The transfer of electron spin states has been implemented in quadruple dot system in GaAs in recent works[93, 94].

Although large-scale quantum computation takes long way to succeed[95], analog quantum simulations have great chance to be performed based on finite number of spin qubits[14, 15]. One outstanding feature of gate defined quantum dot system is that the tunnel barriers can be tuned precisely, allowing analysis of the system characteristics at wide range of coupling strengths. This feature can be used to explore Fermi−Hubbard model[96, 97]. The tunnel coupling between semiconductor quantum dots can be enhanced to favor a larger quantum dot behaving like a small metallic island. This is the finite-size analogue of the interaction-driven Mott metal−insulator transition[98, 99], as shown in Fig. 12(c). The currently available systems present more interesting possibilities. One instance is the simulation of Nagaoka ferromagnets[100] in looped quadruple quantum dots[101]. The simulation requires precise control of tunnel coupling between neighboring quantum dots to change the energy of ferromagnetic state with respect to the low spin state. When the tunnel coupling is much smaller than the Coulomb repulsion, the experiment has demonstrated the emergence of ferromagnetic state with three electron spins in the looped quadruple quantum dots, as shown in Fig. 12(d). The experiment, thus, verified Nagaoka's 50-year-old theory.

The ultracold atom system, a counterpart to semiconductor quantum dot system, relies on the exchange interaction and has achieved more remarkable achievements in quantum simulation[102]. For example, disorder-induced localization (Anderson localization) of matter waves[103] and Mott-insulator phase of the Bose−Hubbard model[104] have been simulated in one-dimensional ultracold atom systems. In two-dimensional ultracold atom array, quantum Monte Carlo simulation has also been preliminarily realized[105]. These achievements owe much to the precise manipulation of the interactions between ultracold atoms. With the scaling-up of the quantum dot array and the improvement of electrical control methods in the exchange interactions between dots, such exotic quantum simulations could also be implemented in a quantum dot array.

The exchange interaction between semiconductor quantum dots plays an important role in implementing single qubit manipulations, two-qubit gate operations, quantum communication and quantum simulations. The review have explored the physical principle of exchange interaction between semiconductor quantum dots based on Heitler−London and Hund−Mulliken approaches. The electrical control method of exchange interaction is then proposed based on Hubbard and constant interaction models. The role of exchange interaction in construction of semiconductor quantum dot based spin qubits and in implementation of two-qubit gate operations is intensively surveyed. The electrical control of exchange interaction is at the heart of succeeding more complex quantum simulations and large-scale quantum computations.

Thanks to Ms. Yujiao Ma of Peking University for her fruitful discussion and technical support to the manuscript preparation. This research was funded by National Natural Science Foundation of China, (Grant Nos. 11974030 and 92165208). We fabricated the devices with the assistance of Peking Nanofab.

Supplementary materials to this article can be found online at https://doi.org/10.1088/1674-4926/24050043.



[1]
Bhat H A, Khanday F A, Kaushik B K, et al. Quantum computing: fundamentals, implementations and applications. IEEE Open J Nanotechnology, 2022, 3, 61 doi: 10.1109/OJNANO.2022.3178545
[2]
DiVincenzo D P. The physical implementation of quantum computation. Fortschritte der Physik, 2000, 48(9-11), 771 doi: 10.1002/1521-3978(200009)48:9/11
[3]
Georgescu I. The DiVincenzo criteria 20 years on. Nat Rev Phys, 2020, 2, 666 doi: 10.1038/s42254-020-00256-4
[4]
Ladd T, Jelezko F, Laflamme R, et al. Quantum computers. Nature, 2010, 464(4), 45 doi: 10.1038/nature08812
[5]
Wei S J, Li H, Long G L. A full quantum eigensolver for quantum chemistry simulations. Research, 2020, 1, 111 doi: 10.34133/2020/1486935
[6]
Shor P W. Algorithms for quantum computation: discrete logarithms and factoring. 35th Annual Symposium on Foundations of Computer Science, 1994, 124 doi: 10.1109/SFCS.1994.365700
[7]
Grover L K. Quantum mechanics helps in searching for a needle in a haystack. Phys Rev Lett, 1997, 79(2), 325 doi: 10.1103/PhysRevLett.79.325
[8]
Hanson R, Kouwenhoven L P, Petta J R, et al. Spins in few-electron quantum dots. Rev Mod Phys, 2007, 79(4), 1217 doi: 10.1103/RevModPhys.79.1217
[9]
Kloeffel C, Loss D. Prospects for spin-based quantum computing in quantum dots. Annu Rev Condens Matter Phys, 2013, 4(1), 51 doi: 10.1146/annurev-conmatphys-030212-184248
[10]
Zhang X, Li H O, Wang K, et al. Qubits based on semiconductor quantum dots. Chin Phys B, 2018, 27(2), 020305 doi: 10.1088/1674-1056/27/2/020305
[11]
Chatterjee A, Stevenson P, Franceschi S, et al. Semiconductor qubits in practice. Nat Rev Phys, 2021, 3(3), 157 doi: 10.1038/s42254-021-00283-9
[12]
Wang N, Wang B C, Guo G P. New progress of silicon-based semiconductor quantum computation. Acta Phys Sin, 2022, 71(23), 230301 doi: 10.7498/aps.71.20221900
[13]
Zhang X, Li H O, Cao G, et al. Semiconductor quantum computation. Natl Sci Rev, 2019, 6(1), 32 doi: 10.1093/nsr/nwy153
[14]
Barthelemy P, Vandersypen L M K. Quantum dot systems: a versatile platform for quantum simulations. Ann Phys, 2013, 525(10-11), 808 doi: 10.1002/andp.201300124
[15]
Burkard G, Ladd T D, Pan A, et al. Semiconductor spin qubits. Rev Mod Phys, 2023, 95(2), 025003 doi: 10.1103/RevModPhys.95.025003
[16]
Mu J W, Huang S Y, Liu Z H, et al. A highly tunable quadruple quantum dot in a narrow bandgap semiconductor InAs nanowire. Nanoscale, 2021, 13, 3983 doi: 10.1039/D0NR08655J
[17]
Zwerver A M J, Krähenmann T, Watson T F, et al. Qubits made by advanced semiconductor manufacturing. Nature Electronics, 2022, 5, 184 doi: 10.1038/s41928-022-00727-9
[18]
Pomorski K, Giounanlis P, Blokhina E, et al. Analytic view on coupled single-electron lines. Semicond Sci Technol, 2019, 34, 125015 doi: 10.1088/1361-6641/ab4f40
[19]
Lawrie W I L, Eenink H G J, Hendrickx N W, et al. Quantum dot arrays in silicon and germanium. Appl Phys Lett, 2020, 116, 080501 doi: 10.1063/5.0002013
[20]
Sánchez R, Gallego Marcos F, Platero G. Superexchange blockade in triple quantum dots. Phys Rev B, 2014, 89, 161402(R doi: 10.1103/PhysRevB.89.161402
[21]
Qiao H, Kandel Y P, Fallahi S, et al. Long-distance superexchange between semiconductor quantum-dot electron spins. Phys Rev Lett, 2021, 126, 017701 doi: 10.1103/PhysRevLett.126.017701
[22]
Kouwenhoven L P, Austing D G, Tarucha S, et al. Few-electron quantum dots. Rep Prog Phys, 2001, 64, 701 doi: 10.1088/0034-4885/64/6/201
[23]
Reimann S M, Manninen M. Electronic structure of quantum dots. Rev Mod Phys, 2002, 74, 1283 doi: 10.1103/RevModPhys.74.1283
[24]
Feng D, Jin G J. Condensed matter physics. Beijing: Higher Education Press, 2013, 1, 1 (in Chinese)
[25]
Heitler W, London F. Wechselwirkung neutraler atome und homöopolare bindung nach der quantenmechanik. Zeitschrift für Physik, 1927, 44, 455 (in German doi: 10.1007/BF01397394
[26]
Hatano T, Amaha S, Kubo T, et al. Manipulation of exchange coupling energy in a few-electron double quantum dot. Phys Rev B, 2008, 77(24), 241301 doi: 10.1103/PhysRevB.77.241301
[27]
Heisenberg V W. Mehrkörperproblem und resonanz in der quantenmechanik. Zeitschrift für Physik, 1926, 38, 411 (in German
[28]
Dirac P A. On the Theory of quantum mechanics. The Royal society, 1926, 661
[29]
Mullin W J, Blaylock G. Quantum statistics: Is there an effective fermion repulsion or boson attraction? Am J Phys, 2003, 71, 1223 doi: 10.1119/1.1590658
[30]
Wannier G H. Statistical physics. New York: Wiley, 1987
[31]
Leighton R B. Principles of modern physics. New York: McGraw−Hill Book Company, 1959
[32]
Griffiths D J, Schroeter D F. Introduction to quantum mechanics (third edition). Cambridge: Cambridge university Press, 2018
[33]
Zeng J Y. Quantum mechanics (fifth edition). Beijing: Science Press, 2013, 1, 1 (in Chinese)
[34]
Van der Wiel W G, Stopa M, Kodera T, et al. Semiconductor quantum dots for electron spin qubits. New J Phys, 2006, 8, 28 doi: 10.1088/1367-2630/8/2/028
[35]
Saraiva A L, Calderón M J, Koiller B. Reliability of the Heitler−London approach for the exchange coupling between electrons in semiconductor nanostructures. Phys Rev B, 2007, 233302 doi: 10.1103/PhysRevB.76.233302
[36]
Liu E K. Semiconductor physics (eighth edition). Beijing: Publishing House of Electronics Industry, 2023 (in Chinese)
[37]
Burkard G, Loss D, DiVincenzo D P. Coupled quantum dots as quantum gates. Phys Rev B, 1999, 59(3), 2070 doi: 10.1103/PhysRevB.59.2070
[38]
DiVincenzo D P, Loss D. Quantum information is physical. Superlattices and Microstructures, 1998, 23(3), 420 doi: 10.1006/spmi.1997.0520
[39]
Fanciulli M. Electron spin resonance and related phenomena in low dimensional structures. Berlin: Springer, 2009
[40]
Tokura Y, Austing D G, Tarucha S. Single-electron tunnelling in two vertically coupled quantum dots. J Phys Condens Matter, 1999, 11, 6023 doi: 10.1088/0953-8984/11/31/310
[41]
Laird E A, Taylor J M, DiVincenzo D P, et al. Coherent spin manipulation in an exchange-only qubit. Phys Rev B, 2010, 82, 075403 doi: 10.1103/PhysRevB.82.075403
[42]
Elzerman J M, Hanson R, Greidanus J S, et al. Few-electron quantum dot circuit with integrated charge read out. Phys Rev B, 2003, 67, 161308(R doi: 10.1103/PhysRevB.67.161308
[43]
Russ M, Burkard G. Three-electron spin qubits. J Phys Condens Matter, 2017, 29, 393001 doi: 10.1088/1361-648X/aa761f
[44]
Hubbard J. Electron correlations in narrow energy bands. Proc R Soc Lond A, 1963, 276(1365), 238 doi: 10.1098/rspa.1963.0204
[45]
Van der Wiel W G, Franceschi S De, Elzerman J M, et al. Electron transport through double quantum dots. Rev Mod Phys, 2003, 75(1), 1 doi: 10.1103/RevModPhys.75.1
[46]
Spałek J. Theory of unconventional superconductivity in strongly correlated systems: real space pairing and statistically consistent mean-field theory- in perspective. Acta Phys Pol A, 2012, 121(4), 764 doi: 10.12693/APhysPolA.121.764
[47]
Taylor J M, Petta J R, Johnson A C, et al. Relaxation, dephasing, and quantum control of electron spins in double quantum dots. Phys Rev B, 2007, 76(3), 035315 doi: 10.1103/PhysRevB.76.035315
[48]
Balachandran A P, Ercolessi E, Morandi G, et al. Hubbard model and anyon superconductivity: a review. Int J Mod Phys B, 1990, 4(14), 2057 doi: 10.1142/S0217979290001030
[49]
Spałek J. Effect of pair hopping and magnitude of intra-atomic interaction on exchange-mediated superconductivity. Phys Rev B, 1988, 37(1), 533 doi: 10.1103/PhysRevB.37.533
[50]
Bravyi S, DiVincenzo D P, Loss D. Schrieffer–Wolff transformation for quantum many-body systems. Ann Phys, 2011, 326(10), 2793 doi: 10.1016/j.aop.2011.06.004
[51]
Assa A. Interacting electrons and quantum magnetism. Berlin: Springer-Verlag, 1998
[52]
DiCarlo L, Lynch H J, Johnson A C, et al. Differential charge sensing and charge delocalization in a tunable double quantum dot. Phys Rev Lett, 2004, 92(22), 226801 doi: 10.1103/PhysRevLett.92.226801
[53]
Hu Y J, Churchill H O H, Reilly D J, et al. A Ge/Si heterostructure nanowire-based double quantum dot with integrated charge sensor. Nat Nanotechnol, 2007, 2, 622 doi: 10.1038/nnano.2007.302
[54]
Wang X M, Huang S Y, Wang J Y, et al. A charge sensor integration to tunable double quantum dots on two neighboring InAs nanowires. Nanoscale, 2021, 13, 1048 doi: 10.1039/D0NR07115C
[55]
Hsiao T K, van Diepen C J, Mukhopadhyay U, et al. Efficient orthogonal control of tunnel couplings in a quantum dot array. Phys Rev Appl, 2020, 13(5), 054018 doi: 10.1103/PhysRevApplied.13.054018
[56]
Mills A R, Zajac D M, Gullans M J. et al. Shuttling a single charge across a one-dimensional array of silicon quantum dots. Nat Commun, 2019, 10, 1063 doi: 10.1038/s41467-019-08970-z
[57]
Hu X D, Sarma S D. Hilbert-space structure of a solid-state quantum computer: two-electron states of a double-quantum-dot artificial molecule. Phys Rev A, 2000, 61(6), 062301 doi: 10.1103/PhysRevA.61.062301
[58]
Ercan H E, Coppersmith S N, Friesen M. Strong electron−electron interactions in Si/SiGe quantum dots. Phys Rev B, 2021, 104, 235302 doi: 10.1103/PhysRevB.104.235302
[59]
Bellucci D, Rontani M, Troiani F, et al. Competing mechanisms for singlet−triplet transition in artificial molecules. Phys Rev B, 2004, 69, 201308 doi: 10.1103/PhysRevB.69.201308
[60]
Giavaras G, Tokura Y. Probing the singlet−triplet splitting in double quantum dots: implications of the ac field amplitude. Phys Rev B, 2019, 100, 195421 doi: 10.1103/PhysRevB.100.195421
[61]
Stopa M, Marcus C M. Magnetic field control of exchange and noise immunity in double quantum dots. Nano Lett, 2008, 8(6), 1778 doi: 10.1021/nl801282t
[62]
Geyer S, Hetényi B, Bosco S, et al. Anisotropic exchange interaction of two hole spin qubits. Nat Phys, 2024 doi: 10.1038/s41567-024-02481-5
[63]
Liu Z H, Entin-Wohlman O, Aharony A, et al. Control of the two-electron exchange interaction in a nanowire double quantum dot. Phys Rev B, 2018, 98, 241303 doi: 10.1103/PhysRevB.98.241303
[64]
Loss D, DiVincenzo D P. Quantum computation with quantum dots. Phys Rev A, 1998, 57(1), 1050 doi: 10.1103/PhysRevA.57.120
[65]
Koppens F H L, Buizert C, Tielrooij K J, et al. Driven coherent oscillations of a single electron spin in a quantum dot. Nature, 2006, 442, 766 doi: 10.1038/nature05065
[66]
Liu H, Wang K, Gao F, et al. Ultrafast and electrically tunable Rabi frequency in a Germanium hut wire hole spin qubit. Nano Lett, 2023, 23(9), 3810 doi: 10.1021/acs.nanolett.3c00213
[67]
Philips S G J, Mądzik M T, Amitonov S V, et al. Universal control of a six-qubit quantum processor in silicon. Nature, 2022, 609, 919 doi: 10.1038/s41586-022-05117-x
[68]
Levy J. Universal quantum computation with spin-1/2 pairs and Heisenberg exchange. Phys Rev Lett, 2002, 89, 147902 doi: 10.1103/PhysRevLett.89.147902
[69]
DiVincenzo D, Bacon D, Kempe J, et al. Universal quantum computation with the exchange interaction. Nature, 2000, 408, 339 doi: 10.1038/35042541
[70]
Zhang T, Liu H, Cao F, et al. Anisotropic g-factor and spin−orbit field in a germanium hut wire double quantum dot. Nano Lett, 2021, 21(9), 3835 doi: 10.1021/acs.nanolett.1c00263
[71]
Wang J Y, Huang S Y, Lei Z J, et al. Measurements of the spin−orbit interaction and Landé g factor in a pure-phase InAs nanowire double quantum dot in the Pauli spin-blockade regime. Appl Phys Lett, 2016, 109(5), 053106 doi: 10.1063/1.4960464
[72]
Petta J R, Johnson A C, Taylor J M, et al. Coherent manipulation of coupled electron spins in semiconductor quantum dots. Science, 2005, 309(5744), 2180 doi: 10.1126/science.1116955
[73]
Wardrop M P, Doherty A C. Exchange-based two-qubit gate for singlet−triplet qubits. Phys Rev B, 2014, 90, 045418 doi: 10.1103/PhysRevB.90.045418
[74]
Johnson A C, Petta J R, Taylor J M, et al. Triplet–singlet spin relaxation via nuclei in a double quantum dot. Nature, 2005, 435, 925 doi: 10.1038/nature03815
[75]
Prance J R, Shi Z, Simmons C B, et al. Single-shot measurement of triplet−singlet relaxation in a Si/SiGe double quantum dot. Phys Rev Lett, 2012, 108, 046808 doi: 10.1103/PhysRevLett.108.046808
[76]
Weinstein A J, Reed M D, Jones A M, et al. Universal logic with encoded spin qubits in silicon. Nature, 2023, 615, 817 doi: 10.1038/s41586-023-05777-3
[77]
Andrews R W, Jones C, Reed M D, et al. Quantifying error and leakage in an encoded Si/SiGe triple-dot qubit. Nat Nanotechnol, 2019, 14, 747 doi: 10.1038/s41565-019-0500-4
[78]
Andrew C D, Matthew P W. Two-qubit gates for resonant exchange qubits. Phys Rev Lett, 2013, 111, 050503. doi: 10.1103/PhysRevLett.111.050503
[79]
Medford J, Beil J, Taylor J M, et al. Self-consistent measurement and state tomography of an exchange-only spin qubit. Nat Nanotechnol, 2013, 8, 654 doi: 10.1038/nnano.2013.168
[80]
Petit L, Russ M, Eenink G H G J, et al. Design and integration of single-qubit rotations and two-qubit gates in silicon above one kelvin. Commun Mater, 2022, 3, 82 doi: 10.1038/s43246-022-00304-9
[81]
Nguyen D Q L, Heinz I, Burkard G. Quantum gates with oscillating exchange interaction. Quantum Science and Technology. 2024, 9, 0150, 20 doi: 10.1088/2058-9565/acef54
[82]
Sigillito A J, Gullans M J, Edge L F, et al. Coherent transfer of quantum information in a silicon double quantum dot using resonant SWAP gates. npj Quantum Information, 2019, 5, 110 doi: 10.1038/s41534-019-0225-0
[83]
Zajac D M, Sigillito A J, Russ M, et al. Resonantly driven CNOT gate for electron spins. Science, 2018, 359(6374), 439 doi: 10.1126/science.aao5965
[84]
Watson T F, Philips S G J, Kawakami E, et al. A programmable two-qubit quantum processor in silicon. Nature, 2018, 555, 633 doi: 10.1038/nature25766
[85]
Nielsen M A, Chuang I L. Quantum computation and quantum information: 10th Anniversary Edition. Cambridge: Cambridge University Press, 2010
[86]
Van Riggelen F, Lawrie W I L, Russ M, et al. Phase flip code with semiconductor spin qubits. npj Quantum information, 2022, 8, 12 doi: 10.1038/s41534-022-00517-3
[87]
Bennett C H, DiVincenzo D P. Quantum information and computation. Nature, 2000, 404, 247 doi: 10.1038/35005001
[88]
Oh S, Shim Y P, Fei J, et al. Resonant adiabatic passage with three qubits. Phys Rev A, 2013, 87, 022332 doi: 10.1103/PhysRevA.87.022332
[89]
Gullans M J, Petta J R. Coherent transport of spin by adiabatic passage in quantum dot arrays. Phys Rev B, 2020, 102, 155404 doi: 10.1103/PhysRevB.102.155404
[90]
Bacon D, Flammia S T. Adiabatic gate teleportation. Phys Rev Lett, 2009, 103, 120504 doi: 10.1103/PhysRevLett.103.120504
[91]
Bose S. Quantum communication through an unmodulated spin chain. Phys Rev Lett, 2003, 91(20), 207901 doi: 10.1103/PhysRevLett.91.207901
[92]
Friesen M, Biswas A, Hu X D, et al. Efficient multiqubit entanglement via a spin bus. Phys Rev Lett, 2007, 98, 230503 doi: 10.1103/PhysRevLett.98.230503
[93]
Kandel Y P, Qiao H, Fallahi S, et al. Coherent spin-state transfer via Heisenberg exchange. Nature, 2019, 573, 553 doi: 10.1038/s41586-019-1566-8
[94]
Qiao H, Kandel Y P, Manikandan S K. et al. Conditional teleportation of quantum-dot spin states. Nat Commun, 2020, 11, 3022 doi: 10.1038/s41467-020-16745-0
[95]
Georgescu I M, Ashhab S, Nori F. Quantum simulation. Rev Mod Phys, 2014, 86, 153 doi: 10.1103/RevModPhys.86.153
[96]
Byrnes T, Kim N Y, Kusudo K, et al. Quantum simulation of Fermi−Hubbard models in semiconductor quantum-dot arrays. Phys Rev B, 2008, 78, 075320 doi: 10.1103/PhysRevB.78.075320
[97]
Hensgens T, Fujita T, Janssen L, et al. Quantum simulation of a Fermi−Hubbard model using semiconductor quantum dot array. Nature, 2017, 548, 70 doi: 10.1038/nature23022
[98]
Imada M, Fujimori A, Tokura Y. Metal−insulator transitions. Rev Mod Phys, 70(4), 1039
[99]
Stafford C A, Sarma S D. Collective Coulomb blockade in an array of quantum dots: a Mott−Hubbard approach. Phys Rev Lett, 1994, 72(22), 3590 doi: 10.1103/PhysRevLett.72.3590
[100]
Nagaoka Y. Ferromagnetism in a narrow, almost half-filled s band. Phys Rev, 1966, 147, 392 doi: 10.1103/PhysRev.147.392
[101]
Dehollain J P, Mukhopadhyay U, Michal V P, et al. Nagaoka ferromagnetism observed in a quantum dot plaquette. Nature, 2020, 579, 528 doi: 10.1038/s41586-020-2051-0
[102]
Maciej L, Anna S, Veronica A, et al. Ultracold atomic gases in optical lattices: mimicking condensed matter physics and beyond. Adv Phys, 2007, 56, 243 doi: 10.1080/00018730701223200
[103]
Roati G, D’Errico C, Fallani L, et al. Anderson localization of a non-interacting Bose−Einstein condensate. Nature, 2008, 453, 895 doi: 10.1038/nature07071
[104]
Damski B, Zakrzewski J. Mott-insulator phase of the one-dimensional Bose−Hubbard model: a high-order perturbative study. Phys Rev A, 2006, 74, 043609 doi: 10.1103/PhysRevA.74.043609
[105]
Wessel S, Alet F, Troyer M, et al. Quantum Monte Carlo simulations of confined bosonic atoms in optical lattices. Phys Rev A, 2004, 70, 053615 doi: 10.1103/PhysRevA.70.053615
Fig. 1.  (Color online) The exchange interaction $J$ is the energy difference between the antisymmetric orbital wave function and the symmetric orbital wave function of two electrons.

Fig. 2.  (Color online) (a) Quartic potential as illustrated in Eq. (2) with $ \hbar\omega_0=3\text{ meV} $ and $ a=19\text{ nm} $. The potential is used to simulate the coupling of two electrons locating in two harmonic wells centered at $ \left( { - a,0} \right) $ and $(a,0)$. The effective Bohr radius of harmonic well is ${a_{\text{B}}} = \sqrt {\hbar /m{\omega _0}} $. (b) The exchange coupling strength $J$ between two spins as a function of the inter-dot spacing $d = a/{a_{\text{B}}}$ with $ \hbar\omega_0=3\text{ meV} $, $ a_{\text{B}}=19\text{ nm} $ and $c = 2.4$ [See Eq. (6)].

Fig. 3.  (Color online) (a) Double quantum dots in GaAs/AlGaAs heterostructure. (b) Two dimensional stability diagram of double quantum dots. The purple dashed-line is the position where $\Delta \varepsilon $ is zero, the yellow dashed-arrow indicates the direction along which $\Delta \varepsilon $ increases.

Fig. 4.  (Color online) (a) Energy of $\left| {{\psi _{\text{A}}}} \right\rangle $ and $\left| {{\psi _{\text{B}}}} \right\rangle $ versus the detuning energy $\Delta \varepsilon $. (b) The electrostatic potential $U$ of quantum dots in Fig. 3(a) along the X-axis at Y ~ $0.08{\text{ }}\mu {\text{m}}$. The barrier height ${E_{\text{B}}}$ between two quantum dots can be manipulated by the barrier gate voltage ${V_{\text{b}}}$. (c) The barrier height ${E_{\text{B}}}$ is negatively proportional to the barrier gate voltage ${V_{\text{b}}}$. The black straight line is a linear fit.

Fig. 5.  (Color online) (a) Infinitely deep double-well model. $a$ is the width of the well, $L$ the width of the barrier between two wells, and ${E_{\text{B}}}$ the barrier height between two wells. For simplicity, the potential out of the double wells is set infinitely high due to Coulomb blockade effect. (b) The tunnel coupling ${t_{\text{c}}}$ as a function of barrier height ${E_{\text{B}}}$. The data points are calculated based on the device shown in Fig. 3(a). The curve is a fit to Eq. (22).

Fig. 6.  (a) The exchange energy $J$ as a function of the barrier gate voltage ${V_{\text{b}}}$ as described in Eqs. (18) and (22). (b) The exchange energy $J$ as a function of detuning energy $\Delta \varepsilon $ as described in Eq. (18).

Fig. 7.  (Color online) (a) The construction of S−T0 qubit. The exchange interaction $J$ between quantum dots, and a magnetic field gradient $\Delta B$ is required to build a S−T0 qubit with double quantum dots. (b) States with two electron spins. The states with color mapping of orange and cyan color are the computational states of S−T0 qubit. The gray color mapped states are the leakage states. (c) Two states of $\left| S \right\rangle $ and $\left| {{T_0}} \right\rangle $ form a two-level subspace. (d) Exchange coupling $J$ and a longitudinal magnetic-field gradient $\Delta B$ provide two orthogonal manipulation axes for S−T0 qubit operation.

Fig. 8.  (Color online) (a) The construction of exchange-only qubit. Only the exchange interaction ${J_{12}}$ and ${J_{23}}$ between quantum dots are required to build the exchange-only qubit with linearly coupled triple quantum dots. (b) States with three electron spins. The states with color mapping of orange and cyan color are the computational states of exchange-only qubit. The gray color mapped states are the leakage states. (c) Two states of $\left| D \right\rangle $ and $\left| {D'} \right\rangle $ form a two-level subspace. $\left| D \right\rangle $ is the mixture of $\left| {{D_{ + 1/2}}} \right\rangle $ and $\left| {{D_{ - 1/2}}} \right\rangle $, $\left| {D'} \right\rangle $ is the mixture of $\left| {D_{ + 1/2}'} \right\rangle $ and $\left| {D_{ - 1/2}'} \right\rangle $. (d) The exchange coupling parameters of ${J_{12}}$ and ${J_{23}}$ provide two independent manipulation axes of the exchange-only qubit manipulation.

Fig. 9.  (Color online) The spin configuration of $ \left| {D'} \right\rangle $ and $ \left| D \right\rangle $. ${S_1}$ is the spin quantum number of electron in the first quantum dot. ${S_{23}}$ is the total spin quantum number of two electrons, which makes the second and the third quantum dot singly occupied.

Fig. 10.  (Color online) (a) The quantum circuit of SWAP gate. (b) Double quantum dots with one single electron in each dot. ${{\boldsymbol{S}}_{1}}$(${{\boldsymbol{S}}_{2}}$) is the spin of electron in left (right) quantum dot. The exchange interaction $J$ is required to obtain the SWAP gate in specific duration.

Fig. 11.  (Color online) (a) The quantum circuit of CNOT gate. (b) Double quantum dots with one single electron in each dot. ${{\boldsymbol{S}}_1}$(${{\boldsymbol{S}}_2}$) is the spin of electron in left (right) quantum dot. The exchange interaction $J$ between two quantum dots, and a magnetic field gradient $\Delta {\boldsymbol{B}}$ are required to achieve the CNOT gate operation. (c) Schematic energy level diagram adopted from Table 5. (d) The spin of electron in the left quantum dot is flipped with absorbing a photo from an alternating magnetic field at frequency of $f_{\left| {{\psi _{\text{R}}}} \right\rangle = \left| \uparrow \right\rangle }^{\text{L}}$ when the spin of electron in the right quantum dot is $\left| \uparrow \right\rangle $.

Fig. 12.  (Color online) (a) Spin-coherent transport through adiabatic passage. (b) Long-range coupler of spins to transfer an arbitrary spin from the left most Alice to the right most Bob[91]. (c) A representation of triple quantum dots for simulation of interaction-driven Mott metal−insulator transition. (d) Energy spectrum of low spin state and ferromagnetic state as a function of tunnel coupling strength, where $S$ is the total spin number of the three electrons.

Table 1.   The eigenstates and the corresponding eigenenergy of the Hamiltonian (Eq. (19)).

EigenstateEigenenergyDeviation
$\left| {{\psi _{\text{A}}}} \right\rangle $${\varepsilon _{\text{M}}} + \sqrt {{{(\Delta \varepsilon )}^2}/4 + t_{\text{c}}^{\text{2}}} $$\Delta = \sqrt {{{\left( {\Delta \varepsilon } \right)}^2} + 4t_{\text{c}}^{\text{2}}} $
$\left| {{\psi _{\text{B}}}} \right\rangle $${\varepsilon _{\text{M}}} - \sqrt {{{(\Delta \varepsilon )}^2}/4 + t_{\text{c}}^{\text{2}}} $
DownLoad: CSV

Table 2.   The eigenstates and corresponding eigenenergies of Hamiltonians $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^1 $ and $ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^2 $ with $\Delta B \equiv {B_2} - {B_1}$.

HamiltonianEigenstateEigenenergy
$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{1}} $$\left| { \uparrow \downarrow } \right\rangle $$ - \mu \Delta B/2$
$\left| { \downarrow \uparrow } \right\rangle $$\mu \Delta B/2$
$\left| { \uparrow \uparrow } \right\rangle $$\mu ({B_1} + {B_2})/2$
$\left| { \downarrow \downarrow } \right\rangle $$ - \mu ({B_1} + {B_2})/2$
$ \hat H_{{\text{S}} - {{\text{T}}_{\text{0}}}}^{\text{2}} $$\left| S \right\rangle $$ - J$
$\left| {{T_0}} \right\rangle $$0$
$\left| { \uparrow \uparrow } \right\rangle $$0$
$\left| { \downarrow \downarrow } \right\rangle $$0$
DownLoad: CSV

Table 3.   The eigenstates and corresponding eigenenergies of Hamiltonian $ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ and $ \hat H_{{\text{E}} - {\text{O}}}^{\text{2}} $[41].

Hamiltonian Eigenstate Eigenenergy
$ \hat H_{{\text{E}} - {\text{O}}}^{\text{1}} $ $\left| {{Q_{ + 3/2}}} \right\rangle $ 0
$ \left| {{Q_{ + 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 3/2}}} \right\rangle $ 0
$\left| {{{\bar D}_{ + 1/2}}} \right\rangle $ 0
$\left| {{{\bar D}_{ - 1/2}}} \right\rangle $ 0
$ \left| {{{\bar D}'}_{ + 1/2}} \right\rangle $ $ - {J_{12}}$
$ \left| {{{\bar D}'}_{ - 1/2}} \right\rangle $ $ - {J_{12}}$
$\hat H_{{\text{E}} - {\text{O}}}^{\text{2}}$ $\left| {{Q_{ + 3/2}}} \right\rangle $ 0
$ \left| {{Q_{ + 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 1/2}}} \right\rangle $ 0
$ \left| {{Q_{ - 3/2}}} \right\rangle $ 0
$\left| {{D_{ + 1/2}}} \right\rangle $ 0
$\left| {{D_{ - 1/2}}} \right\rangle $ 0
$\left| {D_{ + 1/2}'} \right\rangle $ $ - {J_{23}}$
$\left| {D_{ - 1/2}'} \right\rangle $ $ - {J_{23}}$
DownLoad: CSV

Table 4.   The eigenstates and corresponding eigenenergies of Hamiltonian $ {\hat H_{{\text{SWAP}}}}. $

HamiltonianEigenstateEigenenergy
$ {\hat H_{{\text{SWAP}}}} $$\left| S \right\rangle $$ - J$
$\left| {{T_0}} \right\rangle $$0$
$\left| { \uparrow \uparrow } \right\rangle $$0$
$\left| { \downarrow \downarrow } \right\rangle $$0$
DownLoad: CSV

Table 5.   The eigenstates and corresponding eigenenergies of the Hamiltonian ${\hat H_{{\text{CNOT}}}}.$

HamiltonianEigenstateEigenenergy
$ {\hat H_{{\text{CNOT}}}} $$ \left| { \uparrow \uparrow } \right\rangle $$\mu ({B_1} + {B_2})/2$
$ \left| { \Downarrow \Uparrow } \right\rangle $$\left[ { - J + \sqrt {{J^2} + {{\left( {g\Delta B\mu } \right)}^2}} } \right]/2$
$ \left| { \Uparrow \Downarrow } \right\rangle $$\left[ { - J - \sqrt {{J^2} + {{\left( {g\Delta B\mu } \right)}^2}} } \right]/2$
$ \left| { \downarrow \downarrow } \right\rangle $$ - \mu ({B_1} + {B_2})/2$
DownLoad: CSV
[1]
Bhat H A, Khanday F A, Kaushik B K, et al. Quantum computing: fundamentals, implementations and applications. IEEE Open J Nanotechnology, 2022, 3, 61 doi: 10.1109/OJNANO.2022.3178545
[2]
DiVincenzo D P. The physical implementation of quantum computation. Fortschritte der Physik, 2000, 48(9-11), 771 doi: 10.1002/1521-3978(200009)48:9/11
[3]
Georgescu I. The DiVincenzo criteria 20 years on. Nat Rev Phys, 2020, 2, 666 doi: 10.1038/s42254-020-00256-4
[4]
Ladd T, Jelezko F, Laflamme R, et al. Quantum computers. Nature, 2010, 464(4), 45 doi: 10.1038/nature08812
[5]
Wei S J, Li H, Long G L. A full quantum eigensolver for quantum chemistry simulations. Research, 2020, 1, 111 doi: 10.34133/2020/1486935
[6]
Shor P W. Algorithms for quantum computation: discrete logarithms and factoring. 35th Annual Symposium on Foundations of Computer Science, 1994, 124 doi: 10.1109/SFCS.1994.365700
[7]
Grover L K. Quantum mechanics helps in searching for a needle in a haystack. Phys Rev Lett, 1997, 79(2), 325 doi: 10.1103/PhysRevLett.79.325
[8]
Hanson R, Kouwenhoven L P, Petta J R, et al. Spins in few-electron quantum dots. Rev Mod Phys, 2007, 79(4), 1217 doi: 10.1103/RevModPhys.79.1217
[9]
Kloeffel C, Loss D. Prospects for spin-based quantum computing in quantum dots. Annu Rev Condens Matter Phys, 2013, 4(1), 51 doi: 10.1146/annurev-conmatphys-030212-184248
[10]
Zhang X, Li H O, Wang K, et al. Qubits based on semiconductor quantum dots. Chin Phys B, 2018, 27(2), 020305 doi: 10.1088/1674-1056/27/2/020305
[11]
Chatterjee A, Stevenson P, Franceschi S, et al. Semiconductor qubits in practice. Nat Rev Phys, 2021, 3(3), 157 doi: 10.1038/s42254-021-00283-9
[12]
Wang N, Wang B C, Guo G P. New progress of silicon-based semiconductor quantum computation. Acta Phys Sin, 2022, 71(23), 230301 doi: 10.7498/aps.71.20221900
[13]
Zhang X, Li H O, Cao G, et al. Semiconductor quantum computation. Natl Sci Rev, 2019, 6(1), 32 doi: 10.1093/nsr/nwy153
[14]
Barthelemy P, Vandersypen L M K. Quantum dot systems: a versatile platform for quantum simulations. Ann Phys, 2013, 525(10-11), 808 doi: 10.1002/andp.201300124
[15]
Burkard G, Ladd T D, Pan A, et al. Semiconductor spin qubits. Rev Mod Phys, 2023, 95(2), 025003 doi: 10.1103/RevModPhys.95.025003
[16]
Mu J W, Huang S Y, Liu Z H, et al. A highly tunable quadruple quantum dot in a narrow bandgap semiconductor InAs nanowire. Nanoscale, 2021, 13, 3983 doi: 10.1039/D0NR08655J
[17]
Zwerver A M J, Krähenmann T, Watson T F, et al. Qubits made by advanced semiconductor manufacturing. Nature Electronics, 2022, 5, 184 doi: 10.1038/s41928-022-00727-9
[18]
Pomorski K, Giounanlis P, Blokhina E, et al. Analytic view on coupled single-electron lines. Semicond Sci Technol, 2019, 34, 125015 doi: 10.1088/1361-6641/ab4f40
[19]
Lawrie W I L, Eenink H G J, Hendrickx N W, et al. Quantum dot arrays in silicon and germanium. Appl Phys Lett, 2020, 116, 080501 doi: 10.1063/5.0002013
[20]
Sánchez R, Gallego Marcos F, Platero G. Superexchange blockade in triple quantum dots. Phys Rev B, 2014, 89, 161402(R doi: 10.1103/PhysRevB.89.161402
[21]
Qiao H, Kandel Y P, Fallahi S, et al. Long-distance superexchange between semiconductor quantum-dot electron spins. Phys Rev Lett, 2021, 126, 017701 doi: 10.1103/PhysRevLett.126.017701
[22]
Kouwenhoven L P, Austing D G, Tarucha S, et al. Few-electron quantum dots. Rep Prog Phys, 2001, 64, 701 doi: 10.1088/0034-4885/64/6/201
[23]
Reimann S M, Manninen M. Electronic structure of quantum dots. Rev Mod Phys, 2002, 74, 1283 doi: 10.1103/RevModPhys.74.1283
[24]
Feng D, Jin G J. Condensed matter physics. Beijing: Higher Education Press, 2013, 1, 1 (in Chinese)
[25]
Heitler W, London F. Wechselwirkung neutraler atome und homöopolare bindung nach der quantenmechanik. Zeitschrift für Physik, 1927, 44, 455 (in German doi: 10.1007/BF01397394
[26]
Hatano T, Amaha S, Kubo T, et al. Manipulation of exchange coupling energy in a few-electron double quantum dot. Phys Rev B, 2008, 77(24), 241301 doi: 10.1103/PhysRevB.77.241301
[27]
Heisenberg V W. Mehrkörperproblem und resonanz in der quantenmechanik. Zeitschrift für Physik, 1926, 38, 411 (in German
[28]
Dirac P A. On the Theory of quantum mechanics. The Royal society, 1926, 661
[29]
Mullin W J, Blaylock G. Quantum statistics: Is there an effective fermion repulsion or boson attraction? Am J Phys, 2003, 71, 1223 doi: 10.1119/1.1590658
[30]
Wannier G H. Statistical physics. New York: Wiley, 1987
[31]
Leighton R B. Principles of modern physics. New York: McGraw−Hill Book Company, 1959
[32]
Griffiths D J, Schroeter D F. Introduction to quantum mechanics (third edition). Cambridge: Cambridge university Press, 2018
[33]
Zeng J Y. Quantum mechanics (fifth edition). Beijing: Science Press, 2013, 1, 1 (in Chinese)
[34]
Van der Wiel W G, Stopa M, Kodera T, et al. Semiconductor quantum dots for electron spin qubits. New J Phys, 2006, 8, 28 doi: 10.1088/1367-2630/8/2/028
[35]
Saraiva A L, Calderón M J, Koiller B. Reliability of the Heitler−London approach for the exchange coupling between electrons in semiconductor nanostructures. Phys Rev B, 2007, 233302 doi: 10.1103/PhysRevB.76.233302
[36]
Liu E K. Semiconductor physics (eighth edition). Beijing: Publishing House of Electronics Industry, 2023 (in Chinese)
[37]
Burkard G, Loss D, DiVincenzo D P. Coupled quantum dots as quantum gates. Phys Rev B, 1999, 59(3), 2070 doi: 10.1103/PhysRevB.59.2070
[38]
DiVincenzo D P, Loss D. Quantum information is physical. Superlattices and Microstructures, 1998, 23(3), 420 doi: 10.1006/spmi.1997.0520
[39]
Fanciulli M. Electron spin resonance and related phenomena in low dimensional structures. Berlin: Springer, 2009
[40]
Tokura Y, Austing D G, Tarucha S. Single-electron tunnelling in two vertically coupled quantum dots. J Phys Condens Matter, 1999, 11, 6023 doi: 10.1088/0953-8984/11/31/310
[41]
Laird E A, Taylor J M, DiVincenzo D P, et al. Coherent spin manipulation in an exchange-only qubit. Phys Rev B, 2010, 82, 075403 doi: 10.1103/PhysRevB.82.075403
[42]
Elzerman J M, Hanson R, Greidanus J S, et al. Few-electron quantum dot circuit with integrated charge read out. Phys Rev B, 2003, 67, 161308(R doi: 10.1103/PhysRevB.67.161308
[43]
Russ M, Burkard G. Three-electron spin qubits. J Phys Condens Matter, 2017, 29, 393001 doi: 10.1088/1361-648X/aa761f
[44]
Hubbard J. Electron correlations in narrow energy bands. Proc R Soc Lond A, 1963, 276(1365), 238 doi: 10.1098/rspa.1963.0204
[45]
Van der Wiel W G, Franceschi S De, Elzerman J M, et al. Electron transport through double quantum dots. Rev Mod Phys, 2003, 75(1), 1 doi: 10.1103/RevModPhys.75.1
[46]
Spałek J. Theory of unconventional superconductivity in strongly correlated systems: real space pairing and statistically consistent mean-field theory- in perspective. Acta Phys Pol A, 2012, 121(4), 764 doi: 10.12693/APhysPolA.121.764
[47]
Taylor J M, Petta J R, Johnson A C, et al. Relaxation, dephasing, and quantum control of electron spins in double quantum dots. Phys Rev B, 2007, 76(3), 035315 doi: 10.1103/PhysRevB.76.035315
[48]
Balachandran A P, Ercolessi E, Morandi G, et al. Hubbard model and anyon superconductivity: a review. Int J Mod Phys B, 1990, 4(14), 2057 doi: 10.1142/S0217979290001030
[49]
Spałek J. Effect of pair hopping and magnitude of intra-atomic interaction on exchange-mediated superconductivity. Phys Rev B, 1988, 37(1), 533 doi: 10.1103/PhysRevB.37.533
[50]
Bravyi S, DiVincenzo D P, Loss D. Schrieffer–Wolff transformation for quantum many-body systems. Ann Phys, 2011, 326(10), 2793 doi: 10.1016/j.aop.2011.06.004
[51]
Assa A. Interacting electrons and quantum magnetism. Berlin: Springer-Verlag, 1998
[52]
DiCarlo L, Lynch H J, Johnson A C, et al. Differential charge sensing and charge delocalization in a tunable double quantum dot. Phys Rev Lett, 2004, 92(22), 226801 doi: 10.1103/PhysRevLett.92.226801
[53]
Hu Y J, Churchill H O H, Reilly D J, et al. A Ge/Si heterostructure nanowire-based double quantum dot with integrated charge sensor. Nat Nanotechnol, 2007, 2, 622 doi: 10.1038/nnano.2007.302
[54]
Wang X M, Huang S Y, Wang J Y, et al. A charge sensor integration to tunable double quantum dots on two neighboring InAs nanowires. Nanoscale, 2021, 13, 1048 doi: 10.1039/D0NR07115C
[55]
Hsiao T K, van Diepen C J, Mukhopadhyay U, et al. Efficient orthogonal control of tunnel couplings in a quantum dot array. Phys Rev Appl, 2020, 13(5), 054018 doi: 10.1103/PhysRevApplied.13.054018
[56]
Mills A R, Zajac D M, Gullans M J. et al. Shuttling a single charge across a one-dimensional array of silicon quantum dots. Nat Commun, 2019, 10, 1063 doi: 10.1038/s41467-019-08970-z
[57]
Hu X D, Sarma S D. Hilbert-space structure of a solid-state quantum computer: two-electron states of a double-quantum-dot artificial molecule. Phys Rev A, 2000, 61(6), 062301 doi: 10.1103/PhysRevA.61.062301
[58]
Ercan H E, Coppersmith S N, Friesen M. Strong electron−electron interactions in Si/SiGe quantum dots. Phys Rev B, 2021, 104, 235302 doi: 10.1103/PhysRevB.104.235302
[59]
Bellucci D, Rontani M, Troiani F, et al. Competing mechanisms for singlet−triplet transition in artificial molecules. Phys Rev B, 2004, 69, 201308 doi: 10.1103/PhysRevB.69.201308
[60]
Giavaras G, Tokura Y. Probing the singlet−triplet splitting in double quantum dots: implications of the ac field amplitude. Phys Rev B, 2019, 100, 195421 doi: 10.1103/PhysRevB.100.195421
[61]
Stopa M, Marcus C M. Magnetic field control of exchange and noise immunity in double quantum dots. Nano Lett, 2008, 8(6), 1778 doi: 10.1021/nl801282t
[62]
Geyer S, Hetényi B, Bosco S, et al. Anisotropic exchange interaction of two hole spin qubits. Nat Phys, 2024 doi: 10.1038/s41567-024-02481-5
[63]
Liu Z H, Entin-Wohlman O, Aharony A, et al. Control of the two-electron exchange interaction in a nanowire double quantum dot. Phys Rev B, 2018, 98, 241303 doi: 10.1103/PhysRevB.98.241303
[64]
Loss D, DiVincenzo D P. Quantum computation with quantum dots. Phys Rev A, 1998, 57(1), 1050 doi: 10.1103/PhysRevA.57.120
[65]
Koppens F H L, Buizert C, Tielrooij K J, et al. Driven coherent oscillations of a single electron spin in a quantum dot. Nature, 2006, 442, 766 doi: 10.1038/nature05065
[66]
Liu H, Wang K, Gao F, et al. Ultrafast and electrically tunable Rabi frequency in a Germanium hut wire hole spin qubit. Nano Lett, 2023, 23(9), 3810 doi: 10.1021/acs.nanolett.3c00213
[67]
Philips S G J, Mądzik M T, Amitonov S V, et al. Universal control of a six-qubit quantum processor in silicon. Nature, 2022, 609, 919 doi: 10.1038/s41586-022-05117-x
[68]
Levy J. Universal quantum computation with spin-1/2 pairs and Heisenberg exchange. Phys Rev Lett, 2002, 89, 147902 doi: 10.1103/PhysRevLett.89.147902
[69]
DiVincenzo D, Bacon D, Kempe J, et al. Universal quantum computation with the exchange interaction. Nature, 2000, 408, 339 doi: 10.1038/35042541
[70]
Zhang T, Liu H, Cao F, et al. Anisotropic g-factor and spin−orbit field in a germanium hut wire double quantum dot. Nano Lett, 2021, 21(9), 3835 doi: 10.1021/acs.nanolett.1c00263
[71]
Wang J Y, Huang S Y, Lei Z J, et al. Measurements of the spin−orbit interaction and Landé g factor in a pure-phase InAs nanowire double quantum dot in the Pauli spin-blockade regime. Appl Phys Lett, 2016, 109(5), 053106 doi: 10.1063/1.4960464
[72]
Petta J R, Johnson A C, Taylor J M, et al. Coherent manipulation of coupled electron spins in semiconductor quantum dots. Science, 2005, 309(5744), 2180 doi: 10.1126/science.1116955
[73]
Wardrop M P, Doherty A C. Exchange-based two-qubit gate for singlet−triplet qubits. Phys Rev B, 2014, 90, 045418 doi: 10.1103/PhysRevB.90.045418
[74]
Johnson A C, Petta J R, Taylor J M, et al. Triplet–singlet spin relaxation via nuclei in a double quantum dot. Nature, 2005, 435, 925 doi: 10.1038/nature03815
[75]
Prance J R, Shi Z, Simmons C B, et al. Single-shot measurement of triplet−singlet relaxation in a Si/SiGe double quantum dot. Phys Rev Lett, 2012, 108, 046808 doi: 10.1103/PhysRevLett.108.046808
[76]
Weinstein A J, Reed M D, Jones A M, et al. Universal logic with encoded spin qubits in silicon. Nature, 2023, 615, 817 doi: 10.1038/s41586-023-05777-3
[77]
Andrews R W, Jones C, Reed M D, et al. Quantifying error and leakage in an encoded Si/SiGe triple-dot qubit. Nat Nanotechnol, 2019, 14, 747 doi: 10.1038/s41565-019-0500-4
[78]
Andrew C D, Matthew P W. Two-qubit gates for resonant exchange qubits. Phys Rev Lett, 2013, 111, 050503. doi: 10.1103/PhysRevLett.111.050503
[79]
Medford J, Beil J, Taylor J M, et al. Self-consistent measurement and state tomography of an exchange-only spin qubit. Nat Nanotechnol, 2013, 8, 654 doi: 10.1038/nnano.2013.168
[80]
Petit L, Russ M, Eenink G H G J, et al. Design and integration of single-qubit rotations and two-qubit gates in silicon above one kelvin. Commun Mater, 2022, 3, 82 doi: 10.1038/s43246-022-00304-9
[81]
Nguyen D Q L, Heinz I, Burkard G. Quantum gates with oscillating exchange interaction. Quantum Science and Technology. 2024, 9, 0150, 20 doi: 10.1088/2058-9565/acef54
[82]
Sigillito A J, Gullans M J, Edge L F, et al. Coherent transfer of quantum information in a silicon double quantum dot using resonant SWAP gates. npj Quantum Information, 2019, 5, 110 doi: 10.1038/s41534-019-0225-0
[83]
Zajac D M, Sigillito A J, Russ M, et al. Resonantly driven CNOT gate for electron spins. Science, 2018, 359(6374), 439 doi: 10.1126/science.aao5965
[84]
Watson T F, Philips S G J, Kawakami E, et al. A programmable two-qubit quantum processor in silicon. Nature, 2018, 555, 633 doi: 10.1038/nature25766
[85]
Nielsen M A, Chuang I L. Quantum computation and quantum information: 10th Anniversary Edition. Cambridge: Cambridge University Press, 2010
[86]
Van Riggelen F, Lawrie W I L, Russ M, et al. Phase flip code with semiconductor spin qubits. npj Quantum information, 2022, 8, 12 doi: 10.1038/s41534-022-00517-3
[87]
Bennett C H, DiVincenzo D P. Quantum information and computation. Nature, 2000, 404, 247 doi: 10.1038/35005001
[88]
Oh S, Shim Y P, Fei J, et al. Resonant adiabatic passage with three qubits. Phys Rev A, 2013, 87, 022332 doi: 10.1103/PhysRevA.87.022332
[89]
Gullans M J, Petta J R. Coherent transport of spin by adiabatic passage in quantum dot arrays. Phys Rev B, 2020, 102, 155404 doi: 10.1103/PhysRevB.102.155404
[90]
Bacon D, Flammia S T. Adiabatic gate teleportation. Phys Rev Lett, 2009, 103, 120504 doi: 10.1103/PhysRevLett.103.120504
[91]
Bose S. Quantum communication through an unmodulated spin chain. Phys Rev Lett, 2003, 91(20), 207901 doi: 10.1103/PhysRevLett.91.207901
[92]
Friesen M, Biswas A, Hu X D, et al. Efficient multiqubit entanglement via a spin bus. Phys Rev Lett, 2007, 98, 230503 doi: 10.1103/PhysRevLett.98.230503
[93]
Kandel Y P, Qiao H, Fallahi S, et al. Coherent spin-state transfer via Heisenberg exchange. Nature, 2019, 573, 553 doi: 10.1038/s41586-019-1566-8
[94]
Qiao H, Kandel Y P, Manikandan S K. et al. Conditional teleportation of quantum-dot spin states. Nat Commun, 2020, 11, 3022 doi: 10.1038/s41467-020-16745-0
[95]
Georgescu I M, Ashhab S, Nori F. Quantum simulation. Rev Mod Phys, 2014, 86, 153 doi: 10.1103/RevModPhys.86.153
[96]
Byrnes T, Kim N Y, Kusudo K, et al. Quantum simulation of Fermi−Hubbard models in semiconductor quantum-dot arrays. Phys Rev B, 2008, 78, 075320 doi: 10.1103/PhysRevB.78.075320
[97]
Hensgens T, Fujita T, Janssen L, et al. Quantum simulation of a Fermi−Hubbard model using semiconductor quantum dot array. Nature, 2017, 548, 70 doi: 10.1038/nature23022
[98]
Imada M, Fujimori A, Tokura Y. Metal−insulator transitions. Rev Mod Phys, 70(4), 1039
[99]
Stafford C A, Sarma S D. Collective Coulomb blockade in an array of quantum dots: a Mott−Hubbard approach. Phys Rev Lett, 1994, 72(22), 3590 doi: 10.1103/PhysRevLett.72.3590
[100]
Nagaoka Y. Ferromagnetism in a narrow, almost half-filled s band. Phys Rev, 1966, 147, 392 doi: 10.1103/PhysRev.147.392
[101]
Dehollain J P, Mukhopadhyay U, Michal V P, et al. Nagaoka ferromagnetism observed in a quantum dot plaquette. Nature, 2020, 579, 528 doi: 10.1038/s41586-020-2051-0
[102]
Maciej L, Anna S, Veronica A, et al. Ultracold atomic gases in optical lattices: mimicking condensed matter physics and beyond. Adv Phys, 2007, 56, 243 doi: 10.1080/00018730701223200
[103]
Roati G, D’Errico C, Fallani L, et al. Anderson localization of a non-interacting Bose−Einstein condensate. Nature, 2008, 453, 895 doi: 10.1038/nature07071
[104]
Damski B, Zakrzewski J. Mott-insulator phase of the one-dimensional Bose−Hubbard model: a high-order perturbative study. Phys Rev A, 2006, 74, 043609 doi: 10.1103/PhysRevA.74.043609
[105]
Wessel S, Alet F, Troyer M, et al. Quantum Monte Carlo simulations of confined bosonic atoms in optical lattices. Phys Rev A, 2004, 70, 053615 doi: 10.1103/PhysRevA.70.053615

24050043Supporting_Information.pdf

1

Synthesis and characterization of ZnS-based quantum dots to trace low concentration of ammonia

Uma Devi Godavarti, P. Nagaraju, Vijayakumar Yelsani, Yamuna Pushukuri, P. S. Reddy, et al.

Journal of Semiconductors, 2021, 42(12): 122901. doi: 10.1088/1674-4926/42/12/122901

2

High performance GaN-based hybrid white micro-LEDs integrated with quantum-dots

Feifan Xu, Xu Cen, Bin Liu, Danbei Wang, Tao Tao, et al.

Journal of Semiconductors, 2020, 41(3): 032301. doi: 10.1088/1674-4926/41/3/032301

3

Research on the photoluminescence of spectral broadening by rapid thermal annealing on InAs/GaAs quantum dots

Dandan Ning, Yanan Chen, Xinkun Li, Dechun Liang, Shufang Ma, et al.

Journal of Semiconductors, 2020, 41(12): 122101. doi: 10.1088/1674-4926/41/12/122101

4

Perspective: optically-pumped III–V quantum dot microcavity lasers via CMOS compatible patterned Si (001) substrates

Wenqi Wei, Qi Feng, Zihao Wang, Ting Wang, Jianjun Zhang, et al.

Journal of Semiconductors, 2019, 40(10): 101303. doi: 10.1088/1674-4926/40/10/101303

5

Recent progress in epitaxial growth of III–V quantum-dot lasers on silicon substrate

Shujie Pan, Victoria Cao, Mengya Liao, Ying Lu, Zizhuo Liu, et al.

Journal of Semiconductors, 2019, 40(10): 101302. doi: 10.1088/1674-4926/40/10/101302

6

Improvement of tunnel compensated quantum well infrared detector

Chaohui Li, Jun Deng, Weiye Sun, Leilei He, Jianjun Li, et al.

Journal of Semiconductors, 2019, 40(12): 122902. doi: 10.1088/1674-4926/40/12/122902

7

Quantum light source devices of In(Ga)As semiconductor self-assembled quantum dots

Xiaowu He, Yifeng Song, Ying Yu, Ben Ma, Zesheng Chen, et al.

Journal of Semiconductors, 2019, 40(7): 071902. doi: 10.1088/1674-4926/40/7/071902

8

Differential optical gain in a GaInN/AlGaN quantum dot

K. Jaya Bala, A. John Peter

Journal of Semiconductors, 2017, 38(6): 062001. doi: 10.1088/1674-4926/38/6/062001

9

Electric and magnetic optical polaron in quantum dot——Part 1: strong coupling

A. J. Fotue, N. Issofa, M. Tiotsop, S. C. Kenfack, M. P. Tabue Djemmo, et al.

Journal of Semiconductors, 2015, 36(7): 072001. doi: 10.1088/1674-4926/36/7/072001

10

Electro-magnetic weak coupling optical polaron and temperature effect in quantum dot

M. Tiotsop, A. J. Fotue, S. C. Kenfack, N. Issofa, A. V. Wirngo, et al.

Journal of Semiconductors, 2015, 36(10): 102001. doi: 10.1088/1674-4926/36/10/102001

11

The effect of multi-intermediate bands on the behavior of an InAs1-xNx/GaAs1-ySby quantum dot solar cell

Abou El-Maaty M. Aly, A. Nasr

Journal of Semiconductors, 2015, 36(4): 042001. doi: 10.1088/1674-4926/36/4/042001

12

Temperature dependence of InAs/GaAs quantum dots solar photovoltaic devices

E. Garduno-Nolasco, M. Missous, D. Donoval, J. Kovac, M. Mikolasek, et al.

Journal of Semiconductors, 2014, 35(5): 054001. doi: 10.1088/1674-4926/35/5/054001

13

Effect of p-d exchange with an itinerant carrier in a GaMnAs quantum dot

D. Lalitha, A. John Peter

Journal of Semiconductors, 2013, 34(7): 072001. doi: 10.1088/1674-4926/34/7/072001

14

The Origin of Multi-Peak Structures Observed in Photoluminescence Spectra of InAs/GaAs Quantum Dots

Liang Zhimei, Wu Ju, Jin Peng, Lü Xueqin, Wang Zhanguo, et al.

Journal of Semiconductors, 2008, 29(11): 2121-2124.

15

Quantum and Transport Mobilities of a Two-Dimensional Electron Gas in the Presence of the Rashba Spin-Orbit Interaction

Xu Wen

Chinese Journal of Semiconductors , 2006, 27(2): 204-217.

16

Raman Scattering of InAs Quantum Dots with Different Deposition Thicknesses

Zhang Guanjie, Xu Bo, Chen Yonghai, Yao Jianghong, Lin Yaowang, et al.

Chinese Journal of Semiconductors , 2006, 27(6): 1012-1015.

17

Low Voltage Flash Memory Cells Using SiGe Quantum Dots for Enhancing F-N Tunneling

Deng Ning, Pan Liyang, Liu Zhihong, Zhu Jun, Chen Peiyi, et al.

Chinese Journal of Semiconductors , 2006, 27(3): 454-458.

18

Transport Properties of Two Coupled Quantum Dots Under Optical Pumping

Ge Chuannan, Wen Jun, Peng Ju, Wang Baigeng

Chinese Journal of Semiconductors , 2006, 27(4): 598-603.

19

Simulation of Imperfection on Image-Charge Quantum Cellular Automaton Using Image Charge Effect

Wang Yanzhen, Wu Nanjian

Chinese Journal of Semiconductors , 2005, 26(S1): 261-264.

20

Growth of Space Ordered 1.3μm InAs Quantum Dots on GaAs(100) Vicinal Substrates by MOCVD

Liang Song, Zhu Hongliang, Pan Jiaoqing, Wang Wei

Chinese Journal of Semiconductors , 2005, 26(11): 2074-2079.

  • Search

    Advanced Search >>

    GET CITATION

    Zheng Zhou, Yixin Li, Zhiyuan Wu, Xinping Ma, Shichang Fan, Shaoyun Huang. The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. Journal of Semiconductors, 2024, 45(10): 101701. doi: 10.1088/1674-4926/24050043
    Z Zhou, Y X Li, Z Y Wu, X P Ma, S C Fan, and S Y Huang, The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. J. Semicond., 2024, 45(10), 101701 doi: 10.1088/1674-4926/24050043
    shu

    Export: BibTex EndNote

    Article Metrics

    Article views: 812 Times PDF downloads: 104 Times Cited by: 0 Times

    History

    Received: 28 May 2024 Revised: 16 July 2024 Online: Accepted Manuscript: 14 August 2024Uncorrected proof: 16 August 2024Published: 15 October 2024

    Catalog

      Email This Article

      User name:
      Email:*请输入正确邮箱
      Code:*验证码错误
      Zheng Zhou, Yixin Li, Zhiyuan Wu, Xinping Ma, Shichang Fan, Shaoyun Huang. The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. Journal of Semiconductors, 2024, 45(10): 101701. doi: 10.1088/1674-4926/24050043 ****Z Zhou, Y X Li, Z Y Wu, X P Ma, S C Fan, and S Y Huang, The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. J. Semicond., 2024, 45(10), 101701 doi: 10.1088/1674-4926/24050043
      Citation:
      Zheng Zhou, Yixin Li, Zhiyuan Wu, Xinping Ma, Shichang Fan, Shaoyun Huang. The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. Journal of Semiconductors, 2024, 45(10): 101701. doi: 10.1088/1674-4926/24050043 ****
      Z Zhou, Y X Li, Z Y Wu, X P Ma, S C Fan, and S Y Huang, The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing[J]. J. Semicond., 2024, 45(10), 101701 doi: 10.1088/1674-4926/24050043

      The exchange interaction between neighboring quantum dots: physics and applications in quantum information processing

      DOI: 10.1088/1674-4926/24050043
      CSTR: 32376.14.1674-4926.24050043
      More Information
      • Zheng Zhou got his B.S. from University of Science and Technology Beijing in 2022. Now he is a master student at Peking University under the supervision of Prof. Shaoyun Huang. His research focuses on quantum computation and semiconductor quantum devices
      • Yixin Li got her B.S. from Renmin University of China in 2020. Now she is a PhD student at Peking University under the supervision of Prof. Shaoyun Huang and Prof. Shimin Hou. Her research focuses on quantum transport properties and semiconductor quantum devices
      • Zhiyuan Wu got his B.S. from Hebei University of Technology in 2021. Now he is a Master student at Peking University under the supervision of Prof. Shaoyun Huang. His research focuses on gate defined InAs quantum dot and quantum-classical interface
      • Xinping Ma got her B.S. from JiLin University in 2022. Now she is a Master student at Peking University under the supervision of Prof. Shaoyun Huang. Her research focuses on Si/SiGe quantum dot and electrical control of Si/SiGe spin qubit
      • Shichang Fan got his B.S. from East China University of Science and Technology in 2024. Now he is a Master student at Peking University under the supervision of Prof. Shaoyun Huang. His research focuses on the preparation of germanium quantum dots and the theoretical calculation of semiconductor spin qubits
      • Shaoyun Huang received B.S. and M.S. degrees in physics from Nanjing University, Nanjing, China, in September 1997 and 2000, respectively, and Ph.D. degree in semiconductor nanoelectronics from Tokyo Institute of Technology in September 2003. He is an associate professor of School of Electronics, Peking University. He is also a key member of the Beijing key laboratory of quantum devices and the Key Laboratory for the Physics and Chemistry of Nanodevices. His research interests include semiconductor quantum dot spin qubits, solid-state quantum computations, semiconductor low dimensional nanostructure based quantum devices, low-temperature transport
      • Corresponding author: syhuang@pku.edu.cn
      • Received Date: 2024-05-28
      • Revised Date: 2024-07-16
      • Available Online: 2024-08-14

      Catalog

        /

        DownLoad:  Full-Size Img  PowerPoint
        Return
        Return