You are on page 1of 19

Quantum theory of the charge stability diagram of semiconductor double quantum

dot systems
Xin Wang, Shuo Yang, and S. Das Sarma
Condensed Matter Theory Center, Department of Physics,
University of Maryland, College Park, MD 20742

We complete our recently introduced theoretical framework treating the double quantum dot
system with a generalized form of Hubbard model. The e↵ects of all quantum parameters involved
in our model on the charge stability diagram are discussed in detail. A general formulation of the
microscopic theory is presented, and truncating at one orbital per site, we study the implication of
arXiv:1104.5491v2 [cond-mat.mes-hall] 15 Aug 2011

di↵erent choices of the model confinement potential on the Hubbard parameters as well as the charge
stability diagram. We calculate the charge stability diagram keeping three orbitals per site and find
that the e↵ect of additional higher-lying orbitals on the subspace with lowest-energy orbitals only
can be regarded as a small renormalization of Hubbard parameters, thereby justifying our practice
of keeping only the lowest-orbital in all other calculations. The role of the harmonic oscillator
frequency in the implementation of the Gaussian model potential is discussed, and the e↵ect of an
external magnetic field is identified to be similar to choosing a more localized electron wave function
in microscopic calculations. The full matrix form of the Hamiltonian including all possible exchange
terms, and several peculiar charge stability diagrams due to unphysical parameters are presented in
the appendix, thus emphasizing the critical importance of a reliable microscopic model in obtaining
the system parameters defining the Hamiltonian.

PACS numbers: 73.21.La, 03.67.Lx, 71.10.-w, 73.23.Hk

I. INTRODUCTION hand, the quantum e↵ect, in particular the exchange


interaction,38,39 plays an essential role in the qubit
The idea of quantum computation, introduced two manipulation. The exchange interaction constitutes
decades ago,1 has attracted intense research interest be- two-qubit operations in the Loss-DiVincenzo proposal6
cause of its ability to provide novel solutions to cer- (which uses single-electron spin-up/down states as the
tain problems that have been deemed intractable on a qubit). In the double dot singlet/triplet proposal, the ex-
classical computer.2 The fundamental unit of a quan- change interaction controls the singlet-triplet level split-
tum computer is the qubit, which is implemented using ting thus, along with an inhomogeneous magnetic field,
quantum two-level (and sometimes multi-level) systems. provides arbitrary rotation along the Bloch sphere and
Many physical systems have been put forward as candi- thereby achieves full single-qubit control.8,13,23,40 In a
dates on which qubits can be embedded, such as nuclear three-dot scheme,25 the need for an inhomogeneous mag-
magnetic resonance,3 cavity QED,4 and trapped ions.5 netic field is eliminated, and the exchange interaction
Among these proposals, the solid state systems based on alone suffices for universal quantum computation.
semiconductor quantum dots stand out as most promis- The charge stability diagram is fundamental for spin-
ing for the implementation of qubits,6–9 since modern based quantum computation: it is the starting point for
semiconductor industry allows great scalability for these all subsequent qubit manipulation41–43 and information
systems.10–13 Long coherence times have been experi- readout.44,45 Previous studies of the charge stability di-
mentally achieved in both GaAs14–18 and Si19–22 semi- agram have focused on the classical Coulomb e↵ects us-
conductor devices. One-qubit and two-qubit manipula- ing primarily the capacitance circuit model,46–49 which
tions have been demonstrated in the GaAs quantum dot parametrizes the on-site and inter-site Coulomb inter-
systems.9,23–26 In silicon quantum dot systems, despite action as capacitances. The charge stability diagram is
difficulties such as the fabrication-induced disorder27 and found through the electrostatic energy for a given elec-
the valley degree of freedom,28–31 high tunability has tron configuration on the double dot system. Although
been reported32,33 as a milestone toward coherent con- this classical theory has satisfactorily explained many
trol of the qubit. aspects of experiments (especially the cases where the
Two main e↵ects are invariably present in all these quantum fluctuations are weak), there are cases where
experiments involving semiconductor quantum dots: the the quantum e↵ects, inevitably intertwining with the
classical Coulomb interaction and the quantum fluctu- Coulomb interaction, lead to (sometimes substantial) de-
ations. On one hand, due to Coulomb blockade,34 the formation of the charge stability diagram from that pre-
electron configurations in the quantum dot system are dicted by the capacitance circuit model.32,33,50,51 A the-
precisely controlled by electrostatic potentials (gate volt- ory capable of reconciling both the classical and quan-
ages). The details of electron configurations can be ex- tum e↵ects becomes necessary, which will not only help
tracted by an adjacent quantum point contact,35 and vi- us understand the quantum aspect of the charge stability
sualized as the charge stability diagram.36,37 On the other diagram, but will also substantially extend its usefulness.
2

Attempts to explain the quantum aspects of the charge In this paper, we address problems that have not been
stability diagram have been many. Most of them have fully explored in our previous publications.50,51 The pri-
employed the quantum-mechanical two-level model to mary goal of this paper is to provide a comprehensive pic-
study the tunneling of a single electron from one dot to ture of the theoretical framework that we have introduced
the other.37,52–55 From the probability crossover of the in Refs. 50,51, thereby completing this series of study on
two eigenstates (sometimes noted as the “excess charge”) the charge stability diagram of the double quantum dot
one can reliably extract the parameter for quantum- system. In Refs. 50,51, we have proposed that the quan-
mechanical electron hopping (or tunnel coupling) t. A tum fluctuations are essential for a in-depth understand-
somewhat similar work56 has studied the influence of the ing of the charge stability diagram, and have focused our
microscopically calculated exchange interaction on the attention on the tunnel coupling as it is manifestly the
charge stability diagram. On the other hand, interest most significant quantum parameter. However, other pa-
in applying Hubbard-like strongly correlated models to rameters such as the spin-exchange, pair-hopping and the
the quantum dot system57,58 arises in contexts related to occupation-modulated hopping are present as well, al-
the collective Coulomb blockade. In Ref. 59, a Hamil- beit with smaller amplitudes. One of the central aims of
tonian whose form is essentially very similar to our pro- this paper is to quantitatively estimate the e↵ect of all
posed generalized Hubbard model50 has been derived us- these quantum parameters, which define the full quan-
ing the so-called “pocket-state” method. However, the tum Hubbard model, on the charge stability diagram.
power of Hubbard-like models to describe the charge sta- Moreover, although we have outlined the microscopic the-
bility diagram has long been neglected until Gaudreau ory and performed calculation in some specific cases in
and collaborators60,61 have applied the Hubbard model Refs. 50,51, there are several problems that have not been
to triple, but not double, quantum dot system. sufficiently covered in the previous publications. First,
In a previous publication,50 we have introduced a gen- Ref. 50 has focused on the biquadratic model potential
eralized Hubbard model as a quantum generalization of and, since this model is rather special and involves only
the capacitance circuit model. We have shown that, a few parameters, a dimensionless quantity ⌘ combining
with all quantum fluctuations suppressed, the generalized the height of central potential barrier and the energy of
Hubbard model becomes the extended Hubbard model, electron states is identified to quantify the quantum fluc-
which can be mapped to the classical capacitance circuit tuations and completely determine the geometry of the
model exactly. The main advantage of our approach is charge stability diagram. In Ref. 51 the results calcu-
that all terms allowed by symmetry arguments are kept lated from both the biquadratic model and the Gaussian
in the Hamiltonian, which naturally accommodates all model confinement potentials have been compared to ex-
possible kinds of quantum fluctuations. We have further periments side-by-side. However a direct comparison of
recognized that these quantum e↵ects cannot be included the two models in terms of the Hubbard parameters is
in the capacitance model in any ad hoc manner. Be- missing. Although experimentally the electrostatic po-
cause the electron occupancies on each dot are no longer tentials dominate quantitatively, the detailed form of the
good quantum numbers, it breaks the basic assumption actual confinement potential is obviously unknown. Thus
of the capacitance model that the electrostatic energy of it is important to understand the consequences of di↵er-
the system is expressed in terms of electron occupancies ent choices of model potentials, under similar electro-
on each dot, which are assumed to be integer numbers. static situations, on the Hubbard parameters as well as
Therefore, our generalized Hubbard model can be viewed the charge stability diagram. We will therefore compare
as the quantum generalization of the classical capacitance two models, biquadratic and Gaussian, in this paper with
circuit model, with the individual electron occupancy on respect to the charge stability diagram. Second, all calcu-
each dot being a fluctuating quantum variable instead of lations presented in our previous publications50,51 as well
being a fixed number as in the capacitance model. as most of this paper are done assuming that each dot
contains a single orbital and is only allowed to hold up to
The quantum fluctuations may perturb the charge sta- two electrons. We will discuss the case where this con-
bility diagram in a substantial way. In Ref. 50 we dis- straint is lifted to allow three orbitals per site, and pro-
cussed in particular the rounding e↵ect of the boundary jecting to the single-orbital subspace we study the e↵ect
lines due to the electron hopping (tunnel coupling) t. A of additional higher-lying orbitals on this single-orbital
microscopic calculation was outlined to control the pa- subspace. Last, we will also discuss the role of the mi-
rameters in the generalized Hubbard model, and we have croscopic harmonic oscillator frequency in the Gaussian
shown how the charge stability diagram changes with model potential, as well as the e↵ect of external mag-
the variation of quantum fluctuations induced by the de- netic fields. All of these are new and not discussed in our
formation of the microscopic confinement potential. In previous publications.50,51
Ref. 51 we applied our theory to quantitatively explain
two set of experiments32,33 on the silicon system and cal- The remainder of this paper is organized as follows. In
culated, in particular, the tunnel coupling as a function Sec. II we present a general microscopic formulation of
of the height of the central potential barrier. The results the problem. In Sec. III we discuss in detail the general-
are found to well describe the experimental data after ized Hubbard model proposed in Ref. 50. In Sec. IV we
appropriately choosing model parameters. show charge stability diagrams calculated directly from
3

the generalized Hubbard model and discuss the e↵ect of the z axis, which couples to the electrons via the vector
various parameters on the stability diagram. In Sec. V we potential A. Here we have assumed that all electrons are
present a very detailed discussion on the microscopic cal- experiencing the same magnetic field (in actual experi-
culation, including the e↵ects of di↵erent choices of con- ments this may not be the case).
finement potential, the additional higher-lying orbitals, There are di↵erent choices of the detailed form of the
the role of the harmonic oscillator frequency in the Gaus- confinement potential, but we assume that it is approxi-
sian model potential and the e↵ect of the magnetic field. mately parabolic around the minima of the potential well
Sec. VI is the conclusion. The full matrix form of the so that the ground state single-particle wave functions
Hamiltonian and several cases with extreme (and possi- are harmonic oscillator states. In this case, the single-
bly unrealistic) parameters are shown in the appendices. particle Hamiltonian h (r) can be written as the sum
of the Fock-Darwin Hamiltonian HFD,j and some non-
harmonic potential Wj (r), h (r) = HFD,j + Wj (r), with
II. MICROSCOPIC THEORY: GENERAL
FORMULATION 1 2 1 2
HFD,j = [p eA (r)] + m⇤ !j2 (r rj ) , (5)
2m⇤ 2
We start with the general Hamiltonian of a system with 1 ⇤ 2 2
Wj (r) = V (r) m !j (r rj ) . (6)
N electrons, which consists of three parts: 2
Here, !j is the corresponding harmonic oscillator fre-
H (N ) = H0 + HC + HZ . (1)
quency for dot j. We denote r 0j = r r j and express
H0 is the sum of the single-particle Hamiltonian for each r 0j as (rj0 cos ✓j0 , rj0 sin ✓j0 ) in Cartesian coordinates. The
PN eigenfunctions of the Fock-Darwin Hamiltonian at the
electron, H0 = i=1 h (r i ). HC is the Coulomb interac-
jth dot are the Fock-Darwin states satisfying
tion between electrons,
↵ ↵
N
HFD,j 'j,nj ,mj rj0 , ✓j0 = ✏nj+ ,nj 'j,nj ,mj rj0 , ✓j0 .
X1 N
X ke2
HC = , (2) (7)
i=1 i0 =i+1
|r i r i0 |
where
where k = 1/(4⇡"0 "). HZ is the Zeeman energy,
✏nj+ ,nj = ~! j+ (nj+ + 1/2) + ~!j (nj + 1/2) , (8)
N
X 'j,nj ,mj rj0 , ✓j0 = exp [ i (eB/2~) (xj y yj x)]

HZ = g µ B B Siz z
= EB S , (3)
· mj ✓j0 Rnj ,mj rj0 , (9)
i=1

PN with
where EB = g ⇤ µB B, S z = z
i=1 Si . We neglect the 0 p
spin-orbit couplings for simplicity. We further note that mj ✓j0 = eimj ✓j / 2⇡ (10)
in our work, we do not include the interaction between
the electrons on the quantum dots and the environment. and
Therefore Eq. (1) is understood as describing the elec- q
|mj |
trons on the quantum dots only. In fact, many en- Rnj ,mj rj0 = 2nrj !/ nrj + |mj | !/lj0 rj0 /lj0
vironmental factors contribute to the decoherence pro- !
cess of the quantum dot system, such as the hyperfine ⇥ 02 2 ⇤ |mj | rj02
· exp rj /2lj0 Lnr 2 . (11)
interaction with the nuclear spin bath,40,42,43,62,63 the j lj0
coupling to background impurities64,65 and the phonon
modes.66–70 In this work, we concentrate on the electronic The index nj = 0, 1, · · · is the principal quantum num-
interaction in the quantum dot system and highlight its ber, mj = nj , nj + 2, · · · , nj 2, nj is the azimuthal
consequence on the charge stability diagram. Therefore quantum number, and nrj = (nj |mj |) /2 is the ra-
at this early stage of the theory we completely neglect dial quantum number. The pairs of quantum numbers
the coupling to the environment. The elucidation of this (nj , mj ) and (nj+ , nj ) are related by nj = nj + nj+ ,
|m |
problem is important in future studies. mj = nj nj+ . Lnr j denotes the Laguerre polynomi-
The single-particle Hamiltonian h (r) can be written p q j

as als, lj0 = ~/eB 4 1/4 + !j2 /!c2 , and

1 2
q
h (r) = [p eA (r)] + V (r) , (4) !j± = !j2 + !c2 /4 ± !c /2, (12)
2m⇤
where m⇤ is the e↵ective mass of the electrons and V (r) where the Larmor frequency !c = eB/m⇤ .
is the potential (typically double-welled in the case of These Fock-Darwin states can be used either to solve
the triplet/singlet spin qubit) confining electrons on the the Hamiltonian approximately, or to generate an equiv-
xy-plane. We allow for a magnetic field B applied along alent tight-binding model. In Ref. 50, we have obtained
4

a generalized Hubbard model using the Fock-Darwin up to two electrons. E↵ectively this means that we keep
states. There are infinitely many Fock-Darwin states, the s-orbital only, i.e. M = 2 and ⌫ = 1 in Sec. II. The
and it is impossible and unnecessary to keep all of them. case where the p-orbital comes into play will be discussed
For the M -dot system, we truncate the Fock-Darwin ba- in Sec. V B. As discussed in our previous publication,50
sis in each dot to ⌫ levels, i.e., |'l i = |'j,n i, where j labels the system can be described by a generalized form of
quantum dots (1  j  M ) and n denotes the number of the Hubbard model which retains all terms allowed by
energy levels that we have kept (1  n  ⌫). We notice symmetry: the total particle number N and the total spin
that the Fock-Darwin states within one dot are orthogo- Sz are conserved. The one-body part of the Hamiltonian
nal to each other, h'j,n | 'j,n0 i = 0, but those from di↵er- can be written as:
ent dots are in general not orthogonal: h'j,n | 'j,n0 i = 6 0. X X⇣ ⌘
We therefore build a new set of orthogonal bases by mak- H00 = ( µ i ni ) + tc†1 c2 + H.c. (19)
ing the transformation71 i

T where ni = c†i ci , µi denotes the chemical potential (or


(| 1i | 2i ··· | M ⌫ i)
level energy) on site i (i = 1, 2), and t denotes the inter-
1/2 T
=O (|'1 i |'2 i · · · |'M ⌫ i) . (13) site hopping (or tunnel coupling). The Zeeman term is
Here O is the overlap matrix (Ol,l0 = h'l | 'l0 i) generated EB
HZ = (n1" n1# + n2" n2# ) . (20)
by Fock-Darwin states in the single particle subspace. 2
The new basis | l i actually corresponds to, in a second-
The two-body part HC0 in general includes a Coulomb
quantized form, c†l, |0i where c†l, creates an electron on repulsion term HU :
site/orbital l with spin . After the orthogonalization,
the second quantized form of H0 and HC can be built, HU = U1 n1" n1# + U2 n2" n2# + U12 (n1" n2# + n1# n2" )
in a standard way, as + (U12 Je )(n1" n2" + n1# n2# ),
X (21)
H00 = Fl1 ,l2 c†l1 , cl2 , , (14)
l1 ,l2 , and a term HJ including spin-exchange (denoted by
X h (1) Je ), pair-hopping73–75 (denoted by Jp ), and occupation-
HC0 = Gl1 ,l2 ,l3 ,l4 c†l1 ," c†l2 ,# cl3 ," cl4 ,#
modulated hopping terms59,76 (denoted by Jt ):
l1 ,l2 ,l3 ,l4
X i
+
(2)
Gl1 ,l2 ,l3 ,l4 c†l1 , c†l2 , cl3 , cl4 , . (15) HJ = Je c†1# c†2" c2# c1" Jp c†2" c†2# c1" c1#
X (22)
Jti ni c†1 c2 + H.c..
The coupling parameters are i

Z In the case with non-zero t, U1 and U2 but all other


Fl1 ,l2 = dr ⇤l1 (r) h (r) l2 (r) , (16) parameters (U12 , Je , Jp , Jt1 , Jt2 ) being zero, one recov-
Z ers the form of the usual Hubbard model77 (on a lattice
(1) ke2 with only two sites). The inter-site Coulomb repulsion
Gl1 ,l2 ,l3 ,l4 = dr 1 dr 2 ⇤l1 (r 1 ) ⇤l2 (r 2 )
|r 1 r 2 | U12 was introduced in the study of the charge ordering in
· l3 (r 1 ) l4 (r 2 ) , (17) strongly correlated materials78 and the model including
Z the U12 term is usually termed as the “extended Hub-
(2) ⇤ ⇤ ke2 bard model”.79 In our previous work we have mapped
Gl1 ,l2 ,l3 ,l4 = dr 1 dr 2 l1 (r 1 ) l2 (r 2 )
|r 1 r 2 | the widely used capacitance model to the t = 0 case of
·[ l3 (r 1 ) l4 (r 2 ) l4 (r 1 ) l3 (r 2 )] . the extended Hubbard model and have argued that the
(18) extended Hubbard model is the minimal model that ex-
plains the experiment.50
Then the full Hamiltonian can be written down in the The spin-exchange and pair-hopping interaction73–75
second quantization form as H 0 = H00 + HC0 + HZ . have been studied in the context of orbital-selective Mott
We note that the di↵erence between our configuration transition.80 For atomic orbitals simple relations exist be-
interaction method and the traditional molecular orbital tween the parameters (e.g. for d-orbitals in free space,
method72 is that here we have transformed the eigen- U1 = U2 = U , Je = Jp = J, U12 = U 2J and due
value problem with non-orthogonal basis to a standard to orthogonality t = Jt = 0). However, in the quantum
eigenvalue problem with orthogonal basis. dot system these relations need not hold since the system
usually has much lower symmetry than the free space or
a lattice. The occupation-modulated hopping term Jt
III. GENERALIZED HUBBARD MODEL has not been considered much in the literature except in
certain aspects of superconductivity.76
We now consider the specific case of a coupled double The full Hamiltonian is a 16 ⇥ 16 matrix for the four-
quantum dot system with each dot capable of holding electron two-dot system (16 = 42 since each of the two
5

dots allows four possible quantum states). Because the are assumed to be linear combinations of (VR , VL ):60,61
total electron number N and total spin Sz are conserved,
it appears in a block-diagonal form. The details are pre- µ1 = |e| (↵1 VL + 1 VR ) + 1, (23)
sented in Appendix A. For a given (µ2 , µ1 ) one finds the µ2 = |e| (↵2 VR + 2 VL ) + 2; (24)
ground state by diagonalizing the Hamiltonian matrix.
The charge stability diagram, plotted on a plane with where 1 and 2 are constant energy shifts.
µ2 and µ1 as axes, shows how the ground state changes In our previous work50 we have presented a map-
as µ2 and µ1 are varied. Experimentally, the chemical ping between the capacitance model and the extended
potentials of the dots are controlled by the gate voltages Hubbard model, which suggests that the coefficients in
VR and VL and the charge stability diagram is plotted on Eqs. (23) and (24) are related to the parameters by51
the VR -VL plane.
In the case of t = Je = Jp = Jt1 = Jt2 = 0, the (U2 U12 ) U1 (U1 U12 ) U2
↵1 = 2 , ↵2 = 2 , (25)
ground state can be labeled as (n1 , n2 ) since ni (the elec- U1 U2 U12 U1 U2 U12
tron occupancy on dot i) is a good quantum number.
As one or more of the t and J parameters becomes fi- and 1,2 = 1 ↵1,2 . When the two dots are symmetric,
nite, ni ceases to be conserved and the ground state is a we define ↵ = ↵1,2 , = 1,2 and Eq. (25) reduces to50
linear combination of (n1 , n2 ) states. The phase bound-
U
aries between blocks with di↵erent N and Sz are clearly ↵= , (26)
defined as before since there is no mixing between them. U + U12
Within an (N ,Sz ) block we label the charge stability dia- and = 1 ↵.
gram using the (n1 , n2 ) state that dominates the ground We also note that the relation implied by Eqs. (23)
state. For example, t mixes the (1,0) and (0,1) states. and (24) is not a unitary transformation. Since we are
For µ1 > µ2 the (1,0) state is the majority state, and primarily interested in the lengths of segments on lines
vice versa. We then regard the line µ1 = µ2 as the phase µ1 µ2 = const., the lengths do conserve, up to a factor of
separator between two regimes in which (1,0) or (0,1) electron charge e, upon transforming (µ2 , µ1 ) ! (VR , VL )
are the majority states. Experimentally, these boundary in our theory. However, one must be cautious in fitting
lines separating di↵erent states within the same N block the experiments51 since the experimental values of the
are di↵use32,33 due to the hybridization of those states. gate voltages should sometimes be rescaled by a factor
To fully reproduce this fading e↵ect, a conductance cal- in order to apply our theory. This uncertainty between
culation involving excited states is needed.81 In our work the parameter sets (µ2 , µ1 ) and (VR , VL ) is unavoidable
we concentrate on the ground state feature so the fading within the scope of our theory and can only be resolved
of the boundary lines is beyond the scope of this work. with precise experimental information.
In this paper we assume U1 = U2 = U (except in
the discussion of Fig. 2), and Jt1 = Jt2 = Jt . This
means that there is a 1, 2-permutation symmetry and IV. RESULTS FROM THE GENERALIZED
the charge stability diagram is symmetric with respect HUBBARD MODEL
to the line µ1 = µ2 . Under this assumption plus a con-
dition that Jt = 0, the charge stability diagram has a In this section we present charge stability diagrams
particle-hole symmetry (symmetric with respect to the calculated directly from the generalized Hubbard model.
line µ1 + µ2 = U + 2U12 Je ), as can be seen from The parameters U , U12 , t, Je , Jp , and Jt are assumed
the matrix form of the Hamiltonian in Appendix A. In- to be independent of µ2 and µ1 . This is di↵erent from
clusion of Jt e↵ectively changes the values of t in high- the calculations from the microscopic theory where the
occupancy states, thus breaking the particle-hole symme- overlap integrals of wave functions at di↵erent locations
try. Of course allowing U1 6= U2 would lead the stability on the stability diagram are not guaranteed to be the
diagram not being symmetric with respect to the line same and the parameters indeed change slightly at dif-
µ1 = µ2 . ferent locations on the charge stability diagram. We also
The calculated stability diagram from the generalized assume EB = 0 in this section for simplicity. Throughout
Hubbard model will be discussed in the next section. this section we discuss the e↵ect of various parameters in
Analysis of the stability diagram shows that some of the the generalized Hubbard model on the charge stability
parameters can be extracted directly from experimen- diagram. To facilitate the comparison we plot all figures
tal plots. Alternatively, all the parameters of the model within the same µ2 , µ1 range. We also note that although
can be calculated from the microscopic theory [Eqs. (16), we set the unit of parameters to be meV (so that the re-
(17), and (18)] using a microscopic confinement potential sults are comparable with experiments and the results
model. Therefore, the generalized Hubbard model es- from microscopic theory), it can essentially be arbitrar-
tablishes a quantitative correspondence between the mi- ily chosen with the actual energetics being determined
croscopic theory and the experiments. Since the experi- by the details of microscopic confinement.
ments take VR and VL as the basic variable, it is useful The simplest case is the system of two decoupled quan-
to relate VR , VL to µ2 and µ1 . In the literature, (µ2 , µ1 ) tum dots with on-site Coulomb interaction only, i.e.
6

15 15
(2,2) (2,2)
(2,1) A' (2,1)
10 (2,0) B' 10 (2,0) A'
C'
B'
µ1 (meV)

µ1 (meV)
(1,2) C' (1,2)
C
(1,1) C
5 5 (1,1)
C' C'
(1,0) (1,0)
B C B C
0 A (0,2) 0 A (0,2)
(0,1) (0,1)
(0,0) (0,0)

0 5 10 15 0 5 10 15
µ2 (meV) µ2 (meV)

FIG. 1: Charge stability diagram calculated at U = 6.1 meV, FIG. 2: Charge stability diagram calculated at U1 = 5 meV,
U12 = 2.5 meV, and t = Je = Jp = Jt = 0. U2 = 7 meV, U12 = 2.5 meV, and t = Je = Jp = Jt = 0.

U12 = t = Je = Jp = Jt = 0. Its charge stability di- Eq. (26). We note that the readout of U12 is implied
agram has a checkerboard pattern which obviously does by the extended Hubbard model per se, while the value
not fit the experiment. As mentioned in Ref. 50, the ex- of U requires knowledge of the relation between (µ2 , µ1 )
tended Hubbard model (with hopping terms neglected) is and (VR , VL ). While the precise form of this relation is
the minimal model that explains the experiment. There- in general unknown, the mapping between the capaci-
fore we start with the case 0 < U12 < U , while keeping tance model and the Hubbard model provides one in the
t = Je = Jp = Jt = 0. A typical calculation is shown in simple cases where all quantum fluctuations vanish. In
Fig. 1. The parameters U and U12 are chosen to be the the cases where the mapping is not necessarily valid, one
same as Fig. 1 of Ref. 50 and the plot is identical to the must exercise caution.
solid line (t = 0 result) of Fig. 1(b) in Ref. 50. The case 0 < U12 < U considered here is physically
We examine the charge stability diagram in detail, in reasonable, which can be confirmed by calculations using
particular we are interested in the form of the phase the microscopic theory (see, e.g. Fig. 9). For complete-
boundaries and the coordinates of the “triple points”,47 ness we also consider the case with U12 > U and the
defined as the points on the diagram neighboring three results are shown in Appendix B.
di↵erent phases. In this case, all eigenenergies are lin- In Fig. 2 we show the result calculated allowing
early dependent on µi , U , and U12 . Therefore all phase U1 6= U2 . Both 1,2-permutation symmetry and particle-
boundaries are straight lines, which can be seen clearly hole symmetry are broken. We have the coordinates
from Fig. 1. The triple points are lettered on the fig- (µ2 , µ1 ) of the triple points as A(0, 0), B(U12 , U12 ),
ure with notations that exhibit symmetry considera- C(U2 , U12 ), C 0 (U2 +U12 , 2U12 ), C(U12 , U1 ), C 0 (2U12 , U1 +
tions. The 1,2-permutation symmetry does two things: U12 ), B 0 (U2 +U12 , U1 +U12 ), and A0 (U2 +2U12 , U1 +2U12 ).
First it ensures that triple-points A, B, A0 , B 0 lie on Note that 0
p the lengths of line segments AB and CC re-
the line µ1 = µ2 . Second it implies that the coordi- main 2U12 . Since the experimental plot of Ref. 23 does
nates of C and C 0 can be found by interchanging indi- not show the whole range of the stability diagram, the
cies 1 and 2 in the corresponding coordinates of C and explanation of U1 6= U2 is still consistent with the ex-
C 0 . The particle-hole symmetry implies that µ1,2 (X 0 ) = periment, i.e., an alternative fit allowing U1 6= U2 is still
U + 2U12 Je µ2,1 (X), where X = A, B, C, C. valid.
The precise coordinates (µ2 , µ1 ) of the triple points Fig. 3 presents a typical result calculated with a finite
can be readily calculated. We have A(0, 0), B(U12 , U12 ), t while Je = Jp = Jt = 0. The data shown here is iden-
C(U, U12 ), and C 0 (U + U12 , 2U12 ). (Coordinates of other tical to that shown as the red dotted line in Fig. 1(b) of
points can be found by symmetry.) We notice p that the Ref. 50. The main e↵ect of t is to mix states with the
length of the line segments AB = CC 0 = 2U12 and same total electron number N , except N = 0, 4. In this
BC = U . In particular, the length of AB and CC 0 does case one needs to identify the dominant (n1 , n2 ) state in
not change upon transforming from (µ2 , µ1 ) to (VR , VL ). the true ground state to label the corresponding region
Therefore from the experimentally measured charge sta- on the charge stability diagram, as mentioned above. In
bility diagram,23 one can read o↵ the value of U12 from Fig. 3 we use a star as a superscript of (n1 , n2 ) to in-
the length of the phase boundary between (1, 1) and dicate that this is the majority state in a linear com-
(0, 2), as what we have done in Ref. 50. In Ref. 50 we bination rather than the true ground state. Most of the
have also extracted the value of U from the slope of the phase boundaries are curved and the one separating (0, 0)
phase boundary between (1, 1) and (1, 2), according to and (1, 0)/(0, 1) complex can be identified as hyperbola
7

15 15
(2,2) (2,2)
(2,1)* A' (2,1)
(2,0)* (2,0) A'
10 B' 10
C' C'
(1,2)*
µ1 (meV)

µ1 (meV)
C B' (1,2)
C
5 (1,1)S * 5 (1,1)T
C'
(1,0)* (1,0) C'
B
B C C
0 A (0,2)* 0 A (0,2)
(0,1)* (0,1)
(0,0) (0,0)

0 5 10 15 0 5 10 15
µ2 (meV) µ2 (meV)
(a)
15
FIG. 3: Charge stability diagram calculated at t = 0.6 meV, (2,2)
U = 6.1 meV, U12 = 2.5 meV, Je = Jp = Jt = 0. (2,1)
(2,0)*
10

µ1 (meV)
(1,2)
µ1 µ2 = t2 . [The boundary between (2, 2) and (2, 1)/(1, 2)
5 (1,1)T
complex can be found by particle-hole symmetry condi-
(1,0)
tion.] The singlet combination of the (1, 1) states domi-
nates the (1, 1) component of the stability diagram [de- 0 (0,2)*
noted by (1, 1)S ⇤ in the figure], whose energy eigenvalue (0,0)
(0,1)
can be found by diagonalizing Eq. (A-6) where simple
analytical formula does not exist. The coordinates of A 0 5 10 15
and B can be found analytically: p µ2 (A) = µ1 (A) = t, µ2 (meV)
µ2 (B) = µ1 (B) = t+(U +U12 )/2 4t2 + (U U12 )2 /4, (b)
0
while the coordinates of C and C cannot be found ana-
lytically. It is evident, however, that CC 0 is stretched by FIG. 4: Charge stability diagram calculated at Je = 1 meV
the introduction of tunnel coupling t. and (a) Jp = 0, (b) Jp = 1 meV. U = 6.1 meV, U12 = 2.5
meV. t = Jt = 0.
To understand the e↵ect of Je we first switch o↵ t.
Fig. 4(a) shows a typical charge stability diagram calcu-
lated at 0 < Je < U12 with t = Jp = Jt = 0. Since 15
the triplet combination of (1,1) state [labeled by (1, 1)T ] (2,2)
(2,1)*
decouples from Eq. (A-5), all eigenenergies are again lin- (2,0)*
10
ear functions of U , U12 , and Je thus the phase bound-
(1,2)*
µ1 (meV)

aries are linear. The coordinates (µ2 , µ1 ) of the triple


(1,1)
points are A(0, 0), B(U12 Je , U12 Je ), C(U, U12 Je ), S*
5 T
and C 0 (U
p + U12 , 2U12 Je ). The lengthp of line segments (1,0)* S*
AB = 2(U12 Je ), and CC 0 = 2U12 . If the whole
charge stability diagram is measured, one can read o↵ the 0 (0,2)*
value of Je immediately from the di↵erence between the (0,1)*
(0,0)
two segments AB and CC 0 (the e↵ect of other parameters
t, Jt must be assumed to be small, though). Typically Je 0 5 10 15
is smaller than U or U12 by at least an order of magni- µ2 (meV)
tude. Therefore the Je = 1 meV (with U = 6.1 meV,
U12 = 2.5 meV) shown in Fig. 4(a) is slightly exagger-
ated in order to make its e↵ect clear. In actual systems FIG. 5: Charge stability diagram calculated at U = 6.1 meV,
one expects the e↵ect of Je on the charge stability dia- U12 = 2.5 meV, t = 0.6 meV, Je = Jp = 0.2 meV. Jt = 0.
gram to be very small. For a brief discussion of the case
Je U12 , see Appendix B.
The pair hopping interaction mixes the (2, 0) and (0, 2) and (2, 0)⇤ region are indeed curved but the curvature
states. If t + Jt 6= 0, it in addition mixes the (1, 1) sin- is hardly detectable. The boundaries between N = 0, 1
glet with the (2,0) and (0,2) states, as can be seen from blocks and N = 3, 4 blocks remain the same. The cor-
Eq. (A-6). Fig. 4(b) shows a case with a finite but small rection to the coordinates of the triple points C and C 0
Jp . Comparison to Fig. 4(a) reveals that the change im- (and their mirror C and C 0 ) can be shown to be of order
posed by Jp is very small. The boundaries of (0, 2)⇤ Jp2 /(U U12 ).
8

15 At the end of this section we turn on all parameters


(2,2) that we have discussed and consider the e↵ect of Jt .
(2,1)*
(2,0)* Fig. 6(a) shows a case with all parameters being non-
10
zero except Jt . The figure is essentially very similar to
(1,2)*
µ1 (meV)
Fig. 3 since the values of Je and Jp are very small and
5 (1,1)S * have little e↵ect. The particle-hole symmetry is evident.
(1,0)* Fig. 6(b) shows the e↵ect of a relatively large Jt . First,
the N = 0, 1 parts of the stability diagram remain un-
0 (0,2)* changed. Second, the N = 2 parts [(0, 2)/(1, 1)S/(2, 0)
(0,1)* complex] are slightly stretched towards the N = 3 block.
(0,0)
The N = 3 block [(2, 1)/(1, 2) complex] is deformed
0 5 10 15 and both of its boundaries are substantially rounded and
µ2 (meV) smoothened. We expect that this is due to the change
(a) of an e↵ective t to t + Jt (N = 2) or t + 2Jt (N = 3).
15 Overall, it is clear that Jt breaks the particle-hole sym-
(2,2)
metry, making the boundaries between high-occupancy
(2,1)*
states smoother.
10 (2,0)*
(1,2)*
µ1 (meV)

V. RESULTS FROM THE MICROSCOPIC


5 (1,1)S * THEORY
(1,0)*
0 (0,2)* As discussed in Refs. 50 and 51, the microscopic
(0,1)* calculation39,72,82–87 is required to constrain the param-
(0,0)
eters of the Hubbard model in the physically relevant
0 5 10 15 regime. In other words, although the Hubbard model
µ2 (meV) by itself can have arbitrary parameters, a given physical
(b) double-dot system is restricted by the realistic confine-
ment potential which would severely restrict the physi-
FIG. 6: Charge stability diagram calculated at (a) Jt = 0 (b) cal Hubbard parameters for the system. The details of
Jt = 0.5 meV and U = 6.1 meV, U12 = 2.5 meV, t = 0.6 the application of microscopic theory to our model are
meV, Je = Jp = 0.1 meV. described in Sec. II. In this section we discuss the ap-
plication of the microscopic theory to our problem. We
shall primarily use two di↵erent models of potential: a bi-
Comparison of Fig. 3 and Fig. 4(a) shows that the quadratic form [Eq. (28)] and a Gaussian form [Eq. (29)].
(1,1) component of the stability diagram has interesting Both of these model potentials are reasonably realistic
behavior: for t Je the singlet combination dominates and used extensively in describing double-dot systems.
the (1,1) block, while for t ⌧ Je the triplet combination The biquadratic potential is defined as39,88,89
dominates. This can be understood from the eigenvalues n m! 2
0
of Eq. (A-5). The triplet has energy µ1 µ2 + U12 Je VQ (x, y) = Min [(x + a)2 + y 2 ] µ1 ,
while the singlet has energy µ1 µ2 + U12 + Je plus 2 (28)
m!02 o
contributions due to t + Jt . When t + Jt is small the [(x a)2 + y 2 ] µ2 ,
triplet has an energy 2Je lower than the singlet, thus be- 2
coming the ground state. However when t + Jt is large (note that here the two dots are assumed to be symmet-
the correction due to t + Jt exceeds 2Je , making the sin- ric) and the Gaussian potential reads
glet the dominating ground state. One expects an inter- " #
2
mediate value of (t + Jt )/Je such that the singlet and (x + a) + y 2
VG (x, y) = V1 exp 2
triplet co-exists. This situation is shown in Fig. 5, where ld1
the boundaries separating the singlet and the triplet are " #  2 (29)
2
plotted as dashed lines. The condition that the singlet (x a) + y 2 x + y2
V2 exp 2 + Vb exp .
state dominates can be analytically obtained as: ld2 lb2

1 Note that Eqs. (28) and (29) have slightly di↵erent forms
(t + Jt )2 > Je (Je + Jp + U U12 ), (27) than that in Ref. 51, in order to facilitate the comparison
2
of the e↵ect of di↵erent potentials on the charge stability
which is typically satisfied in most microscopic calcula- diagram in Sec. V A. This will be explained in Sec. V A.
tions of double quantum dots. The condition that the Besides, we also consider the e↵ects of the additional p-
triplet dominates without having singlet regimes on its orbitals (Sec. V B), and the role of the harmonic oscilla-
sides cannot be expressed analytically. tor frequency in the calculation with the Gaussian model
9

10
25
ε = 0 meV biquadratic

V (x,0) (meV)
0 15 20
Gaussian

μ1 (meV)
μ1 (meV)
15
-10 10
10
-20 5
5
-100 -50 0 50 100 0 0
(a) x (nm) 0 5 10 15 0 5 10 15 20 25
10 (a) μ2 (meV) (c) μ2 (meV)
20
ε = 10 meV
V (x,0) (meV)

25
0
15 20

μ1 (meV)

μ1 (meV)
-10
15
10
-20 10
5
-100 -50 0 50 100 5
(b) x (nm) 0 0
0 5 10 15 20 0 5 10 15 20 25
(b) μ2 (meV) (d) μ2 (meV)
FIG. 7: (Color online) Profiles of biquadratic (black solid
line) and Gaussian (red dashed line) model potentials under
two di↵erent values of detuning energy " = µ2 µ1 . For the FIG. 8: (Color online) Charge stability diagrams calculated
biquadratic potential, ~!0 = 3.956 meV, a = 33 nm. The from biquadratic [panels (a) and (b)] and Gaussian [panels
parameters of the Gaussian potential are derived from the (c) and (d)] model potential di↵erent inter-dot distances. The
biquadratic one. The gray dotted lines denote the energy parameters for the biquadratic model are ~!0 = 3.956 meV,
reference values of the chemical potential. a = 33 nm (panels (a) and (c)), and a = 28 nm (panels (b)
and (d)). The white crosses [indicating the points at which
the comparison of Hubbard parameters (Fig. 9) is conducted]
are located at µ1 = µ2 = 9.743 meV for panels (a) and (c),
potential (part of Sec. V C). All calculations mentioned
and µ1 = µ2 = 10.095 meV for panels (b) and (d).
above are done without magnetic field, and we consider
the influence of the magnetic field in Sec. V C.
0.02
The dielectric constant for GaAs " = 13.1 and the

Je = Jp (meV)
14
e↵ective mass (m⇤ = 0.067me ) are used throughout the biquadratic
U (meV)

microscopic calculation. 12 Gaussian


0.01
10

8
0
A. Influence of model confinement potentials 28 29 30 31 32 33 28 29 30 31 32 33
(a) a (nm) (d) a (nm)
In the microscopic calculation, the confinement poten- 2
0.04
U12 (meV)

tials in the quantum dot systems are usually represented


Jt (meV)

by models. Since one has several choices of models, it is 1.8 0.02


useful to understand the consequences of di↵erent model
potentials on the charge stability diagram. A similar
1.6 0
comparison has been done for the exchange coupling in 28 29 30 31 32 33 28 29 30 31 32 33
Ref. 90, but in light of the generalized Hubbard model we (b) a (nm) (e) a (nm)
would like to understand the influence on the overall ge-
0.03
ometry of the charge stability diagram and the Hubbard 0.2
t (meV)

parameters. We match the two models of potentials ac- 0.02


t/U

cording to their minima and the central barrier (see Fig. 7 0.1
0.01
for examples at two di↵erent values of detuning energy),
because these quantities are directly related to the elec- 0 0
28 29 30 31 32 33 28 29 30 31 32 33
trostatic potentials that one can experimentally control. (c) a (nm) (f) a (nm)
The biquadratic model potential is chosen first because it
contains only two parameters. The chemical potentials
FIG. 9: (Color online) Parameters in the generalized Hub-
µ1,2 enter the problem in two ways: first, they appear bard model as functions of half of the inter-dot distance a,
in the formal definition of the model [Eq. (28)]; second, calculated at the white crosses shown in Fig. 8, using both
the wave function solution to the Schrödinger equation biquadratic and Gaussian potentials.
implies the values of the chemical potential [see Eqs. (14)
10

and (16)]. In our Hund-Mulliken calculation for the bi- values directly implemented in the model. This tilting
quadratic potential, the chemical potentials defined in e↵ect is not present in the Gaussian model calculation
these two ways are approximately equal; while for the [Fig. 8(c) and (d)], as the chemical potential values are
Gaussian model the chemical potential values calculated chosen to be the ones calculated from the wave functions.
from the wave functions are di↵erent from V1 and V2 de-
noted in Eq. (29). Since we want to directly compare We would also like to understand the Hubbard pa-
the results from the two models under the same electro- rameters as functions of half the inter-dot distance a
static situations, we tune parameters (V1,2 ) such that the for the two models. We choose one particular point on
calculated chemical potential values match. This is the each charge stability diagram in a consistent way to com-
main reason that Eq. (29) is written in a slightly di↵er- pare the parameters. These points are represented as the
ent way than Ref. 51. Since the Gaussian model contains white crosses in the (1,1) components of the charge sta-
more parameters than the biquadratic model, some of the bility diagrams in Fig. 8: they are defined as the mid-way
parameters in the Gaussian model have to be fixed: we point of the (1,1) segment of line µ1 = µ2 for biquadratic
fix lb = 20nm. The reference energy values (shown as model calculations with di↵erent values of a, and those
light gray dotted lines in Fig. 7) correspond to the zero coordinates are inherited in the Gaussian model calcula-
of the chemical potential in Fig. 8. A close examination tions, in spite of the fact that they are not in the center of
of Fig. 7 reveals that the main di↵erence between the two (1,1) regime any more (which is, again, a consequence of
models is that the Gaussian model has narrower potential the more pronounced localization e↵ect of the Gaussian
wells. This implies stronger on-site Coulomb interaction potential). In Fig. 8 the white crosses are actually at the
and weaker inter-dot tunnelling, as shall be seen in the same location for panels (a) and (c), as for panels (b) and
following discussion. (d). We have calculated the Hubbard parameters at these
In Fig. 8 we compare the microscopically calculated points for a range of a values (which are bracketed by
charge stability diagram for di↵erent models of poten- those shown in Fig. 8), and the results are summarized in
tials and inter-dot distances. Note that all charge stabil- Fig. 9. Fig. 9(a) reveals that the on-site Coulomb repul-
ity diagrams shown in this section have an overall energy sion U is independent of the inter-dot distance, indicating
shift relative to those shown in Sec. IV due to the zero that the wave functions are well localized. U is larger in
point energy of the harmonic oscillator states. Panels (a) the Gaussian model than in the biquadratic model by
and (b) show the results calculated from the biquadratic about 35%, which is a consequence of narrower potential
potential, while the results from the Gaussian model are wells in the Gaussian model: the electron wave function
shown in panels (c) and (d). The value of a for panels is more localized than the biquadratic model. For the
(a) and (c) is a = 33 nm, while that for panels (b) and same reason, all parameters characterizing the inter-dot
(d) is a = 28 nm. The parameters a and !0 (given in the interactions are smaller in the Gaussian model than in the
caption) are for biquadratic model, while those for the biquadratic model, as quantitatively shown in Fig. 9(b)-
Gaussian model are derived hereby as explained above. (f). Fig. 9(b) shows that the inter-dot Coulomb inter-
We see that the overall shapes are similar for all panels. action is only slightly decreased in the Gaussian model,
Panels (c) and (d) are shown with a larger x- and y-range while Fig. 9(c) and (e) show that the hopping (tunnel
than panels (a) and (b), because the Gaussian model has coupling) as well as the occupation-modulated hopping
a larger on-site Coulomb interaction than the biquadratic for the Gaussian model is decreased by more than 40%
model at the same interdot distance, shifting the phase from that for the biquadratic model. This implies that
boundaries to higher energies. This is consistent with the di↵erence in the dimensionless ratio t/U can only be
the qualitative argument above that the narrower poten- more pronounced, as shown in Fig. 9(f). In Fig. 9(d),
tial wells of the Gaussian model lead to larger on-site one sees that the spin-exchange and pair-hopping inter-
Coulomb interactions. Moreover, the rounding e↵ects action is substantially smaller in the Gaussian model: it
near the triple points in panels (b) and (d) are more pro- is decreased by an order of magnitude from that for the
nounced than panels (a) and (c), which originates from a biquadratic model.
smaller inter-dot distance 2a, leading to stronger quan-
tum fluctuations. These arguments will also be confirmed Our results indicate that in modeling the double quan-
in the calculated Hubbard parameters below. We have tum dot systems, the particular choice of model confine-
noted that for the biquadratic model, the chemical po- ment potential leads to quantitative changes in the Hub-
tential values derived from Eqs. (14) and (16) are some- bard parameters, although the qualitative behaviors are
times slightly di↵erent from the values in Eq. (28). This similar. In a realistic study of these systems, the first-
implies that in Fig. 8(a) and (b) some of the boundary principles calculation of the exact form of the confine-
lines, which are parallel to the x- or y-axis in the Hubbard ment potential from Poisson equation would be required.
model calculation, are slightly tilted relative to the axes. However, this demands higher precision in both experi-
This e↵ect is more pronounced for a smaller inter-dot dis- ments and theoretical calculations, which should be pos-
tance 2a [Fig. 8(b)], since a stronger overlap between the sible in future studies if more precise information about
wave functions in the two dots leads to a larger devia- the lithographic gates creating the confinement potential
tion of the calculated chemical potential values from the is available.
11

10
20 20

U eff (meV)
μ1 (meV)

μ1 (meV)
10 10
5
s
s+p
0 0 0
0 10 20 0 10 20 -20 -10 0 10 20
(a) μ2 (meV)
(b) μ2 (meV) (a) ε (meV)
6
60

U12eff (meV)
4

50
2

40
0
μ1 (meV)

-20 -10 0 10 20
(b) ε (meV)
30

FIG. 11: (Color online) E↵ective on-site Coulomb interaction


20
[panel (a)] and inter-site Coulomb interaction [panel (b)] of
the results shown in Fig. 10 as a function of the detuning
energy " = µ2 µ1 calculated keeping s-orbitals only (black
10
solid line) and s- and p- orbitals (red dashed line). The traces
where calculations take place are shown as white diagonal
0
lines in Fig. 10(a) and (b). As " is swept from the top left
0 10 20 30 40 50 60 corner to the bottom right corner, the dominating ground
(c) μ2 (meV) state is changed from (2,0) to (1,1), and then to (0,2). The
black dotted lines and red dash-dotted lines denote the phase
boundaries (where the probabilities of states cross) in the case
FIG. 10: (Color online) Charge stability diagrams calculated of s-orbitals only and s- and p-orbitals, respectively.
from the Gaussian potential using (a) s-orbitals only and (b)
s- and p-orbitals. The parameters corresponding to the bi-
quadratic potential are a = 30 nm and ~!0 = 3.956 meV. considered. This can be regarded as taking M = 2 but
Panel (c) shows the full charge stability diagram including all
⌫ = 3 in Sec. II. The full charge stability diagram has
49 components, while panel (b) shows the stability diagram
in the two-electron regime. Panel (b) is a replica of the area 49 components ranging from (0, 0) to (6, 6). We focus
enclosed by the white frame in the lower left corner of panel on the part with population up to two electrons per dot
(c). (2, 2), and study how some of the Hubbard parameters
are renormalized by the additional orbitals.
Fig. 10 compares the charge stability diagram calcu-
lated keeping s-orbitals only and that calculated keeping
B. E↵ect of additional p-orbitals
both s- and p-orbitals. Panel (a) is the result of keeping
the s-orbitals only. Panel (c) shows the full charge sta-
All of the calculations shown in this paper so far, as bility diagram for the case with both s- and p-orbitals,
well as those shown in Refs. 50 and 51, bear the same including all 49 components. Its lower left corner corre-
feature that each quantum dot is only allowed to hold up sponds to the two-electron regime and we replicate the
to two electrons. This means that we have taken M = 2 area enclosed by the white frame as panel (b) and com-
and ⌫ = 1 in the general discussion in Sec. II, keeping pare that to panel (a) side-by-side. Strong similarity be-
the s-orbital (denoting the lowest-lying state) only. In tween panel (a) and panel (b) is evident and the di↵er-
the discussion of silicon devices,51 a multi-electron multi- ences are only in the details. In Fig. 10(b), the stability
band situation is encountered. In Ref. 51 we have reduced diagram are slightly shifted toward the origin, and the
the full multi-band problem to an e↵ective two-electron diamond of the (1, 1) component shrinks its size. This
regime and have argued that the multi-band e↵ect can be means that the Coulomb interaction U is renormalized
regarded as a renormalization of Hubbard parameters. In to a smaller value by the additional orbitals.
this subsection, we examine this idea in a slightly di↵er- To quantify this idea we study the “e↵ective” on-site
ent context: we consider the additional p-orbitals (which Coulomb interaction (U e↵ ) and inter-site Coulomb inter-
e↵
are doubly degenerate and can be labeled as px and py action (U12 ) in Fig. 11. The purpose of studying the
orbitals), the lowest excited states except the s-orbitals e↵ective Coulomb interactions is that we want to intro-
12

duce a method which may be useful for a reduction of the 9.24 and 9.68 meV. This indicates that the main e↵ect of
general multi-band problem to the two-electron regime. the p-orbital is a renormalization of the e↵ective on-site
For the case studied here (retaining three orbitals) it is Coulomb interaction to a slightly smaller value, which
still possible to calculate all parameters in the general- has a very weak dependence on chemical potential val-
ized Hubbard model, but for a general multi-band prob- ues. The results shown in the range |"| < 10 meV are
lem this calculation is cumbersome and unnecessary, as meaningless and will not be discussed here. Fig. 11(b)
e↵
all the qubit manipulations are done in the two-electron shows the calculated U12 and the range |"| < 6 meV is
regime and it is sufficient to study the e↵ect of the addi- directly relevant here. For the case with s-orbital only,
e↵
tional band on this regime. U12 = 1.88 meV, consistent with the value in the Hub-
For a general multi-band problem, the total electron bard model U12 = 1.88 meV. For the case with both s-
e↵
number N and total spin Sz should still be conserved. and p-orbitals, U12 = 2.00 meV. Therefore the additional
The full Hamiltonian is block-diagonalized as per {N, Sz } p-orbitals does not change e↵ective value of U12 apprecia-
and we denote the nth lowest energy eigenvalues in each bly. The calculated values of U e↵ and U12 e↵
are also con-
{N,S } sistent with that extracted directly from Fig. 10(a) and
block as En z . Then when the (2, 0) or (0, 2) is the
dominant ground state, the e↵ective on-site Coulomb in- (b). The main result in this subsection is that the e↵ect of
teraction can be defined as the energy di↵erence between additional orbitals can be regarded as a small renormal-
the doubly occupied state and the singly occupied state: ization of the Hubbard parameters and the main physics
is indeed captured in the two-site one-orbital Hubbard
{2,0} {1,± 12 }
U e↵ = E1 2E1 . (30) model.
We note, that in Fig. 11(b) the U12 value extracted
Note that here we have used the fact that the two dots are from the range |"| > 12 meV is actually related to the
symmetric. When they are asymmetric, Eq. (30) gives Coulomb interaction between s and p orbitals (which can
the Coulomb interaction on the first dot (U1 ) when cal- be denoted as Usp ), and this information is in principle
culated at chemical potential values such that (2,0) dom- encoded in the stability diagram of Fig. 10(c). In fact, the
inates the ground state, and it gives U2 when calculated full characterization of the three-orbital charge stability
in cases where the (0,2) dominates the ground state. diagram in Fig. 10(c) shows rich behavior: the compo-
Moreover, for cases where (1, 1) dominates the ground nents of the stability diagram generically have smaller
e↵
state, the e↵ective inter-site Coulomb interaction U12 is area for highly occupied states than the one-orbital sub-
defined as the di↵erence between the energy of the state space we are considering, for example the (3,3) and
with each dot holding one electron and the total energy (5,5) states spans a substantially smaller range on the
of states that one electron is allowed to fill one and the charge stability diagram. This indicates that the on-site
other dot respectively. This means that in the {N = Coulomb interaction is weaker for higher-lying orbitals
1, Sz = ± 12 } block both the lowest and the second lowest than the s orbitals. In fact, as the p orbitals are popu-
energy eigenvalues are involved, as we want the electron lated, the e↵ective central potential barrier between the
to occupy a di↵erent dot after one of the two dots is two dots are e↵ectively lower since p orbitals themselves
e↵
occupied. Then U12 is expressed as have higher energies, and the corresponding wave func-
{2,±1} {1,± 12 } {1,± 12 } tions are less localized. This naturally leads to the fact
e↵
U12 = E1 E1 E2 . (31) that on-site Coulomb interaction for higher-lying orbitals
is relatively smaller. Moreover, through the experimen-
In Fig. 10(a) and (b) we have drawn two diagonal lines
tal observation of the charge stability diagram, one can
in the N = 2 block of the charge stability diagram, as
in principle infer the e↵ective microscopic confinement
this is the regime which attracts the most interest. Along
potential. The full understanding of this multi-orbital
these lines we have calculated the e↵ective on-site and
problem is, however, beyond the scope of the present
inter-site Coulomb interaction as functions of the detun-
work and is an interesting topic for future investigation.
ing energy " = µ2 µ1 , and the results are shown in
Fig. 11. The black solid lines in Fig. 11 show the case
with s-orbitals only, corresponding to Fig. 10(a), while
the red dashed lines show the case with s- and p-orbitals, C. Harmonic oscillator frequency in the Gaussian
corresponding to Fig. 10(b). One must be cautious when model and the e↵ect of external magnetic field
interpreting Fig. 11, as the definitions of the e↵ective
Coulomb interaction Eqs. (30) and (31) is only valid In this subsection we study the role of the harmonic
for a selective range of parameters that some particu- oscillator frequency in the Gaussian model and the e↵ect
lar states dominates the ground state. In Fig. 11(a), the of external magnetic field.
values shown in the range |"| > 10 meV gives the e↵ective There are intrinsic difficulties when applying the Gaus-
on-site Coulomb interaction: For the case with s-orbital sian model potential in the Hund-Mulliken approxima-
only, U e↵ = 10.38 meV. This is consistent with the value tion: as the chemical potentials change their values, the
in the Hubbard model U = 10.38 meV. For the case with two potential wells are deformed such that a harmonic-
both s- and p-orbitals, U e↵ changes slightly with the de- oscillator-type expansion of the confinement minima
tuning energy but is bound to a narrow range between yields considerably di↵erent values of the harmonic os-
13

25 25 ever, it is useful to briefly study how the charge stability


diagram changes as the harmonic oscillator frequency is
20 20
varied.
μ1 (meV)

μ1 (meV)
15 15

10 10 Fig. 12(a) shows the identical result as that shown in


Fig. 10(a). The harmonic oscillator frequency (energy)
5 5
has been chosen at the locus denoted by a white cross:
0 0 ~!0G = 6.487 meV. In the calculation of Fig. 12(b),
0 5 10 15 20 25 0 5 10 15 20 25
(a) μ2 (meV) (c) μ2 (meV) we artificially multiply the harmonic oscillator energy by
a factor of 1.5, thus further localizing the Fock-Darwin
25 25 wave functions but leaving the confinement potential un-
20 20 changed. Beside an overall shift in energy (due to the
larger ~!0 ), the (1, 1) component of the charge stabil-
μ1 (meV)

μ1 (meV)

15 15
ity diagram clearly expands its area, meaning that in
10 10 this case the on-site Coulomb interaction is increased.
5 5
In Fig. 12(c), the charge stability diagram is calculated
using Fock-Darwin states whose harmonic oscillator fre-
0
0 5 10 15 20 25
0
0 5 10 15 20 25
quencies are carefully calculated according to the precise
(b) μ2 (meV) (d) μ2 (meV) form of Gaussian model at each given point on the charge
stability diagram. We see that the lower left components
(N < 2) are quite similar to that of Fig. 12(a), while
FIG. 12: (Color online) Charge stability diagrams calculated
for N > 2 the spatial variation of the on-site Coulomb
from Gaussian potential with a = 30 nm and ~!0 = 3.956
meV. (a) B = 0, the zero-point energies of Fock-Darwin states interaction U is evident. The shape of the (1,1) com-
are chosen according to the microscopic confinement potential ponent of the charge stability diagram is abnormal: it
at µ1 = µ2 = 9.938 meV (indicated by the white cross) with shrinks its area in proximity to the N = 3 block, while its
~!0G = 6.487 meV. (b) B = 0, the zero-point energies of boundaries with the N = 1 block remain approximately
Fock-Darwin states are ~!0G = 9.731 meV. (c) B = 0, the the same. As a consequence, pairs of the boundaries
harmonic oscillator frequencies of the Fock-Darwin states are of the (1,1) regime are not parallel, in contrast to what
chosen according to the microscopic potentials for each given observed in other calculations and in most experiments.
µ1 and µ2 . (d) B = 6 T, the zero-point energies of Fock- This means that the case considered here, varying the
Darwin states are ~!0G = 6.487 meV. harmonic oscillator frequency of the Fock-Darwin states
as the potential wells are changing, is less relevant to ex-
periments that we have considered. The good correspon-
cillator frequencies. Since the harmonic oscillator fre- dence between Fig. 12(a) and (c) means that the approxi-
quency is the central variable for the Fock-Darwin states, mation that we have employed is reasonable; Rather, the
its variance leads to a change in the Hubbard parameters ~!0 artificially chosen for Fig. 12(b) is too large to accu-
as well as deformation of the charge stability diagram. rately reflect the actual solution. We therefore conclude,
Moreover, the values of the chemical potentials them- since the present calculations already have quantitative
selves rely on the calculated Fock-Darwin states, which correspondence with experiments, that the method we
further depend on how the harmonic oscillator frequency have used in previous sections using Fock-Darwin states
is selected. In our work we have circumvented these dif- with fixed harmonic oscillator frequency is most suitable
ficulties by closely following the biquadratic potential to for our work.
determine the corresponding parameters of the Gaussian
model, as noted in Sec. V A. A fixed harmonic oscilla- In Fig. 12(d) we show a calculation with parameters
tor frequency has been used throughout the calculation same as Fig. 12(a) except that there is a finite external
for a particular charge stability diagram, which corre- magnetic field B = 6 T. The reason that we group this
sponds to a point in the (1, 1) regime where the potential result with others discussed in this section is that the
wells are symmetric. The motivation of this constraint is magnetic field can be regarded to have very similar ef-
that the biquadratic model potential, with !0 fixed, has fect as increasing the harmonic oscillator frequency, as
been shown to well describe the experiment with much can be seen by similarities between Fig. 12(b) and (d).
less number of independent variables than the Gaussian In fact, this can be straightforwardly understood from
model.50,51 We believe that in order to gain physical in- Eq. (12): A finite magnetic field leads to a finite Larmor
sight it is sufficient to manipulate the few most important frequency !c , which consequently enters the energy of
variables. A precise treatment for the Gaussian model, the Fock-Darwin states in a similar way as the harmonic
with the harmonic oscillator frequencies appropriately oscillator frequency (at dot j) !j . Therefore, a magnetic
vary with the deformation of the potential well, gives field e↵ectively localizes the electron wave function, lead-
results that are not quite di↵erent from the calculations ing to stronger on-site Coulomb interactions and weaker
keeping !0 fixed in comparison to experiments. How- inter-site Coulomb interactions.
14

microscopic are completely suppressed, the model is mapped exactly


theory to the capacitance circuit model. On the other hand,
quantum fluctuations perturb the charge stability dia-
gram, which is observed both in experiments32,33 and in
generalized theoretical calculations (see Sec. IV). The intrinsic as-
Hubbard model experiments
sumption of the capacitance circuit model is that the
electrostatic energy of a given system are functions of in-
teger electron occupancies on the dots. We have argued
capacitance that this assumption eliminates the possibility of directly
model including quantum fluctuations in the capacitance cir-
cuit model, as the electron occupancies cease to be good
FIG. 13: Illustration of the connections between di↵erent quantum numbers in the presence of quantum fluctua-
components discussed in this paper as well as Refs. 50 and tions. Therefore, the generalized Hubbard model is the
51. most direct quantum generalization of the capacitance
circuit model.
The generalized Hubbard model contains several pa-
VI. CONCLUSION rameters. Although their values are not determined by
symmetry considerations, they can not be arbitrarily cho-
In this paper, along with our previous publications sen, otherwise unphysical charge stability diagrams may
Refs. 50 and 51, we have developed a theoretical frame- occur (see Appendix B). On one hand, we calculate all
work treating the coupled double quantum dot system parameters of the generalized Hubbard model through
with a generalized form of Hubbard model. The aim of the microscopic theory, detailed in Sec. II. The micro-
our work is to establish conceptual and quantitative con- scopic calculation involves a numerical solution to the
nections between experiments, existing theoretical ap- many-electron Schrödinger equation under some model
proaches and the new theory (see Fig. 13), thus present- confinement potential. The electron wave functions are
ing a unified and transparent description of our current approximated by linear combinations of the Fock-Darwin
understanding of the double quantum dot system and states through the configuration interaction method. On
allowing possibilities for studying more complicated sit- the other hand, some of the parameters can be extracted
uations in future work. The central object of these stud- from experimentally measured charge stability diagrams,
ies, which is invariably present in all these connections, provided that the resolution is sufficiently high. There-
is the charge stability diagram, the visualization of elec- fore the generalized Hubbard model acts as a bridge
tron configurations in the double dot system. Given the which quantitatively connects the experiments and the
essential role of semiconductor double quantum systems microscopic theory. These considerations, together with
in solid state quantum computer architectures, our use the connections to the capacitance model, are shown
of the generalized Hubbard model in describing the sys- symbolically in Fig. 13. Fig. 13 implies that our work in
tem is natural since the necessary entanglement e↵ects this paper as well as Refs. 50 and 51 present a coherent
are embedded in the interaction terms of the Hubbard and complete theoretical picture of our current under-
model. standing of the double quantum dot system including all
Experimentally, the charge stability diagram can be e↵ects of quantum fluctuations and entanglement.
directly measured by a quantum point contact,35 and it In our previous works,50,51 we have introduced the gen-
is therefore a genuine representation of electron states eralized Hubbard model, elaborated the motivation and
in the quantum dot system. It contains two kinds of various simplifications, and applied it to explain several
information: the classical Coulomb e↵ect and quantum specific experiments. We consider the present paper as
fluctuations. The classical aspect of the change stability the concluding one of this three-paper series studying the
diagram is well understood using the capacitance circuit charge stability diagram of double quantum dot system.
model.46–49 On the other hand, microscopic calculations In other words, the goal of this paper is to complete the
are usually carried out for other aspects of the quantum theoretical framework that we have developed, providing
dot system39,72,82–87 but only rarely for the charge sta- detailed information that has not been covered by previ-
bility diagram.56 This means that our understanding of ous papers,50,51 thus forming a comprehensive and coher-
the quantum aspect of the charge stability diagram has ent picture together with the previous works. First, in
largely been limited to the quantum-mechanical two-level Refs. 50,51 we have focused on one of the most important
model,37,52–55 with only a few exceptions.60,61,81 quantum parameters in the generalized Hubbard model,
The generalized Hubbard model (see Sec. III), intro- namely the tunnel coupling, but other quantum param-
duced in Ref. 50, has been naturally written down with eters are important as well. In this paper, we have stud-
only the simplest symmetry considerations in mind. It ied the e↵ect of all quantum parameters on the charge
includes all terms allowed by symmetry, which therefore stability diagram (Sec. IV), and several cases with ex-
means that all possible quantum fluctuation e↵ects are treme parameters are shown (Appendix B). Second, in
included. On one hand, when the quantum fluctuations Ref. 50 we have only used the biquadratic potential as
15

the model confinement potential in the microscopic cal- clear spin bath, making the nuclear spin bath less de-
culations. Although in Ref. 51 results from both the bi- structive and the e↵ect of the impurities more prominent.
quadratic and Gaussian models are shown, a direct com- In some systems such as Si or C, the direct hyperfine cou-
parison between the two models is missing. In this paper, pling can be eliminated by using isotropic purification of
we have directly compared the results of the charge sta- the nuclear elements.95 Phonon modes are the least im-
bility diagram calculated using these two model poten- portant environmental factor for spin qubits working at
tials (Sec. V A). We have found that since the Gaussian low (⇠ 100 mK) temperatures: Firstly, the coupling be-
model has narrower potential wells, the electron states tween the spin qubit and the phonon modes are orders of
are more localized, leading to a stronger on-site Coulomb magnitude weaker than the e↵ect of charged impurities
interaction and weaker inter-site quantum fluctuations. and nuclear spin bath mentioned above. Secondly, the
Third, in all previous works we have retained the lowest- e↵ect of phonon modes can be suppressed by cooling the
lying orbital only. In order to justify this simplification, system down to very low temperatures, while the e↵ects
in this paper we have studied the e↵ect of higher-lying or- due to charged impurities and the nuclear spin bath re-
bitals on the one-orbital subspace that we have previously main important even at low temperatures. In this paper,
considered (Sec. V B). We have found that the additional we focus on the electronic interaction within the quan-
p-orbitals lead to a small renormalization of Coulomb in- tum dot system and study its consequence on the charge
teraction parameters while leaving the main physics un- stability diagram, and quantum decoherence due to en-
changed. This justifies our practice of focusing on the vironmental factors is beyond the scope of the current
one-orbital subspace. Last, we have studied the role of work.
the harmonic oscillator frequency in the Gaussian model
and the e↵ect of external magnetic field (Sec. V C). We Our work suggests several possible directions for future
have found that the electron wave functions with fixed studies. First, it is useful to go beyond the model poten-
value of harmonic oscillator frequency is most suitable tial and Hund-Mulliken approximation, and develop a
for comparison with experiments. The influence of an first-principles Poission-Schrödinger approach for a real-
external magnetic field can be understood as e↵ectively istic calculation of quantum dot confinement. This in-
increasing the harmonic oscillator frequency in the Fock- volves both extracting the detailed form of confinement
Darwin wave function. Both of these are not covered by potentials from experimentally fabricated devices, and
our previous works. an exact treatment of Schrödinger equation under this
realistic confinement potential. However, this would re-
As discussed in Sec. II, we have neglected the inter- quire a precise experimental control of disorder and im-
action of the quantum dot spin qubit with the envi- purities during the fabrication, as well as the knowledge
ronment because environmental decoherence, which is of the underlying Fermi sea. Moreover, the calculation
a huge subject by itself, is well beyond the scope of expenses are much higher than what we have done in
the current paper dealing with the qubit control and this work, although they are not necessarily prohibitive
charge stability diagram in semiconductor quantum dots. since modern numerical techniques, e.g. density func-
Various environment factors contribute to the decoher- tional theory, have already demonstrated their power in
ence process, which include the charge noise,91,92 arising successfully explaining many aspects of complex mate-
from the electromagnetic fluctuations in the surround- rials as well as nanoscale systems. Second, although we
ing gates and the dynamics in unintentional background have argued that the additional higher-lying orbitals pose
impurities,64,65 the hyperfine coupling to the surround- no more than a renormalization of Coulomb parameters
ing nuclear spin bath,40,42,43,62,63 spectral di↵usion of the in the single-orbital subspace, the issue for the full multi-
electron spin due to slow nuclear spin dynamics arising orbital multi-electron problem may not yet be settled. A
from the nuclear spin flip-flops due to dipolar coupling detailed study of the realistic multi-orbital multi-electron
within the nuclear spin bath,62,93,94 the coupling of elec- problem would in particular greatly benefit our under-
tron spins to unintentional background impurity spins in standing of the silicon quantum dot system51 as this
the environment,95 and phonon modes.66–70 In many sit- multi-electron e↵ect may be one of the obstacles of real-
uations, the most important environmental factor is the izing and manipulating the singlet/triplet qubit in silicon
e↵ect of unintentionally introduced background charged double quantum dot system. Last, we have focused on
impurities, since depending on the proximity between equilibrium properties of an isolated quantum dot system
the impurities and the qubit, the interaction with the where the total electron number and the total spin are
impurities could be of the same order of magnitude as conserved. However, non-equilibrium e↵ects are inter-
the on-dot electronic interactions as both are mediated twined with the measurement of the charge stability di-
by the long-range Coulomb coupling.64,65 The environ- agram as well as the manipulation of qubits. The charge
mental nuclear spin bath a↵ects the electron spin both stability diagram is measured through a tunnelling cur-
through a “direct” hyperfine interaction between the spin rent, which itself is a non-equilibrium property. In Ref. 81
qubit and the nuclei, and an “indirect” coupling due to a study of the charge stability diagram in terms of the
the spectral di↵usion process resulting from the dipo- microscopic conductance(admittance) is given, and much
lar interaction between the nuclei. Recent experimental of our theory can also be applied along this direction. Of
progresses18,24 have demonstrated control over the nu- course, such a dynamical theory would be much more
16

complicated than what we have studied here, since the conserved, it appears in a block-diagonal form, as follows.
excited states must be considered in all non-equilibrium (1) N = 0, Sz = 0. On the basis |0i,
calculations. Moreover, the coupling of the quantum dot
system to its environment, which has been neglected in H1 = 0. (A-1)
this paper, is by no means insignificant. Therefore, the n o
assumptions that we have made in this paper as well as (2) N = 1, Sz = 12 . On the basis c†1" |0i , c†2" |0i ,
Refs. 50,51, that the system is isolated and in equilib-
rium, should be lifted in future studies by extending the ✓ ◆
generalized Hubbard model. The theoretical approach µ 1 + EB t
H2 = . (A-2)
that we have presented here provides a foundation for t µ 2 + EB
these future research directions. n o
(3) N = 1, Sz = 1
2. On the basis c†1# |0i , c†2# |0i ,

Acknowledgements ✓ ◆
µ1 EB t
H3 = . (A-3)
t µ2 EB
We thank J. P. Kestner for helpful discussions. This
work is supported by IARPA and LPS.
(4) N = 2, Sz = 1. On the basis c†1" c†2" |0i (labeled as
T+ ),
Appendix A: Matrix form of the generalized
Hubbard model H4 = µ1 µ2 + U12 Je + 2EB . (A-4)

The full Hamiltonian is a 16 ⇥ 16 matrix. Because n (5) N = 2, Sz = 0. On


o the basis
the total particle number (N ) and total spin (Sz ) are c†1" c†2# |0i , c†2" c†1# |0i , c†1" c†1# |0i , c†2" c†2# |0i ,

0 1
µ1
µ2 + U12 Je (t + Jt1 ) (t + Jt2 )
B Je µ1 µ2 + U12 (t + Jt1 ) (t + Jt2 ) C
H5 = @ A. (A-5)
(t + Jt1 ) (t + Jt1 ) 2µ1 + U1 Jp
(t + Jt2 ) (t + Jt2 ) Jp 2µ2 + U2
⇣ ⌘ p ⇣ ⌘ p
Note that the linear combinations c†1" c†2# c†2" c†1# / 2 (labeled as T0 ) and c†1" c†2# + c†2" c†1# / 2 (labeled as S)
decouple the T0 state with energy µ1 µ2 + U12 Je , thus forming a triplet which degenerate at zero magnetic field.
The remaining 3 ⇥ 3 matrix reads
0 p p 1
µ1 pµ2 + U12 + Je 2(t + Jt1 ) 2(t + Jt2 )
He5 = @ A. (A-6)
p2(t + Jt1 ) 2µ1 + U1 Jp
2(t + Jt2 ) Jp 2µ2 + U2

(6) N = 2, Sz = 1. On the basis c†1# c†2# |0i (labeled as T ),

H6 = µ1 µ2 + U12 Je 2EB . (A-7)


n o
(7) N = 3, Sz = 12 . On the basis c†1" c†2" c†1# |0i , c†1" c†2" c†2# |0i ,
✓ ◆
2µ1
µ2 + U1 + 2U12 Je + EB (t + Jt1 + Jt2 )
H7 = . (A-8)
(t + Jt1 + Jt2 ) 2µ2 µ1 + U2 + 2U12 Je + EB
n o
(8) N = 3, Sz = 12 . On the basis c†1" c†1# c†2# |0i , c†2" c†1# c†2# |0i ,
✓ ◆
2µ1 µ2 + U1 + 2U12 Je EB (t + Jt1 + Jt2 )
H8 = . (A-9)
(t + Jt1 + Jt2 ) 2µ2 µ1 + U2 + 2U12 Je EB

(9) N = 4, Sz = 0. On the basis c†1" c†2" c†1# c†2# |0i, Appendix B: Several extreme cases of the
parameters
H9 = 2µ1 2µ2 + U1 + U2 + 4U12 2Je . (A-10)
In this section we present several extreme cases of pa-
rameters. As discussed above, Je < U12 < U in a typical
17

orthogonality of atomic orbitals means that t = 0 while


(2,1) (2,2) Je is still finite. Third, interesting physical phenomena
20 A' happen in some of the parameter regimes, such as the
(2,0) (1,2) U12 > U case discussed in Ref. 96, which is at least of
B'
µ1 (meV) theoretical interest.
10
Fig. 14 shows a typical case for U12 U . The coor-
B
(1,0) (0,2) dinates (µ2 , µ1 ) of the triple points are given by A(0, 0),
0 A B(U, U ), B 0 (2U12 , 2U12 ), A0 (U + 2U12 , U + 2U12 ). The
(0,0) (0,1) (1, 1) component is completely ruled out from the charge
stability diagram since the system always favors either
0 10 20 (0, 2) or (2, 0) state which have smaller Coulomb repul-
µ2 (meV) sion than the (1, 1) state. This charge stability diagram is
not observed experimentally, which means that it is rea-
sonable to set U12 < U for currently implemented double
FIG. 14: Charge stability diagram calculated at U = 6.1 meV, quantum dot systems.
U12 = 7 meV, t = Je = Jp = Jt = 0.
In Fig. 15 we show the case with Je U12 (but still
Je < U ). Fig. 15(a) shows the Je = U12 result, which can
physical system. However it remains interesting to study be understood as the limit of Fig. 4. Fig. 15(b) shows
the cases with U12 U or Je U12 . First, for complete- the result of U12 < Je < U . The topology fundamen-
ness of the understanding of the charge stability diagram, tally changes in that the (1,1) triplet state now shares a
these parameter regimes need be explored. Indeed, as phase boundary directly with the (0,0) state. The coor-
shall be seen below, the topology of the charge stability dinates (µ2 , µ1 ) of the triple points are A(U12 Je , 0),
diagram changes. Second, although the systems we are B(0, U12 Je ), and C(U, U12 Je ), with others found
currently studying impose special limitations of param- by symmetry. p Interestingly the length of line segment
eters, this in general need not be true. For example, in AB
p is still 2(Je U12 ) (which can be expressed as
our study Je ⌧ t. However, in a lattice problem the 2 |U12 Je | formally similar to that of Fig. 4).

1 9
R. P. Feynman, Int. J. Theor. Phys. 21, 467 (1982); Found. F. H. L. Koppens, C. Buizert, K. J. Tielrooij, I. T. Vink,
Phys. 16, 507 (1986); D. Deutsch, Proc. R. Soc. London, K. C. Nowack, T. Meunier, L. P. Kouwenhoven, and L. M.
Ser. A 400, 97 (1985) K. Vandersypen, Nature 442, 766 (2006).
2 10
P. W. Shor, in Proceedings of the 35th Annual Symposium B. E. Kane, Nature 393, 133 (1998).
11
on the Foundations of Computer Science (IEEE Press, Los R. Vrijen, E. Yablonovitch, K. Wang, H. W. Jiang, A.
Alamitos, CA, 1994), pp. 124133; A. Ekert and R. Jozsa, Balandin, V. Roychowdhury, T. Mor, and D. DiVincenzo,
Rev. Mod. Phys. 68, 733 (1996); L. K. Grover, Phys. Rev. Phys. Rev. A 62, 012306 (2000).
12
Lett. 79, 325 (1997); A. Steane, Rep. Prog. Phys. 61, 117 M. Friesen, P. Rugheimer, D. E. Savage, M. G. Lagally, D.
(1998); C. H. Bennett and D. P. DiVincenzo, Nature (Lon- W. van der Weide, R. Joynt, and M. A. Eriksson, Phys.
don) 404, 247 (2000). Rev. B 67, 121301(R) (2003).
3 13
L. M. K. Vandersypen, M. Ste↵en, G. Breyta, C. S. Yan- J. M. Taylor, H.-A. Engel, W. Dür, A. Yacoby, C. M. Mar-
noni, R. Cleve, and I. L. Chuang, Phys. Rev. Lett. 85, 5452 cus, P. Zoller, and M. D. Lukin, Nat. Phys. 1, 177 (2005).
14
(2000); L. M. K. Vandersypen, M. Ste↵en, G. Breyta, C. J. M. Kikkawa and D. D. Awschalom, Phys. Rev. Lett. 80,
S. Yannoni1, M. H. Sherwood, and I. L. Chuang, Nature 4313 (1998).
15
(London) 414, 883 (2001) S. Amasha, K. MacLean, Iuliana P. Radu, D. M. Zumbühl,
4
T. Sleator and H. Weinfurter, Phys. Rev. Lett. 74, 4087 M. A. Kastner, M. P. Hanson, and A. C. Gossard, Phys.
(1995); Q. A. Turchette, C. J. Hood, W. Lange, H. Rev. Lett. 100, 046803 (2008).
16
Mabuchi, and H.J. Kimble, ibid. 75, 4710 (1995). F. H. L. Koppens, K. C. Nowack, and L. M. K. Vander-
5
J. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995); sypen, Phys. Rev. Lett. 100, 236802 (2008).
17
C. Monroe, D. M. Meekhof, B. E. King, W. M. Itano, and C. Barthel, J. Medford, C. M. Marcus, M. P. Hanson, and
D. J. Wineland, ibid. 75, 4714 (1995); C. A. Sackett, D. A. C. Gossard, Phys. Rev. Lett. 105, 266808 (2010).
18
Kielpinski, B. E. King, C. Langer, V. Meyer, C. J. Myatt, H. Bluhm, S. Foletti, I. Neder, M. Rudner, D. Mahalu, V.
M. Rowe, Q. A. Turchette, W. M. Itano, D. J. Wineland, Umansky, and A. Yacoby, Nat. Phys. 7, 109 (2011).
19
and C. Monroe, Nature (London) 404, 256 (2000). A. M. Tyryshkin, S. A. Lyon, A. V. Astashkin, and A. M.
6
D. Loss and D. P. DiVincenzo, Phys. Rev. A 57, 120 Raitsimring, Phys. Rev. B 68, 193207 (2003).
20
(1998). A. Morello, J. J. Pla, F. A. Zwanenburg, K. W. Chan, K.
7
D. P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard, and Y. Tan, H. Huebl, M. Möttönen, C. D. Nugroho, C. Yang,
K. B. Whaley, Nature 408, 339 (2000). J. A. van Donkelaar, A. D. C. Alves, D. N. Jamieson, C.
8
J. Levy, Phys. Rev. Lett. 89, 147902 (2002). C. Escott, L. C. L. Hollenberg, R. G. Clark, and A. S.
18

Phys. Rev. B 80, 205302 (2009); 82, 155312 (2010).


30
(2,0) (2,1) (2,2) S. Goswami, K. A. Slinker, M. Friesen, L. M. McGuire, J.
10 L. Truitt, C. Tahan, L. J. Klein, J. O. Chu, P. M. Mooney,
D. W. van der Weide, R. Joynt, S. N. Coppersmith, and

µ1 (meV)
M. A. Eriksson, Nat. Phys. 3, 41 (2007).
5 (1,2) 31
M. G. Borselli, R. S. Ross, A. A. Kiselev, E. T. Croke, K.
(1,1)T S. Holabird, P. W. Deelman, L. D. Warren, I. Alvarado-
(1,0)
Rodriguez, I. Milosavljevic, F. C. Ku, W. S. Wong, A. E.
0 Schmitz, M. Sokolich, M. F. Gyure, and A. T. Hunter,
Appl. Phys. Lett. 98, 123118 (2011).
(0,0) (0,1) (0,2) 32
C. B. Simmons, M. Thalakulam, B. M. Rosemeyer, B. J.
Van Bael, E. K. Sackmann, D. E. Savage, M. G. Lagally, R.
-5
-5 0 5 10 Joynt, M. Friesen, S. N. Coppersmith, and M. A. Eriksson,
Nano Lett. 9 3234 (2009).
µ2 (meV) 33
N. S. Lai, W. H. Lim, C. H. Yang, F. A. Zwanenburg,
(a) W. A. Coish, F. Qassemi, A. Morello, and A. S. Dzurak,
e-print arXiv:1012.1410 (unpublished).
34
10 (2,0) (2,1) (2,2) C. Livermore, C. H. Crouch, R. M. Westervelt, K. L.
Campman, A. C. Gossard, Science 274, 1332 (1996).
35
J. M. Elzerman, R. Hanson, J. S. Greidanus, L. H. Willems
µ1 (meV)

van Beveren, S. De Franceschi, L. M. K. Vandersypen,


5
(1,0) (1,1)T (1,2) S. Tarucha, and L. P. Kouwenhoven, Phys. Rev. B 67,
161308(R) (2003).
36
A F. Hofmann, T. Heinzel, D.A. Wharam, J.P. Kotthaus, G.
0
B C Böhm, W. Klein, G. Tränkle, and G. Weimann, Phys. Rev.
B 51, 13872 (1995).
(0,0) (0,1) (0,2) 37
L. DiCarlo, H. J. Lynch, A. C. Johnson, L. I. Childress, K.
-5 Crockett, C. M. Marcus, M. P. Hanson, and A. C. Gossard,
-5 0 5 10 Phys. Rev. Lett. 92, 226801 (2004).
38
V. W. Scarola, K. Park, and S. Das Sarma, Phys. Rev.
µ2 (meV) Lett. 93, 120503 (2004); V. W. Scarola and S. Das Sarma,
(b) Phys. Rev. A 71, 032340 (2005).
39
Q. Li, L. Cywiński, D. Culcer, X. Hu, and S. Das Sarma,
FIG. 15: Charge stability diagram calculated at U = 6.1 meV, Phys. Rev. B 81, 085313 (2010).
U12 = 2.5 meV, (a) Je = 2.5 meV = U12 (b) Je = 4 meV > 40
J. M. Taylor, J. R. Petta, A. C. Johnson, A. Yacoby, C. M.
U12 . t = Jp = Jt = 0. Marcus, and M. D. Lukin, Phys. Rev. B 76, 035315 (2007)
41
J. R. Petta, J. M. Taylor, A. C. Johnson, A. Yacoby, M. D.
Lukin, C. M. Marcus, M. P. Hanson, and A. C. Gossard,
Dzurak, Nature (London) 467, 687 (2010). Phys. Rev. Lett. 100, 067601 (2008).
21 42
C. B. Simmons, J. R. Prance, B. J. Van Bael, T. S. Koh, D. J. Reilly, J. M. Taylor, J. R. Petta, C. M. Marcus, M.
Z. Shi, D. E. Savage, M. G. Lagally, R. Joynt, M. Friesen, P. Hanson, and A. C. Gossard, Science 321, 817 (2008).
43
S. N. Coppersmith, and M. A. Eriksson, Phys. Rev. Lett. D. J. Reilly, J. M. Taylor, J. R. Petta, C. M. Marcus, M. P.
106, 156804 (2011). Hanson, and A. C. Gossard, Phys. Rev. Lett. 104, 236802
22
M. Xiao, M. G. House, and H. W. Jiang, Phys. Rev. Lett. (2010).
44
104, 096801 (2010). C. Barthel, D. J. Reilly, C. M. Marcus, M. P. Hanson, and
23
J. R. Petta, A. C. Johnson, J. M. Taylor, E. A. Laird, A. A. C. Gossard, Phys. Rev. Lett. 103, 160503 (2009).
45
Yacoby, M. D. Lukin, C. M. Marcus, M. P. Hanson, A. C. C. Barthel, M. Kjærgaard, J. Medford, M. Stopa, C. M.
Gossard, Science 309, 2180 (2005). Marcus, M. P. Hanson, and A. C. Gossard, Phys. Rev. B
24
S. Foletti, H. Bluhm, D. Mahalu, V. Umansky, and A. 81 161308(R), (2010).
46
Yacoby, Nat. Phys. 5, 903 (2009). C. W. J. Beenakker, Phys. Rev. B 44, 1646 (1991)
25 47
E. A. Laird, J. M. Taylor, D. P. DiVincenzo, C. M. Marcus, W. G. van der Wiel, S. De Franceschi, J. M. Elzerman,
M. P. Hanson, and A. C. Gossard, Phys. Rev. B 82, 075403 T. Fujisawa, S. Tarucha, L. P. Kouwenhoven, Rev. Mod.
(2010). Phys. 75, 1 (2003).
26 48
I. van Weperen, B. D. Armstrong, E. A. Laird, J. Medford, R. Hanson, L. P. Kouwenhoven, J. R. Petta, S. Tarucha,
C. M. Marcus, M. P. Hanson, and A. C. Gossard, Phys. L. M. K. Vandersypen, Rev. Mod. Phys. 79, 1217 (2007).
49
Rev. Lett. 107, 030506 (2011). D. Schröer, A. D. Greentree, L. Gaudreau, K. Eberl, L. C.
27
E. P. Nordberg, G. A. Ten Eyck, H. L. Stalford, R. P. L. Hollenberg, J. P. Kotthaus, and S. Ludwig, Phys. Rev.
Muller, R. W. Young, K. Eng, L. A. Tracy, K. D. Childs, B 76, 075306 (2007).
50
J. R. Wendt, R. K. Grubbs, J. Stevens, M. P. Lilly, M. S. Yang, X. Wang, and S. Das Sarma, Phys. Rev. B 83,
A. Eriksson, and M. S. Carroll, Phys. Rev. B 80, 115331 161301(R) (2011).
51
(2009). S. Das Sarma, X. Wang, and S. Yang, Phys. Rev. B 83,
28
S. Das Sarma, R. de Sousa, X. Hu, and B. Koiller, Sol. St. 235314 (2011).
52
Comm. 133, 737 (2005). J. R. Petta, A. C. Johnson, C. M. Marcus, M. P. Hanson,
29
D. Culcer, L. Cywiński, Q. Li, X. Hu, and S. Das Sarma, and A. C. Gossard, Phys. Rev. Lett. 93, 186802 (2004)
19

53
T. Hatano, M. Stopa, and S. Tarucha, Science 309, 268 Theor. Phys. 30, 275 (1963).
76
(2005). J. E. Hirsch, Physica C 199, 305 (1992); J. E. Hirsch and
54
A. K. Hüttel, S. Ludwig, H. Lorenz, K. Eberl, and J. P. F. Marsiglio, Phys. Rev. B 62, 15131 (2000).
77
Kotthaus, Phys. Rev. B 72, 081310 (2005). J. Hubbard, Proc. R. Soc. London, Ser. A 276, 238 (1963);
55
M. Pioro-Ladrière, M. R. Abolfath, P. Zawadzki, J. La- 277, 237 (1964); 281, 401 (1964).
78
pointe, S. A. Studenikin, A. S. Sachrajda, and P. Hawry- J. Zaanen and M. Oleś, Ann. Phys. (Leipzig) 5, 224 (1996)
79
lak, Phys. Rev. B 72, 125307 (2005). M. Imada, A. Fujimori, Y. Tokura, Rev. Mod. Phys. 70,
56
L.-X. Zhang, D. V. Melnikov, and J.-P. Leburton, Phys. 1039 (1998).
80
Rev. B 74, 205306 (2006). A. Liebsch, Phys. Rev. Lett. 91, 226401 (2003); A. Koga,
57
C. A. Sta↵ord and S. Das Sarma, Phys. Rev. Lett. 72, N. Kawakami, T. M. Rice, and M. Sigrist, Phys. Rev. Lett.
3590 (1994). 92, 216402 (2004); P. Werner and A. J. Millis, Phys. Rev.
58
R. Kotlyar, C. A. Sta↵ord, and S. Das Sarma, Phys. Rev. B 74, 155107 (2006)
81
B 58, R1746 (1998); 58, 3989 (1998); C. A. Sta↵ord, R. A. Cottet, C. Mora, and T. Kontos, Phys. Rev. B 83,
Kotlyar, and S. Das Sarma, ibid. 58, 7091 (1998). 121311(R) (2011).
59 82
J. H. Je↵erson and W. Häusler, Phys. Rev. B 54, 4936 G. Burkard, D. Loss, and D. P. DiVincenzo, Phys. Rev. B
(1996). 59, 2070 (1999).
60 83
L. Gaudreau, S. A. Studenikin, A. S. Sachrajda, P. Za- X. Hu and S. Das Sarma, Phys. Rev. A 64, 042312 (2001).
84
wadzki, A. Kam, J. Lapointe, M. Korkusinski, and P. R. de Sousa, X. Hu, and S. Das Sarma, Phys. Rev. A 64,
Hawrylak, Phys. Rev. Lett. 97, 036807 (2006). 042307 (2001).
61 85
M. Korkusinski, I. P. Gimenez, P. Hawrylak, L. Gaudreau, I. Puerto Gimenez, M. Korkusinski, and P. Hawrylak,
S. A. Studenikin, and A. S. Sachrajda, Phys. Rev. B 75, Phys. Rev. B 76, 075336 (2007).
86
115301 (2007). E. Nielsen, R. W. Young, R. P. Muller, and M. S. Carroll,
62
W. M. Witzel and S. Das Sarma, Phys. Rev. B 74, 035322 Phys. Rev. B 82, 075319 (2010)
87
(2006) E. Nielsen and R. P. Muller, e-print arXiv:1006.2735 (un-
63
L. Cywiński, W. M. Witzel, and S. Das Sarma, Phys. Rev. published)
88
Lett. 102, 057601 (2009); Phys. Rev. B 79, 245314 (2009) M. Helle, A. Harju, and R. M. Nieminen, Phys. Rev. B 72,
64
I. P. Gimenez, C.-Y. Hsieh, M. Korkusinski, and P. Hawry- 205329 (2005)
89
lak, Phys. Rev. B 79, 205311 (2009) J. Pedersen, C. Flindt, N. A. Mortensen, and A.-P. Jauho,
65
N. T. T. Nguyen and S. Das Sarma, Phys. Rev. B 83, Phys. Rev. B 76, 125323 (2007)
90
235322 (2011) J. G. Pedersen, C. Flindt, A.-P. Jauho, and N. A.
66
M. J. Storcz, U. Hartmann, S. Kohler, and F. K. Wilhelm, Mortensen, Phys. Rev. B 81, 193406 (2010).
91
Phys. Rev. B 72, 235321 (2005) X. Hu and S. Das Sarma, Phys. Rev. Lett. 96, 100501
67
V. N. Stavrou and X. Hu, Phys. Rev. B 72, 075362 (2005) (2006)
68 92
P. Stano and J. Fabian, Phys. Rev. Lett. 96, 186602 (2006) M. Stopa, and C. M. Marcus, Nano Lett., 8, 1778 (2008).
69 93
D. Harbusch, D. Taubert, H. P. Tranitz, W. Wegscheider, W. M. Witzel, R. de Sousa, and S. Das Sarma, Phys. Rev.
and S. Ludwig, Phys. Rev. Lett. 104, 196801 (2010) B 72, 161306(R) (2005)
70 94
X. Hu, Phys. Rev. B 83, 165322 (2011) W. M. Witzel and S. Das Sarma, Phys. Rev. B 77, 165319
71
E. Artacho and L. Miláns del Bosch, Phys. Rev. A 43, 5770 (2008)
95
(1991). W. M. Witzel, M. S. Carroll, A. Morello, L. Cywiński, and
72
X. Hu and S. Das Sarma, Phys. Rev. A 61, 062301 (2000). S. Das Sarma, Phys. Rev. Lett. 105, 187602 (2010)
73 96
J. C. Slater, Phys. Rev. 49, 537 (1936). X. Wang and A. J. Millis, Phys. Rev. B 81, 045106 (2010).
74
P. W. Anderson, Phys. Rev. 115, 2 (1959).
75
J. Kanamori, J. Phys. Chem. Solids 10, 87 (1959); Prog.

You might also like