Exact analytical treatment of multiqubit noisy dynamics in exchange-coupled spin

Donovan Buterakos and Sankar Das Sarma Condensed Matter Theory Center and Joint Quantum Institute, Department of Physics, University of Maryland, College Park, Maryland 20742-4111 USA (Dated: May 5, 2021) Charge noise remains the primary obstacle to the development of technolo- gies with semiconductor spin qubits. We use an exact analytical calculation to determine the effects of quasistatic charge noise on a ring of three equally spaced exchange-coupled quantum dots. We calculate the disorder-averaged return probability from a specific initial state, and use it to deter- ∗ mine the coherence time T2 and show that it depends only on the disorder strength and not the mean interaction strength. We also use a perturbative approach to investigate other arrangements of three or four qubits, finding that the return probability contains multiple oscillation frequencies. These oscillations decay in a Gaussian manner, determined by differences in energy levels of the Hamiltonian. We give quantitative values for gate times resulting in several target fidelities. We find that the decoherence time decreases with increasing number of qubits. Our work provides useful analytical insight into the charge noise dynamics of coupled spin qubits.

I. INTRODUCTION However, the fidelities of these two- gates must still be increased substantially (> 99%) in order to meet the Electron spin qubits in semiconductor (particularly, minimum threshold values needed to implement quan- Si) quantum dots is considered to be a promising plat- tum error correcting codes. The two-qubit gate fidelities form for developing quantum information technologies are predominantly limited by charge noise, which can due to its long coherence times and fast gate speed, arise from local charge impurities in the solid or in the as well as the potential scalability and ability to use controlling circuits which create the quantum dot poten- resources from the semiconductor industry which are tial wells [14, 15]. These impurities can affect the elec- currently available. Semiconductor spin qubits remain tron wave functions, and in turn affect the strength of among the most-studied quantum computing platforms the exchange interaction used to perform the two-qubit in the world with laboratories in USA, Australia, Eu- gates, since the exchange interaction depends very pre- rope, Japan, Canada, and China actively pursuing cou- cisely on the wavefunction overlap [14]. Methods have pled semiconductor quantum dot systems for eventual been proposed to reduce the effect of charge noise, such quantum computing applications. There has been much as using quadrupolar exchange-only qubits, which uses experimental progress including the implementation of four electrons in three quantum dots to access a “sweet two qubit gates between singlet-triplet qubits [1], the spot” where qubit operation is robust to first order charge fabrication of linear arrays of nine quantum dots [2], noise [16]. Additionally, there are proposals for dynam- the shuttling of spins across linear arrays [3], and the ical decoupling schemes, which perform rotations using implementation of basic quantum algorithms on a pro- complex pulse shapes which cancel errors to some degree grammable two-qubit processor [4]. A plaquette of four [17, 18]; however, dynamical decoupling requires a pre- quantum dots has also been used to perform simulations cise understanding of the source and form of the noise to of the Hubbard model and observe Nagaoka ferromag- be canceled. None of the proposed techniques for mitiga- netism [5–7]. A four-qubit quantum processor has been tion has been generically successful in eliminating charge constructed with hole qubits in Geranium quantum dots noise, and understanding and eliminating charge noise [8, 9]. Very recent unpublished work from Delft reports remains the main obstacle in the development of semi- the fabrication of a 6-spin qubit Si system, where our conductor spin qubits. work should be relevant [10]. These are just a few exam- The effects of noise have been studied theoretically in ples from the many recent exciting developments in the coupled double quantum dot systems [19–23]. Beginning field. In spite of much impressive progress, the subject with a Hamiltonian for two exchange coupled dots in arXiv:2012.05924v2 [cond-mat.mes-hall] 4 May 2021 faces a rather difficult challenge in mitigating errors aris- the presence of an external magnetic field, the proba- ing from charge noise, which is invariably present in all bility P (t) was calculated that the system initialized in electronic materials, devices, and circuits. In particular, a given state would be measured in the same state after charge noise has prevented semiconductor quantum dot evolution for some time t. Quasistatic noise was mod- platforms from developing multiqubit system operations eled by choosing the magnetic field and exchange interac- with the current limit being 2-4 coupled qubits at most. tion strengths from random Gaussian distributions, and Benchmarking of quantum dot devices has obtained the return probability was analytically averaged over all single qubit gates with fidelities over 99.9% [11], but choices for each parameter. This disorder-averaged re- two-qubit gates have fidelities ranging from 90% to 95%, turn probability showed oscillations within a Gaussian- ∗ with a few specific gates reaching up to 98% [12, 13]. shaped envelope, and the value of T2 for the system was 2 obtained from the decay rate of the envelope [19]. Nu- equilateral triangle. Specifically, we consider the case merical simulations have also been performed for larger where the mean interaction strength between each pair of systems in the same manner, including a system of two dots is identical. This case has a high degree of symmetry, capacitively coupled singlet triplet qubits [20, 21] and ion which allows the dynamics in the presence of noise to be trap spin chains (which have a similar Hamiltonian and calculated exactly analytically. We begin by defining the dynamics)[22]. This method has also been used to deter- model and Hamiltonian; we then show the calculation of mine the fidelity of spin transport across spin chains via the disorder averaged return probability from an initial SWAP gates [23]. These works are all, however, purely state; finally, we discuss the results. numerical, and although useful in their own right, these numerical calculations fail to provide analytical insight into the noise dynamics. A. Model and Hamiltonian In this current work, we analytically calculate the ef- fects of charge noise on various systems of three and four We consider a single half-filled band in a plaquette con- quantum dots. For a ring of three quantum dots, we sisting of a ring of three quantum dot, and thus three perform an exact analytical calculation and obtain ex- electrons are present. We neglect the effect of higher pressions for the disorder averaged return probability, as unoccupied energy orbitals as the energy difference be- well as the spin expectation values and entanglement en- tween bands tends to be much larger than the interaction tropy for each qubit. We find that the shape of the os- strength between dots, which is the relevant energy scale cillation envelope is completely independent of the mean for the dynamics of the system. Similarly, we also as- exchange interaction strength and depends only on the sume a large onsite interaction energy, and thus we ignore disorder strength, and we obtain the exact coherence states where a single dot contains more than one electron ∗ time T2 as a function of disorder strength. For other (these are all experimentally valid and theoretically used systems, we develop a pertubative approach to calculat- approximations for semiconductor quantum dot qubits). ing the noise-induced decoherence effects. We show that Then each dot will contain exactly one electron, and the multiple frequencies appear in the resulting expressions, exchange interaction between dots will give rise to the and that these frequencies depend on the differences be- following Heisenberg Hamiltonian: tween energy levels of the noiseless Hamiltonian. These frequencies each decay in a Gaussian-like manner with ~ ~ ~ ~ ~ ~ their own decay rate, and the decay rates are given by H = J12S1 · S2 + J23S2 · S3 + J13S1 · S3 (1) the first order noise-induced corrections to the energy levels which produce the frequencies. For each system, Let the values of the exchange interaction between two ∗ dots Jij have a Gaussian distribution because of the noise we calculate T2 and find the times tc at which the system drops below various fidelity benchmarks. Our analytical with mean J0 and standard deviation σJ . Here, σJ is a theory provides detailed insight into the noise dynamics measure of the charge noise induced disorder, leading to which are hidden (and therefore difficult to discern) in random Gaussian fluctuations in the exchange coupling. the existing numerical calculations. It is in fact quite In order for this distribution to be sensible, J0 must be an unanticipated finding that the multiqubit dynamics several times larger than σJ , so that the proportion of under charge noise can be obtained analytically. the distribution with a negative value of J is negligible, This paper is organized as follows: in Sec. II, we ex- and we will assume this is the case. It is convenient to amine a ring of three equally spaced quantum dots, and define and work with the deviations ∆ij = Jij −J0. These perform an exact analytical calculation of several quanti- deviations will then have the following distribution ties in the presence of noise, including the disorder aver- 2 aged return probability from a specific initial state. We − ∆ 1 2σ2 give an analysis of the results, discussing the short and f(∆) = √ e J (2) 2πσ long time behavior of the system and its dependence on J the initial Hamiltonian. In Sec. III, we address a system For any function or operator A which depends on Jij, of three quantum dots in a linear geometry by using a we define its disorder averaged expectation [A] as follows: perturbative approach, and we compare our results to a direct numerical evaluation of the same quantities. In Sec Z IV, we use the same perturbative approach to examine a [A] = A f(∆12)f(∆23)f(∆13) d∆12d∆23d∆13 (3) ring of four quantum dots. Finally, we conclude with a discussion in Sec. V. Note that the Hamiltonian in Eq. (1) commutes with the total spin operator S2, as well as its z-projection, and thus both quantities are conserved. There is only II. THREE QUBIT RING one state with Sz = 3/2, and thus its dynamics is trivial. We will focus on the dynamics of the Sz = 1/2 subspace. In this section we calculate the effect of noise on a Our model relies on several assumptions. We assume a quantum dot plaquette with three qubits arranged in an linear noise term, and thus our results may not apply to 3 select systems with quadratic noise terms, such as in Ref. 16, although the majority of experiments do exhibit lin- −1 0 0  −Σ 0 2ξ∗  ear noise. It should be possible to generalize our work to 3J 1 ∆ ∆ H = 0 0 1 0 + 0 Σ 0 (5) different noise terms, but this is beyond the scope of the 4   4  ∆  current work where we establish the general theory and 0 0 −1 2ξ∆ 0 −Σ∆ principles using the usual linear noise model. Addition- where Σ = ∆ + ∆ + ∆ , and ξ = e2πi/3∆ + ally, we use the quasistatic noise approximation, which ∆ 12 13 23 ∆ 12 ∆ + e−2πi/3∆ . Diagonalizing gives the following en- is well-justified in quantum dot systems since gate times 13 23 ergies and eigenvectors: are very fast compared to noise frequencies. In fact, the whole point of semiconductor spin qubits is very high gate speed and the prospect for future scalability. We 3J0 Σ∆ E = + also exclude the Zeeman interaction term in our model, as 0 4 4 our focus for this paper is charge noise. In fact, it is well- 3J Σ |ξ | E = − 0 − ∆ ± ∆ known that field noise arising from nuclear spin coupling ± 4 4 2 is strongly suppressed in Si and can even be eliminated |ψ0i = |φ0i by isotopic purification, making Si spin the ideal semicon- ∗ ±χ∆ 1 ductor material for qubit applications [24], leaving charge |ψ±i = √ |φ−1i + √ |φ+1i (6) noise as the only operational decohering mechanism. The 2 2 effect of the Zeeman interaction on qubit dynamics has where χ = ξ /|ξ |. The initial state |↑↓↑i corre- been studied extensively in other works [19–21], and are ∆ ∆ ∆ √ sponds to the vector (1 1 1)/ 3 in the φ basis, and small when the exchange interaction terms J are large, j ij decomposing into the basis of eigenstates, this becomes: which is when there is the greatest amount of charge noise. In Si spin qubits, field noise is far too weak for it to affect qubit operations at the operational gate frequen- 1 − χ∆ 1 1 + χ∆ |↑↓↑i = √ |ψ−i + √ |ψ0i + √ |ψ+i (7) cies. Additionally, some qubit proposals do not have a 6 3 6 Zeeman term present at all, with both one- and two-qubit operations performed via exchange coupling [25]. Vari- We then calculate the return probability P (t) for a ous other system-dependent deviations from our model particular disorder realization, yielding: may occur, such as changes to the exchange interaction due to valley splitting in Si. However, as long as such 1 |ξ | 3J + Σ 2 P (t) = 2 cos ∆ t + cos 0 ∆ t effects are small, the resulting dynamics will still be qual- 9 2 2 itatively the same. Quantitative numerics require a pre- 2 1 |ξ∆| 3J0 + Σ∆  cise knowledge of the microscopic Hamiltonian, which is + 2 Re χ sin t + sin t (8) 9 ∆ 2 2 system-dependent and generally unknown, and thus we study the qualitative behavior of the system. Our cur- From this expression, the disorder average [P (t)] can rent theory should be thought of as the minimal theory be calculated using eq. (3). In order to compute the which must be the starting point for understanding noisy integral, it is helpful to change the variables of integration qubit dynamics in semiconductor spin systems. from ∆ij to Σ∆ and ξ∆, the latter of which takes values over the whole complex plane. We then write ξ∆ in terms of its magnitude and complex argument ϕ∆ = arg ξ∆, B. Calculation of Return Probability which produces the following integral:

2 2 2 Z −(Σ +2|ξ∆| )/6σ We consider the system prepared in an initial state e ∆ J 2|ξ∆| |↑↓↑i, and define the return probability P (t) to be the [P (t)] = P (t) √ √ dΣ∆d|ξ∆|dϕ∆ ( 2πσ )3 3 3 probability of measuring the system to be in the same J (9) state after it has evolved for some time t. Note that this initial state is the only nontrivial state with individual This integral can be evaluated and expressed in terms spins initialized to ↑ or ↓, since other combinations are of the Dawson function F (x), as follows: equivalent by symmetry. There are three states with to- tal Sz = 1/2: {|↑↑↓i, |↑↓↑i, and |↓↑↑i}. These are not √ spin eigenstates, so we instead use the following basis: 5 4 −3σ2 t2/8 3J0t 1  3σJ t [P (t)] = + e J cos − √ σJ tF 9 9 2 3 3 2 √ 2jπi/3 −2jπi/3 2 −3σ2 t2/8 3J0t  3σJ t e |↑↑↓i + |↑↓↑i + e |↓↑↑i − √ e J σJ t cos F (10) |φji = √ (4) 3 3 2 4 3 Using a similar process, we calculate the disorder av- for j = −1, 0, 1, where |φ0i has spin 3/2, and |φ±1i eraged expectation value of the Z2 operator (the Pauli Z have spin 1/2. In this basis, H is given by: operator on the qubit 2), yielding: 4

[P] [D] 1.0 1.0 σJ /J0=0 σJ /J0=0 J J 0.8 σJ / 0=0.1 0.8 σJ / 0=0.1 J J 0.6 σJ / 0=0.2 0.6 σJ / 0=0.2 σJ /J0=0.3 σJ /J0=0.3 J J 0.4 σJ / 0=0.5 0.4 σJ / 0=0.5 σJ /J0=1 σJ /J0=1

0.2 σJ /J0=2 0.2 σJ /J0=2 J t J t 0 2 4 6 8 10 12 0 0 2 4 6 8 10 12 0 Z [ 1] [Z ] 1.0 2 σJ /J0=0 1.0 σJ /J0=0 σJ /J0=0.1 0.8 σJ /J0=0.1 σJ /J0=0.2 0.5 J 0.6 σJ / 0=0.2 σJ /J0=0.3 J 0.0 σJ / 0=0.3 σJ /J0=0.5 J 0.4 σJ / 0=0.5 σJ /J0=1 J -0.5 σJ / 0=1 0.2 σJ /J0=2 σJ /J0=2 J t J t 0 2 4 6 8 10 12 0 0 2 4 6 8 10 12 0

S1 S2 1.0 1.0 σJ /J0=0 σJ /J0=0 J J 0.8 σJ / 0=0.1 0.8 σJ / 0=0.1 J J 0.6 σJ / 0=0.2 0.6 σJ / 0=0.2 σJ /J0=0.3 σJ /J0=0.3 J J 0.4 σJ / 0=0.5 0.4 σJ / 0=0.5 σJ /J0=1 σJ /J0=1

0.2 σJ /J0=2 0.2 σJ /J0=2 J t J t 0 2 4 6 8 10 12 0 0 2 4 6 8 10 12 0

FIG. 1: The disorder averaged return probability [P (t)], normalized Hamming distance [D(t)], expectation values [Z1(t)] and [Z2(t)], and entanglement entropy of qubits 1 & 2 [S1(t)] and [S1(t)]. Each is plotted for several values of σJ , with J0 held constant.

Additionally, in Fig. 1 we plot the entanglement en-  √ tropy between qubit j and the rest of the system, defined 1 √ 3σJ t [Z ] = − 1 + 2 3σ tF  to as: 2 9 J 2 √ √  −3σ2 t2/8 3J0t 3σJ t  + 4e J cos − 2 + 3σ tF  2 J 4 Sj = − Tr ρj log ρj (13) (11) where ρ(t) is the disorder-averaged density matrix of

The expectation values of the other two spins [Z1] = the system evolved a time t from the initial state |↑↓↑i, [Z3] can be obtained directly from this result, since the and ρj is ρ traced over all qubits except qubit j. three must sum to 1. We plot these expectation val- ues as a function of time with J0 constant in Fig. 1. These expectation values can be used to calculate the C. Analysis expectation of the Hamming distance from the initial to final states. The Hamming distance is a quantity We examine the behavior of the return probability used for error-correction which measures the number of [P (t)] given by Eq. (10) by separating [P (t)] into its qubits which must be flipped to transition between the static and oscillatory parts. We define static part as the two states, and thus is a measure of the error introduced midpoint of the oscillation envelope as a function of time, into the system [22, 26]. In in Fig. 1 we plot normalized and the oscillatory part of [P (t)] is the deviation of [P (t)] Hamming distance defined by: from the envelope’s midpoint, which is thus determined −1 by the envelope’s width. Expanding in (σJ t) , we find 1 1  the long-time asymptotic behavior of the static and os- D(t) = + hZ i − hZ i + hZ i (12) 2 6 1 2 3 cillatory parts as follows, which we plot in Fig. 2: 5

[P] [P] 1.0 envelope 1.0 J 0.8 σJ / 0=0.1 0.8 J σJ / 0=0.2 envelope top 0.6 J σJ / 0=0.3 0.6 envolope bottom J 0.4 σJ / 0=0.5 0.4 static part σJ /J0=1

0.2 σJ /J0=2 0.2 σJ t σJ t 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0 [P] [P] J 1.0000 σJ / 0=0.05 0.5 J 0.9998 σJ / 0=0.1 0.4 σJ /J =0.2 envelope 0 0.9996 σJ /J =0.3 0.3 exponential fit 0 J 0.2 0.9994 σJ / 0=0.5 J J 1 0.1 0.9992 σ / 0= σJ /J0=2 σJ t J t 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.2 0.4 0.6 0.8 1.0 0

FIG. 2: Top Left: The return probability plotted against time with σJ held constant. Top Right: The oscillation envelope −1 plotted alongside the static part [P (t)]st. Note the node at t = 2.13σJ . Bottom Left: The oscillatory part of the envelope and its closest exponential fit. Bottom Right: The return probability plotted for small t, with J0 held constant.

to our system of three qubits, where the oscillation en- velope for the return probability is also Gaussian-like in   4 2 1 nature, with its exact functional form given in Eq. (10). [P (t)]st = − 2 2 + O 4 4 9 27σJ t σJ t The exponential decay approximation simply is invalid   2 2 3J0t 32  1  here. In Fig. 2, we show the half width of the oscilla- −3σJ t /8 [P (t)]osc = e cos − 2 2 + O 4 4 tion envelope alongside its closest exponential fit using 2 27σJ t σJ t (14) a least squares fit. Note that the exponential curve is a very poor approximation of the oscillation envelope. ∗ −1 −2 Thus, we define T2 to be the point where the envelope For t  σJ , the static part varies as t , but the os- 2 width reaches 1/e of its original value, and we stress that cillatory part falls off as e−t , so oscillations are only this only gives the relative time scale on which the sys- detectable on short time scales. Specifically, we note tem decoheres and does not imply exponential behavior. that the Gaussian prefactor for the oscillatory part of ∗ This T2 is simply an operational coherence time. Having the envelope implies that it decays much faster than ∗ a well-defined effective ‘coherence time’ T2 is not depen- the standard exponential approximation often associated dent on the actual details of the decoherence process (i.e. with noise. It is interesting to note that the coefficient of exponential versus Gaussian) as long as it is clearly de- the oscillatory part of the envelope vanishes at the point fined as we do here. Experiments are operationally easier σ t = √4 F −1( √2 ) ≈ 2.13, and the coefficient changes J 3 3 to characterize and describe by a single phenomenological ∗ sign as t crosses this point. This means that there will parameter such as T2 , and this is typically done indepen- always be a node at this point due to the shape of the dent of the details of the decoherence mechanism. Our ∗ envelope, which can be seen in Fig. 2. We also give the definition of T2 does not make any assumption about the −1 short-time behavior by expanding [P (t)] with t  σJ , decay of coherence (exponential, Gaussian, or any other yielding the following: functional form) in our system, which we calculate ex- actly. Note that the oscillation envelope is asymmetric due to the tF (t) term in Eq. (10), and thus defining 5 σ2 t2  4 σ2 t2  3J t ∗ [P (t)] = − J + − J cos 0 + O(σ4 t4) T2 based on the envelope width versus the total enve- 9 6 9 3 2 J lope height will give slightly different results, and either ∗ (15) is perfectly acceptable as long as it is made clear how T2 ∗ ∗ Because coherence time T2 is formally defined in terms is being defined. We choose to define T2 based on the ∗ of the decay constant of an exponential curve, it is often envelope width, and by this definition, T2 depends only implied that noise induces an exponential decay of coher- on the oscillatory part of [P (t)]. ence. In Ref. [19] it was shown that quasistatic charge noise in a two-level system produces a Gaussian decay ∗ rather than an exponential decay. We extend that result Using this definition, T2 is given by: 6

2 J0 ∆12 + ∆23 ∆  T ∗ = 1.127σ−1 (16) E1 = + + O 2 J 2 4 J0 ∆2  E2 = O ∗ J It is interesting that T2 is defined only by σJ with J0 0 2 playing no role. Note from Eq. (10) that the only ef- ∆12 + ∆23 ∆  E3 = −J0 − + O fect that J0 has on the return probability [P (t)] is to set 2 J0 the frequency of oscillations. The shape of the oscillation 1 ∆3  envelope itself is completely independent of J . This is |a |2 = + O 0 1 3 J 3 evident in Fig. 2, where curves with different ratios of 0 2 3 σJ /J0 fill exactly the same envelope. This phenomenon 2 (∆12 − ∆23) ∆  |a2| = 2 + O 3 is due to the symmetry of the system. Specifically, for 8J0 J0 a pair of qubits, an equilateral triangle, or any number 2 3 2 2 (∆12 − ∆23) ∆  of qubits with a complete graph of equal exchange in- |a3| = − + O (18) 3 8J 2 J 3 teractions between every pair of qubits, the unperturbed 0 0 Hamiltonian H0 will be proportional to the total spin The return probability for a particular disorder real- 2 operator S . In order for the shape of the envelope to ization is given by: be affected by J0, there must be some noise term that mixes states with energy difference on the order of J0. However, S2 is conserved regardless of the disorder re- 2 X 2 −iEnt X 2 P (t) = |an| e = |anam| cos(En − Em)t alization, and thus only states with the same spin (and therefore the same energy under H ) can mix. In or- n n,m 0 (19) der for J to affect the shape of the oscillation envelope, 0 Then calculating the disorder average as in Eq. (3), it is necessary that the unperturbed Hamiltonian have we obtain [P (t)] as follows: eigenstates with the same spin and differing energies.

2 2 5 σ 2 2 σ J0t J −σJ t /16 J [P (t)] = − 2 + e 2 cos 9 3J0 6J0 2 2 III. 3-QUBIT LINEAR ARRAY 2 2 σ −σJ t /4 J + e 2 cos J0t 3J0 2 3 2 2 2 4 σ  3J0t σ σ t To demonstrate how qubit geometry affects decoher- −9σJ t /16 J J J + e − 2 cos + O 3 , (20) ence, we now consider a linear array of three dots in the 9 6J0 2 J0 J0 open geometry, given by the following Hamiltonian: In the limit where J0/σJ → ∞, this reduces to 2 2 5 4 −9σJ t /16 3J0t ∗ 9 + 9 e cos 2 , and thus the time T2 for the envelope width to drop to 1/e times its initial width is ~ ~ ~ ~ H = J12S1 · S2 + J23S2 · S3 (17) given by:

4 Because the system is lacking the same symmetry as T ∗ = σ−1, (21) before, an exact analytical expression for [P (t)] is not 2 3 J obtainable. We approach the problem in two ways: first, which is within 20% of the corresponding T ∗ in Eq. analytically by using perturbation theory for σ  J , 2 J 0 (16) for the triangular ring arrangement of the qubits. and second by showing the results of direct numerics. Note that there are three frequencies present in the ex- pression for [P (t)]. In general, these correspond to the energy differences between pairs of eigenstates of the un- perturbed Hamiltonian H0. The Gaussian decay factor A. Perturbative Approach for each frequency is determined by the first-order correc- tion to the energies. We show this by letting the cosine term of Eq. (19) take the form: We begin with the unperturbed Hamiltonian H0 given by Eq. (17) with J12 = J23 = J0. We then add per- turbations ∆ij = Jij − J0, and find the energies of the cos(En − Em)t = cos(δJ + c12∆12 + c23∆23)t eigenstates to order ∆. We also find the overlap a of i = cos δJ t cos(c ∆ + c ∆ )t each eigenstate with the initial state |↑↓↑i. These are 12 12 23 23 given by: − sin δJ t sin(c12∆12 + c23∆23)t (22) 7

Then for each variable ∆ integrated over to compute the disorder average, the cos c∆t term produces a Gaus- Z 1 2 2 2 2 2 2 −(∆ +∆ )/2σ −c t /2σ P (t) √ e 1 3 J δ(∆2 + J0)d∆1d∆2d∆3 sian factor of e J . There is physical meaning be- 2 ( 2πσJ ) hind the connection of these factors to the energies of Z 1 −(Σ2 +2|ξ |2−3J 2)/6σ2 2|ξ∆| the system. Because the oscillation frequencies are deter- = P (t) √ e ∆ ∆ 0 J √ × 2 mined by the unperturbed energies, a disordered system ( 2πσJ ) 3 3 will oscillate with the same frequencies as a clean sys- Σ + 2|ξ | cos ϕ  δ ∆ ∆ ∆ + J dΣ d|ξ |dϕ (23) tem, independent of the disorder strength present. How- 3 0 ∆ ∆ ∆ ever, the rate at which these oscillations decay is deter- mined by the sensitivity of these energies to disorder. where P (t) is given by Eq. (8). This integral is difficult Energy gaps which are less affected by disorder will pro- to evaluate analytically, but can be done numerically. In duce longer-lasting oscillations. This is the essence of Fig. 3, we show the plot of [P (t)] obtained perturbatively noise-induced decoherence in general. from Eq. (20) alongside the same numerical results. For σJ /J0 of 0.1 and 0.2, the plots are nearly identical; how- ever, significant differences can be seen for larger values such as 0.5, where the assumption σJ  J0 is no longer [P] valid, making the perturbation theory inaccurate. From 1.0 the numerical results, we note that for different values of J0, the oscillation envelopes tend to be mostly the same 0.8 in shape. However, there is some small dependence on σJ /J0=0.1 0.6 J0 in the linear chain, as is most evident by the fact that σJ /J0=0.2 the asymptotes of [P (t)] are different as J0 differs. Addi- σJ /J =0.3 0.4 0 tionally, comparing the locations of the peaks in Fig. 3, σJ /J0=0.5 0.2 we see that the peaks corresponding to σJ /J0 = 0.1 (yel- low) are slightly higher than the corresponding peaks for σJ t 0.0 0.5 1.0 1.5 2.0 2.5 3.0 other values of σJ /J0. This demonstrates that the strict independence of [P (t)] from J0 observed in the case of [P] the equilateral triangle above is a result of the symmetry 1.0 of the system, as breaking this symmetry causes some de- viance of the envelopes with J0. However, the results are σJ /J0=0.1 0.8 still reasonably independent of J0 making the analytical J σJ / 0=0.2 results of the triangular ring model approximately appli- 0.6 J σJ / 0=0.3 cable even in the absence of the ring symmetry. J 0.4 σJ / 0=0.5 J σJ / 0=1 [P] 0.2 σJ /J =2 0 1.0 σJ t 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.8 J 0.6 [P]: 0 /σJ =10 FIG. 3: Top: The expectation of return probability [P (t)] for envelope top envelope bottom different values of σJ , calculated from the perturbative result 0.4 given by eq. (20). Bottom: The same quantities calculated 0.2 exactly with numerics from Eq. (23). σJ t 0 1 2 3 4 5 [P] 1.0 0.8 [P]:J 0 /σJ =2 0.6 envelope top B. Numerical Results 0.4 envelope bottom 0.2 An alternative approach is to use the same expressions σJ t as the case in Sec. II with the equilateral triangle, but 0 1 2 3 4 5 enforcing J13 = 0 to be consistent with topology of the linear chain. This will correspond to setting ∆13 = −J0 FIG. 4: The expectation of return probability [P (t)] given by throughout the calculation. Then we arrive at a corre- eq. (25) for J0/σJ = 10 (Top) and J0/σJ = 2 (Bottom). sponding integral to Eq. (9): 8

IV. 4-QUBIT SQUARE 3-qubit ring, the high degree of symmetry in the Hamil- tonian protects states of different energies from mixing For comparison, we also consider a system of four due to disorder, and so the shape of the oscillation en- dots in a square configuration, which gives the following velope is completely independent of the mean exchange Hamiltonian: strength J0, instead depending only on the strength of the disorder σJ . We compare our results to the analogous two-qubit H = J12S~1 · S~2 + J23S~2 · S~3 + J34S~3 · S~4 + J41S~4 · S~1 (24) calculation in Ref. 19, where the return probability was found to be given by: There is some choice of initial state, but we will use an antiferromagnetic state of alternating ↑ and ↓ spins, since 1 1 −σ2 t2/2 this will capture the dynamics of the exchange interac- [P (t)] = + e J cos J0t (26) tion between adjacent spins. This is a reasonable choice, 2 2 but a different choice would not change the theory at all, Here the oscillation envelope is strictly Gaussian, and only the quantitative details of the results. Then begin- we showed that with higher-qubit systems, the envelope ning with the initial state |Ψ0i = |↑↓↑↓i and following a keeps its Gaussian-like characteristics, though the shape perturbative approach identical to Sec. III A, we obtain is no longer exactly Gaussian. In Table I, we calculate the following disorder averaged return probability: the time tc needed for the width of the envelope to reach a fraction of its initial value, for each system and for var- ∗ 7 13σ2 ious coherence fraction. T2 is given by the time required [P (t)] = − J for the envelope width to decay to 1/e of its original value 18 24J 2 0 (the bottom row in Table I), and the times corresponding  2 2  1 2 2 2σ 2 2 5σ 2 2 −σJ t /8 J −σJ t /4 J −σJ t /8 to values such as .99 are used to benchmark gates with + e + 2 e − 2 e cos J0t 3 3J0 16J0 that fidelity. For the 3-qubit line and 4-qubit ring, the  2 2  envelope shape depends partly on J0, so we give results 1 −σ2 t2/2 σJ −5σ2 t2/8 σJ −σ2 t2/2 + e J + e J − e J cos 2J0t in the limit that J /σ → ∞. Because of the Gaussian- 6 3J 2 8J 2 0 J 0 0 like nature of the envelopes, for t  σ−1, the envelopes  2   3 2  J 1 2 2 σ 2 2 σ σj t −9σJ t /8 J −9σJ t /8 J have a parabolic shape, as demonstrated by expansions + e − 2 e cos 3J0t + O 3 , 9 48J0 J0 J0 such as Eq. (15). This is evident in Table I, as changing (25) the target decoherence by a factor of 100 changes tc by roughly a factor of 10. We also see that increasing the The unperturbed Hamiltoninan has energies of J0, 0, number of qubits from two to three decreases the coher- −J0 and −2J0, the differences of which lead to the three ence time, but the 4-qubit ring has increased coherence frequencies present in [P (t)]. A four qubit linear array time from the others. It is important to note that the breaks some symmetry of the system, and causes degen- 3-qubit line and 4-qubit ring have multiple frequencies erate states to split into six different energies, causing present in [P (t)] which decay at different rates, and thus many more frequencies to appear. the coherence times are heavily dependent on the ini- tial state, as a different initial state will lead to different weights for each frequency. The fact that the coherence V. DISCUSSION time decreases by roughly 10% in going from 2 to 3 qubits is of considerable significance indicating that maintaining Using exact analytical methods, we have derived ex- coherence in larger qubit systems would be increasingly a pressions for the disorder averaged return probability challenge unless the system has special symmetries (e.g. from a given initial state for a system consisting of a 3 and 4 qubit rings) which can be exploited to enhance ring of three exchange coupled qubits. We have also cal- coherence. In addition, the fact that decoherence hap- culated quantities such as the normalized Hamming dis- pens through a Gaussian temporal decay and not expo- tance, qubit spin expectation values, and entanglement nentially is also of substantial significance. We should entropy for each qubit. We then used a perturbative mention that scaling up number of qubits necessitates approach to analyze a linear array of three qubits, com- careful consideration of how decoherence changes, since paring our results to direct numerics, and extended the a short decoherence time generates larger errors making same perturbative approach to a ring of four qubits. The fault-tolerant quantum gate operations increasingly more frequencies by which the disorder averaged return proba- challenging. bility oscillates are determined by the energy differences present in the unperturbed Hamiltonian. These frequen- cies have a Gaussian-like decay, in contrast to the usu- Acknowledgments ally assumed exponential decoherence ansatz, with their decay rate generally dependent on the first-order correc- This work is supported by the Laboratory for Physical tion to the corresponding energies. For the system with a Sciences. 9

fraction 2-qubit tcσJ 3-qubit ring tcσJ 3-qubit line tcσJ 4-qubit ring tcσJ 0.9999 0.0141 0.0115 0.0133 0.0163 0.999 0.0447 0.0365 0.0422 0.0517 0.99 0.142 0.116 0.134 0.164 0.9 0.459 0.374 0.433 0.550 1/e 1.41 1.127 1.333 2.389

TABLE I: Decay rates for the oscillation envelopes for systems of various numbers of qubits. The values in the table give the −1 time tc, in units of σJ , for the envelope width to decay to a given fraction of its original value. The fraction 1/e corresponds to ∗ the value of T2 for the given system. For the 3-qubit linear array and the 4-qubit ring, values are given for the limit J0/σJ → ∞.

[1] J. M. Nichol, L. A. Orona, S. P. Harvey, S. Fallahi, G. C. S. Dzurak. Nature 569, pp. 532-536 (2019). Gardner, M. J. Manfra, & A. Yacoby. npj Quant. Inf. 3, [13] X. Xue, T. F. Watson, J. Helsen, D. R. Ward, D. E. Sav- 3 (2017). age, M. G. Lagally, S. N. Coppersmith, M. A. Eriksson, [2] D. M. Zajac, T. M. Hazard, X. Mi, E. Nielsen & J. R. S. Wehner, & L. M. K. Vandersypen. Phys. Rev. X 9, Petta. Phys. Rev. App. 6, 054013 (2016). 021011 (2019). [3] A. J. Sigillito, M. J. Gullans, L. F. Edge, M. Borselli & [14] X. Hu & S. Das Sarma. Phys. Rev. Lett. 96, 100501 J. R. Petta. npj Quant. Inf. 5, 110 (2019). (2006). [4] T. F. Watson, S. G. J. Philips, E. Kawakami, D. R. Ward, [15] O. E. Dial, M. D. Shulman, S. P. Harvey, H. Bluhm, V. P. Scarlino, M. Veldhorst, D. E. Savage, M. G. Lagally, Umansky & A. Yacoby. Phys. Rev. Lett. 110, 146804 M. Friesen, S. N. Coppersmith, M. A. Eriksson & L. M. (2013). K. Vandersypen. Nature 555 pp. 633-637 (2018). [16] M. Russ, J. R. Petta & G. Burkard. Phys. Rev. Lett. [5] J. P. Dehollain, U. Mukhopadhyay, V. P. Michal, Y. 121, 177701 (2018). Wang, B. Wunsch, C. Reichl, W. Wegscheider, M. S. [17] X. Wang, L. S. Bishop, J. P. Kestner, E. Barnes, K. Sun Rudner, E. Demler & L. M. K. Vandersypen. Nature 579, & S. Das Sarma. Nat. Com. 3, 997 (2012). pp. 528-533 (2020). [18] J. Zeng, X.-H. Deng, A. Russo & E. Barnes. New J. Phys. [6] Y. Wang, J. P. Dehollain, F. Liu, U. Mukhopadhyay, M. 20, 033011 (2018). S. Rudner, L. M. K. Vandersypen & E. Demler. Phys. [19] R. E. Throckmorton, E. Barnes & S. Das Sarma. Phys. Rev. B 100, 155133 (2019). Rev. B 95, 085405 (2017). [7] D. Buterakos & S. Das Sarma. Phys. Rev. B 100, 224421 [20] S. Das Sarma, R. E. Throckmorton & Y.-L. Wu. Phys. (2019). Rev. B 94, 045435 (2016). [8] F. van Riggelen, N. W. Hendrickx, W. I. L. Lawrie, M. [21] Y.-L. Wu & S. Das Sarma. Phys. Rev. B 96, 165301 Russ, A. Summak, G. Scappucci & M. Veldhorst. Appl. (2017). Phys. Lett. 118, 044002 (2021). [22] Y.-L. Wu & S. Das Sarma. Phys. Rev. A 93, 022332 [9] N. W. Hendrickx, W. I. L. Lawrie, M. Russ, F. van Rigge- (2016). len, S. L. de Snoo, R. N. Schouten, A. Sammak, G. Scap- [23] R. E. Throckmorton & S. Das Sarma. Phys. Rev. B 102, pucci & M. Veldhorst. arXiv:2009.04268 (2020). 035439 (2020). [10] Lieven Vandersypen, private communication. [24] W. M. Witzel, M. S. Carroll, A. Morello,L.Cywi´nski& [11] J. Yoneda, K. Takeda, T. Otsuka, T. Nakajima, M. R. S. Das Sarma. Phys. Rev. Lett. 105, 187602 (2010). Delbecq, G. Allison, T. Honda, T. Kodera, S. Oda, Y. [25] D. P. DiVincenzo, D. Bacon, J. Kempe, G. Burkard & Hoshi, N. Usami, K. M. Itoh & S. Tarucha. Nature Nan- K. B. Whaley. Nature 408, pp. 339-342 (2000). otechnology 13, pp. 102-106 (2018). [26] R. W. Hamming. Bell Sys. Tech. Journ. 29, 2, pp. 147- [12] W. Huang, C. H. Yang, K. W. Chan, T. Tanttu, B. 160 (1950). Hensen, R. C. C. Leon, M. A. Fogarty, J. C. C. Hwang, F. E. Hudson, K. M. Itoh, A. Morello, A. Laucht & A.