<<

Long-lived in driven many- systems: from two to infinite spatial

Walter Hahn1, 2,a and V. V. Dobrovitski1, 3,b 1QuTech, Delft University of Technology, Lorentzweg 1, 2628 CJ Delft, The Netherlands 2Institute for Quantum and of the Austrian Academy of Sciences, Innsbruck, Austria 3Kavli Institute of Nanoscience, Delft University of Technology, Lorentzweg 1, 2628 CJ Delft, The Netherlands Long-lived coherences, emerging under periodic pulse driving in the disordered ensembles of strongly interacting spins, offer immense advantages for quantum technologies, but the phys- ical origin and the key properties of this phenomenon remain poorly understood. We theoretically investigate this effect in ensembles of different dimensionality, and predict existence of the long-lived coherences in all such systems, from two-dimensional to infinite-dimensional (where every spin is coupled to all others with similar strength), which are of particular importance for quantum sensing and quantum information processing. We explore the transition from two to infinite dimensions, and show that the long- coherence dynamics in all dimensionalities is qualitatively similar, although the short-time behavior is drastically different, exhibiting dimensionality-dependent singularity. Our study establishes the common physical origin of the long-lived coherences in different - alities, and suggests that this effect is a generic feature of the strongly coupled spin systems with positional disorder. Our results lay out foundation for utilizing the long-lived coherences in a range of application, from quantum sensing with two-dimensional spin ensembles, to quantum information processing with the infinitely-dimensional spin systems in the cavity-QED settings.

I. INTRODUCTION The long-lived coherences could be of great benefit for new quantum technology platforms. For instance, Collective quantum coherences of many-spins systems promising platforms for quantum sensing utilize two- play central role in quantum science and technology. But dimensional systems (d = 2, see Fig.1), such as surface quantum coherence is fragile, and extending its lifetime spins or 2D layers of NV spins [14–17]), which can be is a critical problem. For instance, in spin ensembles the brought close to the system being sensed, thus improv- collective coherence (collective transverse polarization) is ing resolution and sensitivity. Employing ensembles of destroyed by dipolar interactions [1–4], and the the spin spins boosts the total signal, and thus greatly improves echo signal, which quantifies coherence, quickly decays on the signal-to-noise ratio, but the collective coherence de- cays quickly due to dipolar coupling between the spins. the time scale T2 [5]. Coherence can be preserved e.g. via pulse and/or continuous-wave decoupling that suppresses Increasing the lifetime of collective coherences would be dipolar interactions [1–3]. Recently, an intriguing alter- of enormous benefit for quantum sensing. On the op- native has attracted much attention: it exploits, rather posite end (d → ∞) are the spin ensembles in a cavity than fights, the spin-spin interactions. Namely, the un- QED-type settings, actively explored for quantum infor- usual many-spin states, which are formed in ensembles of mation applications [18, 19], where each spin is coupled dipolar-coupled spins under periodic driving by π-pulses, to all others with a similar strength via collective pho- exhibit collective spin coherences living up to 105 tonic, phononic, or magnonic mode [20–26]. Taking full ∗ 4 advantage of the long-lived coherences could increase the longer than the T2 and about 10 time longer than the signal-to-noise ratio in these systems by orders of mag- T2 time [6–9]. This phenomenon, along with other sim- ilar effects, has been observed in various -state nu- nitude. In order to achieve that, detailed understanding clear magnetic resonance (NMR) experiments [6,7,9– the long-lived coherences in systems of different spatial 13], but still remains poorly understood. The long-lived dimensionality d is required. So far, even existence of the long-lived coherences at d = 2 or d → ∞ has remained

arXiv:1911.06272v3 [quant-ph] 12 Jul 2021 coherences emerge from the combination of strong dipo- lar coupling, disorder, and pulse imperfections (under- elusive, and their properties have been unknown. stood broadly as deviations of the real π-pulses from per- In this article we predict that the long-lived coherences fect instant 180◦ rotations) [6, 12]. Besides, the long-lived do exist in these important systems, thus opening the coherences demonstrate subharmonic response, i.e. asym- way to employing them in novel quantum information metry in the magnitudes of even and odd echoes [6,9, 12]. platforms. In order to analyze in detail the transition Both effects of long coherence lifetime and its subhar- from d = 2 to d → ∞, and to clarify generic features of monic response are remarkably stable against perturba- the long-lived coherence dynamics, we numerically simu- tions and decoherence. late the dynamics of disordered dipolar-coupled quantum spin ensembles of different dimensionalities, subject to periodic driving by imperfect π-pulses, with the rotation ◦ a [email protected] angle slightly deviating from 180 . For all dimensionali- b [email protected]; author to whom correspondence ties d studied, we observe the long-lived coherences and should be addressed see emerging subharmonic response when the time inter- 2

ish [6,9, 12]. The relation between the long-lived coher- ence and the time dynamics is poorly explored, in spite of its fundamental interest. Besides, the stud- ies of time so far have mostly focused on d = 1 [35, 37, 38, 43–50] and d = 3 systems [39–42, 51], as well as on the systems with d → ∞ [21, 36, 40, 50–52]. For these reasons, we also explore here the time crystal dynamics of the longitudinal spin polarization for differ- ent dimensionalities. We find that it also demonstrates time crystal-like oscillations with a very long lifetime for all d, but the physical features are markedly different from those of the long-lived coherences. More detailed exploration of this issue in the future would be of utmost interest, but is beyond the scope of the paper. Dimensionality plays a key role in the dynamics of dipolar-coupled positionally disordered spin systems. Such systems exhibit characteristic d-dependent dynam- ical singularities: the spin dynamics at short times is strongly singular for d = 2, while for d → ∞ the singular- ity disappears [2, 53–56]. Besides, dimensionality is well known to be decisive in the context of localization and thermalization dynamics in spin ensembles [28, 38, 57– Figure 1. (a) Spin ensembles of different dimensionality: 59]. Since so many key features of the spin dynamics two-dimensional, three-dimensional, and (effectively) infinite- depend on d, one would expect that the properties of dimensional systems. The latter system with all-to-all in- teractions of similar strength is realized e.g. by coupling the long-lived coherences would also strongly depend on d. spins to a detuned collective photon// mode Surprisingly, our results demostrate that this expecta- in a cQED-like setting. (b) Schematic representation of the tion is incorrect. periodic pulse sequence. Imperfect π pulses (P) are applied at times (2n + 1)τ (n ≥ 0), so that a series of echoes (E) is formed at times 2nτ. Each pulse rotates the spins along the II. QUALITATIVE DISCUSSION OF THE x-axis by an angle π(1 + ), where  is the pulse imperfection EFFECT parameter.

We study an ensemble of Ns spins Si = 1/2 (i = 1 ...Ns) in a standard setting of a magnetic resonance- val τ between the pulses increases to become comparable type experiments. Namely, the spin system is placed in to T2, see Fig.1b. Our simulations show that the magni- a strong quantizing magnetic field HQ directed along the tude of the long-lived coherence decreases at larger d, but z-axis [1]; this field induces fast spin precession with Lar- still remains quite large even at d → ∞. By analyzing mor frequency ωQ, which is much larger than all other the Floquet operator, we establish the kinematic origin frequency scales of the problem [60]. Following the stan- of the long-lived coherences, determine that it is similar dard theory of magnetic resonance, we describe spin dy- for all dimensionalities, and identify the states that are namics in the coordinate frame that rotates around the involved in their formation. z-axis with the circular frequency ωQ, and retain only Some aspects of the long-lived coherence resemble the the secular terms in the system’s Hamiltonian, which re- time crystal dynamics in periodically driven spin sys- main static in the rotating frame [61], or vary slowly in tems [27–35], with their characteristic robustness [36– comparison with ωQ [1,2]. 38]. Time crystals have been observed in many spin sys- Initially, by applying a preparatory π/2 pulse, the spins tems, from trapped ions [39] to spin ensembles in dia- are prepared in a state weakly polarized along the x-axis mond [40], including the kind of NMR systems that ex- of the rotating frame, such that the initial ensemble den- hibits the long-lived coherences [35, 41, 42]. However, the sity matrix is ρ(t = 0) ∝ I − µMx, where I is identity physical origins of the two phenomena are different: the matrix, µ  1 is a parameter determining the absolute P coherences are determined by the transverse magnetiza- polarization of the ensemble, and Mx = j Sjx is the tion, while the time crystal dynamics refers to the behav- collective coherence operator. Here and below, Sjα with ior of the longitudinal polarization. They are governed α = {x, y, z} denotes the component of the j-th spin by different processes, and their lifetimes can differ by along the rotating-frame axis α. In experiments, the en- many orders of magnitude [1,2]; e.g., in the case of per- semble coherences along the x- and y-directions are quan- ◦ fect pulses (instantaneous 180 rotations) the time crys- tified by the total transverse magnetizations Mx and My tal oscillations have the longest lifetime and largest am- along the corresponding axes, so we use the terms “coher- plitude, while the long-lived coherences completely van- ence” and “transverse magnetization” interchangeably. 3

1 rotation Siz → −Siz, Siy → −Siy for all spins at once). (a) short τ A pulse applied at t = τ restores collective coherence, 0.8 producing Hahn echo signal at t = 2τ [1,2] (see Fig.1b). periodic driving However, the hard π pulses do not affect the spin-

) Hahn echo spin interaction that destroys the ensemble coherence

t 0.6

( 2 by entangling different spins [1–4]. In relevant experi- x exp( t 3 ) ments, the dominant interaction is the dipolar coupling, 0.4 − M described by the Hamiltonian [1,2]

0.2 Ns X X HI = (Jij/2) (2SizSjz − SixSjx − SiySjy) , (1) 0 i=1 j>i 1 2 3 (b) long τ where Jij = (1 − 3 cos θij)/rij is the coupling constant 0.8 between the spins i and j, rij = |~rij| is the distance between them, and θij is the polar angle of ~rij. Under the periodic driving

) influence of HI, the Hahn echo signal gradually decays as

t 0.6 ∗ ( Hahn echo a function of time t = 2τ at the timescale t ∼ T2  T2 . x 2 Note that the dipolar interaction is long-ranged, so 0.4 3 M exp( t ) that each spin is coupled to all other spins for all systems − considered here, for all values of d from 2 to ∞, and the 0.2 coupling strength Jij decays with distance rij between the spins i and j in the same way for all d. However, the 0 statistical properties of the values Jij greatly vary with 0 5 10 15 20 the system’s dimensionality, thus leading to dramatically t different decay of the Hahn echo signal. The rate of the Hahn echo decay is governed by the

Figure 2. Long-lived coherence Mx(t) for two-dimensional positional disorder of spins. In relevant experiments [9– d = 2 periodically driven disordered spin systems, for short 11, 14–19, 40], the spins are very dilute: they occupy a inter-pulse time delay 2τ (τ/T2 ≈ 0.07) (a) and for long delay small fraction of the lattice sites, and are randomly dis- 2τ (τ/T2 ≈ 0.7) (b). The long-lived train of echoes Mx(t) tributed in the sample. As a result, each spin Si has its (orange lines) extends far beyond the T2 = 1 time. The single- own set of dipolar couplings Jij to the other spins, such pulse Hahn echo (black solid lines) was obtained by simulating that different spins “feel” different environments made of a π pulse applied at time τ = t/2; the analytical result for other spins. In low-dimensional ensembles these spin-to- 2/3 the Hahn echo in d = 2 is Mx(t) = exp(−t ), and is in spin variations are very strong [53–55], leading to a very excellent agreement with numerics. The amplitude of the fast decay of the collective coherence. In AppendixA we long-time tails is larger for short τ. For long τ, the even- show that, in the limit of T  T ∗, when the flip-flop odd echo asymmetry becomes pronounced. The system size 2 2 ∗ terms can be omitted and the Hamiltonian (1) acquires is Ns = 20, other parameters are  = 0.07 and T2 ≈ 0.02. P the form JijSizSjz, the Hahn echo signal hMx(t)i de- cays with time as

The longitudinal polarization, exhibiting time crystal-  d/3 like behavior (see Sec.V), is quantified by the magne- hMx(t)i ∝ exp − | t/T2| . (2) P tization Mz = j Sjz along the z-axis. Note that in this work we vary only the spatial dimensionality d of the en- for d ≤ 5, such that for d = 2 the initial decay is infinitely semble, while spins themselves remain embedded in three fast. For larger d the fluctuations are not as strong, and dimensions, i.e. have three orthogonal components. the singularity of hMx(t)i at t = 0 is weak for d = 4 and In typical experiments, the spins experience random 5. In the limit d → ∞ each spin is coupled to all others quasi-static local magnetic fields, described by the Hamil- with almost uniform coupling, so the echo decay acquires h 2i PNs Gaussian form hM (t)i ∝ exp − (t/T ) , without any tonian HL = j=1 hjSjz; everywhere in this article x 2 we set ~ = 1 and normalize the spins’ gyromagnetic singularity at all. ratio to γ = 1.[62] These fields cause fast dephas- The total Hamiltonian of the system, taking into ac- ing: the x-component of each spin Sjx oscillates at its count the pulse driving, is own rate proportional to hj, and the collective coherence ∗ hMx(t)i = Tr[ρ(t)Mx] vanishes at the timescale T2 . This H(t) = HI + HL + HP(t), (3) is usually too short for practical needs, and dephasing is suppressed by applying a number of hard π pulses, where HP(t) describes periodic driving by a train of (gen- which reverse the sign of the Hamiltonian HL (ideal, i.e. erally imperfect) π-pulses, as shown in Fig.1b. If the instantaneous 180◦ hard pulse along the x-axis performs pulses were ideal, they would suppress the dephasing 4 term HL (see AppendixA for details), and leave the dipo- 1 lar interaction H intact. Thus, the echo signal hM (t)i (a) I x 0.8 d = 5 would not depend on the number of pulses applied during periodic driving ) the time t, and the echo decay would follow Eq.2. t 0.6 ( Hahn echo

Our results show that for non-ideal pulses this is true x 5 at short times: the dynamics drastically depends on d, 0.4 exp( t 3 ) M − with the initial decay rate varying from infinity at d = 2 0.2 to zero at d → ∞. At long times, for spin ensembles of all dimensionalities d, the spin dynamics is controlled 0 by accumulation of the pulse imperfections. This process 1 depends on the specific pulse sequence [6,7,9]; here we (b) 0.8 d = 8 focus on the simple and efficient Carr-Purcell-Meiboom- periodic driving )

Gill (CPMG) protocol shown in Fig.1b. Accumulation t 0.6 ( Hahn echo of the pulse errors, combined with the dipolar spin-spin x coupling, leads to long-lived tails in the echo signal, ex- 0.4 exp( t2) tending far T time, in the region where the Hahn M − 2 0.2 echo has already vanished. The magnitude of the long- lived coherence tails is particularly large when the inter- 0 pulse time delay τ is short compared to T2, as seen in 1 Fig.2a for d = 2. The effect is remarkably robust, and (c) 0.8 d does not disappear even when τ becomes comparable to → ∞ periodic driving T . In that regime, another feature emerges for all val- ) 2 t 0.6 ues of d: the long-lived coherences exhibit pronounced ( Hahn echo x subharmonic response (Fig.2b), with even echoes being 0.4 exp( t2) larger than odd ones. This behavior reflects the fact that M − the period of the operator for CPMG proto- 0.2 col is twice the period of the Hamiltonian HP(t) of the 0 CPMG pulse train [3] (see Sec.III). 0 5 10 15 20 We use direct numerical solution of the many-spin time-dependent Schrödinger equation with the second- t order Suzuki-Trotter [63] or Chebyshev [64] expansion of Figure 3. Long-lived coherence Mx(t) for periodically driven the evolution operator U(t) for up to Ns = 24 spins. disordered spin systems of different spatial dimensionalities d, In all situations that we tested, both methods give con- indicated in the figure. The echoes Mx(t) extend far beyond sistent results. The initial mixed-state density matrix the T2 time. For the single-pulse Hahn echo, the π-pulse is is represented by a random pure state, i.e. as a ran- applied at time τ = t/2. with increasing d, the Hahn echo dom unit-norm complex-valued vector of length 2Ns , uni- approaches Gaussian form. The simulation parameters are Ns ∗ formly sampled from a sphere S2 −1 of unit radius. We Ns = 20, short τ ≈ 0.07,  = 0.07 and T2 ≈ 0.02. calculate the experimentally measured normalized mag- netization response III. DYNAMICS OF THE LONG-LIVED COHERENCES Mα(t) = hMα(t)i/hMα(0)i, (4)

† where α = {x, y, z} and hMα(t)i = Tr[MαU(t)MαU (t)] The behavior of Mx(t) in pulse driven spin ensembles, for initial polarization along the axis α. The spins are as described above, is shown in Fig.2 for d = 2, and randomly placed in a d-dimensional cube, and averag- in Fig.3 for larger d. For reference, we presented the ing is performed over 100–360 realizations of the spatial results for previously unexplored cases of d = 2, 5, 8, and arrangements and values of local fields. The cube’s edge d → ∞. For other dimensionalities d that we studied, all length (i.e. the spin density fs) is adjusted to have T2 = 1 results remain qualitatively the same. [65][66] for each system after averaging (see AppendixA for re- We consider two different values of the inter-pulse de- lation between T2 and fs). T2 is defined as the time lay τ: short τ ≈ 0.07 T2, and long τ ≈ 0.7 T2 (recall when the echo magnitude decreases by the factor e. In that T2 = 1). At short times t . T2, the system’s re- experiments the pulse imperfections may arise from de- sponse Mx(t) closely follows the Hahn echo, and is in liberate or accidental miscalibration, or also due to local excellent agreement with our analytical predictions (see fields and dipolar interactions, which affect the spin rota- AppendixA for details), as seen in the figures (the small tion during the finite-duration pulses [6,9, 12]. For hard differences between the simulated Hahn echo decay and short pulses, imperfections are small, and in this arti- the analytical predictions are due to the finite-size ef- cle we model nonideal pulses as instantaneous rotations fects). At longer times t  T2, the magnetization re- around the x-axis by an angle π(1 + ) with   1. sponse Mx(t) exhibits long-time tails for all dimension- 5 alities d considered, and for both short and long τ. These long-time tails extend far beyond the T2 time, and in experiments are likely to be limited by the spin-lattice relaxation time T1. With increasing d, the overall ampli- tude of the long-time echoes becomes somewhat smaller. Still, even in the limit d → ∞, the long-time tails do not vanish but saturate at a nonzero value. Also, the amplitude of the long-time tails becomes smaller as the inter-pulse delay τ increases. When the inter-pulse delay τ is large, comparable to T2, the long-lived coherences demonstrate pronounced subharmonic dynamics (Fig.2b), where the period of the magnetization response 4τ is twice longer than the period of driving 2τ. The subharmonic response was observed for all values of d we modeled. This feature can be ratio- nalized with the notion of the cycling period of a pulse sequence [3]: for the control Hamiltonian HP(t), the cy- cling period tc is defined by the condition of periodicity of the evolution operator, i.e. U1(t) = U1(t + tc), where R t 0 0 U1(t) = T exp[−i 0 HP(t )dt ]. It is known that the cy- cling period of a pulse sequence can differ from the period of the Hamiltonian [3], e.g. for ideal π-pulses, the cycling period of the CPMG protocol is tc = 4τ, and includes two π-pulses [67], i.e. contains two periods of the underlying control Hamiltonian HP(t). We also note that the even- odd asymmetry is present for short τ, but its amplitude is too small to be easily seen. It is important to point out that the long-lived co- herences and the subharmonic response crucially de- pend on the driving sequence. For instance, for an alternating- Carr-Purcell (APCP) sequence, where jk 2 the direction of driving alternates between the positive Figure 4. The matrix Mx = |hψj |Mx|ψki| for short τ (a) and the negative x-direction, no long-lived tails of Mx(t) and long τ (b) for a two-dimensional d = 2 system. On the emerge [6]. vertical and the horizontal axes, respectively, the quasiener- gies φj and φk of the corresponding Floquet eigenstates |ψj i and |ψki are plotted. For short τ (panel a), a large number of large entries concentrate on the diagonal φ ≈ φ , produc- IV. FLOQUET OPERATOR ANALYSIS OF THE j k ing long-lived coherence with noticeable amplitude. For long LONG-LIVED COHERENCES τ (panel b), the values at semi-diagonals |φj − φk| ≈ π be- come comparable to the values on the diagonal, which leads Periodically driven systems, described by a Hamilto- to noticeable subharmonic response. Overall, the values for nian obeying H(t + 2τ) = H(t), can be analyzed using long τ are smaller than those for short τ. The system size Floquet theory. The stroboscopic time evolution of the is Ns = 14. Only values larger than 0.25 are included in the system’s density matrix, considered only at times that figures. are integer multiples m of the driving period 2τ, can for- m †m mally be written as ρ(2τm) = UF ρ(0)UF , where UF is time evolution operator per one period of the driving so the magnetization response after m driving periods is Hamiltonian (Floquet operator). For the CPMG pulse 4 X i(φj −φk)m 2 sequence considered here, the Floquet operator has the Mx(2τm) = N e |hψj|Mx|ψki| . (6) Ns2 s form UF = UH(τ)Rx[π(1 + )]UH(τ), where Rx[π(1 + )] j,k is the operator of rotation around the x-axis by an angle π(1 + ), and UH(τ) is the operator of evolution under The signal Mx(2τm) is therefore mainly determined by the action of the system’s internal Hamiltonian HI + HL. two quantities: firstly, by the magnitude of the matrix elements M jk = |hψ |M |ψ i|2, and, secondly, by the As a unitary operator, UF possesses a complete or- x j x k distribution P (φ −φ ) of the quasienergy differences φ − thonormal set of eigenstates |ψki, with complex eigen- j k j values eiφk of unit modulus: φk, i.e. by the number of Floquet eigenstates |ψji and |ψki with a given difference in the quasienergies φj and X iφk UF = e |ψkihψk|, (5) φk. The terms with j = k in Eq. (6) do not depend k on m, so the long-time response Mx(2τm) is governed 6 (a) 16000

) 12000 φ

(∆ 8000 P 4000 0 (b) 16000 6000 4000 ) 12000 φ 2000

(∆ 8000

P 0 4000 3.1 π 0 π 0 2 π ∆φ

Figure 6. Histogram of the quasienergy differences ∆φ for a d = 2 disordered spin system with Ns = 14, at short τ (a) and long τ (b). The sharp peaks at ∆φ = 0 and at ∆φ = π correspond to the long-lived coherences and to the subharmonic response, respectively. Specifically, P (∆φ) is the total number of the Floquet eigenstates |ψj i and |ψki having a given difference ∆φ in their quasienergies φj and φk. The quantity ∆φ is downfolded to the interval ∆φ ∈ [0, π] and binned, so that P (∆φ) includes all states whose quasienergies satisfy the condition |φj − φk| ∈ [∆φ − β, ∆φ + β] or 2π − −5 |φj − φk| ∈ [∆φ − β, ∆φ + β], where 2β = π · 10 is the Figure 5. Same as Fig.4 but for a five-dimensional d = 5 width of a bin. To avoid double counting, only the states system. Only values larger than 0.1 are included in the fig- with φj ≥ φk are included in P (∆φ). The results shown are ures. The same features as in the case d = 2 are seen along averaged over many realizations of the disorder. The inset in the diagonal φj ≈ φk and at the semi-diagonals |φj −φk| ≈ π; panel (b) shows the magnified view of the peak at ∆φ ≈ π. these features correspond to the long-lived coherences and The simulation parameters are the same as in Fig.2. to the subharmonic response, respectively. Similarity in the jk structure of the matrices Mx for ensembles of different di- mensionalities confirms similarity in the physical origin and jk behavior of the long-lived coherences for different d. For long τ, the matrix Mx exhibits large entries both on the diagonal, and on the two semi-diagonal lines cor- responding to φj − φk ≈ ±π; the latter correspond to emerging even-odd echo asymmetry seen in Fig.2b. The jj jk by the diagonal elements Mx of the matrix Mx . The semi-diagonals also show comparable contributions from subharmonic response (the even-odd echo asymmetry) is a large number of pairs of Floquet eigenstates. In com- controlled by the pairs of Floquet eigenstates obeying parison with the case of short τ, the diagonal values i(φj −φk)m m jj φj − φk ≈ ±π, such that e ≈ (−1) . Mx on average are smaller for long τ, corresponding to jk 2 smaller amplitude of the long-time tails in Fig.2b. An example of the matrix Mx = |hψj|Mx|ψki| is jk shown in Fig.4 for short τ (a) and long τ (b) for a The matrices Mx for other d exhibit similar structure. jk two-dimensional system, for one typical realization of As an example, Fig.5 exhibits the results for Mx in the jk the positional disorder. The matrix Mx for short τ case of d = 5. The results of diagonalization of the Flo- is dominated by large diagonal elements, whereas the quet operator for different d evidence that the physical off-diagonal elements are almost negligible. This corre- origin of the effect of the long-lived coherences is similar sponds to the pronounced long-lived coherences in Fig.2a for all dimensionalities. and Fig.3, with the almost time-independent amplitude. The other quantity determining the signal Mx(2τm) It is clearly seen that the long-lived coherence contains in Eq. (6) is the distribution P (∆φ) of the quasienergy comparable contributions from a large number of Floquet differences |φj − φk|. In order to take into account the eigenstates, rather than being confined to a small subset symmetries of the summands in Eq.6, we downfold the of some special states. quantity ∆φ to the interval [0, π], i.e. we take ∆φ = 7

|φj − φk| when |φj − φk| ≤ π, and ∆φ = 2π − |φj − φk| 1 when |φj − φk| > π (so that ∆φ is the smallest angular (a) 1 distance between φj and φk on the S circle). The binned distribution P (∆φ) for d = 2 is shown in Fig.6, the bin 0.5 −5 width is 2β = π·10 . A peak at ∆φ ≈ π clearly emerges ) t for long τ. This means that the number of quasienergy ( z 0 pairs with |φj −φk| ≈ π becomes larger for long τ. Hence, the subharmonic response emerges not only because the M jk values of Mx on the semi-diagonals become larger, but 0.5 also because the total number of non-zero entries on the − semi-diagonals increases. 1 − 0 5 10 15 20

V. LONGITUDINAL MAGNETIZATION AND 1 THE INFINITE-TEMPERATURE TIME (b) CRYSTAL DYNAMICS 0.5

Let us now focus on the dynamics of the longitudi- ) t nal polarization Mz(t), which under some circumstances ( z 0 can exhibit robust long-lived infinite-temperature time crystal-like dynamics [27]. M Long-lived coherences and time crystal dynamics share 0.5 a number of similarities: both are induced by strong spin- − spin interactions, robust to experimental imperfections, 1 and demonstrate subharmonic dynamics under appropri- − 0 10 20 30 40 50 60 ate conditions. However, the underlying physics is to- tally different. Since the system’s internal Hamiltonian t H + H conserves the total z-magnetization, the signal I L Figure 7. Long-time tails of the z-component of collective Mz(t) remains constant between the pulses. If the π- magnetization Mz(t) in a d = 2 disordered spin system for pulses were ideal, then Mz(t) would just switch between short τ (a) and long τ (b) (orange lines and symbols), sug- +1 and −1. In contrast, the coherence Mx(t) would gesting time crystal-like behavior. Between the pulses, the quickly vanish under the action of the dipolar spin-spin value of Mz(t) is constant because the internal Hamiltonian coupling, along with the Hahn echo, at the timescale T2, HI + HL conserves the z-component of the total magnetiza- without any long-lived tail. tion. Each π-pulse flips the z-magnetization, so that the sub- For non-ideal pulses, one would expect the longitudi- harmonic response with the period 4τ is the dominating com- nal polarization to decay with increasing the number of ponent of the system’s long-time response. Without dipolar coupling, accumulated pulse error would modulate the magne- applied pulses. Indeed, for  > 0, after an imperfect pulse m the absolute value of M (t) would decrease by a factor tization along the z-axis, and Mz(t) would decay as cos (π) z after m pulses (dashed black line). However, in the presence cos(π), while the y-component increases by sin(π). Ac- of the dipolar coupling, Mz(t) exhibits long-time tails, al- cordingly, if the y-component has irreversibly dephased ternating between the directions “up” and “down” after each to zero during the time 2τ between pulses, then the ab- ∗ π-pulse. The system size is Ns = 20,  = 0.07, and T2 ≈ 0.02. solute value of Mz(t) would be expected to decay mono- tonically as cosm(π) with increasing the number m of applied pulses [35, 41]. The actual behavior of the longitudinal magnetization driving sequence. This long-lived subharmonic response is seen for all dimensionalities we studied: the results for response Mz(t) in a d = 2 disordered spin system with another example, a d = 8 spin ensemble, are presented in Ns = 20 is shown in Fig.7. Initial decay roughly fol- lows the expected cosm(π) pattern for both short and Fig.8, and are very similar, except that the amplitude of long τ (panels (a) and (b), respectively); the additional the long-time tail is somewhat smaller than in the d = 2 modulation seen in panel (a) is likely due to incomplete case. This conclusion is also consistent with the experi- dephasing of the y-component between the pulses due to mental evidence reported for d = 3 spin systems [40–42]. short τ. Note that the initial decay of Mz(t) does not exhibit At later times, however, the absolute magnitude of the singularities present in the short-time dynamics of Mz(t) stabilizes, showing little (if any) decay, and Mz(t) Mx(t), because it is governed by a different physical pro- itself just alternates between positive and negative val- cess, by accumulation of the rotation errors. Likewise, ues. The period of Mz(t) response is 4τ, which is twice the long-time behavior of the longitudinal Mz(t) and the period of the of the driving Hamiltonian HP(t), and the transversal Mx(t) polarization is also qualitatively equals, as expected, the cycling period of the CPMG different. For instance, with increasing the inter-pulse 8

1 (a)

0.5 ) t (

z 0 M 0.5 −

1 − 0 5 10 15 20

1 (b) jk Figure 9. A typical example of the matrix Mz = 2 0.5 |hψj |Mz|ψki| for a two-dimensional d = 2 system and short τ. The vertical and horizontal axes correspond to the quasiener- )

t gies φj and φk of the corresponding Floquet eigenstates |ψj i (

z 0 and |ψki. The entries are concentrated along the semi- diagonals φj − φk ≈ ±π, evidencing time crystal-like dynam- M ics with the doubled period 4τ. For long τ, the structure 0.5 of the matrix M jk is very similar (not shown). Only values − z larger than 1.0 are included in the figure. The system size is Ns = 15. 1 − 0 10 20 30 40 50 60 t exhibit striking similarity: both are induced by spin-spin interactions and demonstrate subharmonic dynamics un- Figure 8. Same as Fig.7 but for an eight-dimensional d = 8 der appropriate conditions. At the same time, these ef- disordered spin system. fects arise in very different regimes, demonstrate very different dynamics, and are differently affected by the pulse imperfections. Understanding the connection be- delay τ, the amplitude of the long-time tails of M (t) in- z tween time crystals (in particular, infinite-temperature creases, while the amplitude of M (t) decreases, thus x time crystals [35]) and long-lived coherences is an inter- clearly demonstrating the difference between the time esting and important, yet rather unexplored problem. crystal dynamics and long-lived coherences. b) The presence of long-time tails along the z-axis We have also analyzed the stroboscopic evolution of which is perpendicular to the pulse driving field implies M (t) by diagonalizing the Floquet operator; an exam- z that the fundamental process for establishing long-time ple of the corresponding matrix M jk = |hψ |M |ψ i|2 is z j z k tails of the magnetization response may not be spin lock- shown in Fig.9 for one typical realization of the posi- ing [1,2] as may appear in analogy with other similar tional disorder. As anticipated, it exhibits nonzero val- effects [10, 11, 68]. ues only at the semi-diagonals φ − φ ≈ ±π, in accor- j k c) Our simulations show that even if the direction of dance with the subharmonic response described above. the driving during the π-pulses is chosen randomly for The diagonal elements are negligible. Similar to the case each spin, the long-lived coherences emerge and persist of long-lived coherences, we see again that many states, for long times, as long as the direction of the driving re- distributed all over the Hilbert , contribute to the mains constant in time. This observation suggests that effect. the origin of the long-lived coherences is primarily kine- For the initial polarization along the y-axis, no long- matic; it may arguably also involve many-body localiza- lived magnetization response My(t) is seen in any di- jk tion [28, 31], transient prethermal regime [32, 69, 70], or mensionality d, and the elements of the matrix My = 2 spin-glass like behaviour. |hψj|My|ψki| are also generally small, without any clear d) The results shown in the present article were ob- structure. tained for a fixed pulse imperfection  = 0.07. Our nu- merical simulations with other values of  demonstrate that qualitatively the same behavior occurs for other VI. FINAL REMARKS AND CONCLUSIONS choices of   1. The long-time spin coherences per- sist even when  is chosen randomly for different spins, Before concluding, let us make some final remarks. as long as it remains constant in time. a) Time crystal dynamics and the long-lived coherences e) We considered in this article spin systems of finite 9 size, with the total number of spins Ns ≈ 20. Our simu- coupled disordered spin systems, including two-, three- lations demonstrate very modest quantitative changes as , and infinite-dimensional systems that are particularly the system size has been varied from Ns = 10 to Ns = 24. relevant for practical applications. Developing specific Besides, our numerical results for d = 3 agree with the protocols for such applications is an exciting avenue for reported experimental results for three-dimensional sys- further research. tems [6,9, 40, 41]. Summarizing, we have investigated disordered dipolar- coupled quantum many-spin systems of different spatial dimensionality subjected to periodic driving (CPMG pro- ACKNOWLEDGEMENTS tocol). Depending on the dimensionality, such systems exhibit singularities in short-time spin dynamics, which are caused by statistical fluctuations in the dipolar spin- The authors would like to thank S. E. Barrett, spin interaction strength. For all dimensionalities, we A. Bleszynski-Jayich, D. Budker, N. de Leon, E. observed the long-lived spin polarization along the driv- B. Fel’dman, B. V. Fine, G. E. Volovik, and N. ing pulses, and along the axis conserved by the internal Yao for stimulating discussions. This work was sup- Hamiltonian (z-axis). The amplitude of the long-lived ported by the DARPA DRINQS program (contract magnetization depends on the inter-pulse time delay and D18AC00015KK1934). This work is part of the research on dimensionality. The Floquet operator analysis shows programme NWO QuTech Physics Funding (QTECH, that the long-lived coherences Mx(t) contain compara- programme 172) with project number 16QTECH02, ble contributions from a large number of Floquet eigen- which is (partly) financed by the Dutch Research Council states. Our results imply that the long-lived coherences (NWO). The work was partially supported by the Kavli and subharmonic response are generic features of dipolar- Institute of Nanoscience Delft.

[1] C. P. Slichter, Principles of Magnetic Resonance [13] Y. Ivanov, B. Provotorov, and E. Fel’dman, Spin dynam- (Springer-Verlag, 1990). ics in multipulse nmr experiments, JETP Letters 27, 153 [2] A. Abragam, Principles of Nuclear Magnetism (Oxford (1978). University Press, 1961). [14] S. Steinert, F. Dolde, P. Neumann, A. Aird, B. Nayde- [3] U. Haeberlen, High Resolution NMR in (Academic nov, G. Balasubramanian, F. Jelezko, and J. Wrachtrup, Press, Inc., 1976). High sensitivity magnetic imaging using an array of spins [4] H. Cho, T. D. Ladd, J. Baugh, D. G. Cory, and C. Ra- in , Review of Scientific Instruments 81, 043705 manathan, Multispin dynamics of the solid-state nmr free (2010). induction decay, Phys. Rev. B 72, 054427 (2005). [15] D. Bluvstein, Z. Zhang, C. A. McLellan, N. R. Williams, [5] E. L. Hahn, Spin echoes, Phys. Rev. 80, 580 (1950). and A. C. B. Jayich, Extending the quantum coherence of [6] D. Li, Y. Dong, R. G. Ramos, J. D. Murray, K. MacLean, a near-surface by coherently driving the paramag- A. E. Dementyev, and S. E. Barrett, Intrinsic origin of netic surface environment, Phys. Rev. Lett. 123, 146804 spin echoes in dipolar solids generated by strong π pulses, (2019). Phys. Rev. B 77, 214306 (2008). [16] S. Sangtawesin, B. L. Dwyer, S. Srinivasan, J. J. Allred, [7] Y. Dong, R. G. Ramos, D. Li, and S. E. Barrett, Con- L. V. H. Rodgers, K. De Greve, A. Stacey, N. Dontschuk, trolling coherence using the internal structure of hard π K. M. O’Donnell, D. Hu, D. A. Evans, C. Jaye, D. A. Fis- pulses, Phys. Rev. Lett. 100, 247601 (2008). cher, M. L. Markham, D. J. Twitchen, H. Park, M. D. [8] M. A. Frey, M. Michaud, J. N. VanHouten, K. L. Insogna, Lukin, and N. P. de Leon, Origins of diamond surface J. A. Madri, and S. E. Barrett, Phosphorus-31 mri of noise probed by correlating single-spin measurements hard and soft solids using quadratic echo line-narrowing, with surface spectroscopy, Phys. Rev. X 9, 031052 (2019). Proceedings of the National Academy of Sciences 109, [17] M. de Wit, G. Welker, J. de Voogd, and T. Oosterkamp, 5190 (2012). Density and T1 of surface and bulk spins in diamond in [9] A. E. Dementyev, D. Li, K. MacLean, and S. E. Barrett, high magnetic field gradients, Phys. Rev. Applied 10, Anomalies in the nmr of silicon: Unexpected spin echoes 064045 (2018). in a dilute dipolar solid, Phys. Rev. B 68, 153302 (2003). [18] D. D. Awschalom, R. Hanson, J. Wrachtrup, and B. B. [10] L. Erofeev and B. Shumm, Experimental investigation Zhou, Quantum technologies with optically interfaced of relaxation processes in multipulse nmr experiments, solid-state spins, Nature Photon. 12, 516 (2018). JETP Letters 27, 149 (1978). [19] T. Astner, S. Nevlacsil, N. Peterschofsky, A. Angerer, [11] W.-K. Rhim, D. P. Burum, and D. D. Elleman, Multiple- S. Rotter, S. Putz, J. Schmiedmayer, and J. Majer, Co- pulse spin locking in dipolar solids, Phys. Rev. Lett. 37, herent coupling of remote spin ensembles via a cavity 1764 (1976). bus, Phys. Rev. Lett. 118, 140502 (2017). [12] W. Zhang, N. Konstantinidis, K. A. Al-Hassanieh, and [20] J. V. Cady, O. Michel, K. W. Lee, R. N. Patel, C. J. V. V. Dobrovitski, Modelling decoherence in quantum Sarabalis, A. H. Safavi-Naeini, and A. C. B. Jayich, Dia- spin systems, Journal of Physics: Condensed 19, mond optomechanical crystals with embedded nitrogen- 083202 (2007). vacancy centers, Quantum Science and Technology 4, 10

024009 (2019). 073028 (2019). [21] Z. Gong, R. Hamazaki, and M. Ueda, Discrete time- [39] J. Zhang, P. W. Hess, A. Kyprianidis, P. Becker, A. Lee, crystalline order in cavity and circuit qed systems, Phys. J. Smith, G. Pagano, I.-D. Potirniche, A. C. Potter, Rev. Lett. 120, 040404 (2018). A. Vishwanath, N. Y. Yao, and C. Monroe, Observation [22] D. Lee, K. W. Lee, J. V. Cady, P. Ovartchaiyapong, and of a discrete time crystal, Nature 543, 217 (2017). A. C. B. Jayich, Topical review: spins and mechanics in [40] S. Choi, J. Choi, R. Landig, G. Kucsko, H. Zhou, J. Isoya, diamond, Journal of Optics 19, 033001 (2017). F. Jelezko, S. Onoda, H. Sumiya, V. Khemani, C. von [23] H. Chen, N. F. Opondo, B. Jiang, E. R. MacQuarrie, Keyserlingk, N. Y. Yao, E. Demler, and M. D. Lukin, Ob- R. S. Daveau, S. A. Bhave, and G. D. Fuchs, Engineer- servation of discrete time-crystalline order in a disordered ing electron-phonon coupling of quantum defects to a dipolar many-body system, Nature 543, 221 (2017). semiconfocal acoustic resonator, Nano Letters 19, 7021 [41] J. Rovny, R. L. Blum, and S. E. Barrett, Observation (2019). of discrete-time-crystal signatures in an ordered dipolar [24] P. Andrich, C. F. de las Casas, X. Liu, H. L. Bretscher, many-body system, Phys. Rev. Lett. 120, 180603 (2018). J. R. Berman, F. J. Heremans, P. F. Nealey, and D. D. [42] S. Pal, N. Nishad, T. S. Mahesh, and G. J. Sreejith, Tem- Awschalom, Long-range spin wave mediated control of poral order in periodically driven spins in star-shaped defect in nanodiamonds, npj Quantum Inf. 3, 28 clusters, Phys. Rev. Lett. 120, 180602 (2018). (2017). [43] P. Ponte, Z. Papić, F. Huveneers, and D. A. Abanin, [25] Y. Dumeige, J.-F. Roch, F. Bretenaker, T. Debuisschert, Many-body localization in periodically driven systems, V. Acosta, C. Becher, G. Chatzidrosos, A. Wickenbrock, Phys. Rev. Lett. 114, 140401 (2015). L. Bougas, A. Wilzewski, and D. Budker, Infrared laser [44] D. V. Else, B. Bauer, and C. Nayak, Floquet time crys- threshold magnetometry with a nv doped diamond intra- tals, Phys. Rev. Lett. 117, 090402 (2016). cavity etalon, Opt. Express 27, 1706 (2019). [45] C. W. von Keyserlingk and S. L. Sondhi, Phase structure [26] K. D. Jahnke, A. Sipahigil, J. M. Binder, M. W. Doherty, of one-dimensional interacting floquet systems. i. abelian M. Metsch, L. J. Rogers, N. B. Manson, M. D. Lukin, symmetry-protected topological phases, Phys. Rev. B 93, and F. Jelezko, Electron–phonon processes of the silicon- 245145 (2016). vacancy centre in diamond, New Journal of Physics 17, [46] C. W. von Keyserlingk and S. L. Sondhi, Phase struc- 043011 (2015). ture of one-dimensional interacting floquet systems. ii. [27] V. Khemani, A. Lazarides, R. Moessner, and S. L. symmetry-broken phases, Phys. Rev. B 93, 245146 Sondhi, Phase structure of driven quantum systems, (2016). Phys. Rev. Lett. 116, 250401 (2016). [47] B. Huang, Y.-H. Wu, and W. V. Liu, Clean floquet time [28] N. Y. Yao, A. C. Potter, I.-D. Potirniche, and A. Vish- crystals: Models and realizations in cold atoms, Phys. wanath, Discrete time crystals: Rigidity, criticality, and Rev. Lett. 120, 110603 (2018). realizations, Phys. Rev. Lett. 118, 030401 (2017). [48] W. C. Yu, J. Tangpanitanon, A. W. Glaetzle, D. Jaksch, [29] N. Y. Yao, C. R. Laumann, S. Gopalakrishnan, M. Knap, and D. G. Angelakis, Discrete time crystal in globally M. Müller, E. A. Demler, and M. D. Lukin, Many-body driven interacting quantum systems without disorder, localization in dipolar systems, Phys. Rev. Lett. 113, Phys. Rev. A 99, 033618 (2019). 243002 (2014). [49] R. E. Barfknecht, S. E. Rasmussen, A. Foerster, and [30] R. Moessner and S. L. Sondhi, Equilibration and order in N. T. Zinner, Realizing time crystals in discrete quan- quantum floquet matter, Nature Physics 13, 424 (2017). tum few-body systems, Phys. Rev. B 99, 144304 (2019). [31] D. A. Abanin, W. D. Roeck, and F. Huveneers, Theory [50] A. Pizzi, J. Knolle, and A. Nunnenkamp, Higher-order of many-body localization in periodically driven systems, and fractional discrete time crystals in clean long-range Annals of Physics 372, 1 (2016). interacting systems (2019), arXiv:1910.07539. [32] D. A. Abanin, W. De Roeck, and F. Huveneers, Expo- [51] W. W. Ho, S. Choi, M. D. Lukin, and D. A. Abanin, nentially slow heating in periodically driven many-body Critical time crystals in dipolar systems, Phys. Rev. Lett. systems, Phys. Rev. Lett. 115, 256803 (2015). 119, 010602 (2017). [33] O. Howell, P. Weinberg, D. Sels, A. Polkovnikov, and [52] A. Russomanno, F. Iemini, M. Dalmonte, and R. Fazio, M. Bukov, Asymptotic prethermalization in periodically Floquet time crystal in the lipkin-meshkov-glick model, driven classical spin chains, Phys. Rev. Lett. 122, 010602 Phys. Rev. B 95, 214307 (2017). (2019). [53] E. B. Fel’dman and S. Lacelle, Configurational averag- [34] K. Sacha and J. Zakrzewski, Time crystals: a review, ing of dipolar interactions in magnetically diluted spin Reports on Progress in Physics 81, 016401 (2017). networks, The Journal of 104, 2000 [35] D. J. Luitz, R. Moessner, S. L. Sondhi, and V. Khemani, (1996). Prethermalization without temperature, Phys. Rev. X [54] S. Lacelle and L. Tremblay, What is a typical dipolar 10, 021046 (2020). coupling constant in a solid?, The Journal of Chemical [36] F. M. Gambetta, F. Carollo, M. Marcuzzi, J. P. Gar- Physics 98, 3642 (1993). rahan, and I. Lesanovsky, Discrete time crystals in the [55] V. V. Dobrovitski, A. E. Feiguin, D. D. Awschalom, and absence of manifest symmetries or disorder in open quan- R. Hanson, Decoherence dynamics of a single spin versus tum systems, Phys. Rev. Lett. 122, 015701 (2019). spin ensemble, Phys. Rev. B 77, 245212 (2008). [37] A. Lazarides and R. Moessner, Fate of a discrete time [56] J. R. Klauder and P. W. Anderson, Spectral diffusion crystal in an open system, Phys. Rev. B 95, 195135 decay in spin resonance experiments, Phys. Rev. 125, (2017). 912 (1962). [38] B. Zhu, J. Marino, N. Y. Yao, M. D. Lukin, and E. A. [57] P. W. Anderson, Absence of diffusion in certain random Demler, Dicke time crystals in driven-dissipative quan- lattices, Phys. Rev. 109, 1492 (1958). tum many-body systems, New Journal of Physics 21, 11

[58] L. S. Levitov, Delocalization of vibrational modes caused [68] D. Suwelack and J. S. Waugh, Quasistationary magne- by electric dipole interaction, Phys. Rev. Lett. 64, 547 tization in pulsed spin-locking experiments in dipolar (1990). solids, Phys. Rev. B 22, 5110 (1980). [59] A. L. Burin, Y. Kagan, L. A. Maksimov, and I. Y. Pol- [69] K. Mallayya, M. Rigol, and W. De Roeck, Prethermal- ishchuk, Dephasing rate in dielectric glasses at ultralow ization and thermalization in isolated quantum systems, temperatures, Phys. Rev. Lett. 80, 2945 (1998). Phys. Rev. X 9, 021027 (2019). [60] In some experimental situations [40, 41, 51] the role of [70] F. Nathan, D. Abanin, E. Berg, N. H. Lindner, and M. S. strong quantizing field is played by strong continuous- Rudner, Anomalous floquet insulators, Phys. Rev. B 99, wave Rabi driving applied at the frequency ωQ. In this 195133 (2019). case, the dynamics is to be considered in a doubly ro- tating frame, where the effective quantization axis is di- rected along the Rabi driving in the primary rotating Appendix A: Statistical analysis of disordered spin frame, and the role of the principal Larmor frequency systems ωQ is played by the Rabi frequency ωR, see e.g. Refs.1 and2 for details. The doubly rotating frame in the theory Let us discuss the effect of disorder in d spatial di- of magnetic resonance is analogous to the dressed state H basis in the context. mensions on the Hahn echo signal Mx (t) (denoted with [61] Note that the conventional rotating-frame description in- the superscript “H”). We are interested in the case of ∗ cludes two equally important components: (i) the non- T2  T2 , when the typical local fields hj are much larger secular Hamiltonian terms are usually either dropped or than the typical dipolar interactions, such that the flip- taken into account perturbatively (like the Bloch-Siegert flop term SixSjx + SiySjy in dipolar Hamiltonian can shift of the resonance line), and, (ii) the observables (such be omitted [1,2, 53, 55, 56], see AppendixB. In this as Mx and My) are also viewed in the rotating coordinate case the π-pulse eliminates the inhomogeneous broad- frame, not in the laboratory frame. Correspondingly, all ening HL, and the system’s response Mx(t) is deter- expectation values should be understood as the expec- mined only by the remaining Ising-like part of the dipo- tation values in the rotating coordinate frame; in the 0 P laboratory frame, even within secular approximation, all lar Hamiltonian HI = j>i JijSizSjz, so the echo sig- H † graphs and formulas would have a different form, with nal hMx (t)i = Tr[MxU0H (t)MxU0H (t)], with U0H (t) = 0 different dependence on time. exp −itHI , can be directly calculated: [62] In this work the local fields hj are taken as independent X Y Gaussian-distributed random numbers with zero mean MH(t) ∝ cos (J t/4). (A1) and variance Γ2, so that the free decay time T ∗ ∼ 1/Γ. x ij 2 i j The exact form of the distribution is not very impor- tant: our simulations with other distributions, such as For an ensemble where each spin Si has its own set of Lorentzian and exponential, gave essentially the same re- ∗ coupling constants Jij, one should average Eq. (A1) over sults. Similarly, the exact values of Γ and T2 are unim- portant: the results remain essentially the same as long all possible positions of the spins Sj in a d-dimensional ∗ sample of volume Vd. For dilute spins we can neglect the as T2 is much shorter than all other timescales (except for the width of the driving pulses, which are modeled as underlying crystal lattice, and replace summation over instantaneous rotations). the lattice by integration over the whole space: [63] M. Suzuki, S. Miyashita, and A. Kuroda, Monte Carlo  Z Ns−1 Simulation of Quantum Spin Systems. I, Progress of The- H 1 3 oretical Physics 58, 1377 (1977). Mx (t) = cos b(θ)t/r dv , (A2) Vd V [64] V. V. Dobrovitski and H. A. De Raedt, Efficient scheme d for numerical simulations of the spin-bath decoherence, where dv is the volume element, and b(θ) = (1 − Phys. Rev. E 67, 056702 (2003). 3 cos2 θ)/4. To calculate the average for a macroscopi- [65] For d = 2 the spins are located on a plane, and the z- cally large spin ensemble, when V → ∞ and N → ∞ axis (direction of the quantizing field) can be aligned at d s with a fixed spin density f ≡ N /V , the integral is re- different angles with respect to this plane. If the z-axis is s s d written as [2, 56] normal to the plane, all spin pairs have θij = π/2, while for the in-plane alignment this angle varies between 0 and  Z Ns−1 π. However, the resulting changes in the Hahn echo are H 1  3 Mx (t) = 1 − 1 − cos b(θ)t/r dv (A3) very small: for a fixed areal spin density fs, the values Vd Vd of T2 differ by only 4%. The long-lived coherences also  Z   3 do not exhibit any qualitative differences between these ≈ exp −fs 1 − cos b(θ)t/r dv , (A4) two cases. Fig.2 presents the results for the case of z-axis Vd normal to the plane. so that the integral in Eq. (A2) is well defined at large r, [66] Due to the nature of the case d → ∞, the averaging for  3 this ensemble is performed only over random local fields where cos b(θ)t/r → 1. Singularity of the spin dynamics is determined by the hj , while the results for finite d also include averaging over the disorder in spin positions. competition of two effects: the dipolar coupling constants [67] F. M. Surace, A. Russomanno, M. Dalmonte, A. Silva, Jij decrease with increasing r, but the number of spins Sj R. Fazio, and F. Iemini, Floquet time crystals in clock which interact with the given spin Si grows with r as the models, Phys. Rev. B 99, 104303 (2019). volume element dv = rd−1drdA (where dA is the surface 12 element of the (d − 1)-dimensional hypersphere of unit 1 radius). For d < 6, this growth is sufficiently slow, so the integration over r can be extended to infinity, yielding the 0.8 d = 2 above-mentioned result (see Eq. (2) of the main text) ) reduced model t

( 0.6

H x full model H  d/3 Mx (t) = exp − |t/T2| (A5) 0.4 2

M exp( t 3 ) − h i−3/d 0.2 with the decay time T = f Λ R |b(θ)|d/3 dA . The 2 s 0 integral over the hypersphere is a numerical factor of or- 0 2 4 6 8 10 der of one, and the quantity t 1 Z ∞ 1 − cos z 1 πd  d Λ = dz = − cos Γ − , d/3+1 H 3 0 z 3 6 3 Figure 10. Hahn echo signal Mx (t) for two-dimensional dis- (A6) ordered dipolar-coupled spin systems, calculated using the which comes from integration over r, is also of order one, full and the reduced models, which differ by the presence of the flip-flop terms in the Hamiltonian. The results of the such that T2 is determined just by the spin density fs. two models are in quantitative agreement with each other. By renormalizing the spin density, we can set T2 = 1 as explained above. The analytic approximation for Hahn echo in d = 2 is M (t) = exp(−t2/3). The system size is N = 18, T = 1, For d = 2, the singularity of the Hahn echo MH(t) = x s 2 x and T ∗ ≈ 0.02. h 2/3i 2 exp − |t/T2| is strong: the initial echo decay is in- finitely fast due to strong fluctuations in the positions √ of the spins at small distances rij. Although the typi- scales as 1/ Ns for large d, in the limit Ns → ∞ the 1/d H h 2i cal distance between spins is of the order of 1/fs = Hahn echo signal is Mx (t) = exp − (t/T2) , free of 1/d (Vd/Ns) , a large fraction of spins has many neighbors singularities, with finite spectral power of the resonance at much smaller distances. Correspondingly, while the line. The effect of local fluctuations vanishes completely typical dipolar coupling is of the order of one (recall that in this case, in stark contrast with d = 2 and d = 3 spin we normalize fs to yield T2 = 1), many realizations of ensembles. the disorder produce very large dipolar couplings Jij. Since the key physical aspects of the spin dynamics Of course, in real crystals, at extremely short times the are drastically different for the systems of different di- initial decay rate is finite, because the distance between mensionalities, it is reasonable to expect that the key spins is limited by the crystal lattice constant, which features of the long-lived coherences would also differ in turn limits the maximal dipolar interaction strength. drastically for different d. Surprisingly, our results have However, the corresponding times are extremely small, demonstrated this expectation to be incorrect. ∗ orders of magnitude smaller than T2 , and are irrelevant for the phenomena considered here. For d = 3, such fluctuations in Jij are less strong, and H Appendix B: The role of the flip-flop processes the singularity is weaker: Mx (t) = exp (− |t/T2|) has a cusp at t = 0, but the initial rate of decay is finite. Still, for both d = 2 and 3, the total spectral power of the In this article, we consider a typical experimental situa- ∗ resonance line is (formally) infinite. For d = 4 and 5, the tion where T2  T2 , i.e. where the random local fields hj fluctuations are less pronounced, the total spectral power are much larger than the dipolar interactions Jij. Strong of the resonance is finite, and the singularity in MH(t) random local fields generally suppress the flip-flop terms x  + − − + is weak. SixSjx + SiySjy = (1/2) Si Sj + Si Sj in the dipolar For d ≥ 6, the integration over r cannot be extended to Hamiltonian HI given in Eq. (1). It is reasonable to ex- infinity: the integral Eq. (A6) diverges at small z (which pect that the magnetization response does not change if correspond to r → ∞). This divergence means that the the flip-flop terms are omitted, but the accumulation of number of the spins Sj coupled to the given spin Si grows the neglected terms at long times may become signifi- too fast with increasing r. The contribution from the cant. Therefore, it is important to directly evaluate the surface of the sample becomes important, dominating at role of the flip-flop terms. H H larger d. The form of Mx (t) then depends on the sample For the Hahn echo signal Mx (t) our simulations show shape and size, but in the limit d → ∞ it again acquires that the impact of the flip-flop terms is insignificant: a universal shape-independent Gaussian form. For any the curves remain practically the same, independent of regular-shaped sample in the limit d → ∞ all spins are whether we include the flip-flop terms or not, see Fig. 10, 0 P located near the surface, and each spin pair is separated where we refer to the Hamiltonian HI = j>i JijSizSjz by almost the same distance, producing all-to-all interac- without the flip-flop terms as the reduced model and to tions with a uniform coupling constant Jij → J. Eq. (A2) the full dipolar Hamiltonian HI as the full model. H Ns−1 yields Mx (t) = cos (2Jt), and, since the value of J The multi-pulse CPMG response Mx(t) for d = 2 13

(a) 1 Appendix C: The role of pulse errors 0.8 d = 2 reduced model In all simulations above, we considered the case where ) t 0.6 full model the pulse error  is constant in time, and is the same (

x for all spins. Fig. 12 demonstrates the role of pulse er- 0.4 rors in producing the long-lived coherences. In panel M (a) we show that ideal pulses, with  = 0, do not pro- 0.2 duce the long-lived coherences: the decay of the multi- 0 pulse CPMG signal Mx(t) is fast, and follows the de- 1 cay of the single-pulse Hahn echo; the latter is in per- (b) fect agreement with the theoretically predicted form 0.8 d H  2/3 → ∞ Mx (t) = exp −(t/T2) . The absence of the long-lived

) reduced model coherences for ideal pulses is a straightforward conse- t 0.6 ( quence of the fact that the dipolar coupling Hamiltonian x full model H = P (J /2) [2S S − S S − S S ] is invari- 0.4 I i>j ij iz jz ix jx iy jy M ant with respect to ideal π-pulses along the x-axis (i.e. 0.2 with respect to simultaneous change Sjz → −Sjz and Sjy → −Sjy for all spins). 0 The panel (b) shows that the pulse errors can be dif- 0 5 10 15 20 ferent for different spins, and still produce long-lived co- t herences, as long as the pulse errors remain constant in time. The orange solid line shows the multi-pulse CPMG Figure 11. Multi-pulse CPMG signals, comparison between signal Mx(t) for the case when the pulse error  is chosen the reduced and the full models for d = 2 (a), and for d → ∞ randomly for each spin (sampled uniformly from the in- (b). The results for the magnetization response Mx(t) are terval [−0.07, 0.07]), but is kept constant in time. In con- almost the same for d = 2, whereas for d → ∞ the ampli- trast, even if the pulse error is the same for all spins, but tudes of the long-lived coherences differ by a factor of ≈ 2. varies in time from one pulse to the next (e.g. sampled The results shown are for N = 20 and short τ; the other s uniformly from the interval [−0.07, 0.07] for each pulse parameters are the same as in Fig.2. anew), the ensemble coherence in CPMG experiment ex- hibits fast decay (black line), coinciding with the decay of the single-pulse Hahn echo (red dashed line). The re- also demonstrates almost identical short-time behavior sults shown in Fig. 12 have been obtained for Ns = 18; and long-lived coherences for the full and for the re- all other parameters are the same as in Fig.2. duced models, as shown in Fig. 11a. However, in the case d → ∞ shown in Fig. 11b, the calculated signals for both full and reduced models coincide only at short Appendix D: Weighted contribution of different times, and exhibit quantitative difference later. Although Floquet states to the long-lived coherences both models clearly demonstrate long-lived coherences, the echo amplitude is approximately twice higher for the In Sec.IV we presented detailed analysis of the Flo- reduced model as compared with the full model. A pos- quet states giving rise to the long-lived coherences. We sible reason for this discrepancy is in the geometry of carefully distinguished and studied separately the two the dipolar couplings. The flip-flop process between two contributions to the CPMG signal Mx(t) (see Eq.6): 2 spins is important if the dipolar coupling between the the values of the matrix elements |hψj|Mx|ψki| between two spins is comparable to, or larger than, the difference different pairs of the Floquet states |ψji and |ψki, see in the respective local fields. In the situation considered Fig.4 for d = 2 and Fig.5 for d = 5, and the number ∗ here, with T2  T2 , the above case can only occur if the P (∆φ) of contributing pairs of states, see Fig.6. local fields of two spins are accidentally similar. In d = 2 Yet another way to quantify the contribution of dif- systems, these two spins must be close to each other to ferent Floquet states to the long-lived coherences is to ensure non-negligible dipolar coupling. In contrast, in consider the combined effect of both contributions, and d → ∞ systems, these two spins can be at any distance study the weighted distribution Σ(∆φ) of the quasienergy to each other because the dipolar interaction is homoge- differences φj − φk, formally defined as neous, coupling each spin to all other spins. Thus, the X 2 h probability for two spins to have accidentally similar local Σ(∆φ) = |hψj|Mx|ψki| δ(|φj − φk| − ∆φ) fields, and to undergo a flip-flop process, is much larger j,k in d → ∞ systems than in d = 2 systems. i + δ(2π − |φj − φk| − ∆φ) , (D1) In this article we use the full model in our simulations, with the exception of the results for the single-pulse Hahn such that each pair of the Floquet states having the given echo shown in Figs.2 and3 of the main text. value of ∆φ enters the sum Σ(∆φ) with the weight coeffi- 14

2 0 −5 cient equal to |hψj|Mx|ψki| ; this corresponds to summa- bins of finite width 2β = π · 10 , i.e. the sum in tion along diagonals of Figs.4 and5. Two δ-functions in Eq. D1 includes all states whose quasienergies satisfy 0 0 the formula reflect the fact that the quantity ∆φ is down- the condition |φj − φk| ∈ [∆φ − β , ∆φ + β ] or 2π − 0 0 folded to the interval [0, π] , i.e. we take ∆φ = |φj − φk| |φj − φk| ∈ [∆φ − β , ∆φ + β ]. The distinctive peak at when |φj − φk| ≤ π and ∆φ = 2π − |φj − φk| when |φj −φk| ≈ π, which corresponds to emergence of the sub- |φj − φk| > π; this downfolding reflects the symmetries harmonic response, appears only for long τ. The peak at of the summands in Eq.6. Note that the quantity Σ(∆φ) φj −φk ≈ 0, which corresponds to the long-lived response, is proportional to the cosine Fourier transform of the is formed by the pairs of eigenstates having almost the CPMG signal Mx(2τm). same quasienergy, and the main contribution to this peak The graphs of Σ(∆φ) are presented in Fig. 13, where comes from the situation j = k, when |ψji and |ψki are instead of an infinitely narrow delta function we used the same. 15

1 (a)  = 0.07 0.8  = 0 2

) exp( t 3 ) t 0.6 − ( x 0.4 M

0.2

0 1 (b)  varies with spin 0.8  varies with time 2

) exp( t 3 ) t 0.6 − ( x 0.4 M

0.2

0 0 5 10 15 20 t

Figure 12. Multi-pulse CPMG signals for a two-dimensional (d = 2) disordered dipolar-coupled spin systems, demonstrat- ing the role of the pulse errors in appearance of the long-lived coherences. (a). Comparison between the decay of coherence Mx(t) for perfect pulses with  = 0 (black solid line) and for imperfect pulses with  = 0.07 (orange solid line). In the case of perfect pulses, the decay of the CPMG signal is fast, and follows the decay of the Hahn echo (red dashed line, showing the theoretically predicted curve exp [−t2/3]), while for the imperfect pulses a long-lived coherence is clearly seen. (b). Comparison between the decay of coherence Mx(t) for im- perfect pulses. Black solid line represents the situation when the pulse error (the value of ) is the same for all spins but randomly varies in time, being sampled uniformly from the interval [−0.07, 0.07] for each new pulse. Orange solid line cor- responds to the case when the pulse errors stay constant in time but vary in a random way from one spin to another (sam- pled uniformly from the interval [−0.07, 0.07]). In the case of the time-varying pulse errors, the decay of the CPMG signal is fast, and follows the decay of the Hahn echo (red dashed line, showing the theoretically predicted curve exp [−t2/3]). However, when the pulse errors remain constant in time, a long-lived coherence is clearly seen, even though the pulse er- rors have different values for different spins. For both panels, the results were obtained for Ns = 18 and averaged over 200 samples; all other parameters are the same as in Fig.2. 16

(a) 10000 short 8000 τ

) 6000 φ 4000 Σ(∆ 2000 0 (b) 1600 long τ 1200 ) φ 800

Σ(∆ 400 0 0 1 1 3 4 π 2 π 4 π π ∆φ

Figure 13. Weighted distribution Σ(∆φ) of the quasienergy differences |φj − φk| for a d = 2 disordered spin system with Ns = 14, at short τ (a) and long τ (b). The sharp peaks at ∆φ = 0 and at ∆φ = π correspond to the long-lived co- herences and to the subharmonic response, respectively. The distribution includes averaging over many realizations of the disorder. The bin width is 2β0 = π · 10−5, i.e. for each value of ∆φ the sum includes all Floquet states with |φj − φk| ∈ 0 0 0 0 [∆φ − β , ∆φ + β ] or 2π − |φj − φk| ∈ [∆φ − β , ∆φ + β ]. To avoid double counting, only the states with φj ≥ φk are included in Σ(∆φ). The normalization is chosen such that R Σ(∆φ)d(∆φ) = 1. All other simulation parameters are the same as in Fig.2.