14
1 The connection between statistical mechanics and the dynamics of isolated many-body quantum systems is typically formulated in terms of the eigenvalue thermalization hypothesis 1-5 (ETH), the idea that complexity in interacting systems prompts ergodicity at the quantum level. Disorder and quantum interference can stymie thermalization, often leading to regimes of sub-diffusive dynamics or suppressed transport, as already identified in a broad class of systems ranging from acoustic and optical waves to cold atomic gases 6-8 . The breakdown of ergodicity can extend even to strongly interacting quantum systems, a regime known as many-body localization 9-11 (MBL). While most MBL experiments thus far have focused on isolated arrays of trapped atoms 12,13 , recent theoretical work argued that the interplay between diffusion and localization also influences the out-of-equilibrium dynamics of driven open systems 14,15 . In particular, these studies built on the extensive body of work on dynamic nuclear polarization 16,17 (DNP) to show that the concept of spin temperature is directly connected (namely, relies on) quantum ergodicity and ETH. Indeed, ensembles of electron and nuclear spins in solids provide a practical experimental platform to investigate thermalization because disorder and long-range interactions compete in ways that can be exposed using alternative spin control techniques. For example, recent experimental work studied spin depolarization in ensembles of nitrogen-vacancy (NV) centers in diamond and revealed sub-exponential, disorder-dependent relaxation associated with critical thermalization dynamics 18,19 . Along similar lines, dynamic nuclear polarization of carbon spins in diamond was exploited to expose electron-spin-mediated nuclear spin diffusion exceeding the value expected for naturally abundant 13 C spins by nearly two orders of magnitude 20 . Here, we resort to nuclear spins in diamond to demonstrate control over the localization/delocalization dynamics of hyperfine-coupled carbons upon variation of the applied magnetic field. We formally capture our observations by considering a model electron-nuclear spin chain featuring magnetic-field-dependent spin transport. Further, we show the dynamics at play can be cast in terms of well-defined dynamic phases that can be accessed by tuning the magnetic field strength and paramagnetic content. The spin state hybridization emerging from the intimate connection between electron and nuclear spins gives rise to otherwise forbidden low-frequency transitions, whose presence underlies the system’s singular spectral response to RF excitation of variable amplitude. In our experiments, we dynamically polarize and probe 13 C spins in a [100] diamond crystal (3×3×0.3 mm 3 ) grown in a high-pressure/high-temperature chamber (HPHT). The system is engineered to host a large (~10 ppm) concentration of NV centers, spin-1 paramagnetic defects that polarize efficiently under green illumination. Coexisting with the NVs is a more abundant group of P1 centers (~50 ppm), spin-1/2 defects formed by substitutional nitrogen atoms. We tune an externally applied magnetic field " in and out of the ‘energy Magnetic-field-induced delocalization in hybrid electron-nuclear spin ensembles Daniela Pagliero 1,* , Pablo R. Zangara 6,7,* , Jacob Henshaw 1 , Ashok Ajoy 3 , Rodolfo H. Acosta 6,7 , Neil Manson 8 , Jeffrey A. Reimer 4,5 , Alexander Pines 3,5 , Carlos A. Meriles 1,2,† 1 Department. of Physics, CUNY-City College of New York, New York, NY 10031, USA. 2 CUNY-Graduate Center, New York, NY 10016, USA. 3 Department of Chemistry, University of California at Berkeley, Berkeley, California 94720, USA. 4 Department of Chemical and Biomolecular Engineering, University of California at Berkeley, Berkeley, California 94720, USA. 5 Materials Science Division Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. 6 Facultad de Matemática, Astronomía, Física y Computación, Universidad Nacional de Córdoba, Ciudad Universitaria, CP:X5000HUA Córdoba, Argentina. 7 Instituto de Física Enrique Gaviola (IFEG), CONICET, Medina Allende s/n, X5000HUA, Córdoba, Argentina. 8 Laser Physics Centre, Research School of Physics and Engineering, Australian National University, Canberra, A.C.T. 2601, Australia * Equally contributing authors. Corresponding author. E-mail: [email protected] We use field-cycling-assisted dynamic nuclear polarization to demonstrate magnetic-field-dependent activation of nuclear spin transport from otherwise isolated strongly-hyperfine-coupled sites. With the help of a toy model comprising electron and nuclear spins, we recast our observations in terms of a dynamic phase diagram featuring zones of active or forbidden nuclear spin current separated by boundaries defined by the interplay between spin Zeeman, dipolar, and hyperfine couplings. Analysis of the polarization transport as a function of the driving field reveals the presence of many-body excitations stemming from the hybrid electron-nuclear nature of the system. These findings could prove relevant in applications to spin-based quantum information processing and nanoscale sensing. Keywords: Nuclear spin diffusion | Dynamic nuclear polarization | Nitrogen-vacancy centers | Thermalization | Many-body localization

Magnetic-field-induced delocalization in hybrid electron ...3 coupling constant for nuclear spin 2=1,2. Note that, for a given set of hyperfine couplings, there are two matching fields

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • 1

    The connection between statistical mechanics and the dynamics of isolated many-body quantum systems is typically formulated in terms of the eigenvalue thermalization hypothesis1-5 (ETH), the idea that complexity in interacting systems prompts ergodicity at the quantum level. Disorder and quantum interference can stymie thermalization, often leading to regimes of sub-diffusive dynamics or suppressed transport, as already identified in a broad class of systems ranging from acoustic and optical waves to cold atomic gases6-8. The breakdown of ergodicity can extend even to strongly interacting quantum systems, a regime known as many-body localization9-11 (MBL). While most MBL experiments thus far have focused on isolated arrays of trapped atoms12,13, recent theoretical work argued that the interplay between diffusion and localization also influences the out-of-equilibrium dynamics of driven open systems14,15. In particular, these studies built on the extensive body of work on dynamic nuclear polarization16,17 (DNP) to show that the concept of spin temperature is directly connected (namely, relies on) quantum ergodicity and ETH.

    Indeed, ensembles of electron and nuclear spins in solids provide a practical experimental platform to investigate thermalization because disorder and long-range interactions compete in ways that can be exposed using alternative spin control techniques. For example, recent experimental work studied spin depolarization in ensembles of nitrogen-vacancy (NV) centers in diamond and revealed sub-exponential, disorder-dependent relaxation associated with critical

    thermalization dynamics18,19. Along similar lines, dynamic nuclear polarization of carbon spins in diamond was exploited to expose electron-spin-mediated nuclear spin diffusion exceeding the value expected for naturally abundant 13C spins by nearly two orders of magnitude20.

    Here, we resort to nuclear spins in diamond to demonstrate control over the localization/delocalization dynamics of hyperfine-coupled carbons upon variation of the applied magnetic field. We formally capture our observations by considering a model electron-nuclear spin chain featuring magnetic-field-dependent spin transport. Further, we show the dynamics at play can be cast in terms of well-defined dynamic phases that can be accessed by tuning the magnetic field strength and paramagnetic content. The spin state hybridization emerging from the intimate connection between electron and nuclear spins gives rise to otherwise forbidden low-frequency transitions, whose presence underlies the system’s singular spectral response to RF excitation of variable amplitude.

    In our experiments, we dynamically polarize and probe 13C spins in a [100] diamond crystal (3×3×0.3 mm3) grown in a high-pressure/high-temperature chamber (HPHT). The system is engineered to host a large (~10 ppm) concentration of NV centers, spin-1 paramagnetic defects that polarize efficiently under green illumination. Coexisting with the NVs is a more abundant group of P1 centers (~50 ppm), spin-1/2 defects formed by substitutional nitrogen atoms. We tune an externally applied magnetic field " in and out of the ‘energy

    Magnetic-field-induced delocalization in hybrid electron-nuclear spin ensembles

    Daniela Pagliero1,*, Pablo R. Zangara6,7,*, Jacob Henshaw1, Ashok Ajoy3, Rodolfo H. Acosta6,7, Neil Manson8,

    Jeffrey A. Reimer4,5, Alexander Pines3,5, Carlos A. Meriles1,2,†

    1Department. of Physics, CUNY-City College of New York, New York, NY 10031, USA. 2CUNY-Graduate Center, New York, NY 10016, USA.

    3Department of Chemistry, University of California at Berkeley, Berkeley, California 94720, USA. 4Department of Chemical and Biomolecular Engineering, University of California at Berkeley, Berkeley, California 94720,

    USA. 5Materials Science Division Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.

    6Facultad de Matemática, Astronomía, Física y Computación, Universidad Nacional de Córdoba, Ciudad Universitaria, CP:X5000HUA Córdoba, Argentina.

    7Instituto de Física Enrique Gaviola (IFEG), CONICET, Medina Allende s/n, X5000HUA, Córdoba, Argentina. 8Laser Physics Centre, Research School of Physics and Engineering, Australian National University, Canberra, A.C.T. 2601,

    Australia *Equally contributing authors. †Corresponding author. E-mail: [email protected]

    We use field-cycling-assisted dynamic nuclear polarization to demonstrate magnetic-field-dependent activation of nuclear spin

    transport from otherwise isolated strongly-hyperfine-coupled sites. With the help of a toy model comprising electron and nuclear spins, we recast our observations in terms of a dynamic phase diagram featuring zones of active or forbidden nuclear spin

    current separated by boundaries defined by the interplay between spin Zeeman, dipolar, and hyperfine couplings. Analysis of the polarization transport as a function of the driving field reveals the presence of many-body excitations stemming from the hybrid

    electron-nuclear nature of the system. These findings could prove relevant in applications to spin-based quantum information processing and nanoscale sensing.

    Keywords: Nuclear spin diffusion | Dynamic nuclear polarization | Nitrogen-vacancy centers | Thermalization | Many-body localization

  • 2

    matching’ condition at ~51 mT, where the Zeeman splitting of the P1 spins coincides with the frequency gap between the 0 and −1 states of the NVs. Following electron and

    nuclear spin manipulation, we monitor the bulk 13C polarization via high-field nuclear magnetic resonance (NMR) upon shuttling the sample into the bore of a 9 T magnet (additional details can be found in Ref. (21)).

    Fig. 1a shows a typical experimental protocol: We continuously illuminate the sample with a green laser (1 W at 532 nm) during a time interval &

    '(= 5 s while

    simultaneously applying radio-frequency (RF) excitation. Figs. 1b and 1c show the resulting spectrum obtained as we measure the bulk 13C NMR signal for different RF frequencies +

    ,- within the range 0.5-160 MHz. Besides the

    dip at 551 kHz — corresponding to the Larmor frequency of bulk 13C at " = 51.5 mT — we find several RF absorption bands, indicative of polarization transport from electron spins to bulk carbons via select groups of strongly hyperfine-coupled nuclei20.

    Rather than relying on NV–P1 cross-relaxation, one can dynamically polarize nuclear spins via the use of chirped micro-wave (MW) pulses, consecutively applied during &

    '(

    22 (Fig. 1d). Upon simultaneous RF excitation at variable frequencies, the spectrum that emerges shows the polarization transport process is now fundamentally distinct. This is shown in Figs. 1e through 1g, where we set the magnetic field to 47.1 mT, a shift of only ~4 mT from the energy matching condition. In particular, we find that the RF impact is mostly limited to a ~1.3 MHz band adjacent to the

    13C Larmor frequency (~0.5 MHz at 47 mT, insert in Fig. 1e). The differences are most striking near 40 MHz and 97 MHz where the dips observed at 51 mT (Figs. 1b and 1c) virtually vanish (Figs. 1e and 1g). Similarly, the small RF dip at ~11 MHz (Figs. 1e and 1f) amounts to only a small fraction of the broad absorption band centered at that frequency, previously apparent under field matching (Fig. 1b).

    To qualitatively understand these observations, we start with the toy model in Fig. 2a, a chain comprising an interacting pair of NV–P1 electron spins, each of them coupled to a neighboring carbon via hyperfine tensors of magnitude 0

    1 with 2 = 1, .2; for illustration purposes, we

    focus on the ‘hyperfine-dominated’ regime where 04~ 0

    6> ℐ

    9> :

    ;, where ℐ

    9 is the NV–P1 dipolar

    coupling constant, and :; is the nuclear Larmor frequency.

    In the absence of hyperfine couplings to the host nitrogen nuclei (a condition assumed here for simplicity) and using <

    4

    (<6) to denote the vector spin operator of the nuclear spin

    coupled to the NV (P1), the 13C spins in the chain are governed by the effective Hamiltonian23

    >?@@= A

    ?@@B4

    C− A

    ?@@B6

    C+ E

    ?@@B4

    FB6

    G+ B

    4

    GB6

    F,(1)

    where A?@@= 2J

    ?" − "

    K is the effective nuclear spin

    frequency offset relative to the matching field "K

    , and J? is

    the electron spin gyromagnetic ratio. Further, the effective coupling between nuclear spins is given by E

    ?@@=

    −:;ℐLsin

    P

    6

    QR

    ST

    QR

    SS RF Q

    R

    ST R, where tan W ≈ 0

    4

    CY04

    CC, and

    01

    CC (01

    CY) denotes the secular (pseudo-secular) hyperfine

    Figure 1 | The role of P1 centers. (a) Static matching field (SMF) protocol. (b) 13C NMR signal amplitude as a function of +

    ,-; the

    external magnetic field is "(F) = 51.5 mT. (c) Zoomed SMF response around ~97 MHz. (d) Dynamic nuclear polarization via micro-wave sweeps (MWS). (e) Same as in (b) but using the MWS protocol to induce nuclear polarization; the external magnetic field is " = 47.1 mT. (f–g) Zoomed 13C response using the MWS protocol. Unlike (b), we see no high-frequency dips. In all experiments, &'(= 5 s, the total number of repeats per point is 8, the driving field amplitude is Ω

    ,-= 4 kHz, and the laser power is 1 W; solid

    traces are guides to the eye. In (d) through (g), the MW power is 300 mW, the sweep range is 25.2 MHz, the sweep rate is 15 MHz ms-1 corresponding to a total of 8333 sweeps during &

    '(.

  • 3

    coupling constant for nuclear spin 2 = 1,2. Note that, for a given set of hyperfine couplings, there are two matching fields yielding analogous dynamics in the four-spin chain23; additional fields are likely in more complex systems.

    Eq. (1) is a nuclear-spin-only Hamiltonian where paramagnetic interactions manifest in the form of field-dependent shifts and effective couplings largely exceeding the intrinsic 13C-13C dipolar couplings (of order 100 Hz for naturally abundant carbon). For example, for the present 50 ppm nitrogen concentration, we have ℐ

    9~3 MHz and thus

    E?@@~30 kHz for 0

    1

    CC~0

    1

    CY~10 MHz, 2 = 1,2. A numerical

    example demonstrating good agreement between the exact and effective nuclear spin evolution is presented in Fig. 2b for three different magnetic fields. It is worth highlighting the amplified sensitivity to field detuning " − "

    K, impacting

    the offset terms in Eq. (1) via the electronic (not the nuclear) spin gyromagnetic ratio. Naturally, comparable ideas apply to the case of larger electron and nuclear spin groups, even if deriving accurate effective Hamiltonians becomes increasingly difficult23,24.

    To more generally capture the nuclear spin dynamics prompted by NV–P1 couplings, we resort to the nuclear spin current operator ^ = 1 2_ B

    4GB6F− B

    4FB6G

    , here serving

    as a measure of delocalization25. Fig. 2c shows an explicit calculation for different combinations of hyperfine couplings as a function of " and ℐ

    9. We find non-zero transport within

    a small region of the parameter space, hence revealing the equivalent of a dynamic phase transition controlled via the applied magnetic field. Two discrete regions emerge in the limit of strong hyperfine couplings (upper plot in the stack), corresponding to conditions where carbons polarize positively or negatively20.

    Similar to the case of carriers subjected to a Hubbard Hamiltonian9, the dynamics in the present spin system can be cast in terms of an interplay between ‘disorder’ — here expressed as site-selective nuclear Zeeman frequencies — and the amplitude of 13C–13C ‘flip-flop’ couplings E

    ?@@ — also

    referred to as the ‘hopping’ term in charge transport studies. This is summarized in Fig. 2d where we compute a weighted average that takes into account the known set of carbon hyperfine couplings with the NV and P1 centers23,26-29, and find non-zero current in the region where E

    ?@@≳ A

    ?@@. We warn

    this latter condition must be understood in a distributional sense, i.e., for a given concentration of paramagnetic centers represented by ℐ

    9, there is a magnetic field range where spin

    diffusion channels become available to the most likely spin arrays in the crystal.

    Interestingly, field-insensitive spin transport between carbons can be mediated via electron spins of the same type through the effective coupling20,23 E

    ?@@

    a≈

    bcd

    RQe

    STQR

    STℐf

    g

    ∆e

    R∆R

    R, where

    ℐ9

    a denotes the homo-electron-spin coupling constant and ∆1

    6= 0

    1

    CC6

    + 01

    CY6

    for 2 = 1,2. Since the number of NVs is typically tied to the P1 concentration, nuclear spin transport away from "

    K can take place in samples with sufficient

    paramagnetic content via NV–NV coupling. For example, assuming an NV concentration of 10 ppm, we have ℐ

    9

    a~0.6

    MHz thus yielding E?@@

    a~6 KHz for the case 0

    1

    CC~0

    1

    CY~10

    MHz, 2 = 1,2. Building on the above ideas, we can now interpret the markedly different frequency response in Figs. 1b and 1e as the manifestation of two complementary spin transport channels, one relying on (field-dependent) matching between NV and P1 resonances, the other emerging from field insensitive interactions between NVs. In particular, we hypothesize this latter mechanism underlies the appearance of the (weak) dip at ~11 MHz in the RF absorption spectrum at 47.1 mT (Figs. 1e and 1f). Similar considerations apply to the regime of moderate and weak hyperfine couplings ℐ

    9> 0

    1, :

    ; (corresponding to RF

    frequencies ≲1–2 MHz), required to describe the flow of polarization from and to more weakly-coupled carbons20.

    Additional information on the dynamics at play can be obtained through the experiments in Fig. 3, where we measure the DNP response under the protocols of Figs. 1a and 1d using RF excitation of variable power. Besides the anticipated gradual growth of the absorption dips, we observe an overall spectral broadening, greatly exceeding that expected from increased RF power alone. This behavior is clearest in the range 5–15 MHz and near 40 MHz (Fig. 3a),

    Figure 2 | Magnetic-field-dependent spin transport. (a) Model spin chain (top) and schematic NV–P1 energy diagram when the magnetic field and NV symmetry axis are aligned, i.e., k=0. (b) Inter-carbon polarization transfer for the chain in (a). The solid (faint) traces in each plot show the calculated evolution under the effective (exact) Hamiltonian. (c) Nuclear spin current ^ for the chain in (a) as a function of " and ℐ

    9 for

    different hyperfine couplings. (d) Same as in (c) but after a weighted average over various configurations of hyperfine couplings (see Ref. 23). In (b) we assume ℐ

    9= 0.7 MHz,

    04

    YC= 13 MHz, 0

    6

    YC= 4 MHz, and 0

    1

    CC

    = 01

    CY for 2 = 1,2.

  • 4

    where all absorption dips grow to encompass several MHz even when the RF Rabi field Ω

    ,- never exceeds 10 kHz.

    To interpret these observations, we resort one more time to the electron–nuclear spin chain of Fig. 2a and model the system dynamics in the presence of a driving RF field under the condition " = "

    K. Figures 3b and 3c show a schematic

    of the energy diagram along with the calculated spectral overlap l between carbons in the chain as a function of +

    ,-

    for variable Rabi amplitude Ω,-

    . Using m1: , 2 = 1,2 to

    denote the Fourier transform of the polarization in each carbon, the spectral overlap l ∝ o:m

    4: m

    6

    ∗: serves

    as a convenient measure of synchronic nuclear spin dynamics, and hence quantifies the disruption caused by the application of RF through the appearance of ‘dips’ at select frequencies (Fig. 3c). A detailed inspection shows that some of these resonances can be associated to ‘zero-quantum’ (i.e., intra-band) transition frequencies in the electron bath. Normally forbidden, these transitions are activated here due to the hybrid, nuclear–electron spin nature of the chain (e.g., transitions (3) and (4) in Fig. 3b, see also Ref. 23). The separation between consecutive dips is determined by the inter-electron and hyperfine couplings, thus leading to complex spectral responses spanning several MHz.

    Fig. 3d shows the calculated spectral overlap change Al ≡ l Ω

    ,-, ν,-

    − 1 as a function of Ω,-

    at select excitation frequencies ν

    ,-: Interestingly, we find that all dips

    — both nuclear and hybrid — grow at comparable rates, a counter-intuitive response given the presumably hindered nature of the zero-quantum transitions23. On the other hand, the transport of nuclear spin polarization — faster for chains featuring greater ℐ

    9 — is more difficult to disrupt if Ω

    ,-≲

    :?@@

    , thus leading to slower growth rates for more strongly coupled chains (Fig. 3e). Correspondingly, the response expected for spins in a crystal (vastly more complex than our toy model) is one where RF excitation of increasing amplitude gradually induces new dips through the perturbation of faster polarization transport channels. The result is an increasingly broader-looking absorption spectrum, in qualitative agreement with our observations.

    Despite the present limitations, our model suggests we should view these hybrid spin systems as a single whole, where nominally forbidden ‘hybrid’ excitations applied locally propagate spectrally to impact groups of spins not directly addressed. Therefore, besides the fundamental aspects, an intriguing practical question is whether, even in the absence of optical pumping, Overhauser-like DNP at higher fields could be simply attained via low-frequency (i.e., RF) manipulation of the electron spins. More generally, these results could prove useful in quantum applications relying on spin platforms, for example, to transport information between remote nuclear qubits or to develop enhanced nanoscale sensing protocols.

    Figure 3 | Dependence with RF power. (a) 13C NMR signal amplitude as a function of the excitation frequency +,-

    using the DNP protocols in Figs. 1a and 1d (respectively, left and right panels) for various RF amplitudes (bottom right in each panel). Horizontal dashed lines indicate the 13C NMR amplitude in the absence of RF excitation and solid traces are guides to the eye. (b) Schematic energy diagram for the electron-nuclear spin chain in the cartoon. States are denoted using projection numbers for the electronic spin and up/down arrows for nuclear spins with primes indicating a dominating hyperfine field. Numbers illustrate some nuclear and electron-nuclear spin transitions; energy separations are not to scale. (c) Spectral overlap l as a function of +

    ,- for different Rabi amplitudes

    Ω,-

    in the case of a spin chain with couplings ℐ9= 30 kHz, 0

    4

    YC= 9 MHz, 0

    6

    YC= 2.5 MHz, and 0

    1

    CC

    = 01

    CY for 2 = 1,2. (d) Spectral overlap change Al at select frequencies (bottom right) as a function of Ω

    ,- for the spin chain in (c). (e) Same as in (d) but for a spin

    chain with couplings ℐ9= 800 kHz, 0

    4

    YC= 9 MHz, 0

    6

    YC= 2.5 MHz, and 0

    1

    CC

    = 01

    CY for 2 = 1,2.

  • 5

    D.P., J.H., and C.A.M. acknowledge support from the National Science Foundation through grant NSF-1903839, and from Research Corporation for Science Advancement through a FRED Award; they also acknowledge access to the

    1 Mark Srednicki, "Chaos and quantum thermalization", Phys. Rev. E. 50, 888 (1994). 2 M. Rigol, V. Dunjko, M. Olshanii, "Thermalization and its mechanism for generic isolated quantum systems", Nature 452, 854 (2009). 3 G. Biroli, C. Kollath, A.M. Läuchli, “Effect of Rare Fluctuations on the Thermalization of Isolated Quantum Systems”, Phys. Rev. Lett. 105, 250401 (2010). 4 R. Nandkishore, D.A. Huse, “Many body localization and thermalization in quantum statistical mechanics”, Ann. Rev. Cond. Matter Phys. 6, 15 (2015). 5 J. Eisert, M. Friesdorf, C. Gogolin, “Quantum many-body systems out of equilibrium”, Nat. Phys. 11, 124 (2015) 6 J. Billy, V. Josse, Z. Zuo, A. Bernard, B. Hambrecht, P. Lugan, D. Clement, L. Sanchez-Palencia, P. Bouyer, A. Aspect, “Direct observation of Anderson localization of matter waves in a controlled disorder”, Nature 453, 891 (2008). 7 T. Schwartz, G. Bartal, S. Fishman, M. Segev, “Transport and Anderson localization in disordered two-dimensional photonic lattices”, Nature 446, 52 (2007). 8 G. Roati, C. D’Errico, L. Fallani, M. Fattori, C. Fort, M. Zaccanti, G. Modugno, M. Modugno, M. Inguscio, “Anderson localization of a non-interacting Bose–Einstein condensate”, Nature 453, 895 (2008). 9 V. Oganesyan, D.A. Huse, “Localization of interacting fermions at high temperature”, Phys. Rev. B 75, 155111 (2007). 10 I.L. Aleiner, B.L. Altshuler, G.V. Shlyapnikov, “A finite-temperature phase transition for disordered weakly interacting bosons in one dimension”, Nat. Phys. 6, 900 (2010). 11 I.V. Gornyi, A.D. Mirlin, D.G. Polyakov, “Interacting electrons in disordered wires: anderson localization and low-T transport”, Phys. Rev. Lett. 95, 206603 (2005). 12 J. Smith, A. Lee, P. Richerme, B. Neyenhuis, P.W. Hess, P. Hauke, M. Heyl, D.A. Huse, C. Monroe, “Many-body localization in a quantum simulator with programmable random disorder”, Nat. Phys. 12, 907 (2016). 13 M. Schreiber, S.S. Hodgman, P. Bordia, H.P. Lüschen, M.H. Fischer, R. Vosk, E. Altman, U. Schneider, I. Bloch, “Observation of many-body localization of interacting fermions in a quasirandom optical lattice”, Science 349, 842 (2015). 14 A. De Luca, A. Rosso, “Dynamic nuclear polarization and the paradox of quantum thermalization”, Phys. Rev. Lett. 115, 080401 (2015). 15 A De Luca, I. Rodríguez-Arias, M. Müller, A. Rosso, “Thermalization and many-body localization in systems under dynamic nuclear polarization”, Phys. Rev. B 94, 014203 (2016). 16 B. Corzilius, “Theory of solid effect and cross effect dynamic nuclear polarization with half-integer high-spin metal polarizing

    facilities and research infrastructure of the NSF CREST IDEALS, grant number NSF-HRD-1547830. J.H. acknowledges support from CREST-PRF NSF-HRD 1827037.

    agents in rotating solids”, Phys. Chem. Chem Phys. 18, 27190 (2016). 17 T.V. Can, Q.Z. Ni, R.G. Griffin, “Mechanisms of dynamic nuclear polarization in insulating solids”, J Magn Reson. 253, 23 (2015). 18 J. Choi, S. Choi, G. Kucsko, P.C. Maurer, B.J. Shields, H. Sumiya, S. Onoda, J. Isoya, E. Demler, F. Jelezko, N.Y. Yao, M.D. Lukin, “Depolarization dynamics in a strongly interacting solid-state spin ensemble”, Phys. Rev. Lett. 118, 093601 (2017). 19 G. Kucsko, S. Choi, J. Choi, P.C. Maurer, H. Zhou, R. Landig, H. Sumiya, S. Onoda, J. Isoya, F. Jelezko, E. Demler, N.Y. Yao, M.D. Lukin, “Critical thermalization of a disordered dipolar spin system in diamond”, Phys. Rev. Lett. 121, 023601 (2018). 20 D. Pagliero, P. Zangara, J. Henshaw, A. Ajoy, R.H. Acosta, J.A. Reimer, A. Pines, C.A. Meriles, “Optically pumped spin polarization as a probe of many-body thermalization”, Science Adv. 6, eaaz6986 (2020). 21 D. Pagliero, K.R. Koteswara Rao, P.R. Zangara, S. Dhomkar, H.H. Wong, A. Abril, N. Aslam, A. Parker, J. King, C.E. Avalos, A. Ajoy, J. Wrachtrup, A. Pines, C.A. Meriles, “Multispin-assisted optical pumping of bulk 13C nuclear spin polarization in diamond”, Phys. Rev. B 97, 024422 (2018). 22 A. Ajoy, K. Liu, R. Nazaryan, X. Lv, P.R. Zangara, B. Safvati, G. Wang, D. Arnold, G. Li, A. Lin, P. Raghavan, E. Druga, S. Dhomkar, D. Pagliero, J.A. Reimer, D. Suter, C.A. Meriles, A. Pines, “Orientation-independent room-temperature optical 13C hyperpolarization in powdered diamond”, Science Adv. 4, eaar5492 (2018). 23 Supplementary Material. 24 P.R. Levstein, H.M. Pastawski, J.L. D'Amato, “Tuning the through-bond interaction in a two-centre problem”, J. Phys.: Condens. Matter 2, 1781 (1990). 25 A. De Luca, M. Collura, J. De Nardis, “Non-equilibrium spin transport in integrable spin chains: Persistent currents and emergence of magnetic domains”, Phys. Rev. B 96, 020403 (2017). 26 C.V. Peaker, M.K. Atumi, J.P. Goss, P.R. Briddon, A.B. Horsfall, M.J. Rayson, R. Jones, “Assignment of 13C hyperfine interactions in the P1-center in diamond”, Diam. Rel. Mater. 70, 118 (2016). 27 K.R.K. Rao, D. Suter, “Characterization of hyperfine interaction between an NV electron spin and a first-shell 13C nuclear spin in diamond”, Phys. Rev. B 94, 060101(R) (2016). 28 B. Smeltzer, L. Childress, A. Gali, “13C hyperfine interactions in the nitrogen-vacancy centre in diamond”, New J. Phys. 13, 025021 (2011). 29 A. Dreau, J.-R. Maze, M. Lesik, J.-F. Roch, V. Jacques, “High-resolution spectroscopy of single NV defects coupled with nearby 13C nuclear spins in diamond”, Phys. Rev. B 85, 134107 (2012).

  • 1

    I. The four-spin model

    A simple spin system to study the main features of the magnetic-field controlled spin transport is presented in Fig. S1. Two 13C nuclear spins are hyperfine-coupled to two paramagnetic impurities, one of them a NV center and the other a P1 center. These two electronic spins, in turn, interact by means of a dipolar coupling. Given the typically large mismatch between the resonance frequencies of hyperfine-coupled and bulk carbons — and thus the corresponding quenching of inter-nuclear flip-flop processes — the 4-spin model above provides a rationale for understanding the dynamics of nuclear polarization in a real crystal. Following the arguments discussed in Ref. [1], these mechanisms lead to an effective, purely nuclear, description of spin diffusion among a large number of 13Cs nuclear spins.

    The Hamiltonian for the four-spin system is given by

    !" = −%&'() − %&'*) + %,-) + %,-.) + / -) * + 0())-)'() + 0()1-)'(1 + 0*))-.)'*) + 0*)1-.)'*1

    +ℐ42 -

    6-.7 + -7-.6 . A. 1

    Here, ;< = = 1,2 stands for the nuclear 13C spin operator, ? is the NV electronic spin operator (- =1), ?. is the P1 electronic spin operator (-. = 1/2), %& = A&B, %, = A, B, and / corresponds to the NV zero-field splitting. Coefficients 0((*))) and 0((*))1 respectively denote the hyperfine tensor components coupling the left (right) 13C and the NV (P1) and ℐ4 stands for the dipolar coupling strength between the NV and the P1 centers. In Eq. (A.1) both hyperfine interactions have been already secularized.

    We restrict our analysis to the vicinity of B = 51mT, where the spin states 0 ↑ for the NV-P1 pair is almost degenerate with −1 ↓ . We focus the analysis of the nuclear spin dynamics

    in the subspaces spanned by these two electronic states, since all other electronic configuration remain energetically inaccessible. Table S1 shows the matrix representation of the Hamiltonian in Eq. (A.1) in such a subspace.

    With the purpose of developing a model of effective 13C-13C interactions, we perform a partial diagonalization of each 13C spin in a local basis given by the hyperfine coupling. More precisely, the quantization axis of each nuclear spin is now determined by the vectors

    Supporting Information

    Magnetic-field-induced delocalization in hybrid electron-nuclear spin ensembles

    Daniela Pagliero1,*, Pablo R. Zangara6,7,*, Jacob Henshaw1, Ashok Ajoy3, Rodolfo H. Acosta6,7, Neil

    Manson8, Jeffrey A. Reimer4,5, Alexander Pines3,5, Carlos A. Meriles1,2

    1 Department. of Physics, CUNY-City College of New York, New York, NY 10031, USA. 2 CUNY-Graduate Center, New York, NY 10016, USA. 3 Department of Chemistry, University of California at Berkeley, Berkeley, California 94720, USA. 4

    Department of Chemical and Biomolecular Engineering, University of California at Berkeley, Berkeley, California 94720, USA. 5 Materials Science Division Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA. 6 Facultad de Matemática, Astronomía, Física y Computación, Universidad Nacional de Córdoba, Ciudad Universitaria, CP:X5000HUA Córdoba, Argentina. 7 Instituto de Física Enrique Gaviola (IFEG), CONICET, Medina Allende s/n, X5000HUA, Córdoba,

    Argentina. 8 Laser Physics Centre, Research School of Physics and Engineering, Australian National University, Canberra, A.C.T. 2601, Australia.

    *Equally contributing authors

  • 2

    K( = 0()1L + 0()) + %& M A. 2

    K*± = 0*)1L + 0*)) ± 2%& M(A. 3)

    for 13Cs 1 and 2, respectively. Notice that K( defines the quantization axis only in the subspace corresponding to the electronic states −1 ↓ , since in the subspace with electronic configuration 0 ↑ the quantization axis remains defined by the external magnetic field (no hyperfine interaction). Additionally, the two possible axes K*± for the second 13C are defined for the 0 ↑ subspace (Eq. (A.3), negative sign) and the −1 ↓ subspace

    (Eq. (A.3), positive sign). The corresponding rotation angles required to transform into such a representation are respectively defined by

    tan S( = 0(TU 0()) + %& (A. 4)

    tan S*± = 0*TU (0*)) ± 2%&).(A. 5)

    The rotations of each 13C quantization axis transform the original Hamiltonian !" into a new one !", whose matrix representation is shown in Table S2. In this new basis, primed labels for the nuclear states indicate the change in the quantization axis when required.

    II. The effective 13C-13C Hamiltonian

    It is clear from the matrix representation of Hamiltonian !" in Table S2 that nuclear spin flip-flops are possible provided that the appropriate energy-matching condition is achieved. For instance, states ↑ 0 ↑↓. and ↓ −1 ↓↑. are degenerate if

    −%&2 +

    %,2 −

    ∆*7

    4 = / −3%,2 −

    ∆*6

    4 +∆(2 ,(A. 6)

    which defines an equation for the ‘matching’ magnetic field B = BY((). When this condition is met, the

    dynamics of the pair of nuclear spins is essentially a flip-flop process ↑↓. ↔ ↓↑. . The time-scale for such a process is given by the matrix element

    ↓ −1 ↓↑. !" ↑ 0 ↑↓. = −ℐ42 [*[((A. 7)

    An equivalent situation can be found when the condition

    %&2 +

    %,2 +

    ∆*7

    4 = / −3%,2 +

    ∆*6

    4 −∆(2 (A. 8)

    is satisfied. In this case, states ↓ 0 ↑↑. and ↑ −1 ↓↓. are degenerate, and the coupling matrix element is the same as in Eq. (A.7). This energy-matching condition corresponds to a different magnetic field B = BY

    (*) as Eq. (A.6) is not equivalent to (A.8). At any of these ‘matching’ fields, an effective description can be proposed for the nuclear spins,

    Figure S1. The four-spin system. Two 13Cs interact with two dipolarly coupled electrons, one NV and one P1 center.

  • 3

    | ↑0↑↑⟩

    | ↑0↑↓⟩

    | ↓0↑↑⟩

    | ↓0↑↓⟩

    | ↑−1↓↑⟩

    | ↑−1↓↓⟩

    | ↓−1↓↑⟩

    | ↓−1↓↓⟩

    ⟨ ↑0↑↑|

    −)*+) , 2

    +. /0

    0 4

    . /02 4

    0 0

    ℐ 4/2

    0

    0 0

    ⟨ ↑0↑↓|

    . /02 4

    ) , 2−. /0

    0 4

    0 0

    0 ℐ 4/2

    0

    0

    ⟨ ↓0↑↑|

    0 0

    ) , 2+. /0

    0 4

    . /02 4

    0 0

    ℐ 4/2

    0

    ⟨ ↓0↑↓|

    0 0

    . /02 4

    )*+) , 2

    −. /0

    0 4

    0 0

    0 ℐ 4/2

    ⟨ ↑−1↓↑|

    ℐ 4/2

    0

    0 0

    −)*+6−3)

    , 2−. /0

    0 4−. 80

    0 2

    −. /0

    2 4

    −. 80

    2 2

    0

    ⟨ ↑−1↓↓|

    0 ℐ 4/2

    0

    0 −. /0

    2 4

    6−3)

    , 2+. /0

    0 4−. 80

    0 2

    0 −. 80

    2 2

    ⟨ ↓−1↓↑|

    0 0

    ℐ 4/2

    0

    −. 80

    2 2

    0 6−3)

    , 2−. /0

    0 4+. 80

    0 2

    −. /0

    2 4

    ⟨ ↓−1↓↓|

    0 0

    0 ℐ 4/2

    0

    −. 80

    2 2

    −. /0

    2 4

    )*+6−3)

    , 2+. /0

    0 4+. 80

    0 2

    Tabl

    e S1

    . Mat

    rix

    repr

    esen

    tatio

    n of

    Ham

    ilton

    ian 9:

    . The

    two

    subs

    pace

    s co

    rres

    pond

    ing

    to e

    ach

    NV

    -P1

    elec

    troni

    c co

    nfig

    urat

    ion

    have

    bee

    n sh

    adow

    ed

    with

    gre

    en (|0↑⟩

    ) and

    gre

    y (| −

    1↓⟩

    ).

  • 4

    | ↑0↑↑; ⟩

    | ↑0↑↓; ⟩

    | ↓0↑↑; ⟩

    | ↓0↑↓; ⟩

    | ↑−1↓↑; ⟩

    | ↑−1↓↓; ⟩

    | ↓−1↓↑; ⟩

    | ↓−1↓↓; ⟩

    ⟨ ↑0↑↑; |

    −)* 2+) , 2

    +∆ /= 4

    0

    0 0

    ℐ 4 2> /> 8

    −ℐ 4 2? /> 8

    −ℐ 4 2> /? 8

    ℐ 4 2? /? 8

    ⟨ ↑0↑↓; |

    0 −)* 2+) , 2

    −∆ /= 4

    0

    0 ℐ 4 2? /> 8

    ℐ 4 2> /> 8

    −ℐ 4 2? /? 8

    −ℐ 4 2> /? 8

    ⟨ ↓0↑↑; |

    0 0

    )* 2+) , 2

    +∆ /= 4

    0

    ℐ 4 2> /? 8

    −ℐ 4 2? /? 8

    ℐ 4 2> /> 8

    −ℐ 4 2? /> 8

    ⟨ ↓0↑↓; |

    0 0

    0 )* 2+) , 2

    −∆ /= 4

    ℐ 4 2? /? 8

    ℐ 4 2> /? 8

    ℐ 4 2? /> 8

    ℐ 4 2> /> 8

    ⟨ ↑−1↓↑; |

    ℐ 4 2> /> 8

    ℐ 4 2? /> 8

    ℐ 4 2> /? 8

    ℐ 4 2? /? 8

    6−3)

    , 2−∆ /@ 4

    −∆ 8 2

    0

    0 0

    ⟨ ↑−1↓↓; |

    −ℐ 4 2? /> 8

    ℐ 4 2> /> 8

    −ℐ 4 2? /? 8

    ℐ 4 2> /? 8

    0

    6−3)

    , 2+∆ /@ 4

    −∆ 8 2

    0

    0

    ⟨ ↓−1↓↑; |

    −ℐ 4 2> /? 8

    −ℐ 4 2? /? 8

    ℐ 4 2> /> 8

    ℐ 4 2? /> 8

    0

    0 6−3)

    , 2−∆ /@ 4

    +∆ 8 2

    0

    ⟨ ↓−1↓↓; |

    ℐ 4 2? /? 8

    −ℐ 4 2> /? 8

    −ℐ 4 2? /> 8

    ℐ 4 2> /> 8

    0

    0 0

    6−3)

    , 2+∆ /@ 4

    +∆ 8 2

    Tabl

    e S2

    . Mat

    rix

    repr

    esen

    tatio

    n fo

    r th

    e H

    amilt

    onia

    n 9A:

    . Ela

    bora

    ting

    on E

    qns.

    (A2-

    A5)

    , the

    not

    atio

    n is

    giv

    en b

    y B C⃗8B=∆ 8

    , BC⃗ /±B =

    ∆ /±, >8=cos(K 8/2),

    ? 8=sin(K 8/2), > /=cos[(K/@−K /=)/2]

    , and

    ?/=sin[(K/@−K /=)/2]

    . Off

    -dia

    gona

    l mat

    rix e

    lem

    ents

    that

    indu

    ce n

    ucle

    ar fl

    ip-f

    lops

    are

    hig

    hlig

    hted

    in y

    ello

    w.

  • 5

    !"## = %"##&'( − %"##&*

    ( + ,"## &'-&*

    . + &'.&*

    - , A. 9

    where

    %"## = 2 5 − 56(',*)

    9:(A. 10)

    ,"## =ℐ>2sin(B'/2) sin[(B*

    - − B*.)/2].(A. 11)

    Figure S2(a-b) exemplifies the comparison of the 13C polarization dynamics for both the complete !F and the effective !"## Hamiltonians.

    From Eq. (A.7) we conclude that the effective 13C-13C coupling ,"## embodies a four-body transition matrix element. This is equivalent to the “hyperfine-dominated” effective coupling described in Ref. [1], where two P1 centers were used to mediate the 13C spins. As a matter of fact, we can use here the same assumption of a ‘hyperfine dominated’ regime to derive a more explicit form for ,"##. Thus, we consider cases where the hyperfine energy is much larger than the Zeeman energy at each 13C, so we can analyze the limit B*- − B*. →0,

    sinB*- − B*

    .

    2≈1

    2tan B*

    - − B*. =

    1

    2

    tan B*- − tan B*

    .

    1 + tan B*- tan B*

    . =1

    2

    K*(L

    (K*(( + 2MN)

    −K*(L

    (K*(( − 2MN)

    1 +K*(L *

    K*(( * − 2MN *

    .(A. 12)

    Then,

    sinB*- − B*

    .

    2≈

    −2MNK*(L

    K*(( * − 2MN * + K*

    (L *.(A. 13)

    If we again use the fact that the hyperfine energy dominates over the Zeeman energy, we can approximate ∆*-~∆*.~ K*(( * + K*(L * ≡ ∆*. Then, the effective flip-flop matrix element can be written as

    ,"## ≈ −ℐ> sinB'2

    MNK*(L

    K*(( * + K*

    (L *,(A. 14)

    which is analogous to the matrix element for the case of a pair of P1 centers in the ‘mediating’ role,

    ,"## ≈ −ℐ>MNK'

    (L

    K'(( * + K'

    (L *

    MNK*(L

    K*(( * + K*

    (L *,(A. 15)

    as derived in Ref. [1]. We warn, however, that the P1-P1 mechanism of electron-mediated nuclear interaction is not dependent on the magnetic field (more precisely, it does not require the field-matching condition), so it is insufficient to provide for a rationale for our experimental observations.

    III. The spin current and the delocalization diagram

    A complementary analysis of the polarization dynamics can be done in terms of the polarization current [2]

    U = 1 2V &'.&*

    - − &'-&*

    . ,(A. 16)

    which provides an observable to quantify the flow of polarization between the two 13C spins. For example, we show in Fig. S2 the time-dependence of the polarization at each 13C along with U for a given set of couplings and satisfying the matching condition 5 = 56

    (').

  • 6

    Being our model strictly finite, both the spin polarization and the current keep oscillating as the two nuclear spins undergo the flip-flop dynamics. A more realistic description would encompass a large number of nuclear spins effectively interacting by means of an appropriate extension of Hamiltonian !"## in Eq. (A.9). In such a case, the polarization would jump from the second 13C to a nearby 13C (also coupled by mediating NV-P1 or P1-P1 pairs), ultimately diffusing away. This physical picture of spin-diffusion from strongly hyperfine-shifted 13Cs to gradually more weakly hyperfine-shifted 13Cs (and then finally to bulk 13Cs) has been extensively discussed Ref. [1]. In Fig. S3(a) we evaluate U as a function of time and the magnetic field 5 for a given choice of hyperfine couplings and ℐ>. In order to capture the dynamical trends at both ‘matching’ fields 56

    (') and 56(*),

    we consider the initial state

    XY =1

    2↑ 0 ↑↓\ ↑ 0 ↑↓\ +

    1

    2↓ 0 ↑↑\ ↓ 0 ↑↑\ .(A. 17)

    Figure S3(b) shows a cross-section of the amplitude of the oscillations in U as a function of 5. Notice the two dominant peaks corresponding to 56

    (') and 56(*), whose widths are associated to higher order processes.

    Figure S3(c) shows the same quantity, the amplitude of U as a function of 5, but here calculated for different ℐ> (Figure 2(c) in the main text shows the same simulation for different sets of hyperfine couplings).

    Figure S2. Polarization and spin current dynamics. (a) Time dependence of polarization '̂ and *̂ at 13Cs C1

    and C2 respectively, using the complete four-spin model evolved with Hamiltonian !_F and initial state |↑ 0 ↑↓\⟩. (b) Same as in (a), but using the effective nuclear interaction given by Hamiltonian !"##, and initial state |↑↓⟩. (c) Time dependence of the spin current for the dynamics induced by !_F as in case (a). (d) Time dependence of the spin current for the effective interaction !"##. In all the cases, we consider K'(( = K'(L = 9MHz, K*(( = K*(L =2.5MHz, ℐ> = 800kHz, 5 = 56

    (')= 51.3085mT.

  • 7

    The amplitude of U is used here as a quantitative indicator for localization/delocalization of nuclear spin polarization. As the amplitude of U increases, the polarization is more efficiently transferred and it can therefore diffuse. This leads us to build a dynamical phase-diagram as shown in Fig. 2(d) of the main text, where we average over many different possible choices of hyperfine couplings (configurations in our 4-spin system). Since the possible NV-13C and P1-13C hyperfine couplings are well known [3–5], we performed a sampling of the most relevant combinations of couplings from 1 to ~14 MHz. In particular, Fig. S4(a) shows the actual bivariate distribution of possible hyperfine configurations and Fig. S4(b) shows the cases we evaluated and accounted for to mimic such a distribution.

    IV. The 4-spin system in the presence of RF driving

    In order to analyze the spectral broadening observed in the DNP signal under RF excitation, we study the dynamics of polarization transfer in our 4-spin model adding the time-dependent perturbation

    i j = Ωlm cos νlmj &'L + &*

    L ,(A. 18)

    Figure S4. Distributions of hyperfine couplings. (a) Bivariate distribution of NV-13C and P1-13C hyperfine couplings as reported in Refs. [3-4] and [5], respectively. (b) Bivariate ‘toy’ distribution employed to mimic (a) and compute the averaged case shown in Fig. 2(d) of the main text.

    Figure S3. Dynamics of spin current q. (a) Time evolution of U as a function of the magnetic field. (b) Amplitude of the oscillations in U (i.e., maximum along the time-axis in (a)) as a function of the magnetic field. (c) Same as in (b), but for a variable dipolar coupling strength ℐ>. In all the cases, we consider K'(( = K'(L =2.9MHz, K*(( = K*(L = 3.2MHz; the initial state is given in Eq. (A.17). In (a-b), ℐ> = 800kHz.

  • 8

    where Ωlm stands for RF amplitude and νlm denotes its frequency. The system’s dynamical response is computed without any explicit approximation using the QuTiP [6] time-dependent solver. Fig. S5(a-b) shows the time dependence of the polarization of each carbon r̂ j (s = 1,2), as a function of the irradiation frequency νlm. Here, we assume resonant polarization transfer (i.e., 5 = 56

    (')) and a weak effective interaction ,"##~1kHz (K'(( = K'(L = 9MHz, K*(( = K*(L = 2.5MHz, ℐ> = 30kHz). The amplitude Ωtu is kept constant at 25kHz.

    Figure S5. Polarization dynamics in the presence of RF irradiation. (a-b) Time response r̂(j) as a function of the irradiation frequency νlm for C1 (a) and C2 (b). (c-d) Fourier spectra '̂

    (Y)(M) and *̂(Y)(M) for C1 and

    C2 respectively (no RF irradiation). (e-f) Fourier spectra '̂(M; νlm) and *̂(M; νlm) for C1 and C2 respectively, in the presence of RF irradiation with νtu ≈ 2.16MHz and Ωtu = 25kHz. In all cases, K'(( = K'(L = 9MHz, K*(( = K*

    (L = 2.5MHz, and ℐ> = 30kHz.

  • 9

    Four possible transitions can be observed as distortions in the synchronized flip-flop dynamics shown in Fig. S5(a-b), each of them distinguished with a label (1,2,3,4). Using the Hamiltonian representation in Table S2, these transitions can be identified according to Table S3. In order to assess the effect of RF excitation, we compute the Fourier transform of r̂ j for a given (Ωlm, νlm). Each normalized Fourier spectrum r̂ M can be then compared to quantify how the two

    13C spins decouple in the presence of a time-dependent perturbation. Figs. S5(c-d) show the unperturbed spectra ̂ '

    (Y)M

    and *̂(Y)

    M for C1 and C2, respectively, obtained in the absence of RF irradiation. Figs. S5(e-f) show the perturbed spectra '̂ M; νlm and *̂ M; νlm for C1 and C2 respectively, obtained in the presence of RF excitation at a frequency νlm corresponding to transition #3 and a given (fixed) RF amplitude Ωtu. This approach provides a systematic way to evaluate the decoupling of the two 13C spins, by computing the overlap between the Fourier spectra r̂ M . Indeed, in cases (c-d) the spectral overlap is ~1 (as in the case of non-resonant RF excitation). Quite on the contrary, the overlap is less than 1 in cases (e-f), which means that the two 13C nuclei are, to some extent, decoupled. This magnitude is ultimately associated to the ‘dip’ in the obtained signal, as explicitly shown in Fig. 3 of the main text. References [1] D. Pagliero, P. R. Zangara, J. Henshaw, A. Ajoy, R. H. Acosta, J. A. Reimer, A. Pines, and C. A.

    Meriles, Sci. Adv. 6, eaaz6986 (2020). [2] A. De Luca, M. Collura, and J. De Nardis, Phys. Rev. B 96, 1 (2017). [3] A. Dréau, J. R. Maze, M. Lesik, J. F. Roch, and V. Jacques, Phys. Rev. B - Condens. Matter Mater.

    Phys. 85, 1 (2012). [4] B. Smeltzer, L. Childress, and A. Gali, New J. Phys. 13, (2011). [5] C. V. Peaker, M. K. Atumi, J. P. Goss, P. R. Briddon, A. B. Horsfall, M. J. Rayson, and R. Jones,

    Diam. Relat. Mater. 70, 118 (2016). [6] J. R. Johansson, P. D. Nation, and F. Nori, Comput. Phys. Commun. 184, 1234 (2013).

    Label States involved Frequency

    1 |↑ 0 ↑↓\⟩ ↔ |↓ 0 ↑↓\⟩ νlm~Mx + y(ℐ>*)

    2 |↑ 0 ↑↓\⟩ ↔ |↑ 0 ↑↑\⟩ νlm~∆*./2 + y(ℐ>*)

    3 |↑ 0 ↑↓\⟩ ↔ |↓ −1 ↓↓\⟩ νlm~∆*-/2 + y(ℐ>*)

    4 |↑ 0 ↑↓\⟩ ↔ |↑ −1 ↓↑\⟩ νlm~∆' + y(ℐ>*)

    Table S3. Transition frequencies excited by the RF irradiation (see Fig. S5).