Solid-state nuclear magnetic resonance can be enhanced using unpaired electron spins with a method known as dynamic nuclear polarization (DNP). Fundamentally, DNP involves ensembles of thousands of spins, a scale that is difficult to match computationally. This scale prevents us from gaining a complete understanding of the spin dynamics and applying simulations to design sample formulations. We recently developed an ab initio model capable of calculating DNP enhancements in systems of up to ∼1000 nuclei; however, this scale is insufficient to accurately simulate the dependence of DNP enhancements on radical concentration or magic angle spinning (MAS) frequency. We build on this work by using ab initio simulations to train a hybrid model that makes use of a rate matrix to treat nuclear spin diffusion. We show that this model can reproduce the MAS rate and concentration dependence of DNP enhancements and build-up time constants. We then apply it to predict the DNP enhancements in core–shell metal-organic-framework nanoparticles and reveal new insights into the composition of the particles’ shells.

The sensitivity of nuclear magnetic resonance (NMR) spectroscopy can be enhanced by transferring polarization from more magnetic electron spins, an approach known as dynamic nuclear polarization (DNP).1–3 The last decade has seen a dramatic rise in magic angle spinning (MAS)-DNP, in large part due to the development and commercialization of high-frequency gyrotron microwave (MW) sources.4,5 With the possible exception of the gyrotron, the greatest leaps in sensitivity have been made through the development of improved radical polarization sources. The most notable of which have been the invention of bisnitroxide polarizing agents that facilitate electron-to-nuclear polarization transfers by way of the three-spin cross-effect (CE) mechanism,6–8 the development of large molecular weight bisnitroxides that have increased relaxation times,9–11 and, most recently, the exploration of asymmetric biradicals.12–14 

Theory has played a large role in the development of new polarization sources. For instance, early on, it was realized that biradicals needed to be designed with relative radical orientations that limited the number of inactive spin pairs.15,16 A deeper understanding of the cross-effect mechanism during MAS revealed new ways to design DNP polarizing agents. Under MAS, the electron paramagnetic resonance (EPR) frequencies of the radicals are modulated due to the g-anisotropy. This rapid change in EPR frequency can lead to four types of anticrossing events,17–19 namely, (1) microwave (MW) events wherein a radical’s EPR frequency crosses the frequency of the applied microwaves, (2) dipolar and exchange (D/J) events that occur when the EPR frequencies of two radicals cross one another, (3) cross-effect6 (CE) events that occur when the difference between two radicals’ EPR frequencies equals the Larmor frequency of the nuclear spins, and finally, (4) solid-effect20,21 (SE) events that occur when the microwave frequency equals the sum or difference of the EPR and NMR frequencies. MW events lead to the saturation or inversion of an EPR resonance, D/J events lead to the exchange of magnetization between electron spins, CE events transfer magnetization to the nuclear spins of up to the difference in magnetization between the two electrons, and SE events lead to simultaneous flips of electron and nuclear spins.

Theoretical investigations revealed a particularly notorious interplay between the D/J and CE events.17,18 Given that D/J events are never 100% efficient, particularly for inter-molecular events, they can equilibrate the polarization of the electron spins, leading to the draining of nuclear magnetization via the cross-effect. This effect has been termed electron-induced depolarization and often leads to the overestimation of DNP enhancement factors.22,23 Inefficient D/J events thus lead to a lower than optimal difference in electron polarization and an overall lower level of hyperpolarization. This revelation has guided the development of improved polarizing agents that aimed to either increase the efficiency of D/J events24,25 or avoid them altogether.12–14,26–28

The role that internuclear interactions play in the overall DNP process is relatively underexplored. It is well understood that spin diffusion mediates the transfer of polarization to the observable nuclei;29–32 however, experiments show that depleting the nuclear spin bath in DNP surface-enhanced NMR spectroscopy (SENS) experiments counterintuitively leads to an increase in DNP enhancements,33 as does the perdeuteration of the polarizing agent,34,35 however, at the cost of a slower DNP build-up rate. It is thought that the increased performance in DNP SENS when using deuterated solvents is made possible by radicals adsorbing onto surfaces,36 coupled with a reduction in the overall number of spins needing to be polarized. Theoretical models to probe such mechanisms in silico are unfortunately difficult to design.

While MAS-DNP processes in small spin systems composed of two electrons and a handful of nuclei can be simulated exactly in Liouville space, the exponential scaling and need for very fine time steps of nanosecond order prevent the expansion of such methods to large spin systems.37–41 The most successful solution to this dilemma has been the use of classical Fickian diffusion to model the transfer of polarization either away from a point source (a radical molecule) or toward the interior of a nanoparticle.27,42–46 Such models have yielded a greater understanding of experimental results but do not provide a means to test new experiments or polarizing agents outside of the laboratory.

To increase the scale of predictive ab initio DNP simulations, Köckenberger proposed the use of restricted state-space methods,47,48 first focusing on static SE DNP. We later expanded this work to MAS-DNP by combining a restricted state-space approach to simulating spin diffusion, known as low-order correlations in Liouville space (LCL),50–52 with a Landau–Zener treatment of the electron spin dynamics. This increased the scale of the simulations to ∼50 nuclei, at the expense of the propagator, given that it would overrun the memory of the computer.53 Since the propagator could not be reused to calculate arbitrary time points, an optimization routine was used to find the steady-state DNP enhancements.53 The use of more aggressive locally restricted (LR) basis sets,54 similar to the IK-1 basis set implemented in Spinach,47,55 enabled for the increase of the spin system size to ∼1000 spins in ab initio simulations.56 These simulations revealed a steep decline in DNP enhancements within the first nanometer from the polarizing agent due to the so-called spin diffusion barrier.57 This abrupt loss in hyperpolarization was found to be dependent on the geometry of the biradical molecule, suggesting the possibility of engineering spin diffusion highways to direct polarization away from the radical.

Despite these successes, LR-LCL-based multinuclear MAS-DNP simulations are highly computationally demanding, requiring on average a week of computation on a computer cluster. Given that the bulk of the compute time is taken up by the treatment of the spin diffusion, we wondered whether this task could be replaced by a rate matrix. Such an approach has been successful in describing proton spin diffusion under MAS58 and has also been applied to large-scale MAS-DNP simulations.59 Rate constants could then be parameterized to reproduce the ab initio result in a small fraction of the time. This idea is closely related to the kinetic Monte Carlo methods developed by Köckenberger for describing static DNP61–62 and Corzilius’ treatment of spin diffusion in MAS-DNP.31 In these methods, however, estimates of the exchange rate constants are calculated from first-principles. The dramatic reduction in matrix sizes would once again enable the use of propagators and the calculation of DNP build-up times in addition to the steady-state enhancements. Here, we will describe the parameterization of the spin diffusion rate matrix and apply the new hybrid quantum-classical MAS-DNP model to simulate the MAS rate and concentration dependence of DNP enhancements and build-up times. Then, we apply this model to study the transfer of polarization into core–shell metal-organic-framework (MOF) nanoparticles.

An ensemble of spins can be described using the density operator, σ̂. Its time evolution, due to the spin Hamiltonian Ĥ(t), is described using the Liouville–von-Neumann equation,

(1)

The commutator between the Hamiltonian and the density operator is known as the Liouvillian (L̂̂), and the Hamiltonian is generally expressed as a sum of independent Hamiltonian operators that describe the various interactions affecting the spins. In the case of MAS-DNP, the Hamiltonian takes the following form in the microwave rotating frame:

(2)

where

(3)
(4)
(5)
(6)
(7)

ĤZ is the Zeeman Hamiltonian and describes the interaction between the spins and the magnetic field. In this expression, gzz,i is the element of the “i” electron’s g tensor that is aligned with the magnetic field, μB is the Bohr magneton, ωMW is the microwave frequency, σzz,j is the component of the “j” nucleus’ magnetic shielding tensor that is aligned with the magnetic field, and ωn is the nuclear Larmor frequency. Ĥe,e, Ĥe,n, and Ĥn,n describe the spin–spin coupling interactions between the electrons and nuclei either through space (d, A, and D) or through the exchange mechanism in the case of the electrons (J). Finally, ĤMW describes the interaction between the electrons and the microwaves, which has a (usually constant) Rabi frequency of ω1.

The evolution of σ̂ in Liouville space uses a propagator, Û̂,

(8)

which is time-dependent due to the MAS rotation. As such, it is necessary to divide time into P short increments of duration Δt over which Ĥ(t) can be assumed to be time-independent,

(9)

where T̂̂ is the Dyson time-ordering operator. If Δt is small, one can also apply the Suziki–Trotter approximation and propagate each of the Hamiltonians independently. To first order, this corresponds to63 

(10)

The exact solution of Eqs. (9) or (10) is typically beyond reach in all but the simplest of spin systems both due to the exponential scaling of σ̂ and the short time steps needed to treat the electron Hamiltonian as time-independent. The following sections will describe the approximations that are made to accelerate the treatment of the nuclei (Sec. II A) and electrons (Sec. II B), the optimization algorithm used to determine the steady-state nuclear polarizations (Sec. II C), and finally the hybrid quantum-classical model (Sec. II D).

The density operator is expressed in a basis set consisting of direct products of single-spin operators,47,50

(11)

where br is the coefficient for the basis operator B̂r and

(12)

In the above summation, qr corresponds to the spin order of the particular basis operator, which is the number of nuclei involved in the product. Îi,r is the operator associated with spin “i,” taking values of

(13)

where Êi is the identity operator.

In this basis, it is possible to distinguish operators by their spin order and coherence order and by the identity of the spins involved in the product. Given that no pulses are applied to the nuclei, we can freely remove the non-zero-quantum operators from the basis set without sacrificing accuracy.64 Given that we can only directly observe single-spin operators and that operators with high-qr must sequentially decrease in qr to contribute to an observable operator (qr = 1), it also follows that operators with low qr values have a larger impact on the spin dynamics.65 Indeed, it has been demonstrated, through comparisons with exact simulations, that limiting qr to a maximum value (qmax) of 4 is sufficient to reproduce an exact simulation in the case of MAS spin diffusion50 and solid effect DNP.67 These two state-space reductions describe the LCL basis set, which scales polynomially in size,47,50 rather than exponentially. At the limit where qmax is equal to the number of nuclei in system (N), the LCL result corresponds to the exact result.

Edwards et al. demonstrated that it is possible to further eliminate product operators involving distant nuclei for the calculation of spin diffusion in liquids without much loss in accuracy,66 an approach which we later adapted for the treatment of spin diffusion in rotating solids.54 In this locally restricted (LR)-LCL basis set, basis operators are only kept if they involve spins that are among the Nmax nearest neighbors of a given spin. A Nmax value of 27 is usually sufficient to reproduce the LCL result. Note that if Nmax equals N − 1, it corresponds to the LCL basis set.

High computational efficiency in LR-LCL calculations can be achieved with the use of a hash map to store the density operator and then propagate one dipolar interaction at a time using a Suzuki–Trotter propagation,50 

(14)

There, a list of the indices of a spin’s nearest neighbors, ni, is used. In practice, the propagation is achieved with the use of a series of four-dimensional subspaces with the corresponding rotation matrices

(15)

for subspaces of the form [ÎizB̂r, ÎjzB̂r, Ii+Îj−B̂r, Îi−Îj+B̂r] and

(16)

for subspaces of the form [ÎB̂r, ÎB̂r, ÎÎjzB̂r, ÎÎiB̂r], as described previously,50 where B̂r is an operator excluding spins i and j and ωD,i,j is the instantaneous dipolar frequency for the interaction between spins i and j,

(17)

RDD,i,j is the dipolar coupling constant between spins i and j, and three sets of Euler angles in Eq. (17) are used to rotate the spin system in the rotor around the magic angle, from the rotor frame to the interaction frame and, finally, to the lab frame.

The chemical shift propagator in this basis is a diagonal matrix with elements of

(18)

where pi is the coherence order of the i spin’s operator in the basis operator (i.e., 0 for Êi and Îi,z, 1 for Îi,+2, and −1 for Îi,2). Equation (18) can also be used to model the effects of hyperfine interactions on the spin diffusion rates by applying an offset corresponding to the strength of the electron-nuclear dipolar coupling at that particular orientation.

An important thing to note is that the number of 4 × 4 operations that need to be performed [Eqs. (14)(16)] is proportional to N, and as such, this type of calculation scales linearly with the size of the spin system, enabling ab initio calculations of spin diffusion in very large spin systems.54 

The explicit treatment of the electron spins has a significant impact on DNP calculations, given that each addition of an electron increases the basis set size by a factor of 4. Fortunately, electron non-zero quantum coherences are largely unpopulated and can be eliminated from the basis set.47,67 Further simplification was made possible in 2012 by Thurber and Tycko who showed that polarization transfers involving electrons occurred in extremely short level anticrossing events brought forth by the MAS rotation.17 By directly calculating the polarization transfers occurring during these events, using the Landau–Zener formula, and making the assumption that the crossings are infinitely short, the electrons could be effectively ignored, save for relaxation, for the bulk of the calculation with minimal losses in accuracy.

Mentink-Vigier later refined this work by representing the Landau–Zener events in the matrix form, which enabled longer calculations that take advantage of the periodicity of the MAS rotation and propagators.18,58 We will reiterate the different rotor events that can occur along with their corresponding propagators. Generally, such a calculation involves the pre-computation of the MAS dependence of the electron Larmor frequencies, ωei, to identify the time steps when a level anticrossing event occurs. The velocities of these events are stored to memory, along with the indices corresponding to the time increments when they occur for use in the later calculation.

When two electron spins’ EPR frequencies cross one another such that ωei = ωej, they will exchange magnetization in a quantity that is proportional to the strength of the dipolar (D) and exchange (J) interactions between them and inversely proportional to the rate of the following crossing:

(19)

where

(20)

if a D/J event occurs and is otherwise zero.

Note that in Eq. (19), the basis operators are ordered as Ê, followed by the electron Îz operators, and then the nuclear operators. Equation (19) corresponds specifically to a D/J event between electrons 1 and 2.

When the EPR frequency of an electron comes into resonance with the frequency of the applied microwaves (ωe,i = ωMW), it is partially saturated or inverted. The corresponding propagator for such an event is as follows in the case of electron 1:

(21)

where

(22)

if a MW event occurs and is otherwise zero.

When the microwave frequency matches either the electron-nuclear zero-quantum or double-quantum frequencies (ωei ± ωnj = ωMW), a solid effect event occurs, and polarization is transferred from the electron to the nuclei. The propagator for such an event is as follows in the case that the irradiated solid-effect condition involves electron 1:

(23)

where

(24)

when a SE event occurs and is otherwise zero.

Finally, if two electrons have a difference in EPR frequencies that equals the nuclear Larmor frequency [±(ωeiωej) = ωn], a cross effect event occurs and a portion of the difference in electron polarization is transferred to the nuclear spins. As an example, if we consider a cross-effect event between electrons 1 and 2 and nucleus 1, the propagator would equal,

(25)

where

(26)

when a CE event occurs and is otherwise zero.

We can thus define a global Landau–Zener propagator, which describes all interactions involving the Nel electrons,

(27)

MAS-DNP simulations are generally performed using the following strategy. We first simulate the evolution of the spin system for the duration of a rotor period and store the propagator,

(28)

where

(29)

and br,eq is the equilibrium coefficient of a given operator: the relative Boltzmann polarization for Iz-type operators, otherwise zero, and Tn is the relaxation rate of a given operator in the product: T1 for Iz-type operators and T2 for raising and lowering operators. Û̂r can then be exponentiated to calculate the spin polarizations at a later time point and predict DNP build-up curves, including the steady-state that yields the DNP enhancement factors in a fully relaxed spin system,

(30)

A problematic feature of this approach is that it requires the use of matrices with dimensions of n, where n is the number of basis operators. In particular, propagators are dense complex matrices and thus require the storage of 2n2 double-precision floating-point numbers. As such, a basis set only needs to contain 11k vectors for the propagator to cost over 2 GB to store, which corresponds to only 11 spins for LCL simulations. It is, nevertheless, possible to simulate far larger basis sets by working in low-dimensional subspaces via Suzuki–Trotter propagation (see Sec. II A), but these approaches are limited in timespan, given that they cannot take advantage of the periodicity of the rotor [Eq. (28)]. For instance, a 10 s simulation takes ten times longer to perform than a 1 s simulation.

It is, nevertheless, possible to exploit rotor periodicity in large-scale calculations of steady-state DNP enhancements by considering the following relationship:53 

(31)

Namely, at the steady-state, the DNP enhancements are invariant at the start and end of a rotor period. The calculation of DNP enhancements can thus be rephrased into a search for a set of nuclear polarizations that satisfy Eq. (31). Whether the nuclear polarizations increase or decrease after a rotor period reflects whether the initial guess of nuclear polarizations was too low or high. We first utilized this feature, in combination with LCL, to calculate DNP enhancements in moderately large spin systems using a simple Monte Carlo optimization algorithm where the nuclear polarizations, P, are stepped in their direction of change by a random step size between 0 and λ, tabulated in vector Λ,53 

(32)

In Eq. (32), the subscript m corresponds to the optimization step. The procedure for calculating steady-state DNP enhancements is then to propagate the density operator from an arbitrary initial guess for a period of tr. The polarizations are then updated according to Eq. (32), and the calculation is continued. This process is repeated until the ΔP values are below a certain threshold. Assuming that λ is sufficiently small, this process is guaranteed to converge to the steady-state DNP enhancements.

We, however, found Eq. (32) to scale poorly in very large spin systems due to the time it takes for a change in polarization from a given nucleus to affect that from a distant one. Far faster convergence is achieved if, in addition to the random jumps from Eq. (32), all nuclear polarizations are stepped in unison in a manner that is proportional to the total change in the nuclear polarizations,56 

(33)

where the factor μ corresponds to the step size. The same process can also be applied to smaller groups of neighboring Nmax spins, leading to

(34)

where

(35)

The step sizes λ, ν, and μ are set to obtain the fastest convergence; for the DNP enhancement of 1H spins, we have found that the following values work well: λ = 0.5, ν = 0.25, and μ = 0.5, with the value of λ decreasing by a factor of 0.999 at each optimization step. Note that the step sizes are normalized to the nuclear Boltzmann polarization.

The combination of LR-LCL and the Landau–Zener model for MAS-DNP has enabled the calculation of DNP enhancements in periodic spin systems containing upward of 1000 spins through the use of the optimization approaches described in Sec. II C.56 In this case, the propagation can be expressed as

(36)

which uses exclusively matrix-vector operations in small subspaces of up to four dimensions. By far, the most expensive part of the calculation is the LR-LCL part, which is needed to treat spin diffusion and the effects that hyperfine interactions have on spin diffusion rates (i.e., the so-called spin diffusion barrier). If the LR-LCL calculation could be replaced with a kinetic model, trained to reproduce the LR-LCL result, the space size could be dramatically reduced by a factor of about 2000 since the basis would only effectively contain the Îz operators and the identity operator. In this case, Eq. (36) would be simplified to

(37)

Here, K̂̂ is a matrix containing the spin diffusion rates between the different nuclei in the system, ki,j. The reduced matrix dimensions also enable the application of Eq. (30) and the calculation of DNP build-up curves in large atomistic models, hopefully, with the same level of accuracy as obtained using the ab initio model.

The application of the simulation approach summarized in Eq. (37) requires that the elements of K̂̂, namely, the spin diffusion rate constants ki,j, be predicted accurately. From perturbation theory, we expect that these rate constants will be proportional to RDD269–71 and inversely proportional to the MAS frequency, νr.72,73 The spin diffusion rates should also have some sort of inverse relationship with the strength of the hyperfine interactions, Azz,PAS, with the different electrons in the sample.74 We, thus, tested the following expression for the calculation of the rate constants:

(38)

The expression has three variable parameters, A, B, and C, which can be tuned to improve the agreement between an LR-LCL-based ab initio calculation and the hybrid model. We first optimized the value of A by setting B to zero and comparing the result to an ab initio calculation that does not apply a hyperfine shift operator. In other words, the spin diffusion barrier is removed. Next, B and C are optimized together against simulations that included the spin diffusion barrier. We were able to obtain a good agreement between the hybrid and ab initio simulations with values of A = 0.000 21, B = 0.0027 s C−1, and C = 1.2; see Fig. 1. The fitting was performed with the MAS rate in a range of 10–40 kHz on 16 mM periodic solutions of TEKPol11 and bTbK16 in 1,1,2,2-tetrachloroethane (TCE).56 

FIG. 1.

Comparison between DNP enhancements calculated for 16 mM periodic bTbK/TCE (a) and TEKPol/TCE (b) solutions containing one biradical molecule per unit cell. Calculations were repeated at different MAS rates (20 and 40 kHz MAS are shown here) both in the absence of a spin diffusion barrier (i) and in its presence (ii). Results from the first-principles calculations are in black, while those from the hybrid model are overlaid in red. Data are plotted as a function of a given nuclear spin’s distance to its nearest electron spin.

FIG. 1.

Comparison between DNP enhancements calculated for 16 mM periodic bTbK/TCE (a) and TEKPol/TCE (b) solutions containing one biradical molecule per unit cell. Calculations were repeated at different MAS rates (20 and 40 kHz MAS are shown here) both in the absence of a spin diffusion barrier (i) and in its presence (ii). Results from the first-principles calculations are in black, while those from the hybrid model are overlaid in red. Data are plotted as a function of a given nuclear spin’s distance to its nearest electron spin.

Close modal

Given that these parameters agree well with the ab initio results in TCE solutions of two different radicals, we expect that they are relatively reliable for TCE solutions, in general. In systems involving different NMR-active isotopes or differences in chemical shifts, it may be necessary to refine the fitted parameters or alter Eq. (38) to suit the particular situation. In certain circumstances, it may also be necessary to include high-order spin diffusion terms, for instance, at very fast MAS frequencies.75,76 In any case, we expect that ab initio simulations of DNP enhancements could be performed on all or part of a system of interest and used to refine Eq. (38) for the particular case, before calculating the build-up rates, or a larger spin system with the hybrid model.

Having optimized the hybrid MAS-DNP model for the reproduction of ab initio results in TCE solutions, we were interested to see whether it could be used to predict DNP enhancements and build-up curves as the radical concentration and MAS rate are altered. In particular, our previous ab initio simulations were limited to periodic models containing only a single biradical molecule. As such, it was not possible to properly assess the concentration dependence, given that inter-radical dipolar events could not be included in the simulation. In the following sections,1–3 we will compare the predicted DNP behavior of single- and multi-biradical solutions to that obtained from the experiment.

1. Modeling conformational and positional disorder

We previously used molecular dynamics simulations of biradical solutions to obtain representative ensembles of radical conformations and solvent proton distributions, both of which are expected to impact the calculated DNP enhancements.56 As we are now interested in simulating DNP in solutions containing upward of 20 000 atoms, this approach proved to be impractical. As such, we instead opted to use the software Balloon77 to search for the lowest-energy TEKPol conformers using a multiobjective genetic algorithm. 86 conformers were identified, and the five lowest-energy, non-duplicate, conformers were optimized using Density functional theory (DFT) in a polarizable continuum model (COSMO, 1,1,2,2-tetrachloroethane).78 The energy difference between the first and fifth lowest-energy conformers was −2.1 kcal/mol. Using these computed energies, we calculated the expected Boltzmann populations of the different conformers in solution at 100 K (Fig. 2).

FIG. 2.

The five lowest energy conformers of TEKPol, when dissolved in TCE, are displayed at the top. The relative populations of these conformers at 100 K are plotted on the bottom. See the supplementary material for more details.

FIG. 2.

The five lowest energy conformers of TEKPol, when dissolved in TCE, are displayed at the top. The relative populations of these conformers at 100 K are plotted on the bottom. See the supplementary material for more details.

Close modal

To generate a simulation model, we then defined a periodic unit cell with its dimensions set such that the radical concentration at room temperature would equal 2, 4, 8, 16, or 32 mM, with a given number of biradical molecules from 1 to 6. Radical molecules were inserted into the unit cell in randomized locations, with randomized orientations. The remainder of the cell was then populated with randomly located solvent protons with a density matching that found in frozen TCE.79 No interatomic contacts below 2 Å were allowed in the model generation.

2. Size-dependence of MAS-DNP computations: Electron-nuclear interactions

LR-LCL calculations scale linearly with the number of nuclei in the spin system since the number of four-dimensional matrix-vector products also scales linearly. In practice, optimization-based DNP calculations scale slightly worse due to the need for performing a greater number of optimization steps in larger spin systems (for example, 250 steps for 1250 nuclei vs 20 steps for 12 nuclei). A 1250-nuclei, two-electron, LR-LCL MAS-DNP calculation may take on the order of 2 days to complete on an AMD Epyc 7662 processor, but the same spin system can be calculated with the hybrid approach [Eq. (37)] in only 10 min. The hybrid calculations, however, scale as O(N3) due to the use of matrix–matrix products for the calculation of a rotor-periodic propagator. This places the practical limit of this simulation approach to roughly 10 000 nuclei, close to the memory limit for the storage of the propagator.

The main bottleneck of the calculations, however, shifts from the treatment of spin diffusion to the application of the numerous cross-effect propagators. As the number of electron spins in the system increases, so does the number of cross-effect events as well, and as such, the calculation also scales quadratically in the number of electron spins.

One way to accelerate the calculations is to reduce the number of cross-effect events such that electrons will only transfer polarization to their immediate neighbors. This concept is based on the fact that the efficiency of DNP transfers has an r−6 dependence, see Eq. (20), while the number of nuclei at a given distance from the radical only has an r2 dependence. While this approach has been used before in DNP simulations,58 it was not clear how many electron-nuclear interactions are needed to retain the accuracy of the calculation. We performed MAS-DNP simulations in a 2 mM TEKPol/TCE solution (2 electrons, 9544 nuclei) as a function of the number of nuclei to which each electron could directly transfer polarization using the cross-effect. If each electron was only allowed to interact with its closest nuclear spin, then the predicted DNP enhancements were lower by 72%. This error is quickly reduced, however, with the inclusion of more nuclei and is below 1% when the electrons each interact with 35 nuclear spins. The total computational time was reduced by a factor of 10 (see Fig. 3).

FIG. 3.

The errors introduced when truncating the polarization transfer to the electrons’ nearest nuclear spins are plotted as a function of the number of nuclei receiving direct polarization transfer through cross-effect. The required computation time for a given simulation is also plotted.

FIG. 3.

The errors introduced when truncating the polarization transfer to the electrons’ nearest nuclear spins are plotted as a function of the number of nuclei receiving direct polarization transfer through cross-effect. The required computation time for a given simulation is also plotted.

Close modal

It would, nevertheless, appear that a full half of the hyperpolarization that is fed to the 9544 nuclei in the spin system was initially generated on only four nuclear spins. These nuclei are situated on the biradical molecule itself, which highlights the importance of properly situating these innermost nuclei and providing them with efficient spin diffusion paths away from the electrons and into the bulk.80 

3. Size-dependence of MAS-DNP computations: Electron–electron interactions

DNP enhancements, relative depolarization levels, and build-up times were measured as a function of the MAS frequency, from 10 to 39 kHz in 5 kHz increments, for TEKPol/TCE solutions with biradical concentrations of 0, 2, 4, 8, 16, and 32 mM. The results are tabulated in Table I. Ion and Ioff are used to refer to the integrated signal volumes in the presence and absence of microwaves, respectively. Ion,10kHz and Ioff,10kHz correspond to the signal intensities at the lowest MAS frequency, 10 kHz, and are used to follow the MAS-dependence of the enhancement and depolarization factors. Ton and Toff are the build-up time constants in the presence and absence of microwaves, respectively. Due to challenges associated with introducing controlled quantities of the solution into the small 1.3 mm rotors, only relative values are reported.

TABLE I.

Experimental MAS and concentration dependence of DNP performance in TEKPol/TCE solutions.

[TEKPol] (mM)νR (kHz)Ion/IoffaIon/Ion,10kHzaIoff/Ioff,10kHzaTonb (s)Toffb (s)
10 0.94 0.94 4.76 4.76 
15 1.00 1.00 5.22 5.22 
20 1.00 1.00 5.74 5.74 
25 0.98 0.98 6.46 6.46 
30 1.03 1.03 7.62 7.62 
35 1.04 1.04 8.37 8.37 
39 0.97 0.97 8.91 8.91 
10 29 1.00 1.00 2.89 2.88 
15 27 0.90 0.97 3.04 3.18 
20 27 0.87 0.94 3.42 3.44 
25 28 0.85 0.88 4.02 3.92 
30 26 0.76 0.85 4.32 4.38 
35 28 0.79 0.82 4.81 4.85 
39 27 0.76 0.82 4.74 4.82 
10 93 1.00 1.00 4.27 3.83 
15 90 0.91 0.94 4.51 4.11 
20 85 0.84 0.92 4.89 4.49 
25 86 0.80 0.86 5.26 4.78 
30 87 0.78 0.83 5.82 5.42 
35 88 0.75 0.80 6.39 5.75 
39 89 0.73 0.76 6.57 5.98 
10 128 1.00 1.00 2.63 2.48 
15 128 0.90 0.90 2.51 2.40 
20 119 0.81 0.88 2.68 2.68 
25 124 0.76 0.79 3.00 2.95 
30 122 0.73 0.76 3.21 3.09 
35 123 0.69 0.72 3.51 3.41 
39 120 0.66 0.71 3.75 3.85 
16 10 150 1.00 1.00 1.95 2.10 
15 141 0.88 0.93 2.00 2.10 
20 132 0.73 0.83 2.10 2.40 
25 131 0.65 0.74 2.40 2.50 
30 134 0.64 0.72 2.40 2.60 
35 133 0.60 0.68 2.70 2.90 
39 134 0.60 0.67 2.70 2.80 
32 10 170 1.00 1.00 1.00 1.11 
15 163 0.88 0.92 1.09 1.08 
20 153 0.77 0.86 1.20 1.20 
25 155 0.73 0.80 1.26 1.30 
30 153 0.67 0.74 1.36 1.34 
35 146 0.66 0.76 1.44 1.46 
39 144 0.62 0.73 1.55 1.53 
[TEKPol] (mM)νR (kHz)Ion/IoffaIon/Ion,10kHzaIoff/Ioff,10kHzaTonb (s)Toffb (s)
10 0.94 0.94 4.76 4.76 
15 1.00 1.00 5.22 5.22 
20 1.00 1.00 5.74 5.74 
25 0.98 0.98 6.46 6.46 
30 1.03 1.03 7.62 7.62 
35 1.04 1.04 8.37 8.37 
39 0.97 0.97 8.91 8.91 
10 29 1.00 1.00 2.89 2.88 
15 27 0.90 0.97 3.04 3.18 
20 27 0.87 0.94 3.42 3.44 
25 28 0.85 0.88 4.02 3.92 
30 26 0.76 0.85 4.32 4.38 
35 28 0.79 0.82 4.81 4.85 
39 27 0.76 0.82 4.74 4.82 
10 93 1.00 1.00 4.27 3.83 
15 90 0.91 0.94 4.51 4.11 
20 85 0.84 0.92 4.89 4.49 
25 86 0.80 0.86 5.26 4.78 
30 87 0.78 0.83 5.82 5.42 
35 88 0.75 0.80 6.39 5.75 
39 89 0.73 0.76 6.57 5.98 
10 128 1.00 1.00 2.63 2.48 
15 128 0.90 0.90 2.51 2.40 
20 119 0.81 0.88 2.68 2.68 
25 124 0.76 0.79 3.00 2.95 
30 122 0.73 0.76 3.21 3.09 
35 123 0.69 0.72 3.51 3.41 
39 120 0.66 0.71 3.75 3.85 
16 10 150 1.00 1.00 1.95 2.10 
15 141 0.88 0.93 2.00 2.10 
20 132 0.73 0.83 2.10 2.40 
25 131 0.65 0.74 2.40 2.50 
30 134 0.64 0.72 2.40 2.60 
35 133 0.60 0.68 2.70 2.90 
39 134 0.60 0.67 2.70 2.80 
32 10 170 1.00 1.00 1.00 1.11 
15 163 0.88 0.92 1.09 1.08 
20 153 0.77 0.86 1.20 1.20 
25 155 0.73 0.80 1.26 1.30 
30 153 0.67 0.74 1.36 1.34 
35 146 0.66 0.76 1.44 1.46 
39 144 0.62 0.73 1.55 1.53 
a

Uncertainties in the enhancement factors are predominantly caused by background signals and are estimated to be ±5%.

b

A χ2 analysis shows that the uncertainties in the build-up times are generally on the order of 3%–10%.

Generally, we observe that while the vertical signal intensity increases with increasing MAS rate, the integrated intensity is fairly constant in the absence of microwaves and radicals. Magnetization losses as the MAS rate is increased, both in the presence and absence of microwaves, are seen to increase slightly with the radical concentration. This signal decrease is similar at the lower radical concentrations but is worse in the presence of microwaves in the higher-concentration samples, leading to a moderate overall reduction in the DNP enhancement factors (Ion/Ioff). Counter to prior reports where narrow-line radicals were used,27,81 we did not observe an increase in DNP efficiency as the MAS rate was increased when using a bisnitroxide. Integrated intensities generally decreased by 25%–40% as the MAS rate was increased from 10 to 39 kHz. Vertical signal intensities, however, did increase, although this was caused by the overall reduction in linewidths (see Table I, Fig. S3).

In all cases, build-up time constants were seen to increase with the MAS frequency and decrease with the radical concentration. Similar time constants were measured in both the presence and absence of microwaves, suggesting that cross-effect-induced nuclear (de)polarization dominates the relaxation processes in both cases. In the absence of radicals, a strong increase in build-up times was again observed as the MAS rate was increased, in agreement with the source–sink model introduced by Chaudhari and co-workers (see Table I).27 

We then performed MAS-DNP calculations, both in the presence and absence of microwaves and in models containing single or multiple biradical molecules per unit cell, to determine how inter-radical interactions impact the accuracy of the simulations. Nuclear relaxation times were set according to the following expression:

(39)

with T1,bulk set to the value measured at the corresponding MAS rate in the 0 mM solution. The electron relaxation times were set to the values measured experimentally on each of the solutions using pulsed X-band EPR spectroscopy (see the supplementary material). The field and orientational dependences of the electron relaxation times were neglected, which may lead to some errors in the calculated DNP behavior.38,82 The microwave power was set to 0.85 MHz, in accordance with previous studies.38,53,83 The results are tabulated in Tables II and III for the single- and multi-biradical models, respectively, and correspond to the average over five randomly generated solutions of a given radical concentration.

TABLE II.

Computed MAS and concentration dependence of DNP performance in TEKPol/TCE solutions for models consisting of a single biradical molecule.

[TEKPol] (mM)νR (kHz)Ion/IoffIon/IBoltzmannIon/Ion,10kHzIoff/IBoltzmannIoff/Ioff,10kHzTon (s)Toff (s)
10 38.07 41.86 1.00 1.10 1.00 4.01 4.12 
15 36.47 39.38 0.94 1.08 0.98 4.37 4.51 
20 35.11 37.54 0.90 1.07 0.97 4.75 4.93 
25 34.58 36.77 0.89 1.06 0.97 5.27 5.5 
30 35.26 37.35 0.90 1.06 0.96 6.08 6.4 
35 34.47 36.14 0.87 1.05 0.95 6.60 6.99 
39 33.45 34.69 0.84 1.04 1.00 6.95 7.41 
10 36.34 59.41 1.00 1.63 0.96 3.42 3.53 
15 34.72 54.37 0.92 1.57 0.90 3.73 3.89 
20 35.51 52.49 0.89 1.48 0.84 4.06 4.25 
25 35.80 49.36 0.84 1.38 0.81 4.50 4.76 
30 38.61 51.40 0.87 1.33 0.78 5.13 5.45 
35 40.02 50.89 0.86 1.27 0.75 5.55 5.94 
39 40.26 49.69 0.84 1.23 1.00 5.84 6.28 
10 115.45 122.55 1.00 1.06 1.00 2.64 2.85 
15 111.03 109.13 0.89 0.98 0.92 2.87 3.12 
20 107.12 100.24 0.82 0.94 0.88 3.10 3.39 
25 103.46 93.59 0.77 0.90 0.85 3.39 3.72 
30 101.48 89.53 0.73 0.88 0.83 3.80 4.17 
35 95.45 82.28 0.68 0.86 0.81 4.08 4.49 
39 90.97 77.21 0.63 0.85 1.00 4.28 4.72 
16 10 164.08 149.08 1.00 0.91 0.95 2.01 2.15 
15 162.64 141.23 0.95 0.87 0.90 2.16 2.31 
20 165.02 135.16 0.91 0.82 0.84 2.30 2.49 
25 169.73 130.60 0.88 0.77 0.78 2.48 2.72 
30 180.91 128.69 0.87 0.71 0.73 2.71 3.00 
35 185.13 124.33 0.84 0.67 0.69 2.88 3.20 
39 189.92 120.34 0.81 0.63 1.00 3.00 3.36 
32 10 316.47 170.34 1.00 0.54 1.00 1.29 2.16 
15 265.78 157.04 0.92 0.59 1.10 1.35 2.00 
20 260.53 155.73 0.92 0.60 1.11 1.44 2.02 
25 259.61 149.46 0.88 0.58 1.07 1.53 2.16 
30 260.29 145.88 0.86 0.56 1.04 1.65 2.30 
35 258.29 140.14 0.83 0.54 1.01 1.73 2.42 
39 254.10 135.66 0.80 0.53 0.99 1.79 2.49 
[TEKPol] (mM)νR (kHz)Ion/IoffIon/IBoltzmannIon/Ion,10kHzIoff/IBoltzmannIoff/Ioff,10kHzTon (s)Toff (s)
10 38.07 41.86 1.00 1.10 1.00 4.01 4.12 
15 36.47 39.38 0.94 1.08 0.98 4.37 4.51 
20 35.11 37.54 0.90 1.07 0.97 4.75 4.93 
25 34.58 36.77 0.89 1.06 0.97 5.27 5.5 
30 35.26 37.35 0.90 1.06 0.96 6.08 6.4 
35 34.47 36.14 0.87 1.05 0.95 6.60 6.99 
39 33.45 34.69 0.84 1.04 1.00 6.95 7.41 
10 36.34 59.41 1.00 1.63 0.96 3.42 3.53 
15 34.72 54.37 0.92 1.57 0.90 3.73 3.89 
20 35.51 52.49 0.89 1.48 0.84 4.06 4.25 
25 35.80 49.36 0.84 1.38 0.81 4.50 4.76 
30 38.61 51.40 0.87 1.33 0.78 5.13 5.45 
35 40.02 50.89 0.86 1.27 0.75 5.55 5.94 
39 40.26 49.69 0.84 1.23 1.00 5.84 6.28 
10 115.45 122.55 1.00 1.06 1.00 2.64 2.85 
15 111.03 109.13 0.89 0.98 0.92 2.87 3.12 
20 107.12 100.24 0.82 0.94 0.88 3.10 3.39 
25 103.46 93.59 0.77 0.90 0.85 3.39 3.72 
30 101.48 89.53 0.73 0.88 0.83 3.80 4.17 
35 95.45 82.28 0.68 0.86 0.81 4.08 4.49 
39 90.97 77.21 0.63 0.85 1.00 4.28 4.72 
16 10 164.08 149.08 1.00 0.91 0.95 2.01 2.15 
15 162.64 141.23 0.95 0.87 0.90 2.16 2.31 
20 165.02 135.16 0.91 0.82 0.84 2.30 2.49 
25 169.73 130.60 0.88 0.77 0.78 2.48 2.72 
30 180.91 128.69 0.87 0.71 0.73 2.71 3.00 
35 185.13 124.33 0.84 0.67 0.69 2.88 3.20 
39 189.92 120.34 0.81 0.63 1.00 3.00 3.36 
32 10 316.47 170.34 1.00 0.54 1.00 1.29 2.16 
15 265.78 157.04 0.92 0.59 1.10 1.35 2.00 
20 260.53 155.73 0.92 0.60 1.11 1.44 2.02 
25 259.61 149.46 0.88 0.58 1.07 1.53 2.16 
30 260.29 145.88 0.86 0.56 1.04 1.65 2.30 
35 258.29 140.14 0.83 0.54 1.01 1.73 2.42 
39 254.10 135.66 0.80 0.53 0.99 1.79 2.49 
TABLE III.

Computed MAS and concentration dependence of DNP performance in TEKPol/TCE solutions for models containing multiple biradical molecules.a

[TEKPol] (mM)νR (kHz)Ion/IoffIon/IBoltzmannIon/Ion,10kHzIoff/IBoltzmannIoff/Ioff,10kHzTon (s)Toff (s)
10 56.57 48.46 1.00 0.86 1.00 3.46 3.67 
15 49.53 41.33 0.86 0.83 0.98 3.79 4.02 
20 43.89 36.17 0.75 0.82 0.97 4.14 4.41 
25 39.81 32.32 0.67 0.81 0.95 4.60 4.91 
30 39.42 31.24 0.65 0.79 0.93 5.26 5.68 
35 36.58 28.65 0.60 0.78 0.92 5.70 6.16 
39 35.62 27.80 0.58 0.78 0.91 6.01 6.53 
10 105.59 82.46 1.00 0.78 1.00 2.91 3.09 
15 92.16 69.73 0.85 0.76 0.97 3.17 3.37 
20 89.21 65.91 0.80 0.74 0.94 3.43 3.67 
25 84.32 60.67 0.74 0.72 0.92 3.76 4.05 
30 85.42 59.09 0.72 0.69 0.88 4.23 4.6 
35 81.79 55.47 0.68 0.68 0.86 4.56 4.96 
39 78.45 52.50 0.65 0.67 0.86 4.79 5.24 
16 10 158.69 111.12 1.00 0.70 1.00 2.14 2.25 
15 145.25 94.74 0.86 0.65 0.93 2.33 2.45 
20 132.86 82.41 0.75 0.62 0.88 2.52 2.67 
25 124.46 73.60 0.67 0.59 0.84 2.75 2.93 
30 123.72 68.84 0.62 0.56 0.79 3.06 3.27 
35 116.87 62.82 0.57 0.54 0.76 3.28 3.51 
39 114.22 60.82 0.55 0.53 0.75 3.43 3.69 
32 10 229.62 134.79 1.00 0.59 1.00 1.33 1.38 
15 222.81 113.53 0.84 0.51 0.86 1.43 1.49 
20 205.40 94.74 0.70 0.46 0.78 1.56 1.60 
25 198.31 83.95 0.62 0.42 0.71 1.63 1.72 
30 202.60 78.15 0.58 0.39 0.64 1.75 1.89 
35 183.09 66.48 0.49 0.36 0.60 1.92 2.01 
39 179.39 64.02 0.48 0.36 0.59 1.98 2.10 
[TEKPol] (mM)νR (kHz)Ion/IoffIon/IBoltzmannIon/Ion,10kHzIoff/IBoltzmannIoff/Ioff,10kHzTon (s)Toff (s)
10 56.57 48.46 1.00 0.86 1.00 3.46 3.67 
15 49.53 41.33 0.86 0.83 0.98 3.79 4.02 
20 43.89 36.17 0.75 0.82 0.97 4.14 4.41 
25 39.81 32.32 0.67 0.81 0.95 4.60 4.91 
30 39.42 31.24 0.65 0.79 0.93 5.26 5.68 
35 36.58 28.65 0.60 0.78 0.92 5.70 6.16 
39 35.62 27.80 0.58 0.78 0.91 6.01 6.53 
10 105.59 82.46 1.00 0.78 1.00 2.91 3.09 
15 92.16 69.73 0.85 0.76 0.97 3.17 3.37 
20 89.21 65.91 0.80 0.74 0.94 3.43 3.67 
25 84.32 60.67 0.74 0.72 0.92 3.76 4.05 
30 85.42 59.09 0.72 0.69 0.88 4.23 4.6 
35 81.79 55.47 0.68 0.68 0.86 4.56 4.96 
39 78.45 52.50 0.65 0.67 0.86 4.79 5.24 
16 10 158.69 111.12 1.00 0.70 1.00 2.14 2.25 
15 145.25 94.74 0.86 0.65 0.93 2.33 2.45 
20 132.86 82.41 0.75 0.62 0.88 2.52 2.67 
25 124.46 73.60 0.67 0.59 0.84 2.75 2.93 
30 123.72 68.84 0.62 0.56 0.79 3.06 3.27 
35 116.87 62.82 0.57 0.54 0.76 3.28 3.51 
39 114.22 60.82 0.55 0.53 0.75 3.43 3.69 
32 10 229.62 134.79 1.00 0.59 1.00 1.33 1.38 
15 222.81 113.53 0.84 0.51 0.86 1.43 1.49 
20 205.40 94.74 0.70 0.46 0.78 1.56 1.60 
25 198.31 83.95 0.62 0.42 0.71 1.63 1.72 
30 202.60 78.15 0.58 0.39 0.64 1.75 1.89 
35 183.09 66.48 0.49 0.36 0.60 1.92 2.01 
39 179.39 64.02 0.48 0.36 0.59 1.98 2.10 
a

The values reported here are averages from five independent simulations with varying radical arrangements. Standard deviations between the averaged polarizations and build-up times were all within 10%.

Figure 4 shows the spatial distribution of the nuclear polarizations in the presence of microwaves, normalized to their Boltzmann polarizations (Ion/IBoltzmann), as a function of the radical concentration and MAS rate for both the single- and multi-biradical solutions. The models containing multiple biradicals show a much greater variation in the polarization levels of the nearest nuclei due to the differing efficiencies of certain biradicals situated perhaps in small clusters. The multi-biradical solutions are also more strongly affected by the MAS rate than their single-biradical counterparts, particularly at higher concentrations. This result may explain why bisnitroxide biradicals are seen to yield diminishing performance at fast-MAS while the narrow-line radical BDPA shows the opposite.27,81 More specifically, inter-radical dipolar events, which lead to the scrambling of electron polarization differences, are the major source of signal loss.

FIG. 4.

Calculated nuclear polarizations, plotted as a function of a given nucleus’ distance to its nearest electron spin, for periodic TEKPol/TCE solutions containing multiple (a) biradical molecules or a single (b) biradical molecule. Calculations were performed for radical concentrations ranging from 32 mM (i) to 2 mM (v) and at an array of MAS frequencies (10, 15, 20, 25, 30, 35, and 39 kHz), with the hotter colors (and lower enhancements) corresponding to the faster MAS rates.

FIG. 4.

Calculated nuclear polarizations, plotted as a function of a given nucleus’ distance to its nearest electron spin, for periodic TEKPol/TCE solutions containing multiple (a) biradical molecules or a single (b) biradical molecule. Calculations were performed for radical concentrations ranging from 32 mM (i) to 2 mM (v) and at an array of MAS frequencies (10, 15, 20, 25, 30, 35, and 39 kHz), with the hotter colors (and lower enhancements) corresponding to the faster MAS rates.

Close modal

In Fig. 4, we also see that the polarization in the bulk is dependent on the radical concentration, while the polarizations of the nearest nuclei are fairly constant. This indicates that T1 of the bulk spins, and the spin diffusion rate, are limiting the hyperpolarization of the bulk.

Comparing the calculations directly against the experimental results (Fig. 5), we can see that the nuclear polarizations are, indeed, more accurately reproduced in the multiradical solutions, particularly in the presence of microwaves. Single-radical solutions underestimate the MAS-dependence of the Ion polarizations. The Ioff polarizations are also considerably underestimated in the 32 mM case.

FIG. 5.

Nuclear polarizations calculated in the absence (a) and presence (b) of microwaves, normalized to the values calculated at 10 kHz MAS, are plotted against the experimental values. Results obtained for the different radical concentrations and models containing 1 or multiple biradical molecules are overlaid.

FIG. 5.

Nuclear polarizations calculated in the absence (a) and presence (b) of microwaves, normalized to the values calculated at 10 kHz MAS, are plotted against the experimental values. Results obtained for the different radical concentrations and models containing 1 or multiple biradical molecules are overlaid.

Close modal

If we compare the enhancement factors, either against the Boltzmann polarization or the off polarization (Fig. 6), we again see that the experiment is better replicated in the multi-biradical model. The Ion/Ioff enhancement, in particular, is highly overestimated in the single-biradical solutions. As can be seen in Fig. 6(b), the trends in the MAS-dependence of the enhancements at 16 and 32 mM (squares and circles) could not be reproduced in the single-biradical models. While this trend is, indeed, reproduced in the case of the multi-biradical models, the enhancement values themselves are not. The experiment is far better correlated with the calculated Ion/IBoltzmann enhancements, perhaps due to the calculated Ion/Ioff values being susceptible to the errors of both Ion and Ioff.

FIG. 6.

DNP enhancements calculated as a function of both the MAS rate and concentration are plotted against the experimental result where the calculated enhancements are given relative to Boltzmann polarizations (a) and the polarization calculated in the absence of microwave irradiation (b).

FIG. 6.

DNP enhancements calculated as a function of both the MAS rate and concentration are plotted against the experimental result where the calculated enhancements are given relative to Boltzmann polarizations (a) and the polarization calculated in the absence of microwave irradiation (b).

Close modal

In contrast to these observations, it would seem that the build-up times are hardly affected by inter-biradical interactions and are nearly identical in both models. There is a strong linear correlation between the calculated and experimental build-up time constants; however, some series appear either above or below the diagonal (Fig. 7). We believe that this discrepancy is predominantly caused by experimental variations in the triplet oxygen concentration in the samples. Given that the T1,bulk and T1e values were determined on physically different samples, any variations in their values may have led to the observed differences. In particular, we measured longer build-up time constants on the 4 mM solution than the 2 mM solution (Table I).

FIG. 7.

Calculated DNP build-up times are plotted against the experimental values. Results obtained for the different radical concentrations and models containing 1 or multiple biradical molecules are overlaid.

FIG. 7.

Calculated DNP build-up times are plotted against the experimental values. Results obtained for the different radical concentrations and models containing 1 or multiple biradical molecules are overlaid.

Close modal

There is significant interest in applying MAS-DNP for the characterization of the core–shell structures of materials.43,44,84 Previous efforts in this area have focused on the adaptation of Fickian models to describe the diffusion of hyperpolarization into the particles, but these results are highly reliant on having an accurate knowledge of spin diffusion and polarization rates. The application of atomistic models based on high-level ab initio theory, and accurate structures, is attractive for this purpose.

To test whether the hybrid quantum-classical model could be used to map the radial structures of nanoparticles, we synthesized a UiO-66-type MOF material possessing well-defined core and shell structures. The core was prepared using a 2,5-dihydroxy-1,4-benzenedicarboxylic acid linker and with Hf as the metal, while the shell was grown using Zr and 2-hydroxy-1,4-benzenedicarboxylic acid. This combination makes it possible to distinguish the core and the shell using both electron microscopy and NMR (Fig. 8). The low contrast Zr-MOF shell of the material was measured to have a thickness of 10.2 ± 1.5 nm, while the high contrast Hf-MOF core had a diameter of 155 ± 23 nm. DNP enhancements and build-up time constants are given in Table IV. As expected, we measured a larger DNP enhancement for the shell, which is in closer proximity to the radical solution. Note that the pores of the material are too small to permit TEKPol from entering and that a deuterated solvent was used to simplify simulations.

FIG. 8.

Chemical composition of the core and shell of the material (a). MAS-DNP NMR spectra acquired with and without the application of microwaves (b). High-angle annular dark-field (HAADF) (c) and energy-dispersive x-ray spectroscopy elemental mapping (EDS) (d) scanning TEM images of the core–shell MOF nanoparticles. The thin Zr-MOF shell has a low contrast than the Hf-MOF core.

FIG. 8.

Chemical composition of the core and shell of the material (a). MAS-DNP NMR spectra acquired with and without the application of microwaves (b). High-angle annular dark-field (HAADF) (c) and energy-dispersive x-ray spectroscopy elemental mapping (EDS) (d) scanning TEM images of the core–shell MOF nanoparticles. The thin Zr-MOF shell has a low contrast than the Hf-MOF core.

Close modal
TABLE IV.

Experimental DNP enhancements and build-up time constants for the core–shell MOF.

CoreShell
Ion/IofftotalIon/IoffTon (s)Toff (s)T1 (s)Ion/IoffTon (s)Toff (s)T1 (s)
3.0 2.5 1.7 1.9 2.1 4.0 0.85 1.7 3.9 
CoreShell
Ion/IofftotalIon/IoffTon (s)Toff (s)T1 (s)Ion/IoffTon (s)Toff (s)T1 (s)
3.0 2.5 1.7 1.9 2.1 4.0 0.85 1.7 3.9 

Unfortunately, despite the efficiency improvements, it is not possible to simulate most 1H-rich nanoparticles using an atomistic basis. The heat transfer properties of a spherical particle are, however, identical to those of a cone cut out from a sphere,85 and as such, it is reasonable to approximate the full particle with what is effectively a conical unit cell (Fig. 9). The area of the base of the cone then defines the radical concentration at the surface of the particle.

FIG. 9.

(a) Depiction of the process used to reduce the model from a spherical particle coated in biradical molecules to a cone cut out from the larger sphere. (b) Calculated radial profile of DNP enhancements in core–shell MOF particles of various diameters. (c) Calculated DNP enhancements for the core and shell nuclei as a function of the total particle diameter. (d) Calculated DNP build-up times for the core and shell nuclei as a function of the total particle diameter.

FIG. 9.

(a) Depiction of the process used to reduce the model from a spherical particle coated in biradical molecules to a cone cut out from the larger sphere. (b) Calculated radial profile of DNP enhancements in core–shell MOF particles of various diameters. (c) Calculated DNP enhancements for the core and shell nuclei as a function of the total particle diameter. (d) Calculated DNP build-up times for the core and shell nuclei as a function of the total particle diameter.

Close modal

Keeping the shell size constant at 12.5 nm (6 unit cells), we varied the size of the core from diameters of 20–300 nm and calculated the propagation of hyperpolarization from a single TEKPol molecule at the interface. The nuclear T1 values were set to the values measured experimentally in the absence of biradicals and the presence of deuterated 1,1,2,2-tetrachloroethane. The A, B, and C constants that describe spin diffusion rate constants were kept fixed to the values determined in Sec. III A. As the core size is increased, the DNP enhancements of all nuclei are reduced. Nuclei beyond 25 nm from the surface are largely not enhanced by DNP, and as such, DNP enhancements quickly converge to the DNP enhancements expected from a particle of infinite size (with the core nuclei having enhancements of 1). Notably, this suggests that a DNP difference experiment could be used to selectively observe the first ∼10 nm from the surface. In the small particle size limit, the build-up time constants of the core nuclei are longer than those of the shell despite them having shorter T1 relaxation sites (Table IV). We expect the surface radical concentration to impact the absolute DNP enhancements, but not necessarily their ratios.

As can be gleaned from the data depicted in Fig. 9, no single particle diameter can reproduce the experimentally measured DNP enhancements and build-up times. The average DNP enhancement over the entirety of the sample (Ion/Iofftotal) is, however, well-predicted by the simulations at 2.7 (vs 3.0 from the experiment). This result, therefore, strongly suggests that there is a partial mixing of the linkers in the shell and the core, with the shells, nevertheless, possessing a higher concentration of the 2-hydroxy-1,4-benzenedicarboxylate linker. This result is not entirely surprising, with linker exchange reactions having been used as an alternate approach for the synthesis of core–shell MOF materials.86 We envision that this method may be further applied to the study of other stratified MOF materials,87,88 where the exact localization of the strata is difficult to determine through microscopy.

Despite showing that accurate atomistic models can be used to probe structure at a macroscale using DNP, these models have clear scale limitations that render them intractable in many materials of interest. Fickian models have no such limitation and do not lead to increasing complexity as a model’s size is increased. Similarly, to our replacement of ab initio spin diffusion with a trained kinetic model, it may be possible to use these hybrid quantum-classical simulations as a means of parameterizing Fickian models and tackle very large problems with minimal parameterization.

A hybrid quantum-classical model for describing MAS-DNP polarization transfer processes in very large spin systems (>10 000 spins) was described. High-level ab initio-calculated DNP enhancements were used to parameterize a fast kinetic model for dealing with the costly nuclear spin diffusion, with the electron spin dynamics treated using the Landau–Zener formula. The accuracy of the method was tested against measurements of the DNP enhancements and build-up times in TEKPol solutions with concentrations ranging from 2 to 32 mM and MAS rates of 10–39 kHz. The inclusion of multiple biradical molecules in the periodic unit cell had a beneficial impact on the calculated DNP enhancements but did not seem to affect the build-up times. Both the DNP enhancements and build-up times (including concentration and MAS dependence) were relatively well predicted by the simulations; however, the depolarization level proved to be more challenging to predict quantitatively.

The model was then used to predict the diffusion of hyperpolarization into stratified MOF nanoparticles and characterize the spatial distribution of linkers. Calculations predicted DNP build-up times and enhancement factors that were in qualitative agreement with the experiment with the average enhancement being in close agreement with the measurements. The deviation between theory and experiment is fully consistent with the partial exchange of linkers between the core and the shell. The strong distance-dependence of DNP enhancements, and the introduction of a method to predict them from first principles, may prove useful in the characterization of other materials with core–shell structures. We suspect that hybrid quantum-classical DNP simulations may, additionally, be helpful for the design of larger-scale Fickian models that describe the distribution of hyperpolarization on a macroscale.

Unless otherwise stated, all reactions were carried out in an ambient atmosphere. Hafnium (IV) chloride (Alfa Aesar, 99.9%), zirconium (IV) oxychloride octahydrate (Acros Organics, 99.5%), formic acid (Alfa Aesar, 97.0%), acetic acid (Alfa Aesar, >99.0%), 2-hydroxyterephthalic acid (Alfa Aesar, >98.0%), 2,5-dihydroxyterephthalic acid (Alfa Aesar, 97.0%), methanol (Alfa Aesar, >99.8%), and N,N-dimethylformamide (DMF, Alfa Aesar, >99.8%) were purchased from the indicated sources and used without further purification. Ultrapure deionized water (deionized water, 18.2 MΩ) was obtained from a Barnstead nanopure system (Thermo Scientific).

2-hydroxyterephthalic acid (27.6 mg) was dissolved in 2 ml DMF in a 20 ml scintillation vial to which 2 ml of acetic acid was added. Then, 1 ml of a Hf-UiO-(OH)2 suspension in DMF was introduced to the above solution and stirred at 600 rpm under ambient conditions for 5 min. Finally, a 1 ml DMF solution containing zirconium (IV) oxychloride octahydrate (10.5 mg) was added. The solution was stirred at 90 °C in an oil bath for 4 h. After cooling, the solid was collected by centrifugation (8000 rpm, 10 min) and washed with DMF (30 ml) three times. The material was then activated by washing with DMF (30 ml) and methanol (30 ml) three times every 12 h. The activated MOF was vacuum dried at room temperature for 24 h.

Transmission electron microscopy (TEM) samples were prepared by drop-casting dilute particle suspensions onto a carbon film stabilized with formvar on 200 mesh Cu grids. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) images were taken using a Titan Themis 300 with a gun monochromator and probe spherical aberration corrector operated at 200 kV. The energy-dispersive x-ray spectroscopy (EDS) mapping was obtained with a super-X EDS detector.

All solid-state NMR experiments were performed using a Bruker AVANCE III 400 MHz/264 GHz MAS-DNP spectrometer equipped with either a 1.3 mm (TCE solutions) or 3.2 mm (MOFs) MAS-DNP probe.

Solutions of TEKPol in 1,1,2,2-tetrachloroethane were transferred to the 1.3 mm rotor using a syringe and spun at a temperature of 100 K. Saturation-recovery experiments were performed using 100 kHz 1H excitation pulses, and 1H spins were detected directly. Reported steady-state polarization corresponds to the plateau value in the saturation-recovery experiment. All spinning sidebands were integrated. The procedure was verified using radical-free solutions, where no variation in the integrated intensity was observed as a function of the MAS rate. Generally, the sample temperature decreases with an increase in the MAS rate due to the higher flow of gas. To ensure that observed differences were primarily due to changes in the polarization transfer mechanisms, and not temperature, the rotor was first allowed to equilibrate at 39 kHz for 30 min. Experiments were then performed in quick succession as the spinning frequency was reduced.

Core–shell MOF materials were impregnated with a 16 mM solution of TEKPol in deuterated 1,1,2,2-tetrachloroethane and spun at a frequency of 10 kHz. 1H enhancements were measured indirectly using 13C{1H} cross-polarization (CP). A 2.5 µs 1H excitation pulse was applied together with a 1 ms CP contact time. Relaxation and build-up times were measured using a saturation recovery experiment. Resonances were assigned by performing DNP-enhanced CPMAS experiments on non-core-shell MOFs.

Electron inversion-recovery experiments of solutions of TEKPol in TCE were performed using Fourier transform (FT) EPR spectroscopy on an X-band ELEXSYS 580 EPR system in an ER 4118X-MD5 dielectric resonator. Experiments were performed at 100 K to match DNP conditions. The microwave power attenuation and optimal pulse lengths were optimized on each formulation, with the typical 90° pulse being ∼32 ns. At least five times T1e was used for all shot repetition times to be quantitative.

The electron inversion-recovery experiments [Fig. S1(a)] were fit using a stretched exponential [Eq. (40)]. Note that due to imperfect inversion, both the intensity at time zero and the fully relaxed intensity were fitted independently. The mean electron longitudinal recovery time, T1e, is defined by using Eq. (41) and is given alongside the stretched exponential fit values in Table S1,

(40)
(41)

Conformer searches were performed using Balloon.77 The structure of the corresponding hydroxylamine of TEKPol was used as a starting point, and 2 × 106 generations were considered. The five lowest-energy, non-duplicate, conformers were selected for further optimization using DFT.

Spin-unrestricted DFT calculations were performed using NWChem.89 Geometries were first optimized in a vacuum at the PBE0/6-31G level of theory90,91 and then refined using COSMO79,92 in TCE.93 

All MAS-DNP simulations were performed using an in-house C/C++ program, which was run on AMD EPYC 7662 dual processors. Calculations were parallelized over the crystal orientations using message passing interface (MPI), with an instance of the software running on each of the 32 L3 caches. The most recent version of the source code can be obtained by contacting the authors.

Powder averaging was achieved using 66 orientations calculated using REPULSION.94 Three γ angles were also used for the ab initio simulations. Relaxation times were set to the values measured experimentally, as described in Sec. III. In most systems, the polarization transfers from the electrons to the nuclei were limited to the nearest 75 nuclei to accelerate the calculations. Euler angles describing the orientation of the electron g tensors, relative to the unit cell, were calculated using the method described by Adiga et al.95 The electron spin was assumed to be localized at the center of the N–O bond, with a negligible exchange constant. Periodic boundary conditions were applied for the simulation of the TCE solutions.

Simulations of the core–shell UiO-66 nanoparticles were performed using a 3 × 3 × N supercell of UiO-6696 and removing any nuclei outside of a cone with a base radius of 2.097 nm and a height of N. A single TEKPol molecule was then placed at the base of the cone in a prone orientation. Simulations did not make use of periodicity. Again, nuclear relaxation times were set to the values measured in the absence of radicals. The electron T1e was assumed to have been unchanged from that measured in the MOF-free 16 mM solution.

See the supplementary material for electron paramagnetic resonance and powder diffraction data and coordinates of the DFT-optimized TEKPol conformers.

Uddhav Kanbur and Dr. Sarah Cady are thanked for their help in conducting the electron paramagnetic resonance measurements. Dr. Lucas Griffiths is thanked for numerous fruitful discussions. This research was supported by the U.S. Department of Energy (DOE), Office of Science, Basic Energy Sciences, Materials Science and Engineering Division. The Ames Laboratory is operated for the DOE by Iowa State University under Contract No. DE-AC02-07CH11358. W.-S.L. acknowledges Boston College for support on the synthesis of core–shell materials.

The authors have no conflicts to disclose.

The data that support the findings of this study are available within the article and its supplementary material or from the corresponding author upon reasonable request.

1.
A. W.
Overhauser
,
Phys. Rev.
92
,
411
(
1953
).
2.
T. R.
Carver
and
C. P.
Slichter
,
Phys. Rev.
92
,
212
(
1953
).
3.
Handbook of High Field Dynamic Nuclear Polarization
, edited by
V. K.
Michaelis
,
R. G.
Griffin
,
B.
Corzilius
, and
S.
Vega
(
John Wiley & Sons
,
Chichester, UK
,
2020
).
4.
T.
Maly
,
G. T.
Debelouchina
,
V. S.
Bajaj
,
K.-N.
Hu
,
C.-G.
Joo
,
M. L.
Mak-Jurkauskas
,
J. R.
Sirigiri
,
P. C. A.
van der Wel
,
J.
Herzfeld
,
R. J.
Temkin
, and
R. G.
Griffin
,
J. Chem. Phys.
128
,
052211
(
2008
).
5.
M.
Rosay
,
L.
Tometich
,
S.
Pawsey
,
R.
bader
,
R.
Schauwecker
,
M.
Blank
,
P. M.
Borchard
,
S. R.
Cauffman
,
K. L.
Felch
,
R. T.
Weber
,
R. J.
Temkin
,
R. G.
Griffin
, and
W. E.
Maas
,
Phys. Chem. Chem. Phys.
12
,
5850
(
2010
).
6.
C. F.
Hwang
and
D. A.
Hill
,
Phys. Rev. Lett.
18
,
110
(
1967
).
7.
K.-N.
Hu
,
H.-h.
Yu
,
T. M.
Swager
, and
R. G.
Griffin
,
J. Am. Chem. Soc.
126
,
10844
(
2004
).
8.
C.
Song
,
K.-N.
Hu
,
C.-G.
Joo
,
T. M.
Swager
, and
R. G.
Griffin
,
J. Am. Chem. Soc.
128
,
11385
(
2006
).
9.
A.
Zagdoun
,
G.
Casano
,
O.
Ouari
,
G.
Lapadula
,
A. J.
Rossini
,
M.
Lelli
,
M.
Baffert
,
D.
Gajan
,
L.
Veyre
,
W. E.
Maas
,
M.
Rosay
,
R. T.
Weber
,
C.
Thieuleux
,
C.
Coperet
,
A.
Lesage
,
P.
Tordo
, and
L.
Emsley
,
J. Am. Chem. Soc.
134
,
2284
(
2012
).
10.
C.
Sauvée
,
M.
Rosay
,
G.
Casano
,
F.
Aussenac
,
R. T.
Weber
,
O.
Ouari
, and
P.
Tordo
,
Angew. Chem., Int. Ed.
52
,
10858
(
2013
).
11.
A.
Zagdoun
,
G.
Casano
,
O.
Ouari
,
M.
Schwarzwälder
,
A. J.
Rossini
,
F.
Aussenac
,
M.
Yulikov
,
G.
Jeschke
,
C.
Copéret
,
A.
Lesage
,
P.
Tordo
, and
L.
Emsley
,
J. Am. Chem. Soc.
135
,
12790
(
2013
).
12.
G.
Mathies
,
M. A.
Caporini
,
V. K.
Michaelis
,
Y.
Liu
,
K. N.
Hu
,
D.
Mance
,
J. L.
Zweier
,
M.
Rosay
,
M.
Baldus
, and
R. G.
Griffin
,
Angew. Chem., Int. Ed.
54
,
11770
(
2015
).
13.
D.
Wisser
,
G.
Karthikeyan
,
A.
Lund
,
G.
Casano
,
H.
Karoui
,
M.
Yulikov
,
G.
Menzildjian
,
A. C.
Pinon
,
A.
Purea
,
F.
Engelke
,
S. R.
Chaudhari
,
D.
Kubicki
,
A. J.
Rossini
,
I. B.
Moroz
,
D.
Gajan
,
C.
Copéret
,
G.
Jeschke
,
M.
Lelli
,
L.
Emsley
,
A.
Lesage
, and
O.
Ouari
,
J. Am. Chem. Soc.
140
,
13340
(
2018
).
14.
F.
Mentink-Vigier
,
G.
Mathies
,
Y.
Liu
,
A.-L.
Barra
,
M. A.
Caporini
,
D.
Lee
,
S.
Hediger
,
R. G.
Griffin
, and
G.
De Paëpe
,
Chem. Sci.
8
,
8150
(
2017
).
15.
K.-N.
Hu
,
C.
Song
,
H.-h.
Yu
,
T. M.
Swager
, and
R. G.
Griffin
,
J. Chem. Phys.
128
,
052302
(
2008
).
16.
Y.
Matsuki
,
T.
Maly
,
O.
Ouari
,
H.
Karoui
,
F.
Le Moigne
,
E.
Rizzato
,
S.
Lyubenova
,
J.
Herzfeld
,
T.
Prisner
,
P.
Tordo
, and
R. G.
Griffin
,
Angew. Chem., Int. Ed.
48
,
4996
(
2009
).
17.
K. R.
Thurber
and
R.
Tycko
,
J. Chem. Phys.
137
,
084508
(
2012
).
18.
F.
Mentink-Vigier
,
Ü.
Akbey
,
Y.
Hovav
,
S.
Vega
,
H.
Oschkinat
, and
A.
Feintuch
,
J. Magn. Reson.
224
,
13
(
2012
).
19.
K. R.
Thurber
and
R.
Tycko
,
Isr. J. Chem.
54
,
39
(
2014
).
20.
C. D.
Jeffries
,
Phys. Rev.
106
,
164
(
1957
).
21.
M.
Abragam
,
R. W.
Kedzie
, and
C. D.
Jeffries
,
Phys. Rev.
106
,
165
(
1957
).
22.
K. R.
Thurber
and
R.
Tycko
,
J. Chem. Phys.
140
,
184201
(
2014
).
23.
F.
Mentink-Vigier
,
S.
Paul
,
D.
Lee
,
A.
Feintuch
,
S.
Hediger
,
S.
Vega
, and
G.
De Paëpe
,
Phys. Chem. Chem. Phys.
17
,
21824
(
2015
).
24.
F.
Mentink-Vigier
,
I.
Marin-Montesinos
,
A. P.
Jagtap
,
T.
Halbritter
,
J.
van Tol
,
S.
Hediger
,
D.
Lee
,
S. T.
Sigurdsson
, and
G.
De Paëpe
,
J. Am. Chem. Soc.
140
,
11013
(
2018
).
25.
A.
Lund
,
G.
Casano
,
G.
Menzildjian
,
M.
Kaushik
,
G.
Stevanato
,
M.
Yulikov
,
R.
Jabbour
,
D.
Wisser
,
M.
Renom-Carrasco
,
C.
Thieuleux
,
F.
Bernada
,
H.
Karoui
,
D.
Siri
,
M.
Rosay
,
I. V.
Sergeyev
,
D.
Gajan
,
M.
lelli
,
L.
Emsley
,
O.
Ouari
, and
A.
Lesage
,
Chem. Sci.
11
,
2810
(
2020
).
26.
T. V.
Can
,
M. A.
Caporini
,
F.
Mentink-Vigier
,
B.
Corzilius
,
J. J.
Walish
,
M.
Rosay
,
W. E.
Maas
,
M.
Baldus
,
S.
Vega
,
T. M.
Swager
, and
R. G.
Griffin
,
J. Chem. Phys.
141
,
064202
(
2014
).
27.
S. R.
Chaudhari
,
D.
Wisser
,
A. C.
Pinon
,
P.
Berruyer
,
D.
Gajan
,
P.
Tordo
,
O.
Ouari
,
C.
Reiter
,
F.
Engelke
,
C.
Copéret
,
M.
Lelli
,
A.
Lesage
, and
L.
Emsley
,
J. Am. Chem. Soc.
139
,
10609
(
2017
).
28.
G.
Stevanato
,
D. J.
Kubicki
,
G.
Menzildjian
,
A.-S.
Chauvin
,
K.
Keller
,
M.
Yulikov
,
G.
Jeschke
,
M.
Mazzanti
, and
L.
Emsley
,
J. Am. Chem. Soc.
141
,
8746
(
2019
).
29.
C.
Ramanathan
,
Appl. Magn. Reson.
34
,
409
(
2008
).
30.
Y.
Hovav
,
A.
Feintuch
, and
S.
Vega
,
J. Chem. Phys.
134
,
074509
(
2011
).
31.
J. J.
Wittmann
,
M.
Eckardt
,
W.
Harneit
, and
B.
Corzilius
,
Phys. Chem. Chem. Phys.
20
,
11418
(
2018
).
32.
Q.
Stern
,
S. F.
Cousin
,
F.
Mentink-Vigier
,
A. C.
Pinon
,
S. J.
Elliot
,
O.
Cala
, and
S.
Jannin
,
Sci. Adv.
7
,
eabf5735
(
2021
).
33.
T.
Kobayashi
,
F. A.
Perras
,
U.
Chaudhary
,
I. I.
Slowing
,
W.
Huang
,
A. D.
Sadow
, and
M.
Pruski
,
Solid State Nucl. Magn. Reson.
87
,
38
(
2017
).
34.
D. J.
Kubicki
,
G.
Casano
,
M.
Schwarzwälder
,
S.
Abel
,
C.
Sauvée
,
K.
Ganesan
,
M.
Yulikov
,
A. J.
Rossini
,
G.
Jeschke
,
C.
Copéret
,
A.
Lesage
,
P.
Tordo
,
O.
Ouari
, and
L.
Emsley
,
Chem. Sci.
7
,
550
(
2016
).
35.
F. A.
Perras
,
R. R.
Reinig
,
I. I.
Slowing
,
A. D.
Sadow
, and
M.
Pruski
,
Phys. Chem. Chem. Phys.
18
,
65
(
2016
).
36.
F. A.
Perras
,
L.-L.
Wang
,
J. S.
Manzano
,
U.
Chaudhary
,
N. N.
Opembe
,
D. D.
Johnson
,
I. I.
Slowing
, and
M.
Pruski
,
Curr. Opin. Colloid Interface Sci.
33
,
9
(
2018
).
37.
F.
Mentink-Vigier
,
Ü.
Akbey
,
H.
Oschkinat
,
S.
Vega
, and
A.
Feintuch
,
J. Magn. Reson.
258
,
102
(
2015
).
38.
D.
Mance
,
P.
Gast
,
M.
Huber
,
M.
Baldus
, and
K. L.
Ivanov
,
J. Chem. Phys.
142
,
234201
(
2015
).
39.
F. A.
Perras
,
A.
Sadow
, and
M.
Pruski
,
ChemPhysChem
18
,
2279
(
2017
).
40.
F.
Mentink-Vigier
,
A.-L.
Barra
,
J.
van Tol
,
S.
Hediger
,
D.
Lee
, and
G.
De Paëpe
,
Phys. Chem. Chem. Phys.
21
,
2166
(
2019
).
41.
F.
Mentink-Vigier
,
Phys. Chem. Chem. Phys.
22
,
3643
(
2020
).
42.
A. J.
Rossini
,
A.
Zagdoun
,
F.
Hegner
,
M.
Schwarzwälder
,
D.
Gajan
,
C.
Copéret
,
A.
Lesage
, and
L.
Emsley
,
J. Am. Chem. Soc.
134
,
16899
(
2012
).
43.
A. C.
Pinon
,
J.
Schlagnitweit
,
P.
Berruyer
,
A. J.
Rossini
,
M.
Lelli
,
E.
Socie
,
M.
Tang
,
T.
Pham
,
A.
Lesage
,
S.
Schantz
, and
L.
Emsley
,
J. Phys. Chem. C
121
,
15993
(
2017
).
44.
A. C.
Pinon
,
U.
Skantze
,
J.
Viger-Gravel
,
S.
Schantz
, and
L.
Emsley
,
J. Phys. Chem. A
122
,
8802
(
2018
).
45.
S.
Björgvinsdóttir
,
B. J.
Walder
,
A. C.
Pinon
, and
L.
Emsley
,
J. Am. Chem. Soc.
140
,
7946
(
2018
).
46.
N. A.
Prisco
,
A. C.
Pinon
,
L.
Emsley
, and
B. F.
Chmelka
,
Phys. Chem. Chem. Phys.
23
,
1006
(
2021
).
47.
I.
Kuprov
,
N.
Wagner-Rundell
, and
P. J.
Hore
,
J. Magn. Reson.
189
,
241
(
2007
).
48.
A.
Karabanov
,
A.
van der Drift
,
L. J.
Edwards
,
I.
Kuprov
, and
W.
Köckenberger
,
Phys. Chem. Chem. Phys.
14
,
2658
(
2012
).
49.
M. C.
Butler
,
J.-N.
Dumez
, and
L.
Emsley
,
Chem. Phys. Lett.
477
,
377
(
2009
).
50.
J.-N.
Dumez
,
M. C.
Butler
, and
L.
Emsley
,
J. Chem. Phys.
133
,
224501
(
2010
).
51.
J.-N.
Dumez
,
M. C.
Butler
,
E.
Salager
,
B.
Elena-Herrmann
, and
L.
Emsley
,
Phys. Chem. Chem. Phys.
12
,
9172
(
2010
).
52.
M. E.
Halse
,
A.
Zagdoun
,
J.-N.
Dumez
, and
L.
Emsley
,
J. Magn. Reson.
254
,
48
(
2015
).
53.
F. A.
Perras
and
M.
Pruski
,
J. Chem. Phys.
149
,
154202
(
2018
).
54.
F. A.
Perras
and
M.
Pruski
,
J. Chem. Phys.
151
,
034110
(
2019
).
55.
I.
Kuprov
,
Magn. Reson. Chem.
56
,
415
(
2018
).
56.
F. A.
Perras
,
M.
Raju
,
S. L.
Carnahan
,
D.
Akbarian
,
A. C. T.
van Duin
,
A. J.
Rossini
, and
M.
Pruski
,
J. Phys. Chem. Lett.
11
,
5655
(
2020
).
57.
J. P.
Wolfe
,
Phys. Rev. Lett.
31
,
907
(
1973
).
58.
B.
Elena
,
G.
Pintacuda
,
N.
Mifsud
, and
L.
Emsley
,
J. Am. Chem. Soc.
128
,
9555
(
2006
).
59.
F.
Mentink-Vigier
,
S.
Vega
, and
G.
De Paëpe
,
Phys. Chem. Chem. Phys.
19
,
3506
(
2017
).
60.
A.
Karabanov
,
D.
Wiśniewski
,
I.
Lesanovsky
, and
W.
Köckenberger
,
Phys. Rev. Lett.
115
,
020404
(
2015
).
61.
D.
Wiśniewski
,
A.
Karabanov
,
I.
Lesanovsky
, and
W.
Köckenberger
,
J. Magn. Reson.
264
,
30
(
2016
).
62.
A.
Karabanov
,
D.
Wiśniewski
,
F.
Raimondi
,
I.
Lesanovsky
, and
W.
Köckenberger
,
Phys. Rev. A
97
,
031404
(
2018
).
64.
H. J.
Hogben
,
P. J.
Hore
, and
I.
Kuprov
,
J. Chem. Phys.
132
,
174101
(
2010
).
65.
L. J.
Edwards
,
D. V.
Savostyanov
,
A. A.
Nevzorov
,
M.
Concistrè
,
G.
Pileio
, and
I.
Kuprov
,
J. Magn. Reson.
235
,
121
(
2013
).
66.
L. J.
Edwards
,
D. V.
Savostyanov
,
Z. T.
Welderufael
,
D.
Lee
, and
I.
Kuprov
,
J. Magn. Reson.
243
,
107
(
2014
).
67.
A.
Karabanov
,
I.
Kuprov
,
G. T. P.
Charnock
,
A.
van der Drift
,
L. J.
Edwards
, and
W.
Köckenberger
,
J. Chem. Phys.
135
,
084106
(
2011
).
69.
D.
Suter
and
R. R.
Ernst
,
Phys. Rev. B
32
,
5608
(
1985
).
70.
P. M.
Henrichs
,
M.
Linder
, and
J. M.
Hewitt
,
J. Chem. Phys.
85
,
7077
(
1986
).
71.
A.
Kubo
and
C. A.
McDowell
,
J. Chem. Phys.
89
,
63
(
1988
).
72.
A.
Kubo
and
C. A.
McDowell
,
J. Chem. Soc., Faraday Trans. 1
84
,
3713
(
1988
).
73.
Z.
Jia
,
L.
Zhang
,
Q.
Chen
, and
E. W.
Hansen
,
J. Phys. Chem. A
112
,
1228
(
2008
).
74.
G. R.
Khutsishvili
,
Sov. Phys. JETP
15
,
909
(
1962
).
75.
L.-S.
Du
,
M. H.
Levitt
, and
C. P.
Grey
,
J. Magn. Reson.
140
,
242
(
1999
).
76.
V.
Agarwal
,
J. Magn. Reson.
311
,
106661
(
2020
).
77.
M. J.
Vainio
and
M. S.
Johnson
,
J. Chem. Inf. Model.
47
,
2462
(
2007
).
78.
A.
Klamt
and
G.
Schüürmann
,
J. Chem. Soc., Perkin Trans. 2
1993
,
799
.
79.
M.
Bujak
and
A.
Katrusiak
,
Z. Kristallogr.
219
,
669
(
2004
).
80.
G.
Stevanato
,
G.
Casano
,
D. J.
Kubicki
,
Y.
Rao
,
L.
Esteban Hofer
,
G.
Menzildjian
,
H.
Karoui
,
D.
Siri
,
M.
Cordova
,
M.
Yulikov
,
G.
Jeschke
,
M.
Lelli
,
A.
Lesage
,
O.
Ouari
, and
L.
Emsley
,
J. Am. Chem. Soc.
142
,
16587
(
2020
).
81.
P.
Berruyer
,
S.
Björgvinsdóttir
,
A.
Bertarello
,
G.
Stevanato
,
Y.
Rao
,
G.
Karthikeyan
,
G.
Casano
,
O.
Ouari
,
M.
Lelli
,
C.
Reiter
,
F.
Engelke
, and
L.
Emsley
,
J. Phys. Chem. Lett.
11
,
8386
(
2020
).
82.
E. M. M.
Weber
,
H.
Vezin
,
J. G.
Kempf
,
G.
Bodenhausen
,
D.
Abergél
, and
D.
Kurzbach
,
Phys. Chem. Chem. Phys.
19
,
16087
(
2017
).
83.
E. A.
Nanni
,
A. B.
Barnes
,
Y.
Matsuki
,
P. P.
Woskov
,
B.
Corzilius
,
R. G.
Griffin
, and
R. J.
Temkin
,
J. Magn. Reson.
210
,
16
(
2011
).
84.
J.
Viger-Gravel
,
A.
Schantz
,
A. C.
Pinon
,
A. J.
Rossini
,
S.
Schantz
, and
L.
Emsley
,
J. Phys. Chem. B
122
,
2073
(
2018
).
85.
H. S.
Carslaw
and
J. C.
Jaeger
,
Conduction of Heat in Solids
(
Oxford at the Clarendon Press
,
UK
,
1959
), pp.
248
253
.
86.
R. A.
Dodson
,
A. P.
Kalenak
, and
A. J.
Matzger
,
J. Am. Chem. Soc.
142
,
20806
(
2020
).
87.
D. R.
Du Bois
and
A. J.
Matzger
,
J. Am. Chem. Soc.
143
,
671
(
2021
).
88.
T.-Y.
Luo
,
C.
Liu
,
X. Y.
Gan
,
P. F.
Muldoon
,
N. A.
Diemler
,
J. E.
Millstone
, and
N. L.
Rosi
,
J. Am. Chem. Soc.
141
,
2161
(
2019
).
89.
M.
Valiev
,
E. J.
Bylaska
,
N.
Govind
,
K.
Kowalski
,
T. P.
Straatsma
,
H. J. J.
Van Dam
,
D.
Wang
,
J.
Nieplocha
,
E.
Apra
,
T. L.
Windus
, and
W. A.
de Jong
,
Comput. Phys. Commun.
181
,
1477
(
2010
).
90.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
,
Phys. Rev. Lett.
77
,
3865
(
1997
).
91.
C.
Adamo
and
V.
Barone
,
J. Chem. Phys.
110
,
6158
(
1999
).
92.
A.
Klamt
,
J. Phys. Chem.
99
,
2224
(
1995
).
93.
P.
Berruyer
,
M.
Lelli
,
M. P.
Conley
,
D. L.
Silverio
,
C. M.
Widdifield
,
G.
Siddiqi
,
D.
Gajan
,
A.
Lesage
,
C.
Copéret
, and
L.
Emsley
,
J. Am. Chem. Soc.
139
,
849
(
2017
).
94.
M.
Bak
and
N. C.
Nielsen
,
J. Magn. Reson.
125
,
132
(
1997
).
95.
S.
Adiga
,
D.
Aebi
, and
D. L.
Bryce
,
Can. J. Chem.
85
,
496
(
2007
).
96.
S.
Øien
,
D.
Wragg
,
H.
Reinsch
,
S.
Svelle
,
S.
Bordiga
,
C.
Lamberti
, and
K. P.
Lillerud
,
Cryst. Growth Des.
14
,
5370
(
2014
).

Supplementary Material