Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

The nuclear magneto-electric response of a chiral molecule via molecular dynamics in a time-dependent electric field

Mateusz A. Słowiński a, Juha Vaara b and Piotr Garbacz *a
aFaculty of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland. E-mail: pgarbacz@uw.edu.pl
bNMR Research Unit, University of Oulu, P.O. Box 3000, FI-90014 Oulu, Finland

Received 16th June 2025 , Accepted 8th September 2025

First published on 3rd October 2025


Abstract

A chiral molecule with a permanent electric dipole moment aligns partially in an external electric field, preventing antisymmetric nuclear spin interactions from averaging out. Molecular dynamics simulations were used to investigate two such interactions – antisymmetric nuclear magnetic shielding and indirect spin–spin coupling in the light fluorinated alcohol, 1,1,1-trifluoropropan-2-ol. The results show that the rate at which a radiofrequency electric field oscillates significantly influences the spin states induced by these interactions, particularly when the frequency approaches a few gigahertz. This effect can be explained by considering dielectric losses in the electromagnetic field, which alter the amplitude and phase of the chirality-sensitive signal. As a result, at sufficiently high frequencies, the signal phase associated with a specific enantiomer may become reversed.


1. Introduction

In nuclear magnetic resonance (NMR) spectroscopy, antisymmetric nuclear interactions are closely linked to molecular chirality.1 Two key interactions exhibiting this antisymmetry are magnetic shielding2 and indirect spin–spin coupling.3 This study extends previous theoretical approaches used in this field by moving beyond tensor symmetry-based statistical averaging over molecular ensembles.4 Instead, we directly analyze detailed molecular trajectories from molecular dynamics (MD) simulations. Notably, we relax the assumption that molecules with a permanent dipole moment instantly align with an external electric field – a simplification that becomes questionable even at relatively low frequencies on the order of a fraction of a gigahertz. For instance, at room temperature, the real part of the dielectric permittivity of propan-2-ol measured at a frequency of 0.6 GHz is only one-half of that obtained for a static electric field.5,6 The model adopted in this work explicitly accounts for the internal friction within the fluid arising from intermolecular interactions. Considering the importance of internal molecular dynamics in NMR, this approach is essential for more realistic results, and we present an application to 1,1,1-trifluoropropan-2-ol (hereafter referred to as TFP; in the following, the results are exemplified by the (R)-enantiomer of TFP).

We employ three computational tools to compute the expected, chirality-dependent nuclear magnetoelectric resonance (NMER) response of a polar liquid (the summary is provided in Table S1 in the SI; tables marked with the letter S, which are referred to later in the text, can similarly be found in the SI). Initially, we used density-functional theory methods (see Section 2.1 for details) to determine the components of the antisymmetric nuclear spin interaction tensors and their dependence on the central intramolecular degree of freedom, namely the dihedral angle defined by the H–C–O–H atoms, which is involved in a large-amplitude motion in this system. Subsequently, we utilized molecular dynamics simulations to track the time evolution of the components of the local reference frame vectors of the molecule, expressed in the laboratory coordinate system. Finally, the parameters obtained from both methods were incorporated into calculations of the time-dependent amplitudes of quantum states using a master equation of the spin system in the Lindbladian form.7–9

In constructing the master equation of the spin system, we focused on the 19F and 1H nuclei in the TFP molecule (see Table S2 for details of the assignment of the nuclei). Specifically, we examined the fluorine nuclei in the CF3 group, considering their antisymmetric shielding, as well as their interactions with the CH proton, in the antisymmetric indirect spin–spin coupling. We assumed that the fluorine atoms in the CF3 group undergo rapid exchange (due to the rotation of the group, which is very fast in the NMR time scale) and can, therefore, be treated as equivalent. Consequently, instead of analyzing a C19F3–C1H four-spin system, we restricted our analysis to a 19F–1H two-spin system with magnetic shielding and spin–spin coupling tensors averaged over the fluorine nuclei. We assumed dipolar relaxation to dominate over other relaxation mechanisms. The dipolar relaxation between 19F nuclei and the CH proton was described using the dipolar relaxation superoperator. The relaxation terms originating from interactions between the 19F nuclei and between the protons of the CH3 group and those of the CH proton were incorporated phenomenologically as the T1 and T2 values (see Table S1 for details).

It is convenient to express the Hamiltonian image file: d5cp02294k-t1.tif of the 19F–1H spin system in a TFP molecule, subjected to both a magnetic field B and an electric field E, using an irreducible tensor decomposition with respect to three-dimensional rotations. Specifically, the isotropic (rank-0) part of a two-index tensor [Fraktur T] is defined as one-third of its matrix trace, i.e., image file: d5cp02294k-t2.tif. In further text, the non-bolded symbols represent the isotropic components of the respective tensors. The star notation image file: d5cp02294k-t3.tif indicates the antisymmetric (rank-1) part of a tensor, defined as half the difference between the matrix and its transpose image file: d5cp02294k-t4.tif; see ref. 2 for a rigorous description. Here, the traceless symmetric (rank-2) components of the shielding tensor and the indirect spin–spin coupling tensor are neglected, as they are small for 1H and 19F in 1,1,1-trifluoropropan-2-ol, as compared to the direct coupling tensor components. By definition, the direct coupling tensor is traceless and symmetric; consequently, both its isotropic and antisymmetric components vanish.10

Let 19F be the first spin and 1H the second. Then, their nuclear spin Hamiltonian is given by

 
image file: d5cp02294k-t5.tif(1)
where the interaction of the nuclear spins with the B field is
 
image file: d5cp02294k-t6.tif(2)
and the indirect and direct interactions between the spins are, respectively,
 
image file: d5cp02294k-t7.tif(3)
 
image file: d5cp02294k-t8.tif(4)

In eqn (2)–(4), Îi are dimensionless spin operators for nuclei i = 1, 2. The parameters γi and σi denote the gyromagnetic ratio and the nuclear magnetic shielding tensor of nucleus i. The tensors J12 and D12 describe the indirect and direct spin–spin coupling interactions between the two nuclei.

One can analyze the electric field E, the magnetic field B, and the spin operators Îi in the laboratory frame, while treating the interaction tensors – σi, J12, and D12 – in the molecular frame.11 The time dependence of the antisymmetric vectors image file: d5cp02294k-t9.tif and image file: d5cp02294k-t10.tif, as well as the D12 tensor is given by

 
image file: d5cp02294k-t11.tif(5)
 
image file: d5cp02294k-t12.tif(6)
 
D12(t) = R(tD12(0)·RT(t)(7)
where R(t) is the proper rotation matrix that transforms the tensor components from their initial molecular orientation to their orientation at time t; det(R) = +1. Notice that the components of the antisymmetric nuclear properties (image file: d5cp02294k-t13.tif and image file: d5cp02294k-t14.tif) transform as (pseudo)vectors, while the rotation of the symmetric components of the nuclear interaction tensors (i.e., D12) is governed by the formula appropriate to tensors of the second rank that are irreducible under three-dimensional rotations.12

The overall spin dynamics was computed by solving the master equation of the spin system

 
image file: d5cp02294k-t15.tif(8)
where ρ is the density matrix of the nuclear spin states, ı is the imaginary unit, and ħ = 1.054571817 × 10−34 J s is the reduced Planck's constant. The Lindbladian operator is expressed in the interaction frame as
 
image file: d5cp02294k-t16.tif(9)
where
 
image file: d5cp02294k-t17.tif(10)
and image file: d5cp02294k-t18.tif for i = B, J, D. The transformation in eqn (9) takes us to the so-called rotating frame – in terms of classical physics, it corresponds to the frame rotating for each spin with its precession frequency. This transformation allows one to avoid solving the first-order differential equation system with time-dependent coefficients, given in eqn (8). Moreover, from an experimental perspective, this transformation eliminates terms that oscillate much faster than the bandwidth of a typical NMR analogue-to-digital signal conversion system at a high magnetic field.

2. Materials and methods

2.1. Quantum chemistry computations

The equilibrium geometry of TFP was optimized at the density-functional theory (DFT) level using the PBE0 exchange–correlation functional,13 the DFT-D3(BJ) dispersion correction14,15 and the def2-QZVPP basis set16 on the ORCA 6.0.1 programme,17 using the integration grid 7 (in the ORCA terminology). The optimization was started from the geometry indicated in Table S3, and the resulting atomic coordinates are given in Table S4.

After the geometry optimization, a relaxed surface scan was performed using ORCA at a similar level of theory as described above, in which the HC(OH) dihedral angle was stepped between the values of 180 and −170 degrees in steps of 10 degrees (altogether 36 different values), and all the other structural parameters were relaxed for each value of the dihedral angle. The 3J(19F,1H) spin–spin coupling tensors for the three distinct fluorine centres (Fa, Fb and Fc) of the CF3 group, as well as the fluorine shielding tensors σ(19F) for the same nuclei, were calculated using the Turbomole programme18 using DFT, the PBE0 functional and integration grid 7 (in the Turbomole terminology), as well as the scalar-relativistic X2C level of theory19 and the x2c-QZVPPall-s basis set.20

When reporting non-scalar molecular properties, such as the NMR tensors or the electric dipole moment vector in this work, one has to specify the used molecule-fixed coordinate frame. This is particularly so when different values of the intramolecular coordinates are used, such as in the case where flexible molecular models are subjected to molecular dynamics simulations21 or, as in the present work, where we want to investigate the anisotropic molecular properties as a function of the HC(OH) dihedral angle, which undergoes large-amplitude motion in TFP. To this end, the calculated properties were transformed, for each relaxed geometry appropriate to the fixed HC(OH) dihedral angle, to the Eckart frame determined by the optimized equilibrium geometry (Table S4) of TFP, omitting the O(H) proton in the transformation. The method and software used in ref. 21 were employed.

2.2. Molecular dynamics

The structure of the lowest-energy (R)-conformer, obtained from DFT computations (Table S5), was used as the initial structure for the TFP molecules. The initial configuration for the MD simulations consisted of a 4.93 × 4.93 × 4.93 nm box containing 828 TFP molecules, prepared using the PACKMOL package.22 The molecular density used in the simulations corresponds to the experimental liquid density of TFP at room temperature, as reported in ref. 23, ρTFP = 1.26 g cm−3.

The energy of this molecular ensemble was minimized using GROMACS 2024.524–26 with the OPLS-AA force field,27,28 employing the parameters obtained from the LigParGen web server,29 which are similar to those described in ref. 30–33. Force field parameters used in our study are listed in Table S6 in the SI and give the permanent electric dipole moment of TFP that reasonably agrees with that obtained from quantum chemistry calculations (see Section 3.3 and Fig. 4 for details).

First, the energy was minimized over approximately 50[thin space (1/6-em)]000 steps, with an initial force tolerance of 12 kJ mol−1 nm−1 and a step size of 0.01 nm, followed by a second minimization step with a force tolerance of 10 kJ mol−1 nm−1 using the Verlet steepest descent algorithm. Subsequently, the system was equilibrated for 300 ps with the constant number of molecules, volume, and temperature (NVT) ensemble at T = 298 K, with periodic boundary conditions and temperature coupling set to 1 ps. The final equilibration step was performed with the constant number of molecules, pressure, and temperature (NPT) ensemble, where the pressure was maintained at approximately p = 1 bar, using the isotropic Parrinello–Rahman pressure coupling method with a coupling constant of 3.0 ps (fast as compared to the inverse of the largest used frequency, i.e., 100 ps), and the temperature was kept at T = 298.15 K by using the Nose–Hoover algorithm. The compressibility of the system was set to 1.332 GPa−1, corresponding to propan-2-ol, the most similar liquid for which experimental data were available.34

The molecular trajectory computations were performed in GROMACS 2021.2,35 where the external electric field was implemented following the procedure described in ref. 36. We performed the computations both without the electric field and for the amplitude of the electric field ranging from 10−3 to 10 V nm−1. The field frequency varied from 0.03 GHz to 10 GHz and oscillated with a cosine time dependence; see Table S7 for further details.

2.3. Spin dynamics

The computations of the changes in time of the spin states were performed using add-on Spin Dynamica 3.0.137 in the Mathematica 11 package.38 The effective Lindbladian operator was calculated analytically. The trajectories of the spin-state amplitudes were computed numerically based on the interpolated dependence of the averaged molecular parameters, image file: d5cp02294k-t19.tif and image file: d5cp02294k-t20.tif, on the frequency of the electric field, E(t) = E0[thin space (1/6-em)]cos(ωEt)êE with E0 = 1 V nm−1. The main magnetic field was assumed to be B0 = B0êZ with B0 = 11.75 T. The density matrix of the 19F–1H spin system in the thermodynamic equilibrium (T = 300 K) was used as the initial mixed spin state. In the studies of antisymmetric magnetic shielding, the pulse sequence consisted only of a single excitation by the electric field of the rectangular envelope, E(t) = E0[thin space (1/6-em)]cos(ωFt)êX, where ωF is the spin precession frequency in the field B0 (approx. 470 MHz). However, in the case of the antisymmetric indirect spin–spin coupling, first an RF-pulse was used to selectively invert the proton spin state (180° pulse), and then the rectangular pulse of the electric field was applied as E(t) = E0[thin space (1/6-em)]cos((ωHωF)t)êZ, where ωH is the spin precession frequency of the proton. The chirality-sensitive states were expressed in the case of the antisymmetric magnetic shielding using the Cartesian product operators, while for the antisymmetric spin–spin coupling, single-transition operators, described in detail in ref. 39 and 40, were used.

3. Results

3.1. Probability distribution of TFP rotamers

Chirality-sensitive NMER effects are determined by the permanent electric dipole moment of the TFP molecule, μe, and the antisymmetric nuclear properties (image file: d5cp02294k-t21.tif and image file: d5cp02294k-t22.tif) which, in turn, are dependent on the molecular conformation; so we started our studies by finding the probability distribution of TFP rotamers. Quantum chemistry calculations indicate that TFP has two low-energy conformers that differ in the orientation of the hydroxyl group. The variation of the molecular energy with this angle is given in Table S8 of the SI. In the lowest-energy conformer, the HC(OH) dihedral angle is −56.4°, while for the second-lowest-energy conformer, this angle is −176.0° (Fig. 1A). From DFT computations it follows that, at 300 K, the ratio of the number of molecules adopting the HC(OH) angle within ±30° of, on the one hand, the first and, on the other hand, the second lowest-in-energy conformers, is approximately 3[thin space (1/6-em)]:[thin space (1/6-em)]1. Molecular dynamics simulations reproduce this result well (Fig. 1B). The lowest-energy rotamer structures agree with those obtained in the infrared/Raman41,42 and microwave43 studies of TFP.
image file: d5cp02294k-f1.tif
Fig. 1 (A) Boltzmann probability distribution of the HC(OH) dihedral angle derived from DFT computations. The distance from the origin is proportional to the Boltzmann probability at temperature T = 300 K of the existence of the conformer with the indicated dihedral angle, with the energy obtained from the DFT calculations. (B) The variation of the dihedral HC(OH) angle (θ) distribution of TFP as a function of the electric-field strength obtained from MD simulations.

3.2. Dependence of image file: d5cp02294k-t23.tif and image file: d5cp02294k-t24.tif on the TFP conformation

Although one might suspect that the permanent electric dipole moment μe should be treated as a global property of the molecule, and the antisymmetric properties (image file: d5cp02294k-t25.tif and image file: d5cp02294k-t26.tif), are mainly determined by the local electronic environment of the nuclei, we have performed systematic calculations of the HC(OH) dihedral angle dependencies of all the three quantities. Such dependences were used further in computing time-variation of μe, image file: d5cp02294k-t27.tif and image file: d5cp02294k-t28.tif, based on the results of MD simulations.

The isotropic parts of the magnetic shielding for the CF3 fluorine nuclei, σiso(19Fi), where i = a, b, c, have roughly comparable magnitudes and they almost do not vary with the HC(OH) angle (Fig. 2). A similar relationship holds for antisymmetric components, image file: d5cp02294k-t29.tif, but the relationship of image file: d5cp02294k-t30.tif on the HC(OH) angle is more visible than for the case of the isotropic contributions. Their Cartesian components are given in Table S9 in the SI. In contrast, the J12 tensor components vary significantly between the fluorine nucleus involved, 19Fa, 19Fb, or 19Fc. The isotropic part of the indirect spin–spin coupling tensor between 19Fc and the CH proton, 3Jiso(19Fc,1H), is noticeably larger than the corresponding couplings 3Jiso(19Fa,1H) and 3Jiso(19Fb,1H). However, the opposite is observed for the antisymmetric component of the indirect spin–spin coupling: the coupling to 19Fc is negligible compared to image file: d5cp02294k-t31.tif and image file: d5cp02294k-t32.tif; see Table S10 in the SI for all their components.


image file: d5cp02294k-f2.tif
Fig. 2 The structure of the lowest-energy conformer of TFP obtained from DFT computations (in the middle of the figure). The Corey–Pauling–Koltun atom colouring was used: hydrogen – white, carbon – black, oxygen – red, fluorine – green. The bonds forming the HC(OH) dihedral angle, θ = −56.4°, are highlighted in blue. Angular dependence of the isotropic and antisymmetric components of the tensors (circular diagrams): fluorine magnetic shielding, σ(19F), and fluorine–proton indirect coupling, 3J(19F,1H) as functions of the HC(OH) dihedral angle. One radial division on the plots represents 200 ppm for σiso(19F), 4 ppm for image file: d5cp02294k-t33.tif, 10 Hz for 3Jiso(19F,1H), and 1 Hz for image file: d5cp02294k-t34.tif.

Although the isotropic and antisymmetric components of both the magnetic shielding tensor and the indirect 19F–1H coupling tensor vary with the rotation of the hydroxyl group, these variations are relatively small in the context of the present study. Instead, the quantity that is quite sensitive to the hydroxyl group rotation angle, is the permanent electric dipole moment of the TFP molecule, μe (Fig. 3).


image file: d5cp02294k-f3.tif
Fig. 3 Three-dimensional view and, in the yz-plane, the projection of the permanent electric dipole moment (μe, green arrow) and (A) the magnetic shielding image file: d5cp02294k-t35.tif for nuclei Fb and Fc and (B) the antisymmetric components of the indirect spin–spin coupling image file: d5cp02294k-t36.tif (orange arrows) for nuclei Fa and Fb. The numbers indicate the value of the HC(OH) dihedral angle. The vectors corresponding to the conformer closest to the most stable conformation are boldfaced, with HC(OH) approximately equal to −60°. We show in the blue colour the ellipsoid spanned by eigenvectors of lengths that are proportional to the absolute values of the eigenvalues of the tensor image file: d5cp02294k-t37.tif, where image file: d5cp02294k-t38.tif is the unity matrix.

3.3. Dependence of the averaged μeof TFP on the electric-field strength

To find the optimal electric-field strength, i.e., such that its effect on the reorientational dynamics of TFP exceeds the stochastic noise but is sufficiently low to make reasonably accurate predictions for the experimentally accessible electric-field strengths (approximately up to 0.01 V nm−1), we performed MD simulations of the partial orientation of the TFP molecules using the static electric field.

If the sample is partially oriented by an externally applied static electric field, E = E0êE, the total dipole moment averaged over the molecules follows the Langevin equation,

 
image file: d5cp02294k-t39.tif(11)
where kB = 1.380649 × 10−23 J K−1 is the Boltzmann constant and T is the temperature, the bracket 〈⋯〉mol means an average over molecules. For small electric fields, such that image file: d5cp02294k-t40.tif, the Langevin equation simplifies to
 
image file: d5cp02294k-t41.tif(12)

If the molecules were spherical, to describe the ordering effect of the electric field, it would be enough to calculate the average dipole electric moment from eqn (11). However, it is not evident to which extent this assumption is met by the TFP molecule due to its non-spherical shape and its internal motion. To more fully describe the field-ordering influence, the orientational probability tensor

 
P = 〈êi1êi2mol,(13)
should be considered. In eqn (13), the symbol ⊗ denotes the outer product, i.e., (êi1êi2)j1,j2 = (êi1)j1(êi2)j2. The unit vector êi of the xyz-molecular frame (with i = x, y, z) components (êi)j is expressed in the XYZ-laboratory frame (with j = X, Y, Z). We chose the molecular frame of the 1,1,1-trifluoropropan-2-ol molecule as follows. The êx vector is from the oxygen to the central carbon atom, the êy vector is perpendicular to the êx vector and in the plane spanned by the O–CH fragment of the molecule, and the êz vector is perpendicular to both the êx and êy vectors. With the static electric field strength 1 V nm−1, the eigenvalues of the tensor P derived from the MD simulations are 0.35, 0.35 and 0.30. Thus, they do not significantly deviate from the isotropic rotational diffusion for which one expects all eigenvalues to equal one third. Consequently, we can consider the influence of an electric field of strength up to 1 V nm−1 as a small perturbation of the rotational dynamics of TFP.

The dependence of the ensemble-averaged molecular-frame axis components, 〈êimol,j, on the strength of the electric field obtained from MD, is shown in Fig. 4 and agrees with eqn (11). Although the orientations of the electric dipole moment of the TFP molecule at the equilibrium geometry, as derived from quantum chemistry calculations and MD differ to some degree, by approximately 30° (see the inset in Fig. 4), to report results consistently for molecular quantities derived from the same kind of computations, we decided to use the μe derived from quantum chemistry. For the electric field strength lower than E0 = 0.1 V nm−1, the degree of orientation was smaller than the amplitude of the fluctuations. On the other hand, a larger electric-field strength, i.e., E0 > 10 V nm−1, eventually causes saturation of the electric polarization of the sample. Moreover, at such high electric fields, most TFP molecules adopt an HC(OH) angle close to +60°, which is not the most stable conformation at lower field strengths (see Fig. 1B and Fig. S1). Therefore, for further studies, we decided to adopt a constant strength of the electric field equal to E0 = 1 V nm−1 (i.e., image file: d5cp02294k-t42.tif), which still corresponds to a linear dependence of the averaged moment 〈μemol on the electric field strength, i.e., it follows eqn (12); see Fig. S2 for the linear approximation compared the components 〈êimol,j obtained from MD simulations.


image file: d5cp02294k-f4.tif
Fig. 4 The components of the ensemble-averaged unit-vector components (〈êimol,j) of the TFP molecular frame vectors (〈êimol) when subjected to a static electric field E0 aligned along the Z-axis of the laboratory frame that were obtained from MD (points). The solid lines represent the best fits to the Langevin function, αL(coth(βLγLE0) − (βLγLE0)−1), with parameters: αL,X = −0.375(28) for 〈êxmol,Z, αL,Y = −0.525(38) for 〈êymol,Z, and αL,Z = 0.044(6) for 〈êzmol,Z. The used value of the parameter βL = 0.811 D−1 V−1 nm corresponds to the temperature T = 300 K. The parameter γL = 2.37(22) D was fitted globally to all data sets above. The left inset depicts the TFP molecule with the permanent electric dipole moment used in molecular dynamics (μeMD) and that obtained from quantum chemistry calculations (μeQC).

3.4. Dependence of the averaged μeof TFP on electric-field frequency

Next, we used an electric field of the optimal strength, i.e., 1 V nm−1, which was found in Section 3.3, and studied the average response of μe of the TFP molecules as a function of the frequency of such a field, utilizing the conformational dependence determined in Section 3.2.

If the electric field is oscillating in time with the frequency ωE,

 
E(t) = E0[thin space (1/6-em)]cos(ωEt)êE,(14)
then the electric dipole moment does not follow the electric field instantaneously. Instead, the dipole moment oscillations are retarded compared to the electric field. Consequently, in the Debye model, the dipole averaged over molecules, 〈μemol, becomes44
 
image file: d5cp02294k-t43.tif(15)
where τ is the diffusion reorientation time. For times sufficient to partially orient the electric field (approximately 1 ns, see Fig. 5), eqn (15) reproduces well the trends observed in MD studies. The rotational diffusion reorientation time of TFP obtained from the MD studies is τ = 94(2) ps.


image file: d5cp02294k-f5.tif
Fig. 5 The evolution of the laboratory-frame components of the averaged electric dipole moment of TFP molecules, 〈μemol, under an external electric field of 1 V nm−1 along the Z-axis: (A) response to a static field E and (B) oscillating field with the frequency f = 0.03 GHz. The points are data obtained from MD, and the lines are the best fits for exponential decay and cosine function, respectively.

Applying the Fourier transform to eqn (15) one obtains the frequency dependence of 〈μemol in a more compact form,

 
image file: d5cp02294k-t44.tif(16)

The dependencies of the amplitude and phase of the Z-component of the averaged dipole moment, 〈μeZmol, on the electric field frequency are shown in Fig. 6A, while the plot of the real and imaginary parts of the moment 〈μeZmol is in Fig. 6B – the green curve.


image file: d5cp02294k-f6.tif
Fig. 6 (A) Dependence of the amplitude and phase shift of the Z-component of the averaged dipole moment, 〈μeZmol, on the electric-field frequency. (B) Dependence of the averaged antisymmetric magnetic shielding, image file: d5cp02294k-t45.tif, and indirect spin–spin coupling, image file: d5cp02294k-t46.tif, involving fluorine nuclei 19Fi (i = a, b, c) in the complex plane on the electric-field frequency. In panel (B), the amplitude of a given nuclear property corresponds to the distance from the origin, while the phase shift is the angle relative to the positive real axis.

3.5. Dependence of the averaged image file: d5cp02294k-t47.tif and image file: d5cp02294k-t48.tif on electric-field frequency

To find out how the antisymmetric image file: d5cp02294k-t49.tif and image file: d5cp02294k-t50.tif depend on the electric-field frequency, we combined the results presented in Section 3.2 with the MD simulations at optimal E-field strength (see Section 3.3) at various frequencies of the E-field. Similarly to the dipole moment μe, the antisymmetric image file: d5cp02294k-t51.tif and image file: d5cp02294k-t52.tif are vector quantities placed in the same molecular frame. Hence, we assumed that they depend on the electric-field frequency analogously as found for the averaged μe of TFP. The results given in Section 3.5 provide frequency dependencies used further as input data in the spin dynamics computations described in Section 3.6.

The dependence of the antisymmetric magnetic shielding averaged over molecules, image file: d5cp02294k-t53.tif, on the frequency of electric field can be found by projecting the vector image file: d5cp02294k-t54.tif onto the vector μe. The electric field orients the permanent electric dipole moment of the molecule, so the scalar projection image file: d5cp02294k-t55.tif, is along the unit vector êE. Therefore, taking into account that the averaging over the molecules introduces a factor image file: d5cp02294k-t56.tif, one finds that

 
image file: d5cp02294k-t57.tif(17)

The analogous reasoning applies to the indirect spin–spin coupling image file: d5cp02294k-t58.tif and it gives

 
image file: d5cp02294k-t59.tif(18)

The frequency dependencies of the molecular parameters image file: d5cp02294k-t60.tif and image file: d5cp02294k-t61.tif for each fluorine nuclei are shown in Fig. 6B.

In general, eqn (17) and (18) reproduce the dependencies of the Z-components of antisymmetric vectors image file: d5cp02294k-t62.tif and image file: d5cp02294k-t63.tif on the frequency of the electric field ωE. However, these properties do not vanish at high electric field frequencies, as expected based on eqn (17) and (18). Such a discrepancy is an artefact of the MD simulation rather than a result of some major deviations from the assumptions of the Debye model (e.g., the absence of intermolecular interactions). Although one has to admit that such assumptions are not fully satisfied in the case of TFP, since the dihedral angle HC(OH) varies by approximately 15° with electric field oscillations (Fig. S3). Consequently, the vector μe, onto which the antisymmetric vectors image file: d5cp02294k-t64.tif and image file: d5cp02294k-t65.tif are projected, follows the electric field in a nontrivial way. It should also be noted that in the case of molecular dynamics simulations that use the electric field directly,45 obtaining accurate results is a challenge compared to the approach using the fluctuation–dissipation theorem.46,47 Moreover, the localized charge of the CF3 group, which significantly contributes to the dielectric dynamics of TFP, is a source of further uncertainties in the performed MD simulations, possibly requiring the refinement of force field parameters.

3.6. Dependence of chirality-sensitive observables on electric-field frequency

To find the frequency dependence of the amplitudes and phases of the chirality-sensitive observables, we used the formulae image file: d5cp02294k-t66.tif and image file: d5cp02294k-t67.tif given in eqn (17) and (18). The results were used as parameters in the spin dynamics computations of the 1H–19F system, described in the Introduction.

In the fast exchange regime, one observes an average over the fluorine nuclei,

 
image file: d5cp02294k-t68.tif(19)
 
image file: d5cp02294k-t69.tif(20)

According to Fig. 6B, the contributions of the nuclei Fa and Fc to image file: d5cp02294k-t70.tif partially cancel out with the contribution of the nucleus Fb. For image file: d5cp02294k-t71.tif, the contributions Fa and Fb partially average out with each other. The resulting frequency dependencies of image file: d5cp02294k-t72.tif and image file: d5cp02294k-t73.tif are shown in Fig. 7.


image file: d5cp02294k-f7.tif
Fig. 7 Time dependence of selected quantum state amplitudes under an oscillating electric field applied at the frequency corresponding to the difference between the proton and fluorine spin precession frequencies (A), and the fluorine spin precession frequency (B), B0 = 11.75 T, i.e., 1 ppm corresponds to approximately 470 Hz, compared with the variation of the corresponding antisymmetric indirect spin–spin coupling, image file: d5cp02294k-t74.tif and magnetic shielding, image file: d5cp02294k-t75.tif. These values were averaged over the three fluorine nuclei. The amplitudes of 〈Î+2,3〉 and 〈Î1+〉 after application of the 180° pulse on the proton followed by the electric-field pulse E0[thin space (1/6-em)]cos((ωHωF)t)êZ (C) and the electric-field pulse E0[thin space (1/6-em)]cos(ωFt)êX (D), respectively. Both coherence amplitudes were normalized dividing by a factor image file: d5cp02294k-t76.tif that corresponds to the amplitude of the coherence 〈Î1x〉 obtained by exciting the sample at thermodynamic equilibrium at T = 300 K by using a 90° radiofrequency pulse.

If the first nucleus is fluorine and the second the proton, the application of the electric field oscillating at the frequency equal to the difference between the proton and fluorine spin precession frequencies, ωHωF, results in the generation of chirality-sensitive spin states given by a single-transition operator Î+2,3 = ÎX2,3 + ıûY2,3 that are dependent on the averaged antisymmetry image file: d5cp02294k-t77.tif.48 The indices of single-transition operators denote the state |↑FH〉 for “2” and the state |↓FH〉 for “3”. In terms of the Cartesian product operators, one finds that Re(Î+2,3) = Î1XÎ2X + Î1YÎ2Y and Im(Î+2,3) = Î1XÎ2YÎ1YÎ2X. The comparison, shown in Fig. 7C, between amplitude-phase dependence of image file: d5cp02294k-t78.tif on the frequency of the electric field (the red curve) and the computed amplitudes of the states (open squares) indicates the frequency dependence of the amplitude of the induced chirality-sensitive spin state by the electric field E

 
image file: d5cp02294k-t79.tif(21)
where the image file: d5cp02294k-t80.tif and image file: d5cp02294k-t81.tif constants account for averaging over the equivalent fluorine nuclei.

Application of the electric field oscillating at the fluorine spin precession frequency, ωF, yields chirality-sensitive spin states Î1X and Î1Y whose amplitudes are dependent on the averaged antisymmetry image file: d5cp02294k-t82.tif shown in Fig. 6B. From the point of view of the spin dynamics studies, an electric field of strength 1 V nm−1 is so strong that it could potentially saturate nuclear magnetisation, which is unrealistic, taking into account the experimentally available electric-field strengths (in practice, lower than several kV mm−1 due to electric breakdown). Therefore, we took advantage of the linear dependence, for E0 < 1 V nm−1, of the liquid response on the field strength and scaled down the image file: d5cp02294k-t83.tif dependence by two orders of magnitude. Analogously to the case of the antisymmetric spin–spin coupling, comparing the amplitude-phase dependence of image file: d5cp02294k-t84.tif on the frequency of the electric field (the blue curve) with the computed amplitudes of the states (open circles), one can see a close correspondence up to a multiplication by a constant (Fig. 7D). Therefore, the frequency dependence of the amplitude of the chirality-sensitive spin state Î1+ = Î1X + ıû1Y is

 
image file: d5cp02294k-t85.tif(22)
with constants image file: d5cp02294k-t86.tif and image file: d5cp02294k-t87.tif. If the discrepancies between the Debye model and the MD simulation results were to be attributed to errors that accumulate during the averaging of fluorine nuclei, then the MD data should be shifted so that image file: d5cp02294k-t88.tif.

4. Discussion and conclusions

The absolute configuration of the molecule primarily governs the phase of the chirality-sensitive NMER signal. The quantities that are influenced by the configuration of the molecule are the pseudovectors of image file: d5cp02294k-t89.tif and image file: d5cp02294k-t90.tif, whose directions are reversed under the transformation of one enantiomer into the other. The largest effects are those where the magnitude of the pseudovector is large. However, the frequency of the externally applied electric field also plays a significant role in modulating this phase. At sufficiently high frequencies, the signal phase may deviate considerably from its limiting behaviour observed at low frequencies, where the electric field varies slowly compared to the timescale of molecular rotational diffusion (Fig. 7A and B).

This frequency-dependent behaviour appears to be an intrinsic feature of chirality-sensitive measurements performed at high magnetic fields. Due to the inherently small magnitude of NMER signals induced by time-dependent electric fields, enhancing the initial magnetization through electric field application becomes a practical necessity for detection. Consequently, the observed signal inevitably reflects the dynamic interplay between molecular motion and the electric-field frequency.

Although the frequency dependence may seem complex, it can be quantitatively described by considering how the key parameters of the NMER effect – namely, the antisymmetric component of the nuclear magnetic shielding tensor and the antisymmetric part of the indirect spin–spin coupling – respond to the applied field frequency. Both quantities, which are vector-like, exhibit frequency dependencies analogous to that of the permanent electric dipole moment of the molecule, see eqn (16)vs.eqn (17) and (18).

By selecting suitable observables – such as the amplitude of the raising operator 〈Î1+〉 in the case of antisymmetric shielding, and the amplitude of the single-transition operator 〈Î1+2,3〉 in the case of antisymmetric spin–spin coupling – one can establish a direct correspondence between the frequency-dependent retardation of the dipole moment and the resulting phase shift in the chirality-sensitive NMER signal (eqn (21) and (22); Fig. 7C and D). This approach is particularly relevant in the radiofrequency range typically employed in NMR experiments, extending up to slightly above 1 GHz. According to the obtained results in our study, the frequency at which the signal phase associated with a specific TFP enantiomer may become reversed is approximately 2 GHz, and it corresponds to the highest point of the red curve in Fig. 7C for image file: d5cp02294k-t91.tif-dependent NMER effect and the lowest point of the blue curve in Fig. 7D for the image file: d5cp02294k-t92.tif-dependent NMER effect. Moreover, the analysis of small, rigid molecules reveals that a tempting assumption – that such molecules reorient rapidly enough for their dipole moments to follow the electric field instantaneously – is not universally valid. This condition may not be satisfied even for molecules with molecular masses around 100 g mol−1.

The impact of the present analysis on both experimental design and expected observables depends critically on the frequency of the applied electric field, which is set chiefly by the spin system considered. For experiments with quasi-static or slowly varying fields (ref. 48), the applicability is limited. Conversely, in W-band EPR (∼100 GHz; ref. 11), the outcome is governed by the extent to which the dipole moment can follow the field.

Rather than a limitation, this sensitivity to reorientational dynamics can be viewed as a valuable source of information. Specifically, the frequency dependence of the phase shift in the NMER signal offers a novel means of probing the rotational mobility of chiral molecules in solution. Furthermore, the detection of chirality-sensitive effects mediated by antisymmetric spin–spin coupling in solution provides a compelling alternative to solid-state approaches, where such effects manifest as subtle perturbations of the spectral lineshapes via the parameter image file: d5cp02294k-t93.tif – a measurement that is experimentally very demanding.49,50

These results underscore the value of integrating quantum chemistry, molecular dynamics and spin dynamics to capture chirality-sensitive nuclear interactions under realistic conditions. By explicitly accounting for finite molecular reorientation times, this framework offers a refined understanding of NMER effects and points toward new spectroscopic strategies for studying molecular motion and chirality in solution.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the SI: (1) Three Wolfram Mathematica 11 files used for computing the amplitudes of the states 〈Î1+〉 (anti-s(19F)_TFP.nb) and 〈Î+2,3〉 (anti-3J(19F,1H)_TFP.nb), and processing data from molecular dynamics (Supplementary_Information.nb); (2) thirteen Mathematica packages used as auxiliary functions described in detail in Table S11; (3) ten text files with molecular parameters (nuclear interaction tensors and permanent electric dipole moment of TFP) described in detail in Table S12, (4) the Origin 9.60 file containing the MD ensemble-averaged components of the TFP molecular frame vectors and the best fits of Langevin curves to them. The Mathematica files can be accessed using the free Wolfram Player software, available for download at https://www.wolfram.com. See DOI: https://doi.org/10.1039/d5cp02294k.

Source data for this article, including results of molecular dynamics simulations performed in Gromacs are available at RepOD at https://doi.org/10.58132/CQYOLA.

Acknowledgements

P. G. would like to acknowledge the European Research Council for financial support through the ERC Starting Grant (project acronym: NMER, agreement ID: 101040164). Molecular dynamics simulations were performed using resources provided by Wroclaw Centre for Networking and Supercomputing (https://wcss.pl). J. V. acknowledges funding from the Research Council of Finland (grant 361326) and the University of Oulu (Kvantum Institute). Quantum chemistry computations were carried out using the resources of CSC-IT Center for Science (Espoo, Finland).

References

  1. P. Garbacz, L. D. Barron and P. Fischer, Fundamentals of Chiral Spectroscopy, in Physical Principles of Chirality in NMR, ed. P. Garbacz, Royal Society of Chemistry, 2024, pp. 33–58 Search PubMed.
  2. P. Garbacz, Antisymmetric Nuclear Magnetic Shielding as an Indicator of Molecular Chirality, in Physical Principles of Chirality in NMR, ed. P. Garbacz, Royal Society of Chemistry, 2024, pp. 158–202 Search PubMed.
  3. P. Garbacz and J. Vaara, Chirality-sensitive Effects Induced by Antisymmetric Spin–Spin Coupling, in Physical Principles of Chirality in NMR, ed. P. Garbacz, Royal Society of Chemistry, 2024, pp. 231–234 Search PubMed.
  4. A. D. Buckingham, J. Chem. Phys., 2014, 140, 011103 CrossRef PubMed.
  5. A. P. Gregory, R. N. Clarke and M. G. Cox, Meas. Sci. Technol., 2009, 20, 075106 CrossRef.
  6. A. P. Gregory and R. N. Clarke, Tables of the complex permittivity of dielectric reference liquids at frequencies up to 5 GHz, NPL Report MAT 23, National Physical Laboratory, 2012, p. 74.
  7. V. Gorini, A. Kossakowski and E. C. G. Sudarshan, J. Math. Phys., 1976, 17, 821–825 CrossRef.
  8. G. Lindblad, Commun. Math. Phys., 1976, 48, 119–130 CrossRef.
  9. C. Bengs and M. H. Levitt, J. Magn. Reson., 2020, 310, 106645 CrossRef CAS PubMed.
  10. J. Kowalewski and L. Mäler, Nuclear spin relaxation in liquids: Theory, experiments, and applications, CRC Press, 2nd edn, 2006, p. 32 Search PubMed.
  11. P. Garbacz and J. Vaara, Chem. Commun., 2021, 57, 8264–8267 RSC.
  12. D. A. Varshalovich, A. N. Moskalev and V. K. Khersonskii, Quantum Theory of Angular Momentum, World Scientific, 1988, p. 29 Search PubMed.
  13. C. Adamo and V. Barone, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
  14. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  15. S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
  16. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297 RSC.
  17. F. Neese, The ORCA program system, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 73–78 CAS.
  18. TURBOMOLE V7.8 2024, a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989–2024, TURBOMOLE GmbH, since 2007, available from https://www.turbomole.com.
  19. Y. J. Franzke, J. Chem. Theory Comput., 2023, 19, 2010–2028 CrossRef CAS PubMed.
  20. Y. J. Franzke, L. Spiske, P. Pollak and F. Weigend, J. Chem. Theory Comput., 2020, 16, 5658–5674 CrossRef CAS PubMed.
  21. T. S. Pennanen, J. Vaara, P. Lantto, A. J. Sillanpää, K. Laasonen and J. Jokisaari, J. Am. Chem. Soc., 2004, 126, 11093–11102 CrossRef CAS PubMed.
  22. L. Martínez, R. Andrade, E. G. Birgin and J. M. Martínez, J. Comput. Chem., 2009, 30, 2157–2164 CrossRef PubMed.
  23. M. Denda, H. Touhara and K. Nakanishi, J. Chem. Thermodyn., 1987, 19, 539–542 CrossRef CAS.
  24. B. Hess, E. Lindahl; contributions from: M. Abraham, A. Alekseenko, V. Basov, C. Bergh, E. Briand, A. Brown, M. Doijade, G. Fiorin, S. Fleischmann, S. Gorelov, G. Gouaillardet, A. Gray, M. E. Irrgang, F. Jalalypour, J. Jordan, C. Kutzner, J. A. Lemkul, M. Lundborg, P. Merz, V. Miletic, D. Morozov, J. Nabet, S. Pall, A. Pasquadibisceglie, M. Pellegrino, H. Santuz, R. Schulz, T. Shugaeva, A. Shvetsov, P. Turner, A. Villa, S. Wingbermuehle, GROMACS 2024.5, Department of Biophysical Chemistry, University of Groningen (1991–2000), Royal Institute of Technology and Uppsala University (2001–2025), https://www.gromacs.org.
  25. M. J. Abraham, T. Murtola, R. Schulz, S. Páll, J. C. Smith, B. Hess and E. Lindahl, SoftwareX, 2015, 1–2, 19–25 CrossRef.
  26. S. Páll, M. J. Abraham, C. Kutzner, B. Hess and E. Lindahl, Tackling exascale software challenges in molecular dynamics simulations with GROMACS, in Solving Software Challenges for Exascale, ed. S. Markidis and E. Laure, Springer International Publishing, London, 2015, pp. 3–27 Search PubMed.
  27. W. L. Jorgensen and J. Tirado-Rives, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 6665–6670 CrossRef CAS PubMed.
  28. L. S. Dodda, J. Z. Vilseck, J. Tirado-Rives and W. L. Jorgensen, J. Phys. Chem. B, 2017, 121, 3864–3870 CrossRef CAS PubMed.
  29. L. S. Dodda, I. Cabeza de Vaca, J. Tirado-Rives and W. L. Jorgensen, Nucleic Acids Res., 2017, 45, W331–W336 CrossRef CAS PubMed.
  30. M. Fioroni, K. Burger and D. Roccatano, J. Chem. Phys., 2003, 119, 7289–7296 CrossRef CAS.
  31. M. Fioroni, K. Burger, A. E. Mark and D. Roccatano, J. Phys. Chem. B, 2003, 107, 4855–4861 CrossRef CAS.
  32. W. Zygmunt and J. J. Potoff, Fluid Phase Equilib., 2016, 407, 314–321 CrossRef CAS.
  33. R. Chitra and P. E. Smith, J. Chem. Phys., 2001, 115, 5521–5530 CrossRef CAS.
  34. W. Haynes, CRC Handbook of Chemistry and Physics, CRC Press, 2014 Search PubMed.
  35. M. Abraham, B. Hess, D. van der Spoel, and E. Lindahl; contributions from: E. Apol, R. Apostolov, P. Bauer, H. J. C. Berendsen, P. Bjelkmar, Ch. Blau, V. Bolnykh, K. Boyd, A. van Buuren, R. van Drunen, A. Feenstra, G. Groenhof, A. Hamuraru, V. Hindriksen, M. E. Irrgang, A. Iupinov, Ch. Junghans, J. Jordan, D. Karkoulis, P. Kasson, J. Kraus, C. Kutzner, P. Larsson, J. A. Lemkul, V. Lindahl, M. Lundborg, E. Marklund, P. Merz, P. Meulenhoff, T. Murtola, S. Páll, S. Pronk, R. Schulz, M. Shirts, A. Shvetsov, A. Sijbers, P. Tieleman, T. Virolainen, Ch. Wennberg, M. Wolf and A. Zhmurov, GROMACS 2021.2, Department of Biophysical Chemistry, University of Groningen (1991–2000), Royal Institute of Technology and Uppsala University (2001–2021), https://www.gromacs.org.
  36. C. Caleman and D. van der Spoel, Angew. Chem., Int. Ed., 2008, 47, 1417–1420 CrossRef CAS PubMed.
  37. C. Bengs and M. H. Levitt, Magn. Reson. Chem., 2018, 56, 374–414 CrossRef CAS PubMed.
  38. S. Wolfram, Mathematica, Version 11.0, Wolfram Research, Inc., Champaign, IL, 2016 Search PubMed.
  39. A. Wokaun and R. R. Ernst, J. Chem. Phys., 1977, 67, 1752 CrossRef CAS.
  40. S. Vega, J. Chem. Phys., 1978, 68, 5518 CrossRef CAS.
  41. J. Murto, A. Kivinen, K. Edelmann and E. Hassinen, Spectrochim. Acta, Part A, 1975, 31A, 479–493 CrossRef CAS.
  42. H. Schaal, T. Häber and M. A. Suhm, J. Phys. Chem. A, 2000, 104, 265–274 CrossRef CAS.
  43. H. Møllendal, J. Phys. Chem. A, 2005, 109, 9488–9493 CrossRef PubMed.
  44. P. Debye, Polar Molecules, Dover Publications, 1929, pp. 89–95 Search PubMed.
  45. M. Woodcox, A. Mahata, A. Hagerstrom, A. Stelson, C. Muzny, R. Sundararaman and K. Schwarz, Chem. Phys., 2023, 158, 124122 CAS.
  46. L. Saiz, E. Guàrdia and J. Padró, J. Chem. Phys., 2000, 113, 2814–2822 CrossRef CAS.
  47. P. Mazur and S. De Groot, Non-Equilibrium Thermodynamics, Dover, 1984, pp. 150–156 Search PubMed.
  48. P. Garbacz, Chem. Phys., 2025, 162, 194304 CAS.
  49. E. R. Andrew and L. F. Farnell, Mol. Phys., 1968, 15, 157–165 CrossRef.
  50. K. J. Harris, D. L. Bryce and R. E. Wasylishen, Can. J. Chem., 2009, 87, 1338–1351 CrossRef CAS.

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.