Skip to Main Content
Skip Nav Destination
This chapter is subject to a Creative Commons CC-BY-NC-ND 4.0 International license.

The chapter covers various topics related to nuclear magnetoelectric resonance (NMER) phenomena capable of direct chirality-sensing. First, the mathematical foundations of antisymmetric properties of a spin system, such as the irreducible spherical decomposition of a tensor, are briefly discussed. Then, the Hamiltonian of a spin system perturbed by an electric field is introduced, and details of three chirality-sensitive NMER effects, including their significance, estimation of the magnitude, and proposed experimental protocols are described. The last part of the chapter is devoted to various experimental methods for achieving the most favourable electromagnetic field distribution for the proposed NMER experiments. These methods encompass devices ranging from shape-based designed electromagnetic cavities and loop-gap resonators to discrete-element systems, exemplified by inductor–capacitor circuits and systems with controllable surface current distribution.

Funding Group:
  • Award Group:
    • Funder(s):
       European Research Council
    • Award Id(s):
      101040164
  • Funding Statement(s):
     Financial support from the European Research Council through an ERC Starting Grant (project acronym: NMER, agreement ID: 101040164) is acknowledged.

Nuclear magnetic resonance (NMR) spectroscopy investigates the interactions of nuclear dipole moments in the presence, at least for the typical studies on the structure and physicochemical properties of molecules, of a strong magnetic field with a magnitude of several teslas. This experimental approach has been successfully applied, over several years, in studies aiming to determine the atomic connectivity and three-dimensional structures of molecules in the liquid phase, which allow us, e.g., to deduce biochemical properties in an environment similar to in vivo conditions.1  However, more subtle structural properties, such as molecular chirality, may be inferred from the results of NMR spectroscopy only in an indirect way, specific to a given group of chemical compounds.2  The range of information obtained from NMR studies can be significantly extended by allowing the presence of an external perturbation in the system under consideration, particularly that induced by an electric field. In this case, according to theoretical predictions discussed further in detail in this chapter, the limitation of NMR spectroscopy in studies of molecular chirality is overcome, and the postulated effects provide a novel method to determine the molecular structural parameters. After a brief introduction to the mathematical formalism used in this work, each of these effects is described in detail incorporating the proposed experimental protocol for the observation of the effect, estimation of the magnitude of the anticipated signal, and discussing the significance of the effect for the determination of molecular structure.

Tensor analysis is frequently referred to in the following discussions. Therefore, let us briefly summarize the relevant tensor properties from the viewpoint of this work. Further details of tensor properties in the context of nuclear magnetic shielding tensor and molecular chirality are given in Section 7.3. In the following text, variables (in italics) that are tensors of rank at least one (e.g., vectors) are boldfaced in contrast to those variables that are plain numbers.

A two-index tensor (a bilinear map) T , such that T : V × V , can be decomposed into a sum,
(7.1)
where T anti ( A , B ) = 1 2 [ T ( A , B ) T ( B , A ) ] is the antisymmetric part and T sym ( A , B ) = 1 2 [ T ( A , B ) + T ( B , A ) ] is the symmetric part of the tensor T . Introducing a scalar product permits us to separate the isotropic part, T iso ( A , B ) = 1 3 A B , and the traceless but symmetric part, T sym T iso .
It is convenient to express the tensor T coordinates in a specific basis, taken here as orthonormal (eα · eβ is 1 if α = β and 0 otherwise), e.g., in the three-dimensional Cartesian coordinate system, V 3 (α, β and γ are the x-axis, y-axis, and z-axis, respectively),
(7.2)
In such a case, explicit forms of the isotropic and antisymmetric parts of the matrix [ T ] are
(7.3a)
(7.3b)
Eqn (7.3a) follows the common convention that uses the same designation for a number [ T iso = 1 3 ( T x x + T y y + T z z ) , unbolded] as for the 2-index tensor ( T iso , bolded) from which it was derived; in eqn (7.3b) for brevity T ( e α , e β ) = T α β . The identity matrix in eqn (7.3a) will be denoted in the further text by the symbol 13.
As in the case of the isotropic part, where it is convenient to use a number instead of a tensor, in the case of the antisymmetric part, it is convenient to describe it as a vector-like quantity T * ,3 
(7.4)
However, the drawback of this assignment is that the coordinates of T * do not transform under the basis change in the same way as vectors (vide infra).

A two-index tensor decomposes to a sum of irreducible components of, at most, the second rank under rotations in three-dimensional space.4  The ranks of the isotropic, antisymmetric, and traceless symmetric parts of tensor T are zero, one, and two, respectively. The total number of tensor elements equals the sum of the numbers of tensor elements into which it has been decomposed under rotations in three-dimensional space, e.g., T iso has one element, T anti has three, and T sym has five, giving a total of nine elements. Thus, the decomposition in isotropic, antisymmetric, and traceless symmetric parts becomes ambiguous for tensors of ranks higher than two. For instance, a reducible third-rank tensor has one zeroth rank component (proportional to the so-called Levi-Civita tensor), three first-rank components, two second-rank components, and one third-rank component.

Usually only the isotropic parts, mainly in liquid-state NMR, and the symmetric parts of the relevant tensors, mainly in solid-state NMR, are experimentally studied.7  However, as will be shown, the presence of the antisymmetric component induces interesting phenomena in chiral molecules.

Tensors provide a convenient description of anisotropic interactions, i.e., those whose amplitude depends on the orientation of the interacting constituents. Although tensors are abstract objects, one can explicitly express them in a given frame of reference. For instance, one can write the nuclear magnetic shielding tensor, denoted as σ, as a 3 by 3 element matrix (see eqn (7.2); the square brackets of [σ] will be omitted in the following text if they do not introduce ambiguity),
(7.5)
The tensor σ is dimensionless (commonly given in parts per million; 1 ppm = 10−6) and describes the interaction between the nuclear dipole magnetic moment μm (J T−1), and the magnetic field B (i.e., magnetic flux density, in T). The energy of the nuclear magnetic moment placed in the magnetic field is
(7.6)
Eqn (7.6) generalizes the energy of a magnetic dipole placed in a magnetic field, −μm · B, to the anisotropic case; the tensor σ describes the extent of this interaction’s anisotropy. The minus sign before the magnetic shielding tensor σ follows a commonly used convention in the NMR community (the elements of the tensor σ can be both positive and negative).

The hat above the Hamiltonian symbol, H ˆ , emphasize that the Hamiltonian is an operator, e.g., the way it acts on the quantum state of nuclear spins allows us to determine how the spin state changes in time. Thus, strictly speaking, one can view eqn (7.6) as the interaction of a magnetic dipole known from classical physics (the resultant of the multiplication is a number), but since the form of the equation is the same as in its quantum-mechanical counterpart, classical and quantum interpretations will be considered as equivalent in the further text – see Section 7.8 for details.

The energy of the interaction between a mole of 1H magnetic moments ( μ H m = 1.41 × 10 26 J T 1 ) in a water molecule (shielding ∼3 × 10−5) and a magnetic field of strength B = 11.75 T, is approximately 3 mJ, which is nearly six orders of magnitude smaller than the thermal energy at T = 300 K.

From eqn (7.6), it follows that the components of the effective magnetic field, (13 − σ) · B, interacting with the magnetic moment μm may not only have a different amplitude compared to the magnetic field B, which occurs via the diagonal elements of the tensor σ, but also experience changes in direction due to off-diagonal elements. This observation may be described more rigorously with the aid of the irreducible tensor decomposition mentioned in Section 7.2. Mathematical details of the decomposition are given in ref. 4. Here, we limit ourselves to the application of the irreducible decomposition of a two-index tensor, an example of which is the nuclear magnetic shielding tensor σ.

To deepen our understanding of tensor antisymmetry, let us look at how its coordinates change under the influence of rotations in three-dimensional space. If R is the rotation matrix, then the components of the two-index tensor transform T under a basis change according to the equation R [ T ] R T . The symbol RT represents the transposition (an interchange of the rows and columns) of the matrix R.

The irreducible decomposition described in Section 7.2 is based on the different transformation properties of tensors under rotations in three-dimensional space. In particular, scalars (zeroth-rank tensors, e.g., energy) remain invariant under rotations. An example of a matrix that transforms under rotations as a scalar is the identity matrix, since the rotation matrix satisfies R · 13 · RT = 13. Vectors (first-rank tensors, e.g., electric field) after rotation are R [ T 1 ] , where T 1 is the first-rank tensor and R is the proper rotation matrix.§ Moreover, the matrix
(7.7a)
transforms with respect to the basis orientation as a vector.
(7.7b)
where the subscript k means the k-th coordinate of the vector and ϵ i j k is the Levi-Civita tensor (in the right-handed Cartesian coordinate system) defined by
(7.8)
since one can directly check that
(7.9)
It is convenient to refer to the quantity σ* as the antisymmetric part of the nuclear magnetic shielding tensor, which is represented either as a matrix (eqn (7.7a)) or a vector (eqn (7.7b)) since its identity relies on its transformation properties rather than on how its components are listed. The correspondence between a 2-form (e.g., the σ* tensor) and a pseudovector is based on the application of the so-called star Hodge operator; mathematical details are described in a very accessible way in ref. 8. It is also the source of motivation for marking the antisymmetric part of nuclear magnetic shielding tensor with a star.

The nuclear site symmetry determines the form of the nuclear magnetic shielding tensor σ, typically resulting in particular components of the tensor becoming zero. For example, highly symmetric nuclear sites such as Td (tetrahedral) and O (octahedral) have only one independent component of σ; in this case, σiso = σxx = σyy = σzz holds true, so such shielding tensors do not have antisymmetric parts. However, less symmetric nuclear environments do not exclude the antisymmetry of the tensor σ. A few examples of the magnetic shielding tensors of nuclei in simple planar molecules are listed in Table 7.1.

Table 7.1

Nuclear magnetic shielding tensors of water, formide (both molecules in the yz-plane), and formic acid (the molecule is in the xy-plane) computed in the CFOUR program using the coupled cluster singles and doubles (CCSD) method in the aug-pCVTZ basis set; the initial molecular geometries were taken from Computational Chemistry Comparison and Benchmark Data Base and further optimized using CCSD/aug-pCVTZ. Tensor elements that contain antisymmetric parts are boldfaced.

HaOHb HaHbCO cis-HaCOaObHb trans-HaCOaObHb
1Ha   ( 23.56 0.00 0.00 0.00 38.92 9 . 61 0.00 7 . 94 30.47 )   ( 21.91 0.00 0.00 0.00 22.28 1 . 64 0.00 - 1 . 56 22.67 )   ( 24.27 2 . 63 0.00 3 . 14 26.67 0.00 0.00 0.00 19.51 )   ( 24.54 2 . 09 0.00 3 . 62 26.97 0.00 0.00 0.00 20.58 )  
1Hb   ( 23.56 0.00 0.00 0.00 38.92 9 . 61 0.00 7 . 94 30.47 )   ( 21.91 0.00 0.00 0.00 22.28 1 . 64 0.00 1 . 56 22.67 )   ( 31.11 4 . 14 0.00 2 . 36 25.34 0.00 0.00 0.00 22.62 )   ( 22.14 0 . 20 0.00 5 . 33 32.60 0.00 0.00 0.00 23.05 )  
13   ( 114.54 0.00 0.00 0.00 81.79 0.00 0.00 0.00 9.79 )   ( 49.56 57 . 37 0.00 75 . 51 26.46 0.00 0.00 0.00 93.83 )   ( 38.59 51 . 93 0.00 79 . 14 28.56 0.00 0.00 0.00 93.40 )  
17Oa   ( 314.23 0.00 0.00 0.00 370.80 0.00 0.00 0.00 327.17 )   ( 429.62 0.00 0.00 0.00 461.54 0.00 0.00 0.00 1065.23 )   ( 382.19 67 . 86 0.00 14 . 85 176.57 0.00 0.00 0.00 299.30 )   ( 307.84 59 . 21 0.00 4 . 62 177.50 0.00 0.00 0.00 287.62 )  
17Ob       ( 127.79 130 . 28 0.00 156 . 91 121.44 0.00 0.00 0.00 194.57 )   ( 204.40 79 . 90 0.00 174 . 00 91.93 0.00 0.00 0.00 175.67 )  
HaOHb HaHbCO cis-HaCOaObHb trans-HaCOaObHb
1Ha   ( 23.56 0.00 0.00 0.00 38.92 9 . 61 0.00 7 . 94 30.47 )   ( 21.91 0.00 0.00 0.00 22.28 1 . 64 0.00 - 1 . 56 22.67 )   ( 24.27 2 . 63 0.00 3 . 14 26.67 0.00 0.00 0.00 19.51 )   ( 24.54 2 . 09 0.00 3 . 62 26.97 0.00 0.00 0.00 20.58 )  
1Hb   ( 23.56 0.00 0.00 0.00 38.92 9 . 61 0.00 7 . 94 30.47 )   ( 21.91 0.00 0.00 0.00 22.28 1 . 64 0.00 1 . 56 22.67 )   ( 31.11 4 . 14 0.00 2 . 36 25.34 0.00 0.00 0.00 22.62 )   ( 22.14 0 . 20 0.00 5 . 33 32.60 0.00 0.00 0.00 23.05 )  
13   ( 114.54 0.00 0.00 0.00 81.79 0.00 0.00 0.00 9.79 )   ( 49.56 57 . 37 0.00 75 . 51 26.46 0.00 0.00 0.00 93.83 )   ( 38.59 51 . 93 0.00 79 . 14 28.56 0.00 0.00 0.00 93.40 )  
17Oa   ( 314.23 0.00 0.00 0.00 370.80 0.00 0.00 0.00 327.17 )   ( 429.62 0.00 0.00 0.00 461.54 0.00 0.00 0.00 1065.23 )   ( 382.19 67 . 86 0.00 14 . 85 176.57 0.00 0.00 0.00 299.30 )   ( 307.84 59 . 21 0.00 4 . 62 177.50 0.00 0.00 0.00 287.62 )  
17Ob       ( 127.79 130 . 28 0.00 156 . 91 121.44 0.00 0.00 0.00 194.57 )   ( 204.40 79 . 90 0.00 174 . 00 91.93 0.00 0.00 0.00 175.67 )  

The first two molecules shown in Table 7.1, water and formaldehyde, have a two-fold symmetry axis, resulting in the absence of an antisymmetric component of the oxygen and carbon shielding tensors, since these nuclei are placed along the C2v axis. The antisymmetric parts do not vanish for the protons, but the planarity of water and formaldehyde molecules constrains the direction of their antisymmetry, σ*(1H). The rotation about the C2v axis, which coincides with the z-axis of the reference frame chosen for the shielding tensor elements given in the table, interchanges the protons. Hence, any element ( σ y * , σ z * ) of the σ*(1H) vector in the yz-plane, in which the atoms are, vanishes. Thus, from computations, it follows that (i) the shielding tensors of protons in a water molecule do have antisymmetry that is perpendicular to the plane containing the two protons and the oxygen atom of the molecule (along the x-axis), and in a water molecule, the antisymmetric part of the shielding tensor of the first proton is oriented antiparallel to the antisymmetric part of the shielding tensor of the second proton, σ*(1Ha) = −σ*(1Hb); see Figure 7.1.

Figure 7.1

From left to right: water (at the top), hydrogen peroxide (at the bottom), and 2,2′-propionyl-3,3′,4,4′-tetramethyl-1,1′-diphosphaferrocene (top and side views of two conformations) molecules. The permanent electric dipole moments μe are depicted as green arrows, while the nuclear magnetic shielding antisymmetries σ* are shown as red and blue arrows. Vector lengths are not to scale.

Figure 7.1

From left to right: water (at the top), hydrogen peroxide (at the bottom), and 2,2′-propionyl-3,3′,4,4′-tetramethyl-1,1′-diphosphaferrocene (top and side views of two conformations) molecules. The permanent electric dipole moments μe are depicted as green arrows, while the nuclear magnetic shielding antisymmetries σ* are shown as red and blue arrows. Vector lengths are not to scale.

Close modal

Nuclei of less symmetric molecules, such as formic acid, may all have antisymmetric components in their shielding tensors. In a very crude approximation, the magnitude of the shielding antisymmetry is correlated with the variation of the shielding for a given element. For example, typically in an organic molecule, the nuclear shielding of proton (1H), carbon (13C), and oxygen (17O) varies in the range of ten, a hundred, and several hundred ppm, respectively. This trend is exemplified in the case of formic acid nuclear shielding antisymmetries, which are less than 3 ppm for protons, about 10 ppm for carbon, and reach almost 40 ppm for oxygen. Several other examples of various nuclear shielding antisymmetries of benzene derivatives are provided in Figure 7.2.

Figure 7.2

The magnitude of the antisymmetric part of carbon-13 nuclear magnetic shielding tensors σ* (in ppm) for benzene derivatives: phenol, toluene, benzonitrile, benzaldehyde, and benzoic acid. The molecules are planar, and consequently, all nuclear shielding antisymmetries are perpendicular to the plane of the figure. The antisymmetries of protons and carbon-13 nuclei of benzene vanish due to the high symmetry of the molecule, i.e., because its symmetry point group is D6h.

Figure 7.2

The magnitude of the antisymmetric part of carbon-13 nuclear magnetic shielding tensors σ* (in ppm) for benzene derivatives: phenol, toluene, benzonitrile, benzaldehyde, and benzoic acid. The molecules are planar, and consequently, all nuclear shielding antisymmetries are perpendicular to the plane of the figure. The antisymmetries of protons and carbon-13 nuclei of benzene vanish due to the high symmetry of the molecule, i.e., because its symmetry point group is D6h.

Close modal

Let us note that if a molecule is non-rigid, then the impact of the flexibility of the molecule may greatly vary with the position of the nucleus. For example, the antisymmetry σ*(1Ha) is almost the same for the two conformers of formic acid, cis-HCOOH and trans-HCOOH, in contrast to the antisymmetry σ*(1Hb) that is more than two times larger for trans-HCOOH in comparison to cis-HCOOH. An analogous observation applies to the pair of 17Oa and 17Ob atoms in the formic acid molecule. Antisymmetry is a much more local property than the permanent molecular electric dipole moment. This is shown by the example of chiral phosphaferrocene depicted on the right side of Figure 7.1. While the dipole moment of the molecule (μe) changes significantly depending on the opening angle between the acrylic substituents, the shielding antisymmetries of the phosphorus nuclei σ*(31P) remain approximately constant and the opening angle only affects their directions.

The next case of a flexible molecule is hydrogen fluoroperoxide, whose σ*(1H) and σ*(19F) antisymmetries are shown in Figure 7.3. As in the case of formic acid, the antisymmetry of the nucleus with a large variability of magnetic shielding, fluorine-19, is roughly an order of magnitude larger than that of nuclei with small shielding variation, i.e., proton. Hydrogen fluoroperoxide is an unstable molecule whose enantiomers interconvert into each other (by the rotation of the oxygen–oxygen bond) on a fast time scale from the point of view of NMR. FOOH molecules with dihedral angles ∠(FOOH) ranging from −180° to 0° are (R)-enantiomers, and those with ∠(FOOH) ranging from 0° to 180° are (S)-enantiomers. Although it is true that the directions of the antisymmetries σ*(1H) and σ*(19F) depend on the molecular conformation and, consequently, on the chirality of a molecule, one cannot use this difference to discriminate between the FOOH enantiomers. The reason for this is that, in a standard liquid-state NMR experiment, only the amplitude of antisymmetry can be measured, e.g., by measuring the rate of nuclear relaxation, which is the same for the positive and negative values of the dihedral angle ∠(FOOH). Moreover, the significant variation of the antisymmetries σ*(1H) and σ*(19F) with the molecular shape indicates that conformational flexibility may be one of the strongest factors determining the direction and amplitude of the antisymmetry, which, in order to reach an agreement with experimental observations, has to be averaged over all the conformations of the molecule that are available at a given temperature.

Figure 7.3

Variation in magnetic shielding antisymmetry of protons and fluorine nuclei in the hydrogen fluoroperoxide molecule with the ∠(FOOH) dihedral angle. Computed using the CFOUR program by the CCSD method in the aug-pCVTZ basis set; molecular geometry rFO = 1.4107 Å, rOO = 1.4246 Å, rOH = 0.9659 Å, ∠(FOO) = 104.55°, ∠(OOH) = 101.97° used for the computation of antisymmetries at the first point [∠(FOOH) = 0°] of the graph. The other structural parameters were reoptimized for each value of the fixed dihedral angle ∠(FOOH).

Figure 7.3

Variation in magnetic shielding antisymmetry of protons and fluorine nuclei in the hydrogen fluoroperoxide molecule with the ∠(FOOH) dihedral angle. Computed using the CFOUR program by the CCSD method in the aug-pCVTZ basis set; molecular geometry rFO = 1.4107 Å, rOO = 1.4246 Å, rOH = 0.9659 Å, ∠(FOO) = 104.55°, ∠(OOH) = 101.97° used for the computation of antisymmetries at the first point [∠(FOOH) = 0°] of the graph. The other structural parameters were reoptimized for each value of the fixed dihedral angle ∠(FOOH).

Close modal

As mentioned at the end of Section 7.3, even though the antisymmetry of the nuclear magnetic shielding tensor is a key component required for direct discrimination between enantiomers, its mere presence is not sufficient for designing an experiment, where the outcome will solely depend on the absolute configuration of a molecule. The significance of nuclear shielding antisymmetry lies in its transformation properties under space inversion.

Let us consider a mirror image of a solenoid through which a current flows. If the axis of the solenoid is vertically oriented, parallel to the surface of the mirror and the current flows counterclockwise, then the magnetic field generated by the solenoid is directed upwards. In the mirror image of a solenoid, the direction of current flow is reversed, i.e., current flows clockwise, and consequently, the magnetic field is directed downwards (Figure 7.4A). Comparing the magnetic-field behaviour with an electric field of a capacitor that has its plates perpendicular to the mirror plane, one finds that the electric field does not alter its direction upon mirror reflection (Figure 7.4B). So, although the magnetic and electric fields are vector quantities, they exhibit different transformation properties under mirror reflection. The manner in which particular elements of the vector change depends on the vector orientation with respect to the mirror plane. If the solenoid axis is perpendicular to the plane, then the direction of the magnetic field will remain the same. However, for a capacitor whose plates are parallel to the plane of the mirror, the direction of the electric field will be reversed. Therefore, it is convenient to use space inversion instead of mirror symmetry since, if the former acts on the magnetic and electric fields, we find that B → B and E → −E.

Figure 7.4

The magnetic field B generated by the current J flowing through the solenoid (A) and the electric field E generated by positive and negative charges on the plates of the capacitor (B) compared with their mirror images. Note the reversal of the direction of the B field.

Figure 7.4

The magnetic field B generated by the current J flowing through the solenoid (A) and the electric field E generated by positive and negative charges on the plates of the capacitor (B) compared with their mirror images. Note the reversal of the direction of the B field.

Close modal

If we would like to introduce mirror reflections into the rule defining a vector A, i.e., its components are transformed according to R · A, where R is the rotation matrix, we need to differentiate proper and improper rotations, where the latter begins with a mirror reflection followed by a proper rotation. A proper rotation matrix has the determinant +1, while the determinant of an improper rotation matrix is −1. Thus, a polar vector, e.g., an electric field, transforms under any rotation (proper or improper) according to the formula R · A, while a pseudovector (alternatively called an axial vector), e.g., magnetic field, transforms as det(R)R · A. These transformation rules allow us to reproduce the result presented at the end of the previous paragraph because the improper rotation matrix corresponding to inversion through a point (S2 = σhC2) is −13 (minus unity 3 by 3 matrix).

For direct chiral discrimination, one needs to find a suitable combination of experimentally available vectors that transform differently for the enantiomers of a given molecule. In order to do that, let us consider two fundamental operations that can be performed on two vectors A and B: a scalar (or dot) product A · B and a vector (or cross) product A × B. These products transform under rotations as follows:
(7.10)
(7.11)
Let us use eqn (7.11) to check whether the magnetic field is a polar vector or a pseudovector. According to the Biot–Savart law, the magnetic field generated by a solenoid is an integral over the vector product of an infinitesimal element of current flow through a wire and the displacement vector from the wire to the point of observation. Both the current and the displacement vector are polar vectors, so their cross-product is a pseudovector according to eqn (7.11).
The antisymmetry of the nuclear magnetic shielding tensor is a pseudovector. In order to see this, let us focus only on the antisymmetric component from eqn (7.6) defining the energy of a magnetic dipole placed in the effective magnetic field at the atomic nucleus. Then, this equation takes the form
(7.12)
Energy is a scalar quantity, and it is assumed here that it is the same for both enantiomers of a given molecule. The nuclear magnetic moment and the magnetic field are pseudovectors. It follows from eqn (7.11) that their cross-product is a pseudovector. Consequently, the antisymmetry must be a pseudovector in order to fulfil eqn (7.10). Eqn (7.12) clearly shows that (i) the antisymmetry has different transformation properties than, for instance, the isotropic shielding (one is a pseudovector, the other a scalar), but (ii) the introduction of an additional interaction is necessary for chiral discrimination. It will be shown in the next section that this interaction occurs with an external electric field used in addition to the magnetic field.
In the following text, it is assumed that all nuclei have the spin quantum number 1 2 , thereby excluding the presence of the quadrupolar interaction, and that the molecule is diamagnetic.** The complete form of the Hamiltonian of a magnetically shielded nucleus written using the decomposition from eqn (7.1) is as follows:
(7.13)
The first term −μm · B, is the Zeeman interaction, i.e., the interaction between the bare nucleus and the magnetic field. The subsequent terms describe the interaction of the spins with the induced magnetic field at the position of the nucleus; i.e., the field is altered by the nuclear magnetic shielding caused by the motion of electrons of the molecule.
Suppose that the sample is a liquid and consists of molecules that have permanent electric dipole moments μe. The energy of a molecule with an electric dipole μe placed in an electric field E, is −μe · E. Therefore, using the Boltzmann average over an ensemble of molecules placed in the electric field, we find that the approximate Hamiltonian of the spin system, which is perturbed by the electric field, is11,12 
(7.14)
where kB = 1.380649 × 10−23 J K −1 is Boltzmann’s constant and T is the temperature. The E and B fields in eqn (7.14) are the local (microscopic) fields, and they may differ from those applied externally in the experiment. These differences may be considerably more pronounced for the electric field than for the magnetic field, as for a typical diamagnetic sample, the relative permittivity (10 < εr < 100) is greater than the relative permeability (μr − 1 < 10−3).
The condition that the molecule bears a permanent electric dipole moment is not necessary for the existence of the predicted NMR effects induced by the electric field E. However, for molecules that do not meet this condition, the amplitudes of the expected effects are too small for practical use;11  thus, their discussion will be omitted. In the case of the absence of a permanent dipole moment, the effects depend on the polarizability of the magnetic shielding tensor σ i j k ( 1 ) in the linear term in the power series expansion of the tensor σ as a function of the electric field E, i.e.,
(7.15)
Phenomena analogous to those described below (i.e., effects E 1 E 3 ) are present even if μe = 0 and can be identified by substituting σ i j ( 0 ) μ k e σ i j k ( 1 ) in the appropriate formulae. For the sake of brevity, the index zero for the unperturbed tensors σ and J is omitted in the following text.
Let us compute the isotropic part of H ˆ E , since in the case of sufficiently fast molecular tumbling, only the isotropic part of the nuclear Hamiltonian manifests itself directly in the NMR spectrum. The terms of H ˆ E containing the molecular quantities, the permanent electric molecular dipole moment μe and the nuclear magnetic shielding tensor σ, that form scalars or pseudoscalars will contribute to the isotropic part of H ˆ E because they are invariant under the rapid (on the NMR time scale) rotational diffusion of molecules. For instance, the isotropic part of the shielding tensor σ is a scalar, so the term σ I i iso μ m B contributes to the Hamiltonian H ˆ E , iso . Besides this, there is another term, resulting from the dot product between the electric moment μe, a polar vector, and antisymmetry σ*, a pseudovector. The result is a pseudoscalar
(7.16)
The pseudoscalar σc (in fm V−1) has the opposite sign for the enantiomers of a molecule and permits one to differentiate between enantiomers directly. This, with the aid of quantum chemistry computations capable of providing the sign of the dot product μe· σ*, opens the way to determine the absolute configuration of the molecule without the need for a chemical modification of the sample, as is the case with classical indirect NMR methods for studying chirality. It follows from eqn (7.16) that the chirality-sensitive, isotropic part of the Hamiltonian H ˆ E given by eqn (7.14) has the form
(7.17)
Let us remark that the chiral sensitivity of the pseudoscalar σc arises because of the choice of a suitably transforming, externally applied quantity, i.e., an electric field E, which is a polar vector, that couples to another property that is associated with the local frame of reference of the molecule. In principle, one could design a chirality-sensitive NMR experiment by choosing a pair of externally controllable and molecular quantities with similar transformation properties under space inversion as those of an electric field.

The amplitude of the pseudoscalar σc may be maximized by making the favourable properties of the antisymmetric part of the shielding tensor and the molecular electric permanent dipole moment μe, i.e., minimization of the angle between μe and σ*, and by maximization of the amplitudes of μe and σ*. In order to illustrate the significance of both mentioned factors, let us take an example of permutations of the substituents in a rigid molecule of a triangular shape, a halogen derivative of cyclopropene HFClBrC3. The magnitude of the permanent electric dipole moment of HFClBrC3 is roughly the same (1.7–2.7 Debye) for its isomers obtained by permutations of hydrogen, fluorine, chlorine, and bromine: HFC–CBr═CCl (where Br and Cl are attached to carbons forming a double bond; the C–C bond closing the ring is implied), ClBrC–CH═CF, and FBrC–CH═CCl. Quantum-chemical computations of 19F nuclear magnetic shielding in HFClBrC3 at the DFT (density-functional theory) level show the following.

  • In HFC–CBr═CCl, μe and σ*(19F) are perpendicular to the cyclopropane ring plane, but the antisymmetry of fluorine shielding is relatively small (a few ppm); thus, even though the relative orientation of μe and σ*(19F) is favorable, ( μ e , σ * ) 12 ° , the pseudoscalar is small, ∣σc∣ < 2 pm V−1.

  • In ClBrC–CH═CF, μe is in the cyclopropane ring plane, while its σ*(19F) is almost perpendicular to the ring’s plane, ( μ e , σ * ) 91 ° , so althought σ*(19F) reaches nearly 70 ppm, the pseudoscalar is even smaller than in the case of HFC–CBr═CCl.

  • In FBrC–CH═CCl σ*(19F) is as large as ClBrC–CH═CF and simultaneously the angle ( μ e , σ * ) is slightly smaller than 30°, resulting in a 30 times larger value of |σc∣ in comparison with HFC–CBr═CCl and ClBrC–CH═CF (3500 pm V−1).

One can view F ̲ Br C ̲ CH ̲ CCl and ClBrC CH ̲ CF ̲ as molecular scaffolds for a two-spin 1H–19F system embedded in an electronic environment resembling an open-book topology of FOOH (also compare with H2O2 antisymmetries shown in Figure 7.1). This is because, in both cases, other atoms, such as heavy halogens, carbon, and oxygen are considered to be either NMR-inactive (12C, 16O) or self-decoupled isotopes (35Cl, 79Br). In the case of ClBrC–CH═CF, the FCCH part of the molecule is planar, and the pseudoscalar σc almost vanishes, similar to the achiral planar conformation of FOOH, for which σ*(19F) is perpendicular to the molecular plane and its permanent electric dipole moment μe is in the direction from the fluorine to the hydrogen atom. In contrast to the case of ClBrC–CH═CF, in FBrC–CH═CCl, the dihedral angle FCCH is about 90°, corresponding to a chiral conformation of FOOH.

Apart for the choice of substituents, the second factor influencing the magnitudes of μe and σ* and the angle ( μ e , σ * ) is molecular flexibility. Frequently, the value of the pseudoscalar σc varies with torsional angles, and in an extreme case, the pseudoscalar σc may reverse its sign for some values of the torsional angle. An example illustrating the influence of molecular flexibility on the sign of the pseudoscalar σc is given in Figure 7.5.

Figure 7.5

Dependence of the dot product (in D ppm) between the permanent electric dipole moment μe of (R)-alanine tripeptide (the substituents R1, R2, and R3 are methyl groups) and the antisymmetric parts of the nitrogen ( σ N * ) and carbon ( σ C * ) nuclear magnetic shielding tensors on the torsional angles ϕ and ψ.

Figure 7.5

Dependence of the dot product (in D ppm) between the permanent electric dipole moment μe of (R)-alanine tripeptide (the substituents R1, R2, and R3 are methyl groups) and the antisymmetric parts of the nitrogen ( σ N * ) and carbon ( σ C * ) nuclear magnetic shielding tensors on the torsional angles ϕ and ψ.

Close modal

The pseudoscalars σ N c and σ C c of nitrogen-15 and carbon-13 of (R)-alanine tripeptide (all three alanine groups are (R)-enantiomers) strongly vary with the conformation of the oligopeptide. The figure shows two regions of the Ramachandran plot: at the bottom, the range of the ϕ and ψ torsional angles corresponds to the structure of a β-sheet, while the top of the plot corresponds to the structure of an α-helix of a protein. Both regions have segments that exhibit positive and negative values of chirality-sensitive pseudoscalars; therefore, determining which enantiomer is present in the sample requires simultaneous knowledge about the conformation of the (R)-alanine tripeptide. On the other hand, the strong dependence of the pseudoscalar value on the structure of the molecule can be useful because it carries information about the detailed three-dimensional structure of a molecule within a certain structural motif. The usefulness lies in the fact, that by finding the values of chirality-sensitive pseudoscalars of nitrogen and carbon nuclei, one can determine the relative orientation of the chemical groups containing 15N and 13C nuclei, e.g., peptide bonds, relative to the permanent electric dipole moment of the molecule. Consequently, this strategy of spatial structure measurement would provide complementary data to those obtained from widely-used biomolecular structure determination techniques such as those utilizing the nuclear Overhauser effect, which gives distances between the atoms, or the Karplus relationship providing the torsional angles.

The relationship between the molecular structure and the amplitude and direction of the antisymmetric part of the nuclear magnetic shielding tensor in Section 7.3 and the current section allows us, to propose a general strategy for maximizing the amplitude of the pseudoscalar σc. This strategy utilizes the facts that the quantities μe and σ* are almost independent in many molecules, and that an achiral molecule can be transformed into a chiral one by attaching a chiral substituent. As mentioned earlier, the chirality-sensitive NMR effects reach maximal amplitude if the angle between the molecular dipole moment μe and the antisymmetric shielding σ* is optimal, and if the amplitudes of the vectors μe and σ* are large. Additionally, the molecule contains other advantageous structural features, e.g., it may be hyperpolarized, which amplifies the chirality-sensitive signals to a level at which they will be easily observed. While in the case of the pseudoscalar σc the optimal angle is 0°, for other effects described further in this chapter, the angle may take other optimal values, since the dependence of the amplitude of the chirality-sensitive effect may be related to the pseudoscalar in another more complicated way (vide infra).

In particular, the strategy is exemplified in Figure 7.6. The core of the chiral molecule (blue sphere) has two substituents, one of which has a large electric dipole moment (polar chemical groups; green cylinder) while the second one is hyperpolarisable and contains a nucleus with a large antisymmetry σ*, (indicated by a red disc). Examples suitable for a SABRE (signal amplification by reversible exchange) experiment are camphor substituted with pyridine (Figure 7.6A) and α-amino acid derivative with a triazole ring (Figure 7.6B). Other target molecules are light fluorinated alcohols and their derivatives with a substituent that has a triple carbon–carbon bond permitting the usage of ALTADENA (adiabatic longitudinal transport after dissociation engenders nuclear alignment) – Figure 7.6C (the latter is not shown in the figure). Note that, in practice, one has to make a trade-off between the molecular complexity and molecular weight since the more molecular features are taken into account, the heavier the molecule becomes, and the molar concentration decreases to which the expected chirality-signal is proportional. This point is clearly visible if we contrast the molecules shown in Figure 7.6A and C. Moreover, this approach is more suitable for a rigid molecule than a flexible one. In the case of flexible molecules, signals may partially cancel out as it happens for (R)-alanine tripeptide shown in Figure 7.5. This is the price that one has to pay for the possible insight into the dynamical behaviour of the molecule, based on the determination of the amplitude of the pseudoscalar σc.

Figure 7.6

The general strategy and examples of chiral molecules that are optimized to maximize the pseudoscalar σc amplitude and consequently expected to provide chirality-sensitive signals.

Figure 7.6

The general strategy and examples of chiral molecules that are optimized to maximize the pseudoscalar σc amplitude and consequently expected to provide chirality-sensitive signals.

Close modal

Let us briefly summarize the results described in the chapter up to now, in view of the other approaches to directly see chirality by NMR. Parity non-conservation does not influence the NMR spectra measured using the currently used NMR spectrometers.13,14  Consequently, one assumes that the Hamiltonian H ˆ CS given in eqn (7.13) is even under the inversion through a point. Therefore, one cannot directly discriminate between the enantiomers through NMR spectroscopy unless the molecule is transformed into a diastereomer (e.g., via Mosher’s method)15  or a chiral environment is present. This limitation is lifted if the electric field operator, which is odd under inversion, is involved in the Hamiltonian.16,17  In this case, some contributions to the Hamiltonian H ˆ E defined in eqn (7.14), e.g., H ˆ CS , c (see eqn (7.17)), change their sign under the inversion, which may be used for a direct discrimination between the enantiomers. Note that the effects that are linear in the electric field are chirality-sensitive, and they vanish for an achiral sample. In contrast, the quadratic effects are present both in the achiral and the chiral samples. We will focus only on the linear effects and show that all of these effects generate NMR signals of the opposite phases (i.e., phase-shifted by 180°) for the enantiomers of a given molecule. The phase dependence of the effects induced by the electric field on chirality differentiates the chirality-sensitive effects described here from the other effects proposed in the literature, which rely on the shift of the NMR lines due to molecular chirality.

In further discussion, it is convenient to introduce the concept of NMR phenomena in which the presence of the electric field is essential. Among others, the effects described in this chapter can be termed nuclear magnetoelectric resonance (NMER; [nmɪr], i.e., NM-ee-R). The adjective magnetoelectric (rather than “electromagnetic”) emphasizes that we are dealing with NMR phenomena either generating an electric field (excluding the electric field associated with the time-varying magnetic field resulting from the Larmor precession of magnetic moments) or induced by the application of an external electric field. This means that interactions such as quadrupolar coupling in which an internal molecular electric field is involved, are also ruled out. NMR is based on the magnetic response of the spin system, while NMER also utilizes the electric response originating from the partial orientation of the sample due to the presence of the antisymmetric part of the nuclear magnetic shielding tensor, σ*. Although the definition outlined above is not entirely strict, the bottom line is that at the essence of NMER lies the slight perturbations of the dynamics of the nuclear spin system associated with the presence of an electric field. Not all NMER signals are chirality-dependent, as may be exemplified in the case of the quadratic effects.

Although work on experimental NMER spectra is still in progress, let us use a 1H NMER spectrum of sucrose calculated using quantum chemistry methods, to indicate the unique features of NMER spectroscopy (see Figure 7.7).

Figure 7.7

Experimental 1H NMR spectrum (A) vs. predicted 1H NMER spectrum (based on DFT computations with the PBE0 functional and the aug-cc-pVTZ basis set; B–d-sucrose, C–l-sucrose) of sucrose showing differences between the NMR spectrum and the NMER spectrum; the two bottom traces are for illustrative purpose only.

Figure 7.7

Experimental 1H NMR spectrum (A) vs. predicted 1H NMER spectrum (based on DFT computations with the PBE0 functional and the aug-cc-pVTZ basis set; B–d-sucrose, C–l-sucrose) of sucrose showing differences between the NMR spectrum and the NMER spectrum; the two bottom traces are for illustrative purpose only.

Close modal

Let us have a look at the first region of the spectra marked by rectangle I. The signals corresponding to NMER signals have opposite phases for d-sucrose and l-sucrose, i.e., for one, the signals are up, while for the other, they are down. In particular, the NMER signals of enantiomers are phase-shifted by ±90° with respect to the NMR signals; the NMER spectra mentioned above are additionally shifted by −90° for clarity. In contrast to NMER, the 1H NMR spectrum of sucrose is the same for both enantiomers. In the case of NMER, the relative intensities of lines and, in particular, integrals of the peaks do not correspond to the molar ratios of protons in the molecule as it is for NMR (see the region marked by rectangle II). Rather than providing information about the number of protons in the individual chemical groups, their intensities depend on the orientation of the given group of protons with respect to the permanent dipole moment of the molecule. Based on this dependence, one may infer the absolute configuration of the molecule through NMER. Finally, rectangle III shows the region with a strong solvent signal (water). This signal would not cover the spectrum of the studied compound in the case of NMR since water is, similar to many other commonly used solvents, an achiral substance. The common feature of NMR and NMER signals is that, in both cases, the frequency of the signal equals the spin precession frequency. The points differentiating NMER from NMR are summarized in Table 7.2.

Table 7.2

Comparison between the ordinary 1H NMR spectrum and anticipated 1H NMER of (R/S)-enantiomers.

NMR NMER
Principle of operation  Magnetic field of precessing nuclear magnetic moment induces a voltage in a receiver coil  Variation of the electric field (of molecular electric dipole moment) generates a voltage in a capacitor-like device; alternatively, the external electric field partially aligns the sample, which results in a perturbation of the dynamics of the spin system 
NMR NMER
Principle of operation  Magnetic field of precessing nuclear magnetic moment induces a voltage in a receiver coil  Variation of the electric field (of molecular electric dipole moment) generates a voltage in a capacitor-like device; alternatively, the external electric field partially aligns the sample, which results in a perturbation of the dynamics of the spin system 
Sample R S Racemic mixture R S Racemic mixture
Observed signal             
Amplitude  A   A   A   B   B  
Phase  0°  0°  0°  +90°  −90°  N/A 
Frequency  ω0   ω0   ω0   ω0   ω0   N/A 
Sample R S Racemic mixture R S Racemic mixture
Observed signal             
Amplitude  A   A   A   B   B  
Phase  0°  0°  0°  +90°  −90°  N/A 
Frequency  ω0   ω0   ω0   ω0   ω0   N/A 
The time evolution of the spin system under the influence of the externally applied electric field E is given by the solution of the Liouville von Neumann equation,
(7.18)
where ρ(t) is the density matrix of the spin ensemble. Introduction of the density matrix ρ(t) allows us to take into account the fact that the sample is macroscopic and the number of spin systems is of the order of the Avogadro number. The detailed solution of eqn (7.18) is beyond the scope of the chapter, so only a brief description of the method used to solve is given here. For technical details, see, for instance, ref. 18 and 19. A quantum-mechanical counterpart of the energy of a shielded magnetic dipole placed in a magnetic field, given in eqn (7.6), can be found by substituting the dipole moment μ m γ N I ˆ , where γN is the nuclear gyromagnetic ratio, e.g., the gyromagnetic ratio of proton is γH ≈ 2π·42.577 MHz T−1,  = 1.054571817 × 10−34 J s is reduced Planck’s constant, and I ˆ is the spin operator represented by, e.g., a vector of Pauli matrices for a spin- 1 2 . In eqn (7.18), ι is the imaginary unit, i.e., ι2 = −1. The bracket [⋯,⋯] represents the symbol of a commutator. In a given basis, eqn (7.18) can be rewritten in a matrix form and consequently solved using methods applicable to a system of first-order differential equations.

It is convenient to use second-order perturbation theory to find tractable analytical solutions of eqn (7.18), averaged over the ensemble of tumbling molecules. This approach is used in the Bloch–Redfield–Wangsness theory.20,21  For the considered case, i.e., isotropic rotational diffusion, the changes in the density matrix ϱ(t) over time may be evaluated using the irreducible spherical tensor decomposition without the need to explicitly average over molecular orientations. This approach was used in the standard NMR case, i.e., without the electric field, in ref. 22. If the coefficients of the obtained relaxation matrix are time-dependent, as in the case of a modulated electric field, then the equation is further approximated by transforming the reference frame to a frame that rotates around the z-axis with the spin precession frequency of each spin, expanding the obtained fundamental matrix of the linear ordinary differential equation system in the Floquet–Magnus series, and truncating the series after the few first terms.23 

Taking into account the incoherent spin system dynamics, the number of irreducible spherical components of the Hamiltonian H ˆ E is larger than 30, even for a two-spin system, which gives at least a thousand of interactions described by the tensors. Therefore, a careful selection of the desired interactions is mandatory. This selection is based on the following three criteria: (i) the amplitude of the effect is sufficiently large for detection in the magnetic and electric fields of strengths accessible by the currently used experimental equipment, (ii) an appropriately designed experiment enables the separation of the effect from other simultaneously present effects, and (iii) experimental observation of the effect reveals interesting information about the molecular structure for chemists and physicists. Based on the general solution of eqn (7.18) and these three criteria, one can propose three chirality-sensitive effects described in subsequent sections: magnetization induced by antisymmetry of the shielding tensor ( E 1 , see Section 7.9), electric polarization induced by the antisymmetry of shielding tensor ( E 2 , see Section 7.10), and the interference between chemical shift anisotropy and dipolar relaxation in an electric field ( E 3 , see Section 7.11). In the following text, we assume that the sample is a liquid and the molecules tumble considerably faster than the electric field varies in time. The postulated effects are listed from those with the highest amplitude ( E 1 ) to those with the lowest amplitude ( E 3 ) .

The chirality-sensitive contribution H ˆ CS,c to the Hamiltonian H ˆ E that does not average out in the liquid phase, is a mixed product σc[E·(B × μm)]; see eqn (7.17).24  Let us take advantage of the fact that the mixed triple product is invariant under cyclic permutations of its components. Therefore, the energy σc[E·(B × μm)] may be interpreted either as the energy −μe,c · E of the induced electric dipole moment μe,c = −σc(B × μm) in the electric field E (the E 2 effect discussed in Section 7.10), or the energy −μm,c · B of the induced magnetic dipole moment μm,c = σc(μm × E) placed in the magnetic field (the E 1 effect described in this section). The latter effect was described by Buckingham and Fischer;25  more details are given in ref. 26 and 27. For this contribution, the equation describing the dynamics of magnetization M in the electric field E and the magnetic field B, is
(7.19)
Here, M is the sum of nuclear magnetic moments μm divided by the sample volume (magnetization in A m−1), T1 is the longitudinal (spin-lattice) relaxation time (in s), T2 is the transverse (spin–spin) relaxation time (in s), Meq is the magnetization of the sample in thermodynamic equilibrium, which has the same direction as the magnetic field B and an amplitude of 1 4 μ m μ m B k B T per sample volume; the one-fourth factor comes from the one-half spin of the nucleus. In eqn (7.19), the influence of the electric field E on the relaxation processes is neglected; the matrix containing relaxation times is written assuming that the magnetic field is oriented along the z-axis. Eqn (7.19) directly follows from eqn (7.18) if one changes the variable μ m I ˆ M and computes the ensemble average over the sample according to the procedure provided in Section 7.8. It generalizes the Bloch equation fundamental for NMR, for the case of chiral molecules placed in an electric field. Moreover, it fully describes the changes in magnetization over time due to a spin that does not interact with the other spins in the molecule, and it is a useful approximation for cases in which such interactions are negligible.

The inductively coupled NMR signal detection requires precession of the nuclear magnetization, which takes place if the vector M has components perpendicular to the main static field B0, used in the experiment. The external magnetic field B0 is commonly assumed to be along the z-axis of the laboratory frame of reference, and we will follow this convention in further text. However, it is worth noting that in some cases it is convenient to orient the field in a different way due to technical reasons, e.g., along the x-axis – this is the case with low magnetic field experiments and, partially, in NMR imaging. In the case of B0 = B0ez, the vector M rotates in the xy-plane with the angular frequency ω0 = γNB0 (in rad Hz).

In order to rotate the magnetization M from its initial orientation resulting from thermodynamic equilibrium, one can use an oscillating magnetic field B1, which is perpendicular to the main field B0 and changes with frequency ωB (called a resonance frequency) over time. Both components of the term B+ σcB × E in eqn (7.19) may be responsible for the rotation of the magnetization M if one applies external, oscillating magnetic B1(t) and electric E1(t) fields. However, even for a sample placed in an electric field of strength E1 = 1 kV mm−1 and for which the pseudoscalar σc is 1 fm V−1, the term σcE1 is considerably smaller than unity. Thus, the angle of rotation of the nuclear magnetization induced by the time-dependent electric field remains a few orders of magnitude smaller than that caused by the time-varying magnetic field.

Let us assume that, at a given point in the sample, the electric field is E1 = E1 sin(ω0t)ey. This field, according to Maxwell equations, generates a magnetic field oscillating at the same frequency, B1 = B1 cos(ω0t)ex, as that of the field E1. In a strong magnetic field B0, a component of the magnetic field B1, which has the same direction as the magnetic field B0, practically does not affect the state of the nuclear spins.17,27  Taking into account nuclear relaxation processes, one can determine a steady-state solution for the nuclear magnetization after the application of the electric field for a long period τ (in practice, it is sufficient to fulfill the condition τ > 3T1),
(7.20)
Here the frequencies of nutation of the nuclear magnetization caused by the electric and magnetic fields are ωE = |σc|E1γNB0/2 and ωB = γNB1/2, and
(7.21)
where the ratio κ = cB1/E1 (c is the speed of light) depends on the electric and magnetic fields generated by the source of the electric field used in the experiment; the quantity κN = σccB0 depends on the studied sample.

The details of the experimental protocol are shown in Figure 7.8. The electric field E1, which oscillates with the spin precession frequency ω0 for the duration τ > 3T1, rotates the nuclear magnetization (Figure 7.8A). A capacitor may generate an electric field E1; however, the oscillating-in-time B1 field saturates the magnetization and decreases the amplitude of the chirality-sensitive signals (more details are given in ref. 17). Therefore, a sufficiently effective suppression of the unwanted B1 requires a dedicated resonator. The resonator depicted in Figure 7.8C consists of a capacitor modified by adding two loops (i.e., loop-gap resonator). The magnetic field generated by these loops cancels the transverse component of the magnetic field generated by the capacitor, and the resulting magnetic field is oriented in the same direction as the field B0 of the NMR spectrometer. The volume of the sample placed at the central point between the plates, where unwanted excitation does not saturate the magnetization, is marked by the red colour (κ < 10−5) in Figure 7.8C, and is sufficiently large for generating a signal approximately 10% of that generated in the standard NMR experiment (Figure 7.8B). The signals obtained from the enantiomers exhibit different phases, contributing to the spectrum as both positive and negative peaks. The positive peak is arbitrarily attributed to the (R)-enantiomer (Figure 7.8C).

Figure 7.8

Experimental protocol for the observation of the magnetization induced by the antisymmetry of the shielding tensor (the effect E 1 ). The pulse sequence generating the effect E 1 (A), the scheme illustrating directions of exciting radiofrequency electric field E1 and the main magnetic field B0 (B), one of the resonators that are capable of providing the desired electromagnetic field distribution – the region of the κ parameter not exceeding 10−5 is marked by the red colour (C).

Figure 7.8

Experimental protocol for the observation of the magnetization induced by the antisymmetry of the shielding tensor (the effect E 1 ). The pulse sequence generating the effect E 1 (A), the scheme illustrating directions of exciting radiofrequency electric field E1 and the main magnetic field B0 (B), one of the resonators that are capable of providing the desired electromagnetic field distribution – the region of the κ parameter not exceeding 10−5 is marked by the red colour (C).

Close modal

Favourable samples for this experiment are the derivatives of cyclopropane (e.g., 1,3-diphenyl-2-fluoro-3-trifluoromethylcyclopropene) and light alcohols (e.g., 1,1,1-trifluoropropan-2-ol), which contain the fluorine nucleus. The latter are more convenient as they are stable at room temperature. Taking into account the possible imperfections in the fabrication of the resonator, the estimated chirality-induced NMR signal that would reach 1–5% of the standard achiral NMR signal, is predicted for an electric field of the strength E1 = 10 V mm−1 and the magnetic field B0 = 10 T.

The E 1 effect allows one to discriminate between the enantiomers, and it has the largest amplitudes among the chirality-sensitive effects described in this and the next chapter. The main limitation in detecting it is the suppression of the unwanted magnetic field generated by the oscillating-in-time electric field, along with convection in the sample caused by dielectric heating from the E1 field.

This effect was predicted by Buckingham16,24  and Buckingham and Fischer.25  The nuclear magnetization M precessing in the magnetic field B0 generates, in a liquid sample containing chiral molecules with non-vanishing permanent electric dipole moment μe, the chirality-sensitive electric polarization
(7.22)
The induced electric polarization P is detected by a capacitor. Any capacitor used at radio frequencies has some parasitic inductance; i.e., it partially acts as a coil and, thus, it can be excited by the variation of magnetic field over time. Therefore, it generates an unwanted signal of nuclear magnetization M, which is the same for both enantiomers.17  The expected signal S ( P ) of the electric polarization P is considerably weaker than the signal S ( M ) of magnetization M. The ratio of these two signals is28 
(7.23)
The ratio cP/M is smaller than 10−4 for diamagnetic samples; thus, it is important to use an experimental setup that has a low value of coefficient κ. Thus, one can use the loop-gap resonator, which allows the suppression of the unwanted signal of the precessing nuclear magnetization.

The experimental protocol for E 2 effect is shown in Figure 7.9. Initially, the magnetization M is rotated from its equilibrium position along the direction of the magnetic field B0 to the xy-plane, followed by precession at the frequency ω (Figure 7.9A). Its precession induces the chirality-sensitive electric polarization P, which, in the reference frame rotating with the frequency ω, aligns with the x-axis (Figure 7.9B). For a sample placed between the plates of the resonator, the signals S ( P ) of polarization P exhibit opposite signs for the enantiomers (shown arbitrarily as a positive peak for the (R)-enantiomer and a negative peak for the (S)-enantiomer). If the symmetry of the resonator is preserved, then the unwanted signal S ( M ) of the magnetization M integrated over the sample is zero (Figure 7.9C). See ref. 17 for more details about the resonator.

Figure 7.9

Experimental protocol for the observation of electric polarization induced by antisymmetry of the shielding tensor (the E 2 effect). The chirality-sensitive polarization is induced by precession of nuclear magnetisation (A). The observed signal could be detected using a capacitor (B) or, in order to supress the unwanted signal of magnetisation, using a loop-gap resonator (C).

Figure 7.9

Experimental protocol for the observation of electric polarization induced by antisymmetry of the shielding tensor (the E 2 effect). The chirality-sensitive polarization is induced by precession of nuclear magnetisation (A). The observed signal could be detected using a capacitor (B) or, in order to supress the unwanted signal of magnetisation, using a loop-gap resonator (C).

Close modal

The favourable samples for the E 2 effect are the same as those for the effect E 1 . The estimated magnitude of the chirality-sensitive electric polarization for these samples is P = 100 aC m−2 at a magnetic field of strength B0 = 10 T. Similar to the E 1 effect, the E 2 effect allows one to discriminate between enantiomers. This is one of the strongest predicted effects. The main advantage of the observation of the E 2 effect is that the sample is not perturbed by dielectric heating. If the sign of the product μe·σ* is known from quantum-chemical computations, then the experimental determination of the sign of the signal S ( P ) allows one to not only distinguish between the enantiomers but also to determine the absolute configuration of the molecule.

The perturbation of the interference between the dipole–dipole (DD) and the chemical shift anisotropy (CSA) relaxation mechanisms by the electric field E in a system consisting of spins I and S allows chirality-sensitive transitions of amplitude (described in detail in ref. 12; see Figure 2 in this reference article)
(7.24)
where ΔS = σzz − (σxx+ σyy)/2 is the anisotropy of the shielding tensor σS of the nucleus S. In eqn (7.24), it is assumed that the components of the shielding tensor σI of the nucleus I are considerably smaller than those for the nucleus S and, thus, they may be neglected. Moreover, the gyromagnetic ratio of the spin I is higher than that for the spin S (I and S could be, e.g., 1H and 13C, respectively). The amplitude ADxCSc of the E 3 effect is redefined here as the inverse of that reported in ref. 12, to obtain a coherent description with the E 1 effect predicted in ref. 29. The functions ψ 1 ( e σ * , e μ e , e b ) and ψ 2 ( e Δ , e μ e , e b ) are defined as follows:
(7.25)
(7.26)
where the unit vectors e σ * = σ * / σ * with σ* being the shielding asymmetry of S, e μ e = μ e / μ e , the unit vector eb is from the nucleus I to the nucleus S, and the unit vector eΔ is along the z-axis of the principal axis system of the symmetric part of the shielding tensor σS of the S nucleus.

The allowed transitions related to E 3 occur in the populations/zero coherence block (M0) and the single-quantum coherence block (M1) of the relaxation matrix of the spin system. From the viewpoint of the experimenter, the transitions within the M0 block are more convenient for detection as they require excitation by the electric field at the difference frequency. The maximal amplitude of the E 3 effect is achieved if the electric field E and the field B0 are parallel to each other. The experimental protocol for the E 3 effect is shown in Figure 7.10. At the beginning of the experiment, the initial state of the spin system, IySx+ IxSy, is created using the modified INEPT pulse sequence (Figure 7.10A). Then, this state evolves under the influence of the electric field E 1 = E 1 sin [ ( ω I ω S ) t ] z ˆ . The final state induced by field E1 is Iz of amplitude 〈Izc, which can be distinguished from the Iz component of amplitude 〈Izac created by the dipolar and CSA relaxation mechanism by subtracting the signal acquired with electric field E1 from that obtained without the E1 field.

Figure 7.10

Experimental protocol for the observation of the interference of CSA-DD relaxation mechanisms in an electric field (the E 3 effect). The chiral-dependent response of the system is generated by the pulse sequence utilizing both electric (the channel E) and magnetic (the channels I and S) excitations; G – the magnetic field gradient along the z-axis (A). The favourable orientation of the capacitor generating the radiofrequency electric field with respect to the static magnetic and a sketch of tensors (shown as ellipsoids) for enantiomers of the studied molecule (B).

Figure 7.10

Experimental protocol for the observation of the interference of CSA-DD relaxation mechanisms in an electric field (the E 3 effect). The chiral-dependent response of the system is generated by the pulse sequence utilizing both electric (the channel E) and magnetic (the channels I and S) excitations; G – the magnetic field gradient along the z-axis (A). The favourable orientation of the capacitor generating the radiofrequency electric field with respect to the static magnetic and a sketch of tensors (shown as ellipsoids) for enantiomers of the studied molecule (B).

Close modal
The magnitude of the effect depends on the strength of the field B0,
(7.27)
where the optimal strength of this field for the experiment is
(7.28)
Let us suppose that the pair of spins consists of a proton and carbon-13 separated by the typical distance of the C–H chemical bond, rCH = 1.1 Å, and the anisotropy ΔS of the magnetic shielding of the carbon nucleus is 200 ppm, then Bopt ≈ 11.75 T, i.e., in the range of the typical strength used in the chemical applications of NMR spectroscopy. For this optimal strength of the B0 field, the permanent electric dipole moment μe = 1 D, and E1 = 1 kV mm−1 oscillating at the frequency ω H 1 ω C 13 equal to 375 MHz, the chirality-sensitive magnetization of 1H–13C pair is approximately 5 × 10−6 of the magnetization of the proton at thermodynamic equilibrium.

In contrast to E 1 and E 2 effects, that vanish if the antisymmetric part of the nuclear shielding tensor is zero, the E 3 effect is non-zero even for σ* = 0 (as long as ψ2 ≠ 0); thus, it may be observable in a broader range of chiral molecules. The modulation of the electric field at the difference frequency permits the use of a simpler resonator for excitation than that required in the experiment aiming to observe the E 3 effect, i.e., a parallel-plate capacitor.

Let us briefly summarize several pieces of information that may be obtained in the studies of the above-described effects E 1 , E 2 , and E 3 .

  • The effects E 1 E 3 are linear in the electric field; consequently, they are only present in chiral samples.

  • The measurements of the amplitudes of E 1 E 3 effects allow us to determine the three-dimensional structure of the molecule. In particular, they offer insight into the relative orientation of local nuclear magnetic properties, i.e., the nuclear magnetic shielding tensors and the permanent electric dipole moment of the molecule. Furthermore, each effect depends in a different manner on the structural parameters and tensor components; thus, a combination of these effects provides comprehensive information about the molecular structure.

  • The E 1 E 3 effects allow one to distinguish directly between the enantiomers. Measurements of E 1 E 3 combined with the results of simple quantum-chemical computations permit the determination of the absolute configuration of the molecule. The quantitative measurement, i.e., determination of the sign (phase) of the signals of the E 1 E 3 effects, is sufficient for assigning the absolute configuration, which is less experimentally demanding.

The predicted effects for a single-spin and a two-spin system differ in terms of their optimal angle between the externally applied electric field E and the field B0, and the frequency of the electric field E, that varies with time. The application of the oscillatory electric field amplifies the E 1 E 3 effects by at least six orders of magnitude as compared to the application of a static field. These factors, favourable samples for the first experiments (predicted from quantum-chemical computations), and the expected signals generated by the effects, are summarized in Table 7.3.

Table 7.3

Predicted NMER effects.

Effect Number of spins Nuclear properties Experimental conditions Favorable sample Expected signal magnitudea
Angle between E and B0 Frequency of P/E oscillations
E 1   σ   90°  ωF     5 × 10−1 
E 2   σ   N/A  ωF   1,1,1-trifluoro-propan-2-ol  10−2 
E 3   σ, Δ  0°  ωH − ωC   alanine  5 × 10−6 
Effect Number of spins Nuclear properties Experimental conditions Favorable sample Expected signal magnitudea
Angle between E and B0 Frequency of P/E oscillations
E 1   σ   90°  ωF     5 × 10−1 
E 2   σ   N/A  ωF   1,1,1-trifluoro-propan-2-ol  10−2 
E 3   σ, Δ  0°  ωH − ωC   alanine  5 × 10−6 
a

The magnitude with respect to that obtained after the excitation of the sample by a 90° pulse; E = 1 kV mm−1, B0 = 10 T.

The available strength of the electric field limits the highest achievable signals of the predicted effects; therefore, a careful selection of the sample for the first set of experiments is required. The fluorine nucleus has favourable properties as it has a high gyromagnetic ratio, large natural abundance, and notable antisymmetry of the nuclear magnetic shielding tensor. The magnitude of the effects for the same electric-field strength field may vary depending on the particular sample. However, for a given group of chemical compounds, one can predict that their magnitudes decrease in the following series: E 1 > E 2 > E 3 . The general requirements for the experimental setup, assuming a molecule with a permanent electric dipole moment of μe = 1 D, are a magnetic field B0 on the order of 10 T and an electric field E = 0.1–1 kV mm−1 which, for favourable samples mentioned in Table 7.3 (e.g., fluorine derivatives of light alcohols), generate expected signals from 5 × 10−1 (the E 1 effect) to 5 × 10−6 (the E 3 effect) of the 1H NMR signal, obtained after excitation of the sample by a 90° pulse at thermodynamic equilibrium. For effects with relatively large amplitudes, it is sufficient to use, at the beginning of the experiment, the nuclear magnetization at thermodynamic equilibrium, whereas in the case of the effects with the smallest amplitude, an application of hyperpolarization techniques may be required, to increase the initial nuclear magnetization (see Chapter 10 for details of hyperpolarization techniques applied to chiral molecules).

Besides the effects that are chirality-sensitive, the studies of the spin dynamics governed by the Hamiltonian H ˆ E and eqn (7.18), have revealed several other effects induced by an external electric field E, but which are not related to molecular chirality. Thus, such phenomena are potentially observable for all NMR-active nuclei. The two effects that are quadratic in the electric field are (i) the transverse relaxation rate induced by the isotropic nuclear magnetic shielding and (ii) induction by the electric field of the transverse component of the magnetization. A detailed description of these effects lies beyond the scope of this book, so the interested reader is referred to the source articles in which these effects are described.11,29 

Briefly speaking, if one takes into account the partial orientation of the sample under consideration due to the application of the electric field, then, unlike in the isotropic case, the isotropic part of the magnetic nuclear shielding tensor can lead to nuclear relaxation. Only the magnetization components that are perpendicular to the magnetic field B0 are affected by the electric field. The optimal conditions for the observation of such an effect are as follows: the magic angle between the electric and the magnetic fields, i.e., arccos ( 1 / 3 ) 54.7 ° , application of a strong, static electric field, i.e., the amplitude of E at least in the order of kVs per mm, and as small a ratio between the isotropic part of the shielding tensor and its anisotropy as possible. As a promising sample for such an experiment, a Pt(CN)6Tl complex containing a 205Tl nucleus was proposed.11 

The second effect arises when the nuclear magnetization returns from the direction that is opposite to the magnetic field.29  Then, the electric field E induces the transverse component of magnetization, which is not present in the absence of the electric field. The emergence of the additional transverse component stems from the disruption of the independence between the longitudinal and transverse nuclear relaxation processes, which is maintained in the absence of the externally applied electric field E. To increase the magnitude of the effect, E can be made to oscillate at one half of the spin precession frequency ω0. Optimally, the electric field is generated by the capacitor tilted by 45° with respect to the field B0, i.e., E1 = E1 sin(ω0t)(ey+ ez), and applied for a duration that is equal to the longitudinal relaxation time of the sample measured in the absence of the electric field E1, i.e., τ = T1. For B0 = 10 T, and E1 = 5 kV mm−1, the estimated amplitude of the effect for the 195Pt nucleus in the cisplatin molecule is 10 mHz, which corresponds to a signal approximately four orders of magnitude smaller than that obtained from the same sample in thermodynamic equilibrium after the application of a 90° pulse without the electric field E.

One may find higher-order effects induced by an external electric field in NMR spectroscopy, but the main limiting factor is a fast decrease in the amplitude of the effects with their order, since the factor μ e E k B T is much smaller then unity in the electric fields of strength that does not exceed the breakdown field in liquids (several kV mm−1). A similar remark applies to the comparison of the magnitudes of the effects, that are due to the coherent dynamics of nuclear spins, and the effects caused by nuclear relaxation – typically, the former have a much larger amplitude than the latter.

As mentioned in Sections 7.9 and 7.10, the interconnection between time-dependent electric and magnetic fields resulting from Maxwell’s equations is the main challenge faced in the experimental detection of chirality-sensitive NMER effects. Although an undesirable magnetic field manifests itself differently in each of the chirality-sensitive phenomena, a coherent description of how to minimize its influence on the NMER signals can be obtained for the analysis of the κ parameter. In the E 1 effect, the oscillating electric field generates an unwanted magnetic field, which results in saturation of the NMER signal, while in the E 2 effect, the source of the time-dependent magnetic field is the precessing nuclear magnetization. In both cases, the lowest possible value of κ parameter is desirable.

Any radiofrequency system that generates an electric field will generate a magnetic field oscillating at the same frequency. Hence, one cannot discriminate between them based on their time dependence. However, one can take advantage of the electric and magnetic fields spatial dependence and, consequently, at a given place, the electric field may dominate over the magnetic field. From this, it follows that the primary aim is to optimise the spatial distribution of E1 and B1 fields by minimizing the E1/B1 ratio and the angle between B1 and B0. For instance, ideally, a capacitor-like source of the electric field with plates perpendicular to the y-axis would have the following fields: E1(t)ey and B1(t)(x2 + z2)(ex − ez) in the sample volume. In practice, the construction of such a device is technically not possible. The main reason is that even a uniform radiofrequency electric field generates a circulating magnetic field, which increases in amplitude with the distance from the centre of the space containing the electric field. Moreover, it is hard to shield the region where the sample is placed from the magnetic field generated by the other parts of the resonance circuit. In the subsequent discussion, let us explore the different possible approaches to achieve a system generating an electromagnetic field distribution with a low value of the κ parameter. The optimal value of the κ parameter is, at most, one hundred thousandths in the volume of at least several cubic millimetres. The maximal acceptable value of the κ parameter depends on how precisely the phase of the NMR signal can be determined in the experimental search of the E 1 effect, alternatively how large is the single-to-noise if we aim for the detection of the E 2 effect. A large signal-to-noise ratio, e.g., S/N ∼ 1000, or sensitive phase detection, e.g., with a resolution better than 0.1°, would increase the maximum acceptable value of the parameter κ by two orders of magnitude, giving κ ≤ 0.001.

The close relationship between the properties of the sample and electromagnetic field distribution quantified by eqn (7.21) and (7.23), is illustrated in Figure 7.11.

Figure 7.11

The profile of the x component of the magnetic field (Bx), the y component of the electric field (Ey), and the reciprocal of the κ parameter. The loop-gap resonator has a radius of 40 cm; its plates are squares of 40 cm side length, and the height of the resonator is 40 cm. The fundamental mode of the frequency of f0 = 468 MHz is shown. The profile is taken along the orange line that lies in the middle between the plates. The points represent the voltage measured in the port located in half of the loop when the resonator is excited by an electric dipole oriented along the y-axis (open circles ○), a magnetic dipole oriented along the x-axis (stars, ★), and the ratio between these two voltages (open squares, □).

Figure 7.11

The profile of the x component of the magnetic field (Bx), the y component of the electric field (Ey), and the reciprocal of the κ parameter. The loop-gap resonator has a radius of 40 cm; its plates are squares of 40 cm side length, and the height of the resonator is 40 cm. The fundamental mode of the frequency of f0 = 468 MHz is shown. The profile is taken along the orange line that lies in the middle between the plates. The points represent the voltage measured in the port located in half of the loop when the resonator is excited by an electric dipole oriented along the y-axis (open circles ○), a magnetic dipole oriented along the x-axis (stars, ★), and the ratio between these two voltages (open squares, □).

Close modal

We will discuss the details of the resonator in Section 7.14.4 and now focus on the general relationship between the usage of the resonator as an electric field generator and as an electric polarization detector. Let us follow the orange line lying in the plane that divides the resonator in half. The region mainly generating the electric field lies between the plates. Therefore, it is a place favourable for locating the sample. Because of the unwanted magnetic field generated by the left part of the resonator, the sample will generate a signal that is a superposition of a chiral (due to the electric field E1) and achiral (due to the magnetic field B1) response. The κ-parameter for this electromagnetic field is the highest on the right side of the space where the magnetic field dominates, κ = c B y ( z ) E y ( z ) (the lines in Figure 7.11). Moreover, one finds that the ratio between the voltages generated by the chirality-independent nuclear magnetization (that acts as a source of the magnetic field) and the chirality-dependent electric polarization (that acts as a source of the electric field), takes exactly the same value, κ = U M ( z ) U P ( z ) (the points in Figure 7.11). Therefore, for finding the most suitable experimental hardware for observing the chirality-sensitive NMER effect, the κ parameter plays a crucial role. In particular, the objective is to find a system that is capable of generating a strong electric field at a given power level (equivalently, very sensitive to a weak excitation by electric polarization) and has a large region in which the κ parameter takes values which are as small as possible.

One of the possible systems that may create a radiofrequency electromagnetic field is a cavity, i.e., a closed space that entirely or in a part is occupied by the sample and surrounded by an electric conductor (usually metallic, e.g., copper). The spatial distributions of the electric and magnetic fields strongly depend on which cavity mode is excited. For instance, the TM010 mode in a cylindrical cavity of radius 0.25 cm and height 5 cm oscillates at 459 MHz, its electric field aligns along the axis of the cylinder, and the magnetic field circulates in the horizontal (perpendicular to that axis) plane. In contrast, the TE011 mode of that cavity oscillates at 734 MHz and has a magnetic field along the axis of the cylinder, and its electric field circulates in the horizontal plane. Thus, the electromagnetic fields of the TM010 and TE011 modes resemble the fields generated by a capacitor and a solenoid, respectively. Considering that the frequencies of the TM010 and TE011 modes of such a cavity are in the range of the proton and fluorine-19 resonance frequencies under magnetic fields of 10–20 T, and that the handy size of the device can possibly further be reduced due to a higher-than-unity dielectric constant of the sample, one can conclude that it fits the superconducting magnets perfectly, providing a highly homogenous magnetic field. In principle, one could place a sample on the cylinder axis and then use a static magnetic field B0 perpendicular to it. Next, aiming to observe the E 1 effect, one could electrically excite the sample using the TM010 mode, and detect the magnetic, chirality-sensitive response of the system by TE011. Alternatively, in the E 2 effect, the magnetic field of the TE011 mode would rotate the nuclear magnetization to the horizontal plane, and then the chirality-sensitive polarization would excite the TM010 mode of the cylindrical cavity. The advantage of using a cavity is its simple construction and fabrication. However, the drawback of this kind of experiment is that one has to change the magnetic field B0 considerably (to satisfy the resonance condition) in a time comparable with the relaxation time of the sample (a fraction to a few seconds). To some extent, this limitation could be overcome by means of a system that transfers the sample between magnetic fields of different magnitudes; however, this could happen at the expense of the mechanical stability of the system.

One could decrease the κ parameter of a cylindrical cavity by making a suitable modification to its shape. In particular, by reducing the ratio of the height to the radius of the cylinder and introducing circular cuts on the side surface of the cylinder, one obtains a region in the middle of the cavity in which the parameter κ is low. A cavity obtained in this way has a cross-section perpendicular to the cylinder axis that resembles a double-bladed axe head (a labrys). Such a modified cylindrical cavity is shown in Figure 7.12.

Figure 7.12

The spatial distribution of the electric field component perpendicular to the plane of the figure (Ez), the magnetic field B, and the κ parameter (from left to right) in the horizontal plane of a labrys resonator. The blue and red colours mark the electric field oriented in front of or behind the drawing plane. A magnetic field of high amplitude is shown in the red colour, while that of low amplitude is shown in the blue colour. The regions marked by red, yellow, and green correspond, respectively, to the κ-parameter values of 10−5, 10−4, and 10−3. The figure shows the fundamental mode of the resonator whose curvature radius on each side is 6.8 cm. The thickness of the resonator is 2.72 cm. The neckings of the cavities are 1.84 cm, 2.64 cm, and 5.30 cm, corresponding to the fundamental frequencies of 11.00 GHz, 7.56 GHz, and 3.74 GHz, respectively. Inside the cavity there is a vacuum. Field distributions were found by numerically solving the Maxwell equations with the aid of the finite-element method implemented in the COMSOL computer program.

Figure 7.12

The spatial distribution of the electric field component perpendicular to the plane of the figure (Ez), the magnetic field B, and the κ parameter (from left to right) in the horizontal plane of a labrys resonator. The blue and red colours mark the electric field oriented in front of or behind the drawing plane. A magnetic field of high amplitude is shown in the red colour, while that of low amplitude is shown in the blue colour. The regions marked by red, yellow, and green correspond, respectively, to the κ-parameter values of 10−5, 10−4, and 10−3. The figure shows the fundamental mode of the resonator whose curvature radius on each side is 6.8 cm. The thickness of the resonator is 2.72 cm. The neckings of the cavities are 1.84 cm, 2.64 cm, and 5.30 cm, corresponding to the fundamental frequencies of 11.00 GHz, 7.56 GHz, and 3.74 GHz, respectively. Inside the cavity there is a vacuum. Field distributions were found by numerically solving the Maxwell equations with the aid of the finite-element method implemented in the COMSOL computer program.

Close modal

In the case of the lowest mode of the labrys cavity, the electromagnetic field is mainly located around its centre: the electric field is in the middle, and the magnetic field circulates approximately along the side walls of the cavity. When the electromagnetic field oscillates in this mode, the charges periodically accumulate on its bases, while the surface current mostly flows along the sides of the cavity necking. The optimal orientation of the labrys cavity with respect to the magnetic field B0 is with the longer dimension aligned along the field. In such an orientation, the region where the parameter κ is minimal resembles a star. The value of the κ parameter mainly depends on how wide the necking of the cavity is and, to a small extent, on its thickness. According to the electromagnetic field distributions obtained from Maxwell equations solved using the finite-element method, the parameter κ may reach a value not exceeding 10−4 in a volume of a few cubic millimetres. In order to lower the resonance frequency of the cavity to the values commonly used in NMR spectroscopy, one could fill it not only with the chiral liquid under investigation, but also with a material with high dielectric permittivity. However, even in this case, the size of the device is quite large for the space available in most NMR spectrometers. Thus, considering its size, it fits better in the microwave frequency range and, therefore, is suitable for exploring chirality-sensitive EPR effects that are analogous to the NMER effects described in Sections 7.9–7.11. The other challenge is coupling the labrys cavity with excitation and detection systems. Inductive or capacitive coupling can be used for this purpose, i.e., a small plate or a coil whose electromagnetic field distribution fits the particular place of the cavity where the coupling element is placed, possibly at several places in the cavity to maintain its symmetry. It is noteworthy that the concept of using electromagnetic cavities in NMER strongly relies on the high quality of fabrication of such devices. Usually, once the cavity is fabricated, tuning its resonance frequency and correcting the spatial distribution of the electromagnetic field require further machining. Consequently, it can be difficult to achieve as high suppression of the magnetic field as that predicted by calculations.

In order to partially mitigate the high demands of a strictly defined shape of the resonance cavity, the solution found for the labrys cavity can be transformed into one that uses the so-called loop-gap resonator. The loop-gap resonator owns its name because, in the simplest configuration, it consists of a cylinder in which a gap is created by cutting its side surface in the longitudinal direction. The side surface of the cylinder generates a solenoid-like magnetic field oriented along the axis of the resonator. The electric field primarily occupies the slit and is directed perpendicularly to the magnetic field. Therefore, similar to the labrys cavity, the optimal configuration involves the axis of the resonator perpendicular to the static magnetic field B0 with the sample placed in the slit (Figure 7.13).

Figure 7.13

From the left to the right side: the classical loop-gap resonator (as used, e.g., in EPR applications), and three low-κ loop-gap resonators consisting of a parallel plate capacitor in which the sample is placed and (successively) two, four, and eight loops providing the spatial distribution of the κ parameter resembling of six-, ten-, and fourteen-arm star pattern, respectively. The electric field generated in the fundamental modes is horizontal and located mainly between the plates. The direction of the main magnetic field B0 is vertical.

Figure 7.13

From the left to the right side: the classical loop-gap resonator (as used, e.g., in EPR applications), and three low-κ loop-gap resonators consisting of a parallel plate capacitor in which the sample is placed and (successively) two, four, and eight loops providing the spatial distribution of the κ parameter resembling of six-, ten-, and fourteen-arm star pattern, respectively. The electric field generated in the fundamental modes is horizontal and located mainly between the plates. The direction of the main magnetic field B0 is vertical.

Close modal

This initial structure may be significantly improved by enlarging the slit to accommodate two parallel rectangular plates. Next, by increasing the symmetry of the resonator by introducing a second cylinder positioned so that its slit connects the plates, one obtains a low-κ loop-gap resonator. For suitably chosen proportions between the sides of the plates and the radius of the loops, there is a space in the middle, between the plates, for which the parameter κ is at most 10−4, which satisfies the optimization condition. The shape of the region for which the parameter κ is minimized is the same as for the labrys cavity. It is worth noting that one can view the low-κ loop-gap resonator as a system in which the circulating magnetic field generated by the radiofrequency electric field between the plates is compensated by the magnetic field of the loops, which circulates in the direction opposite to that of the field generated by the capacitor-like part of the resonator. The same is true for the labrys cavity: the horizontal components of the magnetic field generated by charges on the front and back walls of the cavity are cancelled by the circulating magnetic field generated by the flow of the surface current between the front and back walls of the cavity. As a result, instead of a magnetic field distribution of a parallel-plate capacitor of the vector field, (y2+ z2)(ez − ey), one has a time-dependent magnetic field, B1(t)yez, which is oriented in the same direction as the static magnetic field B0. This design of the low-κ loop-gap resonator also introduces an imposing inductive coupling – through four loops placed symmetrically at the junction between the loops and the capacitor plates. Similar to the labrys cavity, the low-κ loop-gap resonator does not preserve the cylindrical symmetry, so it requires a system capable of compensating for the relatively larger inhomogeneity of the static magnetic field. The region in which the parameter κ has sufficiently low values may increase (equivalently, the parameter κ can be further reduced) if one introduces more loops in the resonator structure. An example could be a pair of loops in the plane perpendicular to the static magnetic field and incorporated into the uppermost and the lowermost parts of the loops of the resonator. In this case, the higher degree of compensation of the unwanted magnetic field is directly visible from the shape of the low-κ region, which becomes a ten-arm instead of a six-arm star. The other arms of the star-like region arise from the compensation of the next spherical harmonic term in the unwanted, oscillating magnetic-field expansion about the centre point of the resonator (Figure 7.14). The reader interested in the rigorous mathematical description of the compensating field can find more details in ref. 17.

Figure 7.14

The spatial distribution of the κ parameter in the case where the size of the plates and loops of the resonator shown in Figure 7.9C are optimal resembles a ten-arm star (on the left side). The variation of the volume for which the κ parameter is less than 10−3 with the dimensions of the plate (xplate and zplate) is shown on the right. The white point in this graph represents an optimum that corresponds to a useful experimental volume of 240 mm3.

Figure 7.14

The spatial distribution of the κ parameter in the case where the size of the plates and loops of the resonator shown in Figure 7.9C are optimal resembles a ten-arm star (on the left side). The variation of the volume for which the κ parameter is less than 10−3 with the dimensions of the plate (xplate and zplate) is shown on the right. The white point in this graph represents an optimum that corresponds to a useful experimental volume of 240 mm3.

Close modal

A resonant system consisting of an inductor (coil) and a capacitor represents an entirely different experimental strategy than that described in Sections 7.14.3 and 7.14.4. In such systems, commonly used in NMR spectroscopy, the electromagnetic energy periodically accumulates in the electric field of the capacitive element and the magnetic field of the inductive element of the system. The most straightforward set-up would consist of rectangular capacitor plates embedded into the side surface of the cylinder and a saddle coil coaxially mounted outside or inside the cylinder. Matching the resonance frequency of such a system with the frequency that results from the fulfilment of the Larmor condition, i.e., ω0 = γNB0, is a challenging task. For instance, the capacitance of two 10 mm square plates on the side wall of a cylinder of 10 mm diameter is ∼1 pF, contributing to the total impedance of the system to a much lesser degree compared to a coil with a typical inductance of dozens of nH. One possible solution would be to divide the total capacitance of the system over two capacitors connected in parallel, i.e., with one of the capacitors excited by the electric field generated by the sample, and the other with variable capacitance. Another mutually non-exclusive solution would be to use two coils connected in series, i.e., one coil excites the sample with its magnetic field, and the other has a variable inductance. As both the capacitance and inductance used for the tuning circuits are low compared to parasitic capacitance and inductance, it is desirable to separate the resonance circuit of a capacitor and that of the coil. Hence, the choice between solutions with one or two resonant circuits is strongly influenced by the operating frequency. The advantage of using a one-circuit solution over a two-circuit solution is that the electric and magnetic fields must oscillate at the same frequency. However, the price one has to pay is the limitation of the maximum resonant frequency to about 200 MHz. It is worth noting that, while this system meets the experimental requirements of the observation of the E 2 effect, it would not be useful in the search for the E 1 effect. In the E 2 effect, it is sufficient to place the sample at the middle of the capacitor to achieve the suppression of the unwanted signal generated by the precession of the nuclear magnetic moment by providing a magnetic field distribution that, on average, vanishes over the sample volume. In contrast, any component of the time-dependent magnetic field that is perpendicular to the direction of the static field B0 will reduce the magnitude of the E 1 effect. Ensuring the appropriate direction of the magnetic field becomes even more challenging if one considers the electric and magnetic fields generated by the tuning parts of the capacitor and the coil surrounding the sample. Since the circuits operating at megahertz frequencies should be considerably smaller than the wavelength, e.g., 500 MHz corresponds to 60 cm, electromagnetic shielding seems more appropriate than increasing the distance between the tuning and the detection/excitation circuits. In summary, the inductor–capacitor circuit offer advantages over the labrys cavity and the loop-gap resonator in that it is much easier to tune, does not place such high demands on the shape control of the system, and its design follows well-established methods in the NMR community. However, its operating frequency is limited and determining the exact distribution of the electromagnetic field is not an obvious task because even between the parts on the excitation–detection system, the capacitor mounted on the cylinder and the saddle coil, there is undesirable but inevitable crosstalk.

The design of a system that will physically realize the desired electromagnetic field distribution, i.e., enables the creation of a field with the electric (magnetic) component perpendicular (parallel) to the static magnetic field, and allows sufficiently fast switching of the field of magnetic component perpendicular the static field, may be reformulated in terms of the electric charges and surface current density. Such design relies on particular solutions of Maxwell’s equations called Jefimenko’s equations.30,31  Jefimenko’s equations make the distribution of the electric and magnetic fields dependent on (i) the density of the electric current Js on the conductive surfaces, forming the system that generates the fields, and (ii) the surface density of the electric charge ρs that results from the continuity equation
(7.29)
Eqn (7.29) applies in the source-free parts of the system. Further discussion will be limited to the case of a system that oscillates at one particular frequency, which coincides with the spin precession frequency ω0. Typical values of the spin precession frequency ω0 rarely reach several GHz and, under these conditions, the system does not significantly generate electromagnetic waves, so one can use the quasistatic approximation. We will also make an assumption about the conducting surface thickness, which will not exceed the skin depth, which for copper is about 3 µm at 500 MHz.

Furthermore, for fixed conductive surfaces that constitute the system, the surface current distribution unambiguously gives the electromagnetic field distribution in the space occupied by the sample. The reverse, if one knows only the electric field spatial distribution, is not valid. In principle, there are infinitely many homogenous oscillating electric fields having various magnetic-field distributions in the vicinity of the sample. Therefore, a capacitor with plates where the divergence of the surface current ∇· Js is constant, with opposite sign on each plate, is generating a homogenous electric field. However, depending on how the sources and sinks of the electric charge are located, the surface current density varies, and the distribution of the magnetic field changes accordingly. In the simplest case, the plates are close to each other, so the electric field is homogenous, the sources and sinks are at the edges of the plates, and the divergence of the surface current is constant over the surface of the plates.

In order to find a unique Js current distribution, one can observe that (i) an irrotational (curl-free) vector field 1 2 ( x e x + y e y ) has a unit divergence, (ii) the field of a source/sink at the point (x0,y0) is a divergence of the function ln((x − x0)2+ (y − y0)2), and (iii) the component of the current Js that is perpendicular to the edge vanishes. Therefore, finding the Js current distribution involves finding a divergence-free vector field in a volume where the field is known on its boundary and this can be further reformulated as solving a partial differential equation.

Let us assume that a system for obtaining the desired electromagnetic field distribution consists of two rectangular parallel surfaces separated by a short distance as compared to their dimensions. If required, the surfaces can also be placed on the side surface of a cylinder, allowing us to partially maintain the cylindrical symmetry of the system. The introduction of curvature to the plates is a compromise between the uniformity of the static magnetic field, which in a typical case decreases rapidly if we move away from the axis of the magnetic field, maximizing the electromagnetic-field strength at a given excitation power. The main issue in the design of such a system is the choice of the position of the sources and sinks of the current and the place where the sample is. The minimal number of connections is two – a source (inlet) connection on one plate and a sink (outlet) connection on the other. The connections may be arranged in two ways: on the same or opposite sides of the plates. If they are on the same side, then the unwanted magnetic field generated by effects contributes to the magnetic field between plates, and obtaining the desired electromagnetic-field distribution is impossible. The surface current on the connectors is higher than on the plates because of their smaller surface than that of the plates. Therefore, the contribution of the magnetic field from the connectors may be as large or even larger than the magnetic field generated by the electric field between the plates (so-called displacement current). This limitation does not appear in the configuration where the plates are connected to opposite connectors (Figure 7.15, on the left side). In this case, the transverse components of the magnetic fields of the inlet and the outlet connectors partially cancel out. In particular, they are zero at the geometric centre of the system, which is the optimal place for the sample, which is assumed to be a small ball, i.e., a solid sphere.

Figure 7.15

The controllable surface current systems having two (on the left side) and eight (on the right side) connections to the plates.

Figure 7.15

The controllable surface current systems having two (on the left side) and eight (on the right side) connections to the plates.

Close modal

It is worth noting that the connections in the opposite configuration only make sense if the connectors are electromagnetically shielded, e.g., cores of coaxial cables, since it is not physically possible to have two sources (one positive and one negative) that do not share a common ground (if we are aiming for observation of the E 1 effect) and if we, instead, consider the E 2 effect, the difference between the voltages on the plates has to be compared in some way, e.g., measured separately with respect to the common ground.

The solution with two connectors would be practically applicable only in the case of the E 2 effect because it provides an electromagnetic-field distribution that, when averaged out over the sample volume, has a vanishing transverse component. In the E 1 effect, the estimated degree of chirality-sensitive signal saturation is too high to make such an arrangement promising for the observation of this effect. However, to facilitate the search for the E 1 effect using a system based on the controllable surface current concept, one can increase the number of connectors. The optimal number may vary depending on the advantages of introducing each pair connector and the difficulties in operating several radiofrequency sources simultaneously at the same frequency, e.g., due to finite directness on the amplifiers. The additional connections may be introduced as inlets/outlets or connections to the ground. The two-connector system described in this section is a balanced circuit since it does not have any explicit ground connection (the ground is a shield surrounding the plates), and both plates are equivalent. For systems with more than two connectors, it is convenient to use the unbalanced configuration in which one plate is grounded, and the power is delivered by several connectors distributed at the edges of the ungrounded plate. While the four-connector system, i.e., two connectors on the top and the bottom of a plate and the other two on the sides of the plate, does not provide any significant benefits over the solution based on two connectors, the six-connector system has a new advantageous feature. The two pairs of connectors placed on the sides of the plate may effectively reorient the transverse time-dependent magnetic field component along the static magnetic field, consequently enabling us to search for the E 1 effect. The other advantage is that one can, by suitably chosen amplitudes and phases on the inlets, not only compensate for the unwanted oscillating magnetic field but also adjust the direction for which the compensation is most effective, to the direction of the static magnetic field.

A further improvement of this system is the introduction of the next two connections; thus, the system has eight connections in total (see Figure 7.14, on the right side). The eight-connector system gives access to the sample along the axis in the vertical symmetric plane of the system, which makes it possible to replace the ball shape of the sample with a long cylindrical shape, resulting in an easier sample replacement. However, using more than eight connectors seems to offer no significant advantages besides making the system more complicated.

Although the systems described in Sections 7.14.2–7.14.6 represent different solutions to the challenge of achieving sufficiently low values of the κ parameter, they share a common feature with respect to the shape of the region where the κ parameter is minimal, i.e., a star-like shape. This is a consequence of the circular magnetic field created by the homogenous electric field (equivalently, by the displacement current). The vertical component of the magnetic field is compensated (or reoriented, since complete suppression of the magnetic field would violate Maxwell’s equations) by additional two elements of the system. These elements may be cavity walls, additional loops, or connectors attached to the capacitors’ plates. To some extent, this type of compensation is similar to the method of obtaining a uniform field in superconducting magnets. This method assumes that any sufficiently small perturbation of the magnetic field in the vicinity of a chosen point can be expanded into a power series, and each term in this expansion may be compensated by a coil that generates an opposite magnetic field with the spatial distribution given by the term. Therefore, one can view the procedure of obtaining a region of low κ-parameter value as the first-order compensation involving a coil generating a horizontal oscillating magnetic field that linearly varies with height, i.e., along the direction of the static magnetic field B0. By compensating for higher-order terms in the series expansion of the time-dependent horizontal magnetic field, the volume exhibiting a small κ parameter is expanded, resulting in a star-like region with more arms, e.g., the first-order compensation has a six-arm pattern, whereas the second-order compensation has a ten-arm pattern. The proposed solutions differ mainly in terms of the tuning range, the requirements for accuracy, and the range of resonance frequencies in which they can in practice be used. In particular, electromagnetic cavities and loop-gap resonators are characterized by the range of high frequencies and high requirements for precision with a small tuning range. Inductor-capacitor circuits offer great tunability up to ∼200 MHz and relatively low machining requirements. However, achieving a predictable electromagnetic-field distribution seems challenging. Systems providing controllable surface current lie somewhere between these two extremes and combine the advantage of a wide range of tuning and controllable electromagnetic-field distribution, but pose significant requirements for obtaining the assumed amplitude and phase at a given point in the system, which can be a difficult condition to meet. The division between the various solutions described in this chapter is not rigid. For instance, the electromagnetic field between the plates of the loop-gap resonator may be satisfactorily explained through an analysis of the surface current that flows first in the outer part of the plates and then generates an electric charge on the inner surface of the plates. Therefore, taking into account both advantages and limitations of each of the described systems in this chapter, it can be presumed that each of them can be used in a certain range of experimental conditions, the most important of which are the size of the area in which the static magnetic field is uniform and the precession frequency of nuclear spins, since these two factors mainly affect the NMR signal strength. Apart from these two imposing factors, one needs to consider the strength of the oscillating electric field generated by the system. However, dielectric heating is probably more practically limiting than the power that an RF amplifier could deliver.

The introduction of an additional external electric field in an NMR experiment may allow direct, non-invasive identification of enantiomers, including determination of the absolute configuration with the aid of low-level, e.g., fast and feasible quantum-chemical computations, even for relatively large molecules. From the theory point of view, this is an unexpected result since it offers an entirely different way to detect molecular chirality using NMR than the previously proposed methods in the literature. The most striking outcome of applying tensor analysis to examine interactions of nuclear magnetic moments is that the antisymmetric parts of such tensors are inherently linked to the molecular chirality due to the symmetry properties, i.e., the antisymmetries are pseudovectors. Detailed analysis of the nuclear spin dynamics in a time-dependent oscillating, electric field and a static magnetic field reveals the existence of at least three different effects, upon which a new branch of NMR spectroscopy, called nuclear magnetoelectric spectroscopy, might be based. Although the proposed effects have not been observed yet, several promising samples with favourable properties for initial experiments have been proposed using quantum-chemical methods, and a few systems capable of generating strong and detecting weak electric fields have been reviewed and described in detail in this chapter. It can be concluded that the theoretical description of the dynamics of nuclear spins in chiral particles placed in a magnetic field, and the description of molecular parameters on which these effects depend, are currently satisfactory. Observing NMER phenomena becomes rather a technical challenge. Meeting the high requirements for the apparatus used to perform the experiment is a challenging task, since the experimental set-up should possess very high sensitivity according to NMR spectroscopy standards, and additional elements related to the electric field are designed in a way that does not follow the commonly used solutions in NMR spectroscopy. Despite these difficulties, NMER spectroscopy seems to constitute a promising field of research as it offers the potential for gaining new insights into the structure and dynamics of many molecules of great importance to chemistry and biology.

PG acknowledges the European Research Council for the financial support through the ERC Starting grant (project acronym: NMER, agreement ID: 101040164), the Max Planck Society for providing access to computational resources (in particular, the COMSOL program), Prof. Juha Vaara (University of Oulu) for carefully reading the chapter and giving extensive feedback, and Dr Katarzyna Grabowska (Faculty of Physics, University of Warsaw) for helpful discussions about tensor transformation properties viewed from the perspective of a mathematical physicist.

This chapter is subject to a Creative Commons CC-BY-NC-ND 4.0 International license. Financial support from the European Research Council through an ERC Starting Grant (project acronym: NMER, agreement ID: 101040164) is acknowledged.

The rank of a tensor T is the minimum number of simple tensors that sum up to T ; see ref. 5 and 6 for details.

§

The determinant of matrix R is +1.

Notice the reversed sign of the Levi-Civita tensor components in the left-handed Cartesian coordinate system.

One can obtain this result by considering the 2-index tensor vol(v, ·, ·), where vol ( v , u , w ) = det ( v x u x w x v y u y w y v z u z w z ) and v , u , w 3 . It is worth noting that, vol(v, u, w) = v · (u × w) and using eqn (7.2) one finds that, in the right-handed Cartesian coordinate system, [ vol ( v , , ) ] = ( 0 v z v y v z 0 v x v y v x 0 ) , thus it has the same form as given in eqn (7.7a), which corresponds to a pseudovector.

**

Chirality-sensitive phenomena in paramagnetic chiral molecules are discussed in ref. 9 and 10.

1
Saitô
H.
,
Ando
I.
,
Ramamoorthy
A.
,
Prog. Nucl. Magn. Reson. Spectrosc.
,
2010
, vol.
57
pg.
181
2
Wenzel
T. J.
,
Wilcox
J. D.
,
Chirality
,
2003
, vol.
15
pg.
256
3
Buckingham
A. D.
,
Pyykkö
P.
,
Robert
J. B.
,
Wiesenfeld
L.
,
Mol. Phys.
,
1982
, vol.
46
pg.
177
4
D. A.
Varshalovich
,
A. N.
Moskalev
, and
V. K.
Khersonskii
,
Quantum Theory of Angular Momentum
,
World Scientific
,
Singapore
,
1989
.
5
N.
Bourbaki
,
Elements of Mathematics
,
Algebra I
,
Herrmann
,
1974
, pp.
309
310
.
6
W.
Hackbusch
,
Tensor Spaces and Numerical Tensor Calculus
,
Springer
,
2019
, pp.
3
8
.
7
H. W.
Spiess
, Rotation of molecules and nuclear spin relaxation, in
Dynamic NMR Spectroscopy
,
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
1978
, pp.
55
214
.
8
T.
Needham
,
Visual Differential Geometry and Forms
,
Princeton University Press
,
Princeton
,
2021
, pp.
376
378
.
9
Soncini
A.
,
Calvello
S.
,
Phys. Rev. Lett.
,
2016
, vol.
116
pg.
163001
10
Calvello
S.
,
Soncini
A.
,
Phys. Chem. Chem. Phys.
,
2020
, vol.
22
(pg.
8427
-
8441
)
11
Garbacz
P.
,
J. Chem. Phys.
,
2016
, vol.
145
pg.
064202
12
Garbacz
P.
,
J. Chem. Phys.
,
2016
, vol.
145
pg.
224202
13
Barra
A. L.
,
Robert
J. B.
,
Wiesenfeld
L.
,
Europhys. Lett.
,
1988
, vol.
5
pg.
217
14
Robert
J. B.
,
Barra
A. L.
,
Chirality
,
2001
, vol.
13
pg.
699
15
Dale
J. A.
,
Dull
D. L.
,
Mosher
H. S.
,
J. Org. Chem.
,
1969
, vol.
34
9
pg.
2543
16
Buckingham
A. D.
,
Chem. Phys. Lett.
,
2004
, vol.
1
pg.
398
17
Garbacz
P.
,
Fischer
P.
,
Krämer
S.
,
J. Chem. Phys.
,
2016
, vol.
145
pg.
104201
18
R. R.
Ernst
,
G.
Bodenhausen
and
A.
Wokaun
,
Principles of Nuclear Magnetic Resonance in One and Two Dimensions
,
Clarendon Press
,
Oxford
,
1997
, pp.
9
69
.
19
F. J. M.
van de Ven
,
Multidimensional NMR in Liquids
,
VCH
,
1995
, pp.
49
98
.
20
A.
Abragam
,
The Principles of Nuclear Magnetism
,
Oxford, London
,
1961
, p.
276
.
21
Goldman
M.
,
J. Magn. Reson.
,
2001
, vol.
149
pg.
160
22
Kuprov
I.
,
Wagner-Rundell
N.
,
Hore
P. J.
,
J. Magn. Reson.
,
2007
, vol.
184
pg.
196
23
Blanes
S.
,
Casas
F.
,
Oteo
J. A.
,
Ros
J.
,
Phys. Rep.
,
2009
, vol.
470
pg.
151238
24
Buckingham
A. D.
,
J. Chem. Phys.
,
2014
, vol.
140
pg.
011103
25
Buckingham
A. D.
,
Fischer
P.
,
Chem. Phys.
,
2006
, vol.
324
pg.
111
26
Walls
J. D.
,
Harris
R. A.
,
Jameson
C. J.
,
Chem. Phys.
,
2008
, vol.
128
pg.
154502
27
Walls
J. D.
,
Harris
R. A.
,
J. Chem. Phys.
,
2014
, vol.
140
pg.
234201
28
Garbacz
P.
,
Cukras
J.
,
Jaszuński
M.
,
Phys. Chem. Chem. Phys.
,
2015
, vol.
17
pg.
22642
29
Garbacz
P.
,
J. Chem. Phys.
,
2018
, vol.
148
pg.
034501
30
Jefimenko
O. D.
,
Am. J. Phys.
,
1990
, vol.
58
pg.
505
31
Griffiths
D. J.
,
Heald
M. A.
,
Am. J. Phys.
,
1991
, vol.
59
(pg.
111
-
117
)
Close Modal

or Create an Account

Close Modal
Close Modal