- 8.1 Introduction
- 8.2 Chirality-sensitive Spin–Spin Coupling
- 8.3 Relationship Between Chirality and Antisymmetric Spin–Spin Coupling
- 8.4 Time-evenness of the Pseudoscalar Jc
- 8.5 Relationship Between Chirality and the Permanent Electric Dipole Moment
- 8.6 Relationship Between the Molecular Structure and Pseudoscalar Jc
- 8.7 Nuclear Magnetoelectric Effective Hamiltonian
- 8.7.1 Rotating Frame Transformation
- 8.7.2 Antisymmetric Nuclear Magnetic Shielding Effective Hamiltonian
- 8.7.3 Antisymmetric Spin–Spin Coupling Effective Hamiltonian
- 8.8 A General Description of the Chirality-sensitive Electric Polarization
- 8.9 Chirality-sensitive Coherences Induced by the Antisymmetric Part of the J-coupling Tensor
- 8.9.1 Description of the
- 8.9.2 Experimental Protocol for the
- 8.9.3 Significance and Estimation of the Magnitude of the
- 8.10 Electric Polarization Induced by the Antisymmetric Part of Indirect Spin–Spin Coupling
- 8.10.1 Description of the
- 8.10.2 Experimental Protocol for the
- 8.10.3 Significance and Estimation of the Magnitude of the
- 8.11 Chirality-sensitive Effects Induced by Antisymmetry of the A-coupling Tensor
- 8.11.1 Description of the
- 8.11.2 Experimental Protocol for the
- 8.11.3 Significance and Estimation of the Magnitude of the
- 8.12 Interference Between Dipolar Relaxation Mechanisms in an Electric Field
- 8.12.1 Description of the
- 8.12.2 Experimental Protocol for the
- 8.12.3 Significance and Estimation of the Magnitude of the
- 8.13 Sources of Radiofrequency Electric Field for J-NMER Experiments
- 8.14 Summary and Conclusions
- References
Chapter 8: Chirality-sensitive Effects Induced by Antisymmetric Spin–Spin Coupling†
-
Published:20 Sep 2024
P. Garbacz and J. Vaara, in Physical Principles of Chirality in NMR, ed. P. Garbacz, Royal Society of Chemistry, 2024, vol. 34, ch. 8, pp. 204-242.
Download citation file:
The chapter introduces the concepts of direct and indirect spin–spin coupling as well as hyperfine spin–spin coupling. It is shown that a chiral molecule with a spin pair coupled by an antisymmetric interaction carries an induced electric dipole moment whose direction depends on the handedness of the molecule. Finding this induced dipole, in turn, serves as the starting point for the derivation of the nuclear magnetoelectric resonance Hamiltonian and analysis of spin dynamics under the influence of an externally applied electric field. Four effects that enable direct chirality sensing in nuclear magnetic resonance spectroscopy are also described with emphasis on the relationships between chirality, antisymmetry of spin–spin coupling, electric dipole moment, and molecular structure. Several capacitive devices capable of generating an electric field with properties suitable for nuclear magnetoelectric resonance experiments, in which the central role is played by spin–spin couplings, are described.
- Award Group:
- Funder(s): National Science Centre
- Award Id(s): 2018/31/B/ST4/02570
- Funder(s):
- Award Group:
- Funder(s): Research Council of Finland
- Award Id(s): 331008
- Funder(s):
- Award Group:
- Funder(s): University of Oulu
- Funder(s):
- Funding Statement(s): Financial support from the National Science Centre, Poland (through OPUS 16 grant No. 2018/31/B/ST4/02570), Academy of Finland (project 331008) and University of Oulu is acknowledged
8.1 Introduction
The mathematical framework, theoretical depiction, and relevant electronic properties required for the description of the interactions between spins of a chiral molecule, are closely related to those mediated by nuclear shielding, as described in Chapter 7. Therefore, in this chapter, we will refer quite often to the results obtained in the previous one, and consequently reading the previous chapter before the current one is recommended.
Many fundamental particles and nuclei have magnetic moments that can interact with each other, so depending on the context, one may name these interactions differently. For instance, in the domain of nuclear magnetic resonance (NMR) spectroscopy, the interactions between two nuclear magnetic dipole moments are called spin–spin coupling. In contrast, electron paramagnetic resonance (EPR) spectroscopists prefer to use the term hyperfine coupling for the interaction between the magnetic dipole moments of an electron and a nucleus or interaction between the magnetic dipole moments of two electrons.1,2 These two terms are frequently used, although both cases involve interactions between the nuclear and electron magnetic moments resulting from their spins. These are hyperfine interactions, i.e., they take place between the magnetic moments and magnetic fields in the molecule. Hereafter, the term “dipole” will be left out in examples where its omission will not introduce ambiguity.
Similarly, we will follow the common convention and use the terms spin and magnetic moment interchangeably, although the magnetic moment is perceived as a concept of classical physics and the spin is considered as a purely quantum-mechanical entity; their correspondence is expressed by the equation , where μI is the magnetic moment expressed in Bohr magnetons for electrons, μB = 9. 274 009 994 × 10−24 J T−1 or nuclear magnetons for nuclei μN = 5.0 507 837 461 × 10−27 J T−1, gI is its g-factor (dimensionless), and is the spin operator (a dimensionless vector of Pauli matrices, in the case of a one-half spin). Alternatively, the proportionality constant between the magnetic moment and the spin operator may be expressed using the gyromagnetic ratio γ (in rad T−1 s−1); , where ℏ = 1.054571817 × 10−34 J s rad−1 is the reduced Planck constant, and the magnetic moment μI is expressed in J T−1.
Two spins may interact in at least two ways. One is direct coupling, which, from the classical physics viewpoint is an interaction between two magnetic dipoles through their magnetic field.3,4 The other is an indirect coupling resulting from the influence of the magnetic field of the magnetic dipole on the molecule’s electrons. The distinction between direct and indirect spin–spin couplings lies in the mediating contribution of the electrons of the molecule, rather than in the presence or absence of a chemical bond. For example, through-space contributions to indirect spin–spin coupling between 19F nuclei due to two lone-pair orbitals on intramolecularly crowded molecules are relatively common.5 In EPR spectroscopy, unlike the convention used in NMR, both interactions are considered together as one interaction.
8.2 Chirality-sensitive Spin–Spin Coupling
Eqn (8.9) can be directly related to Section 1.2, which discusses the relationship between the chiral tetracube handedness and the reference frame in which tetracube’s parameters are described. In eqn (8.9), the indirect spin–spin coupling tensor J corresponds to the parameter specifying the handedness of the tetracube, and information about the handedness of the Cartesian coordinate system is provided by the Levi-Civita tensor. The components of the Levi-Civita tensor for the left-handed Cartesian coordinate system have the same magnitude but opposite signs compared to those given in eqn (7.8) for the right-handed system. Consequently, the sign of the pseudoscalar for an (R)-enantiomer in a right-handed coordinate system is the same as that of an (S)-enantiomer considered in a left-handed coordinate system. For left-handed and right-handed molecules placed in either the left or to the right-handed reference frame, the signs of the pseudoscalar are opposite, analogous to the situation shown in Figure 1.2 for the chiral tetracube.
In terms of the dielectric properties of liquids, the contribution appears because of the partial orientation of molecules bearing a permanent electric dipole moment under the influence of an externally applied electric field. The partial orientation prevents the complete averaging out of the anisotropic parts of the indirect spin–spin coupling tensor J between spins I1 and I2 observed in an isotropic phase.
8.3 Relationship Between Chirality and Antisymmetric Spin–Spin Coupling
8.4 Time-evenness of the Pseudoscalar Jc
8.5 Relationship Between Chirality and the Permanent Electric Dipole Moment
As clearly seen from considerations given in this chapter up to now, as well as in Chapter 7, the molecular permanent electric dipole moment, μe, plays a crucial role in the occurrence of nuclear magnetoelectric resonance (NMER) chirality-sensitive effects. Based on the symmetry arguments given in Section 8.3 and the quantum-mechanics view on chirality given in Section 1.4, we are now in a good position to address the discrepancy in views on μe within the communities of chemists and physicists.
The only vector degree of freedom for a fundamental particle (such as an electron) is its spin and, consequently, the possible electric dipole moment d of the particle may only be oriented either parallel or antiparallel to the spin .18–20 Consequently, the dipole moment has to be proportional to the spin, , and the Hamiltonian of its interaction with an electric field E takes the form . However, under space inversion and time reversal operations, such a Hamiltonian reverses its sign, i.e., equals, and , respectively.21 Therefore, an elementary particle possessing both the spin and electric dipole moments violates P-even and T-even symmetries expected for the Hamiltonian . Measurement for electric dipole moments of electrons,22 neutrons,23 and molecules24 indicate that the electric dipole moment is either zero or immeasurably small; see also ref. 25 for an extensive review of the performed experiments. The same argument can be used for molecules since they are composed of elementary particles.
This observation could raise reasonable doubts regarding the arguments put forward that NMER phenomena (and many others) would allow chirality sensing. On the other hand, the use of molecular electric dipole moments in the chemistry community is very common,26,27 and their existence is frequently considered as an indisputable experimental fact.
So, can molecules have permanent electric dipole moments? The discrepancy in nomenclature results largely from how we define a “molecule”: an object resembling a rigid body embedded in a cloud of diffuse electric charge (the chemical perspective) or an object described by a wave function, consisting of elementary particles that constitute it (the physical perspective). To answer this question, stated and discussed by Klemperer et al. in ref. 28, we have to look deeper into the quantum-mechanical description of a chiral molecule.
If the nuclear spin Hamiltonian is assumed to be parity-conserving, then the energies of both enantiomers are the same. This does not mean, however, that the parity of any initial state of the molecule is preserved over a arbitrarily long period of time. Eigenstates of the Hamiltonian and have, indeed, parity +1 and −1, respectively (see Section 1.4). Depending on the rate of interconversion between the enantiomers, the wavefunction |ψR〉 of a chiral molecule successively evolves in time into |ψ+〉, |ψS〉, −|ψ−〉, … , and returns to the initial wavefunction |ψR〉 (Figure 8.1). The period of the time evolution of the quantum state of the chiral molecule can vary significantly depending on the molecule under consideration, ranging from fractions of a second for hydrogen peroxide29,30 to practically unmeasurably long durations, such as for solid alanine.
The time evolution of the quantum state of a chiral molecule, , in terms of its parity and handedness. On the red axes, the probabilities of two “racemic” states are shown: one of the positive parity, |〈ψ|ψ+〉|2 (the horizontal axis) and the other of the negative parity, |〈ψ|ψ−〉|2 (the vertical axis). The axes rotated by 45° represent probabilities of finding a molecule in a state attributed arbitrarily to the (R)-enantiomer, , and the (S)-enantiomer, . The higher the energetic barrier, the slower the state of the molecule rotates in the plane, resulting in a smaller uncertainty of enantiomeric excess δee in a given time interval δt. The clockwise sense of rotation, assumed in the figure, is also arbitrarily chosen, since it depends on the handedness of the chosen coordinate system.
The time evolution of the quantum state of a chiral molecule, , in terms of its parity and handedness. On the red axes, the probabilities of two “racemic” states are shown: one of the positive parity, |〈ψ|ψ+〉|2 (the horizontal axis) and the other of the negative parity, |〈ψ|ψ−〉|2 (the vertical axis). The axes rotated by 45° represent probabilities of finding a molecule in a state attributed arbitrarily to the (R)-enantiomer, , and the (S)-enantiomer, . The higher the energetic barrier, the slower the state of the molecule rotates in the plane, resulting in a smaller uncertainty of enantiomeric excess δee in a given time interval δt. The clockwise sense of rotation, assumed in the figure, is also arbitrarily chosen, since it depends on the handedness of the chosen coordinate system.
In the case of eigenstates |ψ+〉 and |ψ−〉, the (permanent) electric dipole moment of the molecule vanishes. It is precisely these (pure) states that the conclusions drawn on the basis of symmetry arguments, rooted in physical considerations, are concerned with. From a chemist’s point of view, however, it is more useful to consider the mixed states |ψR〉 and |ψS〉, which are interpreted as enantiomers. In these cases, the molecule actually has an electric dipole moment. We note that the electric field, unlike the magnetic field, mixes molecular states of opposite parities; hence, a linear shift in levels (Zeeman effect) in the presence of a magnetic field is experimentally observed, whereas there is no evidence for the first-order Stark effect in a molecule in its eigenstate. This does not preclude the observation of a first-order Stark effect for a mixed state induced by an external electric field, reported routinely in rotational spectroscopy.31–34 In such a case, from the perspective of molecular eigenstates, the effect of the electric field on the molecule can be more precisely called a second-order Stark effect.35
In the diagram shown in Figure 8.1, the points corresponding to the states |ψ+〉 (parity +1) and |ψ−〉 (parity −1) do not have any electric dipole moment μe, although the antisymmetry J* is non-zero. Consequently, the pseudoscalar Jc vanishes. For convenience, let us assume that the antisymmetry pseudovector points upwards. Since the antisymmetry is P-even, its direction is the same for all states shown in the diagram. The states of the molecule, except for points on the horizontal and vertical axes, do have a dipole moment. If the states in the half-circle above the horizontal axis have the electric dipole moment pointing upwards, resulting in a positive pseudoscalar Jc, then the states in the half-circle below the horizontal axis have the electric dipole moment pointing downwards, yielding a negative pseudoscalar Jc. It is worth noting that, from the assumption of the parity conservation of the Hamiltonian, it follows that neither left-handed nor right-handed coordinate systems are preferred, so the direction of the circulation in Figure 8.1 is arbitrary and depends on the choice of the handedness of the coordinate system. If this were not true, e.g., we could define on the basis of fundamental laws that the enantiomer that first changes the state to that of positive parity |ψ+〉 is designated as the (R)-enantiomer.
8.6 Relationship Between the Molecular Structure and Pseudoscalar Jc
Similar to the antisymmetric part of the shielding tensor σ*, the direction and amplitude of the J* antisymmetry are related to molecular symmetry. The relationships between nuclear site symmetry and the components of the antisymmetric part of the indirect spin–spin coupling tensor have been discussed by Buckingham and his co-workers in ref. 36 and 37. Bryce and Wasylishen analyzed the symmetry properties of indirect nuclear spin–spin coupling tensors of simple fluorine-containing molecules, CFCl3 and F2O.38 An extensive survey of small and medium-sized molecules, with emphasis on the relationship between the molecular structure and directions of the permanent electric dipole moment μe and the antisymmetry J*, can be found in ref. 39. Despite the generally accepted existence of the antisymmetric part of the J tensor resulting from strong theoretical foundations,40 this quantity has not yet been measured experimentally, although experiments aimed at measuring it have been reported.41,42
Water is one of the simplest examples of a molecule with an antisymmetric indirect spin–spin coupling. The Fermi contact, spin–dipole, paramagnetic and diamagnetic spin–orbit contributions to the total J-coupling tensor between the protons and oxygen are listed in Table 8.1.
Indirect spin–spin coupling of water (the molecule is in the yz-plane) computed in the CFOUR program using the coupled cluster singles and doubles (CCSD) method in the aug-pcJ-2 basis set; the initial molecular geometries were taken from Computational Chemistry Comparison and Benchmark Data Base and further optimized at the CCSD/aug-pCVTZ level. Tensor elements that contain antisymmetric parts are boldfaced.
. | 1J(17O, 1Ha) . | 1J(17O, 1Hb) . | 2J(1Ha, 1Hb) . |
---|---|---|---|
Fermi-contact and correlated FC-SD | |||
Spin–dipole | |||
Paramagnetic spin–orbit | |||
Diamagnetic spin–orbit | |||
Total indirect spin–spin coupling |
. | 1J(17O, 1Ha) . | 1J(17O, 1Hb) . | 2J(1Ha, 1Hb) . |
---|---|---|---|
Fermi-contact and correlated FC-SD | |||
Spin–dipole | |||
Paramagnetic spin–orbit | |||
Diamagnetic spin–orbit | |||
Total indirect spin–spin coupling |
In contrast to the other contributions, the Fermi-contact contribution is isotropic, so it does not contribute to the antisymmetry J*. We observe that different contributions may cancel out. For instance, the paramagnetic and diamagnetic terms of 1J(17O, 1Ha) and 1J(17O, 1Hb) couplings are almost equal in magnitude, but their signs are opposite. The amplitudes of total antisymmetries derived at the coupled-cluster level of the theory are 1J*(17O, 1H) = 0.16 Hz and 2J*(1H, 1H) = 1.85 Hz.
The orientation of the antisymmetry with respect to the molecular frame of reference can be deduced from the fact that the molecule is planar. Thus, the only non-vanishing component of the antisymmetry J* must be perpendicular to the plane in which the water molecule lies. As expected, for a non-chiral molecule such as water, the antisymmetry J* is perpendicular to the dipole moment μe and the pseudoscalar Jc is zero. From the C2v symmetry of the water molecule, it follows that the directions of antisymmetric components of the couplings 1J(17O, 1Ha) and 1J(17O, 1Hb) are opposite each other. These structural relationships are also observed in larger molecules of symmetry close to C2v.
Let us take as an example of a molecule of norborn-5-en-2,3-diol shown in Figure 8.2. The two pairs of protons, one methylene (marked by the red color) and the other vicinal (marked by the blue color) are located in quite distinct regions of the molecule. The direction of the antisymmetry 2J*(1H, 1H) for the coupling between the methylene protons is perpendicular to the CH2 plane, aligning with the expected direction based on that for the protons of water. Moreover, the amplitude is comparable to the case of protons of water: 2J*(1H, 1H) = 1.08 Hz. On the other hand, a simple reasoning based only on the analyzed fragment may not give the correct direction of antisymmetry. An example is the antisymmetry 3J*(1H, 1H) for the coupling between the vicinal protons, which approximately lies in the HC═CH plane rather than perpendicular to it, as observed in ethene.
From left to right: molecular structures of norborn-5-en-2,3-diol (top and side views), water, and ethene-d2. The electric dipole moments μe (green arrows), as well as the antisymmetric parts of the proton–proton indirect spin–spin coupling tensors J* (blue and red arrows) are shown in the bottom row. The lengths of these vectors are not to scale. The directions of the vectors are based on the results of density-functional theory computations. Optimal geometries were obtained using the Turbomole computer program43 using the PBE0+D3(BJ) functional44–46 with the def2-QZVPP basis, and spin–spin couplings were computed in the Dalton computer program47 using the PBE0 functional with the aug-pcJ-2 basis.48
From left to right: molecular structures of norborn-5-en-2,3-diol (top and side views), water, and ethene-d2. The electric dipole moments μe (green arrows), as well as the antisymmetric parts of the proton–proton indirect spin–spin coupling tensors J* (blue and red arrows) are shown in the bottom row. The lengths of these vectors are not to scale. The directions of the vectors are based on the results of density-functional theory computations. Optimal geometries were obtained using the Turbomole computer program43 using the PBE0+D3(BJ) functional44–46 with the def2-QZVPP basis, and spin–spin couplings were computed in the Dalton computer program47 using the PBE0 functional with the aug-pcJ-2 basis.48
A similar structural relationship between the direction of the antisymmetry of the spin–spin coupling is observed in a much bigger molecule, a chiral derivative of 1,3-bisdiphenylene-2-phenylallyl radical, the so-called the Koelsch radical,49 frequently abbreviated as BDPA, shown in Figure 8.3. Since the molecule is a radical, the dominating antisymmetric coupling between spins originates from the coupling with the unpaired electron. The highest probability of finding an unpaired electron is in the middle of the molecule, near the carbon atoms joining three groups of rings (two of the tricyclic aromatic hydrocarbon – fluorene, and one of the phenyl substituent). Therefore, the strongest coupling with the electron is exhibited by, among others, the two adjacent carbon nuclei of the fluorenes. The directions of the antisymmetric components of their hyperfine coupling tensors (blue and red arrows in Figure 8.3) are, to a good approximation, not influenced by the labile bond that allows the alanine substituent bridged by the square acid unit to rotate. In contrast, the direction of the permanent electric dipole moment μe of the molecule depends strongly on the conformation that the molecule adopts. If the molecule adopts conformation A shown on the left side of Figure 8.3, the angle between the vectors μe and A* is 47° for the coupling with one of the carbon-13 nuclei, while 127° is obtained for the other. However, if the molecule adopts conformation B shown on the right side of Figure 8.3, the angles between the vectors μe and A* become close to the right angle (94° and 102°, respectively), and the magnitude of the pseudoscalar Ac is, therefore, greatly reduced. It is worth noting that, similar to antisymmetries 1J*(17O, 1H) of water, the antisymmetries of the hyperfine coupling of these two 13C nuclei exhibit opposite directions in the case of the chiral derivative of BDPA shown in Figure 8.3, resulting in opposite signs of the pseudoscalars Ac of the two carbons [Ac is defined in an analogous way to Jc – see eqn (8.11)].
The structures of two dominating conformers of 1,3-bisdiphenylene-2-phenylallyl radical functionalized with a square acid and alanine. The permanent electric dipole moments μe (green arrows) and the antisymmetries of hyperfine coupling tensors A* between an unpaired electron and carbon-13 atoms (blue and red arrows that are attached to these atoms) are shown. The arrow near the square acid unit indicates the bond about which the substituent exhibits a hindered rotation. The optimal geometry was found using the Turbomole computer program using the PBE0+D3(BJ) functional with the def2-QZVPP basis set. The hyperfine couplings were computed in the ReSpect computer program50,51 using the PBE0 functional with the decontracted Jensen’s pcH-2 basis set.52 Atoms are colored according to the Corey–Pauling–Koltun convention: hydrogen – white, carbon – black, oxygen – red, and nitrogen – blue.
The structures of two dominating conformers of 1,3-bisdiphenylene-2-phenylallyl radical functionalized with a square acid and alanine. The permanent electric dipole moments μe (green arrows) and the antisymmetries of hyperfine coupling tensors A* between an unpaired electron and carbon-13 atoms (blue and red arrows that are attached to these atoms) are shown. The arrow near the square acid unit indicates the bond about which the substituent exhibits a hindered rotation. The optimal geometry was found using the Turbomole computer program using the PBE0+D3(BJ) functional with the def2-QZVPP basis set. The hyperfine couplings were computed in the ReSpect computer program50,51 using the PBE0 functional with the decontracted Jensen’s pcH-2 basis set.52 Atoms are colored according to the Corey–Pauling–Koltun convention: hydrogen – white, carbon – black, oxygen – red, and nitrogen – blue.
The last example is trifluoropropan-2-ol shown in Figure 8.4. This molecule is not rigid since the hydroxyl group may rotate in the solution. Quantum chemistry computations indicate that two of the three fluorine nuclei in the CF3 group coupled with the CH̲(OH) proton have antisymmetries of ca. 2 Hz. Their amplitudes vary less than 10% when the group OH rotates, which is much less than the corresponding changes in the magnitude of the permanent electric dipole moment of the molecule. On average, one of the antisymmetries 3J*(19F, 1H) forms an angle of 69° with the moment μe while the other is at the angle of 133°. Hence, effectively a fraction of the maximum possible amplitude of the pseudoscalar Jc, corresponding to antisymmetry 3J*(19F, 1H) oriented parallel to the electric dipole moment μe, could be observed. This situation is also observed in the antisymmetric hyperfine coupling of the chiral derivative of BDPA, as well as in water 1J*(17O, 1H) antisymmetries.
The variation of the permanent electric dipole moment μe (green arrows) and antisymmetries of proton–fluorine indirect spin–spin coupling, 3J*(19F, 1H) (the CH̲(OH) proton; red and blue arrows), with rotation of the hydroxyl group (the black arrow) for the (R)-trifluoropropano-2-ol molecule. The third proton–fluorine coupling is an order of magnitude smaller than the two depicted ones and, consequently, is not shown in the figure. The same molecule is shown from two points of view for better visibility. On the far right: diagrams illustrating the relationship between local symmetry and directions of antisymmetries for 3J*(19F, 1H) of trifluoropropan-2-ol (the upper part) and 1J*(17O, 1H) of water (the lower part). Atoms are colored according to the Corey–Pauling–Koltun convention: hydrogen – white, carbon – black, oxygen – red, and fluorine – green.
The variation of the permanent electric dipole moment μe (green arrows) and antisymmetries of proton–fluorine indirect spin–spin coupling, 3J*(19F, 1H) (the CH̲(OH) proton; red and blue arrows), with rotation of the hydroxyl group (the black arrow) for the (R)-trifluoropropano-2-ol molecule. The third proton–fluorine coupling is an order of magnitude smaller than the two depicted ones and, consequently, is not shown in the figure. The same molecule is shown from two points of view for better visibility. On the far right: diagrams illustrating the relationship between local symmetry and directions of antisymmetries for 3J*(19F, 1H) of trifluoropropan-2-ol (the upper part) and 1J*(17O, 1H) of water (the lower part). Atoms are colored according to the Corey–Pauling–Koltun convention: hydrogen – white, carbon – black, oxygen – red, and fluorine – green.
In all three cases, the relative orientation of the vectors μe, J* (or A*) originates from the presence of a nucleus (or an unpaired electron) on the local two-fold axis/plane of symmetry and a nucleus located off the axis/plane. These symmetry features (C2 – the two-fold axis, a curved arrow, and σ-the mirror plane, a continuous line) are shown diagrammatically for the chiral BDPA derivative, trifluoroisopropan-2-ol, and water on the right sides of Figures 8.3 and 8.4. The influence of other types of local symmetry on the value of the pseudoscalar Jc can be analyzed in a similar way. It seems, however, that the occurrence of local symmetry close to the two-fold axis or plane is one of the most common occurrences in molecules and clearly illustrates the role of local symmetry in the case of the pseudovector J* and the global distribution of electric charge, manifested by the vector μe. See several further examples of achiral and chiral radicals in Figure 8.5.
Neutral radicals with their dipole moments (the green arrows), antisymmetries of some of their hyperfine coupling tensors. The red, blue, and yellow arrows represent the coupling of the unpaired electron with 15N, 13C, and 1H nuclei, respectively. Vectors and pseudovectors are not to scale. Nitroxyls: 1a – TEMPO, i.e., 2,2,6,6-tetramethyl-1-piperidinyloxy, (R)-1b – the TEMPO derivative of (R)-alanine, 2a – 1,3,5-trimethyl-6-oxoverdazyl, (R)-2b – the verdazyl derivative of (R)-phenylalanine; carbon-centered radicals: 3a – triphenylmethyl, 3b – bis(fluoren-9-yl)methyl, and (R)-3c – bisfluorene methyl derivative of (R)-alanine. The radicals 1b, 2b, and 3c are considered chiral from the point of view of EPR and NMR spectroscopy measurement time scales. Reproduced from ref. 17 with permission from the Royal Society of Chemistry.
Neutral radicals with their dipole moments (the green arrows), antisymmetries of some of their hyperfine coupling tensors. The red, blue, and yellow arrows represent the coupling of the unpaired electron with 15N, 13C, and 1H nuclei, respectively. Vectors and pseudovectors are not to scale. Nitroxyls: 1a – TEMPO, i.e., 2,2,6,6-tetramethyl-1-piperidinyloxy, (R)-1b – the TEMPO derivative of (R)-alanine, 2a – 1,3,5-trimethyl-6-oxoverdazyl, (R)-2b – the verdazyl derivative of (R)-phenylalanine; carbon-centered radicals: 3a – triphenylmethyl, 3b – bis(fluoren-9-yl)methyl, and (R)-3c – bisfluorene methyl derivative of (R)-alanine. The radicals 1b, 2b, and 3c are considered chiral from the point of view of EPR and NMR spectroscopy measurement time scales. Reproduced from ref. 17 with permission from the Royal Society of Chemistry.
8.7 Nuclear Magnetoelectric Effective Hamiltonian
8.7.1 Rotating Frame Transformation
8.7.2 Antisymmetric Nuclear Magnetic Shielding Effective Hamiltonian
For instance, for trifluoropropan-2-ol placed in a magnetic field B0 of strength 11.75 T and an electric field of amplitude 10 V mm−1, the nutation frequency for 19F nuclei slightly exceeds 1 mHz. The effective Hamiltonian in eqn (8.20) has the same form as that resulting from the cosinusoidal magnetic field oscillating with frequency ω1 along the x-axis that corresponds to a radiofrequency pulse rotating nuclear magnetization in a standard NMR experiment. However, the effect of the magnetic field on spin dynamics is noticeably stronger in comparison with the influence of the electric field because, in typical NMR experiments, the nutation frequency due to the magnetic field, , amounts to several kHz.
Let us make a brief comment regarding the relationship between the unwanted magnetic field B1 and the electric field E1 generating it, which is applied in experimental protocols aiming to observe the effect described in Section 7.9 of the previous chapter. Maxwell’s equations (Faraday’s law, in particular) dictate that the magnetic field B1 and the electric field E1 have to be perpendicular to each other and shifted in time by a quarter of a period. For instance, if the laboratory frame of reference is suitably chosen, then these fields may be written as B1 = B1 cos(ω1t)ex and . Consequently, the electric and magnetic fields tilt the magnetization into perpendicular directions: –y-axis and the x-axis, respectively.
Let us assume that the reference frame is such that the magnetic field is B1 = B1 cos(ω1t)ex and examine step-by-step how the cross product , the phase shift and spatial orthogonality between the B1 and E1 fields affect the effective rotation operator of the electric field derived from Maxwell’s equations, . This net effect follows from three independent factors among which two cancel out each other. First, if we ignore the phase shift and spatial orthogonality between B1 and E1, then the cross product in the Hamiltonian (compare with the scalar product in a purely magnetic spin Hamiltonian) results in the rotation operator , obtained from eqn (8.17) applied to the electric field E1 = E1 cos(ω1t)ex. Next, let us take into account the second factor, the phase shift between the electric field E1 and magnetic field B1, i.e., consider the electric field . Such a field corresponds to the rotation operator . Finally, by introducing the third factor, the orthogonality of the fields E1 and B1 in space, i.e., the electric field , one obtains the rotation operator ; see eqn (8.20) and (8.21).
8.7.3 Antisymmetric Spin–Spin Coupling Effective Hamiltonian
In the second case, i.e., when the primary goal is to observe the effects induced by the presence of the antisymmetric part of the indirect spin–spin coupling tensor J*, one can consider several factors that influence the design of the experiment. The most basic of these are (i) the choice of the spin system, i.e., homonuclear vs. heteronuclear two-spin system, (ii) the relative orientation of the electric E and magnetic field B0, i.e., one can orient the electric field either parallel or perpendicular to the static magnetic field B0, and (iii) the relative amplitude of the coupling between spins, quantified by the isotropic indirect spin–spin coupling J, to the amplitude of the interaction of the nuclear magnetic moment with the magnetic field, quantified by the spin precession frequency ω. Let us illustrate the various experimental possibilities offered by the Hamiltonian given in eqn (8.18) by three particular examples.
8.7.3.1 Case I
Regardless of whether the terms related to the interaction between the magnetic field and nuclear magnetic moments are present or not, the initial state (vide infra) is not affected by the n-fold application of a spin echo, i.e., the pulse sequence τ–180°–τ, where the delay duration is τ = 1/(4J). Let us assume that the electric field E is static and oriented along the magnetic field B0. Then, the initial state transforms into a singlet–triplet population difference, ∣S0〉〈S0∣ − ∣T0〉〈T0∣, with a period of (2JcE)−1, where and . The advantage of using the state ∣S0〉〈S0∣ − ∣T0〉〈T0∣ is that it is resistant to dipolar relaxation, enabling a longer period of application of the exciting electric field E in comparison with other states of the system, e.g., . Therefore, by using a number of echoes equal to the natural number nearest to J(4JcE)−1, and assuming a sufficiently long relaxation time, one can potentially transform the state completely into the state ∣S0〉〈S0∣ − ∣T0〉〈T0∣. The stronger the system of coupled spins, the lesser the influence of the term related to the interaction of the spin with the magnetic field on the dynamics of the spin system. Hence, the most strongly coupled systems are beneficial for obtaining maximal electric field-induced transfer from the state to the ∣S0〉〈S0∣ − ∣T0〉〈T0∣ state.
8.7.3.2 Case II
8.7.3.3 Case III
8.8 A General Description of the Chirality-sensitive Electric Polarization
In the next sections, a few effects denoted by , , , and that could facilitate chiral discrimination by NMR, are described and briefly summarized in Table 8.2. Each of them is provided with a brief description on how it could be generated, its experimental and hardware requirements, why the effect is important compared to other effects discussed here, and the feasibility of its experimental observation (Table 8.2).
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 . | |||||
2 | J* | 0° | ωH − ωF | 1,1,1-Trifluoro-propan-2-ol | 10−3 | |
2 | A* | 0° | ωe − ωC | 13C-BDPA-alanine | 10−4 | |
2 | J* | N/A | ωH − ωF | 1,1,1-trifluoro-propan-2-ol | 10−5 | |
3 | D | 0° | ωH − ωC | 13C-alanine | 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 . | |||||
2 | J* | 0° | ωH − ωF | 1,1,1-Trifluoro-propan-2-ol | 10−3 | |
2 | A* | 0° | ωe − ωC | 13C-BDPA-alanine | 10−4 | |
2 | J* | N/A | ωH − ωF | 1,1,1-trifluoro-propan-2-ol | 10−5 | |
3 | D | 0° | ωH − ωC | 13C-alanine | 10−6 |
The magnitude with respect to that obtained after the excitation of the sample with a 90° pulse; E = 1 kV mm−1, B0 = 10 T.
8.9 Chirality-sensitive Coherences Induced by the Antisymmetric Part of the J-coupling Tensor
8.9.1 Description of the Effect
The spin dynamics on which the effect is based arises from the quantum-mechanical treatment described in Section 8.7.3 for case II. In particular, the application of a radiofrequency electric field on the two-spin system placed in a chiral molecule induces zero-quantum coherences, whose phase differs for the two enantiomers of that molecule.
The effect relies on the partial alignment of the polar chiral molecules. Therefore, for its occurrence, it is necessary for the molecules to be capable of reorienting under the influence of the E field on a time scale comparable to the time variation of the electric field. Such conditions, for low enough frequency, are met in chiral substances dissolved in the liquid phase and in some molecular solids allowing reorientation in a lattice site, e.g., the chiral derivatives of fullerenes. Typically, the higher the field frequency, the more hindered the reorientation of the sample becomes due to internal friction in a liquid, which manifests itself as the dissipation of energy (sample heating). The frequency dependence strongly varies with the solvent – numerical values of the imaginary part of the dielectric permittivity, on which the losses mainly depend, are listed for several common organic solvents in ref. 55. Moreover, excitation at the differential frequency, as compared to the spin precession frequency, allows for longer molecular reorientation times; the high value of these times is a limiting factor in the case of medium-viscosity solvents.
Even more importantly, dielectric losses increase with the square of the electric field amplitude, which may significantly limit experiments, as shown by the mathematical models in Figure 8.6.
The simulated dielectric heating effect due to the application of a radiofrequency electric field on an NMER probe assembly. The probe head generating the electric field through two parallel plates of the capacitor (marked by the orange color), together with the sample placed inside the coil surrounded by the capacitor (on the left side), the temperature distribution in a vertical cross-section of the assembly (in the middle), the temperature profile along the symmetry axis of the sample depending on the flow and temperature of the cooling gas flowing from the source located at the bottom of the assembly (on the right side).
The simulated dielectric heating effect due to the application of a radiofrequency electric field on an NMER probe assembly. The probe head generating the electric field through two parallel plates of the capacitor (marked by the orange color), together with the sample placed inside the coil surrounded by the capacitor (on the left side), the temperature distribution in a vertical cross-section of the assembly (in the middle), the temperature profile along the symmetry axis of the sample depending on the flow and temperature of the cooling gas flowing from the source located at the bottom of the assembly (on the right side).
8.9.2 Experimental Protocol for the Effect
The experimental protocol for effect is shown in Figure 8.7. The chiral sample is polar and has a heteronuclear pair of spins (I1 and I2). First, the electric field oscillating at the difference frequency, E1 = E1 cos((ω1 − ω2)t)ez, excites the sample for a duration of τ = T2, where T2 is the spin–spin relaxation time. Then, the generated chirality-dependent coherences are converted into an observable magnetization of the spin I2 (Figure 8.7A). In the next step, the chirality-sensitive magnetization induced by the E field of the capacitor is detected by the coil (not shown), generating a signal with opposite phases for the enantiomers (Figure 8.7B).
Chirality-sensitive coherences induced by the antisymmetric part of the J-coupling tensor (the effect): (A) the pulse sequence with the states of the spin system marked after (i) application of the electric field oscillating at the difference frequency (the gray rectangle) and (ii) radiofrequency pulses at spin-precession frequencies ω1 (spin I1) and ω2 (spin I2) flanked by pulses of magnetic-field gradient along the z-axis (the white rectangles); (B) the arrangement of the main magnetic field B0 and electric field E delivered by a capacitor surrounding the sample from above and below and the expected difference in the signals of the chiral molecule enantiomers. In the pulse sequence, the delay τ′ equals 1/(4J). The positive phase of the red signal is attributed to the (R)-enantiomer, and the negative one of the blue signal is attributed to the (S)-enantiomer only for illustrative purposes.
Chirality-sensitive coherences induced by the antisymmetric part of the J-coupling tensor (the effect): (A) the pulse sequence with the states of the spin system marked after (i) application of the electric field oscillating at the difference frequency (the gray rectangle) and (ii) radiofrequency pulses at spin-precession frequencies ω1 (spin I1) and ω2 (spin I2) flanked by pulses of magnetic-field gradient along the z-axis (the white rectangles); (B) the arrangement of the main magnetic field B0 and electric field E delivered by a capacitor surrounding the sample from above and below and the expected difference in the signals of the chiral molecule enantiomers. In the pulse sequence, the delay τ′ equals 1/(4J). The positive phase of the red signal is attributed to the (R)-enantiomer, and the negative one of the blue signal is attributed to the (S)-enantiomer only for illustrative purposes.
8.9.3 Significance and Estimation of the Magnitude of the Effect
The promising samples for experimental observation of the effect are light fluoroalcohols considering, e.g., the two-bond fluorine-proton coupling of (1,1,1)-trifluoropropan-2-ol and other compounds that contain fluorine, e.g., the three-bond fluorine-fluorine coupling of 1,1,1,2-tetrafluoro-2-chloroethane. In these molecules, the pseudoscalar for the proton–fluorine spin system is on the order of 1 nHz m V−1. Let us assume that the spin system is the pair of the proton and fluorine nuclei, the permanent electric dipole moment of the molecule is μe = 1 D, and the sample is placed in the static magnetic field B0 = 10 T, the exciting electric field has a strength of E = 100 V mm−1, and the longitudinal relaxation time T2 is 1 s for both nuclei. Then, the estimated magnitude of the chirality-sensitive nuclear magnetization is 10−4 of the nuclear magnetization signal obtained after applying a 90° pulse to the sample at thermodynamic equilibrium.
Applying the oscillating electric field at the difference frequency minimizes the perturbation of the system by the time-dependent magnetic field generated by the E field. Considering that the magnitude of the effect is proportional to the strength of the electric field E, and the dielectric heating is proportional to the square of the strength of the electric field, one finds that a relatively low-frequency excitation is optimal for the experiment (e.g., for the 1H–19F pair the difference frequency is approx. 30 MHz in a magnetic field of 10 T).
8.10 Electric Polarization Induced by the Antisymmetric Part of Indirect Spin–Spin Coupling
8.10.1 Description of the Effect
8.10.2 Experimental Protocol for the Effect
The experimental protocol for the effect is shown in Figure 8.8. In the first step, the desired quantum state of the two-spin system is obtained using the modified insensitive nuclei enhanced by the polarization transfer (INEPT) pulse sequence (Figure 8.8A). Then, the obtained state, which is described by the density matrix , generates an oscillating electric polarization P at the sum frequency ω1 + ω2. Depending on the enantiomers, the induced electric polarization P has a parallel or an antiparallel direction with respect to the field B0 (Figure 8.8B).
Electric polarization induced by the antisymmetric part of indirect spin–spin coupling (the effect). On the left side, the pulse sequence is shown. The narrow rectangles are 90° pulses, whereas the wide ones are 180° pulses. The red arrow indicates the final state of the spin system generating chirality-sensitive electric polarization of the sample. The radiofrequency pulses are applied on-resonance for both spins I1 and I2. In the pulse sequence, the delay τ′ equals 1/(4J). On the right side, one of the spin state coherences after application of the pulse sequence is depicted. This coherence is the same for both enantiomers. The difference in the directions of electric polarisation P (upwards vs. downwards) arises from the different transformation properties of the vector μe and the pseudovector J* under mirror reflection.
Electric polarization induced by the antisymmetric part of indirect spin–spin coupling (the effect). On the left side, the pulse sequence is shown. The narrow rectangles are 90° pulses, whereas the wide ones are 180° pulses. The red arrow indicates the final state of the spin system generating chirality-sensitive electric polarization of the sample. The radiofrequency pulses are applied on-resonance for both spins I1 and I2. In the pulse sequence, the delay τ′ equals 1/(4J). On the right side, one of the spin state coherences after application of the pulse sequence is depicted. This coherence is the same for both enantiomers. The difference in the directions of electric polarisation P (upwards vs. downwards) arises from the different transformation properties of the vector μe and the pseudovector J* under mirror reflection.
8.10.3 Significance and Estimation of the Magnitude of the Effect
The favourable samples for the observation of the effect are the same as those for the effect described in the previous section. The pseudoscalar for the proton–fluorine spin system is on the order of 1 nHz m V−1, resulting in a chirality-sensitive electric polarization P ≈ 0.1 aC m−2 in a magnetic field of strength B0 = 10 T. In this case, the voltage generated by this electric polarization P in a resonance circuit containing the capacitor is approximately five orders of magnitude smaller than that of the nuclear magnetization corresponding to the equilibrium state of the sample.
The effect can be observed without the need to apply an electric field. Consequently, dielectric heating, which may become an obstacle in observing the effects induced by the electric field E, does not affect the sample. In contrast to the effect, the induced electric polarization P oscillates at the sum frequency and, therefore, does not overlap with the precession frequencies of the spins I1 and I2, simplifying the experimental setup. The detector is a parallel-plate capacitor in the radiofrequency resonant circuit.
8.11 Chirality-sensitive Effects Induced by Antisymmetry of the A-coupling Tensor
The antisymmetric interactions between spins are not limited to the interactions between nuclear spins. In fact, for each of the proposed effects that may occur for nuclear spins, such as those described in Sections 8.9 (the effect) and 8.10 (the effect), an analogous effect can be proposed by replacing one or both nuclear spins with, for instance, electron spins. In this section, we will focus on such interactions that are in the domain of EPR spectroscopy or at the borderline between NMR and EPR. It is tempting to presume that such a generalization is evident due to the similarity of the mathematical framework used to describe the antisymmetries J* and A*; however, the very different technical challenges posed by EPR compared to NMR, lead to the opposite conclusion. Moreover, there is a noticeable gap between the energy scales of the spin–spin interactions between, on the one hand, nuclei and electrons and, on the other hand, among nuclei only. Typically, the sensitivity of the measurements of the electron spin transitions surpasses that of nuclear transitions by several orders of magnitude. Thus, the effects caused by A* place us in a favorable position to achieve direct sensitivity to chirality by EPR.
8.11.1 Description of the Effect
In general terms, the mathematical description of the effect mediated by A* follows those described in Sections 8.9 and 8.10 (vide supra). Similar to the two effects described in those sections, the antisymmetry of the hyperfine coupling yields a chirality-sensitive electric polarization if the initial quantum state of the system is suitably chosen and the electric field, oriented along the main magnetic field, oscillates at the difference frequency and induces zero-quantum coherences; the phase of these coherences depends on the handedness of the studied molecule.
8.11.2 Experimental Protocol for the Effect
The experimental protocols for the effects due to the antisymmetry A* are shown in Figure 8.9.
Two chirality-sensitive effects induced by antisymmetric hyperfine coupling, which lifts the intrinsic insensitivity of electron paramagnetic resonance spectroscopy to detection of molecular chirality (called collectively in the text as the effect). On the top: (R)-BDPA-Ala (in the red color) and (S)-BDPA-Ala (in the blue color) radicals and the chirality-sensitive EPR electric polarization signals that are at a frequency equal to the difference between the resonance frequencies of the 13C nucleus and the electron. An absorption spectrum is shown, in which the frequency is given relative to the electronic transition, and the isotropic hyperfine electron-13C coupling is . On the bottom: the EPR/NMR pulse sequence generating the spin state ( is the electron, while, the 13C nuclear spin operator), whose phase indicates which enantiomer is present in the sample. The spectra on the far right side illustrate the expected signal lineshape in the dispersion mode, i.e., according to the convention used in EPR spectroscopy.
Two chirality-sensitive effects induced by antisymmetric hyperfine coupling, which lifts the intrinsic insensitivity of electron paramagnetic resonance spectroscopy to detection of molecular chirality (called collectively in the text as the effect). On the top: (R)-BDPA-Ala (in the red color) and (S)-BDPA-Ala (in the blue color) radicals and the chirality-sensitive EPR electric polarization signals that are at a frequency equal to the difference between the resonance frequencies of the 13C nucleus and the electron. An absorption spectrum is shown, in which the frequency is given relative to the electronic transition, and the isotropic hyperfine electron-13C coupling is . On the bottom: the EPR/NMR pulse sequence generating the spin state ( is the electron, while, the 13C nuclear spin operator), whose phase indicates which enantiomer is present in the sample. The spectra on the far right side illustrate the expected signal lineshape in the dispersion mode, i.e., according to the convention used in EPR spectroscopy.
The spin precession frequencies of any nucleus and an electron differ a lot, so the use of initial states 〈I1z〉eqI1z ± 〈I2z〉eqI2z becomes irrelevant. The difference between the angular frequency of nuclear spin precession (exemplified here by carbon-13), ωC, and that of the electron, ωe, is comparable to the isotropic hyperfine coupling, . Therefore, there is a risk of a poor spectral separation of the difference frequency from the spin transition of the system. Although the separation is in the range of MHz at fields corresponding to the X-band (e.g., B0 = 0.33 T) or the W-band (ca. 3 T), the duration of the microwave pulses results in a relatively broad excitation. A sufficiently selective excitation provides a major advantage in the absence of a sample response, if there are no chirality-sensitive effects. This happens either because of a suitably chosen initial state resulting in electric polarization, or excitation by the external electric field. Therefore, there is no need to eliminate the interfering achiral background signal.
8.11.3 Significance and Estimation of the Magnitude of the Effect
The success of an experiment heavily depends on the degree of fulfillment of required conditions imposed by the experiment protocol, in its hardware implementation, which may be technically challenging. From the point of view of the optimal sample, the lengthening of the electron relaxation time seems to be crucial. Comparing the nuclear relaxation times (that are on the order of seconds) with those typical in EPR spectroscopy (a few to dozens of ns), we can see that creating a short radio/microwave pulse is a harder task for EPR than for NMR. Therefore, relatively slow relaxation of the chiral radical is very convenient. For, e.g., in BDPA derivatives, one cannot expect significantly longer electron relaxation time than a few microseconds. The second issue is isotope labeling. In several cases, nuclei related by an approximate C2 symmetry axis would give canceling signals, so non-selective labeling does not offer much advantage over samples that are not enriched at all. Taking into account such considerations, one can find that, if the relaxation time of the chiral radical is ca. 10 µs, the ratio between the electric and magnetic fields at the detector is , and when the radical is selectively 13C-enriched, the amplitude of the expected chirality-sensitive electric polarization is on the order of 10−4 of the standard EPR signal.
8.12 Interference Between Dipolar Relaxation Mechanisms in an Electric Field
In contrast to previously postulated chirality-sensitive NMR effects, i.e., the magnetization induced by the electric field (the , and effects) and the electric polarization induced by precessing nuclear magnetic moments (the , , and effects), the effect described in this section does not require computation of the components of the nuclear magnetic shielding tensor and the indirect spin–spin coupling tensor for determination of the sense of molecular chirality. A sufficient condition for determining the absolute configuration of a molecule by observing the dipolar coupling effect is having knowledge about the relative orientation of the observed three nuclei with respect to the permanent electric dipole moment of the molecule.
8.12.1 Description of the Effect
The following discussion will be limited to the interference involving the electric field-induced terms of the direct coupling for only one pair of spins, I1 and I2. Without loss of generality, the other contributions are neglected since their influence on the spin dynamics may be suppressed by suitably chosen experimental conditions.
In order to achieve a chirality-sensitive effect, it is necessary to ensure that the interaction is linear in the electric field, which may be achieved for interference between an E-field induced relaxation mechanism and one that is not perturbed by the electric field. For isotropic tumbling of the molecule, only relaxation mechanisms of the same tensorial rank contribute to cross-correlation effects; thus, only the second-rank part of the contribution , couples with dipolar relaxation, parametrised by a second-rank tensor.
Interference between dipolar relaxation mechanisms in an electric field (the effect). The tensors U12 (in the blue color) and D13 (in the red colour) represented as ellipsoids oriented along their eigenvectors and of semiaxes equal to the absolute values of their eigenvalues, i.e., the ratios of the semiaxes are 0 : 1 : 1 and 1 : 1 : 2, respectively. The orientation of the tensors, which is shown in the figure corresponds to the maximum value of the amplitude of the interference. The permanent electric dipole moment μe is shown as a green arrow.
Interference between dipolar relaxation mechanisms in an electric field (the effect). The tensors U12 (in the blue color) and D13 (in the red colour) represented as ellipsoids oriented along their eigenvectors and of semiaxes equal to the absolute values of their eigenvalues, i.e., the ratios of the semiaxes are 0 : 1 : 1 and 1 : 1 : 2, respectively. The orientation of the tensors, which is shown in the figure corresponds to the maximum value of the amplitude of the interference. The permanent electric dipole moment μe is shown as a green arrow.
8.12.2 Experimental Protocol for the Effect
In order to achieve a chirality-sensitive response from the sample, and one that has sufficient amplitude for experimental observation, it is beneficial to use an oscillating electric field. If the molecule tumbles isotropically and much faster than the speed at which the nuclear spins precess in the magnetic field B0, i.e., at the extreme narrowing limit, the relaxation matrix becomes block diagonal. The excitation of the double and triple coherences requires the field to oscillate at a higher frequency than the spin-precession frequency. At such high frequencies, dielectric heating may severely limit the available amplitude of the electric field. The unwanted magnetic field generated by the electric field oscillating at the spin-precession frequency excites single quantum coherences that are much larger than the postulated effect. Therefore, the most convenient choice is electric field oscillating at the frequency equal to the difference between the precession frequencies of the two spins, which perturb the blocks of the zero quantum coherences.
In order to avoid unnecessary complexity of the solution, let us limit to the case where only the direct dipole–dipole relaxation mechanism contributes to the relaxation matrix. This mechanism is the most pronounced one under typical experimental conditions that prevail in liquid-state physicochemical NMR studies. Moreover, we assume that the dipole–dipole relaxation time for each pair of nuclei is T1. In this case, one finds that the maximal amplitude of the chirality-sensitive signal is reached by the application of the electric field E on the spin state , over the period τ = T1 and then by observation of the phase of the chirality-dependent magnetization of the spin I3. The amplitude is numerically roughly two orders smaller than .
8.12.3 Significance and Estimation of the Magnitude of the Effect
Suppose the electric field E is oscillating at a frequency equal to the difference between the spin precession frequencies of two of the three spins, and that the field E is aligned along the magnetic field B0. In this case, the magnitude of the predicted effect for B0 = 10 T and E = 5 kV mm−1 is 1% of the signal generated by the interference between dipole–dipole relaxation pairs measured without using the electric field E. Although the effect is small in amplitude, its significance lies in the fact that it enables the determination of the absolute configuration of a molecule provided that the relative orientation of these three nuclei with respect to the permanent electric dipole moment of the molecule μe, is known.
8.13 Sources of Radiofrequency Electric Field for J-NMER Experiments
Basically, the magnetic field needed to observe the effects should be parallel to the static magnetic field of the spectrometer, so it requires a different experimental arrangement than that described in Section 7.14 for the effects. Note that, unlike much more complicated systems that suppress the undesirable transverse component of the magnetic field generated by the electric field, in the design of the capacitor relevant to J-NMER experiments, it is sufficient to consider the highest possible homogeneity and amplitude of the generated electric field as primary design objectives. One of the radiofrequency electric field sources is a capacitor with parallel plates, between which the sample is placed, aligning its axis of symmetry along the axis of the capacitor. The advantageous aspect of such an electric field source is not only that the sample can be placed inside the capacitor, but also the surrounding transceiver coil. This type of arrangement is shown in Figure 8.11.
The circular parallel-plate capacitor (in the color of metallic copper) together with the transmitting and receiving saddle coil located inside. The sample (not visible) is placed along the vertical axis of symmetry of the assembly, inside the coil. The remaining elements serve as a frame positioning the individual elements of the system.
The circular parallel-plate capacitor (in the color of metallic copper) together with the transmitting and receiving saddle coil located inside. The sample (not visible) is placed along the vertical axis of symmetry of the assembly, inside the coil. The remaining elements serve as a frame positioning the individual elements of the system.
One can optimize the dimensions of the capacitor using the electric field distribution found by integrating the charge density on the capacitor’s plates; see the corresponding classical electrostatics formula in ref. 56. The capacitor shown in Figure 8.11 has holes that make such a configuration similar to a Helmholtz coil in terms of optimizing its shape. As in the case of a Helmholtz coil, where the highest homogeneity of the magnetic field is achieved when the distance between the coils equals their radii, one can find optimal values of the distance between the capacitor plates L, the radius the plates R and the radius of the holes r in them.
The optimum relationships between the height L and the radius r of the holes both normalized to the radius R of the plates of the capacitor in the capacitor-coil assembly of Figure 8.11. The optimal dimensions in the direction along and perpendicular to the symmetry axis of the capacitor are shown as continuous and dashed lines, respectively. Gray lines mark the asymptotic cases reached for a ring made of a charged wire, which are for the direction parallel to the symmetric axis of the capacitor and for the direction perpendicular to the symmetric axis of the capacitor.
The optimum relationships between the height L and the radius r of the holes both normalized to the radius R of the plates of the capacitor in the capacitor-coil assembly of Figure 8.11. The optimal dimensions in the direction along and perpendicular to the symmetry axis of the capacitor are shown as continuous and dashed lines, respectively. Gray lines mark the asymptotic cases reached for a ring made of a charged wire, which are for the direction parallel to the symmetric axis of the capacitor and for the direction perpendicular to the symmetric axis of the capacitor.
For a capacitor to generate an electric field that oscillates at radio frequencies, its plates must be connected to an appropriate excitation unit. For the capacitor shown in Figures 8.11 and 8.12, this could be a transmission line with a standing wave whose anti-node, i.e., a point where the amplitude of the wave reaches maximum, is at the capacitor plate. Another solution, geometrically compatible with direct connection to the coaxial cable, is shown in Figure 8.13. There is a capacitor with circular and parallel plates, the lower plate of which is connected to the core and the upper plate is connected to the shield of the coaxial cable. Such a solution could be used together with a standard NMR probe. In this case, the capacitor is immersed in a chiral liquid that is placed in a standard NMR tube. The disadvantage incurred in this case is the reduced homogeneity of the electric field and, above all, the static magnetic field due to the mismatch between the magnetic susceptibility of a chiral liquid and the metallic elements constituting the plates, the core, and the shield of the transmission line.
A capacitor directly connected to the coaxial cable (on the left side) together with the cross-section of electric field distribution inside it, taken along the plane containing the symmetry axis of the system (on the right side). The red color marks a high electric-field amplitude, while the field of a low amplitude is shown in the blue color. Arrows mark the direction of the electric field.
A capacitor directly connected to the coaxial cable (on the left side) together with the cross-section of electric field distribution inside it, taken along the plane containing the symmetry axis of the system (on the right side). The red color marks a high electric-field amplitude, while the field of a low amplitude is shown in the blue color. Arrows mark the direction of the electric field.
8.14 Summary and Conclusions
Although from an experimental point of view it would be difficult to consider each of the NMER effects described in this and previous chapter individually, there is an inevitable difference in their amplitudes. If we limit ourselves to the chirality-sensitive effects that are induced by the changes in the spin state over time, the NMER spectrum is a superposition of the effect due to σ* and resulting from J*. The dominant contribution stems from the effect of antisymmetric magnetic nuclear shielding, rather than from antisymmetry of the indirect spin–spin coupling. Therefore, in the domain of experiments aiming for the detection of chirality-sensitive polarization that is small compared to the achiral nuclear magnetization, antisymmetric spin–spin coupling does not offer any features that make the experiment more feasible (Figure 8.14, the upper trace). However, the situation changes if we consider the frequencies that lie far beyond the tiny fragment of the NMR spectrum that is excited on-resonance. Especially the region near the difference frequency seems to be well-suited for achieving the experimental aim of observing chirality-sensitive NMER effects. In this case, the most difficult requirement to fulfill, i.e., suppression of the unwanted magnetic field, is lifted, which permits us to drastically simplify the construction of the device generating the radiofrequency electric field (Figure 8.14, the lower trace). This applies to the effects , , and . Among them, the coherences induced by the antisymmetric J* seem to be the most promising because of their relatively high amplitude.
The various contributions to the chirality-sensitive polarization in a system consisting of a nucleus and an electron interacting through hyperfine coupling. The antisymmetries due to the interaction of the magnetic field with nuclear magnetic moment (the blue pseudovector) and the electron magnetic moment (the red pseudovector), hyperfine coupling between them (the black A* pseudovector) are shown for a chiral molecule and its mirror image (A). The green arrow shows the permanent electric dipole moment μe of the molecule. Expected signals of the enantiomers due to the precession of the magnetization M (the red color) and the oscillation of the chirality-sensitive electric polarizations: (a sum of effects induced by and ) and originating from A*. The spin states are: and . The total signal is shown in black. Reproduced from ref. 17 with permission from the Royal Society of Chemistry.
The various contributions to the chirality-sensitive polarization in a system consisting of a nucleus and an electron interacting through hyperfine coupling. The antisymmetries due to the interaction of the magnetic field with nuclear magnetic moment (the blue pseudovector) and the electron magnetic moment (the red pseudovector), hyperfine coupling between them (the black A* pseudovector) are shown for a chiral molecule and its mirror image (A). The green arrow shows the permanent electric dipole moment μe of the molecule. Expected signals of the enantiomers due to the precession of the magnetization M (the red color) and the oscillation of the chirality-sensitive electric polarizations: (a sum of effects induced by and ) and originating from A*. The spin states are: and . The total signal is shown in black. Reproduced from ref. 17 with permission from the Royal Society of Chemistry.
Moreover, in some special cases, such as in a very strongly coupled spin system, a static electric field could be applied that can be higher in amplitude than the radiofrequency one. A certain obstacle in this case is the difficulty posed by the molecular structure because, in addition to the condition that the spins have to be nearly equivalent, the condition that the permanent electric dipole vector μe and the pseudovector J* cannot be perpendicular to each other, must also be met.
Looking cross-sectionally at the effects described in this section, i.e., , one can notice that certain quantum spin states appear repeatedly regardless of the specific experimental protocol. An example of such a state is , which is antisymmetric when the spins are interchanged. The frequent occurrence of such a quantum state of the system results from its connection with the antisymmetric tensorial properties of the molecule. This leads to the conclusion that the analysis of potential effects depending on the molecular chirality could also be carried out in the order opposite to that presented in this book, i.e., by searching for a quantum state in which chirality can manifest itself directly and then finding its connection with molecular properties.
It could also be observed that one of the effects induced by the spin–spin coupling, namely, the perturbation of dipolar relaxation in a three-spin system by the external electric field (effect ), offers a unique feature: the possibility of determination of the absolute configuration of the molecule, without the need for any quantum-mechanical computation of pseudoscalars. This predicted effect would offer a very rare opportunity for a direct insight into the molecular structure at a comparable level of detail, as mentioned in Chapter 1, such as Coulomb explosion imaging and atomic force microscopy of chiral molecules.
Besides the reasons mentioned above, which motivate experiments aiming for the observation of chirality-sensitive effects, one can perceive studies of such effects not only as a possible practical tool for studying molecular chirality, but also as a means of determining their structure beyond the fact that chiral molecules are distinguishable from their mirror images. Each chirality-sensitive NMER effect has a different dependence on the molecular structure, e.g., vs. or vs. . This allows us to see the molecular structure and dynamics from different perspectives, possibly revealing its different physicochemical properties.
Acknowledgements
PG acknowledges the National Science Centre, Poland, for the financial support through OPUS 16 grant No. 2018/31/B/ST4/02570, the Max Planck Society for providing access to computational resources (in particular, the COMSOL program), and Dr Grzegorz Łach (Faculty of Physics, University of Warsaw) for helpful discussions about the permanent electric dipole moment of a chiral molecule. JV thanks the Academy of Finland (project 331008) and University of Oulu (Kvantum Institute) for the financial support, as well as CSC-the Finnish IT Centre of Science and the Finnish Grid and Cloud Infrastructure project (persistent identifier urn:nbn:fi:researchinfras-2016072533) for computational resources.
This chapter is subject to a Creative Commons CC-BY-NC-ND 4.0 International license. Financial support from the National Science Centre, Poland (through OPUS 16 grant No. 2018/31/B/ST4/02570), Academy of Finland (project 331008) and University of Oulu is acknowledged.