Next Article in Journal
Connecting GSK-3β Inhibitory Activity with IKK-β or ROCK-1 Inhibition to Target Tau Aggregation and Neuroinflammation in Alzheimer’s Disease—Discovery, In Vitro and In Cellulo Activity of Thiazole-Based Inhibitors
Previous Article in Journal
A Comprehensive Analysis of Diversity, Structure, Biosynthesis and Extraction of Biologically Active Tannins from Various Plant-Based Materials Using Deep Eutectic Solvents
Previous Article in Special Issue
Reconversion of Parahydrogen Gas in Surfactant-Coated Glass NMR Tubes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Measurements of Nuclear Magnetic Shielding in Molecules

Laboratory of NMR Spectroscopy, Faculty of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(11), 2617; https://doi.org/10.3390/molecules29112617
Submission received: 29 March 2024 / Revised: 24 May 2024 / Accepted: 30 May 2024 / Published: 2 June 2024

Abstract

:
The origin of nuclear magnetic shielding in diamagnetic molecules is discussed, pointing out various contributions to the shielding from electrons and the effects of intra- and intermolecular interactions. In NMR practice, chemical shifts are determined first as the measure of shielding in observed samples. The descriptions of shielding and chemical shifts are not fully consistent. Gas phase studies permit the withdrawal of intermolecular contributions from shielding and obtaining the magnetic shielding data in isolated molecules. The shielding determination in molecules is possible using at least three methods delivering the reference shielding standards for selected nuclei. The known shielding of one magnetic nucleus can be transferred to other nuclei if the appropriate nuclear magnetic moments are available with satisfactory accuracy. It is possible to determine the nuclear magnetic dipole moments using the most advanced ab initio shielding calculations jointly with the NMR frequencies measurements for small-sized isolated molecules. Helium-3 gas is postulated as all the molecules’ primary and universal reference standard of shielding. It can be easily applied using common deuterium lock solvents as the secondary reference standards. The measurements of absolute shielding are available for everyone with the use of standard NMR spectrometers.

1. Introduction

Electrons always shield atomic nuclei in molecules from the influence of an external magnetic field. This physical phenomenon is described by the difference between the induction of applied field B0 and its value Beff experienced by the nucleus and for isotropic species is expressed as follows:
B e f f = 1 σ B 0
where σ is the shielding parameter dependent on the total electronic structure of the observed molecular system, the above form of Equation (1) means σ = 1/3 xx + σyy + σzz) because in a general case, shielding is a second-rank tensor and the induction of magnetic field is represented by its vectors. Equation (1) is sufficient for spherical systems like atoms or molecules in the gaseous and liquid states where molecular reorientation is not hindered. For atoms, the shielding is just described by the Lamb [1] equation [2]. The shielding theory for molecules is a bit more complex and was first formulated by Ramsey for diatomic molecules, especially for a hydrogen molecule [3,4]. In 1954, Saika and Slichter [5] noted that the magnetic shielding of a nucleus σi can be presented as the sum of three different components
σ i = σ i d + σ i p + i j σ j
In this equation, σid and σip are the local diamagnetic and paramagnetic parts of shielding when the last term is responsible for all the modifications of σi arising from intra- and intermolecular effects. A more detailed description of the shielding in diamagnetic molecules was provided by Pople [6] and other researchers [7,8]. They explained the first two partial terms in Equation (2) as follows [9]:
σ i d = μ 0 4 π e 2 3 m e 0 | n 1 r n | 0
σ i p = μ 0 4 π e 2 3 m e k 0 1 E k E 0 0 | n L n | k k | n L n r n 3 | 0
Equations (3) and (4) contain constants: e—electron charge, me—electron mass, and µ0free space permeability. Then, in the bra-ket notation, there are the described wave functions of the ground state of electrons (0) and all the excited states (k) with their appropriate energies E0 and Ek. L and r are the vectors that represent the orbital angular momentum and the distance from an arbitrary origin for the nth electron, respectively. The important features of Equations (3) and (4) arise from the different signs of these two components to the total magnetic shielding: σid is positive (shielding effect) while σip is always negative (deshielding effect). It means that the total shielding effects for diamagnetic molecules can be positive (σi > 0) or negative (σi < 0), assuming i j σ j = 0 in Equation (2). The last term of Equation (2) is responsible for intra- and intermolecular effects that can be separately measured in the gas phase as described by Jameson [10,11]. All the possible components of Equation (2) are presented in Figure 1 using selected examples from multinuclear NMR spectra.
As seen in Figure 1, the magnetic properties of atoms and molecules are the function of their electronic structures. Atoms are usually more shielded, having more electrons, but atomic electronegativity is also an important feature. Molecules have a more complex distribution of electrons around atomic nuclei, and all the terms of Equation (2) must be remembered. The total shielding can be negative for some atomic nuclei; the F2 molecule is a spectacular example of such a case but is not unique. There are many other examples of the total deshielding effects for diamagnetic molecules, especially observed in 15N, 17O, and 19F NMR spectra. However, more electrons in the neighborhood usually lead to an increase in nuclear magnetic shielding in diamagnetic molecules. Molecular vibrations and rotation cause further changes in shielding because the interatomic distance is enlarged. It results in deshielding effects in the majority of molecules, but some exceptions from the above rule are also known, cf. PH3 molecule shown in Figure 1. All intermolecular interactions also modify nuclear shielding in diamagnetic molecules; usually, the decrease of shielding is observed, but the reverse effects occur for selected atoms containing lone pairs of electrons. Figure 1 illustrates such a case for a 15N nucleus in acetonitrile.
Gas phase studies are crucial for the separation of molecular shielding parameters from all the large intermolecular contributions present in macroscopic samples. The same effects are still present in the gaseous samples, but they are much smaller and can be eliminated by the extrapolation of shielding measurements to the zero-density limit [10,11]. It gives the measurement of shielding equivalent to isolated molecules denoted as σ0. Our review is mostly focused on the shielding in isolated molecules because only such experimental results can be used to connect with the most advanced calculations of the same parameters. Let us add that the description of nuclear magnetic shielding given by Equations (2)–(4) is very useful in the qualitative understanding of shielding in diamagnetic molecules, but such an approximation is old-fashioned and is not applicable to modern quantum chemical calculations. The present state-of-the-art calculations of shielding [12,13,14,15,16,17,18] deliver such excellent results that they can often be treated as the source of the best data even in experimental NMR work, as shown in the recent comparison of experimental and calculated NMR parameters in CH4-nFn molecules [19]. It is nothing unusual, but verifying all available calculated results by experiment is also welcomed because many different approximate quantum chemistry methods are widely applied to shielding calculations.
In molecules, atomic nuclei are always surrounded by electrons, and so far, there is no possibility for the straightforward measurement of the molecular shielding, σi. NMR spectroscopy offers only a reading of the shielding difference between two macroscopic shielded objects, known as a chemical shift. This parameter is extremely helpful for the qualitative analysis of chemical compounds but contains rather limited information on shielding. Nevertheless, chemical shifts are also exploited to determine nuclear magnetic shielding as the first step when the absolute shielding of at least one reference standard is known satisfactorily. The problem of absolute shielding is not easy because it usually requires additional information from other than NMR experiments or quantum chemical calculations, and, often, it must be solved individually for each kind of magnetic nucleus. At least three methods are applied to the above investigations and are discussed in the present review. All the experimental attempts of shielding determination are equally important and precious because they allow for the reliable verification of the determined σ0 data in molecules.

2. NMR Chemical Shifts

NMR spectra are most applied in the qualitative analyses of chemical liquid compounds, and the shielding parameters are observed as the NMR chemical shifts (δi) measured relative to selected reference standards:
δ i = σ r e f σ i 1 σ r e f σ r e f σ i
where σref and σi are the shielding values of reference and investigated compounds, respectively. This is an excellent method for analytical chemistry but requires strict standardization of chemical shift measurements, as described by Harris et al. [20,21].
It is accepted in the scientific literature that NMR chemical shifts were discovered by Proctor and Yu when separate 14N signals from NH4+ and NO3- ions were observed at different resonance frequencies with a constant magnetic field [22]. The same issue of Physical Review also contains Dickinson’s publication [23], which describes the observation of 19F signals of fluorine compounds with various magnetic fields at the fixed resonance frequency. Both the research notes report the same finding of chemical shifts because the NMR experiment can be completed by either changing resonance frequencies with the stable magnetic field (B0) or having a constant electromagnetic frequency (ν0). One can use the variation of a frequency (or a magnetic field) as shown by Equations (6) and (7).
δ i = ν i ν r e f ν r e f   f o r   B 0 = c o n s t .
δ i = B r e f B i B r e f   f o r   ν 0 = c o n s t .
δi is the parameter earlier defined by Equation (5), and the above equations only show how the NMR measurement should be performed. The indexes “i” and “ref” are for the observed and reference nuclei. Chemical shifts are usually in the range of 10−6, and their values are always presented in “parts per million” (in ppm’s). In the first 25–30 years of NMR history, Equation (7) was mostly used for the determination of chemical shifts. Later, the superconducting magnets were introduced to standard NMR spectrometers, and Equation (6) became the formula recommended by IUPAC [20,21].
Let us note that Equations (6) and (7) have different orders of “i” and “ref” indexes in the formulas. It is true because the increase of radiofrequency (B0 = const.) is observed for less shielded nuclei while the magnetic field (ν0 = const.) and magnetic shielding are changing in the same direction. It arises from the basic description of an NMR experiment when the hv quantum of radio-frequency energy is absorbed by the single magnetic nucleus:
h ν = g X μ N 1 σ B
where gX is the g-factor of the observed nucleus (gX = µX/(IXµN), which describes its spin IX, magnetic moment µX, and µN is the nuclear magneton. Over one hundred stable atomic nuclei are magnetic and each kind of them has its magnetic moment and the appropriate gX value. Consequently, there are as many NMR spectroscopic methods as the number of magnetic nuclei. Some of them, e.g., 1H, 13C, 15N, 17O, 19F, and 31P NMR, are especially important in organic chemistry. The others are also intensively explored in experimental studies of chemistry and physics. Table 1 presents only the modest choice of magnetic nuclei and their NMR spectral parameters used for the discussed methods of shielding measurements.
As seen in Table 1, the selected well-known nuclides have mostly a spin number IX equal to ½. Such nuclei are spherical and have no quadrupole electrical momentum. Their NMR signals are usually sharp and permit more precise measurements of chemical shifts. Nevertheless, Table 1 also includes oxygen-17 and deuterium because of their importance in NMR spectroscopy and the presentation of some problems discussed in this review. The next column of Table 1 gives the natural abundance of the magnetic nuclei. Usually, the natural abundance above one percent guarantees good NMR spectra if a modern FT spectrometer is applied. A sample containing a much lower percentage of magnetic nuclei cannot be easily observed at the natural abundance, especially in the gas phase. The magnetic moments and gX factors describe the magnitude of nuclear magnetic properties; their improved values are given in [24].
There is also one more interesting column in Table 1, which reveals the absolute resonance frequencies of selected reference standards if the various NMR experiments are performed when the external magnetic field (B0) is fixed and giving a 1H NMR signal of 1% TMS in CDCl3 precisely at ΞH = 100.000000 MHz [20]. This idea of the absolute chemical shift comes from early double resonance experiments [33,34] and is very interesting because it unifies the important NMR parameter (chemical shift) for all magnetic nuclei. Unfortunately, it requires the double resonance method and involves many other experimental problems that the concept of ΞX has never been popular in everyday use of NMR practice. Limiting the discussion only to proton spectra, we should write the following ΞH values for isolated molecules of methane (100.0002847 MHz), ethane (100.0003593 MHz), and ethylene (100.0008017 MHz) based on existing δi results [35]. It is not so bad when everything is limited only to 1H NMR experiments. However, too many digits present at every NMR measurement make such a description of experimental data impractical and unpopular in everyday NMR experimental work. It is also important that for all nuclei other than protons, the ΞX measurement requires the simultaneous observation of 1H NMR experiment because from the definition ΞH (1% TMS in CD3Cl) must be equal to 100.000000 MHz in every case. It cannot be achieved using a standard NMR spectrometer.

3. Nuclear Magnetic Shielding in Molecules

3.1. General Insight

The range of chemical shifts depends on the electronic structure of atoms in molecules and is different for each kind of magnetic nuclei. Figure 2 illustrates the approximate area of chemical shifts on the scale of the nuclear magnetic shielding for the most popular nuclei and the positions of the recommended reference standards. The diagram refers to Table 1 but lacks two nuclei: 2H and 3He. The range of 2H NMR chemical shifts is practically the same as for 1H NMR if the minimal 2H/1H isotope effects in shielding are neglected [36]. Helium-3 is not present here because it does not form any chemical compound. There is only a special case of 3He NMR chemical shifts when helium atoms are encapsulated in fullerenes; then, the helium-3 shifts are −6.4 ppm for 3He@C60 and −27.9 ppm for 3He@C70 systems relative to pure gas 3He [37,38]. Theoretical models of helium-3 encapsulated in more complex carbon nanostructures predict even larger effects in 3He shielding [39,40], and all recent results on helium-3 NMR were overviewed by Kupka [41] and Krivdin [42].
Figure 2 has been prepared mostly using the available data on chemical shifts given by Bruker Almanac 2010 [43] and the shielding parameters of reference standards shown in Table 1. The blue bars approximately cover the ranges of 1H, 13C, 15N, 17O, 19F, 29Si, 31P, and 77Se magnetic shielding for diamagnetic chemical compounds. There are also marked three selected special points of shielding: σH(HI) = 43.92 ppm (cf. δH = −10.44 ppm for HI in 1H NMR [35,44]), σC(CI4) = 478.7 ppm (δC = −292.3 ppm for 13CI4 [45,46]), and σSi(SiI4) = 730.7 ppm (δSi = −351.7 ppm for 29SiI4 [47]). All the latter chemical shifts are measured relative to TMS and illustrate the unusual increase of shielding due to the presence of so-called “heavy atoms”, an iodide in this case [48,49]. The effect is especially enormous for 13C NMR spectra where the signal of carbon tetraiodide (CI4) is far away from the normal range of carbon-13 chemical shifts. It is so strong that even solvent molecules containing halogen atoms produce extremely large intermolecular shifts in 13C NMR spectra [50,51].
Figure 2 shows only a modest representation of over 100 stable magnetic nuclei, which can be observed by the NMR method, and it proves how different the features of shielding for various magnetic nuclei are. First, there is a different range of shielding for each element. A shielding range of only several ppm for proton spectra and almost three thousand for 77Se, and even more for heavier nuclei like 199Hg or 205Pb. Second, the particular area of shielding belongs to its characteristic region. It is so different if we compare the scales of 17O and 77Se shielding, less positive for oxygen-17 and more positive for selenium-77. Third, the reference standard is usually different for each magnetic nucleus, with one significant exception. Tetramethylsilane (TMS) was introduced as the internal reference standard to 1H NMR by Tiers in 1958 [52] and later accepted also for the referencing of 13C and 29Si spectra. Multinuclear NMR experiments also varied in many other details, such as the natural abundance of each nucleus or the magnitude of its nuclear magnetic moment. As seen in Figure 2, the whole multinuclear NMR spectroscopy is well described and unified by one common parameter—nuclear magnetic shielding.
Therefore, NMR shielding is an important parameter of molecules and can be determined from chemical shifts using Equation (5) if the shielding of at least one small molecule (σref) for each magnetic nuclide is known with satisfactory accuracy. There is a fundamental question of how the σref reliable value for the given nucleus can be determined. It seems that the most direct method is available just from advanced quantum chemical calculations.

3.2. Theoretical Approach to Shielding

The shielding ab initio calculations usually start with a molecule’s computed equilibrium geometry. It requires large basis sets for the satisfactory description of electron correlations in small molecules. The common solution to this problem applies the gauge-included atomic orbitals (GIAO) approach [12,13]. Various approximate methods are used for the shielding parameter calculations (frequently named improperly as the shielding constants): from HF (Hartree-Fock) to FCI (Full Configuration Interaction), MCSCF (Multi-Configuration Self-Consistent-Field), CC (Coupled Claster) approximation and MP (Møller–Plesset) perturbation theory [17]. The reference NMR molecule is usually the smallest one, and for this purpose, we should look for the most advanced ab initio methods with the perturbative-dependent basis set. Electron correlation effects should be calculated at CCSD (Coupled Cluster Singlets and Doubles) or CCSD(T) (Coupled Cluster Singlets and Doubles with Perturbative Triple Corrections) levels of theory [14,15]. The modern state-of-the-art magnetic shielding calculations are fairly advanced at the non-relativistic level [53]. They include all the intra- and intermolecular contributions to shielding [54], as such effects are always present in NMR experiments. To obtain calculated results that are ready for comparison with the experiment, it is necessary to consider the strong dependence of shielding on molecular geometry. It can be achieved at the ZPV (Zero-Point Vibration), or even better, including temperature effects up to the value of experimental work. Finally, the relativistic effects in shielding should be considered, mainly responsible for the extensive shielding of heavy atoms. It requires new methods designed for the description of electron interactions in molecules (four-component Dirac-Coulomb-Breit) [55]. Relativistic calculations are more expensive than any non-relativistic methods, and less advanced descriptions of electrons are usually applied, like HF or density functional theory (DFT) approximation. The relativistic contributions to shielding are mainly responsible for the extensive shielding of 1H in HI, 13C in CI4, and 29Si in SiI4, as shown in Figure 2. It is the so-called HALA effect (heavy-atom effect on light atoms).
Intermolecular effects are difficult for theoretical treatment [17,54], and for this reason, it is better to avoid such effects by performing NMR experiments in the gas phase. Recently, a precise comparison of experimental and calculated shielding for small molecules CH4-nFn has been presented [19]. Earlier, a similar comparison was performed for NF3, PF3, and AsF3 compounds, where the change of absolute 19F shielding in liquid CFCl3 to 197.07 ppm was suggested [56]. It is questionable and requires new experimental proof because the existing last measurement gave 190.0 ppm if a bulk susceptibility correction is excluded [57,58]. The recent comparison of 19F shielding in CH4−nFn [19] rather confirms the old experimental result for liquid CFCl3.
The new methods of shielding calculations are powerful for small molecules, but let us compare the first approximation of proton shielding in the H2 molecule given by Ramsey in 1950 [3] (26.8 ppm) with the available calculated results: Sundholm and Gauss CCSD(T) value is 26.2983 ppm at T = 298 K [59] and Jaszuński et al. CCSD shielding is 26.2980 at T = 300 K [60]. Ramsey’s prediction of 1H shielding in H2 molecule is overestimated only by 1.9 percent. The above example of magnetic shielding in the hydrogen molecule also illustrates the importance of rovibrational effects in shielding [54] that were included in the new calculations and effectively diminished the final H2 result [59,60].
We are mostly interested in the nuclear magnetic shielding observed in small molecules, but quantum chemical calculations are also possible for larger molecular objects if some further approximations are applied. In such a case, the most popular method is DFT (Density-Functional Theory) [16], which can also be applied to studies of shielding at the four-component relativistic level [46]. Another approach is offered by the ONIOM method, in which the shielding calculations are performed with different approximation levels for the selected layers of electrons in a molecule [61]. It permits more precise calculations of shielding in larger molecules, simultaneously saving computer time. Other mixed methods of shielding calculations are applied for solids. Recently, it has been shown that a quantum mechanics/molecular mechanics (QM/MM) method can predict the solid-state NMR shielding for molecular crystals [62].

4. Determination of Absolute Shielding

4.1. The Ramsey–Flygare Method

Equations (2) and (3) reveal that the magnetic shielding in a molecule consists of the positive diamagnetic part (σdia), which depends only on the ground electronic state of the molecule. This shielding part can be relatively easily calculated using modern quantum chemical methods. The second term of shielding described by Equation (4) is much more complex for calculations but is related to the microwave nuclear spin–rotation tensor (C), which is reduced to a coupling constant (cI) for linear molecules. It happens because, in such a case, the paramagnetic shielding part parallel to the bond axis (σp) is equal to zero. Remains the perpendicular part of paramagnetic shielding (σp), which, as shown by Flygare [63,64], is related to the spin-rotation constant for a nucleus in a diatomic molecule as follows:
σ p = 3 2 σ p = m p 2 m e g X c I B 3 2 μ 0 4 π e 2 3 m e Z r
In the above equation, mp is the proton mass, B is the rotational constant of the molecule, Z is the atomic number of the neighbor atom in the molecule, and r is the internuclear distance. The rest of the parameters (me, µ0, e, and gX) are the same as previously defined in Equations (3), (4), and (8), respectively. As seen in Equation (9), there is a link between the spin-rotation interaction of microwave spectroscopy and the absolute nuclear magnetic shielding observed in NMR [65]. However, the procedure of its use requires more additional work: first, the experimental spin-rotation constant must be refined from its vibrational corrections, and then the equilibrium value of shielding is obtained as shown in Equation (9), and finally, the rovibrational corrections should be added to shielding to obtain the σref value at the given temperature (usually 300 K). This relationship was exploited for the determination of absolute nuclear shielding in hydrogen fluoride (HF) delivering σref(19F) [57], in carbon monoxide molecules (13C16O and 12C17O) delivering the important σref(13C) [66,67,68] and σref(17O) values [69,70], respectively. The oxygen-17 case contains an interesting feature: it was initially based on the rotational constant of the 12C17O molecule obtained from the observation of the J = 0 1 l transition in the rotational spectrum observed from interstellar space [71] until a more accurate measurement of the same rotational constant was available from the laboratory experiment in 2002 [72]. It is important to note that the Ramsey–Flygare method is not limited to linear and diatomic molecules [64]. It was possible to obtain the σref(15N) value for ammonia enriched in nitrogen-15 [73] and σref(31P) for PH3 [74].

4.2. Methods Based on 1H NMR Signal of Liquid Water and Shielding Transfers

The second method of shielding determination is based on the proton reference signal from liquid water in a spherical sample at a temperature of 34.7 °C, σref(1H2Oliq.sph) = 25.790(14) ppm [75]. Two experiments standardized this 1H signal: first, the simultaneous observation of the frequencies of an electronic and a proton transition in atomic hydrogen [76]; second, the simultaneous reading of 1H NMR signals from atomic hydrogen and pure water, both in the spherical samples [77]. The σref(1H) value can be directly applied to the scale of proton shielding, but the use of a spherical sample at elevated constant temperature is rather inconvenient in everyday NMR experimental practice. The 1H signal of H2O is extremely dependent on temperature. Figure 3 presents the positions of three isolated molecules (H2, H2O, and TMS) and the sample containing 1% of TMS in CDCl3 on the shielding scale relative to the σref(1H) reference signal. It permits the fast measurements of absolute shielding in 1H NMR spectra for gaseous and liquid compounds using 1H chemical shifts.
The actual σref(1H) parameter can also be used for shielding referencing other than proton nuclei as the universal reference standard of shielding, but it requires knowledge of the other nuclear magnetic dipole moments and the double-resonance experiments. There is also another possibility of transferring the shielding scale from one nucleus to another present in the same molecule, exploring the relaxation T1 time measurements in the gas phase. The latter method is described in detail by Jameson [75] and applied for the determination of the σref(29Si) in a gaseous mixture of SiH4 and SiF4 molecules [78], and the σref(77Se) parameter in H2Se and SeF6 molecules [32].

4.3. Helium-3 Atom as the Universal Shielding Reference

For many reasons, an isolated helium-3 atom is probably the best candidate for the universal reference standard of magnetic shielding in multinuclear NMR spectroscopy. First, the gas phase 3He NMR experiments delivered the resonance frequency of an isolated helium-3 atom at the stable external magnetic field [79]. Second, the 3He measurement is independent of temperature, and no rovibrational corrections are needed for further increase in accuracy. Third, the quantum chemical calculations deliver the most accurate value of the 3He atom, and this result can be accepted as the σref(3He) reference standard equal to 59.967 029(23) ppm [26].
The final step requires just the transfer of shielding from helium-3 to other magnetic nuclei experiments, and such a measurement can be taken using double resonance methods like those previously explored by McFarlane [33,34]. The comparison of Equation (8) for the 3He and another X nucleus in the same magnetic field (B) leads to Equation (10) and means the transfer of shielding from the helium-3 to the X nucleus. The experiment, according to Equation (10), requires only an NMR spectrometer and the nuclear dipole moments (µHe and µX) [80,81,82]:
σ X = 1 ν X ν H e · μ H e μ X · I X I H e 1 σ H e
IX and IHe are the nuclear spin numbers of the X and helium-3 nuclei, respectively. As seen in Equation (10), the knowledge of nuclear magnetic moments is crucial for the determination of shielding, and this problem is generally discussed in the next section.

5. Nuclear Magnetic Dipole Moments

As mentioned in the Introduction, the present state-of-the-art calculations of shielding are very powerful [12,13,14,15,16,17,18] and can often be used to improve our knowledge of nuclear magnetic shielding in molecules. The best theoretical results of magnetic shielding in molecules were also applied for the improvement of nuclear magnetic moments [83]. It was necessary because the existing results of the International Atomic Energy Agency (IAEA) at that time [84] were not reliable for all the heavier nuclei beyond hydrogen and helium-3. The determination of more accurate values of nuclear dipole moments was performed using the best available calculated results of shielding preferably in one molecule (σX and σY) and the gas phase measurements of resonance frequencies for the same isolated molecule (νY and νY) where the index X refers to the reference nucleus, and Y is for the studied nucleus:
μ Y = ν Y ν X · ( 1 σ X ) ( 1 σ Y ) · I Y I X μ X
This way, the nuclear magnetic moment of the X reference nucleus was transferred to the other Y nucleus using NMR results for isolated small molecules and state-of-the-art shielding calculations performed for the same molecules. Protons and helium-3 were mostly used as the reference nuclei because their magnetic moments were repeatedly verified [26,80,81,84,85,86,87,88,89,90,91,92,93,94]. In some cases, the reference molecules could not contain protons, and other nuclei served as the reference X nuclei. The results obtained from the application of Equation (11) to isolated molecules are summarized in Table 2.
An interesting case of boron nuclei (10B and 11B) should be mentioned here, as the mixtures of 3He and BF3 gases were observed in the same gaseous samples [87]. The experiments have shown that such an approach with the direct comparison of shielding with helium-3 is precious, and later, it was also exploited for the measurements of magnetic moments of rare gases [93,94].
A bit less accurate results for heavier nuclei were available when the new experiments could not be performed in the gas phase. The magnetic dipole moments determined from NMR experiments in aqueous solutions and supported by the calculations of hydrated ions deliver fairly good results with the application of the same approach based on Equation (11). The improvements of µX data were possible for numerous nuclei, mostly due to more accurate ab initio calculations: for 6Li, 7Li, 23Na, 39K, 41K, 85Rb, 87Rb, and 133Cs [95]; for 9Be, 25Mg, 43Ca, 87Sr, 135Ba, and 137Ba [96]; for 27Al, 69Ga, 71Ga, 113In, and 115In [97]; and 45Sc, 89Y, 138La and 139La nuclei [98]. All the above-cited data on nuclear magnetic dipole moments [83,86,87,88,89,90,91,92,93,94,95,96,97,98] were recognized by the International Nuclear Data Committee (INDC) as the standards and published in the IAEA documents [99].
Recently, the search for better and more accurate nuclear magnetic moments has continued. The most important study was published by Harding et al. [100]. The authors determined the nuclear magnetic moment with much higher accuracy for an unstable 26Na nucleus (1.1 s half-life time) using an improved version of the β-technique NMR combined with ab initio calculation of nuclear magnetic shielding performed for the stable 23Na reference. New studies and further possible improvements in the determination of magnetic moments for stable nuclei also appeared [101,102,103,104], and three review papers on the same subject were published [105,106,107].

6. Universal Approach to Shielding Measurements

Nuclear magnetic shielding in molecules can be determined using a few different methods, as shown in Section 4. It is very helpful for the cross-checking of available experimental results and permits a better understanding of nuclear magnetic shielding. In our opinion, the application of Equation (10) is very promising because the number of reliable data for nuclear magnetic moments is quickly growing [83,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107]. We also have an excellent reference standard of shielding, which is independent of temperature, the isolated helium-3 atom. However, the idea of the multinuclear experiments requires the simultaneous measurements of two different types of nuclei, the studied (X) and the reference nucleus (3He). Fortunately, it can be achieved using any standard NMR spectrometer with the deuterium lock system when first, the selected 2H NMR signal of lock solvent is calibrated by applying Equation (10), and then the deuterium lock solvent is used as the secondary reference of nuclear magnetic shielding [82]. Let us note that this way, all the shielding measurements remain referenced to an isolated helium-3 atom, σref(3He) = 59.9670 ppm [26]. In our opinion, an isolated helium-3 atom is the best choice for the primary reference standard of shielding in multinuclear NMR experiments. The above method is schematically presented in Figure 4.
Let us note that the new method of shielding measurements is well-calibrated with an isolated helium-3 atom and permits for using any standard NMR spectrometer in the double-resonance experiments according to Equation (12) where νD, µD, and σD are the frequency, 2H magnetic moment, and calibrated 2H shielding of liquid lock solvent, respectively; ID = 1.
σ X = 1 ν X ν D · μ D μ X · I X · 1 σ D
Equation (12) can be generally used for all NMR samples if the lock solvent is separated from the observed substance. It only replaces the measurement of δi in Equation (5) for the direct reading of shielding (σi in Equation (5)). Both the parameters (δi and σi) are based on the same observation of experimental frequencies represented by νi in Equation (6). In the present method of shielding measurements, nothing is modified inside the sample shielding (σi in Equation (5)). Therefore, this method, described by Equation (12), can be safely used for the measurements of shielding in paramagnetic samples.
The σD values of Equation (12) were calibrated for numerous signals from liquid lock solvents and published in original papers [82,107]. Table 3 below shows only the most popular solvents that are frequently used in NMR laboratories.
Equation (12) contains many constants that can be consolidated into one number for practical applications, for example, in the most popular 1H and 13C NMR experiments. The problem is then simplified to the reading of absolute frequencies (νH) for protons or (νC) for carbons-13 and simultaneously for deuterium nuclei (νD) in lock solvent [82]:
σ H = 1 ν H ν D · 0.153506104 · 1 σ D
σ C = 1 ν C ν D · 0.610389782 · 1 σ D
At this point, everything is ready for the recording of multinuclear NMR spectra with the scale of magnetic shielding instead of chemical shifts. We are applying Equations (13) and (14) and the σD parameters of Table 3 to get such a spectrum using a standard NMR spectrometer with a deuterium lock system. It is illustrated in Figure 5, where the 13C and 1H spectra of ethyl crotonate (CH3CH=CHCOOC2H5) dissolved in CDCl3 are shown as examples. It was possible to record them automatically due to a small modification of software in the computer of our 500 MHz VarianINOVA NMR spectrometer.
As shown, the measurements of nuclear magnetic shielding with the application of 1H and 13C NMR spectra are easy and comfortable for everyone. It does not require any special equipment and delivers valuable information on nuclear magnetic shielding. One feature of shielding measurements is especially interesting: the new method allows the measurement of the first-order isotope effects in shielding hydrogen isotopologues [108], which was impossible before [109]. All the isotope effects of hydrogen are illustrated by the original results presented in Table 4. Unexpectedly, the primary isotope effects in shielding 0Δ(2/1H) observed for H2, HD, and D2 molecules are stronger than the secondary isotopic effects 1Δ(2/1H) for the same molecules. It is an important measurement because standard NMR methods cannot be used to determine the primary isotopic effects in shielding. In this case, the direct measurements of 1H and 2H shielding were performed for the gaseous mixtures of hydrogen molecules with helium-3. All the frequency measurements were extrapolated to the zero-density limit, and Equation (10) was applied for the determination of all 1H and 2H shielding data in H2, HD, and D2 molecules [108].
The measurements of 13C magnetic shielding relative to σref(3He) were also extended on solid samples in MAS (Magic Angle Spinning) NMR spectra [111]. It was possible to use spherical samples of liquid TMS, a solution of 1% TMS in CDCl3, and solid fullerene C60 for the 13C shielding measurements in the standard NMR experiments, and the same samples were also observed by the MAS NMR method. Then, the 13C shielding values of popular MAS references like glycine, hexamethylbenzene, and adamantane were obtained by reading carbon-13 chemical shifts for solids [111].
Recently, the 1H, 13C, and 14N magnetic shielding parameters were measured for emodin and chuanxiongzine, plant products that have pharmacological properties and are frequently used in traditional Chinese medicine [112]. The direct measurements of 1H and 13C shielding were also applied in the studies of daidzein and puerarin which have natural anti-oxidant properties [113].

7. Conclusions

The origin of nuclear magnetic shielding in diamagnetic molecules is briefly discussed in the present review article. As shown, the important properties of molecules can be observed via chemical shifts in NMR spectra or can be calculated using advanced quantum chemical methods. The relations between shielding parameters and chemical shifts are rather complex because chemical shifts are separately defined for various magnetic nuclei by their reference standards. On the other hand, the measurements of shielding are badly needed for the direct comparison of experimental and calculated shielding values, also known as shielding constants. An isolated helium-3 atom is the natural choice for the universal reference standard of shielding. It can be easily applied in multinuclear NMR spectroscopy when its shielding is encoded into pure deuterated liquids that are used for the precise stabilization of the external magnetic field (lock system in NMR spectrometers). The method of shielding measurements is completed and can be applied to any liquid or gaseous NMR sample.
As shown in the previous section, the chemical shifts can be completely replaced in the future by the measurement of shielding parameters, and this alternative method of standardization of NMR spectra has numerous experimental advantages. The most important features of the new method are listed as follows. First, it unifies multinuclear methods into one NMR spectroscopy because the values of magnetic shielding have the same meaning independently of observed nuclei. Second, there is no need to use any additional reference standard if the NMR experiment is carried out with the calibrated 2H solvent, as shown in Table 3, because the same original reference standard of shielding is always preserved—an isolated helium-3 atom [82]. Third, the new method allows for the first time the measurement of the first-order isotope effects in shielding, as it has already been shown for hydrogen isotopologues [108]. Fourth, the measurements of 13C shielding relative to σref(3He) can be extended on solid samples in MAS NMR spectra, as shown in ref. [111]. Fifth, the measurements of shielding values are performed with the same precision as the standard determination of chemical shifts because they are based on the same reading of resonance frequencies, cf. Equation (6) vs. Equations (12)–(14). Last but not least, the determined shielding parameters can always be converted back into chemical shifts using Equation (5) if necessary and without the use of any additional reference standard. The simultaneous measurements of both the shielding parameters and appropriate chemical shifts are always available from the same single NMR experiment.
To summarize, the direct measurements of nuclear magnetic shielding are already available for selected light nuclei, and they are relatively easy to apply if the calibrated lock solvents are used as the secondary reference standards. It permits the use of every NMR spectrometer with a 2H lock system for required double-resonance experiments. We believe that progress in the knowledge of accurate nuclear moments will be growing fast, and the exact measurement of shielding will be available for more and more magnetic nuclei soon.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lamb, W.E. Internal Diamagnetic Fields. Phys. Rev. 1941, 60, 817–819. [Google Scholar] [CrossRef]
  2. Pople, J.A.; Schneider, W.G.; Bernstein, H.J. High-Resolution Nuclear Magnetic Resonance; McGraw-Hill: New York, NY, USA, 1959; Chapter 7; pp. 165–183. [Google Scholar] [CrossRef]
  3. Ramsey, N.F. Magnetic Shielding of Nuclei in Molecules. Phys. Rev. 1950, 78, 699–703. [Google Scholar] [CrossRef]
  4. Ramsey, N.F. Dependence of Magnetic Shielding of Nuclei upon Molecular Orientation. Phys. Rev. 1951, 83, 540–541. [Google Scholar] [CrossRef]
  5. Saika, A.; Slichter, C.P. A Note on the Fluorine Resonance Shifts. J. Chem. Phys. 1954, 22, 26–28. [Google Scholar] [CrossRef]
  6. Pople, J.A. Nuclear Magnetic Resonance in Diamagnetic Materials. The Theory of Chemical Shifts. Disc. Faraday Soc. 1962, 34, 7–14. [Google Scholar] [CrossRef]
  7. Karplus, M.; Pople, J.A. Theory of Carbon NMR Chemical Shifts in Conjugated Molecules. J. Chem. Phys. 1963, 38, 2803–2807. [Google Scholar] [CrossRef]
  8. Baker, M.R.; Anderson, C.H.; Ramsey, N.F. Nuclear Magnetic Antishielding of Nuclei in Molecules. Magnetic Moments of F19, N14 and N15. Phys. Rev. 1964, 133, A1533–A1536. [Google Scholar] [CrossRef]
  9. Kalinowski, H.-O.; Berger, S.; Braun, S. Carbon-13 NMR Spectroscopy, 1st ed.; John Wiley & Sons: Chichester, UK; New York, NY, USA; Brisbane, Australia; Toronto, ON, Canada; Singapore, 1988; pp. 91–92. ISBN 0-471-96763-7. [Google Scholar] [CrossRef]
  10. Jameson, C.J. Effects of Intermolecular Interactions and Intramolecular Dynamics on Nuclear Resonance. Bull. Magn. Reson. 1981, 3, 3–29. Available online: https://ismar.org/wp-content/uploads/2021/09/BMR_03_003-028_1981.pdf (accessed on 28 March 2024).
  11. Jameson, C.J. Fundamental Intramolecular and Intermolecular Information from NMR in the Gas Phase. In New Developments in NMR: Gas Phase NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 1; pp. 1–51. [Google Scholar] [CrossRef]
  12. Woliński, K.; Hinton, J.F.; Pulay, P. Efficient Implementation of the Gauge-Independent Atomic Orbital Method for NMR Chemical Shift Calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  13. Gauss, J. Calculation of NMR Chemical Shifts at Second-order Many-body Perturbation Theory Using Gauge-including Atomic Orbitals. Chem. Phys. Lett. 1992, 191, 614–620. [Google Scholar] [CrossRef]
  14. Watts, J.D.; Gauss, J.; Bartlett, R.J. Coupled-cluster Methods with Noniterative Triple Excitations for Restricted Open-shell Hartree–Fock and other General Single Determinant Reference Functions. Energies and Analytical Gradients. J. Chem. Phys. 1993, 98, 8718–8733. [Google Scholar] [CrossRef]
  15. Gauss, J.; Stanton, J.F. Coupled-cluster Calculations of Nuclear Magnetic Resonance Chemical Shifts. J. Chem. Phys. 1995, 103, 3561–3577. [Google Scholar] [CrossRef]
  16. Bühl, M.; Kaupp, M.; Malkina, O.L.; Malkin, V.G. The DFT Route to NMR Chemical Shifts. J. Comput. Chem. 1999, 20, 91–105. [Google Scholar] [CrossRef]
  17. Halgaker, T.; Jaszuński, M.; Ruud, K. Ab Initio Methods for the Calculation of NMR Shielding and Indirect Spin−Spin Coupling Constants. Chem. Rev. 1999, 99, 293–352. [Google Scholar] [CrossRef] [PubMed]
  18. Vaara, J. Theory and Computation of Nuclear Magnetic Resonance Parameters. Phys. Chem. Chem. Phys. 2007, 9, 5399–5418. [Google Scholar] [CrossRef]
  19. Jackowski, K.; Słowiński, M.A. Searching for the Best Values of NMR Shielding and Spin-Spin Coupling Parameters: CH4-nFn Series of Molecules as the Example. Molecules 2023, 28, 1499. [Google Scholar] [CrossRef] [PubMed]
  20. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.M.; Goodfellow, R.; Granger, P. NMR Nomenclature. Nuclear Spin Properties and Conventions for Chemical Shifts (IUPAC Recommendations 2001). Pure Appl. Chem. 2001, 73, 1795–1818, reprinted in Magn. Reson. Chem. 2002, 40, 489–505. [Google Scholar] [CrossRef]
  21. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.M.; Granger, P.; Hoffman, R.E.; Zilm, K.W. Further Conventions for NMR Shielding and Chemical Shifts (IUPAC Recommendations). Pure Appl. Chem. 2008, 80, 59–84, reprinted in Magn. Reson. Chem. 2008, 46, 582–598. [Google Scholar] [CrossRef]
  22. Proctor, W.G.; Yu, F.C. The Dependence of a Nuclear Magnetic Resonance Frequency upon Chemical Compound. Phys. Rev. 1950, 77, 717. [Google Scholar] [CrossRef]
  23. Dickinson, W.C. Dependence of the F19 Nuclear Resonance Position on Chemical Compound. Phys. Rev. 1950, 77, 736–737. [Google Scholar] [CrossRef]
  24. Jackowski, K.; Garbacz, P. Nuclear Magnetic Moments and Measurements of Shielding. In New Developments in NMR: Gas Phase NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 3; pp. 93–125. ISBN 978-1-78262-722-7. [Google Scholar] [CrossRef]
  25. Makulski, W.; Jackowski, K. 1H, 13C and 29Si Magnetic Shielding in Gaseous and Liquid Tetramethylsilane. J. Magn. Reson. 2020, 313, 106716. [Google Scholar] [CrossRef] [PubMed]
  26. Wehrli, D.; Spyszkiewicz-Kaczmarek, M.; Puchalski, M.; Pachucki, K. QED Effects on the Nuclear Magnetic Shielding of 3He. Phys. Rev. Lett. 2021, 127, 263001. [Google Scholar] [CrossRef] [PubMed]
  27. Jameson, C.J.; Jameson, A.K.; Oppusunggu, D.; Wille, S.; Burell, P.M. 15N Nuclear Magnetic Shielding Scale from Gas Phase Studies. J. Chem. Phys. 1981, 74, 81–88. [Google Scholar] [CrossRef]
  28. Komorovsky, S.; Repisky, M.; Malkin, E.; Ruud, K.; Gauss, J. Communication: The Absolute Shielding Scales of Oxygen and Sulfur Revisited. J. Chem. Phys. 2015, 142, 091102. [Google Scholar] [CrossRef] [PubMed]
  29. Makulski, W.; Wilczek, M.; Jackowski, K. 17O and 1H NMR spectral parameters in isolated water molecules. Phys. Chem. Chem. Phys. 2018, 20, 22468–22476. [Google Scholar] [CrossRef] [PubMed]
  30. Jameson, C.J.; Jameson, A.K.; Honarbakhsh, J. 19F Nuclear Magnetic Shielding Scales from Gas Phase Studies. J. Chem. Phys. 1984, 81, 5266–5267. [Google Scholar] [CrossRef]
  31. Jameson, C.J.; De Dios, A.; Jameson, A.K. Absolute Shielding Scale for 31P from Gas-phase NMR Studies. Chem. Phys. Lett. 1990, 167, 575–582. [Google Scholar] [CrossRef]
  32. Jameson, C.J.; Jameson, A.K. Concurrent 19F and 77Se or 19F and 125Te NMR T1 Measurements for Determination of 77Se and 125Te Absolute Shielding Scales. Chem. Phys. Lett. 1987, 135, 254–259. [Google Scholar] [CrossRef]
  33. McFarlane, W. A Magnetic Double Resonance Study of Some Aryl and Pentafluoroaryl Mercury Compounds. J. Chem. Soc. A 1968, 2280–2285. [Google Scholar] [CrossRef]
  34. McFarlane, W. The Determination of Carbon-13 Chemical Shifts in Esters by Hetero-nuclear Magnetic Double Resonance. J. Chem. Soc. B 1969, 28–30. [Google Scholar] [CrossRef]
  35. Garbacz, P.; Jackowski, K.; Makulski, W.; Wasylishen, R.E. Nuclear Magnetic Shielding for Hydrogen in Selected Isolated Molecules. J. Phys. Chem. A 2012, 118, 11896–11904. [Google Scholar] [CrossRef] [PubMed]
  36. Raynes, W.T.; Panteli, N. The Extraction of a Nuclear Magnetic Shielding Function for the H2 Molecule from Spin-Rotation and Isotope Shift Data. Mol. Phys. 1983, 48, 439–449. [Google Scholar] [CrossRef]
  37. Saunders, M.; Jiménez-Vázquez, H.A.; Mroczkowski, S.; Freedberg, D.J.; Anet, F.A.L. Probing the Interior of Fullerenes by 3He NMR Spectroscopy of Endohedral 3He@C60 and 3He@C70. Nature 1994, 367, 256–258. [Google Scholar] [CrossRef]
  38. Saunders, M.; Jiménez-Vázquez, H.A.; Bangerter, B.W.; Cross, R.J.; Mroczkowski, S.; Freedberg, D.J.; Anet, F.A.L. 3He NMR. A Powerful Tool for Following Fullerene Chemistry. J. Am. Chem. Soc. 1994, 116, 3621–3622. [Google Scholar] [CrossRef]
  39. Guan-Wu, W.; Xin-Hao, Z.; Huan, Z.; Qing-Xiang, G.; Yun-Dong, W. Accurate Calculation, Prediction, and Assignment of 3He NMR. Chemical Shifts of Helium-3-Encapsulated Fullerenes and Fullerene Derivatives. J. Org. Chem. 2003, 68, 6732–6738. [Google Scholar] [CrossRef] [PubMed]
  40. Kupka, T.; Stachów, M.; Stobiński, L.; Kaminsky, J. 3He NMR: From Free Gas to its Encapsulation in Fullerene. Magn. Reson. Chem. 2013, 51, 463–468. [Google Scholar] [CrossRef] [PubMed]
  41. Kupka, T. Noble Gases as Magnetic Probes in Fullerene Chemistry. eMagRes 2016, 5, 959–966. [Google Scholar] [CrossRef]
  42. Krivdin, L.B. An Overview of Helium-3 NMR: Recent Developments and Applications. Prog. Nucl. Magn. Reson. Spect. 2023, 136–137, 83–109. [Google Scholar] [CrossRef]
  43. Bruker Almanac 2010, pp. 73–76. Available online: http://www.pascal-man.com/pulseprogram/Almanac2010.pdf (accessed on 28 March 2024).
  44. Schneider, W.G.; Bernstein, H.J.; Pople, J.A. Proton Magnetic Resonance Chemical Shift of Free (Gaseous) and Associated (Liquid) Hydride Molecules. J. Chem. Phys. 1958, 28, 601–607. [Google Scholar] [CrossRef]
  45. Spiesecke, H.; Schneider, W.G. Effect of Electronegativity and Magnetic Anisotropy of Substituents on C13 and H1 Chemical Shifts in CH3X and CH3CH2X Compounds. J. Chem. Phys. 1961, 35, 722–731. [Google Scholar] [CrossRef]
  46. Samultsev, D.O.; Rusakov, Y.Y.; Krivdin, L.B. Normal Halogen Dependence of 13C NMR Chemical Shifts of Halogenomethanes Revisited at the Four-Component Relativistic Level. Magn. Reson. Chem. 2016, 54, 787–792. [Google Scholar] [CrossRef] [PubMed]
  47. Niemann, U.; Marsmann, H.C. Si-Kernresonanzmessungen an Siliciumhalogeniden. 29Si NMR Studies on Silicon Halogenides. Z. Naturforsch. B 1975, 30, 202–206. [Google Scholar] [CrossRef]
  48. Mason, J. The Interpretation of Carbon Nuclear Magnetic Resonance Shifts. J. Chem. Soc. A 1971, 1038–1047. [Google Scholar] [CrossRef]
  49. Wehrli, F.W.; Wirthlin, T. Interpretation of Carbon-13 NMR Spectra; Heyden & Son Ltd.: London, UK; Philadelphia, PA, USA; Reine, Norway, 1978; pp. 34–36. [Google Scholar] [CrossRef]
  50. Jackowski, K.; Kęcki, Z. Molecular Interaction of Hydrocarbon Alkyl Groups with Carbon Tetrachloride Observed in 13C NMR chemical shifts. Ber. Bunsenges. Phys. Chem. 1981, 85, 143–145. [Google Scholar] [CrossRef]
  51. Jackowski, K. Effect of Collision Interactions on 13C Nuclear Magnetic Shielding in Monosubstituted Benzenes in Halogen-Containing Solvents. Org. Magn. Reson. 1984, 22, 263–268. [Google Scholar] [CrossRef]
  52. Tiers, G.D.V. Proton Nuclear Resonance Spectroscopy: I. Reliable Shielding Values by Internal Referencing with Tetramethylsilane. J. Phys. Chem. 1958, 62, 1151–1152. [Google Scholar] [CrossRef]
  53. Antušek, A.; Jaszuński, M. Accurate Non-Relativistic Calculations of NMR Shielding Constants. In New Developments in NMR: Gas Phase NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 6; pp. 186–217. [Google Scholar] [CrossRef]
  54. Faber, R.; Kaminsky, J.; Sauer, S.P.A. Rovibrational and Temperature Effects in Theoretical Studies of NMR Parameters. In New Developments in NMR: Gas Phase NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 7; pp. 218–266. [Google Scholar] [CrossRef]
  55. Repisky, M.; Komorovsky, S.; Bast, R.; Ruud, K. Relativistic Calculations of Nuclear Magnetic Resonance Parameters. In New Developments in NMR: Gas Phase NMR; Price, W.S., Jackowski, K., Jaszuński, M., Eds.; Royal Society of Chemistry: Cambridge, UK, 2016; Chapter 8; pp. 267–303. [Google Scholar] [CrossRef]
  56. Field-Theodore, T.E.; Olejniczak, M.; Jaszuński, M.; Wilson, D.J.D. NMR Shielding Constants in Group 15 Trifluorides. Phys. Chem. Chem. Phys. 2018, 20, 23025–23033. [Google Scholar] [CrossRef]
  57. Hindermann, D.K.; Cornwell, C.D. Vibrational Corrections to the Nuclear-Magnetic Shielding and Spin–Rotation Constants for Hydrogen Fluoride. Shielding Scale for 19F. J. Chem. Phys. 1968, 48, 4154–4161. [Google Scholar] [CrossRef]
  58. Jackowski, K.; Kubiszewski, M.; Makulski, W. 13C and 19F Nuclear Magnetic Shielding and Spin-Spin Coupling in Gaseous Fluoromethane-d3. J. Mol. Struct. 2002, 614, 267–272. [Google Scholar] [CrossRef]
  59. Sundholm, D.; Gauss, J. Isotope and Temperature Effects on Nuclear Magnetic Shieldings and Spin-Rotation Constants Calculated at the Coupled-Cluster Level. Mol. Phys. 1997, 92, 1007–1014. [Google Scholar] [CrossRef]
  60. Jaszuński, M.; Łach, G.; Strasburger, K. NMR Shielding Constants in Hydrogen Molecule Isotopomers. Theor. Chem. Acc. 2011, 129, 325–330. [Google Scholar] [CrossRef]
  61. Karadakov, P.B.; Morokuma, K. ONIOM as an Efficient Tool for Calculating NMR Chemical Shielding Constants in Large Molecules. Chem. Phys. Lett. 2000, 317, 589–596. [Google Scholar] [CrossRef]
  62. Poidevin, C.; Stoychev, G.L.; Riplinger, C.; Auer, A.A. High Level Electronic Structure Calculation of Molecular Solid-State NMR Shielding Constants. J. Chem. Theory Comput. 2022, 18, 2408–2417. [Google Scholar] [CrossRef] [PubMed]
  63. Flygare, W.H. Spin—Rotation Interaction and Magnetic Shielding in Molecules. J. Chem. Phys. 1964, 41, 793–800. [Google Scholar] [CrossRef]
  64. Gierke, T.D.; Flygare, W.H. An Empirical Evaluation of the Individual Elements in the Nuclear Diamagnetic Shielding Tensor by the Atom Dipole Methodology. J. Am. Chem. Soc. 1972, 94, 7277–7283. [Google Scholar] [CrossRef]
  65. Bryce, D.L.; Wasylishen, R.E. Microwave Spectroscopy and Nuclear Magnetic Resonance Spectroscopy-What Is the Connection? Acc. Chem. Res. 2003, 36, 327–334. [Google Scholar] [CrossRef]
  66. Ozier, I.; Crapo, L.M.; Ramsey, N.F. Spin Rotation Constant and Rotational Magnetic Moment of 13C16O. J. Chem. Phys. 1968, 49, 2314–2321. [Google Scholar] [CrossRef]
  67. Jameson, C.J.; Jameson, A.K. Gas Phase 13C Chemical Shifts in the Zero Pressure Limit: Refinements to the Absolute Shielding Scale for 13C. Chem. Phys. Lett. 1987, 134, 461–466. [Google Scholar] [CrossRef]
  68. Raynes, W.T.; McVey, R.; Wright, S.J. An Improved Carbon-13 Nuclear Shielding Scale. J. Chem. Soc. Faraday Trans. 2 1989, 85, 759–763. [Google Scholar] [CrossRef]
  69. Wasylishen, R.E.; Bryce, D.L. A Revised Experimental Absolute Magnetic Shielding Scale for Oxygen. J. Chem. Phys. 2002, 117, 10061–10066. [Google Scholar] [CrossRef]
  70. Wasylishen, R.E.; Mooibroek, S.; Mcdonald, J.B. A More Reliable Oxygen-17 Absolute Chemical Shielding Scale. J. Chem. Phys. 1984, 81, 1057–1059. [Google Scholar] [CrossRef]
  71. Frerking, M.A.; Langer, W.D. A Measurement of the Hyperfine Structure of C17O. J. Chem. Phys. 1981, 74, 6990–6991. [Google Scholar] [CrossRef]
  72. Cazzoli, G.; Dore, L.; Puzzarini, C.; Beninati, S. Millimeter- and Submillimeter-Wave Spectrum of C17O. Rotational Hyperfine Structure Analyzed Using the Lamb-Dip Techniques. Phys. Chem. Chem. Phys. 2002, 4, 3575–3577. [Google Scholar] [CrossRef]
  73. Kukolich, S.G. Proton Magnetic Shielding Tensors from Spin-Rotation Measurements on H2CO and NH3. J. Am. Chem. Soc. 1975, 97, 5704–5707. [Google Scholar] [CrossRef]
  74. Davies, P.B.; Neumann, R.M.; Wofsy, S.C.; Klemperer, W. Radio-Frequency Spectrum of Phosphine (PH3). J. Chem. Phys. 1971, 55, 3564–3568. [Google Scholar] [CrossRef]
  75. Jameson, C.J. Chemical Shift Scales on an Absolute Basis. In Encyclopedia of Nuclear Magnetic Resonance; Grant, D.M., Harris, R.K., Eds.; John Wiley: London, UK, 1996; pp. 1273–1281. [Google Scholar] [CrossRef]
  76. Winkler, P.F.; Kleppner, D.; Myint, T.; Walther, F.G. Magnetic Moment of the Proton in Bohr Magnetons. Phys. Rev. A 1972, 5, 83–114. [Google Scholar] [CrossRef]
  77. Phillips, W.D.; Cooke, W.E.; Kleppner, D. Magnetic Moment of the Proton in H2O in Bohr Magnetons. Phys. Rev. Lett. 1975, 35, 1619–1622. [Google Scholar] [CrossRef]
  78. Jameson, C.J.; Jameson, A.K. Absolute Shielding Scale for 29Si. Chem. Phys. Lett. 1988, 149, 300–305. [Google Scholar] [CrossRef]
  79. Jackowski, K.; Jaszuński, M.; Kamieński, B.; Wilczek, M. NMR Frequency and Magnetic Dipole Moment of 3He Nucleus. J. Magn. Reson. 2008, 193, 147–149. [Google Scholar] [CrossRef]
  80. Mohr, P.J.; Taylor, B.N.; Newell, C.B. CODATA Recommended Values of the Fundamental Physical Constants: 2006. Rev. Mod. Phys. 2008, 80, 633–730. [Google Scholar] [CrossRef]
  81. Mohr, P.J.; Taylor, B.N.; Newell, C.B. CODATA recommended values of the fundamental physical constants: 2010. Rev. Mod. Phys. 2012, 84, 1527–1605. [Google Scholar] [CrossRef]
  82. Jackowski, K.; Jaszuński, M.; Wilczek, M. Alternative Approach to the Standardization of NMR Spectra. Direct Measurement of Nuclear Magnetic Shielding in Molecules. J. Phys. Chem. A 2010, 114, 2471–2475. [Google Scholar] [CrossRef] [PubMed]
  83. Antušek, A.; Jackowski, K.; Jaszuński, M.; Makulski, W.; Wilczek, M. Nuclear Magnetic Dipole Moments from NMR Spectra. Chem. Phys. Lett. 2005, 411, 111–116. [Google Scholar] [CrossRef]
  84. Stone, N.J. Table of Nuclear Magnetic Dipole and Electric Quadrupole Moments. At. Data Nucl. Data Tables 2005, 90, 75–176. [Google Scholar] [CrossRef]
  85. Flowers, J.L.; Petley, B.W.; Richards, M.G. A Measurement of the Nuclear Magnetic Moment of the Helium-3 Atom in Terms of that of the Proton. Metrologia 1993, 30, 75–87. [Google Scholar] [CrossRef]
  86. Makulski, W.; Jackowski, K.; Antušek, A.; Jaszuński, M. Gas-phase NMR measurements, absolute shielding scales and magnetic dipole moments of 29Si and 73Ge nuclei. J. Phys. Chem. A 2006, 110, 11462–11466. [Google Scholar] [CrossRef] [PubMed]
  87. Jackowski, K.; Makulski, W.; Szyprowska, A.; Antušek, A.; Jaszuński, M.; Jusélius, J. NMR Shielding Constants in BF3 and Magnetic Dipole Moments of 11B and 10B Nuclei. J. Chem. Phys. 2009, 130, 044309. [Google Scholar] [CrossRef] [PubMed]
  88. Makulski, W.; Szyprowska, A.; Jackowski, K. Precise Determination of the 13C Nuclear Magnetic Moment from 13C, 3He and 1H NMR Measurements in the Gas Phase. Chem. Phys. Lett. 2011, 511, 224–228. [Google Scholar] [CrossRef]
  89. Lantto, P.; Jackowski, K.; Makulski, W.; Olejniczak, M.; Jaszuński, M. NMR Shielding Constants in PH3, Absolute Shielding Scale and the Nuclear Magnetic Moment of 31P. J. Phys. Chem. A 2011, 115, 10617–10623. [Google Scholar] [CrossRef]
  90. Jaszuński, M.; Repisky, M.; Demissie, T.B.; Komorovsky, S.; Malkin, E.; Ruud, K.; Garbacz, P.; Jackowski, K.; Makulski, W. Spin-Rotation and NMR Shielding Constants in HCl. J. Chem. Phys. 2013, 139, 234302. [Google Scholar] [CrossRef] [PubMed]
  91. Adrjan, B.; Makulski, W.; Jackowski, K.; Demissie, T.B.; Ruud, K.; Antušek, A.; Jaszuński, M. NMR Absolute Shielding Scale and Nuclear Magnetic Moment of 207Pb. Phys. Chem. Chem. Phys. 2016, 18, 16483–16490. [Google Scholar] [CrossRef]
  92. Jaszuński, M.; Olejniczak, M. NMR Shielding Constants in SeH2 and TeH2. Mol. Phys. 2013, 111, 1355–1363. [Google Scholar] [CrossRef]
  93. Makulski, W. 83Kr Nuclear Magnetic Moment in Terms of that of 3He. Magn. Reson. Chem. 2014, 52, 430–434. [Google Scholar] [CrossRef]
  94. Makulski, W. 129Xe and 131Xe Nuclear Magnetic Dipole Moments from Gas Phase NMR Spectra. Magn. Reson. Chem. 2015, 53, 273–279. [Google Scholar] [CrossRef]
  95. Antušek, A.; Kȩdziera, D.; Kaczmarek-Kȩdziera, A.; Jaszuński, M. Coupled Cluster Study of NMR Shielding of Alkali Metal Ions in Water Complexes and Magnetic Moments of Alkali Metal Nuclei. Chem. Phys. Lett. 2012, 532, 1–8. [Google Scholar] [CrossRef]
  96. Antušek, A.; Rodziewicz, P.; Kȩdziera, D.; Kaczmarek-Kȩdziera, A.; Jaszuński, M. Ab Initio Study of NMR Shielding of Alkali Earth Metal Ions in Water Complexes and Magnetic Moments of Alkali Earth Metal Nuclei. Chem. Phys. Lett. 2013, 588, 57–62. [Google Scholar] [CrossRef]
  97. Antušek, A.; Holka, F. Absolute Shielding Scales for Al, Ga, and In and revised nuclear magnetic dipole moments of 27Al, 69Ga, 71Ga, 113In, and 115In nuclei. J. Chem. Phys. 2015, 143, 074301. [Google Scholar] [CrossRef]
  98. Antušek, A.; Šulka, M. Ab Initio Calculations of NMR Shielding of Sc3+, Y3+ and La3+ Ions in the Water Solution and 45Sc, 89Y, 138La and 139La Nuclear Magnetic Dipole Moments. Chem. Phys. Lett. 2016, 660, 127–131. [Google Scholar] [CrossRef]
  99. Stone, N.J. Table of Recommended Nuclear Magnetic Dipole Moments. In IAEA Nuclear Data Section; Vienna International Centre: Vienna, Austria, 2019; Available online: http://www-nds.iaea.org/publications (accessed on 28 March 2024).
  100. Harding, R.D.; Pallada, S.; Croese, J.; Antušek, A.; Baranowski, M.; Bissell, M.L.; Cerato, L.; Dziubinska-Kühn, K.M.; Gins, W.; Gustafsson, F.P.; et al. Magnetic Moments of Short-Lived Nuclei with Part-per-Million Accuracy:Toward Novel Applications of β-Detected NMR in Physics, Chemistry, and Biology. Phys. Rev. X 2020, 10, 041061. [Google Scholar] [CrossRef]
  101. Garbacz, P.; Makulski, W. 183W Nuclear Dipole Moment Determined by Gas-Phase NMR Spectroscopy. Chem. Phys. 2017, 498–499, 7–11. [Google Scholar] [CrossRef]
  102. Antušek, A.; Repisky, M.; Jaszuński, M.; Jackowski, K.; Makulski, W.; Misiak, M. Nuclear Magnetic Dipole Moment of 209Bi from NMR experiments. Phys. Rev. A 2018, 98, 052509. [Google Scholar] [CrossRef]
  103. Makulski, W.; Garbacz, P. Gas-Phase 21Ne NMR Studies and the Nuclear Magnetic Dipole Moment of Neon-21. Magn. Res. Chem. 2020, 58, 648–652. [Google Scholar] [CrossRef] [PubMed]
  104. Makulski, W.; Aucar, J.J.; Aucar, G.A. Ammonia: The Molecule for Establishing 14N and 15N Absolute Shielding Scales and a Source of Information on Nuclear Magnetic Moments. J. Chem. Phys. 2022, 157, 084306. [Google Scholar] [CrossRef] [PubMed]
  105. Makulski, W. Probing Nuclear Dipole Moments and Magnetic Shielding Constants through 3-Helium NMR Spectroscopy. Physchem 2022, 2, 116–130. [Google Scholar] [CrossRef]
  106. Makulski, W.; Słowiński, M.A.; Garbacz, P. Nuclear Dipole Moments and Shielding Constants of Light Nuclei Measured in Magnetic Fields. Magnetochemistry 2023, 9, 148. [Google Scholar] [CrossRef]
  107. Garbacz, P.; Jackowski, K. Referencing of 1H and 13C NMR shielding measurements. Chem. Phys. Lett. 2019, 728, 148–152. [Google Scholar] [CrossRef]
  108. Garbacz, P.; Piszczatowski, K.; Jackowski, K.; Moszyński, R.; Jaszuński, M. Weak intermolecular interactions in gas-phase NMR. J. Chem. Phys. 2011, 135, 084310. [Google Scholar] [CrossRef] [PubMed]
  109. Altman, L.J.; Laungani, D.; Gunnarsson, G.; Wennerstrom, H.; Forsén, S. Proton, deuterium, and tritium nuclear magnetic resonance of intramolecular hydrogen bonds. Isotope effects and the shape of the potential energy function. J. Am. Chem. Soc. 1978, 100, 8264–8266. [Google Scholar] [CrossRef]
  110. Jameson, C.J.; Mason, J. The Chemical Shift. In Multinuclear NMR; Mason, J., Ed.; Plenum Press: New York, NY, USA, 1987; Chapter 3; p. 81. [Google Scholar] [CrossRef]
  111. Jackowski, K.; Makulski, W. 13C shielding scale for MAS NMR spectroscopy. Magn. Reson. Chem. 2011, 49, 600–602. [Google Scholar] [CrossRef]
  112. Yi, Y.; Adrjan, B.; Włodarz, J.; Li, J.; Jackowski, K.; Roszak, S. NMR Measurements and DFT Studies of Nuclear Magnetic Shielding in Emodin and Chuanxiongzine Molecules. J. Mol. Struct. 2018, 1166, 304–310. [Google Scholar] [CrossRef]
  113. Yi, Y.; Adrjan, B.; Li, J.; Hu, B.; Roszak, S. NMR Studies of Daidzein and Puerarin: Active Anti-oxidants in Traditional Chinese Medicine. J. Mol. Model. 2019, 25, 202. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Selected examples from multinuclear NMR spectroscopy qualitatively illustrate the origin and further modifications of nuclear magnetic shielding in diamagnetic molecules. Black arrows mean the increase in shielding, and red arrows represent the deshielding effects.
Figure 1. Selected examples from multinuclear NMR spectroscopy qualitatively illustrate the origin and further modifications of nuclear magnetic shielding in diamagnetic molecules. Black arrows mean the increase in shielding, and red arrows represent the deshielding effects.
Molecules 29 02617 g001
Figure 2. On the scale of nuclear magnetic shielding, the blue bars represent the range of chemical shifts for selected nuclei [43]. The values of chemical shifts can be positive (+) or negative (−) depending on the position of the reference standard (I). Special shielding effects are also highlighted by black triangles for three compounds containing iodide atoms: HI in 1H, CI4 in 13C, and SiI4 in 29Si NMR [35,44,45,46,47].
Figure 2. On the scale of nuclear magnetic shielding, the blue bars represent the range of chemical shifts for selected nuclei [43]. The values of chemical shifts can be positive (+) or negative (−) depending on the position of the reference standard (I). Special shielding effects are also highlighted by black triangles for three compounds containing iodide atoms: HI in 1H, CI4 in 13C, and SiI4 in 29Si NMR [35,44,45,46,47].
Molecules 29 02617 g002
Figure 3. Alternative reference standards of shielding and chemical shifts for 1H NMR: the red dot () represents the shielding of liquid water in a spherical sample at 34.7 °C σref(H2Oliq.sph) = 25.790(14) ppm [77], σ0(H2) = 26.293(5) ppm [35], σ0(H2O) = 30.102(8) ppm [29], and σ0(TMS) = 30.783(5) ppm [25]; σ(1% TMS in liq. CDCl3) = 33.480(5) ppm as shown in Table 1.
Figure 3. Alternative reference standards of shielding and chemical shifts for 1H NMR: the red dot () represents the shielding of liquid water in a spherical sample at 34.7 °C σref(H2Oliq.sph) = 25.790(14) ppm [77], σ0(H2) = 26.293(5) ppm [35], σ0(H2O) = 30.102(8) ppm [29], and σ0(TMS) = 30.783(5) ppm [25]; σ(1% TMS in liq. CDCl3) = 33.480(5) ppm as shown in Table 1.
Molecules 29 02617 g003
Figure 4. The absolute magnetic shielding known for an isolated helium-3 atom σref(3He) can be encoded into popular deuterated lock solvents [82,107]. Then, a standard NMR spectrometer is used as the double nuclear device, which permits the measurement of shielding for other magnetic nuclei. Such an experiment requires the exact reading of two resonance frequencies: νX for the observed nuclei and νD for the deuterium lock solvent. Important, the nuclear magnetic dipole moments must be known with satisfactory accuracy for the above shielding measurements.
Figure 4. The absolute magnetic shielding known for an isolated helium-3 atom σref(3He) can be encoded into popular deuterated lock solvents [82,107]. Then, a standard NMR spectrometer is used as the double nuclear device, which permits the measurement of shielding for other magnetic nuclei. Such an experiment requires the exact reading of two resonance frequencies: νX for the observed nuclei and νD for the deuterium lock solvent. Important, the nuclear magnetic dipole moments must be known with satisfactory accuracy for the above shielding measurements.
Molecules 29 02617 g004
Figure 5. 1H and 13C NMR spectra of liquid ethyl crotonate in CDCl3 on a 500 MHz Varian INOVA spectrometer. The 2H solvent signal of CDCl3 is used as the internal reference standard of nuclear magnetic shielding. The measurements contain all the components of shielding attributed to the particular nuclei in this sample.
Figure 5. 1H and 13C NMR spectra of liquid ethyl crotonate in CDCl3 on a 500 MHz Varian INOVA spectrometer. The 2H solvent signal of CDCl3 is used as the internal reference standard of nuclear magnetic shielding. The measurements contain all the components of shielding attributed to the particular nuclei in this sample.
Molecules 29 02617 g005
Table 1. Spectral NMR parameters and recommended reference standards for selected magnetic nuclei.
Table 1. Spectral NMR parameters and recommended reference standards for selected magnetic nuclei.
NMRNatural Abundance a [%]gX Factor b,cAbsolute Frequency a  ΞX [MHz]Reference Shielding
σ [ppm]
Recommended Reference
Standard a
1H99.98855.585 694 70100.000 00033.480(5) dTMS in CDCl3
2H0.01150.857 438 23115.350 60933.568(5) eTMS-d12 in CDCl3
3He1.37 × 10−4−4.255 250 6276.179 43759.967 f3He gas
13C1.071.404 73925.145 020186.42(10) gTMS in CDCl3
15N0.368−0.566 14110.136 767−135.8 hCH3NO2 liquid
17O0.038−0.757 418 813.556 457290.2 iD2O liquid
19F1005.256 6894.094 011188.7 jCCl3F liquid
29Si4.6832−1.110 10419.867 187379.0(20) gTMS in CDCl3
31P1002.261 8540.480742328.35 kH3PO4 85% in H2O
77Se7.631.06 119.071 5132069 l(CH3)2Se liquid
a Ref. [20]; b ref. [24]; c gX = µX/(IXµN), µN = 5.050 783 53 × 10−27 J T−1; d ref. [25]; e estimated based on 1H results and the actual 2H resonance frequency, ΞD (TMS-d12 in CDCl3); f ref. [26]; g ref. [25]; h ref. [27]; i ref. [28,29]; j ref. [30]; k ref. [31]; l ref. [32].
Table 2. Nuclear magnetic dipole moments are determined from the gas phase NMR measurements and quantum chemical calculations of shielding.
Table 2. Nuclear magnetic dipole moments are determined from the gas phase NMR measurements and quantum chemical calculations of shielding.
Applied NMR MethodsµX TransferObserved MoleculesMagnetic Moments, µX/µNMore Details
1H and 13C1H  13C13CH40.7023694(7)Refs. [83,88]
1H and 14N1H  14N14NH30.4035723(5)Ref. [83]
1H and 15N1H  15N15NH3−0.283057(1)Ref. [83]
1H and 17O1H  17OH217O−1.893547(2)Ref. [83]
1H and 19F1H  19FCH319F2.62834(1)Ref. [83]
1H and 29Si1H  29Si29SiH4−0.555052(3)Ref. [86]
1H and 31P1H  31P31PH31.130925(5)Refs. [83,89]
19F and 33S19F  33S33SF60.64325(2)Ref. [83]
1H and 35Cl1H  35ClH35Cl0.82170(1)Ref. [90]
1H and 37Cl1H  37ClH37Cl0.68398(1)Ref. [90]
1H and 73Ge1H  73Ge73GeH4−0.87824(5)Ref. [86]
1H and 77Se1H  77SeH277Se0.53356(5)Ref. [92]
1H and 207Pb1H  207Pb207Pb(CH3)40.5906(4)Ref. [91]
3He and 10B3He  10B10BF3 + 3He1.8004636(8)Ref. [87]
3He and 11B3He  11B11BF3 + 3He2.688378(1)Ref. [87]
3He and 83Kr3He  83Kr83Kr + 3He−0.970730(3)Ref. [93]
3He and 129Xe3He  129Xe129Xe + 3He−0.77796(2)Ref. [94]
3He and 131Xe3He  131Xe131Xe + 3He0.691845(7)Ref. [94]
Table 3. Deuterium magnetic shielding in popular liquid lock solvents calibrated relative to an isolated 3He atom [82] *.
Table 3. Deuterium magnetic shielding in popular liquid lock solvents calibrated relative to an isolated 3He atom [82] *.
NoLock SolventObserved 2H SignalσD [ppm]
1Cyclohexane-d12-CD2-31.834
2Toluene-d8-CD331.525
3Acetonitrile-d3-CD330.864
4DMSO-d6-CD330.574
5Acetone-d6-CD330.570
6Methanol-d4-CD329.593
7Water-d2-OD28.837
8Nitromethane-d3-CD328.041
9Benzene-d6=CD-26.441
10Chloroform-d-CD26.389
* For liquids observed at the external parallel magnetic field (Bǁ) in 5 mm o.d. spinning NMR sample tubes at 300 K. More the σD parameters are available in ref. [107].
Table 4. The primary and secondary isotope effects from the direct measurements of shielding in H2, HD, and D2 isolated molecules [108] *.
Table 4. The primary and secondary isotope effects from the direct measurements of shielding in H2, HD, and D2 isolated molecules [108] *.
Observed NMR ParametersPrimary Isotope EffectsSecondary Isotope Effects
Shielding in H2 and HD moleculesσ0(H2) = 26.293 ppm
σ0(HD) = 26.239 ppm
σ0(H2) = 26.293 ppm
σ0(HD) = 26.327 ppm
Isotope effects0Δ(2/1H) = −0.046 ppm1Δ(2/1H) = −0.034 ppm
Shielding in HD and D2 moleculesσ0(HD) = 26.327 ppm
σ0(D2) = 26.388 ppm
σ0(HD) = 26.339 ppm
σ0(D2) = 26.388 ppm
Isotope effects0Δ(2/1H) = −0.061 ppm1Δ(2/1H) = −0.049 ppm
* As defined in Ref. [110]: the 0Δ(2/1H) and 1Δ(2/1H) values mean the primary and secondary 2/1H isotope effects, respectively.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jackowski, K.; Wilczek, M. Measurements of Nuclear Magnetic Shielding in Molecules. Molecules 2024, 29, 2617. https://doi.org/10.3390/molecules29112617

AMA Style

Jackowski K, Wilczek M. Measurements of Nuclear Magnetic Shielding in Molecules. Molecules. 2024; 29(11):2617. https://doi.org/10.3390/molecules29112617

Chicago/Turabian Style

Jackowski, Karol, and Marcin Wilczek. 2024. "Measurements of Nuclear Magnetic Shielding in Molecules" Molecules 29, no. 11: 2617. https://doi.org/10.3390/molecules29112617

Article Metrics

Back to TopTop