Next Article in Journal
Intrinsic Fluorometric Reporters of Pteridine Reductase 1, a Target for Antiparasitic Agents
Previous Article in Journal
Assignment of the Vibrational Spectra of Diiron Nonacarbonyl, Fe2(CO)9
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Probing Nuclear Dipole Moments and Magnetic Shielding Constants through 3-Helium NMR Spectroscopy

by
Włodzimierz Makulski
Laboratory of NMR Spectroscopy, Faculty of Chemistry, University of Warsaw, Pasteura 1, 02-093 Warsaw, Poland
Physchem 2022, 2(2), 116-130; https://doi.org/10.3390/physchem2020009
Submission received: 5 February 2022 / Revised: 4 April 2022 / Accepted: 13 April 2022 / Published: 28 April 2022
(This article belongs to the Section Experimental and Computational Spectroscopy)

Abstract

:
Multinuclear NMR studies of the gaseous mixtures that involve volatile compounds and 3He atoms are featured in this review. The precise analyses of 3He and other nuclei resonance frequencies show linear dependencies on gas density. Extrapolation of the gas phase results to the zero-pressure limit gives the ν0(3He) and ν0(nX) resonance frequencies of nuclei in a single 3-helium atom and nuclei in molecules at a given temperature. The NMR frequency comparison method provides an approach for determining different nuclear magnetic moments. The application of quantum chemical shielding calculations, which include a more complete and careful theoretical treatment, allows the shielding of isolated molecules to be achieved with great accuracy and precision. They are used for the evaluation of nuclear moments, without shielding impacts on the bare nuclei, for: 10/11B, 13C, 14N, 17O, 19F, 21Ne, 29Si, 31P, 33S, 35/37Cl, 33S, 83Kr, 129/131Xe, and 183W. On the other hand, new results of nuclear moments were used for the reevaluation of absolute nuclear magnetic shielding in the molecules under study. Additionally, 3He gas in water solutions of lithium and sodium salts was used for measuring 6/7Li and 23Na magnetic moments and reevaluating the shielding parameters of Li+ and Na+ water-solvated cations. In this paper, guest 3He atoms that play a role in probing the electron density in many host macromolecules are also presented.

1. Introduction

Helium (He), a chemical element, is atomic number 2 located by IUPAC in group 18 (or 8A) of the Periodic Table of Elements [1]. It is a colourless, odourless, tasteless, inert, monoatomic gas. Helium is the second most abundant element in the universe. However, in the Earth’s atmosphere and in natural gas, it is only 5.2 ppm by volume. Seven times lighter than air, it is able to escape from the Earth’s atmosphere to space. Radiogenic helium in natural gas fields occurs in concentrations up to ~7% by volume and can be extracted commercially by a low-temperature fractionated distillation process. For the NMR community, the use of liquid helium as a cooling material for superconducting electromagnets in spectrometers is well known. Two stable helium nuclei are known: 3He and 4He.
Both of these helium nuclei play an important role in cryogenics [2]. 4He can be a liquid at a low temperature (4.22 K, or −268.93 °C) and can be a solid at a pressure of 25 atm and temperature of 0.95 K (−272.2 °C). Liquid 3He can be used as a refrigerant in the separated phase solution of 3He in 4He, where a temperature below 0.1 K is reached. In these very low temperatures, it shows characteristic superfluid properties (Helium II) [3]. The natural abundance of 3He is of the order of 0.000134%, but only this isotope with spin number I = 1/2 can be used in NMR and MRI measurements. The use of artificially enriched material is then obligatory. Most of the rare isotope of helium is produced from the decay of tritium during the maintenance of nuclear weapons in the United States and then accumulated in a stockpile. Tritium (T, 3H) is the radioactive isotope of hydrogen with a half-life of 12.32 years. The 3He nucleus—also known as helion (3He+2, symbol h)—is composed of two protons and one neutron that is magnetically active. Two simple requirements are obligatory for successful measurements of helium in the NMR spectrometer. It is necessary to have a suitable transmitter with an available helium frequency and a probe with adequate generated resonance tuned to the helium frequency [4]. In fact, many NMR spectrometers do not meet these conditions. Nevertheless, some practical measurements were performed on helium in its gaseous, liquid and solid states. If samples are enriched in the helium-3 isotope, the sensitivity of the NMR analysis is high enough. The resonance signals are sharp when appearing at a spectral range extending from 10 up to −50 ppm. Since helium does not form any “normal” covalent compounds, the NMR experiments have a physical, rather than chemical, character. In recent years, an experiment with helium-3 revealed a very valuable method of investigation for its specific intermolecular-interactions while in the gas phase [5]. Helium atoms are completely chemically inert and do not form any stable compounds. Interestingly, helium-3 atoms can be included into fullerenes and nanotubes and used as probes of magnetic shielding molecular interiors. This aspect of NMR investigations will be discussed at the end of our review. Novel developments of 3He NMR are connected with nuclear hyperpolarisation techniques, carried out in different ways, to receive signals even 10,000 times greater per atom than those routinely produced at thermal equilibrium [6]. A new, general method of NMR spectra measurements expressed in shielding scales, known as the “Alternative Approach to the Standardization of NMR Spectra”, was proposed using the 3He NMR resonance reference [7]. For this purpose, the 3He NMR spectra were measured simultaneously with the residual proton and carbon signals of different deuterated solvents when helium-3 was dissolved. The experimental shielding of deuterated solvents determined in this way is characterised with high accuracy and precision, and will be used in the next steps of this work. The main part of this review is devoted to the NMR method of NMM measurements. The role of 3He NMR spectroscopy in establishing nuclear magnetic moments and magnetic shielding constants is discussed in detail. We utilised the 3He (helium) atom used in mixtures with simple molecules that possessed the nuclei of interest, e.g., 31PH3, 13CH4, 14NH3, 11BF3, 29SiH4, H37Cl, and some others. Additionally, a few comments on helium-3 NMR achievements and prospects are included. The Scheme 1 of nuclei studied by NMR spectroscopy is shown below:

2. Methodology of Measurements of Nuclear Magnetic Moments

2.1. Methodology of Gas-Phase NMR Experiments

Many physical properties of gases can be conveniently characterised in terms of general virial theorem. In this case, the nuclear magnetic shielding dependence of any nucleus in a pure gas molecule σ(ρ,T) can be described as a virial expansion in the density function (ρ) at a given temperature (T) [8], as follows:
σ(ρ,T) = σ0 + σ1(T)ρ + σ2(T)ρ2 + σ3(T)ρ3 +…
where σ0 is the shielding constant of the nucleus in the “isolated” atom or molecule and σ1, σ2, and σ3 are virial coefficients resulting from many body collisions. Below ~40 atm (density 1.64 mol/L at 297 K), the dependences are strictly linear because only the first two coefficients are present, as follows:
σ(ρ,T) = σ0 + σ1(T)ρ
Analogous equations can be treated when the mixtures of gaseous substances are investigated. These systems can be considered as the binary mixtures of gaseous substance A containing the nucleus X, whose shielding σ(X) is measured, and gas B is the solvent. Equation (1) can be formulated as:
σ(X) = σ0(X) + σAA(X)ρA + σAB(X)ρB + …
where ρA and ρB are the densities of substances A and B. The coefficients σAA(X) and σAB(X) contain the bulk susceptibility corrections ((σA)b and (σB)b) and the terms involving intermolecular interactions during the A–A and A–B collisions (σA-A(X) and σA-B(X)). Again, if the concentration of substance A under investigation is sufficiently low (usually <1%) the σAA(X) can be fully omitted, and the final equation is:
σ(X) = σ0(X) + σAB(X)ρB
Nuclear magnetic shielding constants can be measured also in the liquid solutions of different substances. Here, water solutions of inorganic salts are of the most importance. Several water salt solutions were investigated to establish the nuclear magnetic moments of inorganic isotopes. The analysis of chemical shifts measured in some mixtures of different concentrations is much more complicated than that of gases because of nonlinearity dependences. In this case, polynomial functions are more useful:
σ(X) = σ0(X) + σAB(X)ρB + σAB(X)ρB2
where σ0(X) means the isotope in the solvated cation or anion under study. The final equation forms (2), (4), and (5) will be exercised for numerous individual cases in detail in the next sections.

2.2. Nuclear Magnetic Dipole Moments and Shielding Constants

The main field of the induction resonance technique was to determine the dipole moment of a given nuclei. For atomic nuclei, the intrinsic angular momentum of particle (p) is directly proportional to its magnetic moment (μ) as μ = γ·p, where the proportionality constant γ is the magnetogyric ratio. The dipole magnetic moment is a vector, but only the projection value that is partially in the direction of the magnetic field can be measured in experimental conditions. The nuclear magnetic moment (NMM) μX is related to the nuclear spin number as:
μX = ħ·γX·IX = gX·IX·μN
where γX is the nuclear gyromagnetic ratio, μN (which is equal to 5.050783699(31) × 10−24J/T) is the nuclear magneton, and gX is the g factor of nucleus x. The first measurements of NMM were performed by I.I. Rabi [9], who predicted that magnetic moments could flip their magnetic orientation in field B0 when forced to absorb tiny parts of energy from the additional electromagnetic field B1. These transitions can be detected when spins go to the lower energy orientations. He called this method atomic or molecular beam magnetic resonance (ABMR). This approach was a prototype of the well-known spectroscopic method NMR (nuclear magnetic resonance) carried out in a static magnetic field on bulk macroscopic samples. The NMM precess (rotate) is in the magnetic field around the z axis at a constant rate ωL, which can be observed in the NMR spectrometer as a radio resonance frequency (rf). The equations that determine the resonance frequencies of two different nuclei, νX and νY, in the resonance condition are as follows:
X = (1 – σX)Bz
Y = (1 – σY)Bz
where σX and σY are the absolute shielding constants. In reality, the shielding (screening) effects in the matter can slightly change the Larmor frequency of any given nuclei in the molecule, depending on the chemical environment of the nucleus within a molecule. It is difficult to measure the strength of the magnetic field with high precision. This problem can be avoided if the ratio of frequencies in Equations (7) and (8) is measured. Dividing both equations, it is possible to eliminate the external magnetic field induction B0 and receive the very useful relation:
Δ μ X z = ν X ν Y ( 1 σ Y ) ( 1 σ X ) Δ μ Y z  
often used for evaluation of unknown NMM from other taken as references. The nuclear dipole moment of the reference nucleus should be known with the best possible precision and accuracy. Proper measurements of nuclear magnetic moments require corrections for shielding constants (σ) of the given and reference nucleus f = (1 − σY)/(1 − σX). These correction factors (f) are usually more or less different than 1.00, which means that shielding effects in atoms and molecules change the proper nuclear magnetic moments. New, and now more accessible, sophisticated theoretical calculations of shielding parameters have provided a new impetus for remeasuring moments with great accuracy and precision.
The most excellent result for the nuclear dipole moment was established very recently for the proton µ(1H) = µp. The new physical method for the direct, high-precision measurement of a single isolated proton in a double Penning trap gives the value μpN = g/2 = 2.79284734462(82)μN [10]. The magnetic moment is here expressed in units of nuclear magnetons. The most fruitful method for establishing heavier magnetic dipole moments of stable nuclei is NMR spectroscopy since resonance techniques are able to give fixed values very precisely and accurately. Locating a macroscopic sample in a uniform stable magnetic field, it is possible to measure Larmor frequencies coming from different nuclei present in a sample. These nuclei can originate from the same molecule or different molecules that are simply present in the sample under study. This approach allows the assignment of appropriate frequency relations and, as a consequence, the comparison of two or more NMMs. The basis for further calculations is a good knowledge of one NMM from the pair of chosen nuclei. One can usually use 1H (proton) as a reference. Other referenced nuclei can be deuteron D(2H), helion 3He, or fluorine 19F. In this paper, the second example is mostly used. The other form of Equation (9) can be used to check the consistency of nuclear magnetic moments and shielding factors, as follows:
σ X = 1 ν X ν Y Δ μ Y z Δ μ X z ( 1 σ Y )  
All parameters used in this formula are the same as stated above. It is very convenient to exercise Equation (10) when complex molecules are studied and data for additional nuclei can be applied.

3. Results and Discussion

3.1. 3He Chemical Properties

The helium atom and helion nucleus can be categorised as unusual physical objects. A very small atomic diameter (28 pm, 1 pm = 10−12 m) can cause the glass walls to be penetrated and the glass ampoule to be emptied. This is an indication for fast experimental maintenance of samples containing helium-3 atoms. It is remarkable that the 3He nucleus is the only stable nucleus with a proton number larger than neutrons. The mirror nucleus of helium-3 is the tritium nucleus with two neutrons and one proton (T = 3H). The sign of the 3He nuclear magnetic moment is opposite to that of a proton [11] and negative by convention. Precise knowledge of the helium-3 NMM as a reference is crucial for employing the NMR method for establishing nuclear dipole moments against that of helion. Several attempts have been made to precisely determine the 3He nuclear magnetic moment, and all were undertaken using NMR spectroscopy. The 3He magnetic moment was measured by Anderson and Novick [12], Williams and Hughes (1968) [13], Neronov and Barzakh (1977) [14], Belyi and Shifrin (1986) [15], Hoffman and Becker (2005) [16], Jackowski et al. (2008) [4], Aruev and Neronov (2012) [17], and Makulski (2020) [18]. In our opinion, the best result was achieved by Flowers, Petley, and Richards [19] in pure helium-3 samples against the proton signal in liquid water. The authors measured the ratio of the NMR spin precession frequencies of optically pumped, low pressure helium-3 and protons in water, taking into account the residual inhomogeneities and water diamagnetic corrections accurately. The final results of the nuclear magnetic moment (μ(3He) = −2.127625308(25)μN) were corrected by factors that arose from the shielding parameters of helium-3 (σ0(3He) = 59.96743(10) ppm) [20]. These were included in the latest collections of fundamental physics constants recommended by Stone in the “Table of recommended nuclear magnetic dipole moments” published under the auspices of the INDC (International Nuclear Data Committee) in November 2019 [21]. This value was used for all recalculations of nuclear moments of different nuclei carried out previously and mentioned in this work. It is worth noting that the nuclear moments are rather more experimental than theoretical values. The nuclear moments can be calculated purely theoretically using several nuclear models. The last calculations lead to essentially different results: −1.913042 µN (shell model), −2.2905 µN (QCD model), and −2.004914 µN (Quark model), where QCD refers to the lattice quantum chromodynamics theory [22]. The difference from the best experimental value is rather large (~5% in the last case) and confirms the assumption of nuclear moments’ significance as experimental values. From the quantum mechanical point of view, the 3He nucleus is fermion, while the 4He nucleus is a boson. Several chemical properties of the helium atom and helion nucleus are shown in Table 1. The NMM µ(3He) and spin number (1/2) decide the radiofrequency used for the detection of this nuclei in the stable magnetic field. The Ξ frequency parameter is defined as the ratio of the 3-helium frequency, ν0 (observed), to that for 1H of TMS in CDCl3 in the same magnetic field, expressed as a percentage. For example, in magnetic field B0 = 11.74 T, the 3He radiofrequency is 381.3575 MHz, just between frequencies for 19F (470.910 MHz) and 31P (202.595 MHz). In this way, the Ξ/MHz is 76.178972 [23]. Fortunately, these measurements are available on Varian INOVA (now Agilent, Santa Clara, CA, USA) spectrometers. The relaxation times of the 3He nuclei in the pure gas phase are very long, up to several thousand of seconds. Therefore, the usage of short pulse duration and long relaxation delays are strongly recommended. The linewidth of recorded signals often does not exceed 1 Hz.

3.2. Other Noble Gases: 21Ne, 83Kr, and 129/131Xe

A simple and effective example for handling Equations (2) and (4) for establishing the nuclear magnetic moments is provided by mixtures of noble gas isotopes that possess the non-zero spin numbers. Mixtures of 3He as reference nucleus with other noble gas atoms, including 21Ne [24], 83Kr [25], 129Xr and 131Xe, were carried out using the NMR method [26]. The frequency dependences as functions on the density of the main gaseous ingredients are shown in Figure 1.
The appropriate input data for nuclear magnetic moment calculations—frequency ratios, correction factors, and final nuclear magnetic moments data—are shown in Table 2. Thanks to the high receptivity of 3He NMR signals, the precise evaluation of magnetic moments is mainly a function of shielding correction factors. The correction factor for 21Ne is small (~0.05%), but very valuable for xenon isotopes of 0.7%. The results marked with asterisks were calculated against the deuterium 2H(D) resonance used in the lock system. It can be seen that both kinds of results for helium, krypton, and xenon are in very good agreement. This circumstance justifies using the lock system instead of 3-helium for measuring the 21Ne nuclear moment.
In our spectrometer, low frequency resonances (like 21Ne and 83Kr) should be measured in a broad band probe with test tubes that have a 10 mm external diameter, whereas high frequencies (like 3He) are only available in a special probe with NMR tubes that have a 5 mm external diameter [4]. Both the original lock system and VT facilities were left unchanged in the home-adapted Varian variable temperature probe-head. This probe is notable for its high quality. The duration of the 90° rectangular rf pulse was only 15.5 µs. The lock frequency (2H) was practically constant (76.84640 MHz) for all measurements presented in this work. Our probe generally allows the precise determination of 3He spectra samples with small amounts of helium gas in a short time.
From the final results of NMMs measured against 3He and 2H(D) nuclei (Table 2), it is clear that both are very congruent and can be used interchangeably in these experiments. The accuracy is strictly connected with shielding factors which are large for heavy nuclei and small for light nuclei. The shielding parameters of the lock system (2H,D), for these purposes, show very good quality [4]. It is remarkable that the NMR investigations of other noble gases, mainly argon and radon elements, are not possible. All three stable argon isotopes—36Ar, 38Ar and 40Ar—are not magnetically active because of their spin (I = 0). Otherwise, there are known nuclear magnetic moments of short-lived nuclei such as 35Ar, 37Ar, 39Ar, and 41Ar from physical methods other than typical NMR experiments. Likewise, radon nuclei are all radioactive and short-lived. The electromagnetic properties of radon isotopes, from 209Rn up to 225Rn, are known from physical experiments [18].

3.3. Simple Hydrides: 13CH4, 14NH3, H217O, 29SiH4, 31PH3, and H35/37Cl

The investigations of simple, gaseous hydrides are optimal for obtaining the accurate results of shielding and magnetic moments of nuclei. The range of 1H NMR shifts is very small, practically limited to ~10 ppm, and experimental errors are much less than 0.1 ppm. The shielding constants are very well known from experimental [28] and theoretical [29] studies. The binary hydrides CH4, NH3, SiH4, PH3, and HCl are gases at normal conditions and can be used directly from lecture bottles for the preparation of gaseous samples for NMR measurements. Furthermore, isotopically enriched substances (e.g., 13CH4 and 15NH3) were offered from commercial sources and formerly used in our systematic investigations [30,31,32,33,34]. More problematic were the attempts to prepare water that contained gaseous samples—a 25% enriched preparation was used [35], and next, 90% H217O [36]. In this last case, the samples containing ~5 × 10−4 mol/L enriched water in the buffer gases CH3F and CHF3 were made, and the 17O spectra were measured with success. All suitable input data for the calculation of the nuclear magnetic moments are shown below in Table 3.
Certainly, the measurements of 35Cl and 37Cl NMR spectra were the most challenging of all the gaseous systems presented in Table 3. The relatively large quadrupole moment Q of 35Cl (+0.085 b) (1 barn = 10−28m2) and 37Cl (−0.064 b) leads to effective quadrupolar relaxation times and broad resonance signals with a linewidth of one-hundred hertz (Hz). It is obvious that analogous experiments with other gaseous hydrides—HBr and HJ—dealing with 79Br (Q = +0.51 b), 81Br (Q = 0.26 b), and 129J (Q = −0.696 b) nuclei would be very problematic. Fortunately, water solutions of bromides and iodides with 3He solute can be used in these cases (see, for example, Section 3.5).

3.4. Fluoride Compounds: 10/11BF3, 33SF6, and 183WF6

Unfortunately, not all elements form simple hydrides which are regular gases or volatile liquids at room temperatures. In such cases, one can use the fluorinated analogs of hydrides, which often are available as commercial chemicals. The benefit of using 19F nuclei is their high sensitivity, which is almost equal to that of protons and the universal ability of measurements in most spectrometers. In the case of boron, it is suitable for analysing trifluoroborane BF3 gas rather than diborane B2H6, with a specific structure composed of two bridge H atoms between boron atoms and four H terminal atoms. In the end, we used BF3, which has a trigonal planar structure with D3h symmetry [37]. Its vapour pressure at 20 °C is more than 50 atm and it can be investigated in a large density range. Analogically, similar experiments were performed with sulfur hexafluoride SF6 gas in the 5–25 atm pressure range [38]. A big advantage of using this compound is its high octahedral symmetry consisting of six fluorine atoms connected to one central sulfur atom. The electric field gradient (EFG) on sulfur nuclei is then zero and the 33S signals are very narrow (~1 Hz). Other gaseous sulfur-containing compounds, such as H2S, COS, SO3, and SO2, are notable for their much worse NMR measurement conditions. The gaseous samples containing small amounts of 3-helium atoms dispersed in sulfur hexafluoride were used. Well-resolved sulfur-33 NMR spectra gave the opportunity to recognise the 1J(33S,19F) spin–spin coupling and make use of the INEPT (insensitive enhanced polarisation transfer) sequence to amplify the signal and smooth the baseline in the spectrum. For example, the 33S NMR INEPT spectrum of 21.8 atm (0.89 mol/L) SF6 is shown in Figure 2 as coupled to six fluorine-19 nuclei, as well as fully decoupled.
The linear density dependences of the 3He and 33S radiofrequencies are shown in Figure 3. The observed Larmor frequencies, extrapolated to the zero-pressure limit, belong to the isolated SF6 molecule and helium atom. The application of Equation (9) gives the 33S nuclear magnetic moment in terms of 3He μ(33S) = 0.6432555(10) μN. The new result is in good agreement with previously reported values (μ(33S) = 0.6432474(107) μN [39] and μ(33S) = 0.643251(16) μN [40]), but is more accurate by one order of magnitude. The shielding corrections for the sulfur nuclei (σ(33S) = 392.6 ppm) in SF6 were taken from the first relativistic calculations performed previously [41]. This value, recalculated from Equation (10) using the fluorine parameters—frequency and shielding constants—of 406.2 ppm is in moderate agreement with the calculated one.
Another important modification of the procedure shown above relies on the measurements of the 183WF6 substance where CF4 or C2F6 gaseous buffers were used [42]. They are necessary for lengthening the relaxation times and narrowing the 183W resonance signals. Naturally occurring tungsten (wolfram) contains only one isotope (183W) with a non-zero value of I = 1/2 and a natural abundance of 14.31%, with a rather poor receptivity of 1.06 × 10−5 relative to 1H. This is where it gets weird: 183W is a α-particle-emitter with a long half-life of >1 × 1017 y. The 183W NMR spectra were recorded using the INEPT (insensitive nuclei enhanced by polarisation transfer) sequence for increasing the signal-to-noise ratio and smoothing the baseline distortions. They were decoupled from the 19F nuclei during acquisition, making use of the known spin–spin coupling (1J(F,W) = 43.75 Hz). The isotope effect observed in the 19F NMR spectra connected with the substitution of 183W by other tungsten isotopes is negligible. The 183W nuclear magnetic moment obtained in this experiment was μ(183W) = 0.116953(18) μN, which gives the g-factor gI = μ(183W)/I = 0.233901(18) and gyromagnetic ratio γI = gI μN/ħ = 1.120249 × 107 rad s−1 T−1. The experimental results were corrected by the shielding constant of tungsten in WF6 calculated theoretically (σ(183W) = 6221.0 ppm [41]). The appropriate NMR parameters for the 3He/fluoride compound systems are collected in Table 4.
It is obvious that the 19F nuclear magnetic moment can be calculated from the helium-3 results used in the above-mentioned cases (the BF3/3He, SF6/3He, and WF6/3He systems). If we utilise the measured 19F frequencies and shielding results that come from the best relativistic calculations, the new 19F NMM values are as follows: μ(19F) = 2.6283348(26)μN, μ(19F) = 2.6283711(132) μN, and μ(19F) = 2.6283925(263) μN, respectively. They are in adequate agreement with those used in previously published results (μ(19F) = 2.628321(4) μN [21] and μ(19F) = 2.628335(11) μN [40]) but in disagreement with the result given previously in the literature tables (μ(19F) = 2.6288868(8) μN [40]). The deviations mainly arise from the shielding uncertainties of the fluorine nuclei. It is important to mention the special position of hydrogen fluoride (HF) in fluorine NMR spectroscopy. This simple molecule has one undesirable feature: it forms very corrosive and penetrating fumes with traces of water when hydrofluoric acid is forming. It attacks the glass vessels and vacuum lines mostly used in laboratories for sample preparation and handling. The intermolecular hydrogen bonds in HF are very strong and obtaining monomers, even in the gaseous phase, is very problematic. Nevertheless, the HF molecule was the basic molecule for establishing the 19F NMR absolute shielding scale [43]. The following special conditions were ensured: a thin-walled cell from polyethylenes and a Monel metal vacuum line for the transfer of HF and other gases. The results of this investigation have never been repeated and confirmed. For this reason, the secondary reference substances were usually used (SiF4 and CFCl3 [44]).
In light of these considerations and taking into account the remarks on the fluorine shielding scale, it is strongly recommended to reevaluate this absolute scale in new experimental and theoretical studies. We have recently started the preparatory work for the revision of 19F nuclear magnetic moments and shielding constants in gaseous mixtures of 3He/CH3F and 3He/CF4. The preliminary findings promise good agreement between older results and those of our new investigations.

3.5. Water Solutions: 6/7Li and 23Na Salts

Helium-3 can certainly be used as a reference in NMR measurements of nuclei when the water solutions of different cations are the target of the studies. Neither the lithium nor the sodium elements form any stable gaseous compounds. Li-He, the very weakly interacting Van der Waals chemical compound, is probably only present in low temperatures [45]. The most studied objects in lithium and sodium chemistry are their water solutions, and they were used in the NMM measurements. A total of 1.5 mg of the helium atoms can be dissolved in pure water at 20 °C and 1 bar pressure. Additionally, the shielding correction factors can be known from the theoretical studies of Antušek et al. [46]. These results make use of our new experimental reports on the NMR frequencies of Li+ [47] and Na+ [48] in water solutions of different salts. The concentration dependencies do not show linear functions, but they can be fitted with a quadratic one. In these cases, the frequency results do not belong to a simple isolated cation, but rather to a few aquatic structures of central ions occurring in different mole fractions. In the case of lithium cation, because of its small radius of 90 pm, as a Li(H2O)4+ cation, the shielding correction is 91.69 ppm in the infinitely diluted solution. In the octahedral coordinating lithium cation, this value is 90.89 ppm. Both results can give the 6/7Li nuclear magnetic moments as the lower and upper limits of these numbers. In the case of the 23Na magnetic moment, the situation is much more complicated. The sodium cation Na+ is dynamically transformed between several structures, according to the equation:
Na(H2O)3+ + H2O ↔ Na(H2O)4+ + H2O ↔ Na(H2O)5+ + H2O ↔ Na(H2O)6+ + H2O ↔ Na(H2O)7+
The molecular dynamics (MD) simulation results suggest that the five-coordinated sodium ions is the most-populated structure (~60%) and the six-coordinated structure is the second-most populated (~30%) [49]. In this case, we prefer to consider the dynamic character of water surrounding the central cation by averaging the results for four-, five-, and six-coordination systems, with their appropriate molar fractions. The shielding correction factor of 580.12 ± 10 ppm for sodium cations in liquid water was finally received.
The helium-3 correction factor in lithium water solutions was measured as 2.7675(25) ppm and 2.747(2) for sodium salts against that of the isolated 3He atom in the gas phase in the chemical shift category. Taking into account the volume magnetic shielding susceptibility effect of pure water (4/3χV, where χV is volume susceptibility), the shielding corrections are of the order of 0.24 ppm compared with the gas phase in the relative shielding scale. For 6/7Li and 23Na nuclear magnetic moments, the NMR results are very consistent with the ABMR results achieved previously [50], where shielding corrections were made for the atomic inert species σ(Li) = 147.2 ppm and σ(Na) = 640.62 ppm. The NMMs of lithium and sodium nuclei measured using the NMR and ABMR methods are shown in Table 5. Additionally, the new results published by Neronov [51,52] are included. They were achieved using an interesting method of the simultaneous observation of the two-spin system with an adapted NMR spectrometer, which reduced the random and systematic errors by an order of magnitude.

3.6. 3He Atoms in Different Chemical Environments

It is worth drawing attention to the new spectacular applications of 3-helium spectroscopic achievements. The 3He atom can be inserted into the internal cavities present in fullerene compounds for the investigation of the magnetic shielding environments inside cavities. The first developments dealt with the endohedral fullerenes 3He@C60 and 3He@C70 [53]. 3He shielding values are 6 and 29 ppm here, respectively, relative to the free atom, which indicates substantial diamagnetic ring currents inside fullerene molecules. It is even possible to incorporate two helium-3 atoms in C70 [54]. Interestingly, the reduction of C60 and C70 fullerenes to their hexaanions 3He2@C60−6 and 3He2@C70−6 reduces the magnetic field inside the fullerene, diminishing its aromacity [55]. These observations are in good agreement with the theoretical predictions of [56,57]. 3He NMR spectroscopy was used for probing the small pores in solid substances, e.g., zeolites [58,59], silica aerogels [60], and nanocavities on the crystal surface powders and single crystals [61]. Besides these chemical uses, it is impossible to omit the huge and still growing biological and medicine applications of 3-helium MRI methods. A few remarks about these studies will be added in the conclusion section of this paper.

4. Conclusions

In this review, we have shown the main achievements of the NMR method for establishing the nuclear magnetic moments of different nuclei, which were performed in the gas phase using 3He atoms. It was established that this method is distinguished by its high degree of accuracy and precision. NMR measurements in high magnetic fields, in the spectrometers of advanced technology, and in sophisticated quantum calculation methods leads to unprecedented results. The final nuclear magnetic moment was identified with at least 0.01% accuracy. The interpolation NMR parameters to the zero-point pressure of gases deals with well-defined chemical species, atoms, and molecules. The shielding corrections for these species are often known with great quality from sophisticated theoretical calculations. The experimental conditions in liquid solutions are quite different. The most common water solutions are often complicated systems where cations and anions are solvated by different numbers and time-changing water molecules. The calculation of the shielding values of these different solute compositions is a complicated task and only limited results are available. The references suggested for NMR measurements of nuclear magnetic moments are: 1H (proton), 2H (deuteron), and 3He (helion). Experiments performed in the gaseous phase are strongly recommended. The helion (3He) nuclear magnetic moment belongs to the more accurately known values among all nuclei in the periodic table, with 1.2 × 10−8 relative standard uncertainty. Despite this, we hope that newly planned experiments by physicists, using Penning trap methods in combination with laser cooling, will provide even better results [62]. Helium-3 has unique properties that currently makes it very useful in several fields, and its cryogenic properties are useful in low-temperature physics, magnetic resonance imaging (MRI) with hyperpolarisd atomic nuclei in medicine, and in fast neutron high sensitivity for 3-helium detectors. The gaseous helium-3 magnetic resonance method is even more useful in cryogenic environments [63] and as a probe of the nature of the solid-superfluid interface [64]. 3He can be hyperpolarized by spin-exchange optical pumping and other methods. On the other hand, a helium atom with an 3He nucleus can be used as a probe of electron densities inside macromolecule compounds only if NMR spectra can be measured for an encapsulated helium-3 species. Finally, local magnetic fields in liquid crystals utilising 3He NMR spectroscopy are also known from research [65]. More valuable and prospective are MRI investigations with hyperpolarized helium-3 atoms in medical treatment, firstly demonstrated in the lung images of a guinea pig [66].
This unique chemical nucleus and/or atom is definitely a good probe of the matter being investigated. It is truly fascinating to observe the discovery of the new and surprising purposes of helium-3 in environmental research. We are convinced that this small and curious species will find several spectacular applications in chemistry, physics, and biology in the near future.
Upon completion of this paper, we recognize a very recent publication by Pachucki et al. (2021) [67] on QED effects in calculations of the 3He shielding constant. The final result, which gained in accuracy, is: σ(3He) = 59.967029(23) µN, which is in good agreement with the previous result mentioned in this work. The new value does not essentially influence the helium-3 nuclear magnetic moment.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Grochala, W. On the position of helium and neon in the Periodic Table of Elements. Found. Chem. 2018, 20, 191–207. [Google Scholar] [CrossRef] [Green Version]
  2. Van Sciver, S.W. Helium Cryogenics, 2nd ed.; Springer: New York, NY, USA, 2012; ISBN 978-1-4419-9978-8. [Google Scholar] [CrossRef] [Green Version]
  3. Lee, D.M. The extraordinary phases of liquid 3He. Rev. Mod. Phys. 1997, 69, 645–665. [Google Scholar] [CrossRef] [Green Version]
  4. 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] [PubMed]
  5. Garbacz, P.; Piszczatowski, K.; Jackowski, K.; Moszyński, R.; Jaszuński, M. Weak intermolecular interactions in gas-phase nuclear magnetic resonance. J. Chem. Phys. 2011, 135, 084310. [Google Scholar] [CrossRef]
  6. Parnell, S.R.; Woolley, E.; Boag, S.; Frost, C.D. Digital pulsed NMR spectrometer for nuclear spin-polarized 3He and other hyperpolarized gases. Meas. Sci. Technol. 2008, 19, 045601. [Google Scholar] [CrossRef]
  7. 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]
  8. Jameson, C.J. Chapter 1: Fundamental Intramolecular and Intermolecular Information from NMR in the Gas Phase. In Gas Phase NMR; Jackowski, K., Jaszuński, M., Eds.; RSC: London, UK, 2016; pp. 1–51. [Google Scholar]
  9. Rabi, I.I.; Zacharias, J.R.; Millman, S.; Kusch, P. A New Method of Measuring Nuclear Magnetic Moment. Phys. Rev. 1938, 53, 318. [Google Scholar] [CrossRef]
  10. Schneider, G.; Mooser, A.; Bohman, M.; Schön, N.; Harrington, J.; Higuchi, T.; Nagahama, H.; Sellner, S.; Smorra, C.; Blaum, K.; et al. Double-trap measurement of the proton magnetic moment at 0.3 parts per billion precision. Science 2017, 358, 1081–1084. [Google Scholar] [CrossRef] [Green Version]
  11. Klein, M.P.; Holder, B.E. Determination of the Sign of the He3 Nuclear Magnetic Moment. Phys. Rev. 1957, 106, 837–838. [Google Scholar] [CrossRef]
  12. Anderson, H.L.; Novick, A. Magnetic moment of He3. Phys. Rev. 1948, 73, 919. [Google Scholar] [CrossRef]
  13. Williams, W.L.; Hughes, V.W. Magnetic Moment and hfs Anomaly for He3. Phys. Rev. 1969, 185, 1251–1255. [Google Scholar] [CrossRef]
  14. Neronov, Y.I.; Barzakh, A.E. Determination of the magnetic moment of the He3 nucleus with an error of 2 × 10−6%. Zh. Eksp. Teor. Fiz. 1978, 75, 1521–1540. [Google Scholar]
  15. Belyi, V.A.; Il’ina, E.A.; Shifrin, V.Y. Experimental determination of the magnetic moments ratio for helium-3 and proton nuclei. Meas. Tech. 1986, 29, 613–616. [Google Scholar] [CrossRef]
  16. Hoffman, R.E.; Becker, E.D. Temperature dependence of the 1H chemical shift of tetramethylsilane in chloroform, methanol, and dimethylsulfoxide. J. Magn. Reson. 2005, 176, 87–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Aruev, N.N.; Neronov, Y.I. Gas Samples with a Mixture of Hydrogen Isotopes and 3He for NMR Spectroscopy and Estimation of the Magnetic Moment of the 3He Nucleus. Zh. Tekh. Fiz. 2012, 82, 116–121. [Google Scholar] [CrossRef]
  18. Makulski, W. Explorations of Magnetic Properties of Noble Gases: The Past, Present, and Future. Magnetochemistry 2020, 6, 65. [Google Scholar] [CrossRef]
  19. 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]
  20. Rudziński, A.; Puchalski, M.; Pachucki, K. Relativistic, QED, and mass effects in the magnetic shielding of 3He. J. Chem. Phys. 2009, 130, 244102. [Google Scholar] [CrossRef] [Green Version]
  21. Stone, N.J. Table of Recommended Nuclear Magnetic Dipole Moments; INDC 2019; IAEA Nuclear Data Section, Vienna International Centre: Vienna, Austria, 2019; Available online: http://www-nds.iaea.org/publications (accessed on 1 November 2019).
  22. Pirahmadian, M.H.; Ghahramany, N. Helium-3 magnetic dipole moment determination by using quark constituents for all possible baryon formations. Results Phys. 2017, 7, 2771–2774. [Google Scholar] [CrossRef]
  23. Harris, R.K.; Becker, E.D.; Cabral de Menezes, S.; Granger, P.; Hoffman, R.E.; Zilm, K.W. Further conventions for NMR shielding and chemical shifts (IUPAC Reccomendantions 2008). Magn. Reson. Chem. 2008, 46, 582–598. [Google Scholar] [CrossRef]
  24. Makulski, W.; Garbacz, P. Gas-phase 21Ne NMR studies and the nuclear magnetic dipole moment of neon-21. Magn. Reson. Chem. 2020, 58, 648–652. [Google Scholar] [CrossRef]
  25. Makulski, W. 83Kr nuclear magnetic moment in terms of that of 3He. Magn. Reson. Chem. 2014, 52, 430–434. [Google Scholar] [CrossRef] [PubMed]
  26. Makulski, W. 129Xe and 131Xe Nuclear Magnetic Dipole Moments from Gas Phase NMR Spectra. Magn. Reson. Chem. 2015, 53, 273–279. [Google Scholar] [CrossRef] [PubMed]
  27. Chupp, T.E.; Oteiza, E.R.; Richardson, J.M.; White, T.R. Precision Measurements with polarized 3He, 21Ne and 129Xe atoms. Phys. Rev. A 1988, 38, 3998–4003. [Google Scholar] [CrossRef]
  28. Garbacz, P.; Jackowski, K.; Makulski, W.; Wasylishen, R.E. Nuclear Magnetic Shielding for Hydrogen in Selected Isolated Molecules. J. Phys. Chem. A 2012, 116, 11896–11904. [Google Scholar] [CrossRef] [PubMed]
  29. Kaupp, M.; Bühl, M.; Malkin, V.G. (Eds.) Calculation of NMR and EPR Parameters: Theory and Applications; Wiley: Hoboken, NJ, USA, 2004. [Google Scholar] [CrossRef]
  30. 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]
  31. Makulski, W.; Aucar, J.J.; Aucar, G. Ammonia: Molecule for establishing 14N and 15N absolute shielding scales and a source of information on nuclear magnetic moments. J. Chem. Phys. 2022; sent to publish. [Google Scholar]
  32. Makulski, W.; Garbacz, P. Gas-phase NMR of other nuclei than 1H and 13C. Chemistry, Molecular Sciences and Chemical Engineering. In Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier: Amsterdam, The Netherlands, 2021; pp. 1–16. [Google Scholar] [CrossRef]
  33. Lantto, P.; Jackowski, K.; Makulski, W.; Olejniczak, M.; Jaszuński, M. NMR shielding constants in PH3, absolute shielding scale and the nuclear magnetic moment of 31. J. Phys. Chem. A 2011, 115, 10617–10623. [Google Scholar] [CrossRef]
  34. 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] [Green Version]
  35. Makulski, W.; Wilczek, M.; Jackowski, K. 17O and 1H NMR Spectral Parameters in Isolated Water Molecule. Phys. Chem. Chem. Phys. 2018, 20, 22468–22476. [Google Scholar] [CrossRef]
  36. Makulski, W. Deuterium isotope effects on 17O nuclear shielding in a single molecule from NMR gas phase measurements. Phys. Chem. Chem. Phys. 2020, 22, 17777–17780. [Google Scholar] [CrossRef]
  37. Jackowski, K.; Makulski, W.; Szyprowska, A.; Antušek, A.; Jaszuński, M.; Jusélius, J. NMR shielding constants in BF3 and magnetic dipole moments of 10B and 11B nuclei. J. Chem. Phys. 2009, 130, 044309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Makulski, W.; Garbacz, P. Determination of 33-sulfur nuclear magnetic moment from gas phase NMR studies of 3He and SF6. In Proceedings of the 17-th EUROMAR, Ljubljana, Slovenia, 5–8 July 2021. [Google Scholar]
  39. Antušek, A.; Jackowski, K.; Jaszuński, W.; Makulski, W.; Wilczek, M. Nuclear magnetic dipole moments from NMR spectra. Chem. Phys. Lett. 2005, 411, 11–116. [Google Scholar] [CrossRef]
  40. Jaszuński, M.; Antušek, A.; Garbacz, P.; Jackowski, K.; Makulski, W.; Wilczek, M. The determination of accurate nuclear magnetic dipole moments and direct measurement of NMR shielding constants. Prog. NMR Spectrosc. 2012, 67, 49–63. [Google Scholar] [CrossRef] [PubMed]
  41. Ruud, K.; Demissie, T.B.; Jaszuński, M. Ab initio and relativistic DFT study of spin-rotation and NMR shielding constants in XF6 molecules, X=S, Se, Te, Mo, and W. J. Chem. Phys. 2014, 140, 194308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Garbacz, P.; Makulski, W. 183W nuclear dipole moment determined by gas-phase NMR spectroscopy. Chem. Phys. 2017, 498–499, 7–11. [Google Scholar] [CrossRef]
  43. Hindermann, D.K.; Cornwell, C.D. Fluorine and proton NMR Study of gaseous Hydrogen Fluoride. J. Chem. Phys. 1968, 48, 2017–2025. [Google Scholar] [CrossRef]
  44. Jameson, C.J.; Jameson, A.K. Absolute shielding scale for 29Si. Chem. Phys. Lett. 1988, 149, 300–305. [Google Scholar] [CrossRef]
  45. Tariq, N.; Al Taisan, N.; Singh, V.; Weinstein, J.D. Spectroscopic Detection of the LiHe Molecule. Phys. Rev. Lett. 2013, 110, 153201. [Google Scholar] [CrossRef] [Green Version]
  46. Antušek, A.; Kędziera, D.; Kaczmarek-Kędziera, 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]
  47. Makulski, W. The Radiofrequency NMR Spectra of Lithium Salts in Water; Reevaluation of Nuclear Magnetic Moments for 6Li and 7Li Nuclei. Magnetochemistry 2018, 4, 9. [Google Scholar] [CrossRef] [Green Version]
  48. Makulski, W. Multinuclear Magnetic Resonance Study of Sodium Salts in Water Solutions. Magnetochemistry 2019, 5, 68. [Google Scholar] [CrossRef] [Green Version]
  49. Ikeda, T.; Boero, M.; Terakura, K. Hydration of alkali ions from first principles molecular dynamics revisited. J. Chem. Phys. 2007, 126, 01B611. [Google Scholar] [CrossRef] [Green Version]
  50. Beckmann, A.; Böklen, K.D.; Elke, D. Precision measurements of the nuclear magnetic dipole moments of 6Li, 7Li, 23Na, 39K and 41K. Z. Phys. 1974, 270, 173–186. [Google Scholar] [CrossRef]
  51. Neronov, Y.I. Simultaneous Determination of the Magnetic Moments of 6Li and 7Li Nuclei Using an NMR Spectrometer. Meas. Tech. 2020, 63, 667–673. [Google Scholar] [CrossRef]
  52. Neronov, Y.I. Determination of the Magnetic Moment of a 23Na Nucleus Using an NMR Spectrometer with Simultaneous Detection of Signals from Two Nuclei. Tech. Phys. 2021, 66, 93–97. [Google Scholar] [CrossRef]
  53. Saunders, M.; Jiménez-Vázquez, H.A.; Cross, R.J.; Mroczkowski, S.; Freedberg, D.I.; 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]
  54. Khong, A.; Jiménez-Vázquez, H.A.; Saunders, M.; Cross, R.J.; Laskin, J.; Peres, T.; Lifshitz, C.; Strongin, R.; Smith, A.B. An NMR Study of He2 Inside C70. J. Am. Chem. Soc. 1998, 120, 6380–6383. [Google Scholar] [CrossRef]
  55. Sternfeld, T.; Hoffman, R.E.; Saunders, M.; Cross, J.; Symala, M.S.; Rabinovitz, M. Two Helium Atoms Inside Fullerenes: Probing the Internal Magnetic Field in C606− and C706−. J. Am. Chem. Soc. 2002, 124, 8786–8787. [Google Scholar] [CrossRef] [Green Version]
  56. Kupka, T. Noble gases as magnetic probes in fullerene chemistry. Emagres 2016, 5, 959–966. [Google Scholar] [CrossRef]
  57. Kupka, T.; Stachów, M. 3He NMR: From free gas to its encapsulation in fullerene. Magn. Reson. Chem. 2013, 51, 463–468. [Google Scholar] [CrossRef]
  58. Hayashi, S. Probing the Micropores in Linde-type A Zeolites by Helium-3 NMR. Chem. Lett. 2006, 35, 92–93. [Google Scholar] [CrossRef]
  59. Garbacz, P.; Jackowski, K. NMR shielding of helium-3 in the micropores of zeolites. Microporous Mesoporous Mater. 2015, 205, 52–55. [Google Scholar] [CrossRef]
  60. Tastevin, G.; Nacher, P.-J. NMR measurements of hyperpolarized 3He gas diffusion in high porosity silica aerogels. J. Chem. Phys. 2005, 123, 064506. [Google Scholar] [CrossRef] [Green Version]
  61. Klochkov, A.V.; Naletov, V.V.; Tagirov, M.S.; Tayurskii, D.A.; Yudin, A.N.; Zhdanov, R.S.; Zhdanov, M.R.; Bukharaev, A.A.; Nurgazizov, N.I. NMR and AFM Investigations of Nanocavities on the Double Rare-Earth Fluoride Crystal Surface. Appl. Magn. Reson. 2000, 19, 197–208. [Google Scholar] [CrossRef] [Green Version]
  62. Mooser, A.; Rischka, A.; Schneider, A.; Blaum, K.; Ulmer, S.; Walz, J. A New Experiment for the Measurement of the g-Factors of 3He+ and 3He2+. J. Phys. Conf. Ser. 2018, 1138, 012004. [Google Scholar] [CrossRef]
  63. Fan, X.; Fayer, S.E.; Gabrielse, G. Gaseous 3He nuclear magnetic resonance probe for cryogenic environments. Rev. Sci. Instrum. 2019, 90, 083107. [Google Scholar] [CrossRef] [Green Version]
  64. Hopkins, V.; McKenna, J.; Maynard, J.D. The use of 3He NMR as a probe of the nature of the 4He solid/superfluid interface. Bull. Am. Phys. Soc. 1993, 38, BAPSA6. [Google Scholar]
  65. Seydoux, R.; Muenster, O.; Diehl, P. 3He NMR in Liquid Crystals: Measurement of Local Magnetic Fields. Mol. Cryst. Liq. Cryst. Sci. Technol. A 1994, 250, 99–108. [Google Scholar] [CrossRef]
  66. Middleton, H.; Black, R.D.; Saam, B.; Cates, G.D.; Cofer, G.P.; Guenther, R.; Happer, W.; Hedlund, W.; Johnson, G.A.; Juvan, K.; et al. MR Imaging with Hyperpolarized 3He Gas. Mag. Res. Med. 1995, 33, 271–275. [Google Scholar] [CrossRef]
  67. Wehrli, D.; Spyszkiewicz-Kaczmarek, A.; Puchalski, M.; Pachucki, K. QED Effect on the Nuclear Magnetic Shielding of 3He. Phys. Rev. Lett. 2021, 127, 263001. [Google Scholar] [CrossRef]
Scheme 1. A set of nuclei which were studied by NMR spectroscopy using 3He as a reference for measuring its nuclear magnetic moments.
Scheme 1. A set of nuclei which were studied by NMR spectroscopy using 3He as a reference for measuring its nuclear magnetic moments.
Physchem 02 00009 sch001
Figure 1. NMR frequency dependences on the density of the gaseous noble gases 3He (in Xe), 21Ne, 83Kr, 129Xe, and 131Xe in the magnetic field B0 = 11.7574 T.
Figure 1. NMR frequency dependences on the density of the gaseous noble gases 3He (in Xe), 21Ne, 83Kr, 129Xe, and 131Xe in the magnetic field B0 = 11.7574 T.
Physchem 02 00009 g001
Figure 2. The 33S NMR spectra of gaseous SF6 at 21.8 atm: (A) coupled to six fluorine nuclei, and (B) decoupled fluorine.
Figure 2. The 33S NMR spectra of gaseous SF6 at 21.8 atm: (A) coupled to six fluorine nuclei, and (B) decoupled fluorine.
Physchem 02 00009 g002
Figure 3. 3He and 33S NMR frequencies as a density function of the gaseous mixture 3He-SF6.
Figure 3. 3He and 33S NMR frequencies as a density function of the gaseous mixture 3He-SF6.
Physchem 02 00009 g003
Table 1. Physical and magnetic properties of 3He atoms and nuclei.
Table 1. Physical and magnetic properties of 3He atoms and nuclei.
Property Property
Spin1/2Natural abundances0.000137%
Magnetic moment−2.127625308Isotope mass3.0160293 u
Chemical shift range58 ppmHalf lifestable
Frequency ratio76.18%Boiling point3.19 K
Reference compound3He gasCritical point3.35 K
Linewidth of reference0.3 HzHeat of vaporization0.026 kJ/mol
T1 of reference1000 sMelting pointbelow 1 mK
Receptivity rel.to 1H0.442 when enrichedCovalent radius32 pm
Magnetic susceptibility−1.88 cm3/molVan der Waals radius143 pm
Table 2. NMR parameters for calculating the µ(nX) nuclear magnetic moments of noble gases.
Table 2. NMR parameters for calculating the µ(nX) nuclear magnetic moments of noble gases.
Systemν0(nX)/ν0(3He)Correction Factorμ(nX)References
3He1.00001.000059965−2.127625308(10)[19,20]
−2.127625307(10) *[25,26]
21Ne/3He0.1036381.0004974220.66184(7)[27]
0.6617774(10) *[24]
83Kr/3He0.0505161561(5)1.003529996−0.97072965(32)[25]
−0.97072965(32) *
129Xe/3He0.36309748(3)1.007022726−0.777961(16)[26]
−0.777961(16) *
131Xe/3He0.1076349191.0070227260.691845(7)[26]
0.691845(7) *
* Nuclear magnetic moments from 2H(D) parameters.
Table 3. NMR parameters in simple hydrides for the calculation of the nuclear magnetic moments of 13C, 14N, 17O, 29Si, 31P, and 35/37Cl nuclei from gas-phase experiments.
Table 3. NMR parameters in simple hydrides for the calculation of the nuclear magnetic moments of 13C, 14N, 17O, 29Si, 31P, and 35/37Cl nuclei from gas-phase experiments.
Systemν0(nX)/ν0(3He)Correction Factorμ(nX)References
13CH4/3He0.330074361(2)1.000135070.70236944(68)[30]
0.70236945(68) *
14NH3/3He0.094821748(5)1.000206880.40357377(45)[31]
0.40357367(40) *
H217O/3He0.177949095(15)1.000268521.893553(3)[36]
1.893553(2) *[35]
29SiH4/3He0.260768297(3)1.00042308−0.5550520(3)
−0.5550520(3) *[32]
31PH3/3He0.5312482461.0005551321.1309247(50)[33]
1.1309246(50) *
H35Cl/3He0.128620431.0009171300.821716(5)[34]
0.821721(4) *
H37Cl/3He0.070630371.000917130−0.683997(5)
−0.683997(4) *
* indicates values calculated from 1H NMR data.
Table 4. NMR parameters for calculations for the nuclear magnetic moments of 10/11B, 33S, and 183W against that of 3He.
Table 4. NMR parameters for calculations for the nuclear magnetic moments of 10/11B, 33S, and 183W against that of 3He.
Systemν0(nX)/ν0(3He)Correction Factorμ(nX)μNReference
10BF3/3He0.1410332381.0000379161.8004636(8)[37]
1.80045428 *
11BF3/3He0.4211700451.0000379162.6883781(11)
2.6883642 *
33SF6/3He0.1007448021.00039260.6432555(10)[38]
0.6432467(16) *
183WF6/3He0.0546302641.00619960.116953(18)[42]
1.00622860.116953 *
1.00619890.116950 **
* NMM measured from 19F parameters. ** NMM measured from 19F parameters with the theoretical shielding correction factor.
Table 5. NMR parameters for the calculation of 6/7Li and 23Na nuclear magnetic moments from water solutions. * NMM calculated from 2H(D) parameters.
Table 5. NMR parameters for the calculation of 6/7Li and 23Na nuclear magnetic moments from water solutions. * NMM calculated from 2H(D) parameters.
Systemν(6Li)/ν(3He)Correction Factorμ(6Li)Reference
Li+ water solution/3He0.1931777451.0000315630.8220456(25)[47]
0.958660151.0000624590.8220432(25) *2H(D)
0.8220445(10)[50] ABMR
0.8220454(25)[51]
ν(7Li)/ν(3He)
0.5101643031.0000315633.2564182(98)[47]
2.5317315241.0000624593.2564085(98) *2H(D)
3.2564157(2)[50] ABMR
3.2564171(98)[51]
Na+ water solution/3Heν(23Na)/ν(3He)
0.3472335711.0005207142.2174997(111)[48]
1.7231746111.0005516032.2174962(111) *2H(D)
2.2175019(133)ABMR
2.2175065(55)[52]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Makulski, W. Probing Nuclear Dipole Moments and Magnetic Shielding Constants through 3-Helium NMR Spectroscopy. Physchem 2022, 2, 116-130. https://doi.org/10.3390/physchem2020009

AMA Style

Makulski W. Probing Nuclear Dipole Moments and Magnetic Shielding Constants through 3-Helium NMR Spectroscopy. Physchem. 2022; 2(2):116-130. https://doi.org/10.3390/physchem2020009

Chicago/Turabian Style

Makulski, Włodzimierz. 2022. "Probing Nuclear Dipole Moments and Magnetic Shielding Constants through 3-Helium NMR Spectroscopy" Physchem 2, no. 2: 116-130. https://doi.org/10.3390/physchem2020009

Article Metrics

Back to TopTop