Recent advances in solid-state nuclear magnetic resonance spectroscopy of exotic nuclei

https://doi.org/10.1016/j.pnmrs.2018.08.002Get rights and content

Highlights

  • Progress in SSNMR of exotic nuclei is reviewed.

  • The period 2007–2017 is covered.

  • Advances in hardware are described.

  • A survey of literature data for exotic nuclides is provided.

Abstract

We present a review of recent advances in solid-state nuclear magnetic resonance (SSNMR) studies of exotic nuclei. Exotic nuclei may be spin-1/2 or quadrupolar, and typically have low gyromagnetic ratios, low natural abundances, large quadrupole moments (when I > 1/2), or some combination of these properties, generally resulting in low receptivities and/or prohibitively broad line widths. Some nuclides are little studied for other reasons, also rendering them somewhat exotic. We first discuss some of the recent progress in pulse sequences and hardware development which continues to enable researchers to study new kinds of materials as well as previously unfeasible nuclei. This is followed by a survey of applications to a wide range of exotic nuclei (including e.g., 9Be, 25Mg, 33S, 39K, 43Ca, 47/49Ti, 53Cr, 59Co, 61Ni, 67Zn, 73Ge, 75As, 87Sr, 115In, 119Sn, 121/123Sb, 135/137Ba, 185/187Re, 209Bi), most of them quadrupolar. The scope of the review is the past ten years, i.e., 2007–2017.

Introduction

During the last decade, experiments in the field of solid-state nuclear magnetic resonance (SSNMR) have undergone tremendous technical advances as a result of improvements in pulse sequences, techniques, and hardware. Various NMR methods have become increasingly accessible and widely implemented, enabling the detailed analysis of a multitude of materials including proteins, catalysts, biomaterials, pharmaceuticals, metal-organic materials, and so on. NMR spectroscopy is well known for its ability to probe molecules and materials at an atomic scale. However, considering the nature of various materials of interest, NMR spectroscopy has had to adapt. While Bragg diffraction techniques can determine the long-range periodic structure of powdered samples, SSNMR of a powdered sample offers local, isotope-specific structural information. This complementarity is of great interest for crystallography and chemistry in a broad sense. Facilitated by advances in computational chemistry, a new area of SSNMR spectroscopy has emerged under the term “NMR crystallography” [1]. Combining information gathered from NMR, diffraction, and computational approaches can afford a complete description of the overall chemical and crystallographic structure of a compound. Despite the inherent complications of interpreting some spectra, NMR results can help to understand short range interactions, and is seen as a local probing technique [2].

Although nearly every element in the periodic table has an NMR-active isotope (spin I ≠ 0), some of these isotopes are much less receptive and amenable to NMR, with limited reports available in the literature. Such isotopes are usually referred to as exotic nuclei (see Table 1). The Oxford Dictionary states that ‘exotic’ denotes ‘[] of a kind not ordinarily encountered; [special]; out of the ordinary’. The dedicated expression ‘exotic nuclei’ seems appropriate in the context of SSNMR. In this contribution, we review the recent advances and major breakthroughs made in the field of SSNMR investigations of exotic nuclei, specifically in the last decade (2007–2017). While there have been many contributions in this field, this review aims to capture the major progress and most notable studies. To provide a working definition of the ‘exotic nuclei’ covered herein (see Section 1.3), various properties of nuclides are first discussed below.

The difficulties in studying exotic nuclei arise from several inherent isotopic properties: low gyromagnetic ratio (γ), low natural abundance (NA), high quadrupole moment (Q), and quantum spin number (I). That being said, most isotopes of interest here are quadrupolar (I > 1/2), which represent about 75% of the NMR active nuclei (see Fig. 1). However, this review also covers a number of spin I = 1/2 nuclei that fit into general criteria of being little-studied or difficult to analyse.

When the degeneracy of the Zeeman levels is lifted by the application of an external magnetic field (B0), 2(I + 1/2) energy levels and 2I single quantum transitions result. Therefore, spin-1/2 nuclei only exhibit one transition whereas quadrupolar nuclei exhibit multiple single-quantum transitions. There are two types of single-quantum transitions for half-integer quadrupolar nuclei: the central transition (CT) and the satellite transitions (ST). To first order, only the STs are impacted by the quadrupolar interaction (QI), resulting in anisotropic spectral broadening which is typically much larger than that experienced by the CT. The QI must then be considered to second-order to explain the position and line shapes perturbations of the CT. Quadrupolar nuclei have a non-spherical charge distribution within the nucleus and a concomitant non-vanishing nuclear electrical quadrupole moment, Q (often expressed in millibarn (mb, equal to 10−31 m2)) [3]. The quadrupolar coupling of Q with the electric field gradient (EFG, described by a traceless, symmetric second-rank tensor with principal components: V11, V22, V33) at the nucleus can reach up to several hundreds or thousands of MHz, much larger than all other internal interactions. As a result, the NMR spectra of quadrupolar nuclei are typically much broader than those of spin-1/2 nuclei. Despite this broadening, data extracted from such spectra are of great interest as they offer insight at the location of the nucleus itself, which is very useful for characterizing molecules and materials (vide infra). The magnitude of the nuclear electric QI can be described by the quadrupolar coupling constant (CQ) and the EFG asymmetry parameter (ηQ):CQ=eQV33hηQ=V11-V22V33Here, V11, V22, and V33 (with V11 + V22 + V33 = 0) are the principal components of the traceless electric field gradient (EFG) tensor, with |V33| ≥ |V22| ≥ |V11|, e is the fundamental charge, h is Planck’s constant, and Q is the nuclear quadrupole moment.

The line width factor, (in fm4) [4], is usually defined as follows in solution NMR:=Q2(2I+3)I2(2I-1)

This parameter can offer a rapid and rough estimation of the extent of signal broadening. In order to get a more accurate and suitable value, Wu and Zhu proposed a modified line-width factor for the central transition signal, CT [5], that allows one to predict the CT line width for solid-state signals. CT is expressed in fm4 rad−1 T s in the following equation:CT=1|γ|3Q(1-γ)2I(2I-1)2I(I+1)-34where γ is the Sternheimer antishielding factor [6], [7]; it corresponds to a correction of the EFG at the nucleus caused by the polarization of inner-shell electrons.

In the solid state, the chemical shift anisotropy (CSA) is not averaged by molecular tumbling (Brownian motion) as in the isotropic liquid state, and thus directly impacts the line shape of a given spectral signal. Within the Herzfeld-Berger convention, the magnitude of this interaction can be described by three parameters: the isotropic chemical shift (δiso), the span (Ω), and skew (κ):δiso=δ11+δ22+δ333Ωδ11-δ33(Ω0)κ=3(δ22-δiso)Ω;(-1κ+1)

The principal components of the chemical shift (CS) tensor are ordered as δ11 ≥ δ22 ≥ δ33. The chemical shift range depends on the studied isotope and can be as wide as, e.g., 6800 ppm for 185/187Re or as narrow as 50 ppm for 9Be [8], [9].

The receptivity represents the ease with which a signal of a given isotope may be acquired relative to another one [4]:Receptivityγ3NAI(I+1)

For example, the receptivity of 43Ca with respect to 13C (at natural abundance of 1.11%) is only 0.05 [10], which makes it a rather difficult nucleus to study; for comparison, the receptivity of 9Be relative to 13C is 81.5 [4]. Lowly receptive nuclei are more arduous to study and receptivity values should be considered in any definition of ‘exotic nuclei’. Presented in Table 1 is a liberally defined list of nuclides which may be considered as exotic, for various reasons. For example, 9Be is included due to the dearth of studies in the literature, despite its good receptivity and small quadrupole moment. 17O is listed due its low receptivity; however, it is clear that this isotope is broadly studied in the literature (often with isotopic enrichment) and can hardly be considered as exotic nowadays (vide infra). Other nuclides with good receptivities may be included due to their prohibitively large nuclear electric quadrupole moments (e.g., 115In, 187Re, or 209Bi). Thus, as stated in the table heading, the list summarizes the nuclei that we consider to be exotic; they are further discussed in the following sections of this review. The frequency ratios, Ξ, which corresponds to the ratio of frequency of the reference to that of the protons of TMS at high dilution (1%) in CDCl3 are also given in Table 1 [4].

One major difficulty associated with studying some exotic nuclei is the ubiquitous Achilles’ heel of NMR spectroscopy: sensitivity. Indeed, although SSNMR offers insights at the molecular scale for any spin-active isotope, sensitivity is the most significant challenge that NMR spectroscopy faces when compared with many other analytical methods. In a more general context, materials chemistry deals with new compounds designed to present particular and desired properties. The structures of such materials are typically linked to their properties. X-ray diffraction may not be applicable to certain materials (e.g., amorphous compounds), while SSNMR can still provide nucleus-specific information. However, combined with some of the other properties inherent to particular nuclides (vide supra), sensitivity issues can become very problematic. Various strategies for increasing the sensitivity of the NMR experiment, with a focus on exotic nuclides, are discussed briefly below. An overview of more modern signal enhancement methods is given in Section 2.

The observed SSNMR signal is directly linked to various factors including the natural abundance of the isotope. Oxygen-17 has one of the lowest natural abundances in the periodic table at only 0.038%. Other important elements are also present at a very low natural abundance, e.g., 43Ca and 15N with natural abundances of 0.135% and 0.37%, respectively. One brute force way to counter this problem is to consider isotopic enrichment of the sample. For example, 17O enrichment is strongly advised in order to observe any signal (although there are exceptions to this rule now as a result of advances in dynamic nuclear polarization methodologies, vide infra). This nucleus is of great interest in several areas, from pharmaceuticals [12], [13], [14] to geochemistry [15], inorganic compounds [16] or biomolecules (amino acids, proteins, etc.) [17], [18]. Enrichment is carried out by using an 17O-labelled precursor; however, synthetic yields can be a limiting factor, and the enrichment processes can be challenging. Enriched compounds can be expensive. Moreover, the number of enriched synthetic compounds which are practically accessible is reduced due to the limited number of available enriched precursors. Recently, Métro et al. proposed an appealing synthetic route for 17O enrichment via mechanochemical synthesis [19]. They used a ball mill for the 17O-enrichment of various compounds, from ibuprofen to Sr(OH)2·8H2O. This method requires at most a few hundreds of μL of isotopically-enriched precursor (e.g., H217O) per synthesis, as opposed to a few mL with classical wet chemistry enrichment processes. Enrichment methods are not always trivial and most of the time very costly; therefore, other signal enhancement approaches must absolutely be pursued and implemented.

To characterize chemical shift and quadrupolar coupling tensors, experiments carried out on stationary and magic-angle spinning powders are often useful and complementary. MAS has long been part of the toolkit for the spectroscopist. The advantages offered by this method have greatly helped SSNMR studies over the past decades [20], [21], [22]. As is well known to the readers of this journal, the method consists of spinning the sample at 54.74° (θ-value to cancel the second Legendre polynomial: P2(cos θ) = (3cos2θ − 1)/2) with respect to B0. Magnetic shielding can be described mathematically as a sum of isotropic and anisotropic (and antisymmetric) parts. The dependence of the anisotropic component on P2 implies that it can be averaged by MAS. In favourable cases, the signal is reduced from a wide anisotropic powder pattern to a sharp isotropic peak. However, if the spinning speed is smaller than the magnitude of the CS interaction, spinning sidebands are observed in the spectrum. These sidebands are separated by the spinning frequency, the envelope of the overall spectrum mimicking the static line shape. The QI can also be described in a similar fashion. The first-order interaction is effectively averaged by MAS; nevertheless, spinning sidebands are observed for STs as the QI magnitude is usually much larger than the spinning speed. In both cases of CSA and first-order QI, the spectrum appears as a family of spinning sidebands if the interaction is not sufficiently averaged by MAS. With regards to the second-order QI, the mathematical description shows that it depends on two Legendre polynomials (P2 and P4) which cannot be cancelled simultaneously by sample spinning about a single axis (vide infra). Consequently, the signal will be narrowed but MAS does not completely remove the second-order quadrupolar broadening.

Larger volume MAS rotors allow a larger number of spins to be analysed; however, as can be seen in Fig. 2, the relative sensitivity per unit volume is greater for smaller rotors.

Typical rotor diameters of 7 mm, 4 mm, 3.2 mm, or 2.5 mm can safely spin up to ∼8, ∼15, ∼25 and ∼35 kHz, respectively. However, these speeds are not always fast enough to entirely average the CSA or the QIs, and so experimentalists might consider so-called ultrafast probes. Nowadays, ultrafast (>40–50 kHz) experiments are becoming more routine thanks to tremendous technical advances in NMR probe design, motivated in large part by the advantages of averaging of 1H–1H dipolar couplings. In order to be able to spin faster, the rotor diameter is reduced (e.g., 0.7 mm and 0.75 mm o.d. probes are offered by Bruker and JEOL which can spin in the 110–120 kHz range). Recent breakthroughs in the development of MAS NMR probes with spinning frequencies exceeding 100 kHz have opened a wide range of potential applications [24], [25], [26], [27]. During MAS experiments, it is well-known that in the absence of external temperature regulation, the sample heats up, especially if the spinning speed is high [26], [28]. This also applies to the inner pressure arising from the centrifugal forces. Therefore, attention should be paid to the thermal stability of the compound and to the temperature regulation, when the spinning speed reaches very high frequencies. As already mentioned, working at ultrafast spinning speed presents a major trade-off, e.g., a low spin number; such a method is thus often used for high resolution NMR (typically of protons) and benefits from dipolar coupling averaging.

The MAS NMR spectra of heavy spin-1/2 nuclides like platinum-195 can be rendered complex by very large CSA that results in numerous spinning sidebands. In 2017, Perras et al. published an article describing a pulse sequence (D-HMQC-MAT) that allows one to record MAS spectra corresponding to infinite speed of heavy spin-1/2 nuclides by indirect detection [29]. They applied it to Pt-containing metal-organic frameworks (MOFs) which display ultra-wide spectra (Fig. 3) and successfully resolved Pt coordination environments in complex mixtures.

The comparison between the three pulse sequences, CT-D-HMQC, D-HMQC-aMAT and D-HMQC-MAT, gives 1D 1H signal intensity ratios of 1.00, 0.52 and 0.06, respectively (first slice) (Fig. 3). However, the MAT advantage resides in the two-dimensional experiment, as shown in Fig. 3: sensitivity is greatly improved by folding the spinning sidebands into the centerband. Moreover, the centerband intensity is now greatly increased. The authors also mention that this strategy could be competitive and complementary to Dynamic Nuclear Polarization (DNP).

Larmor frequencies are directly proportional to the applied external magnetic field, B0 (νL=|γ|B0/2π). This implies that ultrahigh magnetic fields (UHF) provide a tremendous advantage for low-γ nuclei. From the Boltzmann distribution, it is easily understood that the spin polarization is related to the Larmor frequency, thus implying that low-γ nuclei will display a poorer sensitivity. Working at UHF not only improves the S/N directly; the effects of the QI and CSA also depend on B0. Second-order quadrupolar broadening is reduced (as inversely proportional to B0), thus implying a resolution enhancement, but the CSA is proportional to the external magnetic field (in frequency units) and is therefore increased at UHF. The highest magnetic field currently available for NMR experiments is about 45 T and is used for experiments where high homogeneity is not needed [30]; however, a new series-connected resistive/superconducting hybrid magnet operating with good homogeneity up to 35.2 T (corresponding to 1.5 GHz for νL(1H)) in the National High Magnetic Field Laboratory, Tallahassee, has recently provided very promising results and potential for future applications [31]. Multifield experiments can also be of great help, and are often essential, to refine the quadrupolar and CS parameters extracted from spectra if the sensitivity is good enough [32]. NQR experiments are quite often a good alternative to measure the quadrupolar coupling constant of exotic nuclei, especially when the quadrupole moment is large [33].

This very brief presentation summarizes the basic strategies commonly accessible for SSNMR experiments on difficult nuclei. To briefly recapitulate, when considering SSNMR experiments, one should first of all carefully think about the pulse sequence, the rotor size and probe used, the magnetic field available and the experiments (MAS, static) that are required before analysing the sample. These practical considerations are particularly important for exotic nuclei. Further description of more advanced pulse sequences and techniques constitutes a separate section of this review wherein we will elaborate on the advantages and possibilities offered by some specific sequences.

Nuclides with integer spin quantum number (I = n with n ∈ Z+) represent a small fraction of quadrupolar isotopes in the periodic table. Their particularity lies in the fact that no central transition can be observed as the Zeeman interaction leads to an odd number of energy levels (2I + 1, see Fig. 4). Therefore, their NMR spectra will be highly affected by the first-order quadrupolar coupling. The most commonly studied integer spin nuclides are hydrogen-2 (deuterium), nitrogen-14, and lithium-6; however, others exist including boron-10, potassium-40, vanadium-50, lanthanum-138, and lutetium-176 (with I = 3, 4, 6, 5, and 7, respectively). Only the first three will be briefly reviewed here as the other ones are extremely rare if not absent from the literature during the last ten years.

As an illustration, consider 14N, a spin I = 1 nucleus. It displays two different kinds of transitions between the three energy levels: two single-quantum (SQ, +1 ↔ 0 and 0 ↔ −1) and one double quantum (DQ, +1 ↔ −1) transitions.

Fig. 4 shows that SQ transitions are highly affected by first-order QI whereas DQ transitions will only be affected by second and higher order QIs. The SQ transitions thus generally lead to broad signals whereas the DQ transitions present narrower spectra. Direct generation of DQ coherence is called overtone transition (OT) NMR spectroscopy (with rf irradiation at 2νL(14N)). A review by Dib et al. has nicely recapitulated recent works focused on nitrogen-14 [35]. As stated in the beginning of their review, spectroscopists usually prefer to study 15N despite its very low natural abundance (0.37% vs. 99.63% for 14N). This is due partially to the relative ease of labelling with this isotope, but also because 15N is a spin-1/2 nuclide which circumvents the difficulties arising from the quadrupolar 14N isotope. Readers are therefore directed to this review article if they are interested in more details on this isotope.

The second interesting integer spin nuclide is hydrogen-2 (2H or 2D) which is also a spin-1 nuclide. This isotope has also a very low NA (0.011%), which when combined with a medium γ (4.11 × 107 rad s−1 T−1) and a small Q (2.8 mb) results in a receptivity of 1.1 × 10−6 relative to 1H. In the last 10 years, various articles have used 2H SSNMR in order to study dynamics [36], [37], [38], adsorption [39], or dehydration [40], [41]. Finally, lithium-6 (I = 1) is cited here but will not be further discussed as we will not include it in our definition of exotic nuclei; indeed extensive literature is easily found, along with work on the isotope 7Li (I = 3/2).

For the purposes of this article, we propose a practical definition for exotic nuclei which includes those which suffer from (i) typically moderate to large nuclear quadrupolar coupling constants, (ii) low natural abundance, (iii) low gyromagnetic ratio, or combinations thereof. We also discuss some nuclides which for other reasons have been rarely studied, thereby qualifying them as ‘exotic’ as well.

In the first part of the remainder of this article, some techniques to counter the sensitivity problems of exotic nuclei are discussed. These range from manipulating the satellite transitions with the help of frequency sweeps to hyperpolarization of the sample via dynamic nuclear polarization (DNP). In the second part, we will survey the literature of the past decade for so-called exotic nuclides. The review covers roughly the past ten years, but is not meant to be exhaustive. Readers are referred to excellent review articles relating to specific nuclei where appropriate.

Section snippets

Sensitivity enhancement

There are numerous pulse sequences available to maximize the information available from SSNMR experiments. Selection of a particular experiment allows one to focus the study on the desired NMR interactions in a given compound. Most of nuclei covered in this review are quadrupolar, which may require pulse sequences that differ from those used for spin-1/2 nuclides for optimum signals. The standard cross-polarization (CP) sequence used for spin-1/2 nuclides is not as easily or broadly applicable

Survey of data: s-block nuclei

The s-block of nuclei comprises the first two groups (Group 1 & Group 2 + helium) of the periodic table (see Fig. 1). From a SSNMR perspective, all of these elements have at least one NMR active isotope except for the two radioisotopes, francium and radium. The s-block is composed of two groups (alkali metals and alkaline earth metals) which will be discussed separately.

Survey of data: d-block nuclei

Transition and post-transition metals compose the d-block. NMR interaction tensors can offer valuable information pertaining to the bonding geometries of such elements; this can be valuable given the range of oxidation states available to d-block elements and their corresponding coordination environments. While metallic materials give rise to spectra dominated by the Knight shift, and metals in particular oxidation states can lead to paramagnetism of molecular species, the focus below is on

Survey of data: p-block nuclei

The p-block contains familiar elements such as carbon, oxygen, nitrogen, aluminium, silicon, phosphorus, and so on. Such elements are very important for a broad array of systems, ranging from biological molecules to glasses and inorganic materials. Unsurprisingly, there is a wealth of NMR spectroscopic data for these elements. In this section, we discuss the exotic nuclei of groups 13, 14, 15, and 16 of the periodic table: icosagens, tetrels (or crystallogens), pnictogens, and chalcogens.

Concluding remarks

We have provided a survey of experimental techniques and experimental data relating to solid-state NMR of exotic nuclei, covering the period 2007–2017. This is a very broad field and as such the review is not meant to be exhaustive. Looking forward, we can speculate on the future of SSNMR spectroscopy of exotic nuclei. Ultrahigh magnetic fields will continue to offer advantages for isotopes with low natural abundances and/or large nuclear quadrupole moments. The series-connected hybrid (SCH)

Acknowledgements

We are grateful to Scott Southern, Patrick Szell, and Yijue Xu for helpful comments on the manuscript. D.L.B. thanks the Natural Sciences and Engineering Research Council of Canada for funding.

References (300)

  • K.J. Harris et al.

    Broadband adiabatic inversion pulses for cross polarization in wideline solid-state NMR spectroscopy

    J. Magn. Reson.

    (2012)
  • R. Fu et al.

    Frequency-modulated cross-polarization for fast magic angle spinning NMR at high fields: relaxing the Hartmann-Hahn condition

    Chem. Phys. Lett.

    (1997)
  • R. Fu et al.

    Recoupling of heteronuclear dipolar interactions in solid state magic-angle spinning NMR by simultaneous frequency and amplitude modulation

    Chem. Phys. Lett.

    (1997)
  • A.J. Vega

    CP/MAS of quadrupolar S = 3/2 nuclei

    Solid State Nucl. Magn. Reson.

    (1992)
  • P.J. Barrie

    Distorted powder lineshapes in 27Al CP/MAS NMR spectroscopy of solids

    Chem. Phys. Lett.

    (1993)
  • X. Zhao et al.

    Heteronuclear polarization transfer by symmetry-based recoupling sequences in solid-state NMR

    Solid State Nucl. Magn. Reson.

    (2004)
  • J.D. van Beek et al.

    Symmetry-based recoupling of 17O–1H spin pairs in magic-angle spinning NMR

    J. Magn. Reson.

    (2006)
  • D. Lee et al.

    Is solid-state NMR enhanced by dynamic nuclear polarization?

    Solid State Nucl. Magn. Reson.

    (2015)
  • F. Mentink-Vigier et al.

    Fast passage dynamic nuclear polarization on rotating solids

    J. Magn. Reson.

    (2012)
  • J.H. Lee et al.

    Sensitivity enhancement in solution NMR: emerging ideas and new frontiers

    J. Magn. Reson.

    (2014)
  • A. Bornet et al.

    Microwave frequency modulation to enhance dissolution dynamic nuclear polarization

    Chem. Phys. Lett.

    (2014)
  • P. Berthault et al.

    Biosensing using laser-polarized xenon NMR/MRI

    Prog. Nucl. Magn. Reson. Spectrosc.

    (2009)
  • A.S. Lilly Thankamony et al.

    Dynamic nuclear polarization for sensitivity enhancement in modern solid-state NMR

    Prog. Nucl. Magn. Reson. Spectrosc.

    (2017)
  • Z. Yao et al.

    Sensitivity enhancement of the central transition NMR signal of quadrupolar nuclei under magic-angle spinning

    Chem. Phys. Lett.

    (2000)
  • D. Iuga et al.

    Population and coherence transfer induced by double frequency sweeps in half-integer quadrupolar spin systems

    J. Magn. Reson.

    (2000)
  • A.P. Kentgens et al.

    Advantages of double frequency sweeps in static, MAS and MQMAS NMR of spin I = 3/2 nuclei

    Chem. Phys. Lett.

    (1999)
  • L.A. O’Dell et al.

    Acquisition of ultra-wideline NMR spectra from quadrupolar nuclei by frequency stepped WURST–QCPMG

    Chem. Phys. Lett.

    (2009)
  • A.W. MacGregor et al.

    New methods for the acquisition of ultra-wideline solid-state NMR spectra of spin-1/2 nuclides

    J. Magn. Reson.

    (2011)
  • L.A. O’Dell

    The WURST kind of pulses in solid-state NMR

    Solid State Nucl. Magn. Reson.

    (2013)
  • R. Siegel et al.

    Signal enhancement of NMR spectra of half-integer quadrupolar nuclei in solids using hyperbolic secant pulses

    Chem. Phys. Lett.

    (2004)
  • R.K. Harris et al.

    NMR Crystallography

    (2009)
  • S.E. Ashbrook et al.

    Recent developments in solid-state NMR spectroscopy of crystalline microporous materials

    Phys. Chem. Chem. Phys.

    (2014)
  • P. Pyykkö

    Year-2008 nuclear quadrupole moments

    Mol. Phys.

    (2008)
  • R.K. Harris et al.

    NMR nomenclature. Nuclear spin properties and conventions for chemical shifts (IUPAC Recommendations 2001)

    Pure Appl Chem.

    (2001)
  • R. Sternheimer

    On nuclear quadrupole moments

    Phys. Rev.

    (1950)
  • R. Sternheimer

    On nuclear quadrupole moments

    Phys. Rev.

    (1951)
  • C.M. Widdifield et al.

    Definitive solid-state 185/187Re NMR spectral evidence for and analysis of the origin of high-order quadrupole-induced effects for I = 5/2

    Phys. Chem. Chem. Phys.

    (2011)
  • D.L. Bryce et al.

    Beryllium-9 NMR study of solid bis(2,4-pentanedionato- O,O’)beryllium and theoretical studies of 9Be electric field gradient and chemical shielding tensors. First evidence for anisotropic beryllium shielding

    J. Phys. Chem. A

    (1999)
  • P. Pyykkö

    Year-2017 nuclear quadrupole moments

    Mol. Phys.

    (2018)
  • X. Kong et al.

    Solid-state 17O NMR of pharmaceutical compounds: salicylic acid and aspirin

    J. Phys. Chem. B

    (2013)
  • F.G. Vogt et al.

    17O solid-state NMR as a sensitive probe of hydrogen bonding in crystalline and amorphous solid forms of diflunisal

    Mol. Pharm.

    (2013)
  • X. Kong et al.

    A solid-state 17O NMR study of platinum-carboxylate complexes: carboplatin and oxaliplatin

    Can. J. Chem.

    (2015)
  • S.K. Lee et al.

    Oxygen-17 nuclear magnetic resonance study of the structure of mixed cation calcium−sodium silicate glasses at high pressure: implications for molecular link to element partitioning between silicate liquids and crystals

    J. Phys. Chem. B

    (2008)
  • S.E. Ashbrook et al.

    Solid state 17O NMR—an introduction to the background principles and applications to inorganic materials

    Chem. Soc. Rev.

    (2006)
  • O.N. Antzutkin et al.

    Hydrogen bonding in Alzheimer’s amyloid-β fibrils probed by 15N{17O} REAPDOR solid-state NMR spectroscopy

    Angew. Chemie Int. Ed.

    (2012)
  • E.G. Keeler et al.

    17O MAS NMR correlation spectroscopy at high magnetic fields

    J. Am. Chem. Soc.

    (2017)
  • T.-X. Métro et al.

    Unleashing the potential of 17O NMR spectroscopy using mechanochemistry

    Angew. Chemie Int. Ed.

    (2017)
  • E.R. Andrew et al.

    Nuclear magnetic resonance spectra from a crystal rotated at high speed

    Nature.

    (1958)
  • E.R. Andrew et al.

    Removal of dipolar broadening of nuclear magnetic resonance spectra of solids by specimen rotation

    Nature

    (1959)
  • I.J. Lowe

    Free induction decays of rotating solids

    Phys. Rev. Lett.

    (1959)
  • Cited by (35)

    • A review of exotic quadrupolar metal nmr in mofs

      2023, Comprehensive Inorganic Chemistry III, Third Edition
    • Solid-state NMR of quadrupolar nuclei: Selected new methods and applications

      2022, Annual Reports on NMR Spectroscopy
      Citation Excerpt :

      In particular, in addition to the more favourable Boltzmann distribution at higher field, the well-known narrowing of the central transition (CT) due to a reduction in second-order quadrupolar broadening effects confers improved spectral resolution. A number of relatively recent articles and reviews have addressed different aspects of the field of solid-state NMR of quadrupolar nuclei [1–5]. Perras et al. reviewed a range of signal enhancement methods in 2013 [1].

    View all citing articles on Scopus
    View full text