Draft · 2020. 6. 23. · a sterically unfavourable C-O-C bond angle of about 60o),oxiranes are...

35
Draft A combined solid-state 17O NMR, crystallographic, and computational study of oxiranes Journal: Canadian Journal of Chemistry Manuscript ID cjc-2020-0114.R1 Manuscript Type: Article Date Submitted by the Author: 13-Apr-2020 Complete List of Authors: Rinald, Andrew; Queen's University Terskikh, Victor; University of Ottawa, Chemistry Schatte, Gabriele; Queen's University Wu, Gang; Queen's University, Is the invited manuscript for consideration in a Special Issue?: Not applicable (regular submission) Keyword: Oxirane, solid-state 17O NMR, quadrupole coupling tensor, chemical shift tensor https://mc06.manuscriptcentral.com/cjc-pubs Canadian Journal of Chemistry

Transcript of Draft · 2020. 6. 23. · a sterically unfavourable C-O-C bond angle of about 60o),oxiranes are...

  • Draft

    A combined solid-state 17O NMR, crystallographic, and computational study of oxiranes

    Journal: Canadian Journal of Chemistry

    Manuscript ID cjc-2020-0114.R1

    Manuscript Type: Article

    Date Submitted by the Author: 13-Apr-2020

    Complete List of Authors: Rinald, Andrew; Queen's UniversityTerskikh, Victor; University of Ottawa, ChemistrySchatte, Gabriele; Queen's UniversityWu, Gang; Queen's University,

    Is the invited manuscript for consideration in a Special

    Issue?:Not applicable (regular submission)

    Keyword: Oxirane, solid-state 17O NMR, quadrupole coupling tensor, chemical shift tensor

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    1

    A combined solid-state 17O NMR, crystallographic, and

    computational study of oxiranes

    Andrew Rinald,1 Victor Terskikh,1,2 Gabriele Schatte,1 and Gang Wu1*

    1Department of Chemistry, Queen’s University, 90 Bader Lane,

    Kingston, Ontario, Canada K7L 3N6; 2Department of Chemistry, University of Ottawa,

    Ottawa, Ontario, Canada K1A 0R6

    *Corresponding author: [email protected]

    Page 1 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    2

    Abstract

    We report synthesis and solid-state 17O NMR characterization of three 17O-labeled

    oxiranes: (2S*, 3S*)-2,3-bis(4-nitrophenyl)-[17O]oxirane, (2S*, 3R*)-2,3-bis(4-

    nitrophenyl)-[17O]oxirane, and 2,2,3-triphenyl-[17O]oxirane. In addition, we have

    determined the crystal structure of (2S*, 3R*)-2,3-bis(4-nitrophenyl)oxirane by X-ray

    crystallography. When the experimentally determined 17O NMR tensors for oxiranes

    (where the C-O-C bond angle is about 60) are compared with those for dimethyl ether

    (where the C-O-C bond angle is 113) and other R-O-R functional groups, we found that

    the highly constrained geometry of oxiranes results in distinct tensor orientations in the

    molecular frame of reference. The experimental results are complemented by quantum

    chemical computations. This study represents the first time that 17O chemical shift and

    quadrupole coupling tensors are simultaneously determined for oxirane compounds.

    Keywords: oxirane, solid-state 17O NMR, quadrupole coupling tensor, chemical shift tensor

    Page 2 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    3

    1. Introduction

    Oxiranes (also known as 1,2-epoxides) are a class of organic compounds

    containing a three membered heterocyclic ring of two carbon atoms and one oxygen

    atom. Oxiranes are important building blocks used in organic syntheses. An example of

    their utility is in the use of ring opening mechanisms to make regioselective

    ethers/alcohols.1 Due to the strain present in the three membered ring system (containing

    a sterically unfavourable C-O-C bond angle of about 60o), oxiranes are very reactive

    electrophiles that can serve as useful reaction intermediates. Ring opening of epoxides is

    the basis for the formation of epoxy glues and glycols. Oxiranes are also biologically

    important molecules. They are present in numerous biological hormones, notably in

    many juvenoids and insect sex pheromones.2-3 An example of oxirane incorporation into

    insect sex pheromones is (+/-)-disparlure. Gypsy moth pheromone-binding proteins

    enantioselectively recognize (+)-disparlure, whereas (-)-disparlure cancels the attraction

    of (+)-disparlure. Clearly the conformation of the oxirane ring is directly linked to

    disparlure function, but the specific source of the enantioselectivity is unknown.3

    The C-O-C bond angle in oxiranes is approximately 60o, which is highly unusual,

    because it is considerably smaller than the angle between orbitals with 100% p-character

    (90o). Coulson and Moffit4 in 1949 described the bonding in three-membered rings as

    “bent”, as depicted in Scheme 1, referring to chemical bonds that contain hybrid orbitals

    whose maxima do not lie in the directions of the bonds. They used this explanation to

    justify the strain in non-linear molecules, and how it is possible to have bond angles

    smaller than 90o. They defined “strain energy” to be the difference between the observed

    heat of formation and that of a strainless reference compound. More recent calculations

    Page 3 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    4

    suggested the primary source of strain energy in oxiranes to be due to the distortion of

    electron populations within the three-membered ring as compared with open chain

    structures.5 Also related to the synthetic utility of oxiranes is that their bonding properties

    are much closer to what are observed for olefins than for typical aliphatic systems. This

    was first observed in early electron diffraction studies that found large bond shortening

    effects in three membered oxygen-containing rings.6-7 The carbon hybridization was also

    found to be closer to sp2 than sp3. A later UV-Vis spectroscopic study also found

    uncharacteristically high π-character for the oxygen atoms valence electrons in oxiranes.8

    Due to the inherent topological features of three-membered rings, there is a high level of

    electron density in the plane of the ring, which leads to a level of surface delocalization,

    strengthening the bonds within the ring. Increasing the electronegativity of the X group

    in the three-membered ring (for C2H4X), increases the electron donation to the X atom

    and decreases the back-donation to the carbon atoms, thus the oxygen atom in oxiranes is

    more nucleophilic than one would expect, which leads to its utility as a nucleophilic

    attack and coordination site.9

    ~ 60o

    Page 4 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    5

    Scheme 1. Cartoon representation depicting the formation of bent bonds in oxiranes.

    More recent studies have analyzed the effect that substituting electron withdrawing

    groups on oxiranes have on the structural parameters of the central oxirane ring.

    Computational and experimentally derived results suggest that increasing the electron

    withdrawing character of oxirane residues (thus removing electron density from the

    oxirane ring) lengthens the C-C bond lengths (and makes the bond less covalent), widens

    the C-O-C angles, and in general weakens the C-O bond strength (the effect is additive).

    It was also found that by adding asymmetric substituents to the oxirane ring leads to

    asymmetric local electron densities within the oxirane ring that mirror the orientation of

    the electron withdrawing ability of the substituents.10

    While 1H and 13C NMR studies of oxiranes have been widely reported,11-21 very

    limited information about 17O NMR parameters for oxiranes is available in the literature.

    In an early microwave spectroscopic study of ethylene oxide (the smallest oxirane

    molecule), Creswell and Schwendeman reported the following 17O quadrupole coupling

    (QC) tensor: zz = 12.6, yy = –7.4, and xx = –5.2 MHz (or CQ = 12.6 MHz and Q =

    0.17).22 There were also several early 17O NMR studies of oxiranes in solution.23-25

    Owing to the unusual oxygen bonding in oxiranes, several computational studies were

    aimed at calculating accurate CQ(17O) values.26-29 In addition, the 17O QC tensor in

    oxiranes was analyzed with the aid of the Townes-Dailey model.30-31 Oxiranes were also

    used to test magnetically corrected basis sets in magnetic shielding calculations.32 In the

    present work, we set out to achieve two goals. First, we investigated synthetic methods

    Page 5 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    6

    for 17O-labeling of oxiranes. In this regard, we report herein synthesis of three

    representative oxiranes: (2S*,3S*)-2,3-bis(4-nitrophenyl)-[17O]oxirane (Compound 1),

    (2S*,3R*)-2,3-bis(4-nitrophenyl)-[17O]oxirane (Compound 2), and racemic 2,2,3-

    triphenyl-[17O]oxirane (Compound 3); see Scheme 2. The second objective of this study

    was to experimentally determine 17O chemical shift (CS) and quadrupole coupling (QC)

    tensors in these oxiranes. To the best of our knowledge,33-35 no solid-state 17O NMR

    studies have ever been reported on oxirane compounds.

    1

    2

    3

    Scheme 2. Molecular structures of (2S*,3S*)-2,3-bis(4-nitrophenyl)-[17O]oxirane

    (Compound 1), (2S*,3R*)-2,3-bis(4-nitrophenyl)-[17O]oxirane (Compound 2), and [17O]-

    2,2,3-triphenyl-[17O]oxirane (Compound 3). Note that Compound 3 prepared in this study

    is a racemic mixture.

    Page 6 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    7

    2. Experimental section

    Synthesis

    All common chemicals and solvents were purchased from Sigma-Aldrich (Oakville,

    ON). Oxygen-17 enriched water (40% 17O, 40% 18O, 20% 16O) and oxygen-18 enriched

    water (97% 18O, 3% 16O) were purchased from CortecNet (Voisons-Le-Bretonneux,

    France).

    Synthesis of [17O]-(2S*,3S*)/(2S*,3R*)-2,3-bis(4-nitrophenyl)oxirane. p-Nitro-

    [17O]benzaldehyde was first prepared following a literature method.36 In particular, p-

    nitrobenzaldehyde (478 mg, 3.14 mmol) was dissolved in a minimum of dichloromethane

    and 120 µL 40% H217O (6.25 mmol). The solution was stirred vigorously for three days

    at room temperature. Upon removal of the solvent, the resultant white powders were

    further dried in vacuo. (480 mg, 100% yield). 17O NMR (67 MHz, CH2Cl2), δ = 585 ppm.

    (2S*,3S*)/(2S*,3R*)-2,3-Bis(4-nitrophenyl)-[17O]oxirane were prepared by using a

    ZnBr2-mediated reaction of p-nitro-[17O]benzaldehyde with triethyl phosphite as reported

    by Raju et al.37 Triethylphosphite (1.05 g, 6.37 mmol) was added to a mixture of p-Nitro-

    [17O]benzaldehyde (502 mg, 3.30 mmol) and zinc bromide (87 mg, 0.33 mmol) under

    nitrogen atmosphere. After the solution was stirred for three hours, it was poured over 50

    g ice and left for two hours. The organics were extracted with 2 50 mL ethyl acetate,

    then washed with 2 25 mL 20% brine. The solvent was then removed in vacuo.

    Isolation of the (2S*,3S*) isomer (or cis-isomer) was achieved by treating the resultant

    solid with 10 mL cold methanol, leaving behind the cis-isomer as a white precipitate that

    Page 7 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    8

    was collected by filtration. The filtrate was then dried in vacuo and the (2S*, 3R*)

    isomer (or trans-isomer) was isolated as a white powder using silica column

    chromatography (mobile phase 10% ethyl acetate: 90% hexanes). The trans compound

    was recrystallized from hexanes. (2S*,3S*)-2,3-bis(4-nitrophenyl)-[17O]oxirane: (120

    mg, 27% yield). 1H NMR (300 MHz, CDCl3): δ = 8.24 (d, 3JHH = 8.66 Hz, 4H), 7.54 (d,

    3JHH = 8.66 Hz, 4H), 3.98 (s, 2H) ppm. 13C NMR (75.4 MHz, CDCl3): δ = 148.38,

    143.22, 126.38, 124.07, 62.06 ppm. 17O NMR (67 MHz, CDCl3) δ = 35 ppm. (2S*,

    3R*)-2,3-bis(4-nitrophenyl)-[17O]oxirane: (70 mg, 16% yield). 1H NMR (300 MHz,

    CDCl3): δ = 8.11 (d, 3JHH = 8.66 Hz, 4H), 7.40 (d, 3JHH = 8.66 Hz, 4H), 4.55 (s, 2H) ppm.

    13C NMR (75.4 MHz, CDCl3) δ = 147.5, 140.7, 127.5, 123.4, 59.0 ppm. 17O NMR (67

    MHz, CDCl3) δ = 9 ppm. The level of 17O isotopic enrichment was determined to be

    about 9% for both compounds by mass spectrometry performed on a Micromass GCT

    (GC-EI TOF Mass Spectrometer) with + polarity. Details are provided in the Supporting

    Information.

    Synthesis of [Hydroxy(mesyloxy)iodo]benzene. (Diacetoxy)iodobenzene (1.950 g,

    6.0931 mmol) was suspended in 11.25 mL acetonitrile, to which were added

    methanesulfonic acid (1.20 g, 12.5 mmol) and water (225 mg, 12.5 mmol) in 2.5 mL

    acetonitrile. The mixture was stirred overnight. The off white powder was then filtered,

    washed with acetone and ether, and dried in vacuo. (1.591 g, 82% yield) 1H NMR (300

    MHz, DMSO-d6): δ = 9.75 (s, 1H), 8.22 (d, 3JHH = 8.20 Hz, 2H), 7.72 (t, 3JHH = 7.50 Hz,

    1H), 7.612 (dd, 3JHH = 8.20, 7.50 Hz, 2H), 2.29 (s, 3H) ppm. 13C NMR (75.4 MHz,

    DMSO-d6): δ = 137.57, 131.15, 128.18, 39.96 ppm.

    Page 8 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    9

    Synthesis of [18O]-Iodosylbenzene. Because 18O-labeled water is considerably

    cheaper than 17O-labeled water, we first tested the synthesis with 18O-labeled water.

    [Hydroxyl(mesyloxy)iodo]benzene (350 mg, 1.0937 mmol) was dissolved in H218O (500

    mg, 25 mmol) contained in a centrifuge tube. Sodium hydroxide (50 mg, 1.25 mmol) was

    added and the tube was shaken and centrifuged. The liquid layer was removed and the

    powder was washed repeatedly with ether and then water (206 mg, 84% yield). FTIR

    (powder): 3049 s, 1566 m, 1434 s, 734 s, 689 s, 565 w, 485 m, 424 m cm-1. The

    observation of IR bands at 565 and 424 cm-1 confirmed the > 90% 18O isotopic

    enrichment in iodosylbenzene.

    Synthesis of [17O]-Iodosylbenzene. [Hydroxyl(mesyloxy)iodo]benzene (350 mg,

    1.0937 mmol) was dissolved in H217O (500 mg, 26.3 mmol) contained in a centrifuge

    tube. Sodium hydroxide (50 mg, 1.25 mmol) was added and the tube was shaken and

    centrifuged. The liquid layer was removed and the powder was washed repeatedly with

    ether and then water. (153 mg, 63% yield). FTIR (powder): 3049 s, 1566 m, 1434 s, 733

    s, 689 s, 587 w, 487 m, 436 m cm-1.

    Synthesis of 2,2,3-Triphenyl-[17O]oxirane: Triphenylethylene (200 mg, 0.780

    mmol) was combined with Jacobsen’s catalyst (50 mg, 0.078 mmol).38 [17O]-

    iodosylbenzene (172 mg, 0.778 mmol) was added over two minutes and the solution was

    refluxed for 7 days. 17O-iodosylbenzene (83 mg, 0.38 mmol) was then added and the

    mixture was stirred for three days further. The solvent was removed in vacuo and the

    crude product was isolated using a silica column (mobile phase 10% ethyl acetate: 90%

    hexanes). The remaining impurities were dissolved using a minimum of hexanes. (104.8

    mg, 49% yield). 1H NMR (300 MHz, CDCl3), δ = 7.37 (m, 6H), 7.24 (m, 3H), 7.18 (m,

    Page 9 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    10

    3H), 7.08 (m, 3H), 4.37 (s, 1H) ppm. 13C NMR (75.4 MHz, CDCl3), δ = 141.03, 135.83,

    135.45, 127.87, 127.82, 127.75, 127.68, 127.60, 127.53, 126.79, 126.37, 68.70, 68.09

    ppm. 17O NMR (67 MHz, CDCl3) δ = 47 ppm.

    X-ray crystallography

    Among the three oxirane compounds investigated in this study, only Compound 1 has

    its crystal structure reported in the literature.37 We were able to obtain a single crystal of

    Compound 2 and determine its crystal structure. Attempts to obtain a single crystal of

    Compound 3 were unsuccessful. A colorless, rod-like crystal of Compound 2 having the

    approximate dimensions of 0.125 0.118 0.075 mm, coated with oil (Paratone 8277,

    Exxon), was collected onto the aperture of a mounted MicromountTM (diameter of the

    aperture: 100 microns; MiTeGen - Microtechnologies for Structural Genomics; USA) and

    quickly transferred to the cold nitrogen gas stream of the Oxford Cryostream 800

    operating at 93.16 °C. The mounted MicromountTM had previously been inserted into a

    reusable magnetic goniometer base (B3S-R, MiTeGen - Microtechnologies for Structural

    Genomics; USA). All measurements were made on a Bruker AXS D8 Venture Duo

    diffractometer using Mo K radiation ( = 0.71073 Å) generated by a high brilliance

    Incoatec Is microfocus tube equipped with a HELIOS multilayer mirror optics (power:

    50 kV 1 mA). Data were recorded with a Bruker AXS PHOTON II Charge-Integrating

    Pixel Array Detector (CPAD) (frame size: 768 × 1024). The structure was solved using

    direct methods and refined by full-matrix least-squares method on F2 with SHELXL-

    2014 using ShelXle as the graphical user interface.39 Hydrogen atoms of the CH groups

    were included at geometrically idealized positions (C-H bond distances: 0.95 Å) and

    Page 10 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    11

    were not refined. The isotropic thermal parameters of these hydrogen atoms were fixed

    at 1.2 times (CH-groups) that of the preceding carbon atom. The hydrogen atoms

    attached to the carbon atoms labeled as C(7) and C(8) were located in the difference

    Fourier map. The coordinates and isotropic parameters for these hydrogen atoms were

    refined. Data collection and refinement conditions are listed in Table 1. Other detailed

    crystallographic data are given in the Supporting Information.

    Solid-state 17O NMR

    Solid-state 17O NMR spectra were recorded at 14.1 and 21.1 T, operating at the 17O

    Larmor frequencies of 81.38 and 122.02 MHz, respectively. Experiments performed at

    14.1 T utilized a 4 mm Bruker HX probe and samples were packed into 4 mm o.d.

    zirconia rotors. On this probe, the B1 field at the 17O Larmor frequency was about 50

    kHz. All experiments at 21.1 T were performed at the National Ultrahigh-Field NMR

    Facility for Solids (Ottawa, Ontario, Canada) using a home-built solenoid 5 mm HX

    static probe for static samples and a 3.2 mm MAS Bruker HX probe for magic angle

    spinning (MAS). On both probes, the B1 field at the 17O Larmor frequency was

    approximately 42 kHz. For the static experiments, a 5-mm Teflon tube was used as

    sample holder. In the MAS experiments, the sample spinning frequency was 22 kHz. In

    the 17O MAS experiment for Compound 3, the sample was cooled to 5 oC using a Bruker

    BCU05 cooling unit while its MAS spectrum was recorded. NMR spectra were analysed

    using Bruker Topspin 2.0. Solid-state 17O NMR spectra were fitted using DMFit.40

    Quantum chemical computations

    Page 11 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    12

    All property calculations were performed on molecules that were either optimized

    using the same method/basis set combination as was used for the specific property

    calculation, or were performed on original crystal structures. All quantum chemical

    calculations were carried out using Gaussian 09.41 The licensing for the software

    packages is provided by the Center for Advanced Computing (CAC) (Queen’s

    University, Kingston, Ontario, Canada). Calculations are submitted directly to the

    Frontenac Cluster. Typically, 12 processors were used for each calculation. All

    optimization and NMR calculations were performed for gas phase molecules. Structural

    optimizations and NMR calculations were performed using B3LYP/6-311++G(3df,3pd)

    and the GIAO method. Structural optimizations and NMR calculations for ethylene oxide

    and dimethyl ether were performed using B3LYP/6-311G(d,p) and the GIAO method.

    3. Results and discussion

    Crystal structure of Compound 2

    In this section, we describe the new crystal structure of Compound 2. Figure 1 shows

    the molecular structure of Compound 2. The geometrical parameters around the central

    three-membered ring are: rC7-O1 = 1.435, rC8-O1 = 1.435, rC7-C8 = 1.480 Å, C7-O1-C8 =

    62.08. These are very similar to those found in the crystal structure of the cis isomer:37

    rC7-O1 = 1.432, rC8-O1 = 1.441, rC7-C8 = 1.471 Å, C7-O1-C8 = 61.60. It is also

    interesting to note that the two crystal structures are also similar to the geometry of the

    parent compound, ethylene oxide (Compound 4), which was determined by microwave

    spectroscopy in the gas phase: rCO = 1.434, rCC = 1.470 Å, ∠COC = 61.67o.42

    Page 12 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    13

    NMR spectral analysis

    Figure 2 shows the solid-state 17O NMR spectra of Compounds 1, 2, and 3 recorded at

    two magnetic fields, 14.1 and 21.1 T, under both MAS and static conditions. In general,

    the major features in the static 17O NMR spectra for these oxirane compounds are typical

    of those arising from the second-order quadrupole interactions.33 This observation

    immediately suggests that the 17O chemical shift anisotropy is rather small for these

    compounds. At 21.1 T, as the second-order quadrupole broadening is reduced, we were

    able to obtain high-quality 17O MAS NMR spectra for these compounds. As seen from

    Figure 2, the experimental 17O MAS NMR spectra display characteristic line shapes

    suggesting the presence of relatively large CQ but rather small ηQ values. We should

    point out that the spectral quality for Compound 3 is rather low because of the low 17O

    enrichment level of the product. Nonetheless, all solid-state 17O NMR spectra obtained

    for Compounds 1-3 can be properly simulated and the final results for the 17O QC and CS

    tensors are listed in Table 2. Some geometric parameters are also given in Table 2.

    Among the three oxirane compounds studied (Compounds 1-3), it appears that changing

    the regiochemistry and the nature of the substituents on the oxirane ring does not have a

    large effect on its structural features. We also found that Compounds 1-3 display very

    similar 17O QC and CS tensors. For example, the CQ(17O) values in these oxirane

    compounds are about 12-13 MHz and the asymmetry parameters are around 0.2. In

    addition, the 17O CS tensors have rather small span ( = 11 – 33 ≈ 200-250 ppm). Table

    2 also lists computational results for the 17O QC and CS tensors for Compounds 1-3. In

    general, the agreement between the experimental and computed 17O NMR results is

    reasonably good.

    Page 13 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    14

    Now we can compare the observed 17O QC and CS tensor data in 1-3 with those

    previously reported for the R-O-R (R, R = H or C) type of oxygen atoms. The CQ(17O)

    values in oxiranes (≈ 13 MHz) are found to be considerably larger than those in water (H-

    O-H, CQ ≈ 7-10 MHz),43-49 hydroxyl groups (C-O-H, CQ ≈ 9-10 MHz),50-55 and ethers (C-

    O-C, CQ ≈ 10-12 MHz),56-57 but still smaller than those found in the H-O-X type (X = N

    and O) of bonding (for example, the hydroxylammonium cation, HONH3+, has

    CQ(17O) = 14.7 MHz;58 hydrogen peroxide, HOOH, has CQ(17O) = 16.3 MHz).59

    Another distinct feature of the 17O QC tensors in 1-3 is that the ηQ value is rather small,

    ηQ ≈ 0.2. In comparison, all the compounds mentioned above have ηQ > 0.6. We will

    further discuss this difference in a later section. In the literature, there are fewer studies

    where 17O CS tensors are reported for the R-O-R (R, R = H or C) type of oxygen atoms.

    For the water molecules in crystalline hydrates, the spans of the 17O CS tensor, , are

    typically less than 80 ppm.47, 49 Similarly, the value of (17O) in the hydronium ion,

    H3O+, is also small, 87 ppm.60 For the C-O-H groups in phenol and hemiketal

    compounds, (17O) is typically on the order of 70-90 ppm,52, 54 which is still smaller than

    those found in Compounds 1-3. It is interesting to note that the 17O NMR tensor

    parameters in 1-3 are somewhat similar to those reported for the ether type O atom in [5-

    17O]-D-glucose.61 In general, compounds with high CQ but small values would benefit

    the most by going to ultrahigh magnetic fields such as 35.2 T.62

    One of the advantages of the solid-state 17O NMR method employed in this study is

    that one can obtain information about tensor orientation in the molecular frame of

    reference. Here we will further examine the 17O QC and CS tensor orientations in

    oxiranes. In particular, we are interested in comparing the tensor orientations in the

    Page 14 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    15

    highly strained geometry of oxiranes (with a C-O-C angle of 60) with those in a model

    ether, dimethyl ether (5), where the C-O-C angle is about 113. The computed 17O NMR

    tensors for dimethyl ether are also listed in Table 1. As seen from Figure 3, for oxiranes,

    the least shielded component, 11, of the 17O CS tensor lies in the molecular plane

    bisecting the C-O-C angle and the most shielded component, 33, is perpendicular to the

    oxirane ring. In contrast, for dimethyl ether, 11 lies in the plane but along the tangent

    direction of the C-O-C angle and 22 is perpendicular to the C-O-C plane. Because the

    17O chemical shift anisotropies are rather small in both oxiranes and dimethyl ether, it is

    not possible to identify predominant molecular orbital contributions to individual tensor

    components. For this reason, we will not further discuss them. As seen from Figure 3,

    the 17O QC tensors of oxirane and dimethyl ether share a common feature in that both

    tensors have their largest components, zz, aligning perpendicular to the molecular plane.

    Computations also confirm that the sign of CQ(17O) is positive in both cases; see Table 2.

    However, one key difference between the 17O QC tensors in these two systems is the

    interchange of orientations of xx and yy. As seen in Figure 3, in oxiranes, yy bisects

    the C-O-C angle whereas in dimethyl ether it is xx. We note that the 17O QC tensor

    orientation in dimethyl ether is the same as that observed for the water molecule.43 The

    difference in the 17O QC tensor orientations shown in Figure 3 between dimethyl ether

    and ethylene oxide is also related to the fact that oxirane compounds often exhibit ηQ ≈ 0

    but dimethyl ether has ηQ ≈ 1. The distinct 17O QC tensor orientation in oxiranes (in

    particular, ethylene oxide) was also noted in earlier studies by Gready.63-64 In the

    following section, we will provide an explanation as to the origin of the different 17O QC

    tensors in oxirane and dimethyl ether.

    Page 15 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    16

    On the basis of the Townes-Dailey model,65 it is well-established that, for oxygen

    atoms in the R-O-R type of bonding geometry, ηQ of the 17O QC tensor depends on the

    bond angle, , in the following fashion:66-68

    [1] 𝜂𝑄 =3(1 + 𝑐𝑜𝑠𝜃)1 ― 3𝑐𝑜𝑠𝜃 for 109.47° ≤ θ ≤ 180°

    [2] 𝜂𝑄 = ―3𝑐𝑜𝑠𝜃 for 90° ≤ θ ≤ 109.47°

    Using the above equations, one predicts that ηQ is 0.84 for dimethyl ether ( = 113),

    which is in good agreement with the ab initio computational results shown in Table 2.

    For ethylene oxide, although the above equation is not strictly valid (as the three

    membered ring has < 90), the observed small ηQ value does resemble the extreme case

    of 90 (ηQ = 0). Of course, for the “bent” bonds, the hybrid atomic orbitals are no longer

    aligned with the bond directions. To further understand the 17O QC tensors observed for

    dimethyl ether and ethylene oxide, it is important to analyze the electron lone-pairs

    around the oxygen atom under study. The results from a Natural Bonding Orbital (NBO)

    analysis are shown in Figure 4. The oxygen atom in each compound has two electron

    lone-pairs. The common feature between the two compounds is that they both have one

    electron lone-pair, LP(2), with an essentially full occupancy (1.92 electrons) in the pure

    pz atomic orbital pointing in the direction perpendicular to the molecular plane. Because

    this direction clearly has an excess of valence p-orbital populations, it corresponds to the

    direction along which the largest 17O QC tensor component lies, as we explained in a

    recent study using the concept of valence p-orbital population anisotropy (VPPA).31 The

    key difference between the two compounds, however, lies in the configuration of LP(1).

    As seen from Figure 4, while dimethyl ether has its LP(1) in an sp1.39 hybrid orbital (42%

    2s and 58% 2p) pointing along the y-axis, the LP(1) of ethylene oxide has much less p

    Page 16 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    17

    characters (an sp0.39 hybrid orbital has 72% 2s and 28% 2p). Because of the lack of p-

    character in LP(1) of ethylene oxide, the valence p-orbital populations in the px and py

    orbitals are primarily from those making the C-O bonds, which are also mainly pure p

    atomic orbitals (the hybrid orbital on the oxygen atom to make the C-O bond has 23% 2s

    and 87% 2p). As a result, the valence p-orbital populations in the px and py orbitals are

    essentially the same, which leads an approximately axially symmetric 17O QC tensor in

    ethylene oxide (i.e., ηQ 0). For dimethyl ether, on the other hand, because its LP(1) has

    a considerable amount of p-character, the valence p-orbital population in py is

    significantly greater than that in px, resulting in a high ηQ value; see data in Table 2.

    4. Conclusions

    In this study, we have reported synthesis of three 17O-labeled oxirane compounds.

    Using solid-state 17O NMR, we were able to measure the 17O QC and CS tensors in these

    compounds. During the course of this study, we have also determined the crystal

    structure for (2S*,3R*)-2,3-bis(4-nitrophenyl)oxirane (Compound 2). The 17O QC tensor

    in an oxirane molecule typically displays CQ = 12-13 MHz and ηQ = 0.1-0.2, whereas the

    17O CS tensor has a relatively small anisotropy ( = 200-250 ppm). We found that the

    oxirane structural parameters and the 17O NMR tensors are similar in all the three oxirane

    compounds examined in this study, despite the differences in regiochemistry and

    substituents on the oxirane ring. In general, oxirane compounds show similar 17O NMR

    parameters as those of the R-O-R (R, R = H and C) type of functional groups. However,

    when compared with dimethyl ether, oxiranes display distinct 17O QC and CS tensor

    orientations as a result from the constrained C-O-C bond angle. In addition, dimethyl

    Page 17 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    18

    ether and oxirane also represent two extreme cases of the 17O QC tensor in terms of their

    asymmetry parameters (dimethyl ether has ηQ 1, but oxiranes have ηQ 0). In the field

    of solid-state 17O NMR for organic and biological molecules, there are continuing efforts

    to synthesize new 17O-labeled functional groups and to characterize 17O NMR tensors.

    These studies will ultimately lead to understanding of the relationship between 17O NMR

    tensors and molecular structure/chemical bonding.

    Acknowledgement

    This work was supported by the Natural Sciences and Engineering Research Council

    (NSERC) of Canada. A.R. thanks the Government of Ontario for an Ontario Graduate

    Scholarship. Access to the 900 MHz NMR spectrometer was provided by the National

    Ultrahigh Field NMR Facility for Solids (Ottawa, Canada), a national research facility

    funded by a consortium of Canadian universities, National Research Council Canada and

    Bruker BioSpin and managed by the University of Ottawa (http://nmr900.ca).

    Supporting Information Available. MS data analysis for Compounds 1 and 2. Detailed

    crystallographic data for Compound 2. A crystallographic information file (CIF) for

    Compound 2 has been deposited to the Cambridge Crystallographic Data Centre (CCDC-

    1997472).

    Page 18 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

    http://nmr900.ca

  • Draft

    19

    Tables and Figure Captions

    Table 1. Crystallographic data for (2S*, 3R*)-2,3-bis(4-nitrophenyl)oxirane (Compound 2).

    Empirical formula C14H10N2O5Formula weight 286.24Crystal color, habit colorless, rod-likeCrystal dimensions (mm) 0.125 0.118 0.071Crystal system triclinicSpace group P1Unit cell parameters

    a (Å) 7.2465(3)7.9199(4)b (Å)

    c (Å) 11.4449(5)α (°) 77.367(2)β (°) 80.038(2)γ (°) 81.173(2)V (Å3) 630.73(5)Z 2

    F(000) 380Density (ρcalcd) 1.507 Mg/m3

    Absorption coefficient (μ) 0.117 mm-1

    Page 19 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    20

    Table 2. Experimental and computed (G09) 17O NMR parameters and selected structural parameters for Compounds 1, 2, and 3, ethylene oxide (4), and dimethyl ether (5). The uncertainties in the experimental 17O CS tensor components were estimated to be 10 ppm by visual inspection of the agreement between experimental and simulated spectra.

    Compd δiso (ppm)

    δ11 (ppm)

    δ22 (ppm)

    δ33 (ppm)

    CQ (MHz)

    ηQ ∠COC (o)

    rCO (Å)

    1 Exp. 30 2 160 –15 –55 12.1 0.2 0.18 0.10 61.60a 1.436aCal. 30 193 –33 –70 13.4 0.22 61.59b 1.432b

    2 Exp. 30 2 180 –33 –70 13.4 0.2 0.22 0.10 62.09b 1.435bCal. 31 171 4 –81 13.3 0.22 62.97b 1.428b

    3 Exp. 43 2 173 –9 –35 12.8 0.2 0.20 0.10 Cal. 42 183 –13 –44 12.9 0.17 63.14b 1.424b

    4 Exp. 12.6c 0.17c 61.67d 1.434dCal. –11 176 –52 –157 14.1 0.28 59.46b 1.490b

    5 Cal. –16 7 –5 –49 12.3 0.90 113.0b 1.450baFrom ref. 37. bThis work. cFrom ref. 22. dFrom ref. 42.

    Page 20 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    21

    Figure 1. Crystal structure of Compound 2. Thermal ellipsoids are shown at the 30% probability level.

    Page 21 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    22

    * *

    δ(17O)/ppm-10000

    δ(17O)/ppm

    -1000-50005001000

    (a)

    0500 -500 -10001000

    δ(17O)/ppm

    * *

    1000 500 -500

    (b) (c)

    MAS at 21.1 T

    Static at 21.1 T

    Static at 14.1 T

    *

    Figure 2. Experimental (blue trace) and simulated (red trace) 17O solid-state NMR spectra of (a) 1, (b) 2, and (c) 3. The peak marked by * was due to the ZrO2 rotor used in MAS experiments. The sample spinning frequency was 22 kHz. All spectra were recorded with a Hahn-echo pulse sequence with central-transition selective pulses and a recycle delay of 10 s. Other data acquisition parameters are given below. In (a), 8000 transients for the MAS at 21.1 T; 13000 transients for the static experiment at 21.1 T; 41432 transients in the static experiment at 14.1 T. In (b), 24000 transients for the MAS at 21.1 T; 14000 transients for the static experiment at 21.1 T; 17262 transients in the static experiment at 14.1 T. In (c), 33000 transients for the MAS at 21.1 T; 16000 transients for the static experiment at 21.1 T; 21960 transients in the static experiment at 14.1 T.

    Page 22 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    23

    δ11

    δ22 χxx

    χyy(a) (b)

    (c) (d)

    113° 113°

    δ11

    δ33

    χyy

    χxx

    60° 60°

    Figure 3. Depiction of the 17O CS (a and c) and QC (b and d) tensor orientations in the molecular frames of reference of oxirane (a and b) and dimethyl ether (c and d). Tensor components that are perpendicular to the paper plane are not shown for clarity.

    Page 23 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    24

    (a) (b)

    O

    y

    x

    LP(1): spy0.39

    LP(2): pz

    O

    y

    x

    LP(1): spy1.39

    LP(2): pz

    113° 60°

    Figure 4. NBO results on electron lone-pairs in (a) dimethyl ether and (b) ethylene oxide.

    TOC graphics

    17O NMRO

    C C

    References

    1. Chitwood, H. C.; Freure, B. T. The Reaction of Propylene Oxide with Alcohols. J. Am. Chem. Soc. 1946, 68, 680-683.2. Mumby, S. M.; Hammock, B. D. Stability of Epoxide-Containing Juvenoids to Dilute Aqueous Acid. J. Argric. Food. Chem. 1979, 27, 1223-1228.3. Sanes, J. T.; Plettner, E. Gypsy moth pheromone-binding protein-ligand interactions: pH profiles and simulations as tools for detecting polar interactions. Arch. Biochem. Biophys. 2016, 606, 53-63.4. Coulson, C. A.; Moffitt, W. E. I. The properties of certain strained hydrocarbons. Philos. Mag. 1949, 40, 1-35.5. Vila, A.; Mosquera, R. A. AIM interpretation of strain energy of oxiranes. Chem. Phys. 2003, 287, 125-135.6. Donohue, J.; Humphrey, G. L.; Schomaker, V. The Structure of Spiropentane. J. Am. Chem. Soc. 1945, 67, 332-335.7. Bastiansen, O.; Hassel, O. The Molecular Structure of the So-called "cis" Decalin. Nature 1946, 157, 765-765.8. Walsh, A. D. The structures of ethylene oxide, cyclopropane, and related molecules. Trans. Faraday Soc. 1949, 45, 179-190.

    Page 24 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    25

    9. Cremer, D.; Kraka, E. Theoretical Determination of Molecular Structure and Conformation. 15. Three-Membered Rings: Bent Bonds, Ring Strain, and Surface Delocalization. J. Am. Chem. Soc. 1985, 107, 3800-3810.10. Grabowsky, S.; Schirmeister, T.; Paulmann, C.; Pfeuffer, T.; Luger, P. Effect of electron-withdrawing substituents on the epoxide ring: An experimental and theoretical electron density analysis of a series of epoxide derivatives. J. Org. Chem. 2011, 76, 1305-1318.11. Allen, G.; Blears, D. J.; Webb, K. H. Nuclear magnetic resonance spectra of some substituted oxiranes. J. Chem. Soc. 1965, 810-11.12. Lyle, G. G.; Keefer, L. K. The configurations at C-9 of the cinchona alkaloids. N.M.R. spectral study of the derived oxiranes. Tetrahedron 1967, 23, 3253-63.13. Villa, L.; Schenetti, L.; Taddei, F. Proton NMR spectra of oxiranes. cis-Arylmethyloxiranes substituted in the phenyl ring. Org. Magn. Reson. 1973, 5, 593-4.14. Paulson, D. R.; Tang, F. Y. N.; Moran, G. F.; Murray, A. S.; Pelka, B. P.; Vasquez, E. M. Carbon-13 nuclear magnetic resonance spectroscopy. Quantitative correlations of the carbon chemical shifts of simple epoxides. J. Org. Chem. 1975, 40, 184-186.15. Kleinpeter, E.; Kuhn, H.; Muhlstada, M. NMR-Untersuchungen an dicyclopentadienederivaten. Org. Magn. Reson. 1977, 9, 312-317.16. Kingsbury, C. A.; Durham, D. L.; Hutton, R. 3JCH coupling constants in oxiranes, thiiranes, and cyclopropanes. J. Org. Chem. 1978, 43, 4696-4700.17. Lesage, S.; Perlin, A. S. Carbon-13 nuclear magnetic resonance spectra of oxiranes. Configuration of 1,2-epoxypropyl side chains in fungal metabolites. Can. J. Chem. 1978, 56, 3117-20.18. Christol, H.; Laffite, C.; Plenat, F.; Renard, G. 1H and 13C NMR study of the effects exerted by an oxiarne ring in the epoxybicyclo[2.2.2]octane series. Org. Magn. Reson. 1981, 17, 110-117.19. Matoba, Y.; Kagayama, T.; Ishii, Y.; Ogawa, M. Carbon-13 NMR spectra of some epoxides derived from dicyclopentadiene. Org. Magn. Reson. 1981, 17, 144-147.20. Benner, S. A.; Maggio, J. E.; Simmons, H. E. Rearrangement of a geometrically restricted triepoxide to the first topologically nonplanar molecule: A reaction path elucidated by using oxygen isotope effects on carbon-13 chemical shifts. J. Am. Chem. Soc. 1981, 103, 1581-1582.21. Krivdin, L. B.; Sauer, S. P. A.; Peralta, J. E.; Contreras, R. H. Non-empirical calculations of NMR indirect carbon–carbon coupling constants: 1. Three-membered rings. Magn. Reson. Chem. 2002, 40, 187-194.22. Creswell, R. A.; Schwendeman, R. H. Centrifugal Distortion and Oxygen-17 Quadrupole Coupling in Ethylene Oxide. Chem. Phys. Lett. 1974, 27, 521-524.23. Iwamura, H.; Sugawara, T.; Kawada, Y.; Tori, K.; Muneyuki, R.; Noyori, R. 17O NMR Chemical Shifts Versus Structure Relationships in Oxiranes. Tetrahedron Letters 1979, 20, 3449-3452.24. Sauleau, A.; Sauleau, J.; Monti, J. P.; Faure, R. Oxygen-17 Nuclear Magnetic Resonance Study of Some Oxirane Derivatives. Org. Magn. Reson. 1983, 21, 403-404.25. Monti, J. P.; Faure, R.; Sauleau, A.; Sauleau, J. 13C and 17O NMR of some substituted oxiranes. Chemical shifts and quantitative correlations. Magn. Reson. Chem. 1986, 24, 15-20.

    Page 25 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    26

    26. Ludwig, R.; Weinhold, F.; Farrar, T. C. Effective O-17 quadrupole moments for the calibrated computation of quadrupole coupling parameters at different levels of theory. J. Chem. Phys. 1996, 105, 8223-8230.27. Eggenberger, R.; Gerber, S.; Huber, H.; Searles, D. Calculations of 17O Nuclear Quadrupole Coupling Constants. J. Mol. Spectrosc. 1992, 151, 474-481.28. Bailey, W. C. B3LYP Calculation of 17O Quadrupole Coupling Constants in Molecules. Chem. Phys. Lett. 1998, 292, 71-74.29. De Luca, G.; Russo, N.; Koster, A. M.; Calaminici, P.; Jug, K. Density functional theory calculations of nuclear quadrupole coupling constants with calibrated 17O quadrupole moments. Mol. Phys. 1999, 97, 347-354.30. Balakina, M. Y.; Fominykh, O. D.; Morozova, I. D.; Osokin, D. Y. The Townes and Dailey approach to the interpretation of NQR parameters in the light of non-empirical calculations. Three-membered rings. Magn. Reson. Chem. 1995, 33, 761-766.31. Rinald, A.; Wu, G. A Modified Townes-Dailey Model for Interpretation and Visualization of Nuclear Quadrupole Coupling Tensors in Molecules. J. Phys. Chem. A 2020, 124, 1176-1186.32. Rossikhin, V. V.; Okovytyy, S. I.; Kasyan, L. I.; Voronkov, E. O.; Umrikhina, L. K.; Leszczynski, J. An investigation of the 17O NMR chemical shifts in oxiranes using magnetically corrected basis sets. J. Phys. Chem. A 2002, 106, 4176-4180.33. Wu, G. Solid-state 17O NMR studies of organic and biological molecules. Prog. Nucl. Magn. Reson. Spectrosc. 2008, 52, 118-169.34. Wu, G. Solid-State 17O NMR studies of organic and biological molecules: Recent advances and future directions. Solid State Nucl. Magn. Reson. 2016, 73, 1-14.35. Wu, G. 17O NMR studies of organic and biological molecules in aqueous solution and in the solid state. Prog. Nucl. Magn. Reson. Spectrosc. 2019, 114-115, 135-191.36. Wu, G.; Mason, P.; Mo, X.; Terskikh, V. Experimental and computational characterization of the 17O quadrupole coupling and magnetic shielding tensors for p-nitrobenzaldehyde and formaldehyde. J. Phys. Chem. A 2008, 112, 1024-32.37. Raju, P.; Gobi Rajeshwaran, G.; Nandakumar, M.; Mohanakrishnan, A. K. Unusual Reactivity of Aryl Aldehydes with Triethyl Phosphite and Zinc Bromide: A Facile Preparation of Epoxides, Benzisoxazoles, and α-Hydroxy Phosphonate Esters. Eur. J. Org. Chem. 2015, 2015, 3513-3523.38. Adam, W.; Humpf, H.-U.; Roschmann, K. J.; Saha-mller, C. R.; Saha-mo, C. R. Enantioselective Epoxidation with Chiral Mn (salen) Catalysts : Kinetic Resolution of Aryl-Substituted Allylic Alcohols Enantioselective Epoxidation with Chiral Mn III ( salen ) Catalysts : Kinetic Resolution of Aryl-Substituted Allylic Alcohols. J. Org. Chem. 2001, 5796-5800.39. Sheldrick, G. M. SHELXT-2014, Program for the Solution of Crystal Structures. Acta Crystallogr. 2015, A71, 3-8.40. Massiot, D.; Fayon, F.; Capron, M.; King, I.; Le Calvé, S.; Alonso, B.; Durand, J.-O.; Bujoli, B.; Gan, Z.; Hoatson, G. Modelling one-and two-dimensional solid-state NMR spectra. Magn. Reson. Chem. 2002, 40, 70-76.41. Frisch, M. J., et al. Gaussian 09, Revision A.02, Gaussian, Inc., Wallingford, CT, 2016.42. Hirose, C. Microwave Spectra and r0, rs, and rm Structures of Ethylene Oxide. Bull. Chem. Soc. Jpn. 1974, 47, 1311-1318.

    Page 26 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    27

    43. Verhoeven, J.; Dymanus, A.; Bluyssen, H. Hyperfine Structure of HD17O by Beam‐Maser Spectroscopy. J. Chem. Phys. 1969, 50, 3330-3338.44. Spiess, H. W.; Garrett, B. B.; Sheline, R. K.; Rabideau, S. W. Oxygen-17 quadrupole coupling parameters for water in its various phases. J. Chem. Phys. 1969, 51, 1201-1205.45. Zhang, Q. W.; Zhang, H. M.; Usha, M. G.; Wittebort, R. J. 17O NMR and crystalline hydrates. Solid State Nucl. Magn. Reson. 1996, 7, 147-154.46. Wu, G.; Rovnyak, D.; Huang, P. C.; Griffin, R. G. High-Resolution Oxygen-17 NMR Spectroscopy of Solids by Multiple-Quantum Magic-Angle Spinning. Chem. Phys. Lett. 1997, 277, 79-83.47. Michaelis, V. K.; Keeler, E. G.; Ong, T.-C.; Craigen, K. N.; Penzel, S.; Wren, J. E. C.; Kroeker, S.; Griffin, R. G. Structural Insights into Bound Water in Crystalline Amino Acids: Experimental and Theoretical 17O NMR. J. Phys. Chem. B 2015, 119, 8024-8036.48. Keeler, E. G.; Michaelis, V. K.; Griffin, R. G. 17O NMR Investigation of Water Structure and Dynamics. J. Phys. Chem. B 2016, 120, 7851-7858.49. Nour, S.; Widdifield, C. M.; Kobera, L.; Burgess, K. M. N.; Errulat, D.; Terskikh, V. V.; Bryce, D. L. Oxygen-17 NMR spectroscopy of water molecules in solid hydrates. Can. J. Chem. 2016, 94, 189-197.50. Sefzik, T. H.; Houseknecht, J. B.; Clark, T. M.; Prasad, S.; Lowary, T. L.; Gan, Z.; Grandinetti, P. J. Solid-state 17O NMR in carbohydrates. Chem. Phys. Lett. 2007, 434, 312-315.51. Zhu, J.; Geris, A. J.; Wu, G. Solid-state 17O NMR as a sensitive probe of keto and gem-diol forms of [small alpha]-keto acid derivatives. Phys. Chem. Chem. Phys. 2009, 11, 6972-6980.52. Zhu, J.; Lau, J. Y. C.; Wu, G. A solid-state 17O NMR study of L-tyrosine in different ionization states: implications for probing tyrosine side chains in proteins. J. Phys. Chem. B 2010, 114, 11681-11688.53. Kong, X.; Shan, M.; Terskikh, V.; Hung, I.; Gan, Z.; Wu, G. Solid-State 17O NMR of Pharmaceutical Compounds: Salicylic Acid and Aspirin. J. Phys. Chem. B 2013, 117, 9643-9654.54. Kong, X.; Dai, Y.; Wu, G. Solid-state 17O NMR study of 2-acylbenzoic acids and warfarin. Solid State Nucl. Magn. Reson. 2017, 84, 59-64.55. Rees, G. J.; Day, S. P.; Barnsley, K. E.; Iuga, D.; Yates, J. R.; Wallis, J. D.; Hanna, J. V. Measuring multiple 17O–13C J-couplings in naphthalaldehydic acid: a combined solid state NMR and density functional theory approach. Phys. Chem. Chem. Phys. 2020, 22, 3400-3413.56. Hsieh, Y.; Koo, J. C.; Hahn, E. L. Pure nuclear quadrupole resonance of naturally abundant oxygen-17 in organic solids. Chem. Phys. Lett. 1972, 13, 563-6.57. Butler, L. G.; Cheng, C. P.; Brown, T. L. Oxygen-17 Nuclear Quadrupole Double Resonance. 6. Effects of Hydrogen Bonding. J. Phys. Chem 1981, 85, 2738-2740.58. Lu, J.; Kong, X.; Terskikh, V.; Wu, G. Solid-State 17O NMR of Oxygen−Nitrogen Singly Bonded Compounds: Hydroxylammonium Chloride and Sodium Trioxodinitrate (Angeli's Salt). J. Phys. Chem. A 2015, 119, 8133-8138.59. Lumpkin, O.; Dixon, W. T. Oxygen-17 pure quadrupole resonances in hydrogen peroxide. J. Chem. Phys. 1979, 71, 3550-3551.

    Page 27 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    28

    60. Wu, G.; Hook, A.; Dong, S.; Yamada, K. A Solid-State NMR and Theoretical Study of the 17O Electric Field Gradient and Chemical Shielding Tensors of the Oxonium Ion in p-Toluenesulfonic Acid Monohydrate. J. Phys. Chem. A 2000, 104, 4102-4107.61. Shen, J.; Terskikh, V.; Wu, G. Observation of the Second-Order Quadrupolar Interaction as a Dominating NMR Relaxation Mechanism in Liquids: The Ultraslow Regime of Motion. J. Phys. Chem. Lett. 2016, 7, 3412-3418.62. Gan, Z.; Hung, I.; Wang, X.; Paulino, J.; Wu, G.; Litvak, I. M.; Gor'kov, P. L.; Brey, W. W.; Lendi, P.; Schiano, J. L.; Bird, M. D.; Dixon, I. R.; Toth, J.; Boebinger, G. S.; Cross, T. A. NMR spectroscopy up to 35.2 T using a series-connected hybrid magnet. J. Magn. Reson. 2017, 284, 125-136.63. Gready, J. E. The Relationship between nuclear quadrupole coupling constants and asymmetry parameter. The interplay of theory and experiment. J. Am. Chem. Soc. 1981, 103, 3682-3691.64. Gready, J. E. Theoretical study of the variability of the electric field gradient tensor of oxygen nuclei in organic molecules. J. Phys. Chem. 1984, 88, 3497-3503.65. Townes, C. H.; Dailey, B. P. Determination of Electronic Structure of Molecules from Nuclear Quadrupole Effects. J. Chem. Phys. 1949, 17, 782-796.66. Vega, S. On the asymmetry parameter and the principal directions of the electric field gradient tensor in NQR spectroscopy On the asymmetry parameter and the principal directions of the electric field gradient tensor in NQR spectroscopy. J. Chem. Phys. 1974, 60, 3884-3888.67. Poplett, I. J. F. 1H/2H and 1H/17O Nuclear Quadrupole Double-Resonance Study of Several Hydroxide Compounds. II. The Water Molecule J. Magn. Reson. 1982, 50, 397-408.68. Sternberg, U. The bond angle dependence of the asymmetry parameter of the oxygen-17 electric field gradient tensor. Solid State Nucl. Magn. Reson. 1993, 2, 181-190.

    Page 28 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft~ 60o

    Page 29 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft1

    2

    3

    Page 30 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    Page 31 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    * *

    δ(17O)/ppm -1000 0

    δ(17O)/ppm -1000 -500 0 500 1000

    (a)

    0 500 -500 -1000 1000 δ(17O)/ppm

    * *

    1000 500 -500

    (b) (c)

    MAS at 21.1 T

    Static at 21.1 T

    Static at 14.1 T

    *

    Page 32 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft

    δ11

    δ22 χxx

    χyy (a) (b)

    (c) (d)

    113° 113°

    δ11

    δ33

    χyy

    χxx

    60° 60°

    Page 33 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry

  • Draft(a) (b)

    O

    y

    x

    LP(1): spy0.39 LP(2): pz

    O

    y

    x

    LP(1): spy1.39 LP(2): pz

    113° 60°

    Page 34 of 34

    https://mc06.manuscriptcentral.com/cjc-pubs

    Canadian Journal of Chemistry