Gregory J.
Rees
a,
Stephen P.
Day
a,
Kristian E.
Barnsley
a,
Dinu
Iuga
a,
Jonathan R.
Yates
b,
John D.
Wallis
c and
John V.
Hanna
*a
aDepartment of Physics, University of Warwick, Gibbet Hill Road, Coventry, CV4 7AL, UK. E-mail: j.v.hanna@warwick.ac.uk
bDepartment of Materials, Oxford University, 16 Parks Rd, Oxford, OX1 3PH, UK
cSchool of Science and Technology, Nottingham Trent University, Clifton Lane, Nottingham, NG11 8NS, UK
First published on 24th January 2020
A combined multinuclear solid state NMR and gauge included projected augmented wave, density functional theory (GIPAW DFT) computational approach is evaluated to determine the four heteronuclear 1J(13C,17O) couplings in solid 17O enriched naphthalaldehydic acid. Direct multi-field 17O magic angle spinning (MAS), triple quantum MAS (3QMAS) and double rotation (DOR) experiments are initially utilised to evaluate the accuracy of the DFT approximations used in the calculation of the isotropic chemical shifts (δiso), quadrupole coupling constants (CQ) and asymmetry (ηQ) parameters. These combined approaches give δiso values of 313, 200 and 66 ppm for the carbonyl (CO), ether (–O–) and hydroxyl (–OH) environments, respectively, with the corresponding measured quadrupole products (PQ) being 8.2, 9.0 and 10.6 MHz. The geometry optimised DFT structure derived using the CASTEP code gives firm agreement with the shifts observed for the ether (δiso = 223, PQ = 9.4 MHz) and hydroxyl (δiso = 62, PQ = 10.5 MHz) environments but the unoptimised experimental XRD structure has better agreement for the carbonyl group (δiso = 320, PQ = 8.3 MHz). The determined δiso and ηQ values are shown to be consistent with bond lengths closer to 1.222 Å (experimental length) rather than the geometry optimised length of 1.238 Å. The geometry optimised DFT 1J(13C,17O) coupling to the hydroxyl is calculated as 20 Hz and the couplings to the ether were calculated to be 37 (O–C
O) and 32 (O–C–OH) Hz. The scalar coupling parameters for the unoptimised experimental carbonyl group predict a 1J(13C,17O) value of 28 Hz, whilst optimisation gives a value of 27 Hz. These calculated 1J(13C,17O) couplings, together with estimations of the probability of each O environment being isotopically labelled (determined by electrospray ionisation mass spectrometry) and the measured refocussable transverse dephasing (T2′) behaviour, are combined to simulate the experimental decay behaviour. Good agreement between the measured and calculated decay behaviour is observed.
The high chemical specificity and ability to probe short range interactions enables solid state NMR to investigate the relationship between a negative O-centred nucleophile and a sp2 hybridized electrophilic C centre. In this study solid naphthalaldehydic acid (1) and its 2H-naphthalaldehydic acid (2) counterpart, depicted in Fig. 1(a) and (b), respectively, will be examined using this technique to characterise C–O bonding arrangements with varying strength. This molecule represents a unique structural motif which possesses three O-based functionalities which are in close proximity within the six-membered lactone crown of the naphthalaldehydic structure, thus introducing numerous C–O bonding scenarios. A recent report describes a series of salts in which the lactone ring has opened to produce isolated aldehyde and carboxylate groups, and the n–π* C⋯O interactions and H-bonding scenarios between them.7 Some prominent examples of this overall phenomenon included the efficacy and structure–function properties of aspirin,8 and the stabilisation of chirality in protein and polymer structures.9–13
![]() | ||
Fig. 1 (a) The structures of the naphthalaldehydic acid variants with the hydroxyl group (1) and the deuterated hydroxyl group (2) as evaluated in this study. The numbering scheme and labels defining the carbonyl (C![]() |
The measurement of scalar (or J-) couplings between directly bonded or closely proximate NMR active nuclei are routinely exploited to aid molecular structure determination.14 Scalar coupling values are quantifiable entities emanating from the hyperfine interactions between neighbouring nuclei and is transmitted through the bonding electrons, providing information about connectivity and dynamics between atomic positions.15 In typical solution state NMR experiments on organic and inorganic compounds, the determination of such couplings can be attained from basic 1D experiments; this is due to fast Brownian motion (on the timescale of the NMR experiment) removing the anisotropic components of all interactions present, thus resulting in high resolution isotropic data. However, it should be emphasised that J-couplings in solution involving the 17O nucleus (particularly 1J(13C,17O) couplings) are typically elusive due to the fast quadrupole relaxation of the 17O nucleus.16 In contrast, the residual inhomogeneous broadenings commonly experienced in solids usually preclude J couplings between common nuclei (1H, 13C, 15N, 31P, 29Si, 17O, etc.) from being measurable in direct observation experiments. Solid state NMR experiments rarely succeed in the achieving the spectral resolution attainable in the solution state because of the existence of higher order cross terms which cannot be completely averaged by MAS, residual dipolar interactions and distributions of isotropic chemical shifts due to structural disorder.17 However, developments in high speed magic angle spinning (MAS),18 effective heteronuclear decoupling and pulse sequence development,19 higher magnetic field strengths, and the GIPAW DFT computation of NMR parameters20–23 have all contributed to improved resolution and introduced new rationales for interpreting solid state NMR data. These advances have stimulated a renewed interest in undertaking J-coupling measurements in the solid state.
The aim of this study is to establish a combined methodology which uses first principles density functional theory (DFT) J-coupling calculations using the CASTEP code and heteronuclear (13C–17O) spin echo-based experiments to determine the 1J(13C,17O) couplings within the various C–O moieties of naphthalaldehydic acid which is a complex lactone arrangement with three O-containing functionalities in close proximity. This approach provides a natural complement to single crystal X-ray structure determinations and charge density studies that can collectively interrogate the relative strength of C–O bonds and could be extended to C⋯O interactions and bond formation scenarios. Solid state NMR J-coupling measurements are typically achieved using J-resolved or INADEQUATE experiments, or specific variants of these sequences.24–29 A key feature of these spin echo-based experiments is that they refocus the evolution of all terms that appear as offsets, including those that are directly induced by chemical shift distributions.30 This has critical implications for solid state NMR experimentation as the utilisation of these methods allows the indirect detection of J-couplings, even when inhomogeneous broadenings impede its direct observation and measurement from the more routinely acquired 1D data. As the 17O nucleus has an extremely low natural abundance (0.037%), there exists a dearth of measured 1J(13C,17O) couplings in both the solid and solution NMR literature. Hung et al. demonstrated 1J(13C,17O) measurements from a doubly labelled (13C,17O) solid glycine·HCl sample, however no further investigations have been reported from solid systems.16 This work represents the first study to implement a combined computational and experimental approach to measure multiple 1J(13C,17O) couplings in a more complex organic system to further investigate C–O bond characteristics.
The naphthalaldehydic acid (100 mg) was deuterated by dissolving the product in dichloromethane (20 ml), and the filtered solution shaken with D2O (99%, 2 × 2 ml). The dichloromethane solution was dried with anhydrous sodium sulphate, filtered and evaporated to give the mono-deuterated material.
1H NMR (400 MHz, (CD3)2CO, 292.1 K): δ = 8.36 (1H, dd, 3J = 7.2, 1.0 Hz, Ar–H1), 8.31 (1H, dd, 3J = 8.4, 1.0 Hz, Ar–H1), 8.09 (1H, d, 3J = 8.2 Hz, Ar–H1), 7.71–7.81 (3H, m, Ar–H3), 7.20 (1H, d, 3J = 6.8 Hz, OH), 6.93 (1H, br d, 3J = 6.0 Hz, CHOH); 13C NMR (100 MHz, (CD3)2CO, 292.1 K): δ = 163.6 br (CO), 134.4, 132.9, 130.8, 129.8, 129.1, 128.0, 127.7, 127.2, 126.1, 121.2 (Ar–C10), 97.0 br (CHOH).
IR (ATR): 3270 br, 1663, 1623, 1597, 1588, 1516, 1407, 1373, 1356, 1316, 1263, 1248, 1189, 1179, 1109, 1070, 993, 984, 919, 904, 843, 771, 739, 728, 687, 652 cm−1; note that the CO stretch in unlabelled material is 1675 cm−1.
MS: (LTQ Orbitrap XL, nano-ESI-negative): found: 203.0555 (13), 202.0522 (100), 201.0484 (67), 200.0444 (15), corresponds to ca. 6:
4
:
1 ratio of triply, doubly and singly 17O labelled naphthalaldehydate anions. Calculated for C12H7[17O]3: 202.0527; for C12H7O1[17O]2: 201.0485; C12H7O2[17O]: 200.0443.
The 13C MAS NMR data was acquired at 11.7 T using a Bruker Avance-III 500 spectrometer operating at a 13C Larmor frequency of 125.8 MHz. A 1H–13C cross-polarisation MAS (CPMAS) experiment was implemented using a Bruker 4 mm double resonance probe which enabled a rotation frequency of 10 kHz. A variable amplitude (70–100%) CPMAS experiment employed an initial 1H π/2 pulse of 2.5 μs, a contact period of 5 ms, a recycle delay of 600 s and a 1H decoupling field of ∼100 kHz during acquisition. The 13C CPMAS NMR data was referenced to the primary standard of tetramethylsilane (TMS, δiso = 0 ppm) via a secondary solid alanine reference (δiso(CH3) = 20.5 ppm, δiso(CH) = 50.5 ppm and δiso(COOH) = 177.8 ppm). The measurement of the 1J(13C,17O) coupling constants were performed using a modified spin-echo experiment as previously described by Hung et al.16 A 1H–13C CP preparation of magnetisation (see above) was followed by a refocussing π pulse (CP–τ–π–τ) on both the 13C and 17O channels thus forming an echo. The 13C and 17O π pulse duration on each channel was 6 μs. The τ delays were incremented from 0 to 15 ms, with 100 kHz of 1H decoupling applied throughout. The 2D 1H–13C Lee–Goldberg heteronuclear correlation experiment used the same cross polarisation preparation as described above and implemented homonuclear Lee–Goldberg 1H–1H decoupling during acquisition. No additional resolution enhancement in the 1H spectrum was observed with the implementation of homonuclear decoupling.
For the quadrupolar 17O (I = 5/2) nucleus MAS NMR studies were performed at 14.1, 16.5 and 20.0 T using Bruker Avance-II+ 600, Bruker Avance-III HD 700 and Bruker Avance-III 850 instruments operating at 17O Larmor frequencies of 81.3, 94.9 and 115.2 MHz, respectively. A Bruker 3.2 mm double resonance probe was used at each field which facilitated a MAS frequency of 20.0 kHz for each measurement. All 17O MAS NMR data were acquired using a rotor-synchronised Hahn echo experiment (θ–τ–2θ–τ–acquire). A ‘non-selective’ (solution) π/2 pulse time of 3 μs was calibrated using 10% H217O which represented a ‘selective’ π/2 (solid) pulse of 1 μs; a θ = π/2 = 1 μs/2θ = π = 2 μs combination was used for all measurements. A recycle delay of 5 s was used throughout; this was carefully checked against acquisitions using recycle delays of up to 60 s. All 17O apparent shifts (centre-of-gravity, δcg) and isotropic chemical shifts (δiso) are reported against a primary reference of H217O (δiso = 0 ppm).32 Additional 17O double rotation (DOR) experiments using odd order sideband suppression33 were undertaken at 9.4, 11.7 and 14.1 T using Bruker Avance-III HD 400, Bruker Avance-III 500 and Bruker Avance-II+ 600 spectrometers operating at the 17O Larmor frequencies of 54.1, 67.8 and 81.3 MHz, respectively. The measurements at each field were performed using Samoson designed DOR probes which functioned with external rotor spinning frequencies of 1.3–1.9 kHz and internal rotor spinning frequencies of 5.9–9.7 kHz.
All central transition spectra from the 17O MAS NMR data of 1 acquired at 14.1, 16.5 and 20.0 T were simulated using the DMFit34 simulation package to determine the experimental isotropic chemical shift (δiso) and quadrupole (CQ, ηQ) NMR parameters (see Table 1 and Fig. 3(a), (c) and (e), respectively). The CASTEP calculated NMR parameters for δiso, 1JOC, CQ and ηQ (see Tables 2 and 3) were used as input into the SIMPSON 4.2.135 simulation package to generate a direct comparison between experimental and calculated data. The SIMPSON input files are given in Table S3 (ESI‡).
Site | 17O MAS NMR | 17O DOR NMR | 17O(2H) DOR NMR | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
δ iso /ppm | C Q /MHz | C Q /MHz | η Q | η Q | P Q /MHz | P Q /MHz | δ cg /ppm | δ cg/ppm | P Q /MHz | P Q/MHz | δ cg /ppm | δ cg/ppm | P Q /MHz | P Q/MHz | |
Error | Error | Error | Error | Error | Error | Error | |||||||||
a IUPAC isotropic shift convention averaged from the multiple field simulations: δiso = (δ11 + δ22 + δ33)/3. b Error derived from average of deconvolutions at three magnetic fields. c EFG tensor conventions: |V33| ≥ |V22| ≥ |V11|, CQ = e2qQ/h = eQV33/h, ηQ = (V11 − V22)/V33 (1 ≥ ηQ ≥ 0). d Quadrupole parameter: δcg (ppm) = δiso − (3/500) (PQ2/ν02) × 106 and PQ (MHz) = CQ (1 + ηQ2/3)1/2. | |||||||||||||||
C![]() |
313 | 8.2 | 0.1 | 0.03 | 0.02 | 8.2 | 0.1 | 285 | 19 | 7.3 | 1.7 | 308 | 2 | 8.1 | 0.1 |
–O– | 200 | 8.9 | 0.1 | 0.25 | 0.01 | 9.0 | 0.1 | 173 | 20 | 8.2 | 1.5 | 199 | 1 | 9.0 | 0.1 |
–OH/–OD | 66 | 9.6 | 0.4 | 0.81 | 0.08 | 10.6 | 0.4 | 63 | 19 | 10.4 | 0.1 | 16 | 32 | 8.6 | 2.5 |
O site | GIPAW DFT calculated 17O NMR parameters (experimental (XRD) structure) | GIPAW DFT calculated 17O NMR parameters (geometry optimised structure) | ||||||
---|---|---|---|---|---|---|---|---|
δ iso/ppm | C Q/MHz | η Q | P Q/MHZ | δ iso/ppm | C Q/MHz | η Q | P Q/MHz | |
C![]() |
320 | 8.3 | 0.09 | 8.3 | 307 | 8.1 | 0.31 | 8.2 |
–O– | 210 | 9.3 | 0.30 | 9.4 | 223 | 9.3 | 0.25 | 9.4 |
–OH/–OD | 26 | 9.7 | 0.65 | 10.4 | 62 | 9.5 | 0.82 | 10.5 |
Coupling constant | 13C nuclei | 17O nuclei | Bond length/Å | Bond angle/°O–C–O | 1 J co/Hz | Optimised bond length/Å | Optimised bond angle/°O–C–O | Optimised 1Jco/Hz |
---|---|---|---|---|---|---|---|---|
1 J CO1 | C11 | C![]() |
1.222 | 117.5 | 28 | 1.238 | 117.4 | 32 |
1 J CO2 | –17O– | 1.342 | 117.5 | 37 | 1.338 | 117.4 | 37 | |
1 J CO3 | C12 | –17O– | 1.477 | 106.7 | 32 | 1.490 | 107.4 | 32 |
1 J CO4 | –17OH | 1.386 | 106.7 | 21 | 1.379 | 107.4 | 20 |
Geometry optimisation and H position optimisation calculations were performed on all systems using the over-converged values described previously to ensure accurate forces. For these structural optimisations, the lattice parameters were kept fixed whilst the atomic positions were optimised to a force tolerance of 0.05 eV Å−1, a maximum ionic displacement of 1 × 10−3 Å and a total energy change of 2 × 10−5 eV per atom. NMR parameter calculation of the chemical shift,22,36 electric field gradient tensors20 and J-couplings23,37 invoked the gauge included projector augmented wave (GIPAW) DFT method which extended the pseudopotential (valence electron approximation) approximation to recover all electron charge densities. This approach demonstrated that the NMR parameters depend more sensitively upon the electron density than the total energy, so each NMR parameter will be more sensitive to basis set truncation errors than to total energies. This is compensated for by the over-convergence of the basis set to a level which experience suggests will produce fully converged NMR parameters. Consequently, these calculations were completed using the plane wave cut-off and k-point Monkhorst–Pack grids described earlier.
In order to calculate isotropic chemical shifts for the 17O nucleus the relationship δiso = −[σ − σref] was employed, where σ is the calculated isotopic shielding calculated against a bare atom, and σref is the previously determined reference isotropic shielding of 262 ppm when compared against a bare atom.38 These calculations used the same basis set convergence (0.4 mH) as the original calculations to minimise the propagation of errors.
As observed in Fig. 1(b), the crystalline salt naphthalaldehydic acid 1 has a cyclic lactone structure with only one crystallographically unique molecule comprising the asymmetric unit of the crystal. There exists intermolecular H-bonding between the CO functionality and the –OH moiety on an adjacent molecule constrained by a twofold screw axis, with the two symmetry related H-bonds associated with each molecule assuming an O⋯O distance of 2.754(4) Å.7 There is an anomeric interaction between the two O atoms attached to the methine C leading to a short C–OH bond length of 1.387 Å and a long C–O(C
O) bond length of 1.477 Å due to donation of lone pair electron density from the hydroxyl O atom into the adjacent methine CH–O(C
O) bond.7 In comparison, a typical unperturbed C–O bond length is usually observed to fall between these values at ∼1.43 Å.39
The 1H MAS NMR data from naphthalaldehydic acid measured at 14.1 T under fast MAS conditions (νr = 60 kHz) is shown in Fig. 2(a). This spectrum shows a single broad resonance with no individual proton environments able to be resolved from this signal, despite the acquisition of this data under high field and fast MAS conditions. The aromatic, hydroxyl and methine 1H species are all expected to appear at similar isotropic chemical shifts (δ 6.8–8.4 ppm in solution), and the strong homogeneous 1H–1H dipolar interaction broadens all resonances thus resulting in a single unresolved feature. It is possible that both the 1H and 13C resonances suffer from anisotropic bulk magnetic susceptibility (ABMS) broadening, however this broadening typically only increases the linewidth by ∼1 ppm.40 These problems are commonly observed in highly complex aromatic systems and heavy (but not complete) deuteration of the sample is required to significantly increase spectral resolution.41
The 17O labelled naphthalaldehydic acid sample 1 was selectively deuterated at the hydroxy position to form a doubly 17O/2H enriched variant 2 (see Fig. 1(a)) to investigate (and reduce) the potential of 1H motional modulation effects on the 17O MAS and DOR NMR data (vide infra). The 2H MAS NMR data from this quadrupolar (I = 1) nucleus shown in Fig. 2(b) depicts a broad satellite transition spectrum with a 2H quadrupole coupling constant (CQ) of ∼200 kHz being measured from the simulation of this data. Previous measurements of 2H CQ values from various enriched –OD arrangements demonstrate that CQ occurs within the range of ∼190–290 kHz;42,43 the CQ characterising the –OD moiety in this hydroxyl-lactone species represents a value at the low end of this range thus suggesting that partial motional averaging of this interaction is evident. This value is significantly smaller than the predicted CQ value of ∼308 kHz as computed using the CASTEP code (see Table S1 in the ESI‡). These DFT calculations in their basic form do not account for rotation, vibration or libration of the –OD bond, hence any subsequent motional averaging of the static quadrupole interaction is not modelled. Recently, it has been suggested that dueteration can also reduce the point symmetry of the moiety.44 It is important to note that the δiso(2H) for the –OD species is ∼8.0 ppm, and this is corroborated by the 1D 1H MAS and 2D 1H–13C HETCOR MAS NMR data depicted in Fig. 2(a) and (d). The 2D 1H–13C HETCOR data of Fig. 2(d) clearly shows that all protonated C species are correlated with the broad 1H resonance at ∼8.0 ppm, indicating that the δiso(–OH) value is not resolved from the more intense 1H resonances associated with the aromatic protons. Collectively, the observations of a more deshielded 1H chemical shift (δiso(1H) ∼8.9 ppm) and a partial motional averaging of the 2H quadrupole interaction (CQ ∼200 kHz) indicates that the bond covalence characterising the –OH group is of intermediate strength, while the intermolecular H-bonding interaction inherent to the crystal structure of 1 (influencing the –OH proton) is weak.31,38
Solid state 17O MAS NMR data from 1 measured at 20.0 T (ν0 = 115.2 MHz), 16.5 (ν0 = 94.9 MHz) and 14.1 T (ν0 = 81.3 MHz) presented in Fig. 3(a)–(c), respectively, shows the three O species present in the six membered hydroxyl-lactone ring form of the naphthalaldehydic acid lactone crown are all influenced by substantial second order quadrupole interactions. The highly characteristic second order broadened quadrupole resonances are fully resolved at 20.0 T, however due to the 1/B0 dependence of the quadrupole interaction the observed linewidths increase with decreasing B0, consequently reducing the resolution of the lower field data. Corresponding 2D 17O 3QMAS data from 1 acquired at 16.5 T shown in Fig. 4 clearly demonstrates the chemical purity of this system, and that very high structural order is reflected by the absence of any distributions in the chemical shift and/or distributions of quadrupole NMR parameters. Previous studies have established that H-bond motion can modulate the multiple quantum excitation and evolution processes in the MQMAS experiment thus inducing detrimental effects on the resolution observed in the F1 dimension, however this is also not observed here.38
![]() | ||
Fig. 3 The 17O MAS NMR data and spectral deconvolution from 17O enriched naphthalaldehydic acid 1 acquired at (a) 20.0 T (ν0 = 115.2 MHz, νr = 20 kHz), (c) 16.45 T (ν0 = 94.9 MHz, νr = 20 kHz), (e) 14.1 T (ν0 = 81.3 MHz, νr = 20 kHz), compared against their corresponding SIMPSON45 simulations (b), (d), (f) generated using GIPAW DFT calculations of the NMR parameters from the single crystal naphthalaldehydic acid 1 structure using the CASTEP code. In each case, the green resonance represents the carbonyl (C![]() |
Simulation of the 17O 1D data for the CO species (i.e. the most deshielded 17O environment) yields an isotropic chemical shift of δiso = 313 ppm and quadrupole parameters of CQ = 8.2 MHz/ηQ = 0.03. A 17O δiso value of this magnitude is low compared to a ketone group, but it is within the reported limits of typical carboxylic acid and lactone functionalities.46,47 Conversely, the observed 17O resonance from the hydroxyl (–OH) species is the most shielded (δiso = 66 ppm) and represents a larger quadrupole interaction with CQ = 9.6 MHz/ηQ = 0.81. Both sets of 17O quadrupole parameters measured for both the C
O and the –OH groups are in firm agreement with previously reported data in the literature observed for amino acid·HCl salts.38,46,48 The high ηQ value for the –OH species, and corresponding low ηQ value for the C
O species are typical for 17O nuclei bonded to larger conjugated aromatic moieties.38,49 In previous studies on amino acids, it has been reported that the 17O asymmetry parameter (ηQ) is close to zero for C⋯O distances in a C
O functionality of ∼1.22 Å, while it appears to increase rapidly for longer bond lengths. Simulation of the 17O resonance from the ether (–O–) linkage (located between the C
O and –OH resonances) in this lactone structure yields an isotropic chemical shift of δiso = 200 ppm and quadrupole parameters of CQ = 8.9 MHz/ηQ = 0.25. All isotropic chemical shift and quadrupole parameters characterising the three O species comprising this lactone structure are in agreement with previously reported values for the lactone of 2-[1,2,3-17O3]acetylbenzoic acid which is the only other reported labelled lactone structure studied by solid state NMR.47 Despite the use of high magnetic fields in this study, no chemical shift anisotropy (CSA) contributions were required to facilitate the accurate simulation the data in Fig. 3(a)–(c), suggesting that CSA influences are small in comparison to the dominant quadrupole interaction and subsequently eliminated by MAS.
For moderate-weak H-bonded systems, a previously established empirical relationship between the 17O quadrupole coupling constant of the O species involved in X–H⋯O (X = N, O) H bonds and the X⋯O distance across the whole X–H⋯OH bond takes the form:46,50,51
CQ ≈ 1.5199(R(X⋯O)) + 3.9987 | (1) |
In addition, there is a strong correlation between the CO bond length and the carbonyl 17O isotropic chemical shift (∼1200 ppm Å−1), with a shift of ∼313 ppm representing a bond length of 1.222 Å.48 A comparison of the C
O bond lengths to C
17O δiso values compiled by Pike et al.,22 together with the GIPAW DFT data reported by Gervias et al.,52 produces an overall empirical correlation relating the 17O δiso value to the carbonyl bond length (R(C
O)):
δiso(C![]() ![]() | (2) |
From inspection of the spectral and computationally generated data exhibited in Fig. 3 and Tables 1 and 2, there exists a very strong correlation between the directly measured and simulated 17O MAS NMR data (see Fig. 3(a), (c) and (e), and Table 1) and the GIPAW DFT calculated 17O NMR interaction parameters from a geometry optimised structure of 1 using the CASTEP code (see Fig. 3(b), (d) and (f) and Table 2). In particular, the experimentally determined –OH data (δiso = 66 ppm, CQ = 9.6 MHz, ηQ = 0.81), –O– data (δiso = 200, CQ = 8.9, ηQ = 0.25) and CO data (δiso = 313 ppm, CQ = 8.2 MHz and ηQ = 0.03) show an excellent agreement with the corresponding DFT results (–OH, δiso = 62 ppm, CQ = 9.5 MHz, ηQ = 0.85; –O–, δiso = 223ppm, CQ = 9.3 MHz, ηQ = 0.25; C
O, δiso = 307 ppm, CQ = 8.1 MHz and ηQ = 0.31) when geometry optimisation is implemented. These correlations are marginally stronger than those derived from calculations using non-geometry optimised structures; this is particularly evident when comparing the MAS NMR measured and DFT calculated –OH δiso values when geometry optimisation is omitted. In this case, a discrepancy of ∼40 ppm exists. Nevertheless, the strength of the above correlations suggests that the GIPAW DFT method represents a very sound approximation for describing the electronic configuration of O species in organic solids, and this approach thus presents a solid foundation on which to underpin 1J(13C,17O) coupling calculations in the naphthalaldehydic acid system 1. The DFT calculations also highlight that both geometry optimised and direct calculations on the X-ray derived structure need to be considered when calculating the 1J(13C,17O) couplings in this system.
Unlike the conventional MAS NMR technique, the DOuble Rotation (DOR) experiment involves spinning the sample about two angles to enable averaging of both the P2 and P4 terms of the 4th order Legendre polynomial, which describes the angular dependences of the second order quadrupole interaction. The first order quadrupole broadening is reduced by sample rotation around the familiar magic angle of 54.7° with respect to B0, while the second order quadrupole broadening is eliminated by simultaneous rotation at 54.7° and a second angle of 30.6° with respect to the magic angle. The two major advantages that the DOR NMR technique offers are, (a) the observed (featureless) centre-of-gravity shift (δcg) yields greater resolution than conventional MQMAS experiments which are limited by the number of attainable slices defining the F1 dimension, and (b) the DOR experiment uses a single pulse excitation experiment in comparison to two-pulse or three-pulse MQMAS experiments which possess limited pulse efficiencies. The greater resolution and superior signal-to-noise ratio afforded by the DOR technique is met with technical challenges; sample rotation at two angles simultaneously is difficult to stabilise, and the limited rotational rates achievable by the inner and outer rotors introduces large residual spinning sideband manifolds which necessitate acquisitions at multiple spinning rates to isolate the central transition.33,54
For a quadrupolar nucleus such as 17O, the apparent centre-of-gravity shift (δcg) is comprised of both field independent isotropic chemical shift (δiso) and field dependent second-order quadrupole shift (δ(2)Q,iso) components:
δcg = δiso + δ(2)Q,iso(I,m) | (3) |
δ(2)Q,iso(I,m) = (3CQ2/(40νo2I2(2I − 1)2))[I(I + 1) − 9m(m − 1) − 3](1 + ηQ2/3) | (4) |
δcg(ppm) = δiso(ppm) – (3/500)(PQ2/ν02) × 106 | (5) |
PQ = CQ√(1 + ηQ2/3) | (6) |
High resolution data afforded by the 17O DOR technique allows some subtle aspects of motion affecting the O positions to be detected. The characterisation of these properties is an important prerequisite to the measurement and calculation of 1JCO couplings. Table 1 summarises the δiso and PQ data elucidated from the variable field (14.1, 11.7 and 9.4 T) 17O DOR measurements and the graphical approach outlined above for 1 (see Fig. 5(a)–(d)) and 2 (see Fig. 5(e)–(h)). Each spectrum shows the centre-of-gravity shifts (δcg) identified; these were isolated from the spinning sideband manifolds by varying the spinning frequency of the outer rotor over a range of 1.3–1.9 kHz. From the undeuterated system 1 the determined 17O δiso values for the carbonyl (CO), ether (–O–) and hydroxyl (–OH) species are 285, 173 and 63 ppm, respectively, while the analogous δiso values of 308, 199 and 16 ppm are obtained from the deuterated version 2. Upon deuteration, a marked deshielding of ∼23–26 ppm in the 17O δiso values are observed for the C
O and –O– moieties, which is in contrast with the increased shielding of ∼47 ppm observed for the –OH species.
Previous studies have demonstrated that the deuteration of protonated oxo functionalities in solid organic systems improves the resolution obtained in the 17O DOR experiment. This is achieved by damping the heteronuclear dipolar proton coupling which homogenously broadens the 17O resonance.60–63 This is corroborated by the data in Fig. 5 where the 17O resonance for the –OD (see Fig. 5(e)–(h) for 2) moiety is significantly narrower and of greater intensity than its –OH counterpart (see Fig. 5(a)–(d) for 1), especially at the higher fields of 11.7 and 14.1 T. As observed from Table 1 and Fig. 5(d) and (h), the data for the CO and –O– moieties from the undeuterated sample 1 show an inferior overall correlation on the linear regression and greater errors on each individual measurement in comparison to their counterparts from the deuterated sample 2. This is a consequence of the broader 17O resonances due to proton motion in the H-bond which induces additional cooperative modes of motion throughout the other O functionalities. Furthermore, for the C
O and –O– moieties the measured δiso and PQ values from the DOR experiment closely corroborate those reported from the MAS, MQMAS and geometry optimised DFT (GO-DFT) data. In comparison, for both samples 1 and 2 the overall correlations described in the linear regression for the –OH species are inferior to those described above for the C
O and –O– moieties. As suggested above, a reduced correlation for the –OH data (sample 1) is expected to emanate from proton heteronuclear dipolar coupling in the H-bonded arrangement; however, the –OD data (sample 2) exhibits a more reduced correlation and much larger errors. This contradicts previously reported evidence that proposes sample deuteration reduces these errors, promoting more resolved 17O DOR data, and enables more accurate measurements of the δiso and PQ parameters.33 This anomaly suggests that the reduced correlations/larger errors from 2 originate from lower point energy of the OD bond and results in an isotope effect on the NMR parameters.62,63 The deuteration procedure of 17O enriched sample 1 (using D2O) to produce sample 2 does not yield a homogeneous doubly labelled 2D/17O system. Some proportion of the original 17O enrichment has exchanged with 16O producing a complex mixture of 2D/17O, 2D/16O, 1H/17O and 1H/16O species, thus producing a complex distribution of isotope effects. This phenomenon is demonstrated in Fig. 4(d) where 17O MAS NMR data from the deuterated sample 2 (which originated as sample 1) exhibits greatly inferior signal/noise in comparison to sample 1 shown in Fig. 4(b). These complex isotope affects clearly directly influence upon the 17O δiso and PQ parameters characterising each O environment, and the deuteration of molecule 1 can produce measured NMR parameters which are not directly comparable to those calculated from the single crystal derived structure. The process of deuteration of hydroxyl environments to reduce heteronuclear dipolar coupling, and potentially reduce the point symmetry of the moiety, is occasionally employed in the literature. Here, we show this is not best employed when trying to accurately compare diffraction derived DFT NMR parameters and the experimental MAS data. Overall, the GIPAW DFT-derived CQ and ηQ parameters show a strong correlation with the MAS and MQMAS achieved parameters, therefore for the following 1J determinations the experimentally determined parameters will be used.
The spin echo dephasing data from the (undeuterated) naphthalaldehydic acid system 1 is presented in Fig. 6, and the measured heteronuclear 1J(13C,17O) coupling constants are summarised in Table 3. This study represents only the second measurement of 1J(13C,17O) coupling constants in the solid state. While the first was demonstrated on a doubly 13C/17O labelled solid glycine·HCl sample,16 this work focusses on a uniformly labelled 17O system which possesses multiple O positions which are in close proximity.
![]() | ||
Fig. 6 The experimental (black) and DFT predicted (red, SIMPSON simulated) 13C/17O heteronuclear spin echo dephasing behaviour for (a) C11 and (b) C12 acquired at 11.7 T (νr = 11 kHz) as a function of evolution time τ (0 to 16 ms). The predicted 13C/17O heteronuclear spin echo dephasing behaviour (red line) is derived using GIPAW DFT (CASTEP) calculations of the one-bond dominant 1J(13C,17O) and 1J(13C,13C) coupling constants (1JC11![]() ![]() ![]() |
A CT-selective 17O π-pulse in a heteroncuelar spin echo experiment only refocuses the central-transition (CT) spin states, consequently confining the dephasing behaviour of the 13C signals observed in Fig. 6 to only those13C spins bonded to 17O spins in the +1/2 or −1/2 CT spin states. Presuming 100% 17O enrichment of an isolated 13C–17O two spin system, only 1/3 of 13C spins attached to 17O CT (+1/2 or −1/2) spin states will show a cosine J-modulation, while the other 2/3 of the 13C spins will show no cosine modulation as they couple to the satellite transition (ST) +5/2, +3/2, −3/2 or −5/2 spin states. Hence, a J-resolved experiment utilising a CT-selective 17O pulse for a 13C–17O spin pair will exhibit a cosine modulation (presuming no T2′ contribution) that goes from an initial signal intensity of 1, down to a minimum signal of S(τ)het = 0.33, with a 0.66 signal intensity for a J-evolution time of 1/2J. However, the relatively broad nature (∼41 kHz) of the multiple 17O environments in naphthalaldehydic acid means uniformly exciting the three oxygen environments required non-selective pulses, therefore accurate SIMPSON simulations are required to evaluate the data with respect to the pulse lengths used, and fitting the decay using a function is challenging.
All data presented in Fig. 6 were acquired using a modified spin echo sequence as detailed by Hung et al.,16 with the simulation of each multi-component spin echo dephasing curve being achieved using SIMPSON.45 An analytical expressions for deriving isolated 1J-couplings have been developed by Duma et al., with the quadrupolar contributions to the scalar coupling explored by Perras and Bryce.64,65 SIMPSON 4.2.1 cannot simulate multiple 1J-couplings with multiple quadrupole coupling constants. Therefore, the simulation was broken up into their individual one-bond components and multiplied together. The π pulse length (6 μs, 41 kHz), was also considered in the SIMPSON calculation to account for any errors or issues caused by finite pulse lengths, this pulse length was chosen as it is the same as the linewidth of the three oxygen environments.
The output of the SIMPSON simulations were then scaled to represent the oxygen enrichment, which was determined by high-resolution ESI-MS. This gave the following probabilities, p:
pO1O2 = pO3O4 = 80%; probabilities that C11 or C12 is directly bonded to two enriched 17O species, |
pO1 = pO2 = pO3 = pO4 = 7.5%; probabilities that C11 or C12 is directly bonded to one enriched 17O species, |
pnone = 5%; probability that C11 or C12 are not directly bonded to any enriched 17O species. |
The GIPAW DFT calculated 1J(13C,17O) coupling constants for the O positions are derived from the single crystal structure determination of 1,7 with these calculations being performed on both the geometry optimised and the experimental structures, coupled with associated dispersion corrections. These results are summarised in Table 3. Key to the effectiveness of this computational approach for these 1J(13C,17O) coupling constants is the highly accurate characterisation of 1 involving the more routinely GIPAW DFT calculated NMR parameters such as δiso, CQ and ηQ as presented in Fig. 3, and Tables 1 and 2, which demonstrates that the ultrasoft ‘on-the-fly’ O pseudopotential is particularly optimised for 17O NMR measurements in organic solids. This provides the simulation of the dephasing characteristics for C11 and C12 shown in Fig. 6(a) and (b) using the calculated 1J(13C,17O) coupling constants (1JCO1, 1JCO2, 1JCO3 and 1JCO4) with greater statistical importance.
As the C11 and C12 positions are not 13C labelled, the spin echo intensities at larger evolution times are difficult to determine, as they cannot be accurately discriminated from the noise. The error in each data point grows cumulatively for increasingly longer evolution times; as observed from Fig. 6, these dephasing measurements were truncated at a maximum evolution time of 15 ms. The contributions from neighbouring proton species are removed by high power 1H (100 kHz) decoupling during the evolution and acquisition periods; however, the XJ couplings to neighbouring 13C spins are not suppressed. The 1J-coupling from C11 to C1 and C12 to C9 are 58 and 39 Hz, respectively. As the natural abundance of the 13C isotope is 1.1%, these species will induce minor contributions to the dephasing curve and therefore these were considered in the DFT derived SIMPSON simulations. Extended homonuclear 1J(13C,13C) contributions are predicted by CASTEP to have an insignificant contribution to the decay (Table S2, ESI‡), and as there is no 13C labelling in the structure, the maximum contribution of this 1J(13C,13C) coupling is 1.1% of cos(πJCCτ) to the overall result, these may account for a minor source of error in the observed decay. Using high power decoupling limits the acquisition times you can safely record the data, therefore the cosine modulation feature which is predicted at a τ between 20 to 24 ms was not possible to observe. This would also be challenging to reproduce accurately due to reduced signal-to-noise at these larger τ delays and the need for high power decoupling at these extreme lengths is likely to cause damage to the probe.
Despite these sources of error, the results presented in Fig. 6 suggest that overall methodology invoked above appears to be a robust approach for calculating multiple 1J(13C,17O) couplings in a complex arrangement containing multiple oxo functionalities in the solid state. In particular, it is evident that GIPAW DFT calculations of 1J(13C,17O) coupling constants are accurate as observed by the strong correlation between the experimental dephasing behaviour and the simulated S(τ)het behaviour (adopting the isolated spin system approximations above) using calculated values of 1JCO1, 1JCO2, 1JCO3 and 1JCO4 from CASTEP. The correlation between the calculated data using the proposed model and the experimental data is excellent for the three nuclear spin system involving C12 which is directly bonded to the methine (–O–) and hydroxyl (–OH) moieties (i.e.1JCO3 and 1JCO4). It should be noted that the reduced correlation describing C11 involves 1JCO1 coupling to the CO moiety which is influenced by intramolecular H-bonding to the –OH species in an adjacent molecule. Furthermore, the errors in the C11 heteronuclear spin echo experiment were larger, due to insufficient signal to noise, possibly due to less efficient cross polarisation.
It is interesting to note that the magnitudes of the two 1J(13C,17O) couplings to the central ether –O– linkage reflect the corresponding bond lengths involved. The 1J(13C,17O) coupling constant for the shorter bond length CO species is 37 Hz (carbonyl C–O bond length: 1.338 Å), which is larger than the coupling constant of 32 Hz to the more distant CH–OH (methine C–O bond length: 1.490 Å). In contrast, the 1J(13C,17O) coupling constant between the methine C and the –OH species is smaller than that to the ether –O– atom (20 Hz cf. 32 Hz) despite having a significantly shorter bond length (1.379 Å cf. 1.490 Å). As mentioned above, –OH motion within the weak H-bonding arrangement can modulate and diminish the magnitude of this 1J(13C,17O) coupling, however the bond angle disposition as proposed by the Karplus relationship may also contribute to this marked reversal in this trend.15,66 Although calculations67 and measurements68 of 1J(13C,17O) coupling constants characterising the CO and CO2 gas phase systems have been reported (measured/calculated; 16.4/17.2 Hz and 16.1/17.1 Hz, respectively), the only existing measurement of 1J(13C,17O) couplings ascertained from a real solid state system pertains to glycine·HCl where 1J(C
O) = 25 Hz and 1J(C–OH) = 28 Hz.16 The determined 1J (13C,17O) coupling constants that characterise solid naphthalaldehydic acid 1 summarised in Table 3 are comparable to these previous values, and this study reports the first 1J(13C,17O) measurements on ether –O– linkages. This work has demonstrated that measurements of 1J(13C,17O) coupling constants is not limited to selectively labelled highly isolated spin systems in small molecules, and that this information can be elucidated from larger and more complex spin systems which are isotopically enriched by more routine synthetic methods.
The geometry optimised DFT structure derived using the CASTEP code gives a firm agreement with the experimentally measured shifts observed for the ether (δiso = 223, PQ = 9.4 MHz) and hydroxyl (δiso = 62, PQ = 10.5 MHz) groups, however the unoptimised experimental structure has better agreement for the carbonyl group (δiso = 320, PQ = 8.3 MHz). The determined δiso and ηQ values for the latter are shown to be consistent with bond lengths closer to 1.222 Å emanating from the unoptimised experimental XRD structure, rather than the geometry optimised length of 1.238 Å. The geometry optimised DFT 1J(13C,17O) coupling to the hydroxyl is calculated as 20 Hz and the 1JCO couplings to the ether oxygen were calculated to be 37 (O–CO) and 32 (O–C–OH) Hz. These measurements are all within error of the experimental echo decays.
Footnotes |
† The experimental data for this study are provided as a supporting dataset from WRAP, the Warwick Research Archive Portal at http://wrap.warwick.ac.uk/130520. |
‡ Electronic supplementary information (ESI) available. See DOI: 10.1039/c9cp03977e |
This journal is © the Owner Societies 2020 |