Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jun 3.
Published in final edited form as: J Am Chem Soc. 2020 Aug 19;142(35):14877–14889. doi: 10.1021/jacs.0c02810

Higher Magnetic Fields, Finer MOF Structural Information: 17O Solid-State NMR at 35.2 T

Vinicius Martins ‡,#, Jun Xu $,*,#, Xiaoling Wang §, Kuizhi Chen §, Ivan Hung §, Zhehong Gan §,*, Christel Gervais #, Christian Bonhomme #,*, Shijia Jiang $, Anmin Zheng , Bryan EG Lucier , Yining Huang ‡,*
PMCID: PMC8175036  NIHMSID: NIHMS1705751  PMID: 32786791

Abstract

The spectroscopic study of oxygen, a vital element in materials, physical, and life sciences, is of tremendous fundamental and practical importance. 17O solid-state NMR (SSNMR) spectroscopy has evolved into an ideal site-specific characterization tool, furnishing valuable information on the local geometric and bonding environments about chemically distinct and, in some favorable cases, crystallographically inequivalent oxygen sites. However, 17O is a challenging nucleus to study via SSNMR, as it suffers from low sensitivity and resolution, owing to the quadrupolar interaction and low 17O natural abundance. Herein, we report a significant advance in 17O SSNMR spectroscopy. 17O isotopic enrichment and the use of an ultrahigh 35.2 T magnetic field has unlocked identification of many inequivalent carboxylate oxygen sites in the as-made and activated phases of the metal–organic framework (MOF) α-Mg3(HCOO)6. The subtle 17O spectral differences between the as-made and activated phases yield detailed information about host–guest interactions, including insight into nonconventional O…H─C hydrogen bonding. Such weak interactions often play key roles in applications of MOFs, such as gas adsorption and biomedicine, and are usually difficult to study via other characterization routes. The power of performing 17O SSNMR experiments at ultrahigh magnetic field of 35.2 T for MOF characterization is further demonstrated by examining activation of the MIL-53(Al) MOF. The sensitivity and resolution enhanced at 35.2 T allows partially- and fully-activated MIL-53(Al) to be unambiguously distinguished, and also permits several oxygen environments in the partially-activated phase to be tentatively identified. This demonstration of the very high resolution of 17O SSNMR recorded at the highest magnetic field accessible to chemists to date illustrates how a broad variety of scientists can now study oxygen-containing materials and obtain previously inaccessible fine structural information.

Graphical Abstract

graphic file with name nihms-1705751-f0001.jpg

1. Introduction

The element of oxygen is ubiquitous across nearly all scientific fields. Therefore, characterization of oxygen local electronic and geometric environments is very important. 17O solid-state NMR (SSNMR) spectroscopy has become an ideal site-specific characterization tool for probing oxygen local environments, as 17O is sensitive to the chemical shift and quadrupolar interactions,1-11 has a large diagnostic chemical shift range,12-25 and is influenced by coupling to neighboring NMR-active nuclei (e.g., 1H, 13C, and 15N).26-31 There has been tremendous progress made in NMR methodology and technology in recent years, yet the potential of 17O SSNMR for uncovering detailed structural and bonding information in oxygen-containing compounds has been limited by the inherently low sensitivity and resolution resulting from the very low natural abundance (0.038%), relatively low gyromagnetic ratio (γ = −5.774 MHz·T−1), and quadrupolar nature (spin I = 5/2) of 17O.32

The sensitivity problem associated with the low 17O natural abundance can be mitigated by isotopic enrichment.16-22 To address the relatively low γ and quadrupolar nature of 17O, NMR measurements can be performed at high magnetic fields; this not only inherently enhances sensitivity but also reduces spectral line broadening associated with the second-order quadrupolar interaction. A new series-connected resistive/superconducting hybrid magnet operating at a record-high magnetic field strength of 35.2 T (1H Larmor frequency of 1.5 GHz) has recently entered service,33 which promises very high 17O SSNMR resolution in biomolecules and minerals.33-35 In this work, taking advantage of the state-of-the-art magnet and rf technology, we targeted microporous α-Mg3(HCOO)6 to demonstrate that very high spectral resolution of 17O SSNMR spectra can be achieved at 35.2 T, providing an excellent opportunity for characterizing promising materials such as metal-organic frameworks (MOFs).

MOFs are a fascinating family of hybrid organic-inorganic porous materials with many practical applications.36-37 SSNMR spectroscopy of MOFs has proven to be a powerful tool for characterizing the immediate environment about metal centers and probing the local structure of organic linkers.38-40 SSNMR can also provide information on the behavior of adsorbed guests, which is critically important for many applications.41-42 For example, MOFs are promising materials for the removal of greenhouse gas such as CO2 and storage of fuels such as H2 and CH4. SSNMR can provide information on the location of guest species,10, 43-47 which is critically important for practical applications, as the location of guest gas molecules can be directly linked to binding site positions and strength, respectively. Similarly, ascertaining the location of guest species in MOFs is key to understanding their applications in sensing and drug delivery.48-51 Oxygen present in various carboxylate ligands, which are the most extensively used organic linkers, is a key constituent of many important MOFs.52-56 Oxygen anions (O2−) are also associated with the metal clusters of the frameworks (e.g., MOF-5).52 Hydroxyl groups are common linkers bridging metal clusters (e.g., MIL-53)53 and exist as part of the secondary building units (e.g., UiO-66).54 Water molecules can directly bond to the metal center, with well-known examples include as-made MOF-74 and HKUST-1.55-56 These oxygen species play critical roles in applications such as guest adsorption/separation,57 sensing,48 catalysis,58 solid-state conductors,59 and biomedicine,49-51 rendering oxygen a key target for MOF characterization. Although 17O SSNMR has been utilized to examine some MOFs,10, 15-17 potential successes in molecular-level characterization and site assignment have been limited by spectral resolution.

Microporous α-Mg3(HCOO)6 is an attractive target for 17O SSNMR characterization for several reasons.

  1. Microporous α-Mg3(HCOO)6 is a commercially available and low-cost MOF with good molecular selectivity, such as a preference for C2H2 over CO2.60-61 It is a representative small pore MOF suitable for gas adsorption.

  2. This MOF presents a very challenging case for characterization by 17O SSNMR, as the crystal structure features twelve inequivalent carboxylate oxygen sites across two bonding modes. Using our highest available field of 21.1 T at the time, only two 17O NMR signals corresponding to the two different oxygen bonding modes of formate anions could be observed (vide infra).16 As increasing the magnetic field from 21.1 to 35.2 T leads to an improvement in resolution by a factor of 2.8 (the second-order quadrupolar broadening in ppm varies as the inverse ratio of the fields squared),1 resolving many of these inequivalent oxygen sites at 35.2 T should be possible.

  3. The applications of MOFs require the removal of the solvent molecules from inside the pores of as-made MOFs in a process known as “activation”. Activation often leads to changes in the framework structure. While significant changes can be detected by X-ray diffraction, activation-induced changes for MOFs including α-Mg3(HCOO)6 can be subtle (i.e., small changes in unit cell parameters), and the specific molecular-level alterations cannot be detected by diffraction-based methods. Thus, it is important to develop SSNMR as a method complementary to XRD. The high spectral resolution that can be achieved at 35.2 T holds promise for providing fine details on oxygen local environments, reflecting the changes resulting from activation.

  4. MOFs have many potential applications across fields such as biology and medicine.62-63 Thus, nonconventional O…H─C hydrogen bonding involving guest species and bioactive components in the framework can play important roles for host–guest interactions in BioMOFs, a new subclass of MOFs,49-50 or MOF-based drug delivery systems.51 Since nonconventional O…H─C hydrogen bonding has been observed in as-made α-Mg3(HCOO)6 (C6H6Mg3O12·C3H7NO, C3H7NO is N,N´-dimethylformamide or DMF),64 this system has been selected as a model compound to explore the possibility of using 17O SSNMR at 35.2 T to directly probe this weak host–guest interaction in MOF systems.

Our results show that the 17O SSNMR spectra of α-Mg3(HCOO)6 samples acquired at 35.2 T indeed exhibit very high resolution, allowing many inequivalent framework oxygen sites to be identified. The high resolution and sensitivity realized at 35.2 T not only lead to ultrafine information on oxygen local environment and corresponding subtle changes upon activation, but also make it possible to detect weak host–guest interactions such as nonconventional O…H─C hydrogen bonding. To further illustrate the benefits of performing 17O SSNMR at an ultrahigh field of 35.2 T, we examined two activated MOF MIL-53(Al) samples; the high spectral resolution and sensitivity allow us to unambiguously distinguish partially-activated from fully-activated MIL-53(Al) samples. The very high resolution and sensitivity of 17O SSNMR achievable at 35.2 T detailed in this work illustrates the great potential of SSNMR for unlocking fine structural information in solids as higher magnetic fields become increasingly available.

2. Experimental Methods

2.1. Synthesis and characterization of MOF samples.

As-made 17O-enriched α-Mg3(HCOO)6 was synthesized according to the reported procedure.16 The starting materials were used as received without further purification. A mixture of Mg(NO3)2·6H2O (3 mmol, Aldrich) and HCOOH (6 mmol, Aldrich) was dissolved in 10 mL of DMF and 0.25 mL of 17O-enriched H2O (6 mmol, CortecNet, 35 atom%) in a 23 mL Teflon-lined autoclave. The container was sealed and heated at 383 K for 2 days. After cooling the autoclave to room temperature, the white powder product was collected, washed with DMF and dried overnight at 363 K. Activated 17O-enriched α-Mg3(HCOO)6 was obtained by heating as-made 17O-enriched α-Mg3(HCOO)6 at 423 K overnight under dynamic vacuum. The PXRD patterns of 17O-enriched α-Mg3(HCOO)6 samples are shown in Figure S1, Supporting Information (SI).

In the experimental PXRD patterns of as-made and activated α-Mg3(HCOO)6, the two low angle reflections at 2θ = 9.7° (−1 0 1) and 9.9° (1 0 1) are much weaker than the reflection at 10.7° (0 1 1), which strays from the comparable intensities apparent in the simulated PXRD spectra. A previous study demonstrated that the PXRD patterns of α-Mg3(HCOO)6 crystals prepared under different synthetic conditions may exhibit different intensity patterns.65 In particular, the three low-angle reflections at 9.7°, 9.9°, and 10.7° have different relative intensities if water is involved. For α-Mg3(HCOO)6 samples prepared in the presence of water, the relative intensity pattern for the three above-mentioned low-angle diffractions looks very similar to that seen in Figure S1, but distinct from the simulated PXRD pattern based on the structure determined from a single crystal prepared in non-aqueous DMF solvent.60 In the present work, 17O-enriched α-Mg3(HCOO)6 was prepared in the presence of 17O-enriched H2O. Although the synthesis of α-Mg3(HCOO)6 is very straightforward and highly reproducible, and we have made this MOF routinely in many studies with and without water,16, 43-46, 64, 66-68 to unambiguously confirm the identity of the samples used in the present study we repeated the sample preparation several times. These additional samples were prepared under the exact same conditions used for preparing 17O-enriched MOF, except that normal water was used rather than 17O-enriched water. Figure S1 indicates that when the samples were prepared in the presence of a small amount of water, the relative intensities of the first two low-angle reflections are significantly lower than that of the third one, which is consistent with the literature.65 We performed a Le Bail fit of the PXRD data using the GSAS II package (Figure S2, see the SI for details),69 and obtained unit cell parameters of the samples prepared in the presence of water that are comparable to those reported in the literature (Table S1); this data indicates that although the relative intensities of reflections may differ, the samples indeed share the same crystal structure.

The activation process for this MOF has been proven to be robust.60, 64, 66, 70 To verify the solvent DMF molecules occluded inside the channels are completely removed upon activation, TGA profiles and 1H-13C CP/MAS SSNMR spectra were obtained (Figure S3 and S4, with experimental details in Section S2). The results unanimously agree that the activation process is very effective and the solvent molecules are completely removed.

The synthesis of 17O-enriched MIL-53(Al) samples and corresponding PXRD patterns (Figure S5) can be found in Section S3 of the Supporting Information.

2.2. 17O solid-state NMR measurements.

The 1D rotor-synchronized spin-echo spectrum of activated α-Mg3(HCOO)6 was recorded at 21.1 T (17O Larmor frequency of 122.0 MHz) on a Bruker Avance II spectrometer at the National Ultrahigh-Field NMR Facility for Solids in Ottawa, Canada. A 4 mm H/X MAS Bruker probe and a spinning frequency of 18 kHz were used. The recycle delay was 4 s. A π/2 pulse of 4 μs was used.

1D and 2D 17O SSNMR experiments at 35.2 T were performed on the series-connected hybrid (SCH) magnet (17O Larmor frequency of 203.4 MHz) at the National High Magnetic Field Laboratory (NHMFL) in Tallahassee, USA.33 A Bruker Avance NEO console and a NHMFL home-built single-resonance 3.2 mm low-gamma MAS probe were used, along with a spinning frequency of 18 kHz and a pulse delay of 0.1 s. The pulse delay was optimized to achieve the highest S/N and no significant changes in NMR lineshapes were observed when employing different pulse delays. A π/2 pulse of 5.0 μs was used in 1D rotor-synchronized spin echo experiments. All rotor-synchronized 3QMAS spectra were acquired using the shifted-echo pulse sequence.71 The 3Q excitation and conversion pulses were 3.0 and 1.0 μs, respectively. The number of t1 increments was 34. Since the shifted-echo is phase modulated, the number of time increments was also 34, corresponding to a maximum t1 evolution time of ~1.9 ms. Note that during spectral acquisition, the number of t1 increments were carefully chosen to include a significant portion beyond the point where the signal dropped to the baseline noise level, ensuring that the shifted-echo 3QMAS spectra were acquired with the highest S/N during the limited SCH magnet time without compromising the resolution. The 3QMAS spectra were acquired using rotor-synchronized t1 increments to avoid spinning sidebands, which are significant due to the chemical shift anisotropy (CSA) at the very high field and 3Q excitation/conversion modulation along the indirect F1 dimension. However, this restricts the F1 window to the spinning frequency (18 kHz), which is not wide enough to cover the spread of resonances and spinning side bands (SSBs). Fortunately, this problem can be solved by the Q-shearing method to the k = 3 Q-representation.72 The F1 spectral window can then be zero-filled and expanded to whatever limits are needed for shearing back to the isotropic representation with a large and unfolded F1 window. In the present case, the F1 spectral window was zero-filled eight times and expanded to give an unfolded F1 window of 8 × 18 kHz after Q-shearing. Consequently, the F1 digital resolution is equivalent to 8 × 34 t1 increments with the final expanded F1 window. The 17O spectra were referenced to 18 atom% 17O-enriched H217O(l) or distilled water at 0 ppm.

2.3. Spectral simulations.

For quadrupolar nuclei such as 17O (spin I > 1/2), their electric quadrupole moments interact with the surrounding EFG, resulting in broad powder patterns rather than sharper resonances. The interplay between the quadrupolar interaction with the CSA effect makes the shapes of powder patterns more complicated and difficult to simulate. The dmfit software package was used to simulate SSNMR spectra using the Int2QUAD mode, including both the quadrupolar and CSA effects.73 In dmfit, the EFG tensor is described by three principal components in the following order: ∣VYY∣ ≤ ∣VXX∣ ≤ ∣VZZ∣. The quadrupolar coupling constant (CQ) and asymmetry parameter (ηQ) describe the spherical and cylindrical symmetry of the EFG tensor, respectively, and are defined as follows: CQ = (eQVZZ/h) × 9.7177 × 1021 (in Hz) and ηQ = (VYYVXX)/VZZ, where e is the electric charge, Q is the quadrupole moment (−2.558 × 10−30 m2),32 and a conversion factor of 9.7177 × 1021 V·m−2 is used during the calculation of CQ to convert from atomic units to Hz. The chemical shift (CS) tensor is also described by three principal components such that ∣δ22δiso∣ ≤ ∣ δ11δiso∣ ≤ ∣ δ33δiso∣, and the isotropic chemical shift δiso= (δ11 + δ22 + δ33)/3 relates to the bonding modes. The CSA parameters are defined by ΔCS = δ33δiso and ηCS = ∣( δ22δ11)/ ΔCS∣. Three Euler angles (ϕ, χ, ψ) are employed to describe the orientations of the CS tensor with respect to the EFG principal (fixed) axis frame using the ZYZ convention: the corresponding transformation matrix was used to deduce the new directional characteristics of the CS tensor with respect to the EFG system. As a result, eight independent parameters, CQ, ηQ, δiso, ΔCS, ηCS, ϕ, χ, and ψ are required to characterize a single 17O site when both the quadrupolar and the CSA effects are considered. All uncertainties in NMR parameters were estimated by bidirectional variation of the parameter of interest in both directions from the best-fit value while holding all other NMR parameters constant.

2.4. Theoretical calculations.

The unit cell parameters were set to the single-crystal XRD parameters60 and kept fixed during geometry optimizations to ensure consistency between experimental and optimized structures. Proton positions were then optimized using the VASP (Vienna Ab-initio Simulation Package) code74 based on the Kohn-Sham Density Functional Theory (DFT) and using a plane-wave pseudopotential approach. The NMR parameters were then calculated within Kohn-Sham DFT using the QUANTUM-ESPRESSO code.75 The PBE generalized gradient approximation76 was used and the valence electrons were described by norm-conserving pseudopotentials77 in the Kleinman Bylander form.78 The wave functions were expanded on a plane wave basis set with a kinetic energy cut-off of 80 Ry. The integral over the first Brillouin zone was performed using a Monkhorst–Pack 2 × 2 × 2 k-point grid for the charge density and chemical shift tensor calculation. The magnetic shielding tensor was computed using the Gauge-Including Projector Augmented Wave (GIPAW) approach,79-81 which enables the reproduction of the results of a fully converged all-electron calculation. The isotropic chemical shift δiso is defined as δiso = [σisoσiso(ref)], where σiso is the isotropic magnetic shielding and σiso(ref) is the isotropic magnetic shielding of the same nucleus in a reference compound. In the present case, the fit of the linear correlation between the experimental δiso and the calculated σiso values of 17O for Na2SiO3, α-Na2Si2O5, α- and γ-glycine, and α-SrSiO3 enabled the determination of the relation between δiso and calculated σiso for the 17O nucleus, as described previously.21 It is worth noting that for most MOFs, the solvent molecules are disordered inside the framework. Very often, disordered solvent molecules have to be removed before calculation. In the present case, the DMF molecules are orderly distributed within the channels. This rare situation makes the calculation not only simpler, but also more accurate.

3. Results and Discussion

The three-dimensional framework of microporous α-Mg3(HCOO)6 is formed by corner- and edge-sharing MgO6 octahedra interconnected by formate ligands (Figure 1), and features zig-zag channels measuring 4.5 × 5.5 Å.60 Among the twelve crystallographically distinct framework oxygen sites, six are associated with the carboxylate group and adopt a μ2-O bonding mode (sites O1, O3, O5, O7, O9, and O11), and the other six are associated with the carboxylate group and are in a μ1-O bonding mode (sites O2, O4, O6, O8, O10, and O12). The C─μ1-O bonds are of shorter length and more double-bond character than the C─μ2-O bonds. However, the local environments of all six μ1-O sites are almost identical, while the six μ2-O sites are also very similar. As a result, only two signal groups were resolved in previous 17O 1D magic-angle spinning (MAS) spectrum of as-made α-Mg3(HCOO)6 at 21.1 T (Figure 1).16 Those two 17O signals were simulated reasonably well by two 17O powder patterns at 21.1 T and were assigned to groups of μ1- and μ2-O sites, respectively (Figure S6, Supporting Information).

Figure 1.

Figure 1.

Top (left to right): Representations of the framework of activated α-Mg3(HCOO)6, twelve framework oxygen sites, and two different oxygen bonding modes. Color coding: Mg, turquoise; C, grey; H, white; O, red; μ1-O, orange; μ2-O, pink. Bottom: 17O 1D MAS NMR spectra of 17O-enriched α-Mg3(HCOO)6 at fields of 35.2 T (blue) and 21.1 T (black) acquired at a spinning frequency of 18 kHz. The asterisk (*) denotes spinning sidebands (SSBs). The 17O 1D MAS NMR spectrum of as-made α-Mg3(HCOO)6 at 21.1 T is adapted with permission from the American Chemical Society.16

As-made MOFs typically consist of solvent molecules occupying their pores and channels; thus the creation of permeable spaces in MOFs by evacuating the solvent (i.e., the activation process) is a pre-requisite for many applications. Therefore, it is of fundamental importance to understand the effect of activation on the MOF structure, especially regarding the subtle local environment changes that are invisible in diffraction-based techniques but very important for applications. The single-crystal XRD data of as-made and activated α-Mg3(HCOO)6 phases indicate that the local oxygen environments only undergo very minor changes upon removal of DMF solvent during activation, as the Mg ions are coordinately saturated and this MOF framework is fairly rigid.60 Consequently, detecting the very small activation-induced structural changes via 17O SSNMR is quite challenging, as evidenced by the nearly identical 17O 1D MAS spectra of two phases at 21.1 T (Figure 1). It is apparent that higher spectral resolution is necessary to detect the subtle difference in oxygen local environment. The activation method employed in this work for DMF solvent removal from α-Mg3(HCOO)6 is well established.60, 70 Based on our previous experience, we are certain that this activation process completely removes all residual DMF guests from the as-made MOF.64, 66 To confirm, we carried out TGA alongside 1H─13C CP/MAS experiments, and the results unambiguously prove that the activation is complete (see Section S2 for details).

The newly-obtained 17O 1D MAS NMR spectra of as-made and activated α-Mg3(HCOO)6 phases at 35.2 T are shown in Figure 1 alongside the 21.1 T spectra. At 35.2 T, the resonances of both phases are considerably narrower, owing to the reduced second-order quadrupolar broadening. The spectral envelope containing all overlapping signals of the μ1-O group is now completely separated from that of the μ2-O group at 35.2 T, and several diagnostic spectral features including the “edges” and “horns” of individual 17O SSNMR resonances have emerged. Thus, this spectral envelope must consist of several overlapping powder patterns corresponding to multiple μ1-O sites. Similarly, the spectral envelope of the μ2-O group should also be simulated using multiple powder patterns. However, due to the severe overlap of powder patterns, it is very challenging to determine the number of oxygen sites and their corresponding NMR parameters using only a 1D MAS spectrum. Furthermore, the 17O 1D MAS spectra of as-made and activated α-Mg3(HCOO)6 phases are now distinctly different at 35.2 T, implying that the activation-induced structural changes not apparent at 21.1 T are now discernable, owing to the much higher spectral resolution achieved at 35.2 T. It is worth mentioning that the changes in the μ1-O spectral envelope are also more significant than those in the μ2-O one, suggesting that the local environments of μ1-O sites are more influenced by activation versus μ2-O sites. The spectra exhibit very intense spinning sidebands due to the chemical shift anisotropy (CSA) enhanced at 35.2 T.

As mentioned earlier, the 17O 1D MAS spectra at 35.2 T feature overlapping resonances arising from multiple 17O sites, indicating that the maximum achievable 1D spectral resolution is still not high enough to resolve fine features. This is because the conventional 1D MAS experiments only partially averages the 17O second-order quadrupolar interaction. To further enhance spectral resolution, 2D 3QMAS experiments were performed:82 this technique can eliminate the 17O second-order quadrupolar broadening along the indirect F1 dimension and thus separate the overlapping signals observed in 1D MAS spectra. The 17O 2D 3QMAS spectra of as-made and activated α-Mg3(HCOO)6 at 35.2 T are shown in Figure 2. There are a number of well-resolved signals along the F1 dimension in the spectra of the as-made and activated phases, respectively. Since the number of resolved signals in the isotropic dimension (seven for the as-made phase and eight for the activated phase) is smaller than the twelve crystallographically distinct framework oxygen sites, some signals along the F1 dimension must correspond to very similar signals arising from multiple oxygen sites with almost identical NMR parameters.

Figure 2.

Figure 2.

17O 2D 3QMAS NMR spectra of 17O-enriched α-Mg3(HCOO)6 at 35.2 T. Black dashed lines correspond to the slices examined. Blue and red solid lines denote experimental and simulated spectra, respectively. The asterisk (*) denotes the SSBs. The 3QMAS spectra were acquired using a shifted-echo MQMAS pulse sequence and rotor synchronized t1, and processed using the Q-shearing method to avoid spectral folding of the peaks and SSBs.72 The 3QMAS spectra without markups are shown in Figure S7 for clarity.

For each F2 cross-section extracted at δ1 along the F1 dimension, the isotropic chemical shift, δiso (in ppm), and the quadrupolar product, PQ = CQ(1 + ηQ2/3)1/2 (in MHz), can be obtained directly from the spectral center of gravity (δ2) along the F2 dimension17 using the following equations83

δiso=1727δ1+1027δ2
PQ={17081[4I(2I1)]2[4I(I+1)3](δ1δ2)}12v0×103

where ν0 is the Larmor frequency and I is the spin quantum number.

The δiso and PQ values derived from each peak along the F1 dimension are given in Table 1. These two values are determined accurately from the resonance positions in F1 and F2 dimensions, without the need of fitting the F2 cross-section, under the S/N obtained with the limited magnet time.

Table 1.

Experimental 17O NMR parameters, calculateda δiso values, and peak assignments of α-Mg3(HCOO)6.

Sample O Type δ1 (ppm) PQ (MHz) δiso (ppm) δiso, calc (ppm) Assignment CQ (MHz) ηQ
As-made μ2-O 224(1) 7.2(5) 220(2) 227.1 O3 6.5(4) 0.80(10)
μ2-O 227(1) 6.6(5) 223(2) 230.5 O9 6.0(4) 0.88(10)
μ2-O 234(1) 6.6(5) 230(2) 236.1 O1
238.9 O11
239.4 O5
μ2-O 241(1) 7.8(5) 236(2) 256.4 O7 7.0(4) 0.70(10)
μ1-O 278(1) 7.5(5) 273(2) 288.7 O10
291.2 O6
293.2 O4
μ1-O 293(1) 8.8(5) 286(2) 295.5 O12
297.2 O8
μ1-O 296(1) 8.3(5) 290(2) 306.5 O2 7.9(4) 0.55(10)
DMF 305(1) 9.1(5) 298(2) 308.3 O1S
Activated μ2-O 225(1) 6.6(5) 221(2) 227.7 O9 6.0(4) 0.85(10)
μ2-O 230(1) 7.5(5) 225(2) 232.0 O3
232.0 O5
μ2-O 232(1) 7.2(5) 228(2) 234.7 O1 6.5(4) 0.80(10)
μ2-O 235(1) 6.9(5) 231(2) 238.7 O11 6.1(4) 0.85(10)
μ2-O 239(1) 7.2(5) 235(2) 243.5 O7 6.4(4) 0.80(10)
μ1-O 290(1) 7.8(5) 285(2) 285.3 O6 7.7(4) 0.20(10)
μ1-O 295(1) 8.1(5) 289(2) 288.0 O8
289.9 O10
μ1-O 302(1) 7.5(5) 297(2) 296.4 O12
296.5 O4
298.5 O2

The numbers in the parentheses are the estimated uncertainty of the last significant figure.

a

The complete calculated 17O NMR parameters are shown in Table S2.

For the peaks along the isotropic dimension corresponding to a single oxygen site, CQ and ηQ values can be extracted by fitting the F2 cross-section. If a peak along the F1 dimension originates from multiple oxygen sites, δiso and PQ represent average values. To assign each resolved signal in the isotropic dimension to a single or multiple oxygen sites, gauge-including projector augmented wave (GIPAW) density functional theory (DFT) calculations were carried out,79-81 as this approach has proven to be very reliable for 17O NMR spectral assignments in various systems.13-15, 20-21, 24 Although the calculated δiso value may not exactly match the experimental value, assignments of multiple signals based on relative calculated δiso values are typically valid.13-15, 20-21, 24 Accordingly, calculated δiso values were utilized for the assignment of each signal in the F1 dimension (Table 1) to the framework oxygen site(s). For example, the order of calculated δiso values for μ2-O sites in as-made α-Mg3(HCOO)6 is O3 < O9 < O1 ≈ O11 ≈ O5 < O7. The 17O signal with the lowest measured δiso of 220 ppm (δ1 = 224 ppm) is thus assigned to O3, the 17O signal with the second-lowest δiso of 223 ppm (δ1 = 227 ppm) is assigned to O9, the 17O signal with the third-lowest δiso of 230 ppm (δ1 = 234 ppm) and significantly higher intensity is assigned to O1, O5 and O11, and the 17O signal with the highest δiso of 236 ppm (δ1 = 241 ppm) is assigned to O7. The 17O signals of μ1-O sites are assigned in a similar fashion. Since the 17O signals at δ1 = 224 ppm, 227 ppm, and 241 ppm each correspond to a single oxygen site, their CQ and ηQ values can be further extracted by fitting the F2 cross-sections (Table 1). The 2D 3QMAS spectrum of activated α-Mg3(HCOO)6 was analyzed with the same approach and the results are also shown in Table 1.

The spinning sidebands in 1D spectra are particularly intense, suggesting a very large chemical shift anisotropy (CSA) is present at the ultrahigh magnetic field of 35.2 T. Efforts were made to simulate 17O 1D MAS spectra to estimate the 17O CSA. Theoretically, fitting a 1D MAS spectrum with 12 sites requires 96 independent parameters if both the EFG and CSA effects are considered. Therefore, to practically simulate the spectrum, some approximations must be employed to reduce the number of fitting parameters. Thus, the experimental CQ, ηQ, and δiso values of oxygen sites that were resolved in the F1 dimension (e.g., O2, O3, O7, and O9 sites of as-made phase) were directly used in simulations without adjustment. For oxygen sites that are unresolved in the F1 dimension (e.g., O1, O5, and O11 sites of as-made phase), the PQ and δiso are average values. However, it is evident from the lineshape of cross-sections that these sites give rise to very similar NMR parameters. For the 1D spectral simulations, the calculated ηQ values (Table S2) were used without further adjustment and the corresponding CQ values were obtained from the known relationship between PQ and CQ/ηQ. The δiso values were obtained using the average δiso as a starting point and making slight adjustments. Keeping the CQ and ηQ constant during the simulation is reasonable as the anisotropy from the EFG is much smaller compared to the CSA at the ultrahigh magnetic field. The small variation of the EFG parameters among different sites contributes very little to the intensity of the SSBs (Figure S8).

To include the CSA effects, additional NMR parameters must be incorporated: two CSA parameters including the reduced anisotropy ΔCS and the chemical shift asymmetry parameter ηCS, along with three Euler angles (ϕ, χ, ψ) describing the orientations of the chemical shift tensor with respect to the EFG tensor principal axis frame (see Experimental).73 It is reasonable to assume that the 6 μ1-O sites have the same ΔCS, since they reside in very similar chemical environments and the ΔCS value for all 6 μ2-O sites have the same value yet are distinct from that of μ1-O sites. The ΔCS values for μ1-O and μ2-O were modified during simulations, but the calculated ηCS values and Euler angles (Table S2) were kept constant.

The final simulated 1D MAS spectra are shown in Figure 3, illustrating that the isotropic regions and SSBs can be simulated reasonably well by considering both the quadrupolar and the CSA effects (see Table S3 for the final NMR parameters for simulation). If only the second-order quadrupolar interaction is considered, the isotropic regions of 17O signals can still be simulated accurately, but the simulated SSBs are far too low in intensity (Figure S8). It is also worth noting that there is a very weak 17O signal at δ1 = 305 ppm (PQ = 9.1 MHz, δiso = 298 ppm) in the 3QMAS spectrum. The signal position suggests that it arises from DMF molecules within MOF channels.84 Apparently, DMF oxygen atoms are also subject to 17O exchange under the reaction conditions. The enhanced sensitivity and resolution of 17O SSNMR at 35.2 T permit detection of the DMF signal, thus it was included in the simulation of 1D MAS spectrum of as-made α-Mg3(HCOO)6 (i.e., the O1S site).

Figure 3.

Figure 3.

Experimental and simulated 17O 1D MAS NMR spectra of 17O-enriched α-Mg3(HCOO)6 at 35.2 T. Both the quadrupolar and CSA effects are considered in these simulations, using the parameters shown in Table S3. In each phase, the signal intensity of each individual oxygen in the μ1-O sites is approximately equal, and the same is true for signals arising from the μ2-O sites. The asterisk (*) denotes SSBs.

In general, the calculated δiso values are slightly overestimated compared to experimental values. Several factors can be responsible for these discrepancies, including (i) limitations of the GIPAW method, (ii) possible inaccuracy of the crystal structure, (iii) temperature effects on the crystal structure, and (iv) dynamics within the crystal structure.85 In this case, the single crystal structures are based on the XRD data obtained at 100 K. In addition, our DFT calculations do not consider molecular motions such as the librational motions of formate anions reported for the related β-Ca(HCOO)2 at room temperature,86 while NMR experiments are subject to their influence.

17O 2D 3QMAS experiments at an ultrahigh magnetic field strength of 35.2 T has unlocked the identification of twelve inequivalent framework oxygen sites in both the as-made and activated phases of α-Mg3(HCOO)6. The experimental PQ and δiso values of μ1-O sites range from 7.5 to 8.8 MHz and 273 to 297 ppm, respectively; these ranges are 6.6 to 7.8 MHz and 220 to 236 ppm for μ2-O sites. Both sets of these ranges are typical for C═O and C─O environments of carboxylates.3, 8

The very high resolution achieved at 35.2 T permits observation of small changes in 17O NMR parameters such as δiso at each oxygen site upon activation, which were not observable at a lower field of 21.1 T. Such changes reflect the influence and interaction of guest DMF solvent molecules with the MOF host. Activation of the as-made α-Mg3(HCOO)6 phase, and corresponding removal of DMF molecules from the pores, only results in subtle changes in local bond angles and distances while long-range order is preserved.60 A comparison between the experimental 17O δiso values of as-made and activated α-Mg3(HCOO)6 phases reveals that more significant changes occur at μ1-O sites versus μ2-O sites. This disparity in local structural changes is because each μ2-O site is firmly anchored to the framework by two Mg and one C atoms, thus the degree of perturbation on their local oxygen coordination spheres by guest molecules is not as evident as the coordination spheres of μ1-O sites, which are only bound to the framework via one Mg and one C atom. The most significant activation-induced changes are associated with the δiso values of O4 and O10 (> 15 ppm). Figure 4 illustrates the orderly arrangement of DMF molecules along the zig-zag channels of α-Mg3(HCOO)6. Each DMF molecule interacts with two adjacent framework formate anions containing O4 and O10 sites. According to the criteria for the formation of O…H─C nonconventional hydrogen bonds (O…H distance < 2.72 Å and O…H─C angle > 130°),87 both the O1S…H5 distance of 2.38 Å and the O1S…H5─C5 bond angle of 157° in as-made α-Mg3(HCOO)6 are in favor of weak hydrogen bonding. Thus, the significant change in the δiso value of O10, bound to C5 via a C─μ1-O bond, is due to the O1S…H5─C5 hydrogen bonding interaction. In the case of O4, although both the O1S…H2 distance of 2.83 Å and the O1S…H2─C2 bond angle of 88° are not very favorable for O1S…H2─C2 hydrogen bond formation, the O1S…C2 distance of 3.00 Å is considerably shorter than the summation of their van der Waals radii (3.22 Å),88 pointing toward van der Waals forces between O1S and C2 as responsible for the significant change in the δiso value of O4 upon activation.

Figure 4.

Figure 4.

Schematic illustrations of DMF guest molecules within the zig-zag channels of as-made α-Mg3(HCOO)6. Only the Mg nodes are shown for clarity in the bottom diagram. The top inset shows a DMF molecule and two adjacent formate anions. The distances listed were extracted from the DFT-optimized structures. Color coding: Mg, turquoise; N, blue; C, grey; H, white; O1S of DMF, red; μ1-O, orange; μ2-O, pink.

The formation of O1S…H5─C5 hydrogen bond is also evident from the δiso value of the DMF amide oxygen site (O1S) in as-made α-Mg3(HCOO)6. As demonstrated in the literature,84 the δiso value of the DMF amide oxygen in the absence of hydrogen bonding is 323 ppm, and this value decreases with increasing hydrogen bonding strength. The δiso value of 298 ppm for O1S corresponds to a hydrogen bonding strength between those of infinitely diluted DMF in ethanol (299.3 ppm) and infinitely diluted DMF in methanol (292.5 ppm).84 It is the high sensitivity and high resolution achieved at an ultrahigh field of 35.2 T that make it possible to detect the site-specific O…H─C nonconventional hydrogen bonding involving O10 and O1S and estimate the strength of this interaction. As mentioned earlier, nonconventional O…H─C hydrogen bonding can play important roles in the host–guest interactions in BioMOFs or MOF-based drug delivery systems.49-51 Thus, the ability to detect this type of weak interaction and estimate its strength by 17O SSNMR at ultrahigh field, as demonstrated in this study, provides researchers a useful tool for investigating this type of host–guest interaction in MOF systems.

To further demonstrate the power of ultrahigh magnetic field on MOF characterization using 17O SSNMR, we examined two activated MOF MIL-53(Al) samples. MIL-53(Al) is a well-studied MOF with high thermal and chemical stability, and is very promising in guest separation.89 As Figure 5a illustrates, the channels of as-made MIL-53(Al) are occupied by the unreacted linker precursor molecules, 1,4-benzenedicarboxylic acid (H2BDC). After activation, the channel dimension increases from 7.3 × 7.7 Å2 to 8.5 × 8.5 Å2 due to the removal of the hydrogen bonding interaction between H2BDC and the carboxylate of BDC2− linkers in the framework. Upon activation, the framework topology is retained, but the crystal symmetry changes from Pnma in the as-made phase to Imma in the activated phase.47 There are two reports on 17O SSNMR studies of MIL-53 conducted at lower magnetic fields.16-17

Figure 5.

Figure 5.

(a) 17O 1D MAS NMR spectra of 17O-enriched fully- (Sample A) and partially- (Sample B) activated MIL-53(Al) samples at 35.2 T. The blue and red solid lines denote experimental and simulated spectra, respectively. Only the carboxylate oxygen regions are shown for clarity. The full spectra are shown in Figure S9. The asterisk (*) and dollar sign ($) denotes the SSBs of −COO and μ2-OH, respectively. The structures of as-made and activated MIL-53(Al) phases are shown at top. (b) 17O 2D 3QMAS NMR spectrum of partially-activated (Sample B) 17O-enriched MIL-53(Al) at 35.2 T. Black dashed lines correspond to the slices examined for analyses.

Activation of MIL-53 is not very straightforward. Early on, activation of as-made MIL-53(Al) was performed by directly calcination in air at 503 K for 3 days.47 The conditions were harsh and led to the reduced crystallinity. To avoid such harsh activation conditions, a milder alternate route was developed by first exchanging the trapped H2BDC with DMF under solvothermal conditions (e.g., 423 K) for an extended period (e.g., 12 h or longer) and then heating the DMF-exchanged MOF at 473 K under dynamic vacuum (≤ 1 mbar) overnight.89 We prepared two 17O-enriched MIL-53(Al) samples which were activated under identical conditions except the DMF exchange time (see Section S3): 24 h for Sample A and 12h for Sample B. According to literature, an exchange time of 12 h should be sufficient for the activation of MIL-53(Al).90 Although the PXRD patterns of Sample A and B look very similar (Figure S5), the 17O 1D MAS spectra at 35.2 T of Samples A and B look distinctly different in the carboxylate region (Figure 5a). The spectrum of Sample A exhibits a relatively narrow pattern that can be well fitted with a single set of 17O NMR parameters (Table S4), indicating that the signal corresponds to a single oxygen site. This is consistent with the crystal structure of the activated phase which only has one carboxylate oxygen site in the unit cell.47 Thus, the 17O MAS spectrum clearly indicates that the sample exchanged with DMF for 24 h is fully activated.

The spectrum of Sample B displays a rather broad profile in the carboxylate region that cannot be simulated by a single oxygen site, implying the existence of multiple carboxylate oxygen environments in this sample. It appears that this MOF sample is only partially-activated. The 17O 3QMAS spectrum of Sample B obtained at 35.2 T provides detailed information. A long spectral “ridge” is observed for the carboxylate oxygen sites, implying that the partial activation can contribute to the distributions on the quadrupolar coupling and chemical shift, consequently line broadening. Such a situation is not unexpected since MIL-53(Al) is a flexible MOF that undergoes a phase transition from Pnma in the as-made phase to Imma upon full activation without breaking any bonds. It appears that Sample B represents an intermediate state between these two phases. Furthermore, the unreacted linker precursors in the channels are disordered and seem to be randomly removed. Nonetheless, five peaks are extracted above the “ridge” at δ1 = 212, 217, 226, 237, and 244 ppm, respectively. The corresponding PQ and δiso values are shown in Table S4. Based on the δiso values, the peaks at δ1 = 212 and 217 ppm are assigned to the ─OH and C═O of H2BDC molecules within the MOF channels, respectively.33 The signal at δ1 = 237 ppm is attributed to the oxygen site in the empty channels because its δiso value (232 ppm) is similar to that of activated MIL-53(Al), while the resonances at δ1 = 226 and 244 ppm are tentatively assigned to the oxygen sites in the occupied channels with local environments similar to those in as-made MIL-53(Al).16-17 The high resolution and sensitivity of 17O NMR gained at 35.2 T not only allow for unambiguously distinguishing partially-activated from fully-activated MIL-53(Al) samples, but also permit observation of several oxygen sites in empty and occupied channels in partially-activated MIL-53(Al).

4. Conclusions

In summary, the very high spectral resolution and sensitivity achieved at an ultrahigh magnetic field of 35.2 T in this work represents an advance in 17O SSNMR spectroscopy. At 35.2 T, many inequivalent carboxylate oxygen sites have been identified in 17O SSNMR spectra of both the activated and as-made α-Mg3(HCOO)6 MOFs. The very high resolution achieved at 35.2 T enables the observation of subtle changes in 17O SSNMR spectra of as-made and activated α-Mg3(HCOO)6 phases. These alterations arise from weak site-specific interactions between DMF guests and the MOF framework, such as hydrogen bonding and van der Waals forces. The investigation of these weak interactions is important for MOF applications in various fields including gas adsorption and biomedical applications. The advantage of performing 17O SSNMR experiments at 35.2 T for MOF characterization is further illustrated by activation of MIL-53(Al). The partially- and fully-activated phases of MIL-53(Al) can be unambiguously distinguished. Several oxygen sites with different local environments in the partially-activated phase are tentatively identified.

This work illustrates how a wide variety of organic and inorganic compounds are now viable targets for 17O SSNMR at an ultrahigh magnetic field of 35.2 T. The sensitivity and resolution afforded at this field strength greatly extend the volume and quality of structural and chemical information available from 17O SSNMR spectroscopy, as much of this data is unavailable at or below magnetic fields of 21.1 T.

Supplementary Material

supporting information

ACKNOWLEDGMENTS

J.X. thanks the financial support from the National Natural Science Foundation of China (Project 21904071) and the Open Funds (KF1818) of the State Key Laboratory of Fine Chemicals. Y.H. thanks the Natural Sciences and Engineering Research Council (NSERC) of Canada for a Discovery Grant. A portion of this work was performed at the NHMFL, which is supported by NSF DMR-1644779 and the State of Florida. In addition, development of the SCH magnet and NMR instrumentation was supported by the NSF (DMR-1039938 and DMR-0603042) and NIH (BTRR 1P41 GM122698). We thank Dr. Victor V. Terskikh at University of Ottawa for acquiring 17O SSNMR spectra at 21.1 T.

Footnotes

Supporting Information.

Powder XRD patterns, TGA profiles, and 1H─13C CP/MAS NMR spectra of α-Mg3(HCOO)6, experimental details on 17O-enriched MIL-53(Al) preparation and the powder XRD patterns of this MOF, additional NMR data and analyses. The Supporting Information is available free of charge on the ACS Publications website.

REFERENCES

  • (1).Ashbrook SE; Smith ME, Solid State 17O NMR - An Introduction to the Background Principles and Applications to Inorganic Materials. Chem. Soc. Rev 2006, 35 (8), 718–735. [DOI] [PubMed] [Google Scholar]
  • (2).Brownbill NJ; Gajan D; Lesage A; Emsley L; Blanc F, Oxygen-17 Dynamic Nuclear Polarisation Enhanced Solid-State NMR Spectroscopy at 18.8 T. Chem. Commun 2017, 53, 2563–2566. [DOI] [PubMed] [Google Scholar]
  • (3).Wu G, Solid-State 17O NMR Studies of Organic and Biological Molecules. Prog. Nucl. Magn. Reson. Spectrosc 2008, 52 (2-3), 118–169. [DOI] [PubMed] [Google Scholar]
  • (4).Yamada K, Chapter 3 - Recent Applications of Solid-State 17O NMR. Annu. Rep. NMR Spectrosc 2010, 70, 115–158. [Google Scholar]
  • (5).Gerothanassis IP, Oxygen-17 NMR Spectroscopy: Basic Principles and Applications (Part II). Prog. Nucl. Magn. Reson. Spectrosc 2010, 57 (1), 1–110. [DOI] [PubMed] [Google Scholar]
  • (6).Castiglione F; Mele A; Raos G, Chapter Four - 17O NMR: A “Rare and Sensitive” Probe of Molecular Interactions and Dynamics. Annu. Rep. NMR Spectrosc 2015, 85, 143–193. [Google Scholar]
  • (7).Ohlin CA; Casey WH, Chapter Five - 17O NMR as a Tool in Discrete Metal Oxide Cluster Chemistry. Annu. Rep. NMR Spectrosc 2018, 94, 187–248. [Google Scholar]
  • (8).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. [DOI] [PubMed] [Google Scholar]
  • (9).Buannic L; Blanc F; Middlemiss DS; Grey CP, Probing Cation and Vacancy Ordering in the Dry and Hydrated Yttrium-Substituted BaSnO3 Perovskite by NMR Spectroscopy and First Principles Calculations: Implications for Proton Mobility. J. Am. Chem. Soc 2012, 134 (35), 14483–14498. [DOI] [PubMed] [Google Scholar]
  • (10).Wang WD; Lucier BEG; Terskikh VV; Wang W; Huang Y, Wobbling and Hopping: Studying Dynamics of CO2 Adsorbed in Metal–Organic Frameworks via 17O Solid-State NMR. J. Phys. Chem. Lett 2014, 5 (19), 3360–3365. [DOI] [PubMed] [Google Scholar]
  • (11).Holmes ST; Schurko RW, Refining Crystal Structures with Quadrupolar NMR and Dispersion-Corrected Density Functional Theory. J. Phys. Chem. C 2017, 122 (3), 1809–1820. [Google Scholar]
  • (12).Alam TM; Nyman M; Cherry BR; Segall JM; Lybarger LE, Multinuclear NMR Investigations of the Oxygen, Water, and Hydroxyl Environments in Sodium Hexaniobate. J. Am. Chem. Soc 2004, 126 (17), 5610–5620. [DOI] [PubMed] [Google Scholar]
  • (13).Pedone A; Gambuzzi E; Menziani MC, Unambiguous Description of the Oxygen Environment in Multicomponent Aluminosilicate Glasses from 17O Solid State NMR Computational Spectroscopy. J. Phys. Chem. C 2012, 116 (27), 14599–14609. [Google Scholar]
  • (14).Romao CP; Perras FA; Werner-Zwanziger U; Lussier JA; Miller KJ; Calahoo CM; Zwanziger JW; Bieringer M; Marinkovic BA; Bryce DL; White MA, Zero Thermal Expansion in ZrMgMo3O12: NMR Crystallography Reveals Origins of Thermoelastic Properties. Chem. Mater 2015, 27 (7), 2633–2646. [Google Scholar]
  • (15).Kong X; Terskikh VV; Khade RL; Yang L; Rorick A; Zhang Y; He P; Huang Y; Wu G, Solid-State 17O NMR Spectroscopy of Paramagnetic Coordination Compounds. Angew. Chem. Int. Ed 2015, 54 (16), 4753–4757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (16).He P; Xu J; Terskikh VV; Sutrisno A; Nie H-Y; Huang Y, Identification of Nonequivalent Framework Oxygen Species in Metal–Organic Frameworks by 17O Solid-State NMR. J. Phys. Chem. C 2013, 117 (33), 16953–16960. [Google Scholar]
  • (17).Bignami GPM; Davis ZH; Dawson DM; Morris SA; Russell SE; McKay D; Parke RE; Iuga D; Morris RE; Ashbrook SE, Cost-Effective 17O Enrichment and NMR Spectroscopy of Mixed-Metal Terephthalate Metal-Organic Frameworks. Chem. Sci 2018, 9, 850–859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Chen C-H; Gaillard E; Mentink-Vigier F; Chen K; Gan Z; Gaveau P; Rebière B; Berthelot R; Florian P; Bonhomme C; Smith ME; Métro T-X; Alonso B; Laurencin D, Direct 17O Isotopic Labeling of Oxides Using Mechanochemistry. Inorg. Chem 2020, doi: 10.1021/acs.inorgchem.0c00208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (19).Champouret Y; Coppel Y; Kahn ML, Evidence for Core Oxygen Dynamics and Exchange in Metal Oxide Nanocrystals from In Situ 17O MAS NMR. J. Am. Chem. Soc 2016, 138 (50), 16322–16328. [DOI] [PubMed] [Google Scholar]
  • (20).Wang M; Wu X-P; Zheng S; Zhao L; Li L; Shen L; Gao Y; Xue N; Guo X; Huang W; Gan Z; Blanc F; Yu Z; Ke X; Ding W; Gong X-Q; Grey CP; Peng L, Identification of Different Oxygen Species in Oxide Nanostructures with 17O Solid-State NMR Spectroscopy. Sci. Adv 2015, 1 (1), e1400133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (21).Métro T-X; Gervais C; Martinez A; Bonhomme C; Laurencin D, Unleashing the Potential of 17O NMR Spectroscopy Using Mechanochemistry. Angew. Chem. Int. Ed 2017, 56 (24), 6803–6807. [DOI] [PubMed] [Google Scholar]
  • (22).Griffin JM; Clark L; Seymour VR; Aldous DW; Dawson DM; Iuga D; Morris RE; Ashbrook SE, Ionothermal 17O Enrichment of Oxides Using Microlitre Quantities of Labelled Water. Chem. Sci 2012, 3 (7), 2293–2300. [Google Scholar]
  • (23).Trease NM; Clark TM; Grandinetti PJ; Stebbins JF; Sen S, Bond Length-Bond Angle Correlation in Densified Silica—Results from 17O NMR Spectroscopy. J. Chem. Phys 2017, 146 (18), 184505. [Google Scholar]
  • (24).Pavón E; Osuna FJ; Alba MD; Delevoye L, Natural Abundance 17O MAS NMR and DFT Simulations: New Insights into the Atomic Structure of Designed Micas. Solid State Nucl. Magn. Reson 2019, 100, 45–51. [DOI] [PubMed] [Google Scholar]
  • (25).Pustogow A; Luo Y; Chronister A; Su YS; Sokolov DA; Jerzembeck F; Mackenzie AP; Hicks CW; Kikugawa N; Raghu S; Bauer ED; Brown SE, Constraints on the Superconducting Order Parameter in Sr2RuO4 from Oxygen-17 Nuclear Magnetic Resonance. Nature 2019, 574 (7776), 72–75. [DOI] [PubMed] [Google Scholar]
  • (26).Peng L; Liu Y; Kim N; Readman JE; Grey CP, Detection of Bronsted Acid Sites in Zeolite HY with High-Field 17O-MAS-NMR Techniques. Nat. Mater 2005, 4 (3), 216–219. [DOI] [PubMed] [Google Scholar]
  • (27).Hung I; Uldry A-C; Becker-Baldus J; Webber AL; Wong A; Smith ME; Joyce SA; Yates JR; Pickard CJ; Dupree R; Brown SP, Probing Heteronuclear 15N─17O and 13C─17O Connectivities and Proximities by Solid-State NMR Spectroscopy. J. Am. Chem. Soc 2009, 131 (5), 1820–1834. [DOI] [PubMed] [Google Scholar]
  • (28).Chen L; Lu X; Wang Q; Lafon O; Trébosc J; Deng F; Amoureux J-P, Distance Measurement between A Spin-1/2 and A Half-Integer Quadrupolar Nuclei by Solid-State NMR Using Exact Analytical Expressions. J. Magn. Reson 2010, 206 (2), 269–273. [DOI] [PubMed] [Google Scholar]
  • (29).Carnahan SL; Lampkin BJ; Naik P; Hanrahan MP; Slowing II; VanVeller B; Wu G; Rossini AJ, Probing O─H Bonding through Proton Detected 1H─17O Double Resonance Solid-State NMR Spectroscopy. J. Am. Chem. Soc 2019, 141 (1), 441–450. [DOI] [PubMed] [Google Scholar]
  • (30).Jakobsen HJ; Bildsøe H; Brorson M; Wu G; Gor’kov PL; Gan Z; Hung I, High-Field 17O MAS NMR Reveals 1J(17O-127I) with its Sign and the NMR Crystallography of the Scheelite Structures for NaIO4 and KIO4. J. Phys. Chem. C 2015, 119 (25), 14434–14442. [Google Scholar]
  • (31).Perras FA; Kobayashi T; Pruski M, Natural Abundance 17O DNP Two-Dimensional and Surface-Enhanced NMR Spectroscopy. J. Am. Chem. Soc 2015, 137 (26), 8336–8339. [DOI] [PubMed] [Google Scholar]
  • (32).Harris RK; Becker ED; Cabral De Menezes SM; Goodfellow R; Granger P, NMR Nomenclature. Nuclear Spin Properties and Conventions for Chemical Shifts (IUPAC Recommendations 2001). Pure Appl. Chem 2001, 73 (11), 1795–1818. [DOI] [PubMed] [Google Scholar]
  • (33).Gan Z; Hung I; Wang X; Paulino J; Wu G; Litvak IM; Gor'kov PL; Brey WW; Lendi P; Schiano JL; Bird MD; Dixon IR; Toth J; Boebinger GS; Cross TA, NMR Spectroscopy up to 35.2 T Using A Series-Connected Hybrid Magnet. J. Magn. Reson 2017, 284, 125–136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (34).Keeler EG; Michaelis VK; Colvin MT; Hung I; Gor’kov PL; Cross TA; Gan Z; Griffin RG, 17O MAS NMR Correlation Spectroscopy at High Magnetic Fields. J. Am. Chem. Soc. 2017, 139 (49), 17953–17963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (35).Keeler EG; Michaelis VK; Wilson CB; Hung I; Wang X; Gan Z; Griffin RG, High-Resolution 17O NMR Spectroscopy of Structural Water. J. Phys. Chem. B 2019, 123 (14), 3061–3067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (36).Zhou H-C; Long JR; Yaghi OM, Introduction to Metal–Organic Frameworks. Chem. Rev 2012, 112 (2), 673–674. [DOI] [PubMed] [Google Scholar]
  • (37).Furukawa H; Cordova KE; O’Keeffe M; Yaghi OM, The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341 (6149), 1230444. [DOI] [PubMed] [Google Scholar]
  • (38).Hoffmann HC; Debowski M; Mueller P; Paasch S; Senkovska I; Kaskel S; Brunner E, Solid-State NMR Spectroscopy of Metal–Organic Framework Compounds (MOFs). Materials 2012, 5, 2537–2572. [Google Scholar]
  • (39).Sutrisno A; Huang Y, Solid-State NMR: A Powerful Tool for Characterization of Metal–Organic Frameworks. Solid State Nucl. Magn. Reson 2013, 49–50, 1–11. [DOI] [PubMed] [Google Scholar]
  • (40).Lucier BEG; Chen S; Huang Y, Characterization of Metal–Organic Frameworks: Unlocking the Potential of Solid-State NMR. Acc. Chem. Res 2018, 51 (2), 319–330. [DOI] [PubMed] [Google Scholar]
  • (41).Witherspoon VJ; Xu J; Reimer JA, Solid-State NMR Investigations of Carbon Dioxide Gas in Metal–Organic Frameworks: Insights into Molecular Motion and Adsorptive Behavior. Chem. Rev 2018, 118 (20), 10033–10048. [DOI] [PubMed] [Google Scholar]
  • (42).Wong YTA; Martins V; Lucier BEG; Huang Y, Solid-State NMR Spectroscopy: A Powerful Technique to Directly Study Small Gas Molecules Adsorbed in Metal–Organic Frameworks. Chem. –Eur. J 2019, 25 (8), 1848–1853. [DOI] [PubMed] [Google Scholar]
  • (43).Lucier BEG; Zhang Y; Huang Y, Complete Multinuclear Solid-State NMR of Metal-Organic Frameworks: The Case of α-Mg-formate. Concept. Magnetic Res. A 2016, 45A (6), e21410. [Google Scholar]
  • (44).Lucier BEG; Zhang Y; Lee KJ; Lu Y; Huang Y, Grasping Hydrogen Adsorption and Dynamics in Metal–Organic Frameworks Using 2H Solid-State NMR. Chem. Commun 2016, 52 (48), 7541–7544. [DOI] [PubMed] [Google Scholar]
  • (45).Lu Y; Lucier B; Zhang Y; Ren P; Zheng A; Huang Y, Sizable Dynamics in Small Pores: CO2 Location and Motion in the α-Mg Formate Metal-Organic Framework. Phys. Chem. Chem. Phys 2017, 19, 6130–6141. [DOI] [PubMed] [Google Scholar]
  • (46).Zhang Y; Lucier BEG; Fischer M; Gan Z; Boyle PD; Desveaux B; Huang Y, A Multifaceted Study of Methane Adsorption in Metal–Organic Frameworks by Using Three Complementary Techniques. Chem. –Eur. J 2018, 24 (31), 7866–7881. [DOI] [PubMed] [Google Scholar]
  • (47).Loiseau T; Serre C; Huguenard C; Fink G; Taulelle F; Henry M; Bataille T; Férey G, A Rationale for the Large Breathing of the Porous Aluminum Terephthalate (MIL-53) upon Hydration. Chem. –Eur. J 2004, 10 (6), 1373–1382. [DOI] [PubMed] [Google Scholar]
  • (48).Kitagawa S; Uemura K, Dynamic Porous Properties of Coordination Polymers Inspired by Hydrogen Bonds. Chem. Soc. Rev 2005, 34 (2), 109–119. [DOI] [PubMed] [Google Scholar]
  • (49).McKinlay AC; Morris RE; Horcajada P; Férey G; Gref R; Couvreur P; Serre C, BioMOFs: Metal–Organic Frameworks for Biological and Medical Applications. Angew. Chem. Int. Ed 2010, 49 (36), 6260–6266. [DOI] [PubMed] [Google Scholar]
  • (50).Cai H; Huang Y-L; Li D, Biological Metal–Organic Frameworks: Structures, Host–Guest Chemistry and Bio-Applications. Coord. Chem. Rev 2019, 378, 207–221. [Google Scholar]
  • (51).Wu M-X; Yang Y-W, Metal–Organic Framework (MOF)-Based Drug/Cargo Delivery and Cancer Therapy. Adv. Mater 2017, 29 (23), 1606134. [DOI] [PubMed] [Google Scholar]
  • (52).Li H; Eddaoudi M; O'Keeffe M; Yaghi M, Design and Synthesis of an Exceptionally Stable and Highly Porous Metal–Organic Framework. Nature 1999, 402 (6759), 276–279. [Google Scholar]
  • (53).Millange F; Serre C; Férey G, Synthesis, Structure Determination and Properties of MIL-53as and MIL-53ht: The First CrIII Hybrid Inorganic–Organic Microporous Solids: CrIII(OH)·{O2C─C6H4─CO2}·{HO2C─C6H4─CO2H}x. Chem. Commun 2002, 2002 (8), 822–823. [DOI] [PubMed] [Google Scholar]
  • (54).Hafizovic Cavka J; Jakobsen S; Olsbye U; Guillou N; Lamberti C; Bordiga S; Lillerud KP, A New Zirconium Inorganic Building Brick Forming Metal Organic Frameworks with Exceptional Stability. J. Am. Chem. Soc 2008, 130 (42), 13850–13851. [DOI] [PubMed] [Google Scholar]
  • (55).Rosi NL; Kim J; Eddaoudi M; Chen B; O'Keeffe M; Yaghi OM, Rod Packings and Metal–Organic Frameworks Constructed from Rod-Shaped Secondary Building Units. J. Am. Chem. Soc 2005, 127 (5), 1504–1518. [DOI] [PubMed] [Google Scholar]
  • (56).Chui SS-Y; Lo SM-F; Charmant JPH; Orpen AG; Williams ID, A Chemically Functionalizable Nanoporous Material [Cu3(TMA)2(H2O)3]n. Science 1999, 283 (5405), 1148–1150. [DOI] [PubMed] [Google Scholar]
  • (57).Ahmed I; Jhung SH, Applications of Metal-Organic Frameworks in Adsorption/Separation Processes via Hydrogen Bonding Interactions. Chem. Eng. J 2017, 310, 197–215. [Google Scholar]
  • (58).Roberts JM; Fini BM; Sarjeant AA; Farha OK; Hupp JT; Scheidt KA, Urea Metal–Organic Frameworks as Effective and Size-Selective Hydrogen-Bond Catalysts. J. Am. Chem. Soc 2012, 134 (7), 3334–3337. [DOI] [PubMed] [Google Scholar]
  • (59).Meng X; Wang H-N; Song S-Y; Zhang H-J, Proton-Conducting Crystalline Porous Materials. Chem. Soc. Rev 2017, 46 (2), 464–480. [DOI] [PubMed] [Google Scholar]
  • (60).Rood JA; Noll BC; Henderson KW, Synthesis, Structural Characterization, Gas Sorption and Guest-Exchange Studies of the Lightweight, Porous Metal–Organic Framework α-[Mg3(O2CH)6]. Inorg. Chem 2006, 45 (14), 5521–5528. [DOI] [PubMed] [Google Scholar]
  • (61).Fischer M; Hoffmann F; Froeba M, New Microporous Materials for Acetylene Storage and C2H2/CO2 Separation: Insights from Molecular Simulations. ChemPhysChem 2010, 11 (10), 2220–2229. [DOI] [PubMed] [Google Scholar]
  • (62).Lian X; Fang Y; Joseph E; Wang Q; Li J; Banerjee S; Lollar C; Wang X; Zhou H-C, Enzyme–MOF (Metal–Organic Framework) Composites. Chem. Soc. Rev 2017, 46 (11), 3386–3401. [DOI] [PubMed] [Google Scholar]
  • (63).Novio F; Simmchen J; Vázquez-Mera N; Amorín-Ferré L; Ruiz-Molina D, Coordination Polymer Nanoparticles in Medicine. Coord. Chem. Rev 2013, 257 (19), 2839–2847. [Google Scholar]
  • (64).Xu J; Terskikh VV; Chu Y; Zheng A; Huang Y, Mapping Out Chemically Similar, Crystallographically Nonequivalent Hydrogen Sites in Metal–Organic Frameworks by 1H Solid-State NMR Spectroscopy. Chem. Mater 2015, 27 (9), 3306–3316. [Google Scholar]
  • (65).Hu J; Sun T; Ren X; Wang S, HF-Assisted Synthesis of Ultra-Microporous [Mg3(OOCH)6] Frameworks for Selective Adsorption of CH4 over N2. Microporous Mesoporous Mater. 2015, 204, 73–80. [Google Scholar]
  • (66).Xu J; Terskikh VV; Huang Y, Resolving Multiple Non-equivalent Metal Sites in Magnesium-Containing Metal–Organic Frameworks by Natural Abundance 25Mg Solid-State NMR Spectroscopy. Chem. –Eur. J 2013, 19 (14), 4432–4436. [DOI] [PubMed] [Google Scholar]
  • (67).Mao H; Xu J; Hu Y; Huang Y; Song Y, The Effect of High External Pressure on Structure and Stability of MOF α-Mg3(HCOO)6 Probed by in Situ Raman and FT-IR Spectroscopy. J. Mater. Chem. A 2015, 3, 11976–11984. [Google Scholar]
  • (68).Xu J; Terskikh VV; Chu Y; Zheng A; Huang Y, 13C Chemical Shift Tensors in MOF α-Mg3(HCOO)6: Which Component is More Sensitive to Host-Guest Interaction? Magn. Reson. Chem 2020, doi: 10.1002/mrc.4944. [DOI] [PubMed] [Google Scholar]
  • (69).Toby BH; Von Dreele RB, GSAS-II: The Genesis of a Modern Open-Source All Purpose Crystallography Software Package. J. Appl. Crystallogr 2013, 46 (2), 544–549. [Google Scholar]
  • (70).Rood JA; Henderson KW, Synthesis and Small Molecule Exchange Studies of a Magnesium Bisformate Metal–Organic Framework: An Experiment in Host-Guest Chemistry for the Undergraduate Laboratory. J. Chem. Educ 2013, 90 (3), 379–382. [Google Scholar]
  • (71).Massiot D; Touzo B; Trumeau D; Coutures JP; Virlet J; Florian P; Grandinetti PJ, Two-Dimensional Magic-Angle Spinning Isotropic Reconstruction Sequences for Quadrupolar Nuclei. Solid State Nucl. Magn. Reson 1996, 6 (1), 73–83. [DOI] [PubMed] [Google Scholar]
  • (72).Hung I; Trébosc J; Hoatson GL; Vold RL; Amoureux J-P; Gan Z, Q-Shear Transformation for MQMAS and STMAS NMR Spectra. J. Magn. Reson 2009, 201 (1), 81–86. [DOI] [PubMed] [Google Scholar]
  • (73).Massiot D; Fayon F; Capron M; King I; Le Calve S; Alonso B; Durand J-O; Bujoli B; Gan Z; Hoatson G, Modeling One- and Two-Dimensional Solid-State NMR Spectra. Magn. Reson. Chem 2002, 40 (1), 70–76. [Google Scholar]
  • (74).Kresse G; Hafner J, Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal--Amorphous-Semiconductor Transition in Germanium. Phys. Rev. B 1994, 49 (20), 14251–14269. [DOI] [PubMed] [Google Scholar]
  • (75).Giannozzi P; Baroni S; Bonini N; Calandra M; Car R; Cavazzoni C; Ceresoli D; Chiarotti GL; Cococcioni M; Dabo I; Dal Corso A; de Gironcoli S; Fabris S; Fratesi G; Gebauer R; Gerstmann U; Gougoussis C; Kokalj A; Lazzeri M; Martin-Samos L; Marzari N; Mauri F; Mazzarello R; Paolini S; Pasquarello A; Paulatto L; Sbraccia C; Scandolo S; Sclauzero G; Seitsonen AP; Smogunov A; Umari P; Wentzcovitch RM, QUANTUM ESPRESSO: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys.: Condens. Matter 2009, 21 (39), 395502. [DOI] [PubMed] [Google Scholar]
  • (76).Perdew JP; Burke K; Ernzerhof M, Generalized Gradient Approximation Made Simple. Phys. Rev. Lett 1996, 77 (18), 3865–3868. [DOI] [PubMed] [Google Scholar]
  • (77).Troullier N; Martins JL, Efficient Pseudopotentials for Plane-Wave Calculations. Phys. Rev. B 1991, 43 (3), 1993–2006. [DOI] [PubMed] [Google Scholar]
  • (78).Kleinman L; Bylander DM, Efficacious Form for Model Pseudopotentials. Phys. Rev. Lett 1982, 48 (20), 1425–1428. [Google Scholar]
  • (79).Charpentier T, The PAW/GIPAW Approach for Computing NMR Parameters: A New Dimension Added to NMR Study of Solids. Solid State Nucl. Magn. Reson 2011, 40 (1), 1–20. [DOI] [PubMed] [Google Scholar]
  • (80).Bonhomme C; Gervais C; Babonneau F; Coelho C; Pourpoint F; Azaïs T; Ashbrook SE; Griffin JM; Yates JR; Mauri F; Pickard CJ, First-Principles Calculation of NMR Parameters Using the Gauge Including Projector Augmented Wave Method: A Chemist’s Point of View. Chem. Rev 2012, 112 (11), 5733–5779. [DOI] [PubMed] [Google Scholar]
  • (81).Pickard CJ; Mauri F, All-Electron Magnetic Response with Pseudopotentials: NMR Chemical Shifts. Phys. Rev. B 2001, 63 (24), 245101. [Google Scholar]
  • (82).Frydman L; Harwood JS, Isotropic Spectra of Half-Integer Quadrupolar Spins from Bidimensional Magic-Angle Spinning NMR. J. Am. Chem. Soc 1995, 117 (19), 5367–5368. [Google Scholar]
  • (83).Amoureux J-P; Huguenard C; Engelke F; Taulelle F, Unified Representation of MQMAS and STMAS NMR of Half-Integer Quadrupolar Nuclei. Chem. Phys. Lett 2002, 356 (5), 497–504. [Google Scholar]
  • (84).Gerothanassis IP; Vakka C, 17O NMR Chemical Shifts as a Tool to Study Specific Hydration Sites of Amides and Peptides: Correlation with the IR Amide I Stretching Vibration. J. Org. Chem 1994, 59 (9), 2341–2348. [Google Scholar]
  • (85).Michaelis VK; Keeler EG; Ong T-C; Craigen KN; Penzel S; Wren JEC; Kroeker S; Griffin RG, Structural Insights into Bound Water in Crystalline Amino Acids: Experimental and Theoretical 17O NMR. J. Phys. Chem. B 2015, 119 (25), 8024–8036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (86).Hallock KJ; Lee DK; Ramamoorthy A, The Effects of Librations on the 13C Chemical Shift and 2H Electric Field Gradient Tensors in β-Calcium Formate. J. Chem. Phys 2000, 113 (24), 11187–11193. [Google Scholar]
  • (87).Desiraju GR; Steiner T, The Weak Hydrogen Bond in Structural Chemistry and Biology. Oxford University Press: Oxford/New York, 1999. [Google Scholar]
  • (88).Bondi A, van der Waals Volumes and Radii. J. Phys. Chem 1964, 68 (3), 441–451. [Google Scholar]
  • (89).Trung TK; Trens P; Tanchoux N; Bourrelly S; Llewellyn PL; Loera-Serna S; Serre C; Loiseau T; Fajula F; Férey G, Hydrocarbon Adsorption in the Flexible Metal Organic Frameworks MIL-53(Al, Cr). J. Am. Chem. Soc 2008, 130 (50), 16926–16932. [DOI] [PubMed] [Google Scholar]
  • (90).Zhang Y; Lucier BEG; Huang Y, Deducing CO2 Motion, Adsorption Locations and Binding Strengths in a Flexible Metal–Organic Framework without Open Metal Sites. Phys. Chem. Chem. Phys 2016, 18 (12), 8327–8341. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

supporting information

RESOURCES